Paper 2
Paper 2
(2023) 37:61–82
https://doi.org/10.1007/s00162-023-00643-4
O R I G I NA L A RT I C L E
Received: 10 January 2023 / Accepted: 1 February 2023 / Published online: 16 February 2023
© The Author(s) 2023
Abstract We study the compression of spatial and temporal features in fluid flow data using multimedia
compression techniques. The efficacy of spatial compression techniques, including JPEG and JPEG2000 (JP2),
and spatiotemporal video compression techniques, namely H.264, H.265, and AV1, in limiting the introduction
of compression artifacts and preserving underlying flow physics are considered for laminar periodic wake
around a cylinder, two-dimensional turbulence, and turbulent channel flow. These compression techniques
significantly compress flow data while maintaining dominant flow features with negligible error. AV1 and H.265
compressions present the best performance across a variety of canonical flow regimes and outperform traditional
techniques such as proper orthogonal decomposition in some cases. These image and video compression
algorithms are flexible, scalable, and generalizable holding potential for a wide range of applications in fluid
dynamics in the context of data storage and transfer.
1 Introduction
High-fidelity simulations and experiments within the field of fluid dynamics produce exceedingly large amounts
of data. As the need for higher-fidelity simulations and advanced experimental resources expands, storage and
transfer requirements for spatiotemporal data from simulations become a major challenge. To address this issue,
spatiotemporal redundancies or repeated dominant flow features can be exploited by a variety of compression
techniques to alleviate memory constraints for fluid flow data storage. A variety of compression techniques,
including modal analysis [1–3], sub-sampling and local re-simulation [4], and deep learning [5–8] have been
considered in an effort to reduce the size of fluid flow data. Although effective, these techniques can be
application-specific and struggle to achieve substantial compression ratios without introducing undesirable
compression artifacts such as discontinuities or deletions of flow features.
In comparison, multimedia compression techniques are general and simple to use, and have benefited from
demand for the modern technologies of high-resolution video streaming [9–11] and video-conferencing [12–
14]. These compression techniques are classified into two groups: lossless compression and lossy compres-
sion [15]. With lossless techniques, the data retrieved from or reconstructed from the compressed state is
identical to that preceding the application of a compression algorithm. Hence, this is preferred for archival
purposes and used for medical imaging [16] and technical drawings [17]. In contrast, processed data with
lossy techniques do not necessarily match the original data, enabling a significant data-size reduction in the
compressed state. Since this may introduce compression artifacts such as discontinuities in image data or the
Communicated by Vassilis Theofilis.
V. Anatharaman · J. Feldkamp · K. Fukami (B) · K. Taira
Department of Mechanical and Aerospace Engineering, University of California, Los Angeles, CA 90095, USA
E-mail: [email protected]
62 V. Anatharaman et al.
loss of high spatial frequency information, it is suitable for natural images such as photographs in applica-
tions where imperceptible loss may be acceptable [18]. We consider here the impacts of such losses on fluid
mechanics simulation data to assess the costs of applying lossy techniques. In 2003, Schmalzl [19] considered
multimedia data compression for fluid flows with an example of Rayleigh-Bénard convection. With multimedia
compression technologies having undergone significant advances in the last two decades, we reassess image
and video compressions with modern algorithms for applications to fluid flow data.
Lossy techniques of interest typically involve frequency-domain transformation, filtering, and entropy
coding as components in the compression process. The development of the discrete cosine transform (DCT) [20,
21] has played a crucial role in image compression and is the basis of Joint Photographic Experts Group
(JPEG) [22]. The emergence of JPEG enabled efficient image compression in a wide range of communities
and became a generally accepted format for digital images. After the development of DCT, wavelet transforms
began to be utilized for image compression in such algorithms as JPEG2000 (JP2) [23], which achieves better
compression than the DCT of JPEG as a result of multi-scale properties of wavelets. It is worth pointing out
that there have been studies on computational fluid dynamics that leverage wavelet methods to efficiently
decompose multi-scale features for applications in turbulence modeling and simulations [24].
In tandem with the growth of image compression techniques, advancement in video compression tech-
nologies followed suit since video data can be characterized as a time series of image frames. Generally, these
time frames include both spatial and temporal redundancies. In fact, we often see the similarities (redun-
dancies) between temporally adjacent frames or spatially adjacent pixels. Video compression algorithms are
designed to remove such redundancies and obtain a compact form of the original information. Current video
compression technologies are generally based on the DCT [25]. Although other candidates including fractal
compression [26,27], matching pursuit [28], and discrete wavelet transform (DWT) have been investigated
as the subject of some studies, these are still not used in practical products. Moving Picture Experts Group
(MPEG) series have been traditionally used for video compression of high-definition television [29–31].
H.2xx series was then developed, and they have achieved significant compression compared to the conven-
tional MPEGs [32,33]. Especially in the recent versions such as H.264 and H.265, motion compensation,
quantization, and entropy coding are applied for efficient video compression. More recently, AOMedia Video
1 (AV1), an open, royalty-free video coding format, was released in 2018, achieving enhanced compression
compared to the aforementioned techniques [34,35].
To meet the demand for these image and video compression tools, significant investment and research have
produced compression techniques of impressive efficiency and usability in addition to free video encoders [36]
to promote widespread accessibility. As such, leveraging these multimedia-inspired compression techniques
should also be of particular interest to the fluid dynamics community given the massive scale of data produced,
stored, and transferred. A standardization on one or more multimedia compression formats for storing fluid
flow data in a compressed representation can yield dividends in research output by allowing greater access to
high-fidelity fluid flow datasets and by removing memory constraints as a barrier to entry.
This paper investigates the effectiveness of these image and video compression techniques on fluid flow data.
Spatial image compression techniques, such as JPEG and JP2, alongside spatiotemporal video compression
techniques, namely H.264, H.265, and AV1, are examined for various flow fields, including laminar cylinder
flow, two-dimensional turbulence, and turbulent channel flow. Field variables from simulation data, such as
streamwise velocity and vorticity, are represented as grayscale images, and multiple snapshots are packaged
into a video. These videos are then encoded into a compressed form using the aforementioned multimedia
compression methods. Modern techniques can compress flow data well below 10% of the original file size with
negligible error and preserve the underlying physics of the flow. Although this paper focuses on applications to
canonical fluid flows, the flexibility and scalability of these algorithms suggest an expansive potential within
this field.
Compression is a process in which data are compressed (encoded) into a representation that uses less data,
and decompressed (decoded) into identical data in the case of lossless compression or nearly identical data
in the case of lossy compression. Through this procedure, a compression method reduces bits of the original
data q(x, t) by eliminating statistical redundancies that may be contained within temporally adjacent frames
and spatially adjacent pixels. In general, a data compression algorithm is referred to as an encoder φ while one
that performs the decompression is called a decoder ψ,
Fig. 1 a Spatial compression: an example velocity field of flow over a cylinder q(x) is represented as a grayscale image, encoded
using an image-based technique to a compressed form, and reconstructed as q̃(x) using a decoder. b Spatiotemporal compression:
multiple snapshots of this flow field data q(x, t) are represented as a grayscale video and are compressed to q̃(x, t) with both
spatial and temporal techniques
where γ (x, t) ∈ Rm is the compressed data corresponding to the original data q(x, t) ∈ Rn with m
n. Depending on the extent of compression, the data, and a choice of encoder/decoder, the reconstruction
q̃(x, t) ∈ Rn generally includes some amount of error.
The data compression process is illustrated in Fig. 1 for both image and video compressions. Figure 1a
depicts a lossy spatial image compression technique, involving quantization of the image data in a compressed
space and producing a reconstruction in the image space showing the operations of JPEG and JP2. Figure 1b
provides a visualization of a spatiotemporal compression technique, exploiting a redundant block of a frame
that remains consistent across subsequent frames, similar to H.264, H.265, and AV1. As these algorithms
originated in the multimedia industry, they are optimized for human viewers and involve the removal of high-
frequency components in the data and down-sampling of the color spectrum such that the eyes cannot easily
distinguish compressed data from the original data. For the purposes of this study, we only consider grayscale
images and videos, which are comprised only of a single-component field data matrix, denoted as q̃(x). This is
in contrast to full-color data, which requires red, green, and blue components, and is unnecessary for the current
analysis as we are interested in considering field variables individually. Herein, we consider the application of
five compression techniques on grayscale images and videos. The encoding schemes, which package the data
into a compressed binary form, are detailed in what follows.
64 V. Anatharaman et al.
2 Compression techniques
2.1.1 JPEG
Let us first describe JPEG, which is a standard lossy spatial compression used for encoding image data based
on the discrete cosine transform (DCT) [20]. An example of a JPEG compression process with a vorticity field
of two-dimensional decaying isotropic turbulence is presented in Fig. 2. The images are partitioned into 8 × 8
blocks in a left-to-right, top-to-bottom scan. Pixel values within blocks are quantized to values of [−128, 127]
from [0, 255]. The forward DCT is individually performed at each block and outputs compressed data. The
DCT for 8 × 8 blocks is mathematically expressed as
7
7
1 (2i x + 1)k x π (2i y + 1)k y π
F(k x , k y ) = C(k x )C(k y ) f (i x , i y ) cos cos , (2)
4 16 16
i x =0 i y =0
7
1
7
(2i x + 1)k x π (2i y + 1)k y π
f (i x , i y ) = C(k x )C(k y )F(k x , k y ) cos cos , (3)
4 16 16
k x =0 k y =0
where √
1/ 2 for k = 0
C(k) = (4)
1 otherwise.
Here, F(k x , k y ) denotes the DCT coefficient corresponding to the horizontal wavelength k x and vertical
wavelength k y and f (i x , i y ) describes the pixel value at the location corresponding to i x and i y . In other
words, the forward DCT takes as input a discrete signal of 64 points and produces coefficients for a linear
combination of 64 unique basis signals, each denoting a specific spatial wavelength. Most of the spatial domain
information is concentrated across lower wavelength because of slow spatial variation from one pixel to the next
in image data. This quality permits lossy quantization, which refers to constant values in a quantization table
Q(k x , k y ) with 64 elements. The DCT coefficient is normalized by a constant Q(k x , k y ) in an element-wise
manner,
F(k x , k y )
F (k x , k y ) =
Q
(5)
Q(k x , k y )
where F Q (k x , k y ) is a normalized coefficient and the operation · denotes rounding to the nearest integer.
Quantization tables are provided by the Joint Photographics Experts Group. Note that dividing the DCT
coefficients by values in the quantization table reduces high-wavenumber coefficients to 0, which permits
efficient entropy coding (explained later) to perform the cutoff at high frequencies [37]. The resulting quantized
DCT coefficients form a matrix of size 8 × 8 with low-wavenumber components generally located in the top-left
of the matrix and high-wavenumber coefficients at the bottom-right, as a consequence of the similar distribution
Reconstructed image
Input DCT for 8 8 block (Keeping 8.31% of the DCT coefficients)
8 8 block
Fig. 2 JPEG compression process with an example of two-dimensional isotropic turbulent vorticity
Image and video compression of fluid flow data 65
of spatial modes to which these coefficients correspond. Subsequently, quantized coefficients are ordered from
low to high wavenumbers.
To reduce the data size, entropy coding [38,39], a lossless method of compressing bitstreams with redun-
dancies, is then performed for the output of DCT. The idea of entropy coding is used not only for JPEG but
also other image/video compression techniques such as JP2, H.2xx series, and AV1. To express the encoding-
based data compression, let us consider a message of DAEBCBACBBBC (12 characters). Since this message
includes five different characters, it needs to prepare 3 bits to convert these characters to bits or binary digits
representation. Here, we use the following conversion table,
A B C D E
000 001 010 011 100
With this table, the message is expressed as
D A E B C B A C B B B C
011 000 100 001 010 001 000 010 001 001 001 010
As shown, the number of bits is 36. The idea of the encoding-based compression is to prepare an adaptive
conversion table assigning a shorter bit length for characters that appear in a high probability and a longer bit
length for characters that barely appear. For example, the following adaptive table can be used:
A B C D E
110 0 10 1110 1111
With this new table, the message can be expressed as
D A E B C B A C B B B C
1110 110 1111 0 10 0 110 10 0 0 0 10
The current table can save the total number of bits for the message from 36 to 25. Modern data compression
techniques efficiently find such an adaptive conversion table for saving image and video sizes.
The presence of a better adaptive conversion can be proven with the source coding theorem [40]. For any
data, the expected code length should satisfy the relationship,
where l is the number of symbols in a message, d is the coding function, b is the number of symbols in a
table, and P is the probability of the original symbol. An entropy coding method attempts to approach the
lower bound. For JPEG compression, Huffman coding [41] is used to determine an adaptable table composed
of the estimated probability of occurrence for each possible value. Huffman coding uses binary trees [42] for
efficient encoding.
JPEG2000 (JP2) is a successor to JPEG. JP2 operates using a similar four-step process to JPEG, involving
image partitioning, frequency-domain transformation, quantization, and entropy coding. In contrast to JPEG,
JP2 introduces more advanced dynamic tiling algorithms based on variable-sized macroblocks. This makes use
of discrete wavelet transform (DWT) [43,44] and performs additional preprocessing prior to entropy coding.
Tiling in this context refers to the partitioning of the source image into several non-overlapping rectangular
blocks, each of which is processed distinctly. Whereas JPEG restricts tile sizes to 8 × 8, tiles in JP2 can be
of arbitrary size up to the image dimensions. The DWT is applied to each tile in a manner similar to DCT,
decomposing a signal into a linear combination of wavelet functions. The coefficients in this linear combination
correspond to a specific wavelet basis function in the signal. A wavelet can be defined as a scale and shift of a
basis wavelet [45]. Child wavelets [46] are generally considered for DWT, given by
1 s − 2g r
ψg,r (s) = g/2 ψ (7)
2 2g
where g is a scaling factor, r is a shift factor, and s corresponds to the index of the one-dimensional representation
of an image. In other words, the flow field snapshot is converted to a one-dimensional representation and the
66 V. Anatharaman et al.
Fig. 3 An example of the two-level discrete wavelet transform for two-dimensional homogeneous turbulent vorticity field, used in
JP2 compression. High-pass filtering yields three large images. Low-pass filtering and downscaling are then performed, producing
the three small images. a The final approximation image. b, e Vertical, c, f horizontal, and d, g diagonal coefficients of the second
b–d and the first levels e–g are shown
independent variable upon which this one-dimensional signal f (s). The DWT coefficient given a wavelet of
the preceding definition is then
∞
Fg,r = f (s)ψg,r ds (8)
−∞
where f (s) is a one-dimensional signal. The signal can be reconstructed through the summation of the product
of each coefficient with the corresponding wavelet. In a discrete interpretation, this is written as
∞
∞
f (s) = Fg,r ψg,r . (9)
g=−∞ r =−∞
The summation bounds for both the calculation of the coefficients and reconstruction of the signal can be set
to finite values and can still produce lossless reconstructions assuming that the wavelets contain the maximum
and minimum wavelengths within the source image.
An example of the DWT for the vorticity field of two-dimensional decaying turbulence is presented in Fig. 3.
The DWT can be applied recursively to one-dimensional signals to produce higher-fidelity representations of
Image and video compression of fluid flow data 67
Motion vector
Fig. 4 Motion compensation (MC), used in temporal compression of H.264, H.265, and AV1, is exemplified by the calculation of
a motion vector field, describing the translation of pixels between successive frames. As an example, a streamwise velocity field
u of three-dimensional turbulent channel flow is considered. Subtracting the prediction frame from the reference frame yields a
residual image, which is stored
data. As such, successive high-pass filters are applied on down-sampled images, producing a higher-fidelity
representation of spatial frequencies in the source image. The DWT can be extended to higher dimensions by
applying the one-dimensional DWT on rows and columns. The recursive application of the DWT produces 2n
distinct filtered images where n is the number of times the DWT is applied.
Similar to JPEG compression, the DWT coefficients are quantized following the transformation on the
wavespace,
Q |Fa,b |
Fa,b = sign(Fa,b ) , (10)
b
where b is defined as the quantization step. DWT coefficients within the range (−b , b ) are quantized
to 0. Following quantization, the coefficients are processed in preparation for entropy coding. Arithmetic
coding [47] is used for entropy coding in JPEG2000. While Huffman coding separates the original data into
component symbols and replaces each with a code in a table, arithmetic coding encodes the entire message
into a single number represented with an arbitrary-precision fraction pa , where 0 ≤ pa < 1.
The H.264 video compression includes a multi-step process, consisting generally of prediction, transformation
(a set of frequency-domain representation and quantization in image compression), and entropy encoding on
the encoder side. A similar process for file reconstruction is performed on the decoder side. Prediction in
video compression amounts to an operation to remove redundancies in the given signal. H.264 supports a
range of prediction options such as intra-prediction used for data within the current frame, inter-prediction
for motion compensation, and multiple block-size-based predictions. An accurate prediction implies that the
residual contains very little information, amounting to good compression performance.
Video data are first partitioned into macroblocks of dimension 16 × 16 pixels. A prediction of the current
macroblock is formed using 4 × 4 and 16 × 16-sized blocks in the case of intra-frame prediction, referring to
predictions from surrounding blocks within the same frame. A range of block sizes from 4 × 4 to 16 × 16 are
also considered in the case of inter-frame prediction, referring to predictions from previously coded frames.
Macroblock prediction is further discretized into intra-prediction with neighboring blocks in the current frame,
blocks in a previously coded frame, and blocks from up to two previously coded frames.
In intra-prediction, the size of prediction macroblocks can be three cases: 16 × 16, 8 × 8, or 4 × 4 pixels [32].
The choice of block size is made primarily based on prediction efficiency. One of several prediction modes
where each prediction mode indicates a direction in which to extrapolate pixel values or to average across all
pixels. In the case of inter-prediction, the reference frame, where the prediction block is situated, is chosen
from several previously decoded frames. A motion vector is obtained for the current macroblock based on
the offset from the prediction block or from previously coded motion vectors. The prediction frame is then
calculated using the motion vector and the reference frame. The residual between the reference and prediction
frames is finally stored, as illustrated in Fig. 4. These motion vectors are optionally weighted to account for
temporal proximity between frames, and are sent into the data stream. Generally, a deblocking filter is further
68 V. Anatharaman et al.
applied to each frame to store in a decoded format for subsequent inter-frame predictions in smoothing the
sharp edges caused by the use of block coding [48].
Transformation involves the conversion of blocks to frequency-domain representations and quantization
of coefficients corresponding to high wavelength data. The DCT step is given by the transformation of block
X by matrix A into DCT coefficients Y for each macroblock. For the case of 4 × 4 blocks, these matrices
X, Y ∈ R4×4 are expressed as
⎡ ⎤ ⎡ ⎤
a a a a a b a c
⎢b c −c −b⎥ ⎢a c −a −b⎥
Y = AX A T = ⎣ ,
a −a −a a ⎦ ⎣a −c −a b ⎦
X (11)
c −b b −c a −b a −c
√ √
where a = 1/2, b = 1/2 cos(π/8), c = 1/2 cos(3π/8). The rows of A are orthonormal. To calculate
Eq. 11 on a practical processor, the approximation for b and c is required. This is achieved with a fixed-point
approximation, which is equivalent to scaling each row of A by 2.5 and rounding to the nearest integer. A core
transform C f 4 , which scales each term of A by 2.5 and rounds to the nearest integer, and a scaling matrix S f 4 ,
which restores norms of row of C f 4 to 1 by scaling, are, respectively, defined as,
⎡ ⎤ ⎡ ⎤
c1 c1 c1 c1 s1 s2 s1 s2
⎢c c1 −c1 −c2 ⎥ ⎢s s3 s2 s3 ⎥
Cf4 = ⎣ 2 , Sf4 = ⎣ 2 ,
c1 ⎦ s1 s2 s1 s2 ⎦
(12)
c1 −c1 −c1
c1 −c2 c2 −c1 s2 s3 s2 s3
√
where c1 = 1, c2 = 2, s1 = 1/4, s2 = 1/(2 10), and s3 = 1/10. Matrix Y is then determined as
Y = [C f 4 XC Tf 4 ]S f 4 . (13)
Similar DCT approximations are specified for other block sizes, involving C f 8 , S f 8 , and others [32]. A
quantization mechanism similar to JPEG is applied, with a quantization table specified for various block
sizes. The quantized DCT coefficients are then traversed in an oscillating, “zig-zag” manner from low- to
high-wavenumber components. Entropy coding is finally applied to the output of DCT.
H.265 was released in 2013 as the successor to H.264. While the fundamental architecture is unchanged from
H.264, H.265 makes use of coding tree units which are similar to macroblocks but expand the range of possible
dimensions, with variable dimensions selected by the encoder, allowing coding tree units to be divided into
sub-blocks. Predictions and reconstructions are performed on coding tree units and supported sub-block sizes
range from 64 × 64 to 4 × 4 pixels. Motion vectors are predicted based on those of adjacent units or blocks in
the case of intra prediction, or previous encoded frames in the case of inter prediction. As in H.264, the DCT
is performed at the coding tree block level, and the resultant coefficients are subjected to scalar quantization
and entropy coding. Instead of deblocking filter in H.264, a sample adaptive offset filter is applied within the
prediction loop to improve the quality of the compressed data [49].
2.2.3 AV1
AV1 was released in 2018 in an effort to replace the H.2XX series of video compression algorithms. In
AV1, much of the fundamental architecture from H.2XX is maintained. Frames are partitioned using a 4-way
partition tree with dimensions ranging from 128 × 128 to 4 × 4. AV1 supports 56 directional spatial modes
for intra-frame prediction with finer angle variations than that provided by H.2XX. AV1 extends the number
of reference frames that any given frame can use to perform predictions to seven references for inter-frame
prediction, thus enabling more accurate encoding of data with rich temporal characteristics. Motion vector field
formation is improved by expanding the spatial search domain for vector candidates and through the utilization
of a temporal motion field estimation system. AV1 also extends frequency-domain transform algorithms to
include the asymmetric discrete sine transform with a richer set of transform kernels for varying block sizes.
Entropy coding and scalar quantization are also used as well as H.2XX compression algorithms. To perform
deblocking, a constrained directional enhancement filter and loop restoration filters are applied [34]. These
filters are able to effectively remove artifacts without causing blurring, compared to conventional deblocking
filters [50].
Image and video compression of fluid flow data 69
3 Flow fields
Let us apply the compression techniques presented above to representative fluid flow datasets from numerical
simulations. We describe herein the problem setup of the example flow fields we analyze and the simulation
approach used to generate them.
Bluff body flow forms a large class of problems, such as the vortex shedding around a cylinder. We first apply
the compression techniques to the two-dimensional cylinder wake obtained by direct numerical simulation
(DNS) [51,52]. The governing equations are the incompressible Navier–Stokes equations,
∇ · u = 0, (14)
∂u 1
+ u · ∇u = −∇ p + ∇ u,
2
(15)
∂t Re D
where u and p are the non-dimensionalized velocity vector and pressure, respectively. All variables are non-
dimensionalized with the fluid density ρ, the uniform velocity U∞ , and the cylinder diameter D. The Reynolds
number is defined as Re D = U∞ D/ν = 100 with ν being the kinematic viscosity. We consider five nested
levels of multi-domains with the finest grid level covering (x, y) = [−1, 15] × [−8, 18] and the largest
domain being (x, y) = [−5, 75] × [−40, 40]. The time step for DNS is t = 2.50 × 10−3 . We extract the
domain around a cylinder body over (x ∗ , y ∗ )/D = [−0.7, 15] × [−5, 5] with (N x , N y ) = (500, 300) and
(x, y) = (0.0314, 0.0333). The flow exhibits vortex shedding with a single period with 21 snapshots. For
the compression analysis, 160 temporal snapshots of grayscale images of the streamwise velocity field u are
considered.
To examine the data compression performance by the present techniques for more complex turbulent flows,
we also consider a two-dimensional decaying homogeneous isotropic turbulence. This time-varying flow can
be regarded as a canonical fluid flow example for a broad range of turbulent flows. The dataset is prepared by
DNS using the two-dimensional vorticity transport equation [53]. We set the initial Reynolds number Re0 ≡
u ∗l0∗ /ν = 80.4, where u ∗ is the characteristic velocity obtained by the square root of the spatially averaged
initial kinetic energy, l0∗ = [2u 2 (t0 )/ω2 (t0 )]1/2 is the initial integral length, and ν is the kinematic viscosity.
The computational domain and the numbers of grid points are set to L x = L y = 1 and N x = N y = 128,
respectively. We use 1000 snapshots in an eddy turn-overtime of t ∈ [2, 6] with a time interval of t = 0.004.
For the data compression analysis, 128 × 128 grid with grayscale contours of the vorticity field ω are used.
To further demonstrate the present data compression techniques, we also examine turbulent channel flow at
a friction Reynolds number of Reτ = u τ δ/ν = 180, where u τ is the friction velocity, δ is the half-width
of the channel, and ν is the kinematic viscosity. This flow involves a broader range of spatiotemporal flow
scales and fewer redundancies compared to the previous two examples. The datasets are prepared by a three-
dimensional DNS [54,55], numerically solving the incompressible Navier–Stokes equations. The present DNS
has been validated by comparison with spectral DNS data of Moser et al [56]. The streamwise, wall-normal,
and spanwise spatial coordinates are denoted by x, y, and z, respectively. The size of the computational domain
and the number of grid points here are (L x , L y , L z ) = (4πδ, 2δ, 2πδ) and (N x , N y , Nz ) = (256, 96, 256),
respectively. The grids in the streamwise and spanwise directions are taken to be uniform, while that in the
y direction is a non-uniform grid. The no-slip boundary condition is imposed on the walls and a periodic
boundary condition is applied to the x and z directions. The flow is driven by a constant pressure gradient. In
what follows, we denote wall-unit quantities with the superscript +.
For the present study, an x − z cross-sectional streamwise velocity u at y + = 13.2 is analyzed, where
representative streak structures are present [57]. Fifty temporal snapshots of a 256 × 256 spatial grid over
t + ∈ [0, 63] are formatted into grayscale data and are used for compression assessment.
70 V. Anatharaman et al.
4 Results
For image compression, matrices of flow field data corresponding to specific temporal snapshots are represented
as uncompressed grayscale images with 8 bits. We have confirmed that the loss of precision in converting to the
grayscale image does not significantly affect the statistics of fluid flows. These images are compressed using
JPEG and JP2 encoders. For video compression, the matrices of flow field data are represented as grayscale
images and concatenated into uncompressed, grayscale videos. This raw video file is used as the input to
the H.264, H.265, and AV1 encoders. To obtain control over the outputted file size for the purposes of this
analysis, a two-pass encoding scheme is considered. In the two-pass encoding, the encoder runs twice. The
first run is used to collect some information and statistics such as how many bytes would be needed for data
compression and the second run performs the actual encoding. These two processes enable the use of the
information collected in the first run to achieve enhanced compression.
This study uses FFmpeg [36], a free and open-source software consisting of various libraries for handling
video, audio, and other multimedia files. These libraries can be easily used from the command line ffmpeg.
Default encoding settings for the relevant FFMpeg library are used to maintain consistency across all tests.
To establish a baseline performance for comparison, individual flow snapshots are compressed using singular
value decomposition (SVD) [58]. Individual snapshots are decomposed into left and right singular vector
matrices U and V T , and a diagonal matrix containing the singular values. Snapshots are reconstructed by
retaining the r leading modes. Here, the compression ratio η in the SVD context is then defined as
r (m + n + 1)
η= , (16)
mn
where m and n are the snapshot dimensions in the horizontal and vertical directions and a value of η = 1
corresponds to the original, uncompressed snapshot. In the present study, the snapshot dimensions for each
flow example are set as (m, n) = (500, 300) for cylinder wake, (128, 128) for two-dimensional decaying
turbulence, and (256, 256) for turbulent channel flow, respectively.
Let us compare the image compression techniques with the cylinder wake example. As for the data attribute,
we use a streamwise velocity u. The compressed wake fields and the spatial absolute error distributions with
η ≈ 0.05 are presented in Fig. 5. The L 2 error norm of reconstruction ε = || f Ref − f Comp ||2 /|| f Ref ||2 ,
where f Ref and f Comp are, respectively, the reference and compressed flow fields, is also shown underneath
the decoded fields. The SVD produces negligible error for the entire flow field, although slight discontinuities
100
50
0
-50
-100
{ , } = {0.00670, 0.0532} { , } = {0.0176, 0.0548} { , } = {0.00370, 0.0507}
Spatial error distribution
0.05
0.04
0.03
0.02
0.01
0
Fig. 5 Comparison of image compression techniques for cylinder wake at Re D = 100. A streamwise velocity field u is considered.
The L 2 error norm of the reconstruction ε and the compression ratio η are shown underneath each flow field contour. Spatial
absolute error distribution for each compression technique is also presented
Image and video compression of fluid flow data 71
100
50
0
-50
-100
{ , } = {0.00820, 0.282} { , } = {0.0350, 0.282} { , } = {0.0233, 0.270}
Fig. 6 Comparison of image compression techniques for two-dimensional decaying turbulence. A vorticity field ω is considered.
The L 2 error norm of the reconstruction ε and the compression ratio η are shown underneath each flow field contour
Fig. 7 Comparison of image compression techniques for turbulent channel flow at Reτ = 180. A streamwise velocity field u is
considered. The L 2 error norm of the reconstruction ε and the compression ratio η are shown underneath each flow field contour
are observed in the wake region. By comparison, JPEG compression introduces some compression artifacts
including discontinuities of grayscale contours and granulated vortical structures. This is due to fixed-size areas
upon which the DCT is performed and elementary anti-blocking features. This indicates that compression based
on the fixed block of 8 × 8 pixels is not appropriate for flow fields that include fine-scale spatial variations.
In contrast, the compressed flow field with the JP2 algorithm retains wake features even while achieving
significant data compression, with the L 2 error of 0.00370. These results support the effectiveness of the
adaptive block size of DWT in JP2 compression for bluff body wake datasets.
Next, we apply the image compression techniques to two-dimensional decaying homogeneous isotropic
turbulence, as shown in Fig. 6. We use a vorticity field ω as a quantity of interest and compare the compression
results with η ≈ 0.280. Similar to the cylinder example, the SVD can provide a smooth field and small error
for the entire flow field. Although SVD can achieve a reasonable compression for the laminar cylinder wake
and two-dimensional turbulence that are mainly composed of large vortical structures, we discuss later how
the presence of fine-scale turbulent structures alters the compression performance. For this two-dimensional
turbulence, the effect of 8×8 pixel blocks can be clearly observed in JPEG compression. Such pixelized artifacts
on the flow field can be mitigated by using the JP2 compression technique, analogous to the observation with
the cylinder example.
The limitation of the SVD and the efficacy of the DWT-based process in the JP2 algorithm are further
emphasized in the example of more complex turbulence. Here, the compression techniques are applied to
a streamwise velocity u of the three-dimensional channel flow. The compression results with η ≈ 0.025
are compared in Fig. 7. As shown, the SVD-based compression cannot retain the important features of the
streaks. Compared to SVD, JPEG provides a better reconstruction although it also introduces discontinuities
that obscure small spatial length scales in the flow field. Surprisingly, JP2 produces non-negligible artifacts
and maintains an L 2 error norm less than half that of SVD. The channel flow field at this low η remains nearly
indistinguishable from the uncompressed flow field, also preserving the streak spacing of the reference DNS
field [57,59]. These observations suggest the effectiveness of the JP2 algorithm for image compression of
complex fluid flow data.
Building on these assessments, the L 2 error between compressed and uncompressed flow fields is evaluated
across compression ratios, as shown in Fig. 8a. The error is averaged over all temporal snapshots of each flow
72 V. Anatharaman et al.
(a) (b)
: SVD : Cylinder
: JPEG : 2D turbulence
: JP2 : Channel
SVD SVD
SVD
JP2
JP2
JP2
JPEG
JPEG
JPEG
Fig. 8 Relationship between a the L 2 error norm ε, b SSIM, and image compression ratio η. Zoom-in view of η-SSIM curve for
c cylinder wake, d two-dimensional turbulence, and e turbulent channel flow
example. In general, all compression algorithms produce an asymptotically decaying L 2 error. JPEG introduces
appreciable error at low η, in the same order as SVD compression. It is worth pointing out that JP2 performs
especially well at low η while SVD compression produces the lowest L 2 error for high η for all flow fields.
As an additional metric for quantifying the error introduced by each compression method, the localized
structural similarity index (SSIM) [60] is computed between compressed and uncompressed flow fields. SSIM
can capture spatial correlation around pixels and is less sensitive against a pixel-wise error caused by translation
and rotational difference compared to the L 2 error. Hence, SSIM is suited for the image and video-based
compression analysis. The SSIM χ is defined as
χ = l(i x , i y )c(i x , i y )s(i x , i y ) (17)
where
2μx μ y + C1 2σx σ y + C2 σx y + C 3
l(i x , i y ) = , c(i x , i y ) = 2 , s(i x , i y ) = (18)
μ2x + μ2y + C1 σx + σ y2 + C2 σx σ y + C 3
with μx and σx defined as the mean and standard deviation of i x , respectively, σx y being the covariance of
i x and i y , and c1 , c2 , and c3 being constants to stabilize division. We set {C1 , C2 , C3 } = {0.16, 1.44, 0.72}
following Wang et al. [60]. The resultant value lies between 0, representing no similarity, and 1, representing
an identical image. The relationship between the image compression ratio and the L 2 error is depicted in
Fig. 8b. Generally, JP2 and SVD produce a negligible decrease in the SSIM at low compression ratios and
asymptotically approach an SSIM value of 1 at higher compression ratios. The SSIM value of the cylinder flow
field with JPEG compression applied decays by approximately 10%, as a result of significant discontinuities
produced by JPEG.
We also present the zoom-in view of the relationship between SSIM and the compression ratio η for each
flow, in Fig. 8c–e. Similar to the observation in the η − ε curves in Fig. 8a, SVD and JP2 provide high SSIM
scores compared to JPEG. Especially at excessive compression (low η) of three-dimensional turbulent channel
flow, JP2 can provide better reconstructions than the other two cases. Although scalar metrics such as the L 2
error ε and SSIM are useful, we note that monitoring not only scalar values but also decoded flow fields with
Image and video compression of fluid flow data 73
Fig. 9 Kinetic energy spectra for two-dimensional decaying homogeneous isotropic turbulence using a JPEG and d JP2. b, e
Streamwise and c, f spanwise kinetic energy spectrum of three-dimensional turbulent channel flow compressed with b, c JPEG
and e, f JP2
statistics is important in assessing how vortical structures can be retained through data compression because
the influence of local structures are averaged.
We are additionally interested in whether finer structures in flow images can still be retained through the
present compression process. To examine this aspect, we consider the kinetic energy spectrum of both two-
and three-dimensional turbulence examples, as summarized in Fig. 9. The kinetic energy spectrum E(k) for
two-dimensional decaying turbulence is
1
E(k) = (u i u i ), (19)
2
where u i are the components of the fluctuating velocity and the overbar denotes an averaging operation in
space and time. For three-dimensional turbulent channel flow, the one-dimensional streamwise and spanwise
spectra are evaluated
z,t x,t
E uu (k x+ ; y + ) = û ∗ û , E uu (k z+ ; y + ) = û ∗ û , (20)
ˆ denotes the one-dimensional Fourier transformed variable.
where (·)∗ represents the complex conjugate and (·)
Here, we compare three compression ratios, denoted as low, medium, and high, for each turbulent flow.
JP2 demonstrates a strong adherence to the kinetic energy spectrum of the uncompressed flow field in
both the x and z directions while JPEG compression at low η introduces non-negligible errors at higher wave
numbers. This is a consequence of the quantization step of JPEG compression that removes high wavelength
scales from the image. Considering the overestimation of E uu (k x+ ) as seen in Fig. 9 when using JPEG, this
is likely caused by the absence of a deblocking filter, producing more high wavelength artifacts in the image
than what exists in the uncompressed data. Similarly, the underestimation of E(k) by JP2 can be attributed
to adaptive block sizes that produce a lower peak signal-to-noise ratio, indicative of lower quality. In general,
JP2 is more adept at preserving high-wavenumber structures.
From the perspective of information, fluid flows are inherently temporally redundant—as such, video compres-
sion algorithms that perform temporal compression are a powerful tool, achieving compression performance
that outperforms the previously analyzed image-based techniques. This section assesses the capabilities of
74 V. Anatharaman et al.
Fig. 10 Comparison of video compression techniques applied on a streamwise velocity field u of cylinder wake at Re D = 100,
compressed using H.264, H.265, AV1, and POD compression algorithms. The L 2 error norm of the reconstruction ε and the
compression ratio η are shown underneath each flow field contour
video compression techniques such as H.264, H.265, and AV1 compression algorithms for time-varying fluid
flow data. Additionally, proper orthogonal decomposition (POD) compression [61] is considered to compare
this familiar method of compression within the fluid dynamics community with those analyzed herein [3,62].
POD is used to decompose a matrix of vectorized, temporally evolving flow field data into a set of basis modes
and eigenvalues that contain coherent flow structures
n and can be used for flow field reconstruction. Formally, a
flow field q(x, t) − q(x) can be represented as j=1 a j φ j where a j is the temporal coefficient for mode φ j .
The value of a j is the inner product between the mode φ j and the mean-subtracted flow field, q(x,t) − q(x).
This modal representation can be truncated to r modes, such that the flow field is approximated by rj=1 a j φ j .
This study uses the snapshot POD method [63] for comparison to the other video compression techniques.
The compression ratio for a reconstructed flow field containing r modes is evaluated as
r (m + n) + m
η= , (21)
n(m + n) + m
where m is the total number of pixels in the flow field and n is the total number of flow snapshots. Note that
POD is hereafter used for time series of flow snapshots in comparing to the video compression techniques
while we performed an instantaneous SVD for image compression,
The results of video compression for laminar cylinder wake at Re D = 100 are shown in Fig. 10. Here,
we compare the decoded streamwise velocity field u with η ≈ 0.02 − 0.03. POD compression introduces
negligible error, likely as a result of the temporally redundant nature of periodic wake and the larger coherent
modal structures that POD is able to extract. By comparison, H.264 compression produces significant artifacts
at low η. H.264 struggles likely because inter-frame prediction candidates are chosen from a shallow time
range. We also observe that H.265 fails to improve in terms of error level over H.264 for the cylinder wake.
This is due to the employment of a similar inter-frame prediction and selection algorithm to that of H.264.
Compared to these H.2XX series, the AV1 algorithm provides much better compression, achieving a lower
L 2 error than that achieved by H.264 and H.265. This highlights the enhanced capability of AV1 to compress
laminar and temporally redundant flow fields.
Image and video compression of fluid flow data 75
Fig. 11 Comparison of video compression techniques applied on two-dimensional isotropic turbulent vorticity field, compressed
using H.264, H.265, AV1, and POD compression algorithms. The L 2 error norm of the reconstruction ε and the compression
ratio η are shown underneath each flow field contour
Fig. 12 Comparison of video compression techniques applied on a streamwise velocity field u of three-dimensional turbulent
channel flow at Reτ = 180, compressed using H.264, H.265, AV1, and POD compression algorithms. The L 2 error norm of the
reconstruction ε and the compression ratio η are shown underneath each flow field contour
We next examine the video compression techniques for two-dimensional decaying homogeneous isotropic
turbulence, as summarized in Fig. 11. The flow fields are compared for the compression ratios of η ≈ 0.02 −
0.03. Similar to the cylinder case, POD compression provides a reasonable reconstruction, likely because
large-scale vortical structures are dominant at this particular time. It is, however, easily anticipated that the
error of this time-varying flow relies on the presence of a range of length scales, as the small length scales
disappear with the progress of the decay over time [64,65]. The dependence of the compression performance
over time for decaying flow will be examined later. While H.2xx compression techniques provide a reasonable
reconstruction, AV1 provides better compression without suffering from pixelized artifacts. These results
suggest the powerful capabilities of novel deblocking filters for fluid flow applications.
The video compression techniques are also applied to the x − z sectional streamwise velocity field u of
three-dimensional turbulent channel flow at Reτ = 180, as depicted in Fig. 12. The compressed flow fields are
compared for η ≈ 0.150. In contrast to the other flow examples, POD compression produces significant visible
artifacts and a high error value for turbulent channel flow because of a complex temporal evolution of the flow
field. POD requires a greater number of modes for adequate reconstruction [66–68]. Although H.265 improves
over H.264 significantly for turbulent channel flow, this still produces few observable discontinuities. This is
76 V. Anatharaman et al.
(a) (b)
: POD : Cylinder
: H264 : 2D turb.
: H265 : Channel
: AV1
(c) (d)
POD
POD
4
H26
H264
AV1 H265
AV1
H265
Fig. 13 Relationship between a the L 2 error norm ε, b SSIM, and video compression ratio η. Zoom-in view of η-SSIM curve
for c cylinder wake and d two-dimensional isotropic turbulence
likely caused by adaptive tiling in macroblocks for prediction procedures, allowing lower η with similar flow
field representation. AV1 exceeds the performance of H.264 and H.265 consistently and POD compression
on turbulent channel flow. Flow fields compressed using AV1 are indistinguishable from uncompressed flow
fields at high η.
The L 2 error and SSIM are evaluated across a range of compression ratios for each type of flow field,
as presented in Fig. 13. In general, all video compression algorithms produce asymptotically decaying L 2
error values with increasing η. AV1 performs well at low η, especially for the cylinder wake. Additionally,
all compression algorithms perform well for two-dimensional turbulence, likely as a result of the flow field
snapshots holding slow changes from one frame to the next due to the decaying nature of the flow. Moreover, as
observed with samples at various compression ratios, the L 2 error for all algorithms plateau at non-negligible
values for the cylinder wake flow field. SSIM values generally diverge from the asymptotic limit at low η.
The exceptional cases include cylinder wake and two-dimensional turbulence compressed using AV1, which
introduces negligible error at low η. AV1 outperforms H.264 and H.265 on turbulent channel flow as well, due
to the improved blocking techniques.
In addition, the time evolution of the L 2 error is examined to gain insight into the performance of video
compression algorithms for individual snapshots. The temporal evolution of the L 2 error norm ε for two-
dimensional decaying turbulence is shown in Fig. 14. H.2xx compression techniques exhibit repeated temporal
structures in its L 2 error evolution, likely as a consequence of inter-frame prediction selecting frames to make
predictions from at relatively similar intervals. AV1 compression provides a distinctive reduction in the L 2
error over time for medium and low η, indicating improved accuracy as snapshots begin to show redundancies
due to the vortex field decaying and exhibiting similar large-scale coherent structures from one snapshot to
the next. We also observe that H.264 produces a high error at low η for early flow field snapshots. This relates
to the time-varying flow nature of the present decaying turbulence, as mentioned above. The presence of
finer structures at the high Taylor Reynolds number Reλ (t) portion of the flow likely causes the difficulty in
compressing vortical flow data.
Image and video compression of fluid flow data 77
Fig. 14 L 2 error norm ε of vorticity field ω for two-dimensional decaying homogeneous isotropic turbulence over time. a H.264,
b H.265, and c AV1. (i) and (ii) in each case are chosen due to their employment in inter-frame prediction in each algorithm
Fig. 15 L 2 error norm ε of streamwise velocity field u for turbulent channel flow over time. a H.264, b H.265, and c AV1. (i)
and (ii) in each case are chosen due to their employment in inter-frame prediction in each algorithm
We also examine the L 2 error norm ε and the flow fields over time for turbulent channel flow, as depicted
in Fig. 15. H.264 generally produces a larger L 2 error compared to the other techniques, as we also observed
with the visual assessments in Fig. 12. With low η of H.264 compression, the error decreases over time, likely
as a result of a later snapshot being selected for inter-frame prediction. Compared to H.264, H.265 provides
better compression over time. Similar to the observation with H.264, the L 2 error significantly varies over time
at a low η. This is likely due to the inter-frame selection of an early frame from which further predictions were
made. AV1 produces a negligible error at a high η while the errors increase as η decreases.
We are also interested in the performance of video compression algorithms in preserving high wavenumber
structures in the compressed state. The general performance of each compression algorithm with regard to
78 V. Anatharaman et al.
Fig. 16 Kinetic energy spectra E(k) for a, d, g two-dimensional decaying homogeneous isotropic turbulence using H.264, H.265,
and AV1. b, e, h Streamwise E uu (k x+ ) and c, f, i spanwise kinetic energy spectra E uu (k z+ ) of three-dimensional turbulent channel
flow
kinetic energy spectra of each flow field is investigated, as shown in Fig. 16. H.264 performs well for two-
dimensional turbulence, but produces a noticeable error at all η in both the stream- and spanwise directions
of the kinetic energy spectrum of turbulent channel flow. A similar divergence from the expected data can be
observed at low η in the spanwise direction as well. H.265 performs comparatively well for two-dimensional
decaying isotropic turbulence, and for turbulent channel flow in the spanwise direction. However, it produces a
non-negligible error at high wavenumbers when compressed at low η in the spanwise direction. This indicates
an over-representation of high-wavenumber components due to blocking as a result of the adaptive subblock
sizes of H.265. Generally, AV1 is the best-suited algorithm for preserving spatial frequency information,
particularly at high wavenumbers, for both two and three-dimensional turbulent flow fields. At higher η, the
energy contents at each wavenumber are almost indistinguishable.
At last, we investigate whether the video compression techniques can preserve the temporal evolution
of complex turbulent flows. Here, let us examine the temporal two-point correlation for three-dimensional
channel flow compressed using all three video compression algorithms. The temporal two-point correlation
coefficient at a given t + is defined as Ruu
+ (t + )/R + (0) [69,70] and is depicted in Fig. 17. The assessment of
uu
temporal two-point correlation provides insight into the relations of flow snapshots to preceding snapshots.
Consistent with the insights gained from the kinetic energy spectrum, H.264 compression at a η exhibits
disagreement with the reference curve at t + values between 5 and 30, and above 50. This is indicative of a
de-correlation of the velocity field and is likely a result of poor performance in capturing high-wavenumber
information. Except for this particular case, all compression algorithms generally perform well, with temporal
two-point correlation coefficients closely following that of the uncompressed flow field. These results suggest
that these novel video compression techniques capture the spatiotemporal redundancies well even for complex
turbulent flows and also significantly reduce data size while preserving their physics.
Image and video compression of fluid flow data 79
Fig. 17 Normalized temporal two-point correlation coefficients Ruu (t + )/Ruu (0) for three-dimensional turbulent channel flow
using H.264, H.265, and AV1
5 Conclusion
We compressed flow field data from canonical flow examples using a number of widely available multimedia
compression techniques. The performance of the JPEG and JP2 spatial image compression techniques and the
H.264, H.265, and AV1 spatiotemporal video compression techniques were considered for simulated laminar
cylinder flow, decaying isotropic turbulence, and turbulent channel flow. Streamwise velocity and vorticity
field data were represented as grayscale images and videos, and were compressed using the aforementioned
techniques.
All techniques, with the exception of JPEG, were shown to compress flow data below 10% of the original file
size while introducing negligible error and preserving underlying flow physics. AV1 and H.265 compression
were shown to have the best performance across a variety of flow regimes. The spatial error distributions
were concentrated on the cylinder surface and directly behind the cylinder for the streamwise velocity data
compression and in the vortex shedding wake for the vorticity data. Turbulence statistics in the form of kinetic
energy spectra were preserved under compression for all methods except JPEG.
For single snapshots of data represented as an image, JP2 compression was shown to far outperform JPEG
compression, with a tolerable increase in computational complexity. For multiple temporal snapshots of data
represented as a video, the choice of compression method becomes more nuanced. JP2 compression was
shown to achieve the lowest compression error as temporal compression adds slight error to the data. The
AV1 algorithm maximizes η at the expense of computational complexity and non-negligible encoding time.
This algorithm is new and emerging from the research environment, so future optimizations could bring this
encoding time to a manageable level. The H.265 algorithm provided excellent compression performance at
a fast encoding time, and appears as a promising algorithm for current fluid dynamics applications. H.264
provided acceptable compression performance, but was largely triumphed by the AV1 and H.265 algorithms.
AV1 exhibited a significant difference of compression performance compared to conventional techniques for
complex turbulence data. Hence, it can be argued that AV1 can be especially recommended in compressing
complex turbulent flows.
We have shown that modern multimedia compression algorithms provide robust performance in a variety
of fluid flow applications. The implementation of these techniques becomes especially pertinent as simulations
within computational fluid dynamics become exceedingly data intensive, a trend that decreases the accessibility
to high-fidelity models. These methods are free, easily accessible, regularly updated and supported, and provide
flexible and scalable compression performance. Multimedia compression can also support fast data transfer
from open fluid-flow databases [71–73], promoting data-based analysis in fluid mechanics. As such, the
implementation of these compression techniques has exciting potential across the fluid dynamics community
for data storage and transfer with minimal loss.
80 V. Anatharaman et al.
Acknowledgements KT acknowledges the support from the US Army Research Office (W911NF-21-1-0060), the US Air
Force Office of Scientific Research (FA9550-21-1-0178), and the US Department of Defense Vannevar Bush Faculty Fellowship
(N00014-22-1-2798). We also thank Professor Koji Fukagata (Keio University) for sharing his DNS code.
Authors’ Contribution KT designed research. VA, JF, and KF performed research and analyzed data. VA, JF, KF, and KT wrote
the paper. KT supervised.
Data availability The data that support the findings of this study are available from the corresponding author upon reasonable
request.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use,
sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original
author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other
third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit
line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted
by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To
view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
Declarations
Conflict of interest The authors declare that they have no conflict of interest.
The increased performance of new compression algorithms comes at a cost; non-negligible increases in com-
putational complexity should be considered when implementing these algorithms. In fact, in a paper from
2000 on compressing three-dimensional flows with the JPEG and JP2 algorithms [74], the added complexity
of the JP2 algorithm caused JPEG to be recommended over JP2, despite losing clear performance benefits. The
recommendation of the present study reverses that statement. As such, it is important to quantify the encoding
time of these algorithms at the time of writing this study
The decoding time is observed to be negligibly small for all compression codecs; thus, this appendix focuses
on encoding. The streamwise velocity and vorticity data are encoded for both the laminar cylinder flow and
turbulent channel flow cases at the same bitrate (100 KB/s) for all compression algorithms and the encoding
time is measured. The encoding is performed with a 2.5GHz i7 Intel Core processor and 8 GB RAM. The
results are summarized in Table 1. Encoding time per frame is observed to be larger for the turbulent channel
flow than the laminar cylinder flow, indicating that the algorithms struggle to encode multiscale turbulent flow
data. Across encoding algorithms, JPEG and H.264 compression are the fastest, a testament to the maturity and
low complexity of these methods. JP2 and H.265 encoding are generally several times slower, but still relatively
fast, justifying their added compression performance. AV1 is observed to be far slower in encoding than the
other methods: over 100 times slower than JPEG and H.264, and over 25 times slower than JP2 and H.265. This
severe encoding time increase limits the practicality of implementing this algorithm in large-scale applications
and perhaps justifies the use of H.265 over AV1. As the algorithm was released only a few years prior to the
writing of this paper, advances in computing and algorithm development could increase its practicality in the
near future
Table 1 Encoding time (s) flow regimes, compressed at 100 KB/s bitrate
References
1. Holmes, P., Lumley, J.L., Berkooz, G., Rowley, C.W.: Turbulence, Coherent Structures, Dynamical Systems and Symmetry,
2nd edn. Cambridge University Press (2012)
2. Schmid, P.J.: Dynamic mode decomposition of numerical and experimental data. J. Fluid Mech. 656, 5–28 (2010)
3. 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), 4013–4041 (2017)
4. Wu, Z., Zaki, T.A., Meneveau, C.: Data compression for turbulence databases using spatiotemporal subsampling and local
resimulation. Phys. Rev. Fluids 5(6), 064607 (2020)
5. Liu, Y., Wang, Y., Liang, D., Wang, F., Liu, F., Lu, Y., Li, S.: A novel in situ compression method for CFD data based on
generative adversarial network. J. Vis. 22, 10 (2018)
6. Glaws, A., King, R., Sprague, M.: Deep learning for in situ data compression of large turbulent flow simulations. Phys. Rev.
Fluids 5(11), 114602 (2020)
7. Mohan, A.T., Tretiak, D., Chertkov, M., Livescu, D.: Spatio-temporal deep learning models of 3D turbulence with physics
informed diagnostics. J. Turb. 21(9–10), 484–524 (2020)
8. Momenifar, M., Diao, E., Tarokh, V., Bragg, A.D.: Dimension reduced turbulent flow data from deep vector quantisers. J.
Turb. 23(4–5), 232–264 (2022)
9. Apostolopoulos, J.G., Tan, W.-T., Wee, S.J.: Video streaming: concepts, algorithms, and systems. HP Laboratories, Report
HPL-2002-260, pp. 2641–8770 (2002)
10. Rao, A., Legout, A., Lim, Y.-S., Towsley, D., Barakat, C., Dabbous, W.: Network characteristics of video streaming traffic.
In: Proceedings of the Seventh Conference on Emerging Networking Experiments and Technologies, pp. 1–12 (2011)
11. Jiang, X., Yu, F.R., Song, T., Leung, V.C.M.: A survey on multi-access edge computing applied to video streaming: some
research issues and challenges. IEEE Commun. Surv. Tutor. 23(2), 871–903 (2021)
12. Egido, C.: Video conferencing as a technology to support group work: a review of its failures. In: Proceedings of the 1988
ACM Conference on Computer-Supported Cooperative Work, pp. 13–24 (1988)
13. Augestad, K.M., Lindsetmo, R.O.: Overcoming distance: video-conferencing as a clinical and educational tool among
surgeons. World J. Surg. 33(7), 1356–1365 (2009)
14. Mpungose, C.B.: Lecturers’ reflections on use of Zoom video conferencing technology for e-learning at a South African
university in the context of coronavirus. Afr. Ident. 1–17 (2021)
15. Said, A., Pearlman, W.A.: An image multiresolution representation for lossless and lossy compression. IEEE Trans. Image
Process. 5(9), 1303–1310 (1996)
16. Liu, F., Hernandez-Cabronero, M., Sanchez, V., Marcellin, M.W., Bilgin, A.: The current role of image compression standards
in medical imaging. Information 8(4), 131 (2017)
17. Arora, K., Shukla, M.: A comprehensive review of image compression techniques. Int. J. Comput. Sci. Inf. Technol. 5(2),
1169–1172 (2014)
18. Guo, J., Chao, H.: Building dual-domain representations for compression artifacts reduction. In: European Conference on
Computer Vision, pp. 628–644. Springer (2016)
19. Schmalzl, J.: Using standard image compression algorithms to store data from computational fluid dynamics. Comput.
Geosci. 29(8), 1021–1031 (2003)
20. Ahmed, N., Natarajan, T., Rao, K.R.: Discrete cosine transform. IEEE Trans. Comput. 100(1), 90–93 (1974)
21. Ahmed, N.: How I came up with the discrete cosine transform. Digit. Signal Process. 1(1), 4–5 (1991)
22. Mitchell, J.: Digital compression and coding of continuous-tone still images: requirements and guidelines. ITU-T Recom-
mend. Trans. 81 (1992)
23. Taubman, D., Marcellin, M.: JPEG2000 Image Compression Fundamentals, Standards and Practice: Image Compression
Fundamentals, Standards and Practice, vol. 642. Springer (2012)
24. Schneider, K., Vasilyev, O.V.: Wavelet methods in computational fluid dynamics. Annu. Rev. Fluid Mech. 42, 473–503
(2010)
25. Hoffman, R.: Data Compression in Digital Systems. Springer (2012)
26. Fisher, Y.: Fractal image compression. Fractals 2(03), 347–361 (1994)
27. Fisher, V.: Fractal Image Compression: Theory and Application. Springer (2012)
28. Mallat, S.G., Zhang, Z.: Matching pursuits with time-frequency dictionaries. IEEE Trans. Signal Process. 41(12), 3397–3415
(1993)
29. Tudor, P.N.: MPEG-2 video compression. Electron. Commun. Eng. J. 7(6), 257–264 (1995)
30. Haskell, B.G., Puri, A., Netravali, A.N.: Digital Video: An Introduction to MPEG-2. Springer (1996)
31. Bosi, M., Brandenburg, K., Quackenbush, S., Fielder, L., Akagiri, K., Fuchs, H., Dietz, M.: ISO/IEC MPEG-2 advanced
audio coding. J. Audio Eng. Soc. 45(10), 789–814 (1997)
32. Wiegand, T., Sullivan, G.J., Bjontegaard, G., Luthra, A.: Overview of the H.264/AVC video coding standard. Circuits Syst.
Video Technol. 13(7), 560–576 (2003). (IEEE Trans)
33. Pastuszak, G., Abramowski, A.: Algorithm and architecture design of the H.265/HEVC intra encoder. Circuits Syst. Video
Technol. 26(1), 210–222 (2015). (IEEE Trans)
34. Chen, Y., Murherjee, D., Han, J., Grange, A., Xu, Y., Liu, Z., Parker, S., Chen, C., Su, H., Joshi, U., Chiang, C.-H., Wang,
Y., Wilkins, P., Bankoski, J., Trudeau, L., Egge, N., Valin, J.-M., Davies, T., Midtskogen, S., Norkin, A., de Rivaz, P.: An
overview of core coding tools in the AV1 video codec. In: 2018 Picture Coding Symposium (PCS), pp. 41–45 (2018)
35. Han, J., Li, B., Mukherjee, D., Chiang, C.-H., Grange, A., Chen, C., Su, H., Parker, S., Deng, S., Joshi, U., Chen, Y., Wang,
Y., Wilkins, P., Xu, Y.: A technical overview of AV1. Proc. IEEE 109(9), 1435–1462 (2021)
36. FFmpeg.: A complete, cross-platform solution to record, convert and stream audio and video (2022). https://www.ffmpeg.
org/. Accessed 25 Nov 2022
37. Wallace, G.K.: The JPEG still picture compression standard. IEEE Trans. Consum. Electron. 38(1), 34 (1992)
38. Sze, V., Marpe, D.: Entropy coding in HEVC. In: High Efficiency Video Coding (HEVC), pp. 209–274. Springer (2014)
82 V. Anatharaman et al.
39. Duda, J., Tahboub, K., Gadgil, N.J., Delp, E.J.: The use of asymmetric numeral systems as an accurate replacement for
Huffman coding. In: 2015 Picture Coding Symposium (PCS), pp. 65–69. IEE (2015)
40. Shannon, C.E.: A mathematical theory of communication. Bell Syst. Tech. J. 27(3), 379–423 (1948)
41. Huffman, D.A.: A method for the construction of minimum-redundancy codes. Proc. IRE 40(9), 1098–1101 (1952)
42. Knuth, D.E.: Fundamental algorithms (1973)
43. Daubechies, I.: Orthonormal bases of compactly supported wavelets. Commun. Pure Appl. Math. 41(7), 909–996 (1988)
44. Mallat, S.G.: Multiresolution approximations and wavelet orthonormal bases of l 2 (r ). Trans. Am. Math. Soc. 315(1), 69–87
(1989)
45. Ricker, N.: Wavelet contraction, wavelet expansion, and the control of seismic resolution. Geophysics 18(4), 769–792 (1953)
46. Heil, C.E., Walnut, D.F.: Continuous and discrete wavelet transforms. SIAM Rev. 31(4), 628–666 (1989)
47. Rissanen, J., Langdon, G.G.: Arithmetic coding. IBM J. Res. Dev. 23(2), 149–162 (1979)
48. List, P., Joch, A., Lainema, J., Bjontegaard, G., Karczewicz, M.: Adaptive deblocking filter. IEEE Trans. Circuits Syst. Video
Technol. 13(7), 614–619 (2003)
49. Sullivan, G.J., Ohm, J.-R., Han, W.-J., Wiegand, T.: Overview of the high efficiency video coding (HEVC) standard. IEEE
Trans. Circuits Syst. Video Technol. 22(12), 1649–1668 (2012)
50. Midtskogen, S., Valin, J.-M.: The AV1 constrained directional enhancement filter (CDEF). In: 2018 IEEE International
Conference on Acoustics, Speech and Signal Processing (ICASSP), pp. 1193–1197. IEEE (2018)
51. Taira, K., Colonius, T.: The immersed boundary method: a projection approach. J. Comput. Phys. 225(2), 2118–2137 (2007)
52. Colonius, T., Taira, K.: A fast immersed boundary method using a nullspace approach and multi-domain far-field boundary
conditions. Comput. Methods Appl. Mech. Eng. 197, 2131–2146 (2008)
53. Taira, K., Nair, A.G., Brunton, S.L.: Network structure of two-dimensional decaying isotropic turbulence. J. Fluid Mech.
795, R2 (2016)
54. Fukagata, K., Kasagi, N., Koumoutsakos, P.: A theoretical prediction of friction drag reduction in turbulent flow by super-
hydrophobic surfaces. Phys. Fluids 18, 051703 (2006)
55. Fukami, K., Fukagata, K., Taira, K.: Machine-learning-based spatio-temporal super resolution reconstruction of turbulent
flows. J. Fluid Mech. 909, A9 (2021)
56. Moser, R.D., Kim, J., Mansour, N.N.: Direct numerical simulation of turbulent channel flow up to Reτ = 590. Phys. Fluids
11(4), 943–945 (1999)
57. Kim, J., Moin, P., Moser, R.: Turbulence statistics in fully developed channel flow at low Reynolds number. J. Fluid Mech.
177, 133–166 (1987)
58. Stewart, G.W.: On the early history of the singular value decomposition. SIAM Rev. 35(4), 551–566 (1993)
59. Smith, C.R., Metzler, S.P.: The characteristics of low-speed streaks in the near-wall region of a turbulent boundary layer. J.
Fluid Mech. 129, 27–54 (1983)
60. Wang, Z., Bovik, A.C., Sheikh, H.R., Simoncelli, E.P.: Image quality assessment: from error visibility to structural similarity.
IEEE Trans. Image Process. 13(4), 600–612 (2004)
61. Lumley, J.L.: The structure of inhomogeneous turbulent flows. In: Yaglom, A.M., Tatarski, V.I. (eds.) Atmospheric Turbulence
and Radio Wave Propagation. Nauka (1967)
62. Taira, K., Hemati, M.S., Brunton, S.L., Sun, Y., Duraisamy, K., Bagheri, S., Dawson, S., Yeh, C.-A.: Modal analysis of fluid
flows: applications and outlook. AIAA J. 58(3), 998–1022 (2020)
63. Sirovich, L.: Turbulence and the dynamics of coherent structures I. Coherent structures. Q. Appl. Math. 45(3), 561–571
(1987)
64. McWilliams, J.C.: The emergence of isolated coherent vortices in turbulent flow. J. Fluid Mech. 146, 21–43 (1984)
65. Yeh, C.-A., Gopalakrishnan Meena, M., Taira, K.: Network broadcast analysis and control of turbulent flows. J. Fluid Mech.
910, A15 (2021)
66. Alfonsi, G., Primavera, L.: The structure of turbulent boundary layers in the wall region of plane channel flow. Proc. R. Soc.
A 463(2078), 593–612 (2007)
67. Muralidhar, S.D., Podvin, B., Mathelin, L., Fraigneau, Y.: Spatio-temporal proper orthogonal decomposition of turbulent
channel flow. J. Fluid Mech. 864, 614–639 (2019)
68. Fukami, K., Nakamura, T., Fukagata, K.: Convolutional neural network based hierarchical autoencoder for nonlinear mode
decomposition of fluid field data. Phys. Fluids 32(9), 095110 (2020)
69. Fukami, K., Nabae, Y., Kawai, K., Fukagata, K.: Synthetic turbulent inflow generator using machine learning. Phys. Rev.
Fluids 4, 064603 (2019)
70. Quadrio, M., Luchini, P.: Integral space-time scales in turbulent wall flows. Phys. Fluids 15(8), 2219–2227 (2003)
71. Li, Y., Perlman, E., Wan, M., Yang, Y., Meneveau, C., Burns, R., Chen, S., Szalay, A., Eyink, G.: A public turbulence database
cluster and applications to study Lagrangian evolution of velocity increments in turbulence. J. Turb. 9, 31 (2008)
72. Wu, X., Moin, P.: A direct numerical simulation study on the mean velocity characteristics in turbulent pipe flow. J. Fluid
Mech. 608, 81–112 (2008)
73. Towne, A., Dawson, S., Brès, G.A., Lozano-Durán, A., Saxton-Fox, T., Parthasarathy, A., Jones, A.R., Biler, H., Yeh, C.-A.,
Patel, H.D., Taira, K.: A database for reduced-complexity modeling of fluid flows. AIAA J. (2023)
74. Schmalzl, J.: Using standard image compression algorithms to store data from computational fluid dynamics. Comput.
Geosci. 29(8), 1021–1031 (2003)
Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional
affiliations.