Skip to main content
Log in

Characterizing and correcting for the effect of sensor noise in the dynamic mode decomposition

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

Dynamic mode decomposition (DMD) provides a practical means of extracting insightful dynamical information from fluids datasets. Like any data processing technique, DMD’s usefulness is limited by its ability to extract real and accurate dynamical features from noise-corrupted data. Here, we show analytically that DMD is biased to sensor noise, and quantify how this bias depends on the size and noise level of the data. We present three modifications to DMD that can be used to remove this bias: (1) a direct correction of the identified bias using known noise properties, (2) combining the results of performing DMD forwards and backwards in time, and (3) a total least-squares-inspired algorithm. We discuss the relative merits of each algorithm and demonstrate the performance of these modifications on a range of synthetic, numerical, and experimental datasets. We further compare our modified DMD algorithms with other variants proposed in the recent literature.

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

Similar content being viewed by others

References

  • Bagheri S (2013) Koopman-mode decomposition of the cylinder wake. J Fluid Mech 726:596–623

    Article  MathSciNet  MATH  Google Scholar 

  • Bagheri S (2014) Effects of weak noise on oscillating flows: linking quality factor, Floquet modes, and Koopman spectrum. Phys Fluids 26(9):094104

  • Bai Z, Silverstein JW (2009) Spectral analysis of large dimensional random matrices. Springer, Berlin

    MATH  Google Scholar 

  • Belson BA, Tu JH, Rowley CW (2013) A parallelized model reduction library. ACM T Math Software

  • Chen KK, Tu JH, Rowley CW (2011) Variants of dynamic mode decomposition: boundary condition, Koopman, and Fourier analyses. J Nonlinear Sci 22(6):887–915

    Article  MathSciNet  MATH  Google Scholar 

  • Cheng X, Singer A (2013) The spectrum of random inner-product kernel matrices. Random Matrices Theory Appl 2(04):1350010

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

    Article  MathSciNet  MATH  Google Scholar 

  • Duke D, Soria J, Honnery D (2012) An error analysis of the dynamic mode decomposition. Exp Fluids 52(2):529–542

    Article  MATH  Google Scholar 

  • Eckart C, Young G (1936) The approximation of one matrix by another of lower rank. Psychometrika 1(3):211–218

    Article  MATH  Google Scholar 

  • Epps BP, Techet AH (2010) An error threshold criterion for singular value decomposition modes extracted from PIV data. Exp Fluids 48(2):355–367

    Article  Google Scholar 

  • Golub GH, Van Loan CF (2012) Matrix computations, vol 3. JHU Press, Baltimore

    MATH  Google Scholar 

  • Goulart PJ, Wynn A, Pearson D (2012) Optimal mode decomposition for high dimensional systems. In CDC, pp 4965–4970

  • Grosek J, Kutz JN (2014) Dynamic mode decomposition for real-time background/foreground separation in video. arXiv preprint arXiv:14047592

  • Hemati MS, Williams MO, Rowley CW (2014) Dynamic mode decomposition for large and streaming datasets. Phys Fluids 26(11):111701

    Article  Google Scholar 

  • Hemati MS, Rowley CW, Deem EA, Cattafesta LN (2015) De-biasing the dynamic mode decomposition for applied koopman spectral analysis. arXiv preprint arXiv:150203854

  • Ho BL, Kalman RE (1965) Effective construction of linear state-variable models from input/output data. In: Proceedings of the 3rd annual Allerton conference on circuit and system theory, pp 449–459

  • Jardin T, Bury Y (2012) Lagrangian and spectral analysis of the forced flow past a circular cylinder using pulsed tangential jets. J Fluid Mech 696:285–300

    Article  MathSciNet  MATH  Google Scholar 

  • Jovanović MR, Schmid PJ, Nichols JW (2014) Sparsity-promoting dynamic mode decomposition. Phys Fluids 26(2):024103

  • Juang JN, Pappa RS (1985) An eigensystem realization algorithm for modal parameter identification and model reduction. J Guid Control Dyn 8(5):620–627

    Article  MATH  Google Scholar 

  • Juang JN, Pappa RS (1986) Effects of noise on modal parameters identified by the eigensystem realization algorithm. J Guid Control Dyn 9(3):294–303

    Article  Google Scholar 

  • Juang JN, Cooper J, Wright J (1988) An eigensystem realization algorithm using data correlations (ERA/DC) for modal parameter identification. Control Theory Adv Technol 4(1):5–14

    MathSciNet  Google Scholar 

  • Koopman BO (1931) Hamiltonian systems and transformation in hilbert space. Proc Natl Acad Sci USA 17(5):315

    Article  MATH  Google Scholar 

  • Mezić I (2005) Spectral properties of dynamical systems, model reduction and decompositions. Nonlinear Dyn 41(1–3):309–325

    MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Noack BR, Afanasiev K, Morzynski M, Tadmor G, Thiele F (2003) A hierarchy of low-dimensional models for the transient and post-transient cylinder wake. J Fluid Mech 497:335–363

    Article  MathSciNet  MATH  Google Scholar 

  • Pan C, Xue D, Wang J (2015) On the accuracy of dynamic mode decomposition in estimating instability of wave packet. Exp Fluids 56(8):1–15

    Article  Google Scholar 

  • Provansal M, Mathis C, Boyer L (1987) Bénard-von Kármán instability: transient and forced regimes. J Fluid Mech 182:1–22

    Article  MATH  Google Scholar 

  • Pyatykh S, Hesser J, Zheng L (2013) Image noise level estimation by principal component analysis. IEEE Trans Image Process 22(2):687–699

    Article  MathSciNet  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Rowley CW, Mezić I, Bagheri S, Schlatter P, Henningson DS (2009) Spectral analysis of nonlinear flows. J Fluid Mech 641(1):115–127

    Article  MathSciNet  MATH  Google Scholar 

  • Schmid P, Li L, Juniper M, Pust O (2011) Applications of the dynamic mode decomposition. Theor Comput Fluid Dyn 25(1–4):249–259

    Article  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Schmid PJ (2011) Application of the dynamic mode decomposition to experimental data. Exp Fluids 50(4):1123–1130

    Article  Google Scholar 

  • Schmid PJ, Sesterhenn J (2008) Dynamic mode decomposition of numerical and experimental data. In: 61st annual meeting of the APS Division of Fluid Dynamics, American Physical Society

  • Singer A, Wu HT (2013) Two-dimensional tomography from noisy projections taken at unknown random directions. SIAM J Imaging Sci 6(1):136

    Article  MathSciNet  MATH  Google Scholar 

  • Stewart G (1991) Perturbation theory for the singular value decomposition. In SVD and signal processing, 11: algorithms, analysis and applications

  • Stewart M (2006) Perturbation of the SVD in the presence of small singular values. Linear Algebra Appl 419(1):53–77

    Article  MathSciNet  MATH  Google Scholar 

  • Taira K, Colonius T (2007) The immersed boundary method: a projection approach. J Comput Phys 225(2):2118–2137

    Article  MathSciNet  MATH  Google Scholar 

  • Tao T, Vu V (2012) Random covariance matrices: universality of local statistics of eigenvalues. Ann Probab 40(3):1285–1315

    Article  MathSciNet  MATH  Google Scholar 

  • Tu JH, Rowley CW, Kutz JN, Shang JK (2014a) Spectral analysis of fluid flows using sub-Nyquist-rate PIV data. Exp Fluids 55(9):1–13

    Article  Google Scholar 

  • Tu JH, Rowley CW, Luchtenburg DM, Brunton SL, Kutz JN (2014b) On dynamic mode decomposition: theory and applications. J Comput Dyn 1(2):391–421

    Article  MathSciNet  MATH  Google Scholar 

  • Williams MO, Rowley CW, Kevrekidis IG (2014) A kernel approach to data-driven Koopman spectral analysis. arXiv preprint arXiv:14112260

  • Williams MO, Kevrekidis IG, Rowley CW (2015) A data-driven approximation of the Koopman operator: extending dynamic mode decomposition. J Nonlinear Sci 25(6):1307–1346

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

    Article  MathSciNet  Google Scholar 

  • Wynn A, Pearson DS, Ganapathisubramani B, Goulart PJ (2013) Optimal mode decomposition for unsteady flows. J Fluid Mech 733:473–503

    Article  MathSciNet  MATH  Google Scholar 

  • Zhao Z, Singer A (2013) Fourier–Bessel rotational invariant eigenimages. JOSA A 30(5):871–877

    Article  Google Scholar 

Download references

Acknowledgments

The authors gratefully acknowledge the support for this work from the Air Force Office of Scientific Research grant FA9550-14-1-0289, J. Shang for making available the PIV data for flow past a cylinder, J. Tu for insightful discussions, and the anonymous reviewers for helpful comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Scott T. M. Dawson.

Additional information

This work was supported by the Air Force Office of Scientific Research Grant FA9550-14-1-0289.

Appendix: Quantifying the size of the bias in DMD

Appendix: Quantifying the size of the bias in DMD

We seek to quantify the magnitude of this bias present in DMD that was derived in Sect. 2.2, subject to certain simplifying assumptions on the nature of the data and noise. If the noise is uniform, and spatially and temporally independent, then \(\mathbb {E}( \tilde{N}_X\tilde{N}_X^* ) = \mathbb {E}(U^*N_X N_X^*U) = U^*m \sigma _N^2U = m \sigma _N^2 I\), where \(\sigma _N^2\) is the variance of each independent component of the noise matrix. Furthermore, if we assure that we are projecting onto the POD modes of the clean data, then \((\tilde{X} \tilde{X}^*) = \Sigma ^2\), where \(U \Sigma V^*\) is the singular value decomposition of X. Thus, with these assumptions, Eq. (6) can be simplified to give

$$\begin{aligned} \mathbb {E}(\tilde{A}_m) \approx \tilde{A}(I - m\sigma _N^2\Sigma ^{-2}). \end{aligned}$$
(24)

The (diagonal) entries \(\Sigma _i^2\) of \(\Sigma ^2\) have the interpretation of being the energy content of the ith POD mode. We then should expect that \(\Sigma _i^2 \sim mnq_i \sigma _X^2\), where \(\sigma _X^2\) is the RMS value of the elements in the data matrix X, and \(q_i = \frac{\Sigma _i^2}{\text {Trace}(\Sigma ^2)}\) is the proportion of the total energy of the system contained in the ith POD mode. For this scaling, we make the assumption that adding/removing rows and columns of data (i.e., varying m and n) does not affect either \(\sigma _X\) or \(q_i\). The bias term \(m\sigma _N^2\Sigma ^{-2}\) is a diagonal matrix whose ith entry has a size \((e_b)_i\) proportional to

$$\begin{aligned} (e_b)_i \sim \frac{1}{n q_i SNR^2}, \end{aligned}$$
(25)

where SNR is the signal-to-noise ratio. Thus, sensor noise has the effect of reducing the diagonal entries of the computed \(\tilde{A}_m\) matrix by a multiplicative factor of \(1-\frac{\sigma _N^2}{n q_i \sigma _X^2}\), which means that POD coefficients are predicted to decay more rapidly than they actually do. This effect is most pronounced for lower-energy modes, for which the \(q_i\) is smaller. We thus expect to identify with DMD (continuous-time) eigenvalues that are further into the left half plane than they should be (or would be if we applied DMD to noise-free data). Duke et al. (2012) argue in the case of periodic data that the growth rate of the eigenvalues should typically be the most challenging to identify, since there are a range of preexisting methods that can identify frequencies. Here, we have argued that it is precisely this growth rate that is most affected by noise. Importantly, the amount of bias is independent of m, which suggests that the bias component of the error will be particularly dominant when we have a large number of low-dimensional snapshots. Importantly, this suggests that one cannot always effectively reduce the effect of noise by simply using more snapshots of data, since the bias error will eventually become the dominant error.

While we can now quantify the magnitude of the bias in DMD, we do not as yet know how it compares to the random component of the error that would arise from a given realization of noise. To do this, we will estimate the typical size of the variance of individual entries of \(\tilde{A}\), using the standard definition

$$\begin{aligned} var\left[ \tilde{A}_{ij} \right]&= \mathbb {E}\left\{ \left( (\tilde{Y}_m \tilde{X}_m^+)_{ij} - \mathbb {E}\left[ (\tilde{Y}_m \tilde{X}_m^+)_{ij}\right] \right) \right. \\&\quad \quad \left. \times \left( (\tilde{Y}_m\tilde{X}_m^+)_{ij} - \mathbb {E}\left[ (\tilde{Y}_m \tilde{X}_m^+)_{ij}\right] \right) \right\} . \end{aligned}$$
(26)

Referring back to Eq. (3), if we exclude terms that are quadratic or higher in noise and assume that the noise covariance matrix is sufficiently close to its expected value, we find that

$$\begin{aligned} (\tilde{Y}_m \tilde{X}_m^+)&- \mathbb {E}\left[ (\tilde{Y}_m \tilde{X}_m^+)\right] \\&= (\tilde{Y} + \tilde{N}_Y)(\tilde{X} + \tilde{N}_X) (\tilde{X} \tilde{X}^* + \tilde{X}\tilde{N}_X^* + \tilde{N}_X \tilde{X}^*+\tilde{N}_X\tilde{N}_X^*)^{-1} \\&\quad \quad - \tilde{Y}\tilde{X}^+ - \mathbb {E}(\tilde{N}_X\tilde{N}_X^*)\Sigma ^{-2}\\&= \left[ \tilde{Y}\tilde{X}^+(\tilde{X} \tilde{N}_X^* + \tilde{N}_X\tilde{X}^*)+ \tilde{N}_Y\tilde{X}^* + \tilde{Y}\tilde{N}_X^*\right] \Sigma ^{-2}. \end{aligned}$$

Elements of the terms \(\tilde{X} \tilde{N}_X^*,\ \tilde{N}_X\tilde{Y}^*, \ \tilde{N}_Y\tilde{X}^*\), and \(\tilde{Y}\tilde{N}_X^*\) are uncorrelated sums over m random terms, with each term in the sum having variance \(n q_i\sigma _X^2\sigma _N^2\) where as before \(q_i\) is the energy fraction in the \(i^{\text{ t }h}\) POD mode. This means that the sum will have variance \(mn q_i\sigma _X^2\sigma _N^2\). Assuming that \(\tilde{Y}\tilde{X}^+ (=\tilde{A})\) does not greatly change the magnitude of quantities that it multiplies and assuming that \(q_i\) remains constant when varying m and n, this means that we find that

$$\begin{aligned} var\left[ \tilde{A}_{ij} \right] \sim \frac{\sigma _N^2}{mn\sigma _X^2}. \end{aligned}$$
(27)

Thus, the expected size of the random error in applying DMD to noisy data is

$$\begin{aligned} e_r \sim \frac{1}{m^{1/2}n^{1/2}SNR}. \end{aligned}$$
(28)

Comparing Eq. (28) with Eq. (25), we propose that the bias in DMD will be the dominant source of error whenever

$$\begin{aligned} m^{1/2}SNR > n^{1/2}. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dawson, S.T.M., Hemati, M.S., Williams, M.O. et al. Characterizing and correcting for the effect of sensor noise in the dynamic mode decomposition. Exp Fluids 57, 42 (2016). https://doi.org/10.1007/s00348-016-2127-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-016-2127-7

Keywords

Navigation