You are on page 1of 8

Exp Fluids (2011) 50:1123–1130

DOI 10.1007/s00348-010-0911-3

RESEARCH ARTICLE

Application of the dynamic mode decomposition to experimental


data
Peter J. Schmid

Received: 7 January 2010 / Revised: 14 May 2010 / Accepted: 5 June 2010 / Published online: 2 February 2011
Ó Springer-Verlag 2011

Abstract The dynamic mode decomposition (DMD) is a Experiments and numerical computations are the two pil-
data-decomposition technique that allows the extraction of lars upon which this transfer relies. In the numerical field,
dynamically relevant flow features from time-resolved ever-increasing computer resources have recently enabled
experimental (or numerical) data. It is based on a sequence the analysis of rather complex flows as to their global
of snapshots from measurements that are subsequently stability, receptivity and controllability (Edwards et al.
processed by an iterative Krylov technique. The eigen- 1994; Lehoucq and Scott 1997; Theofilis 2000). Modern
values and eigenvectors of a low-dimensional representa- techniques, mainly adapted from fast iterative algorithms
tion of an approximate inter-snapshot map then produce for linear algebra problems, are becoming commonplace in
flow information that describes the dynamic processes many large-scale and multi-physics calculations. On the
contained in the data sequence. This decomposition tech- experimental side, remarkable progress has been made in
nique applies equally to particle-image velocimetry data the measurement of flow quantities. In particular, particle-
and image-based flow visualizations and is demonstrated image velocimetry (PIV) techniques are becoming more
on data from a numerical simulation of a flame based on a and more powerful, and the degree of spatial and temporal
variable-density jet and on experimental data from a lam- resolution often rivals equivalent numerical simulations. In
inar axisymmetric water jet. In both cases, the dominant addition, the use of high-speed cameras and image-based
frequencies are detected and the associated spatial struc- visualization techniques has provided unprecedented
tures are identified. insight into fast fluid dynamic processes. In contrast to
numerical simulations, however, the quantitative evalua-
tion of these experimental data lags behind the possibilities
1 Introduction for computational results. The reason for this gap lies in the
above-mentioned algorithms. Nearly all the algorithms in
An objective and quantitative description of flow behavior current use in quantitative fluid dynamics rely on a model
is the underlying principle of many scientific inquiries in equation, commonly the (linearized) Navier–Stokes equa-
computational and experimental fluid dynamics. A pro- tions and their variants. This model appears in the algo-
found understanding of all relevant processes also builds rithms as a matrix-vector multiplication, which is pivotal in
the foundation for the transfer of fundamental fluid effects maintaining orthogonality or in adding robustness to the
into technological applications and fluid-based devices. respective method (Lehoucq and Scott 1997). This neces-
sary model information is easily furnished by a numerical
code, but is impossible to access in experimental mea-
surements. In physical experiments, only the measured data
are available. To achieve the same level of quantitative
description of fluid processes, the common algorithms have
P. J. Schmid (&)
to be modified to eliminate their reliance on a model and to
Laboratoire d’Hydrodynamique (LadHyX),
Ecole Polytechnique, Palaiseau 91128, France restrict their input to data only. The dynamic mode
e-mail: peter@ladhyx.polytechnique.fr decomposition (DMD) (Schmid and Sesterhenn 2008;

123
1124 Exp Fluids (2011) 50:1123–1130

Schmid 2010) achieves this goal by extracting the relevant that, for experimental data, the snapshots vj are produced
dynamics from a sequence of data; similar to the Arnoldi by a nonlinear process. Invoking a linear-tangent
algorithm, a high-degree matrix polynomial is fitted to the approximation, a linear mapping A from one snapshot to
data sequence. An inter-snapshot linear map is then iden- the next is assumed; this mapping is furthermore taken as
tified that acts as a low-dimensional approximation of the constant over the data sequence. We can thus write
system dynamics. This linear map is computed by pro- vjþ1 ¼ Avj : ð2Þ
cessing the data sequence, generated by a nonlinear pro-
cess, and represents the optimal linear operator (in a least- Following the idea underlying Krylov techniques and, in
squares sense) that describes the evolution of the flow over particular, the Arnoldi method (Greenbaum 1997;
a small time interval. The eigenvalues and eigenvectors of Trefethen and Bau 1997), we express and approximate
this map then capture the principal dynamics contained in the linear map A using an (N - 1)-dimensional snapshot
the snapshot basis. basis VN-1
1 . This step can be expressed as
The representation of a nonlinear process by a linear
sample-to-sample map is closely linked to the concept of a fv2 ; v2 ; . . .; vN g ¼ Afv1 ; v2 ; . . .; vN1 g
Koopman operator, an analysis tool for dynamical systems.  fv1 ; v2 ; . . .; vN1 gS ð3Þ
This type of spectral analysis of nonlinear processes pro- or
vides the mathematical foundation of the dynamic mode
decomposition and has recently been applied to complex VN2 ¼ AVN1  VN1 S ð4Þ
1 1
fluid flows (Rowley et al. 2009).
Snapshot-based analysis of fluid flows has previously with S as a companion matrix that simply shifts the
been accomplished by the proper orthogonal decomposi- snapshots 1 through N - 1 (via the subdiagonal matrix
tion (POD) (Lumley 1970; Sirovich 1987; Berkooz et al. entries) and approximates the last snapshot N by a linear
1993), which gained widespread popularity by extracting combination of the previous N - 1 snapshots (Ruhe 1984).
coherent structures from experimental or numerical data. It is thus necessary to best express the last measurement by
This statistical technique provides a hierarchy of spatial a linear combination of the previous measurements. This
structures that are mutually orthogonal and cumulatively procedure will result in the low-dimensional system matrix
capture the energy content of the flow. However, the S. It is then known that the eigenvalues of S, also referred
extraction of the temporal dynamics is not immediate, and to as the Ritz values, approximate some of the eigenvalues
the interpretation of each structure’s contribution to the of the full system matrix A. The associated eigenvectors of
overall dynamic process is not trivial. Nevertheless, POD S provide the coefficients of the linear combination that is
structures are widely used in model reduction and have necessary to express the modal structure within the
shown significant success in feedback control applications snapshot basis. We obtain
(Noack et al. 2003).
min vN  VN1 1 s ð5Þ
s

2 Background where s denotes the last column of the companion matrix S.


The above equation is solved by a straightforward least-
The dynamic mode decomposition (DMD) is based on a squares procedure, i.e., with VN-1 1 = QR as the QR-
time-resolved sequence of flow field measurements. These decomposition of VN-1
1 , we obtain the solution to the above
measurements can consist of visualizations of passive equation as
tracers, image-based techniques or particle-image veloci-
s ¼ R1 QH vN : ð6Þ
metry (PIV) data. The extracted structures will reflect the
type of input data; the temporal dynamics, on the other where QH is the complex conjugate transpose of Q from
hand, will be independent of the furnished data. the QR-decomposition of VN-1 1 . It is important to note that
We will assume general flow field data and denote each at no point in the above procedure do we need the explicit
such field by a vector vj. A sequence of N snapshots is then form of the system matrix A. Only its low-dimensional
written as proxy—which can be extracted solely from the data
fields—is required. This feature is pivotal in adapting an
VN1 ¼ fv1 ; v2 ; . . .; vN g ð1Þ
iterative stability technique to a data-based experimental
where an equal time interval Dt is assumed between two setting. The complete algorithm (for a full-rank data
subsequent snapshots vj and vj?1. It should be kept in mind sequence) is given below.

123
Exp Fluids (2011) 50:1123–1130 1125

Algorithm 1 Dynamic mode decomposition


mapping from one snapshot in space to the next at a dis-
tance Dx further downstream. Consequently, the eigen-
Require: a sequence of n snapshots {v1, v2, ..., vn} sampled value decomposition of S produces spatial stability
equispaced in time with Dt information.
Output: dynamic mode spectrum kj and associated dynamic modes Additional applications of the dynamic mode decom-
DMj with j = 1, ..., n - 1
position (DMD), both in a temporal or spatial setting, cover
Vn-1
1 / {v1, v2, ..., vn-1}
numerous aspects of familiar data processing needs for
Vn2 / {v2, v3, ..., vn}
experimental measurements, among them, the low-dimen-
[Q,R] / qr(Vn-1
1 , 0)
sional representation of a dynamic process, the filtering of
S / R-1QHVn2
raw data based on structural and dynamic coherence, and
[X,D] / eig(S)
the recovery of data from gappy measurements or signals.
kj logðDjj Þ=Dt For a temporally ordered data sequence from a nonlinear
DMj / Vn-1
1 X(:,j) process, we expect neutrally stable eigenvalues as the data
sequence becomes sufficiently long. This is due to the fact
that instabilities will have nonlinearly saturated while
decaying processes will have vanished from the data
The principal parameters of the algorithm are the sep- sequence; only neutrally stable fluid elements will remain
aration Dt between the snapshots and the number N of and will be identified. Growing or decaying eigenvalues stem
processed snapshots. Whereas the latter can be determined from either insufficient data sequences or the decomposition
by monitoring the residual of the least-squares step, the of temporally transient phenomena. It is worth pointing out,
former requires input from the user. For a limited number though, that, within a spatial setting, regions in the flow
of snapshots, the separation between samples has to reflect domain exist where the flow can be aptly described by a
the characteristic time-scale of the fluid phenomenon under linear process. In this case, spatially growing or decaying
investigation. A too low (below the Nyquist frequency) or eigenvalues and dynamic modes are possible and physical.
too high sampling frequency will yield unsatisfactory In contrast to the proper orthogonal decomposition
results. where a spatial orthogonality of the identified structures is
An additional advantage of the dynamic mode decom- enforced at the expense of a mixture of frequencies in each
position lies in the processing of limited or spatially individual POD mode, the dynamic mode decomposition
restricted data (Schmid 2010). In many realistic applica- aims at an orthogonality in time (by identifying pure fre-
tions, multiple flow domains and instability mechanisms quencies). The resulting dynamic modes are in general
are present. For example, for a jet in cross flow (Bagheri non-orthogonal.
et al. 2009) Kelvin–Helmholtz-like instabilities of the
counter-rotating vortex sheet co-exist with wake-vortices
and the breakdown of horse-shoe vortices. These local 3 Demonstration on numerical data
instabilities act on different time-scales and are difficult to
extract and separate within a global stability analysis. Before applying the dynamic mode decomposition (DMD)
However, by extracting measurements from the respective to experimental data, it seems instructive to treat synthetic
localized regions, the DMD algorithm is capable of iden- data in order to familiarize ourselves with its output, features
tifying the system matrix S for each individual subdomain and capabilities. For this reason, we use snapshots generated
and is thus able to represent the dominant dynamical fea- by a numerical simulation of a flame based on a variable-
tures of a multi-scale process more effectively and density fuel jet. Image-based analysis will be demonstrated
efficiently. by tracking a passive tracer in form of a colormap. Selected
Even the analysis of spatially evolving disturbance sequences of a shedding/instability cycle are shown in Fig. 1
dynamics (Schmid and Henningson 2001; Schmid 2010) is for a subregion of the full computational domain. The roll-up
possible. In this case, time-resolved measurements in, say, of a shear instability, originating at the nozzle and advecting
an (x, y)-plane are taken. However, the full spatio-temporal downstream, is clearly visible. Anticipating results of the
‘‘data cube’’ Vðx; y; tÞ is reshaped such that each snapshot subsequent decomposition, it is worth pointing out the
vj represents a (y, t)-field at a given location xj in the x- coexistence of large-scale structures (the rolled-up vortex)
direction. By aligning multiple snapshots at consecutive and small-scale features (sharp gradients).
(equispaced) x-locations, we can produce a spatial The two-dimensional colormap of each snapshot is
sequence of data fields. Feeding this data sequence into the rearranged in a one-dimensional vector, which then form
DMD algorithm yields a low-dimensional representation S the columns of the data matrix VN1 with N = 80 for our
of the spatial evolution matrix A, which describes the case. The companion matrix S is then determined by

123
1126 Exp Fluids (2011) 50:1123–1130

Fig. 1 Selected (equispaced)


snapshots of a passive tracer
from a numerical simulation of
a reactive flow. One shedding
cycle is shown in a subregion of
the full computational domain

expressing the last column v80 by a linear combination of the second mode (DM2) is characterized by large structures
the previous 79 snapshot vectors according to (5). This spanning the mean shear layer; its imaginary part (not
matrix then represents an optimal linear snapshot-to-snap- shown) is shifted by 90°. Both real and imaginary part,
shot map that can further be analyzed. The complex when combined with the temporal behavior of the mode
eigenvalues of S describe the temporal dynamics contained exp(x t) with x & 5.97i, represent a large-scale structure
in the data sequence. In accordance with standard con- advected in the downstream direction. This structure has
vention, the eigenvalues of S are logarithmically mapped been identified as the most relevant within the processed
onto the complex plane, i.e., x ¼ logðeigðSÞÞ=Dt with data sequence. Subsequent dynamic modes, DM4,6,8,10
Dt ¼ 0:053; such that their real part xr represents expo- show increasingly smaller spatial features (reminiscent of
nential growth or decay (depending on the sign) and their higher harmonics), but also increasingly higher frequencies
imaginary part xi contains the temporal frequency (see x4,6,8,10 &i{11.94, 17.91, 23.88, 29.84}.
Fig. 2a). Since real data have been processed, complex In order to further demonstrate the decomposition into
eigenvalues appear as conjugate pairs resulting in a sym- dynamic modes, a superposition of the most relevant
metric spectrum with respect to the axis xi = 0. modes is attempted. This is accomplished by computing the
The spectrum (see Fig. 2a) displays an eigenvalue near amplitude of each mode contained in the original data
the origin, which signifies a corresponding structure that is sequence. The temporal evolution of, say, the k-th mode is
steady in time. The associated dynamic mode (DM1) shows then given by the dynamic mode DMk, properly scaled by
a structure reminiscent of the mean flow. The real part of its amplitude and multiplied by its temporal dependence

a Spectrum b DM1 DM2 DM4 DM6 DM8 DM10


40

30

20

10

ω
i
0

−10

−20

−30

ωr
−40
−0.04 −0.02 0 0.02

Fig. 2 Results from a dynamic mode decomposition of a data sequence (with N = 80) illustrated in Fig. 1. a Dynamic mode spectrum,
logarithmically mapped to the complex plane. b Selected dynamic modes, visualized by the real part of the passive tracer

123
Exp Fluids (2011) 50:1123–1130 1127

Fig. 3 Animation of a
superposition of dynamic
modes. (top) The three most
dominant dynamic modes have
been used; (bottom) eleven
modes, based on all modes
shown in Fig. 2, have been
used. The time instances
correspond to the ones chosen
for the original sequence,
displayed in Fig. 1

exp(xk t). In this manner, each structure can be animated structures—or for model reduction—by including only the
individually resulting in a hierarchy of dynamic processes. most relevant structures of interest. The above demonstra-
Alternatively, a superposition of structures gradually reconsti- tion has been intended for instructive purposes; the pro-
tutes the full dynamic of the original data sequence. Figure 3 cessing of experimental data is more challenging due to the
shows the result of two animations of such superpositions, with presence of noise and uncertainties. The decomposition of an
three and eleven dynamic modes, respectively. The time image-sequence taken from experiments in a water channel
instances correspond to the ones depicted in Fig. 1 for the will be dealt with in the next section.
original data sequence. In the reduced model based on only
three dynamic modes (mean mode DM1, DM2 and its complex
conjugate DM3) the overall large-scale structures are, as 4 Application to an axisymmetric water jet
expected, captured both in space and in time; small-scale details
and sharper features, however, are still missing with this severe The dynamic mode decomposition will be applied to an
a truncation. The superposition based on eleven dynamic image-based flow visualization of a laminar axisymmetric
modes (all modes shown in Fig. 2: mean mode DM1 and jet. In this configuration, a low Reynolds number axisym-
DM2,4,6,8,10 with their complex conjugates) recovers the origi- metric water jet enters, without swirl, into a quiescent pool
nal process to an astonishing degree. of water (Fig. 4). An axisymmetric vortex sheet develops,
This demonstration shows that the dynamic mode which subsequently rolls up according to a Kelvin–Helm-
decomposition (DMD) is effective in detecting and extract- holtz-type instability (Gallaire and Chomaz 2003). Due to
ing the relevant dynamic features of the flow. No assumption the slow speed of the flow, no three-dimensional effects
of a energy-based ranking nor the use of second-order sta- have been observed, and the flow is consequently taken as
tistics, as required for a proper orthogonal decomposition, axisymmetric throughout the window of observation.
has been necessary. The DMD simultaneously provides Snapshots from a sequence of images are then extracted
temporal information (expressed in the spectrum) and spatial by a video camera at a rate of 25 Hertz. These snapshots
details (contained in the modes) of the dynamic process. A are subsequently processed into a pixelized colormap rep-
superposition of the modes, weighted by their amplitude and resenting the evolution of a passive dye tracer. A localized
temporal behavior, gives a low-dimensional representation 3 px 9 3 px Laplacian filter is applied to the raw images to
of the full dynamics that can equally be used for filtering—by eliminate noise, mostly in the freestream. The resulting
neglecting modes with low amplitudes or noisy spatial images used in this experiment are shown in Fig. 4. They

123
1128 Exp Fluids (2011) 50:1123–1130

Fig. 4 Sketch of flow


configuration (left) and selected
representative snapshots taken
from the experiment

show the typical roll-up of the outer jet shear layer due to axial direction is observable, in addition to a more pro-
Kelvin–Helmholtz instabilities. Even in this short sequence nounced slanting of the coherent structures. The axial
of images, the combined advective and diffusive nature of extent of this mode covers nearly the entire observation
the instability is clearly visible, and the scale and speed of window and shows structure even close to the nozzle exit.
the underlying instability can be estimated by inspection Again, a typical staggering of the real and imaginary
alone. components of this mode appears. The eigenvalue corre-
Nevertheless, the images by themselves do not represent sponding to this dynamic mode has a larger temporal fre-
an objective and quantitative means to gain insight into the quency; this tendency supports the fact that the dynamic
prevalent perturbation dynamics. Moreover, only the most mode decomposition has successfully extracted a disper-
dominant features can be observed, whereas more subtle sion relation of the flow field from the furnished data
and smaller-scale instabilities may be missed entirely. For sequence.
this reason, we apply the dynamic mode decomposition The next higher mode, shown in Fig. 5 4r,4i, has even
(DMD) outlined above. Each image is converted into a smaller coherent features, again concentrated in regions of
column-vector, and a rectangular matrix of these vectors maximum mean shear. Significant amounts of structure can
(from 50 images taken) is formed, which represents the be observed upstream near the nozzle. The spatial structure
matrix VN1 with N = 50 from the derivation above. The reveals scales of small to moderate size further downstream
system matrix S is subsequently extracted, and its eigen- of the nozzle, which substantiates the observation that
values, again logarithmically mapped, give a quantitative small-scale features appear further downstream in the axial
picture of the underlying temporal disturbance dynamics. direction. The corresponding eigenvalue shows an expec-
The spatial features corresponding to the identified (and ted larger temporal frequency.
numbered) eigenvalues are shown in Fig. 5. The fifth and final dynamic mode is displayed in Fig. 5
As before, an eigenvalue appears at the origin. This 5r,5i. It shows a higher frequency and a nearly neutral
eigenvalue (which indicates neither exponential growth/ stability behavior. It is thus expected to be readily obser-
decay nor oscillatory behavior) accounts for a steady state vable over the processed time interval.
contained in the data sequence. As such, it represents the The moderate fluid velocities in this jet experiment
mean flow or the time-invariant component of the flow. make the decomposition into dynamic modes via DMD and
This steady component is shown in Fig. 5 (top right). A the interpretation of the results a relatively straightforward
characteristic spreading of the shear layer is observed in the exercise. For high-speed flows, this may become increas-
downstream (top-to-bottom) direction, but no spatial length ingly difficult due to the presence of very small-scale
scale in the axial jet direction can be detected. incoherent fluid elements. Nevertheless, the DMD tech-
More interesting flow features are contained in the nique is able to optimally identify and extract the dynam-
higher modes. These DMD modes now represent the per- ically most relevant structures. The relevance of structures
turbation (rather than the mean flow) dynamics. The second among the many identified modes can be further quantified
DMD mode (Fig. 5 2r,2i) shows a characteristic pattern by computing the amplitude of each structure contained in
located at the shear layer, which represents the roll-up of the original data sequence. In this manner, it is possible to
the axisymmetric vortex sheet and the resulting entrain- discern dynamically irrelevant structures and focus on the
ment process of quiescent ambient fluid into the jet. The more coherent fluid elements and their associated temporal
real and imaginary part of the modal patterns are staggered dynamics.
due to a phase difference of 90o between them. A clear It should be mentioned that in our case, the extracted
spatial scale in the downstream direction is visible, which spectrum is symmetric with respect to the temporal fre-
readily translates into a corresponding (local) wavenumber quency. This is a consequence of processing real-valued
in the axial coordinate. data that yields eigenvalues that are either purely real (non-
The third mode, depicted in Fig. 5 3r,3i, shows even oscillatory) or come in complex conjugate pairs. The
more sophisticated patterns. A slightly smaller scale in the dynamic mode decomposition, however, can also be

123
Exp Fluids (2011) 50:1123–1130 1129

Fig. 5 Results of the DMD DMD-Spectrum


analysis of a N = 50 snapshot 40

sequence of an axisymmetric
water jet. (top left) DMD 30
mean flow
spectrum, logarithmically 5
mapped to the complex plane. 20
(top right) Dynamic mode DM1, 4
representing the mean flow. 3
10
Dynamic modes (2 through 5)
2
corresponding to the labeled 1
eigenvalues; middle row real 0

ωi
part, bottom row imaginary part
−10

−20

−30

−40
−4 −2 0
ωr
(2r) ( 3r) ( 4r) ( 5r)

(2i) ( 3i) ( 4i) ( 5i)

applied to snapshots that have been preprocessed by, e.g., a 5 Summary and conclusions
Fourier transform in a homogeneous coordinate direction
and are thus complex valued. In this case, the DMD A novel decomposition technique, referred to as dynamic
spectrum will be non-symmetric with respect to the real mode decomposition (DMD), has been introduced and
axis. demonstrated on image-based flow visualizations. It relies

123
1130 Exp Fluids (2011) 50:1123–1130

on a time-resolved sequence of snapshots and extracts, in References


an objective and robust manner, flow characteristics that
describe the fluid dynamics over the processed time inter- Bagheri S, Schlatter P, Schmid PJ, Henningson DS (2009) Global
stability of a jet in crossflow. J Fluid Mech 624:33–44
val. Mathematically, the technique is related to the Arnoldi Berkooz G, Holmes P, Lumley JL (1993) The proper orthogonal
method but does not need the explicit or implicit avail- decomposition in the analysis of turbulent flows. Ann Rev Fluid
ability of a stability matrix or flow model. Whereas this Mech 25:539–575
fact somewhat compromises the convergence properties of Edwards WS, Tuckerman LS, Friesner RA Sorensen DC (1994)
Krylov methods for the incompressible Navier-Stokes equations.
the method (when compared to the standard Arnoldi J Comp Phys 110:82–102
method), it allows the treatment of experimental data, the Gallaire F, Chomaz J-M (2003) Mode selection in swirling jet
analysis of subdomains where localized instabilities are experiments: a linear stability analysis. J Fluid Mech
expected or observed and the formulation of a fluid prob- 494:223–253
Greenbaum A (1997) Iterative methods for solving linear systems.
lem using a spatial framework. The application of the SIAM Publishing, Philadelphia
dynamic mode decomposition to a sequence of flow images Lehoucq RB, Scott JA (1997) Implicitly restarted Arnoldi methods
of a slow jet entering a quiescent fluid showcased the and subspace iteration. SIAM J Matrix Anal Appl 23:551–562
detection of dynamically relevant coherent structures that Lumley JL (1970) Stochastic tools in turbulence. Academic Press,
USA
play an important role in characterizing the fluid behavior Noack BR, Afanasiev K, Morzynski M, Tadmor G, Thiele F (2003) A
over the processed time interval. The dynamic mode hierarchy of low-dimensional models for the transient and post-
decomposition has clear advantages over the proper transient cylinder wake. J Fluid Mech 497:335–363
orthogonal decomposition as the former strives for a rep- Rowley C, Mezic I, Bagheri S, Schlatter P, Henningson DS (2009)
Spectral analysis of nonlinear flows. J Fluid Mech 641:115–127
resentation of the dominant flow features within a tempo- Ruhe A (1984) Rational Krylov sequence methods for eigenvalue
rally orthogonal framework (i.e. pure frequencies), while computation. Lin Alg Appl 58:279–316
the latter is based on a spatially orthogonal model. It should Schmid PJ, Henningson DS (2001) Stability and transition in shear
be particularly suited for fluid phenomena that are char- flows. Springer Verlag, New York
Schmid PJ, Sesterhenn JL (2008) Dynamic mode decomposition of
acterized by distinct frequency bands numerical and experimental data. Bull Am Phys Soc, San
It is hoped that the dynamic mode decomposition Antonio/TX
(DMD) will provide a helpful tool for experimentalist to Schmid PJ (2010) Dynamic mode decomposition of numerical and
detect and quantify important mechanisms in time-resolved experimental data. J Fluid Mech 656:5–28
Sirovich L (1987) Turbulence and the dynamics of coherent
measurements of complex fluid flow and help in furthering structures. Quart. Appl Math 593:333–358
our understanding of fundamental fluid processes. Theofilis V (2000) Advances in global linear instability analysis of
nonparallel and three-dimensional flows. Prog Aerosp Sci
Acknowledgments Support from the Agence Nationale de la 39:249–315
Recherche (ANR) through their ‘‘chaires d’excellence’’ program and Trefethen LN, Bau D (1997) Numerical linear algebra. SIAM
from the Alexander-von-Humboldt Foundation is gratefully Publishing, Philadelphia
acknowledged.

123

You might also like