Skip to main content
Log in

Data-driven modeling of the chaotic thermal convection in an annular thermosyphon

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

Abstract

Identifying accurate and yet interpretable low-order models from data has gained a renewed interest over the past decade. In the present work, we illustrate how the combined use of dimensionality reduction and sparse system identification techniques allows us to obtain an accurate model of the chaotic thermal convection in a two-dimensional annular thermosyphon. Taking as guidelines the derivation of the Lorenz system, the chaotic thermal convection dynamics simulated using a high-fidelity computational fluid dynamics solver are first embedded into a low-dimensional space using dynamic mode decomposition. After having reviewed the physical properties the reduced-order model should exhibit, the latter is identified using SINDy, an increasingly popular and flexible framework for the identification of nonlinear continuous-time dynamical systems from data. The identified model closely resembles the canonical Lorenz system, having the same structure and exhibiting the same physical properties. It moreover accurately predicts a bifurcation of the high-dimensional system (corresponding to the onset of steady convection cells) occurring at a much lower Rayleigh number than the one considered in this study.

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

Similar content being viewed by others

Notes

  1. A video of these dynamics is available online at https://tinyurl.com/y55buvnc.

References

  1. Holmes, P., Lumley, J.L., Berkooz, G.: Turbulence, Coherent Structures, Dynamical Systems and Symmetry. Cambridge University Press, Cambridge (1996). https://doi.org/10.1017/CBO9780511622700

    Book  MATH  Google Scholar 

  2. Fabbiane, N., Semeraro, O., Bagheri, S., Henningson, D.S.: Adaptive and model-based control theory applied to convectively unstable flows. Appl. Mech. Rev. (2014). https://doi.org/10.1115/1.4027483

    Article  Google Scholar 

  3. Brunton, S.L., Noack, B.R.: Closed-loop turbulence control: progress and challenges. Appl. Mech. Rev. 67(5), 050801 (2015). https://doi.org/10.1115/1.4031175

    Article  Google Scholar 

  4. Sipp, D., Schmid, P.J.: Linear closed-loop control of fluid instabilities and noise-induced perturbations: a review of approaches and tools. Appl. Mech. Rev. 68(2), 020801 (2016). https://doi.org/10.1115/1.4033345

    Article  Google Scholar 

  5. Rowley, C.W., Dawson, S.T.M.: Model reduction for flow analysis and control. Annu. Rev. Fluid Mech. 49(1), 387 (2017). https://doi.org/10.1146/annurev-fluid-010816-060042

    Article  MathSciNet  MATH  Google Scholar 

  6. Taira, K., Brunton, S.L., Dawson, S.T.M., Rowley, C.W., Colonius, T., McKeon, B.J., Schmidt, O.T., Gordeyev, S., Theofilis, V., Ukeiley, L.S.: Modal analysis of fluid flows: an overview. AIAA J. 55(12), 4013 (2017). https://doi.org/10.2514/1.J056060

    Article  Google Scholar 

  7. Lorenz, E.N.: Deterministic nonperiodic flow. J. Atmos. Sci. 20(2), 130 (1963). https://doi.org/10.1175/1520-0469(1963)020<0130:DNF>2.0.CO;2

    Article  MathSciNet  MATH  Google Scholar 

  8. Sirovich, L.: Turbulence and the dynamics of coherent structures. I. Coherent structures. II. Symmetries and transformations. III. Dynamics and scaling. Q. Appl. Math. 45, 561 (1987). https://doi.org/10.1090/qam/910462

    Article  MATH  Google Scholar 

  9. Berkooz, G., Holmes, P., Lumley, J.L.: The proper orthogonal decomposition in the analysis of turbulent flows. Annu. Rev. Fluid Mech. 25(1), 539 (1993). https://doi.org/10.1146/annurev.fl.25.010193.002543

    Article  MathSciNet  Google Scholar 

  10. Noack, B.R., Afanasiev, K., Morzyński, M., Tadmor, G., Thiele, F.: A hierarchy of low-dimensional models for the transient and post-transient cylinder wake. J. Fluid Mech. 497, 335 (2003). https://doi.org/10.1017/s0022112003006694

    Article  MathSciNet  MATH  Google Scholar 

  11. Juang, J.N., Pappa, R.S.: An eigensystem realization algorithm for modal parameter identification and model reduction. J. Guidance 8(5), 620 (1985). https://doi.org/10.2514/3.20031

    Article  MATH  Google Scholar 

  12. Rowley, C.W., Mezić, I., Bagheri, S., Schlatter, P.: Spectral analysis of nonlinear flows. J. Fluid Mech. 641, 115 (2009). https://doi.org/10.1017/S0022112009992059

    Article  MathSciNet  MATH  Google Scholar 

  13. Schmid, P.J.: Dynamic mode decomposition of numerical and experimental data. J. Fluid Mech. 656, 5 (2010). https://doi.org/10.1017/S0022112010001217

    Article  MathSciNet  MATH  Google Scholar 

  14. Tu, J.H., Rowley, C.W., Luchtenburg, D.M., Brunton, S.L., Kutz, J.N.: On dynamic mode decomposition: theory and applications. J. Comput. Dyn. 1(2), 391 (2014). https://doi.org/10.3934/jcd.2014.1.391

    Article  MathSciNet  MATH  Google Scholar 

  15. Billings, S.A.: Nonlinear system identification: NARMAX methods in the time, frequency, and spatio-temporal domains (Paperbackshop UK import (2013). https://www.ebook.de/de/product/20764167/stephen_a_billings_nonlinear_system_identification_narmax_methods_in_the_time_frequency_and_spatio_temporal_domains.html

  16. Bongard, J., Lipson, H.: Automated reverse engineering of nonlinear dynamical systems. Proc. Natl. Acad. Sci. USA 104(24), 9943 (2007). https://doi.org/10.1073/pnas.0609476104

    Article  MATH  Google Scholar 

  17. Schmidt, M., Lipson, H.: Distilling free-form natural laws from experimental data. Science 324(5923), 81 (2009). https://doi.org/10.1126/science.1165893

    Article  Google Scholar 

  18. Brunton, S.L., Proctor, J.L., Kutz, J.N.: Discovering governing equations from data by sparse identification of nonlinear dynamical systems. Proc. Natl. Acad. Sci. USA 113(15), 3932 (2016). https://doi.org/10.1073/pnas.1517384113

    Article  MathSciNet  MATH  Google Scholar 

  19. Takens, F.: in Lecture Notes in Mathematics, pp. 366–381. Springer, Berlin (1981). https://doi.org/10.1007/BFb0091924

    Book  Google Scholar 

  20. Broomhead, D.S., King, G.P.: Extracting qualitative dynamics from experimental data. Physica D 20(2–3), 217 (1986). https://doi.org/10.1016/0167-2789(86)90031-x

    Article  MathSciNet  MATH  Google Scholar 

  21. Yorke, J.A., Yorke, E.D., Mallet-Paret, J.: Lorenz-like chaos in a partial differential equation for a heated fluid loop. Physica D 24(1–3), 279 (1987)

    Article  MathSciNet  Google Scholar 

  22. Ehrhard, P., Müller, U.: Dynamical behaviour of natural convection in a single-phase loop. J. Fluid Mech. 217, 487 (1990)

    Article  Google Scholar 

  23. Louisos, W.F., Hitt, D.L., Danforth, C.M.: Chaotic natural convection in a toroidal thermosyphon with heat flux boundaries. Int. J. Heat Mass Transf. 88, 492 (2015)

    Article  Google Scholar 

  24. Anishchenko, V.S., Vadivasova, T.E., Okrokvertskhov, G.A., Strelkova, G.I.: Correlation analysis of dynamical chaos. Physica A 325(1–2), 199 (2003)

    Article  MathSciNet  Google Scholar 

  25. Saltzman, B.: Finite amplitude free convection as an initial value problem I. J. Atmos. Sci. 19(4), 329 (1962)

    Article  Google Scholar 

  26. Jovanović, M.R., Schmid, P.J., Nichols, J.W.: Sparsity-promoting dynamic mode decomposition. Phys. Fluids 26(2), 024103 (2014)

    Article  Google Scholar 

  27. 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(6), 1307 (2015). https://doi.org/10.1007/s00332-015-9258-5

    Article  MathSciNet  MATH  Google Scholar 

  28. Kutz, J.N., Brunton, S.L., Brunton, B.W., Proctor, J.L.: Dynamic Mode Decomposition: Data-Driven Modeling of Complex Systems. SIAM-Society for Industrial and Applied Mathematics, New York (2016). https://www.amazon.com/Dynamic-Mode-Decomposition-Data-Driven-Modeling/dp/1611974496?SubscriptionId=AKIAIOBINVZYXZQZ2U3A&tag=chimbori05-20&linkCode=xm2&camp=2025&creative=165953&creativeASIN=1611974496

  29. 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 (2016). https://doi.org/10.1007/s00348-016-2127-7

    Article  Google Scholar 

  30. 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 (2016). https://doi.org/10.1017/jfm.2016.678

    Article  MathSciNet  MATH  Google Scholar 

  31. Le Clainche, S., Vega, J.M.: Higher order dynamic mode decomposition. SIAM J. Appl. Dyn. Syst. 16(2), 882 (2017). https://doi.org/10.1137/15m1054924

    Article  MathSciNet  MATH  Google Scholar 

  32. Hirsh, S.M., Harris, K.D., Kutz, J.N., Brunton, B.W.: Centering data improves the dynamic mode decomposition (2019). arXiv:1906.05973

  33. Koopman, B.O.: Hamiltonian systems and transformation in Hilbert space. Proc. Natl. Acad. Sci. USA 17(5), 315 (1931)

    Article  Google Scholar 

  34. Izenman, A.J.: Reduced-rank regression for the multivariate linear model. J. Multivar. Anal. 5(2), 248 (1975)

    Article  MathSciNet  Google Scholar 

  35. De la Torre, F.: A least-squares framework for component analysis. IEEE Trans. Pattern Anal. Mach. Intell. 34(6), 1041 (2012)

    Article  Google Scholar 

  36. Héas, P., Herzet, C.: Low-rank dynamic mode decomposition: optimal solution in polynomial-time. e-print arXiv:1610.02962 (2016)

  37. Tegmark, M.: Latent representations of dynamical systems: when two is better than one. e-print arXiv:1902.03364 (2019)

  38. Loiseau, J.C., Brunton, S.L.: Constrained sparse Galerkin regression. J. Fluid Mech. 838, 42 (2018). https://doi.org/10.1017/jfm.2017.823

    Article  MathSciNet  MATH  Google Scholar 

  39. Tibshirani, R.: Regression shrinkage and selection via the lasso. J. R. Stat. Soc. Ser. B (Methodol.) 73, 267 (1996). https://doi.org/10.1111/j.1467-9868.2011.00771.x

    Article  MathSciNet  MATH  Google Scholar 

  40. Mangan, N.M., Brunton, S.L., Proctor, J.L., Kutz, J.N., Trans, I.E.E.E.: Inferring biological networks by sparse identification of nonlinear dynamics. Mol. Biol. Multi Scale Commun. 2(1), 52 (2016). https://doi.org/10.1109/tmbmc.2016.2633265

    Article  Google Scholar 

  41. Brunton, S.L., Proctor, J.L., Kutz, J.N.: Sparse identification of nonlinear dynamics with control (SINDYc). IFAC Pap. Online 49(18), 710 (2016). https://doi.org/10.1016/j.ifacol.2016.10.249

    Article  Google Scholar 

  42. Brunton, S.L., Brunton, B.W., Proctor, J.L., Kaiser, E., Kutz, J.N.: Chaos as an intermittently forced linear system. Nat. Commun. 8, 1 (2017). https://doi.org/10.1038/s41467-017-00030-8

    Article  Google Scholar 

  43. Rudy, S.H., Brunton, S.L., Proctor, J.L., Kutz, J.N.: Data-driven discovery of partial differential equations. Sci. Adv. 3(4), e1602614 (2017). https://doi.org/10.1126/sciadv.1602614

    Article  Google Scholar 

  44. Kaiser, E., Kutz, J.N., Brunton, S.L.: Sparse identification of nonlinear dynamics for model predictive control in the low-data limit. Proc. R. Soc. A 474(2219), 20180335 (2018)

    Article  MathSciNet  Google Scholar 

  45. Loiseau, J.C., Noack, B.R., Brunton, S.L.: Sparse reduced-order modelling: sensor-based dynamics to full-state estimation. J. Fluid Mech. 844, 459 (2018). https://doi.org/10.1017/jfm.2018.147

    Article  MATH  Google Scholar 

  46. John Soundar, J., Chomaz, J.M., Huerre, P.: Transient growth in Rayleigh–Bénard–Poiseuille/Couette convection. Phys. Fluids 24(4), 044103 (2012)

    Article  Google Scholar 

  47. Podvin, B., Sergent, A.: Proper orthogonal decomposition investigation of turbulent Rayleigh–Bénard convection in a rectangular cavity. Phys. Fluids 24(10), 105106 (2012)

    Article  Google Scholar 

  48. Castillo-Castellanos, A., Sergent, A., Podvin, B., Rossi, M.: Cessation and reversals of large-scale structures in square Rayleigh–Bénard cells. J. Fluid Mech. 877, 922–954 (2019). https://doi.org/10.1017/jfm.2019.598

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  50. Manohar, K., Brunton, B.W., Kutz, J.N., Brunton, S.L.: Data-driven sparse sensor placement. IEEE Control Syst. Mag. 38(3), 63 (2018). https://doi.org/10.1109/MCS.2018.2810460

    Article  Google Scholar 

  51. Clark, E., Askham, T., Brunton, S.L., Nathan Kutz, J.: Greedy sensor placement with cost constraints. IEEE Sens. J. 19(7), 2642 (2019). https://doi.org/10.1109/JSEN.2018.2887044

    Article  Google Scholar 

  52. Kennel, M.B., Brown, R., Abarbanel, H.D.I.: Determining embedding dimension for phase-space reconstruction using a geometrical construction. Phys. Rev. A 45, 3403 (1992). https://doi.org/10.1103/PhysRevA.45.3403

    Article  Google Scholar 

  53. Cao, L.: Dynamics from multivariate time series. Phys. D Nonlinear Phenomena 110(1), 43 (1997). https://doi.org/10.1016/S0167-2789(97)00118-8

    Article  MathSciNet  Google Scholar 

  54. Krakovská, A., Mezeiová, K., Budáčová, H.: Use of false nearest neighbours for selecting variables and embedding parameters for state space reconstruction. J. Complex Syst. 2015, 12 (2015)

    Google Scholar 

  55. Kamb, M., Kaiser, E., Brunton, S.L., Kutz, J.N.: Time-delay observables for Koopman: theory and applications. Preprint arXiv:1810.01479 (2018)

Download references

Acknowledgements

This work is adapted from a contribution to the long program Machine Learning for Physics and the Physics of Learning, organized by the Institute for Pure and Applied Mathematics (IPAM) at UCLA (Los Angeles, USA) in 2019 and made possible thanks to the financial support of IPAM, which is supported by the National Science Foundation (NSF Grant No. DMS-1440415). The author would also like to gratefully thank Steven Brunton, Jared Callahan, Kathleen Champion, Onofrio Semeraro, Alessandro Bucci and many others for the fruitful dicussion on this topic.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jean-Christophe Loiseau.

Additional information

Communicated by Steven L. Brunton.

Publisher's Note

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

Appendices

Derivation of the optimal solution to the DMD problem

1.1 Equivalence between norm-minimization and trace-maximization formulation of the DMD problem

This section aims at deriving the equivalence between the norm-minimization and trace-maximization formulations of the DMD problem and the corresponding solution. Given two data sequences \( {\varvec{X}} \) and \( {\varvec{Y}} \) related by

$$\begin{aligned} {\varvec{y}}_k = {\varvec{f}}({\varvec{x}}_k), \end{aligned}$$

the aim of DMD is to find the low-rank operator \( {\varvec{A}} \) that best approximates the possibly nonlinear function \( {\varvec{f}} \) in the least-squares sense. As discussed in Sect. 4.1, this can be formulated as the following rank-constrained minimization problem

$$\begin{aligned} \begin{aligned} {{\,\mathrm{minimize~}\,}}_{{\varvec{P}}, {\varvec{Q}}}&\quad \Vert {\varvec{Y}} - {\varvec{PQ}}^H {\varvec{X}} \Vert _F^2 \\ {{\,\mathrm{subject~to~}\,}}&\quad \text {rank } {\varvec{P}} = \text {rank } {\varvec{Q}} = r \\&\quad {\varvec{P}}^H {\varvec{P}} = {\varvec{I}}, \end{aligned} \end{aligned}$$
(4)

where we use the fact the unknown low-rank operator can be factorized as \( {\varvec{A}} = {\varvec{PQ}}^H \). Introducing the cost function

$$\begin{aligned} \begin{aligned} \mathcal {J}({\varvec{P}}, {\varvec{Q}})&= \Vert {\varvec{Y}} - {\varvec{PQ}}^H {\varvec{X}} \Vert _F^2 \\&= \text {Tr} \left( ({\varvec{Y}} - {\varvec{PQ}}^H {\varvec{X}})({\varvec{Y}} - {\varvec{PQ}}^H {\varvec{X}})^H \right) \end{aligned} \end{aligned}$$

for the sake of notational simplicity, the solution \( \left( {\varvec{P}}^*, {\varvec{Q}}^* \right) \) of Eq. (4) is given by the stationary points of \( \mathcal {J}({\varvec{P}}, {\varvec{Q}}) \), i.e.,

$$\begin{aligned} \begin{aligned} \frac{\partial \mathcal {J}}{\partial {\varvec{P}}}&= -\,2 {\varvec{YX}}^H {\varvec{Q}} + 2 {\varvec{PQ}}^H {\varvec{XX}}^H {\varvec{Q}}&= 0\\ \frac{\partial \mathcal {J}}{\partial {\varvec{Q}}}&= -\,2 {\varvec{XY}}^H {\varvec{P}} + 2 {\varvec{XX}}^H {\varvec{QP}}^H {\varvec{P}}&= 0 \end{aligned} \end{aligned}$$

Introducing the orthogonality constraint on \( {\varvec{P}} \) in the second equation above yields the following expression for \( {\varvec{Q}} \)

$$\begin{aligned} {\varvec{Q}}^H = {\varvec{P}}^H {\varvec{C}}_{yx} {\varvec{C}}_{xx}^{-1}, \end{aligned}$$

where \( {\varvec{C}}_{yx} = {\varvec{YX}}^H \) and \( {\varvec{C}}_{xx} = {\varvec{XX}}^H \) are the empirical variance–covariance matrices introduced in Sect. 4.1. Inserting this expression for \( {\varvec{Q}} \) in Eq. (4) yields

$$\begin{aligned} \begin{aligned} {{\,\mathrm{minimize~}\,}}_{{\varvec{P}}}&\quad \Vert {\varvec{Y}} - {\varvec{PP}}^H {\varvec{C}}_{yx} {\varvec{C}}_{xx}^{-1} {\varvec{X}} \Vert _F^2 \\ {{\,\mathrm{subject~to~}\,}}&\quad \text {rank } {\varvec{P}} = r \\&\quad {\varvec{P}}^H {\varvec{P}} = {\varvec{I}}, \end{aligned} \end{aligned}$$
(5)

From this point, it is straightforward to show that the above problem can be rewritten as

$$\begin{aligned}&{{\,\mathrm{minimize~}\,}}_{{\varvec{P}}} \text { Tr }\left( {\varvec{C}}_{yy}\right) -2 \text {Tr} \left( {\varvec{P}}^H{\varvec{C}}_{yx}{\varvec{C}}_{xx}^{-1}{\varvec{C}}_{xy}{\varvec{P}}\right) \nonumber \\&{{\,\mathrm{subject~to~}\,}}\text { rank } {\varvec{P}}=r\nonumber \\&{\varvec{P}}^{H}{\varvec{P}}={\varvec{I}} \end{aligned}$$
(6)

\(\text {Tr}\left( {\varvec{C}}_{yy}\right) \) being constant, we thus established the equivalence between the norm minimization problem (4) and the following trace maximization problem

$$\begin{aligned} \begin{aligned} {{\,\mathrm{maximize~}\,}}_{{\varvec{P}}}&\quad \text {Tr}\left( {\varvec{P}}^H {\varvec{C}}_{yx} {\varvec{C}}_{xx}^{-1} {\varvec{C}}_{xy} {\varvec{P}}\right) \\ {{\,\mathrm{subject~to~}\,}}&\quad \text {rank } {\varvec{P}}=r \\&\quad {\varvec{P}}^H {\varvec{P}} = {\varvec{I}}. \end{aligned} \end{aligned}$$
(7)

Finally, recalling that the trace of a matrix is the sum of its eigenvalues, basic linear algebra is sufficient to prove that the solution of the above optimizations problem is given by the first r leading eigenvectors of the symmetric positive definite matrix \( {\varvec{C}}_{yx} {\varvec{C}}_{xx}^{-1} {\varvec{C}}_{xy} \). Once \( {\varvec{P}} \) has been computed, the matrix \( {\varvec{Q}} \) is simply given by \( {\varvec{Q}}^H = {\varvec{P}}^H {\varvec{C}}_{yx} {\varvec{C}}_{xx}^{-1} \) as stated previously.

Interestingly, it can be noted that if \( {\varvec{X}} \) is a full rank \( m \times m \) matrix (i.e., as many linearly independent samples as degrees of freedom), then

$$\begin{aligned} \begin{aligned} {\varvec{C}}_{yx} {\varvec{C}}_{xx}^{-1} {\varvec{C}}_{xy}&= {\varvec{Y}} {\varvec{X}}^H \left( {\varvec{XX}}^H \right) ^{-1} {\varvec{XY}}^H \\&= {\varvec{YY}}^H \\&= {\varvec{C}}_{yy}. \end{aligned} \end{aligned}$$

In this particular case (hardly encountered in fluid dynamics due to the memory footprint of storing m snapshots of a m-dimensional vector), the columns of \( {\varvec{P}} \) are simply the first r leading eigenvectors of the variance–covariance matrix \( {\varvec{C}}_{yy} \), i.e., the POD modes of the output matrix \( {\varvec{Y}} \). Alternatively, if \( {\varvec{X}} \) is a skinny \(m \times n\) matrix (with \( m > n \)) or a rank-deficient \( m \times m \) matrix, one can introduce its economy-size singular value decomposition

$$\begin{aligned} {\varvec{X}} = {\varvec{U}}_{{\varvec{X}}} \varvec{\Upsigma }_{{\varvec{X}}} {\varvec{V}}^H_{{\varvec{X}}}. \end{aligned}$$

The columns of \( {\varvec{P}} \) are then given by the first r left singular vectors of the matrix \( {\varvec{YV}}_{{\varvec{X}}}{\varvec{V}}_{{\varvec{X}}}^H \) where \( {\varvec{V}}_{{\varvec{X}}} {\varvec{V}}_{{\varvec{X}}}^H \) is the projector onto the rowspan of \( {\varvec{X}} \).

1.2 Computing the DMD modes and eigenvalues from the low-rank factorization of \( {\varvec{A}} \)

It must be noted that the columns of \( {\varvec{P}} \) are not the so-called DMD modes although they span the same subspace. The DMD modes can nonetheless be easily computed from the low-rank factorization \( {\varvec{A}} = {\varvec{PQ}}^H \) of the DMD operator. The \(i{\mathrm{th}}\) DMD mode is solution to the eigenvalue problem

Since needs to be in the columnspan of \( {\varvec{A}} \), it can be expressed as a linear combination of the columns of \( {\varvec{P}} \), i.e.,

where \( {\varvec{b}}_i \) is a small r-dimensional vector. Introducing this expression into the DMD eigenvalue problem yields after some simplifications

$$\begin{aligned} {\varvec{b}}_i \lambda _i = {\varvec{Q}}^H {\varvec{P}} {\varvec{b}}_i. \end{aligned}$$

Although \( {\varvec{A}} \) is a high-dimensional \( m \times m \) matrix, \( {\varvec{Q}}^H {\varvec{P}} \) is a small \( r \times r \) matrix whose eigenvectors \( {\varvec{b}}_i \) and eigenvalues \( \lambda _i \) can easily be computed using direct solvers thus enabling the computations of the eigenvalues and eigenvectors of \( {\varvec{A}} \) at a reduced cost.

Similarly, the adjoint DMD modes \( \varvec{\Upphi }_i \) are solution to the adjoint DMD eigenvalue problem

$$\begin{aligned} \varvec{\Upphi }_i \bar{\lambda }_i = {\varvec{A}}^H \varvec{\Upphi }_i, \end{aligned}$$

where the overbar denotes the complex conjugate. These now live in the rowspan of \( {\varvec{A}} \) and can thus be expressed as

$$\begin{aligned} \varvec{\Upphi }_i = {\varvec{Qc}}_i \end{aligned}$$

where once again \( {\varvec{c}}_i \) is a small r-dimensional vector. The corresponding low-dimensional eigenvalue problem then reads

$$\begin{aligned} {\varvec{c}}_i \bar{\lambda }_i = {\varvec{P}}^H {\varvec{Q}} {\varvec{c}}_i \end{aligned}$$

with \( {\varvec{P}}^H {\varvec{Q}} \) a small \( r \times r \) matrix. It must be noted however that, from a practical point of view, converging the adjoint DMD modes \( \varvec{\Upphi }_i \) may require significantly more data than needed to the converge the direct DMD modes .

Broomhead–King embedding for attractor reconstruction

Attractor reconstruction from a single time-series has been instrumental in highlighting the connection between the dynamics of the thermosyphon and that of the canonical Lorenz system (see Figs. 34). The aim of this section is to briefly introduce inexperienced readers to the Broomhead–King attractor reconstruction technique used herein. For more details, please refer to the original paper [20]. For the sake of simplicity and reproducibility, we will apply this attractor reconstruction technique using the Lorenz system

$$\begin{aligned} \begin{aligned} \dot{x}&= \sigma \left( y - x \right) \\ \dot{y}&= x \left( \rho - z \right) - y \\ \dot{z}&= xy - \beta z \end{aligned} \end{aligned}$$

with \( \left( \sigma , \rho , \beta \right) = \left( 10, 28, {8}\big /{3} \right) \) so that it exhibits chaotic dynamics. The properties of this dynamical system have been discussed in Sect. 3. Figure 17 depicts its well-known strange attractor along with representative time-series of x(t) , y(t) and z(t) . In the rest of this appendix, we will work exclusively with the time-series of x(t) . The whole dataset consists of the evolution of x(t) from \(t = 0\) to \(t = 400\) with a sampling period \( \Delta t = 0.001 \) resulting in 400,000 samples.

Although the Takens embedding theorem [19] provides theoretical guarantees to reconstruct a multidimensional attractor from a single time-series, this approach might break when applied to real data that may be contaminated by measurement noise or it might be quite sensitive on the sampling period used to acquire the signal. Additionally, once the lag \( \tau \) has been determined, the number of time-lagged versions of x(t) needed to reconstruct the attractor needs to be determined using technique-based nearest neighbors [52,53,54] which might provide conflicting estimates of the attractor’s dimension and different reconstructions. As to overcome these limitations, Broomhead and King [20] have proposed instead to reconstruct the attractor based on the singular value decomposition of the Hankel matrix constructed from time-lagged version of x(t)

$$\begin{aligned} {\varvec{H}} = \frac{1}{\sqrt{N}} \begin{bmatrix} x_1 &{}\quad x_2 &{}\quad x_3 &{}\quad \cdots &{}\quad x_d \\ x_2 &{}\quad x_3 &{}\quad x_4 &{}\quad \cdots &{}\quad x_{d+1} \\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \cdots &{}\quad \vdots \\ x_{N-d+1} &{}\quad x_{N-d+2} &{}\quad x_{N-d+3} &{}\quad \cdots &{}\quad x_{N} \\ \end{bmatrix} \end{aligned}$$

where \( x_k = x(t_k) \), N is the total length of the time-series (i.e., 400,000 in our case) and d is the maximum number of time-lagged considered. The choice of d will be discussed later on. Decomposing this Hankel matrix as \( {\varvec{H}} = {\varvec{U}} \varvec{\Upsigma } {\varvec{V}}^H \), the dimension of the attractor can then be inferred from the distribution of the singular values \( \sigma _i \). The columns of \( {\varvec{U}} \) then provide an orthonormal basis for the embedding space, while the columns of \( {\varvec{V}} \) describe the evolution of the system within this embedding space, thus enabling the reconstruction we aimed for.

The choice of the number of lags d needed to construct the Hankel matrix is of crucial importance as it will determine the window length \( \tau _w = d \Delta t\) used to embed the dynamics which, in turn, will determine the structure of the singular value spectrum. In practice, the window length \( \tau _w \) needs to be large enough to average out the possible noise contaminating the data. Note however that, in the limit \( \tau _w \rightarrow \infty \), the singular value decomposition of the Hankel matrix converges to the discrete Fourier transform of x(t) . Chaotic systems being characterized by a continuous spectrum, an infinite number of singular components would thus be needed to reconstruct the signal, grossly overestimating the dimension of the attractor. In their paper, Broomhead and King [20] recommend to use a window length \( \tau _w \) of approximately one tenth of the period of oscillation about the unstable saddle-foci. In this work, this window length has been chosen as one twentieth of this oscillating period, even though choosing one tenth does not change qualitatively (nor quantitatively) the results presented. Note that the same rationale has been used when applying this technique to the data from the thermosyphon. For our choice of parameters, Fig. 18 provides the temporal evolution of the system within the embedding space and the corresponding reconstruction of the attractor. The theoretical properties of such time-delay embedding of attractors have been recently studied by Kamb et al. [55].

Fig. 17
figure 17

Projection of the Lorenz attractor on the (xz) plane (left) and corresponding time-series of x(t) , y(t) and z(t) (right). The parameters of the system are chosen as \( \left( \sigma , \rho , \beta \right) = \left( 10, 28, {8}\big /{3} \right) \)

Fig. 18
figure 18

Projection of the Lorenz attractor reconstructed using the Broomhead–King technique on the \((v_1, v_2)\) plane (left) and corresponding time-series of \( v_1(t) \), \(v_2(t)\) and \(v_3(t)\) (right). A window length \( \tau _w = 0.032 \), corresponding to roughly one twentieth of the oscillating period around the unstable saddle-foci, has been chosen

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Loiseau, JC. Data-driven modeling of the chaotic thermal convection in an annular thermosyphon. Theor. Comput. Fluid Dyn. 34, 339–365 (2020). https://doi.org/10.1007/s00162-020-00536-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00162-020-00536-w

Keywords

Navigation