Skip to main content
Log in

The fluidic pinball with symmetric forcing displays steady, periodic, quasi-periodic, and chaotic dynamics

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

Abstract

We numerically investigate the fluidic pinball under symmetric forcing and find seven flow regimes under different rotation speeds. The fluidic pinball consists of three rotatable cylinders placed at the vertices of an equilateral triangle pointing upstream in a uniform oncoming flow. The starting point is the unforced asymmetric periodic vortex shedding at Reynolds number Re = 100 based on the cylinder diameter. The flow is symmetrically actuated by rotating the two rear cylinders at constant speed |b| up to three times the oncoming velocity in both directions. Counterclockwise (b > 0) and clockwise (b < 0) rotation of the bottom cylinder correspond to boat tailing and base bleeding, respectively. A total of seven distinct flow regimes are observed, including a steady flow, three symmetric/asymmetric periodic types of shedding, two symmetric/asymmetric quasi-periodic behaviors, and a chaotic dynamics. The vortex shedding features multiple coupled oscillator modes, including in-phase, anti-phase, and out-of-phase synchronization and non-synchronization. These shedding regimes are analyzed employing the temporal evolution of the aerodynamic forces and a dynamical mode decomposition of the wake flow. The kaleidoscope of unforced and forced dynamics promotes the fluidic pinball as a challenging modeling and control benchmark.

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

Similar content being viewed by others

References

  1. Williamson, C.H.: Vortex dynamics in the cylinder wake. Annu. Rev. Fluid Mech. 28, 477–539 (1996)

    Article  MathSciNet  Google Scholar 

  2. Mittal, R., Balachandar, S.: Effect of three-dimensionality on the lift and drag of nominally two-dimensional cylinders. Phys. Fluids 7, 1841–1865 (1995)

    Article  Google Scholar 

  3. Cox, J.S., Brentner, K.S., Rumsey, C.L.: Computation of vortex shedding and radiated sound for a circular cylinder: subcritical to transcritical Reynolds numbers. Theoret. Comput. Fluid Dyn. 12, 233–253 (1998)

    Article  Google Scholar 

  4. Roshko, A.: Experiments on the flow past a circular cylinder at very high Reynolds number. J. Fluid Mech. 10, 345–356 (1961)

    Article  Google Scholar 

  5. Noack, B.R., Eckelmann, H.: A global stability analysis of the steady and periodic cylinder wake. J. Fluid Mech. 270, 297–330 (1994)

    Article  Google Scholar 

  6. Canuto, D., Taira, K.: Two-dimensional compressible viscous flow around a circular cylinder. J. Fluid Mech. 785, 349–371 (2015)

    Article  MathSciNet  Google Scholar 

  7. Sarpkaya, T.: A critical review of the intrinsic nature of vortex-induced vibrations. J. Fluids Struct. 19, 389–447 (2004)

    Article  Google Scholar 

  8. Marquet, O., Sipp, D., Jacquin, L.: Sensitivity analysis and passive control of cylinder flow. J. Fluid Mech. 615, 221–252 (2008)

    Article  MathSciNet  Google Scholar 

  9. Felippa, C.A., Park, K.-C., Farhat, C.: Partitioned analysis of coupled mechanical systems. Comput. Methods Appl. Mech. Eng. 190, 3247–3270 (2001)

    Article  Google Scholar 

  10. Derakhshandeh, J., Alam, M.M.: A review of bluff body wakes. Ocean Eng. 182, 475–488 (2019)

    Article  Google Scholar 

  11. Crawford, J.D., Knobloch, E.: Symmetry and symmetry-breaking bifurcations in fluid dynamics. Annu. Rev. Fluid Mech. 23, 341–387 (1991)

    Article  MathSciNet  Google Scholar 

  12. Alam, M.M., Meyer, J.P.: Two interacting cylinders in cross flow. Phys. Rev. E 84, 056304 (2011)

    Article  Google Scholar 

  13. Chen, W., Ji, C., Xu, W., Liu, S., Campbell, J.: Response and wake patterns of two side-by-side elastically supported circular cylinders in uniform laminar cross-flow. J. Fluids Struct. 55, 218–236 (2015)

    Article  Google Scholar 

  14. Williamson, C.H.: Evolution of a single wake behind a pair of bluff bodies. J. Fluid Mech. 159, 1–18 (1985)

    Article  Google Scholar 

  15. Bearman, P., Wadcock, A.: The interaction between a pair of circular cylinders normal to a stream. J. Fluid Mech. 61, 499–511 (1973)

    Article  Google Scholar 

  16. Sumner, D.: Two circular cylinders in cross-flow: a review. J. Fluids Struct. 26, 849–899 (2010)

    Article  Google Scholar 

  17. Carini, M., Giannetti, F., Auteri, F.: On the origin of the flip-flop instability of two side-by-side cylinder wakes. J. Fluid Mech. 742, 552–576 (2014)

    Article  MathSciNet  Google Scholar 

  18. Cantwell, C.D., et al.: Nektar++: an open-source spectral/hp element framework. Comput. Phys. Commun. 192, 205–219 (2015)

    Article  Google Scholar 

  19. Chen, W., Ji, C., Xu, D.: Vortex-induced vibrations of two side-by-side circular cylinders with two degrees of freedom in laminar cross-flow. Comput. Fluids 193, 104288 (2019)

    Article  MathSciNet  Google Scholar 

  20. Deng, N., Noack, B.R., Morzyński, M., Pastur, L.R.: Low-order model for successive bifurcations of the fluidic pinball. J. Fluid Mech. 884, A37 (2020)

    Article  MathSciNet  Google Scholar 

  21. Chen, W., Ji, C., Alam, M.M., Williams, J., Xu, D.: Numerical simulations of flow past three circular cylinders in equilateral-triangular arrangements. J. Fluid Mech. 891, A14 (2020)

    Article  MathSciNet  Google Scholar 

  22. Deng, N., Noack, B.R., Morzyński, M., Pastur, L.R.: Galerkin force model for transient and post-transient dynamics of the fluidic pinball. J. Fluid Mech. 918, A4 (2021)

    Article  MathSciNet  Google Scholar 

  23. Deng, N., Noack, B.R., Morzyński, M., Pastur, L.R.: Cluster-based hierarchical network model of the fluidic pinball-cartographing transient and post-transient, multi-frequency, multi-attractor behaviour. J. Fluid Mech. 934, A24 (2022)

    Article  MathSciNet  Google Scholar 

  24. Güemes, A., Sanmiguel Vila, C., Discetti, S.: Super-resolution generative adversarial networks of randomly-seeded fields. Nat. Mach. Intell. 4, 1165–1173 (2022)

    Article  Google Scholar 

  25. Li, S., Li, W., Noack, B.R.: Machine-learned control-oriented flow estimation for multi-actuator multi-sensor systems exemplified for the fluidic pinball. J. Fluid Mech. 952, A36 (2022)

    Article  Google Scholar 

  26. Farzamnik, E., et al.: From snapshots to manifolds-a tale of shear flows. J. Fluid Mech. 955, A34 (2023)

    Article  MathSciNet  Google Scholar 

  27. Peitz, S., Otto, S.E., Rowley, C.W.: Data-driven model predictive control using interpolated Koopman generators. SIAM J. Appl. Dyn. Syst. 19, 2162–2193 (2020)

    Article  MathSciNet  Google Scholar 

  28. Wang, X., Deng, N., Cornejo Maceda, G.Y., Noack, B.R.: Cluster-based control for net drag reduction of the fluidic pinball. Phys. Fluids 35, 023601 (2023)

    Article  Google Scholar 

  29. Cornejo Maceda, G.Y., Li, Y., Lusseyran, F., Morzyński, M., Noack, B.R.: Stabilization of the fluidic pinball with gradient-enriched machine learning control. J. Fluid Mech. 917, A42 (2021)

    Article  MathSciNet  Google Scholar 

  30. Raibaudo, C., Zhong, P., Noack, B.R., Martinuzzi, R.J.: Machine learning strategies applied to the control of a fluidic pinball. Phys. Fluids 32, 015108 (2020)

    Article  Google Scholar 

  31. Li, Y., et al.: Explorative gradient method for active drag reduction of the fluidic pinball and slanted ahmed body. J. Fluid Mech. 932, A7 (2022)

    Article  MathSciNet  Google Scholar 

  32. Feng, H., Wang, Y., Xiang, H., Jin, Z., Fan, D.: How to control hydrodynamic force on fluidic pinball via deep reinforcement learning. Phys. Fluids 35, 045157 (2023)

    Google Scholar 

  33. Pino, F., Schena, L., Rabault, J., Mendez, M.A.: Comparative analysis of machine learning methods for active flow control. J. Fluid Mech. 958, A39 (2023)

    Article  MathSciNet  Google Scholar 

  34. Benner, P., Gugercin, S., Willcox, K.: A survey of projection-based model reduction methods for parametric dynamical systems. SIAM Rev. 57, 483–531 (2015)

    Article  MathSciNet  Google Scholar 

  35. Brunton, S.L., Noack, B.R., Koumoutsakos, P.: Machine learning for fluid mechanics. Annu. Rev. Fluid Mech. 52, 477–508 (2020)

    Article  MathSciNet  Google Scholar 

  36. Fukami, K., Fukagata, K., Taira, K.: Assessment of supervised machine learning methods for fluid flows. Theoret. Comput. Fluid Dyn. 34, 497–519 (2020)

    Article  MathSciNet  Google Scholar 

  37. Ratha, N.K., Chen, S., Jain, A.K.: Adaptive flow orientation-based feature extraction in fingerprint images. Pattern Recogn. 28, 1657–1672 (1995)

    Article  Google Scholar 

  38. Jain, A.K., Duin, R.P.W., Mao, J.: Statistical pattern recognition: a review. IEEE Trans. Pattern Anal. Mach. Intell. 22, 4–37 (2000)

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  40. Taira, K., et al.: Modal analysis of fluid flows: an overview. AIAA J. 55, 4013–4041 (2017)

    Article  Google Scholar 

  41. Carlberg, K., Farhat, C., Cortial, J., Amsallem, D.: The gnat method for nonlinear model reduction: effective implementation and application to computational fluid dynamics and turbulent flows. J. Comput. Phys. 242, 623–647 (2013)

    Article  MathSciNet  Google Scholar 

  42. Amsallem, D., Zahr, M.J., Farhat, C.: Nonlinear model order reduction based on local reduced-order bases. Int. J. Numer. Meth. Eng. 92, 891–916 (2012)

    Article  MathSciNet  Google Scholar 

  43. Benner, P., Ohlberger, M., Cohen, A., Willcox, K.: Model Reduction and Approximation: Theory and Algorithms. SIAM, New Delhi (2017)

    Book  Google Scholar 

  44. Baur, U., Benner, P., Feng, L.: Model order reduction for linear and nonlinear systems: a system-theoretic perspective. Arch. Comput. Methods Eng. 21, 331–358 (2014)

    Article  MathSciNet  Google Scholar 

  45. Holmes, P.: Turbulence, Coherent Structures, Dynamical Systems and Symmetry. Cambridge University Press, Cambridge (2012)

    Book  Google Scholar 

  46. Schlatter, P., Bagheri, S., Henningson, D.S.: Self-sustained global oscillations in a jet in crossflow. Theoret. Comput. Fluid Dyn. 25, 129–146 (2011)

    Article  Google Scholar 

  47. Willcox, K.: Unsteady flow sensing and estimation via the gappy proper orthogonal decomposition. Comput. Fluids 35, 208–226 (2006)

    Article  Google Scholar 

  48. Alomar, A., Nicole, A., Sipp, D., Rialland, V., Vuillot, F.: Reduced-order model of a reacting, turbulent supersonic jet based on proper orthogonal decomposition. Theoret. Comput. Fluid Dyn. 34, 49–77 (2020)

    Article  MathSciNet  Google Scholar 

  49. Karhunen, K.: Zur Spektraltheorie stochastischer Prozesse. Ann. Acad. Sci. Fennicae, AI 34, 90 (1946)

    MathSciNet  Google Scholar 

  50. 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. Theoret. Comput. Fluid Dyn. 31, 349–368 (2017)

    Article  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  52. Kutz, J.N., Brunton, S.L., Brunton, B.W., Proctor, J.L.: Dynamic Mode Decomposition: Data-Driven Modeling of Complex Systems. SIAM, New Delhi (2016)

    Book  Google Scholar 

  53. Lumley, J.L.: The structure of inhomogeneous turbulent flows. Atmos. Turbul. Radio Wave Propag. 166–178 (1967)

  54. Mezić, I.: Analysis of fluid flows via spectral properties of the Koopman operator. Annu. Rev. Fluid Mech. 45, 357–378 (2013)

    Article  MathSciNet  Google Scholar 

  55. Williams, M.O., Kevrekidis, I.G., Rowley, C.W.: A data-driven approximation of the Koopman operator: extending dynamic mode decomposition. J. Nonlinear Sci. 25, 1307–1346 (2015)

    Article  MathSciNet  Google Scholar 

  56. Schmid, P.J., Li, L., Juniper, M.P., Pust, O.: Applications of the dynamic mode decomposition. Theoret. Comput. Fluid Dyn. 25, 249–259 (2011)

    Article  Google Scholar 

  57. Tu, J.H.: Dynamic mode decomposition: theory and applications. Ph.D. thesis, Princeton University (2013)

  58. Kou, J., Zhang, W.: An improved criterion to select dominant modes from dynamic mode decomposition. Eur. J. Mech. B/Fluids 62, 109–129 (2017)

    Article  MathSciNet  Google Scholar 

  59. Liu, B., Jaiman, R.: Dynamics and stability of gap-flow interference in a vibrating side-by-side arrangement of two circular cylinders. J. Fluid Mech. 855, 804–838 (2018)

    Article  MathSciNet  Google Scholar 

  60. Yan, Y., Ji, C., Srinil, N.: On wake modulation and interaction features of a pair of dual-step circular cylinders in side-by-side arrangements. Phys. Fluids 33, 093604 (2021)

    Article  Google Scholar 

  61. Zhang, H.-Q., Fey, U., Noack, B.R., König, M., Eckelmann, H.: On the transition of the cylinder wake. Phys. Fluids 7, 779–795 (1995)

    Article  Google Scholar 

  62. Koda, Y., Lien, F.-S.: Aerodynamic effects of the early three-dimensional instabilities in the flow over one and two circular cylinders in tandem predicted by the lattice boltzmann method. Comput. Fluids 74, 32–43 (2013)

    Article  MathSciNet  Google Scholar 

  63. Carmo, B.S., Assi, G.R., Meneghini, J.R.: Computational simulation of the flow-induced vibration of a circular cylinder subjected to wake interference. J. Fluids Struct. 41, 99–108 (2013)

    Article  Google Scholar 

  64. Rowley, C.W., Mezić, I., Bagheri, S., Schlatter, P., Henningson, D.S.: Spectral analysis of nonlinear flows. J. Fluid Mech. 641, 115–127 (2009)

    Article  MathSciNet  Google Scholar 

  65. Noack, B.R., Stankiewicz, W., Morzyński, M., Schmid, P.J.: Recursive dynamic mode decomposition of transient and post-transient wake flows. J. Fluid Mech. 809, 843–872 (2016)

    Article  MathSciNet  Google Scholar 

  66. Brunton, S.L., Proctor, J.L., Tu, J.H., Kutz, J.N.: Compressed sensing and dynamic mode decomposition. J. Comput. Dyn. 2, 165–191 (2016)

    Article  MathSciNet  Google Scholar 

  67. Kang, S.: Characteristics of flow over two circular cylinders in a side-by-side arrangement at low Reynolds numbers. Phys. Fluids 15, 2486–2498 (2003)

    Article  Google Scholar 

  68. Noack, B.R., Afanasiev, K., MORZY\(\acute{{\rm N}}\)SKI, M., Tadmor, G., Thiele, F.: A hierarchy of low-dimensional models for the transient and post-transient cylinder wake. J. Fluid Mech. 497, 335–363 (2003)

  69. Um, E., et al.: Phase synchronization of fluid-fluid interfaces as hydrodynamically coupled oscillators. Nat. Commun. 11, 5221 (2020)

    Article  Google Scholar 

  70. Bui-Thanh, T., Damodaran, M., Willcox, K.: Aerodynamic data reconstruction and inverse design using proper orthogonal decomposition. AIAA J. 42, 1505–1516 (2004)

    Article  Google Scholar 

Download references

Acknowledgements

We appreciate the support from Marek Morzyński for the numerical simulations, and valuable discussions with Md. Mahbub Alam, Stefano Discetti, Chang Hou, Andrea Ianiro, Chunning Ji, François Lusseyran, Guy Y. Cornejo Maceda, Luc R. Pastur, and the HIT-Hanghua team.

Funding

This work is supported by the National Natural Science Foundation of China under Grants 12172109, 12172111, and 12202121, by the China Postdoctoral Science Foundation under grants 2023M730866 and 2023T160166, by the Guangdong Basic and Applied Basic Research Foundation under grant 2022A1515011492, and by the Shenzhen Science and Technology Program under Grant JCYJ20220531095605012.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Nan Deng or Bernd R. Noack.

Ethics declarations

Author contributions

BRN conducted the project management and funding acquisition. DN and BRN conceived the project, designed the research, and supervised the project process. WX provided numerical support. LY and DN performed the research, data curation, and formal analysis. LY, DN, and BRN discussed and investigated the results. L.Y. prepared the validation and visualization. LY and DN wrote the original draft. WX, DN, and BRN reviewed the original draft. DN and BRN edited the manuscript.

Data availability

The data that support the findings of this study are available from the corresponding authors on request.

Code availability

The code that supports the findings of this study is available from the corresponding authors on request.

Conflict of interest

The authors declare that there is no conflict of interest regarding the publication of this paper.

Ethical approval

Not applicable.

Consent to participate

Not applicable.

Consent for publication

Not applicable.

Additional information

Publisher's Note

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

Appendices

Transient and post-transient dynamics

In this work, we start numerical simulations from a steady solution with \(\text {Re}=100\) and activate the actuation after \(t=5\). This symmetric solution is one of three unstable steady solutions obtained by solving the steady Navier–Stokes equations [20]. As shown in Fig. 14, the unforced flow takes a long time to reach a stable post-transient regime, starting from the steady solution with a tiny initial perturbation. Before that, the transient flow successively undergoes the destabilization around the steady solution and the transition from the steady solution to the attracting set, associated with the nonlinear saturation of the instability of the system.

Fig. 14
figure 14

Transient and post-transient dynamics of the unforced fluidic pinball at \(\text {Re}=100\), illustrated by the temporal evolution of the resultant lift coefficient and representative snapshots. It has been divided into four regimes: (I) the steady solution, (II) the destabilization around the steady solution, (III) the transition from the steady solution to the attracting set, and (IV) the post-transient regime

Proper orthogonal decomposition

The flow is equidistantly sampled with a time interval \(\Delta t\), yielding M velocity field snapshots \(\varvec{u}^m (\varvec{x}):= \varvec{u}(\varvec{x}, t^m)\) at \(t^m = m \Delta t, m = 1, \ldots , M\). The velocity field is decomposed into a base flow \(\varvec{u}_0\) and the fluctuating contribution \(\varvec{v}^m\),

$$\begin{aligned} \varvec{u}^m = \varvec{u}_0 + \varvec{v}^m, \end{aligned}$$
(B1)

where the base flow is defined as the time-averaged field in this work:

$$\begin{aligned} \varvec{u}_0 = \frac{1}{M} \sum _{m=1}^{M} \varvec{u}^m. \end{aligned}$$
(B2)

The inner product of two velocity fields \(\varvec{v}\), \(\varvec{w}\) in the observation domain \(\Omega \) reads

$$\begin{aligned} \left( \varvec{v}, \varvec{w}\right) _\Omega = \int _\Omega d\varvec{x}\, \varvec{v}(\varvec{x}) \cdot \varvec{w}(\varvec{x}) . \end{aligned}$$
(B3)

The POD modes are computed with the snapshot method, i.e. require the correlation matrix \(\textbf{C}\) of the fluctuating contribution. The elements of \(\textbf{C}\) read

$$\begin{aligned} C^{mn} = \frac{1}{M} (\varvec{v}^m, \varvec{v}^n)_\Omega , \quad m,n = 1, \ldots , M. \end{aligned}$$
(B4)

The spectral decomposition of the positive semi-definite matrix \(\textbf{C}\) yields non-negative eigenvalues \(\lambda _i\), \(i=1,\ldots ,M\) and orthogonal eigenvectors \(\varvec{e}_i\),

$$\begin{aligned} \textbf{C} \varvec{e}_i = \lambda _i \varvec{e}_i, \quad i = 1, \ldots , M. \end{aligned}$$
(B5)

The ith normalized eigenvector has M components \(\varvec{e}_i = \left[ e^1_i, e^2_i, \cdots , e^M_i\right] ^\intercal \). The ith POD mode is a linear combination of the fluctuations,

$$\begin{aligned} \varvec{u}_i = \frac{1}{\sqrt{M \lambda _i}} \sum _{m=1}^{M} e^m_i \varvec{v}^m, \end{aligned}$$
(B6)

and the corresponding temporal coefficients are rescaled eigenvectors of the correlation matrix,

$$\begin{aligned} a_i^m = {\sqrt{M \lambda _i}} \, e_i^m. \end{aligned}$$
(B7)

In the end, the fluctuating contribution is represented as the sum of spatial modes weighted by temporal coefficients. These modes are ranked according to their energetic contribution. A low-dimensional approximation \(\varvec{\tilde{v}}^m\) of the fluctuating part \(\varvec{{v}}^m\) reads

$$\begin{aligned} \varvec{\tilde{v}}^m = \sum _{i=1}^{N} a_i^m\, \varvec{u}_i, \end{aligned}$$
(B8)

where \(N \le M \) is the selected number of leading POD modes.

The POD modes compress the velocity snapshots to N dimensions and minimize the representation error. As in Eqs. (B1) and (B8), the POD modes and the corresponding temporal coefficients allow for the linear reconstruction [70] of the flow field at any instant.

Dynamic mode decomposition

The state snapshot matrices \(\textbf{Y}\) and \(\textbf{Y}^\prime \) read,

$$\begin{aligned} \textbf{Y}= & {} \left[ \varvec{a}^1, \varvec{a}^2, \cdots , \varvec{a}^{M-1} \right] , \end{aligned}$$
(C9a)
$$\begin{aligned} \textbf{Y}^\prime= & {} \left[ \varvec{a}^2, \varvec{a}^3, \cdots , \varvec{a}^{M} \right] , \end{aligned}$$
(C9b)

where the snapshot matrices are of size \(N \times (M-1)\).

We assume that the time step \(\Delta t\) is small enough for a linear discrete dynamics where \(\varvec{a}^{m+1}\) can be obtained from the current state \(\varvec{a}^m\) by a full-state mapping matrix \(\textbf{A} \in \mathcal {R}^{N \times N}\),

$$\begin{aligned} \varvec{a}^{m+1} = \textbf{A} \> \varvec{a}^m. \end{aligned}$$
(C10)

We also assume that the matrix \(\textbf{A}\) does not change with increasing time. The latter assumption is legitimate for the considered post-transient regimes where DMD approximates a Fourier analysis. Such time-invariant linear dynamics are clearly invalid for transients from steady to unsteady solutions [65].

Equation (C10) implies for Eq. (C9)

$$\begin{aligned} \textbf{Y}^\prime = \left[ \textbf{A} \varvec{a}^1, \textbf{A} \varvec{a}^2, \cdots , \textbf{A} \varvec{a}^{M}\right] = \textbf{A} \textbf{Y}. \end{aligned}$$
(C11)

When the time step is small enough, the system can satisfy this linear approximation.

We can optionally find a low-rank \(r < M-1\) projection of the full-state \(\textbf{A}\) by singular value decomposition,

$$\begin{aligned} \textbf{Y} = \textbf{U} \varvec{\Sigma } \textbf{V}^\intercal , \end{aligned}$$
(C12)

where \(\textbf{U}\) and \(\textbf{V}\) are square unitary matrices of sizes \(N \times N\) and \((M-1) \times (M-1)\), the superscript ‘\(\intercal \)’ indicates the transpose, and \(\varvec{\Sigma }\) is an \(N \times (M-1)\) rectangular diagonal matrix. For the low-rank projection, we keep the first r columns of \(\textbf{U}\) and \(\textbf{V}\) and the first \(r \times r\) elements in \(\varvec{\Sigma }\), then the approximated matrix \(\widetilde{\textbf{A}}\) of the full-state \(\textbf{A}\) is

$$\begin{aligned} \widetilde{\textbf{A}} = \textbf{U}^\intercal \textbf{A} \textbf{U}, \end{aligned}$$
(C13)

which can be seen as the least-squares fit or minimum-norm solution of \(\textbf{A}\) for \(\textbf{Y}^\prime = \textbf{A} \textbf{Y}\),

$$\begin{aligned} \mathop {\text {min}}\limits _{\widetilde{\textbf{A}}} \Vert \textbf{Y}^\prime - \textbf{U} \widetilde{\textbf{A}} \varvec{\Sigma } \textbf{V}^\intercal \Vert _2 , \end{aligned}$$
(C14)

then, \(\widetilde{\textbf{A}}\) can be written as,

$$\begin{aligned} \widetilde{\textbf{A}} = \textbf{U}^\intercal \textbf{Y}^\prime \textbf{V} \Sigma ^{-1}. \end{aligned}$$
(C15)

The eigenvalues \(\varvec{\Lambda }\) and eigenvectors \(\textbf{W}\) are obtained by eigendecomposition of \(\widetilde{\textbf{A}}\):

$$\begin{aligned} \widetilde{\textbf{A}} \textbf{W} = \textbf{W} \varvec{\Lambda }. \end{aligned}$$
(C16)

The diagonal elements of \(\varvec{\Lambda }\) are Ritz eigenvalues \(\mu _j\), \(j = 1, \ldots , r\). The \(\mu _j\) are used to calculate the growth rate \(\sigma _j = \textrm{Re} (\ln \mu _j / \Delta t)\) and the oscillating frequency \(\omega _j = \textrm{Im} (\ln \mu _j / \Delta t)\). The DMD modes are the columns of the matrix \(\varvec{\Phi }\) of size \(N \times r\), which can be computed by,

$$\begin{aligned} \varvec{\Phi } = \textbf{Y}^\prime \textbf{V} \varvec{\Sigma }^{-1} \textbf{W}. \end{aligned}$$
(C17)

Finally, the spatial velocity field of DMD modes is reconstructed following Eq. (B8), by using the DMD mode as the temporal coefficients of the leading N POD spatial modes \(u_i\), which reads

$$\begin{aligned} \varvec{u}^\textrm{DMD}_j = \sum _{i=1}^{N} {\Phi }_{i, j}\, \varvec{u}_i . \end{aligned}$$
(C18)

Figure 15 depicts the eigenvalues obtained from the DMD analysis for varying values of b in the range \(\left[ -3, 2\right] \). The real and imaginary parts of the eigenvalues are represented along the horizontal and vertical axes, respectively. The eigenvalues corresponding to the vortex-shedding frequency are marked with red dots. Notably, for the chaotic and quasi-periodic cases, the eigenvalues corresponding to the base-bleeding jet frequency are marked with blue dots.

Fig. 15
figure 15

Eigenvalues for varying \(b \in \left[ -3, 2\right] \), with (a\(\sim \) (l) similar to Fig. 5. The real and imaginary parts of the eigenvalues are shown along the horizontal and vertical axes, respectively. The red points represent eigenvalues related to the vortex shedding, and the blue points correspond to the base-bleeding jet eigenvalues. The corresponding modes are plotted in Figs. 1112 and 13 (colour figure online)

As illustrated in Fig. 5, the power spectral density (PSD) curves of the lift coefficient exhibit a harmonic series of peaks. The dominant peak and the lowest peak align with the vortex-shedding frequency and the base-bleeding jet frequency, respectively. Further insight into the eigenvalue distribution is presented in Fig. 15, where the real and imaginary parts of the eigenvalues are plotted. The points falling outside the unit circle have positive growth rates, indicating divergence and instability. The points falling inside have negative growth rates, indicating convergence and stability, and the points on the circle have zero growth rates, indicating stability. Most of the points lie very close to the unit circle in the spectrums. The eigenvalues associated with the characteristic frequencies are marked with red points on the unit circle, calculated as \(\omega _j=2\pi f_j\). Each frequency is represented by two points on the DMD modal eigenvalue distribution, sharing the same real part and opposite imaginary parts. The eigenvalues of the extracted modes cluster around the unit circle, indicating their stability. In particular, for a higher rotation speed \(\left| b \right| \), more points fall within the unit circle, signifying faster convergence to a stable state.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, Y., Deng, N., Noack, B.R. et al. The fluidic pinball with symmetric forcing displays steady, periodic, quasi-periodic, and chaotic dynamics. Theor. Comput. Fluid Dyn. (2024). https://doi.org/10.1007/s00162-024-00693-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00162-024-00693-2

Keywords

Navigation