Estimation of motion using a correlation-based method and an “optical-flow”

advertisement
Estimation of motion using a PIV correlation-based method and an “optical-flow”
one for two experimental flows : quantitative and qualitative comparison.
Thomas Corpetti1 , Dominique Heitz1 , Georges Arroyo1 , Étienne Mémin2 , Alina Santa-Cruz1
1
: Cemagref, 17 av de Cucillé, CS64427, 35044 Rennes Cedex, France
2
: IRISA/INRIA, Campus Universitaire de Beaulieu, 35042 Rennes Cedex, France
E-mail : dominique.heitz@cemagref.fr
abstract
Families of ”Optical Flow” methods are frequently used in the Computer Vision community to estimate
the velocity from a pair of images, because of their ability to estimate a reliable and dense (i.e. one vector
per pixel) motion field.
Even if many extensions of usual PIV correlation-based methods are able to obtain more and more
information associated to a “sub pixel” precision, let us remark that it is always difficult to estimate correctly
the motion when some particular event appears such as a loss of pairing in the interrogation area, an image
truncation due to its finite size, a velocity gradient in the interrogation window, ... Furthermore, only the
“most probable” displacement is extracted.
From this point of view, the idea of using a motion estimation technique based on optical-flow for fluid
images is appealing.
In this paper, a qualitative and quantitative study is done to compare motion fields obtained from an
usual PIV method and with a particular optical-flow method, dedicated to fluid motion. A mixing layer and
the near wake of a circular cylinder are analyzed and significant parameters are extracted for both methods.
Statistical results are compared with hot-wire anemometry measurements.
1
15
1
10
0.5
5
0
y/D
y/D
0.5
-5
0
-10
-15
0
-0.5
-0.5
-1
-0.5
-1
0
0.5
1
1.5
2
x/D
2.5
3
3.5
4
0.5
4.5
1
1.5
2
2.5
3
3.5
4
x/D
1
15
1
10
0.5
5
0
y/D
y/D
0.5
-5
0
-10
-15
0
-0.5
-0.5
-1
-0.5
-1
0
0.5
1
1.5
2
x/D
2.5
3
3.5
4
4.5
0.5
1
1.5
2
2.5
3
3.5
4
x/D
F IG . 1 – View of instantaneous near wake vortices at Re=3000 – TOP : Optical-flow treatement : left : vector field ; right : map
of iso-vorticity ; BOTTOM : piv treatement : left : vector field ; right : map of iso-vorticity
1
Introduction
In a certain number of domains requiring experimental fluid flows visualization and measurements, Particle Image Velocimetry (PIV) techniques or related methods plays an important role and are intensively used to provide motion measurements.
These techniques are based on a spatio-temporal cross-correlation under a consistancy assumption of the flow within a local interrogation window. The interrogation windows contains generally several particle features with different motions. Considering
a unique motion vector for all these particles may lead in some circonstances to very poor local motion representation.
In the Computer Vision community, “optical flow” methods enable to extract a dense representation of the motion field (i.e.
one vector per pixel). These methods, introduced by Horn & Schunck Horn & Schunck (1981) in the early eighties, consist to
estimate a vectorial function by minimizing an objective functional. This functionnal is composed of two terms. The first one, the
data model, is an adequacy term between the unknown motion function and the data. It generally relies on brightness consistancy
assumption. Similarly to correlation techniques, this assumption state that a given point keeps its intensity along its trajectory.
The second term, named “regularization” one, promotes a global smoothness of the motion field over the image plane.
These techniques have been devised for the analysis of quasi-rigid motions with stable salient features. Unfortunately, there
is no such thing as an “object” in most imaged fluid flows which contain mainly deformable and transient brightness patterns. On
such sequences, techniques based on standard computer vision ingredients are thus intrinsically limited. The design of alternative
approaches dedicated to fluid motion thus remains a widely open domain of research.
In this paper, we present the basis of an optical-flow method which permits to extract reliable dense motion fields. This
method is dedicated to the analysis of fluid flows image sequences. It includes in the data-model term and in the smoothness term
some general considerations about the flow motion in the image plane.
This article is organized as follows. In a first section, we present the standrad approaches frequently used by the fluid mechanical community to extract velocity measurements from PIV images. We then present, in a second section, the proposed alternative.
In a last part, experiments on two classical flows (a mixing layer and the near wake of a circular cylinder) are presented and
analysed. We provides some elements of comparizon of our method with a state of the art industial PIV system.
2
PIV approaches
Since the 70s, numerous methods have been proposed to extract the velocity of the flow from particle images. See for instance
Adrian (1991); Raffel et al. (2000) for an extended bibliography on the subject. In the next sections, we will further describe
common PIV approaches.
2.1
Computation of the motion field
Most of the PIV systems relies on correlation functions such as “auto-correlation” (if only one image containing the traces of
particles is available) or “cross-correlation” (if the temporal information is separated in two images). Due to its better stability
against noise and as it provides a complete information on the motion (velocity magnitude and direction), this latter technique is
usually prefered.
The image frames are interrogated by the computation of the spatial cross-correlation in a M × N interrogation region :
R(d) =
1
kW(x)|
X
I1 (y) × I2 (y + d(x)),
(1)
y∈W(x)
where I1 and I2 are the first and second discrete image frames, W(x) being a window centered on the point x. The motion
associated to the center of the interrogation window d(x) = (u(x), v(x)) corresponds to a pic of the correlation surface.
As they lead to real time computations, it is commom to compute all correlations in the Fourier space using a Fast Fourier
Transform FFT technique. These techniques requires to be applied to 2p × 2q periodic images. In practice, this condition is often
not verified and the original images have to be accordingly adjusted. To prevent aliasing artifacts, Nyquist criteria must be in
addition respected : the considered displacement d must be such that d < (N/2, M/2). If necessary, a lowering of the velocity
magnitude is commonly performed by growing up the size of the interrogation window or by decreasing the time interval between
two laser pulses. These manipulations may affect significantly the quality of the estimated motion field and a compromise has
to be generally defined between a direct computation which do not have such restrictions and fast computations in the Fourier
spaceMcKenna & McGillis (2002).
Another inherent difficulty of correlation method lies in the discrete data nature. As a matter of fact, with such approaches
the estimated displacement vector is intrinsically discretized on the image grid. Infinitesimal brightness variation can indeed not
be identified. This results in the well known “peak-locking” effect which spoil the accuracy of the velocity measurement. Many
studies has been done to attenuate this phenomenon. One possibility consists to increase particle size (whose diameter must
be 2 or 3 pixels in the image) or to pay a particular attention to the correlation maxima localization (by sound interpolations
for example). Several methods have been proposed to provide “sub-pixel” accuracy Lourenco & Krothapalli (1995); Willert &
Gharib (1991).
Actually, many extensions of the basic cross-correlation principle have been proposed to improve the estimation quality in
terms of dynamic rank, accuracy and spatial resolution. For example, Scarano & Reithmuller Scarano & Reithmuller (2000)
and Leucona et al Lecuona et al. (2002) use an iterative prediction of the unknown velocity in the estimation scheme, whereas
Lourenco & Krothapalli Lourenco & Krothapalli (2000) or Wereley & Meinhart Wereley & Meinhart (2001) directly modified
2
the correlation criterion, using a central difference approximation (of order 2) instead of a forward difference approximation (of
order 1) as commonly done. One can also note that the use of a super-resolution technique as in Nogueira et al. (2001) allows an
uncertainty reduction and attenuates the effect of peak-locking.
2.2
Discussion
Even if the principle of correlation techniques authorizes to extract only displacements vectors on the image lattice (i.e.
vectors of the type d = (u × pix, v × pix), where u and v are integers and pix is the pixel size), many extended methods have
been proposed to achieve “sub pixel” accuracy. All these correlation techniques shares nevertheless some common limitations
which prevents a comfortable use Lourenco & Krothapalli (2000) :
1. due to the finite size of the interrogation area, a “loss of pairing” may alter the estimation. In this case, the maximum of
correlation does not corresponds to the actual motion. The definition of the size of the interrogation window is indeed very
problematic to extract a relevant motion in accordance with the observed phenomenon ;
2. The existence of “velocity and speeding gradient in the interrogation region” introduces a bias towards the lower displacements and higher seeded sub-regions as a result of the more frequent pairings. For large interrogation windows, this is
very problematic ;
3. due to its statistical nature, only the “most probable” displacement is extract.
All these limitations can hardly be avoided as they are intrinsically linked to the dicrete and local nature of correlation
methods.
Optical-Flow approaches, originally introduced by Horn & Schunck Horn & Schunck (1981), enable to extract in a natural
way a dense motion field (i.e. one vector per pixel). Such techniques take into account the whole image plane to estimate
a vectorial continuous function representing the velocity field. As a consequence they should allow intrinsically to avoid the
difficulties mentionned above. The idea which consists to use optical-flow methods for PIV images appears then very attractive.
Unfortunately such techniques designed for computer vision applications have also some limitations which prevents their direct
use in an experimental fluid mechanics context.
In the following section, we present further the basis of generic optical-flow schemes. We then discuss their limitations
with respect to fluid image sequences and present an extension of such techniques specifically dedicated to fluid flows velocity
measurements.
3
3.1
Optical flow scheme
Generic optical-flow estimator
The most accurate techniques to address the generic problem of estimating the apparent motion from image sequences are
based on the seminal work of Horn and Schunck Horn & Schunck (1981). These techniques are based on the minimization of a
global cost function H composed of two terms. The first one, named “observation term”, is derived from a brightness constancy
assumption and assumes that a given point keeps the same intensity along its trajectory. It is expressed through the well known
optical flow constraint equation (OFCE) :
ZZ
∂E(x, t)
dx,
(2)
Hobs (E, v) =
f1 ∇E(x, t) · v(x, t) +
∂t
Ω
where v(x, t) = (u, v)T is the unknown velocity field at time t and location x = (x, y) in the image plane Ω, E(x, t) being
the image brightness, viewed for a while as a continous function. This first term relies on the assumption that the visible points
conserve roughly their intensity in the course of a small displacement.
∂E
dE
= ∇E · v +
≈ 0.
dt
∂t
(3)
The associated penalty function f1 is often the L2 norm. However, better estimates are usually obtained by choosing a “softer”
penalty function Black (1994); Mémin & Pérez (1998). Such functions, arising from robust statistics Huber (1981), limit the
impact of the many locations where the brightness constancy assumption does not hold, such as on occlusion boundaries.
This single (scalar) observation term does not allow to estimate the two components u and v of the velocity. In order to solve
this ill-posed problem, it is common to employ an additional smoothness constraint Hreg . Usually, this second term enforces a
spatial smoothness coherence of the flow field. It relies on a contextual assumption which enforces a spatial smoothness of the
solution. This term usually reads :
ZZ
Hreg (v) =
f2 |∇u(x, t)| + |∇v(x, t)| ,
(4)
Ω
As with the penalty function in the data term, the penalty function f2 was taken as a quadratic in early studies, but a softer penalty
is now preferred in order not to smooth out the natural discontinuities (boundaries, ...) of the velocity field Black (1994); Cohen
3
& Herlin (1999); Kornprobst et al. (1999); Mémin & Pérez (1998). Based on (2) and (4), the estimation of motion can be done
by minimizing :
ZZ
ZZ
∂E(x, t)
H(E, v) =
f1 ∇E(x, t) · v(x, t) +
dx + α
f2 |∇u(x, t)| + |∇v(x, t)| ,
(5)
∂t
Ω
Ω
where α > 0 is a parameter controlling the balance between the smoothness constraint and the global adequacy to the observation
assumption.
It is important to outline that such an estimator defined as the minimizer of H is generic. It is only based on the assumption of
luminance conservation and of first-order spatial smoothness of the motion. Even if this kind of estimator has been successfully
used for fluid motion Bannehr et al. (1996); Cohen & Herlin (1999); Larsen et al. (1998); Mémin & Pérez (1999); Wallace
& Foss (1995), the uderlying assumption are far to be sound hypothesis for such an applicative domains. As a matter of fact,
image sequences representing fluid phenomena exhibit areas where the luminance function undergoes high temporal variations
along the motion. These areas are often the center of tridimensional motions that cause the appearance or the disappearance of
fluid matter within the bidimensional visualization plane. These regions are in additon associated to divergent motions which
influence greatly the shape of the velocity field in large surrounding areas. An accurate estimation of the 2D apparent motion in
such regions is therefore of the highest importance and is hardly possible with the optical-flow constraint.
3.2
Fluid dedicated optical-flow estimator
In a recent paper Corpetti et al. (2002), we have proposed an optical-flow techniques specifically dedicated to image sequences
depicting fluid phenomenon. This specialized optical-flow estimator relies on adaptation of the functionnal data-model and
smoothness term.
3.3
Continuity equation and data model
Instead of sticking to the intensity conservation assumption, the data-model that has been considered relies on the fluid law
of mass conservation :
∂χ
+ div(χv) = 0,
(6)
∂t
∂v
∂w
where χ denotes the density of the fluid, v its 3D velocity and divv = ∂u
∂x + ∂y + ∂z stands for the divergence of the vector field
v = (u, v, w). Simple manipulations yield the alternative rewriting :
dχ
+ χdivv = 0.
(7)
dt
When the divergence of the 3D apparent flow vanishes, this equation is of the same form as the 2D optical flow constraint on
luminance. The continuity equation, originally introduced in Schunk (1984) as a data model for motion estimation of intensity
time varying images, has been since successfully incorporated in several works Béréziat et al. (2000); Corpetti et al. (2002);
Zhou et al. (2000); Wildes et al. (1997); Song & Leahy (1991); Amini (1994).
The use of continuity equation for image sequences analysis relies on two hypotheses. First, the luminance function is
assumed to be directly related to a passive quantity transported by the fluid. Secondly, the continuity equation which holds in
3D, is assumed to hold as well for the bidimensional motion field captured by the image sequence. This latter assumption has
been theoretically established in the case of transmittance imaging by Fitzpatrick Fitzpatrick (1988) and extended by Wildes
et al in Wildes et al. (1997). As the brightness consistancy is obviously not verified in PIV images, the equation of continuity
provides us with an interesting alternative data-model. Instead of expressing a point-wise conservation of the luminance along
the motion, this alternative model assumes the conservation of the total luminance of any moving elements of the image. The
new optical-flow constraint then reads :
ZZ
dE(x, t)
Hobs (d) =
f1
+ E(x, t)div v(x, t) .
(8)
dt
Ω
It is important to note that, due to its differential nature, this new constraint based on the continuity equation is not valid in
case of large displacements. To handle large displacements, one can use this expression in an integrated way and solve the new
resulting problem through a succession of linearizations embedded within a multiresolution scheme. In Corpetti et al. (2002), the
reader will find details of these methods.
We now turn to the definition of the smoothness prior to be used in conjunction with this new data energy term.
3.4
Second order Div-Curl regularization
By using Euler-Lagrange conditions of optimality, it is readily demonstrated that the standard first-order regularization functional :
ZZ
Hreg (v) =
|∇u(x)|2 + |∇v(x)|2 dx
Ω
4
is equivalent from the minimization point of view, to the so-called div-curl regularization functional Suter (1994) :
ZZ
Hreg (v) =
div2 v(x) + curl2 v(x) dx,
(9)
Ω
where div and curl are respectively the divergence and the vorticity of the motion field v = (u, v).
A first-order regularization therefore penalizes the amplitude of both the divergence and the vorticity of the vector field. For
fluid motion estimation, this does not seem appropriate since the apparent velocity field usually exhibits compact areas with high
value of vorticity and/or divergence. Then, it would seem more appropriate to rely on second-order div-curl regularization Gupta
& Prince (1996); Suter (1994) :
ZZ Hreg (d) =
(10)
|∇divv(x)|2 + |∇curlv(x)|2 dx.
Ω
This regularization tends to preserve the divergence and the vorticity of the displaced motion field v to estimate. In Corpetti et al.
(2002), the reader will find numerical details for the implementation of this fluid motion estimator.
4
4.1
Experimental details
Plane turbulent mixing layer
The mixing layer studied here has been generated in a closed circuit, subsonic wind tunnel R300 of the CEAT (Centre
d’Études Aérodynamiques et Thermiques, University of Poitiers, France). This wind tunnel has a square test section of crosssectional dimensions 300 mm × 300 mm and length 2 m, with a contraction ratio of 10. The one stream blow generated by a
fan is divided with a splitter plate and flows through two different head loss devices, followed by a fine mesh screen, to give a
two-stream flow with low turbulence level. Tripping wire were used on both sides of the splitter plate to ensure rapid and uniform
transition of the boundary layer well upstream of the trailing edge. Hence, the mixing layer is initially turbulent. Wind-tunnel
ceiling was tilted to achieve a negligible streamwise pressure gradient. For more details about the facility, the reader can refer
to Heitz (1999). The velocity ratio of the two streams r = Ua /Ub (Ua = 9 m s−1 and Ub = 6 m s−1 ) is 0.67 with an average
advective velocity Um = (Ua + Ub )/2 of about 7.5 m s−1 . The turbulence intensity corresponding to free stream velocity is less
than 0.3%. The location of the origin of the x axis satisfies the condition δω = 15 mm and was found 16δω downstream of the
trailing edge of the splitter plate. In all the figures this value was employed to reduce x and y coordinates.
4.2
Wake of a circular cylinder
The experiments in the near wake of a circular cylinder were carried out in a small opened wind tunnel having a test section
142 mm wide 142 mm high, and 1100 mm long. The wind tunnel consist of a blower supplying air to a conditioning system
composed of a honeycomb and a porous media (foam) followed by a 2 :1 contraction, which provided a free-stream uniformity
and a turbulence intensity less than 0.1%. The flow velocity was 4.5 m s−1 . The circular cylinder of cross-sectional diameter
D = 10 mm and of L = 142 mm long, has an aspect ratio of L/D = 14.2. The Reynolds number based on the diameter D was
of Re = 3000. The cylinder was mounted vertically and not bounded with end plates.
4.3
PIV settings
For the two flow configurations, data were acquired from two differents PIV systems.
For the mixing layer the illumination setup comprised a Nd :YAG double-pulses laser system (Quantel) with an output
energy of 30 mJ per pulse at a green wavelength λ = 532 mm. The flow was seeded with particles of oil. The spray generator
was located downstream of the test section in the closed loop of the wind tunnel, to ensure good homogeneity of the seeding in
the view plane. The images were captured by a CCD Kodak camera (type 700) of 1008 × 984 pixels resolution and 8 bit dynamic
range. The camera was placed so that the middle-height of the image corresponded to the height of the splitter plate, and so that
the vorticity thickness of the mixing layer (δω ) at the inlet of the image was equal to 15 mm. The image size in the physical
space was Lx × Ly = 84.5 × 82.5 mm2 = 5.6δω × 5.5δω . The laser pulse rate was adjustable up to 20 Hz but the acquisition
frequency was limited to 15 Hz by the camera frame rate. The pulse delay was fixed at ∆t = 50µs in order to keep particle
displacements in a reasonable range. Synchronization of the camera apertures with the laser pulses was achieved using LaVision
hardware system. The software of LaVison, Davis, was also used to calculate the cross-correlation of the two pairs of grey level
images acquired by the camera. For comparison with optical flow results two interrogation procedures, named PIV I and PIV II
respectively, were applied. A 32 × 32 pixels interrogation window with 50% overlap (PIV I) was used, leading to a grid spacing
of 16 × 16 pixels (1.34 × 1.34 mm2 ), and a multipass technique with two refinement steps and 50% overlap (PIV II) was used to
obtain interrogation regions as small as 16 × 16, yielding a grid spacing of 8 × 8 pixels (0.67 × 0.67 mm2 i.e. 0.05 × 0.05 δω2 ).
For the wake of the circular cylinder the illumination setup comprised a Nd :YAG double-pulses laser system (New Wave)
with an output energy of 30 mJ per pulse at a green wavelength λ = 532 mm. The flow was seeded with particles of oil. The
spray generator was located upstream of the blower, to ensure good homogeneity of the seeding in the view plane. The images
were captured by a CCD LaVision camera of 1280 × 1024 pixels resolution and 12 bit dynamic range. The image size in the
physical space was Lx × Ly = 98.7 × 83.4 mm2 = 9.9D × 8.3D. The laser pulse rate was adjustable up to 20 Hz but the
acquisition frequency was limited to 15 Hz by the camera frame rate. The pulse delay was fixed at ∆t = 69µs in order to keep
5
particle displacements in a reasonable range. Synchronization of the camera apertures with the laser pulses was achieved using
LaVision hardware system. The software of LaVison, Davis, was also used to calculate the cross-correlation of the two pairs of
grey level images acquired by the camera. A multipass technique with refinement steps was used to obtain interrogation regions
as small as 16 × 16 pixels, leading to a grid spacing of 12 × 12 pixels (0.93 × 0.93 mm2 i.e. 0.093 × 0.093 D2 ).
Both instantaneous and averaged quantities are employed to represent the flow structure. Mean quantities were computed with
the whole sequence of 520 and 540 instantaneous vector fields for the mixing layer and the circular cylinder wake respectively.
4.4
Optical flow settings
For the optical-flow approach, regularization parameters used are α = 300 and λ = 300. For the penalty function f1 , we
prefer to choose the Leclerc penalty function f1 (x) = 1 − exp(−τ1 x2 ) instead of the quadratic one in reason of the nonsystematic correspondence between the brightness intensity and the concentration of particles. Parameter τ1 has then to be set
and we choose its value as τ1 = 1.6. The choice of these three parameters has been done once on synthetic images Corpetti et al.
(2002) and we used to keep the same set of parameters values what ever the input data-s are. For f2 penalty function dedicated
to the div-curl regularization, we choose a quadratic penalisation function as we have a physical spatial-continuity both on the
divergence and the vorticity of the flow.
Optical-flow technique yields a grid spacing of 1 × 1 pixel. Therefore the values of spatial resolution in the plane of the laser
sheet, for the mixing layer and the circular cylinder wake, were 0.08 × 0.08 mm2 i.e. 0.006 × 0.006 δω2 and 0.08 × 0.08 mm2
i.e. 0.008 × 0.008 D2 respectively.
5
5.1
Results for the mixing layer
Instantaneous velocity field
Images of instantaneous vectors fields are shown in figure 2-(a) and in figure 3-(a), for Optical-flow and PIV approach
respectively. Only one vector out of eight have been represented for Optical-flow. Results exhibits the footprint of the primary
structures of the mixing layer. Based on vector fields both approaches give comparable informations but with higher density
for Optical-flow. Figure 2-(b) and 3-(b) show contours of constant spanwise vorticity ωz? . It was noted in the litterature that the
error in the measurement of the vorticity depends on the truncation error associated with the finite difference scheme used, and
on the measurement uncertainty, ε, which is directly proportional to the grid spacing, that is ε/∆x (Lourenco & Krothapalli
(1995); Raffel et al. (2000)). Furthermore, for PIV approach, if the interrogation window overlap exceed 50%, the velocities
are correlated, causing the differential estimates to be biased. In the present study the average vorticity was estimated within an
enclosed area of 3 × 3 pixels by calculating the local circulation around this area (Raffel et al. (2000)). When compared with
PIV results, Optical-flow approach seems to exhibit more localized informations. Nevertheless, vorticity yields noisy contours
and higher levels of vorticity (see Fig. 2-(c) and 3-(c)) which are due to the measurement noise. The quality of the input images
can explain these noisy vorticity contours : for the mixing layer, a 8 bit CCD Kodak camera was used whereas a 12 bit CCD La
Vision one was employed for the study of the wake of the circular cylinder. In the §6, related to the latter experiment, one can see
that vorticity maps extracted are significantly less noisy.
1
y/δω
0.5
0
-0.5
-1
-1
0
1
2
x/δω
3
4
5
-1
0
1
2
x/δω
3
4
5
(a)
1
y/δω
0.5
0
-0.5
-1
(b)
F IG . 2 – Views of instantaneous turbulent plane mixing layer structure at Reδω = 7000. Optical-flow : (a), Instantaneous vector
field (one vector out of eight were plotted) ; (b), Instantaneous spanwise isocontour of vorticity ωz? = (ωz ∆U )/δω (ωz?min = −9,
ωz?max = −3, ∆ωz? = 0.5).
6
1
y/δω
0.5
0
-0.5
-1
-1
0
1
2
x/δω
3
4
5
-1
0
1
2
x/δω
3
4
5
(a)
1
y/δω
0.5
0
-0.5
-1
(b)
F IG . 3 – Views of instantaneous turbulent plane mixing layer structure at Reδω = 7000. PIV II - multipass 32-16 treatement : (a),
Instantaneous vector field ; (b), Instantaneous spanwise isocontour of vorticity ωz? = (ωz ∆U )/δω (ωz?min = −9, ωz?max = −1,
∆ωz? = 0.6).
5.2
Mean quantities
In a mixing layer, the mean streamwise velocity component can be expressed with the theoretical solution as :
U − Ub
1
= (1 − erf(ση))
Ua − Ub
2
(11)
where η = (y − yo )/(x − xo ) with (xo , yo ) the coordinate of virtual origin of the mixing layer and the spreading parameter σ is
constant.
For each streamwise location x and for both approaches the experimental values fit the theoretical solution. Results gathered
in table 1 show that the spreading parameter σ adjusted from Optical-flow and PIV II treatement was of the order of 49.8 and 43.6
respectively. Optical-flow value compare well with the value of 52.7 obtained by Heitz (1999) with pitot tube measurements,
while PIV II gives poor agreement. Whereas, for both approaches, the evolution of the vorticity thickness δω , defined as δω =
−∆U/(∂U/∂y)max where ∆U = Ua − Ub , supports the observation of a linear growth of the mixing layer. As for the spreading
parameter PIV II results exhibit a higher growth of the mixing layer. This behaviour can be explained by the interrogation interval
in which the velocity estimated is an average value, in a region where there are high displacements gradients.
r
λ
σ
dδw /dx
Heitz (1999)
0.67
0.2
52.7
0.0336
PIV
0.67
0.2
43.55
0.0407
optical-flow
0.67
0.2
49.78
0.0356
TAB . 1 – Comparison on main charcateristics parameters of the mixing layer for Pitot-tube measurements (Heitz (1999)), PIV
and Optical-flow approaches
II
To have a finer analysis, it is of primary interest to observe the superior order moments. To that end, the distribution of
the Reynolds stresses are represented in figure 4. Data are compared with Heitz (1999) fourth hot-wire measurements. Results
show that PIV I approach overestimates u02 and understimates v 02 and u0 v 0 . PIV II approach gives better results, with a slight
overestimation of u02 and u0 v 0 and a slight underestimation of v 02 . Optical-flow approach seems to fit hot-wire measurements.
It should be noted that for PIV and Optical-flow approaches the second order moments are computed with 520 samples,
whereas for hot-wire measurements more than 100000 samples were collected. Hence the Reynolds stresses are not statistically
converged values. Nevertheless, we can say that both PIV and Optical-flow approaches give similar results in good agreement
with the litterature.
5.3
Discussions of results
This first experience proves the ability of the method presented in this article to recover dense motion fields from PIV images.
Main characteristics of the flow (mixing layer mean characteristics and second order moments) are recovered in a similar way
with both approaches (PIV and the proposed Optical-flow). A major difference between the two approaches is the number of
vectors that the Optical-flow technique is able to extract. Let us nevertheless remark that, from a time computation point of
view, the presented method takes nearly 120 s for the estimation of an instantaneous dense motion field on a PC-linux (2.8GHz,
1.5Go-RAM) whereas the estimation with the PIV approach based on cross-correlation takes nearly 15 s.
7
0.045
0.04
0.035
u02 /∆U 2
0.03
0.025
0.02
0.015
0.01
0.005
0
-1
-0.5
0
0.5
1
0.5
1
0.5
1
y/δω
0.025
v 02 /∆U 2
0.02
0.015
0.01
0.005
0
-1
-0.5
0
y/δω
0.018
0.016
0.014
−u0 v 0 /∆U 2
0.012
0.01
0.008
0.006
0.004
0.002
0
-0.002
-1
-0.5
0
y/δω
F IG . 4 – Reynolds stresses distributions of the mixing layer at x/δω = 4. Each symbol representing a velocity measurement
technique : +, Optical-flow ; , PIV II ; ◦, PIV I ; •, Heitz (1999).
8
6
6.1
Results on the wake of a circular cylinder
Instantaneous velocity field
Instantaneous velocity field, vorticity map and contours are shown in figures 5 and 6 for the proposed Optical-flow approach
and for the PIV approach respectively.
The motion fields obtained from the two methods are very similar and represent the physical phenomenon well. The main
vortex launched in the middle of the image appears clearly in the velocity field as well as in different vorticity maps. These maps,
represented in figures 5 and 6, are very interesting to analyze. Indeed, it can be observed that both share the main phenomena but
that the major difference appears in the spatial resolution, where that of the optical flow approach is highly superior. Nevertheless,
a more precise study on this over-information must be carried out to reach its relevance. A work on that subject is currently in
progress. Let us notice, as we previously mentioned, that the higher quality of the input images used for this experiment (a 12
bit CCD camera vs a 8 bit one for the mixing layer) is probably responsible of the better results, in term of noise, of the vorticity
map issued form the optical-flow velocity field.
1
y/D
0.5
0
-0.5
-1
-0.5
0
0.5
1
1.5
2
x/D
2.5
3
3.5
4
4.5
1
y/D
0.5
0
-0.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
x/D
1
15
10
0.5
5
y/D
0
-5
0
-10
-15
-0.5
-1
0.5
1
1.5
2
2.5
3
3.5
4
x/D
F IG . 5 – View of instantaneous near wake vortices at Re=3000 – Optical-flow treatement : (a) : vector field (1 vector over 12 is
shown) ; (b), iso-contour of vorticity ωz? = (ωz U )/D (|ωz?min | = 0.5, |ωz?max | = 10, ∆ωz? = 1) ; (c), map of iso-vorticity ωz? .
6.2
Mean quantities
In this section, all motion fields form the 540 pairs of images of the whole sequence are computed. For both methods, mean
velocity field and their corresponding streamlines are represented in figures 7 and 8.
In these figures, the recirculation region clearly appears. Concerning the density of the resulting velocity field, one can observe
that the one issued from the optical flow is superior.
9
1
y/D
0.5
0
-0.5
-1
-0.5
0
0.5
1
1.5
2
x/D
2.5
3
3.5
4
4.5
1
y/D
0.5
0
-0.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
4
x/D
1
15
10
0.5
5
y/D
0
-5
0
-10
-15
-0.5
-1
0.5
1
1.5
2
2.5
3
3.5
4
x/D
F IG . 6 – View of instantaneous near wake vortices at Re=3000 – PIV treatement : (a), vector field ; (b), iso-contour of vorticity
ωz? = (ωz U )/D (|ωz?min | = 0.5, |ωz?max | = 10, ∆ωz? = 1) ; (c), map of iso-vorticity ωz? .
10
1
y/D
0.5
0
-0.5
-1
-0.5
0
0.5
1
1.5
2
x/D
2.5
3
3.5
4
4.5
1
y/D
0.5
0
-0.5
-1
0
1
2
3
4
x/D
F IG . 7 – Mean velocity field and corresponding streamlines in the near wake of a circular cylinder at Re = 3000 with Opticalflow approach (one vector out of twelve was plotted).
1
y/D
0.5
0
-0.5
-1
-0.5
0
0.5
1
1.5
2
x/D
2.5
3
3.5
4
4.5
1
y/D
0.5
0
-0.5
-1
0
1
2
3
4
x/D
F IG . 8 – Mean velocity field and corresponding streamlines in the near wake of a circular cylinder at Re = 3000 with
approach.
11
PIV
x/D = 1.56
x/D = 3.6
1.4
1.1
1.2
1
1
0.9
U/U∞
U/U∞
0.8
0.6
0.8
0.7
0.4
0.6
0.2
0.5
0
-0.2
0.4
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
-2
-1.5
-1
-0.5
y/D
0
0.5
1
1.5
2
0.5
1
1.5
2
0.5
1
1.5
2
0.5
1
1.5
2
0.5
1
1.5
2
y/D
0.08
0.15
0.06
0.1
0.04
0.05
V /U∞
V /U∞
0.02
0
-0.02
0
-0.05
-0.04
-0.1
-0.06
-0.08
-0.15
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
-2
-1.5
-1
-0.5
y/D
0
y/D
0.02
0.03
0.018
0.025
0.016
0.02
0.012
2
u02 /U∞
2
u02 /U∞
0.014
0.01
0.008
0.015
0.01
0.006
0.004
0.005
0.002
0
0
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
-2
-1.5
-1
-0.5
0.01
0.1
0.009
0.09
0.008
0.08
0.007
0.07
0.006
0.06
0.005
0.004
0.05
0.04
0.003
0.03
0.002
0.02
0.001
0.01
0
0
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
-2
-1.5
-1
-0.5
y/D
0
y/D
0.004
0.02
0.003
0.015
0.01
2
u0 v 0 /U∞
0.002
2
u0 v 0 /U∞
0
y/D
2
v 02 /U∞
2
v 02 /U∞
y/D
0.001
0
-0.001
0.005
0
-0.005
-0.01
-0.002
-0.015
-0.003
-0.02
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
-2
y/D
-1.5
-1
-0.5
0
y/D
F IG . 9 – Comparison of profiles of mean velocities and Reynolds stresses at two locations downstream of a cylinder : (a),
x/D = 1.56 ; (b), x/D = 3.6.
12
To have a quantitative idea of the relevance of different results, we represent in figure 9 some profiles for the mean velocity
and for the Reynolds stresses at two locations downstream of the cylinder (x/D = 1.56 and x/D = 3.6). Concerning the
mean velocities, both algorithms yields the same results. Concerning Reynolds stresses, the shape of all curves are quite similar
for x/D = 3.6 but their local extrema are differents, especially for x/D = 1.56. For the longitudinal velocity, the two local
extrema for x/D = 1.56 are indeed more important with the PIV algorithm. Let us recall that this region corresponds to the free
shear layers, which for the subcritical regime concerned, exhibits low vorticity thickness with high velocity range. Therefore the
vertical variation of the longitudal velocity is very large (strong velocity gradient). The proposed algorithm, which is based on a
global regularization, tends to smooth the velocity when its spatial gradient is too important. Hence, this can explain the lower
value of each extrema located in this area. In such area, the proposed div-curl regularization of the optical-flow approach tries
to estimate an homogenous vorticity area. Then, artificial small vortices may be introduced (by introducing small values of the
vertical component), which can be responsible of the more important value of the fluctuation of the transversal velocity. In a
future work, a hot-wire measurement will be done in this area to have another comparison.
Another interesting parameter to extract is the vortex formation length Lf . Two main approaches can be distinguished : the
distance from the base of the cylinder to the point with null longitudinal mean velocity (U = 0) on the centreline of the flow
(y = 0) and the point where the u0 2 + v 0 2 is maximum on the centreline. The latter criterion defined by Bloor & Gerrard (1966)
enables to compare results with hot-wire measurements. Indeed, measurements were carried out using a single hot-wire with the
wire parallel to the axis of the cylinder y−axis, so that it was more sensitive to the velocities induced by the formatioon of the
Kármán vortices. Based on this, we represent in table 2 all values extracted from the two algorithms. It is shown that results are
very similar. This proves the ability of the proposed approach to extract motion fields that are of great relevance.
U =0
u0 2 + v 0 2
Hot-wire
2.8
PIV
2.57
2.70
Optical-flow
2.50
2.80
TAB . 2 – Vortex formation length measurements Lf .
7
Conclusion
In this paper, we have experimentally evaluated a new method for estimating instantaneous velocities of fluid flows from
image sequences. This method is an extension of the standard optical-flow based approaches used in Computer Vision where a
two-fold robust objective function is minimized. The two parts (i.e, the data term and the regularizer) of the novel cost function
have been specifically designed to suit image sequences of fluid flows.
The data term is based on a continuity equation, as a more physically-grounded alternative to the usual brightness constancy
assumption. To be compatible with large displacements, we propose to use it in an integrated form. Such situations occur when
the imaged flow is fast (like in a number of fluid experiments), or when the temporal sampling rate is low (as with satellite
images).
As for the regularization, we argue that only a second-order regularizer is able to preserve completely the divergence and
vorticity structures of the flows. Based on the div-curl formalism, we thus introduced a robust second-order regularizer which
capture the divergence and vorticity of the unknown flow.
The developed approach has been tested on two experimental flows respectively representing a mixing layer and the near
wake of a circular cylinder. In each case, we have compared our results with ones issued from PIV. It can be point out that
main parameters values extract with both methods have the same order of magnitude. The relevance of obtained motion field is
then similar as PIV velocity fields. A major difference comes from the number of vectors that the presented technique is able to
estimate. Let us recall that a dense motion field is obtained (i.e. one vector per pixel).
Acknowledgments
The authors would like to thank Joel Delville (University of Poitiers, France) and Beatriz Camano (Rio Grande do Sul Federal
University, Brazil) for their valuable contribution on mixing layer experiments.
Références
A DRIAN , R. 1991 Particle imaging techniques for experimental fluid mechanics. Annal Rev. Fluid Mech. 23, 261–304. 2
A MINI , A. 1994 A scalar function formulation for optical flow. In Proc. Europ. Conf. Computer Vision, pp. 125–131. 4
BANNEHR , L., ROHN , R. & WARNECKE , G. 1996 A functionnal analytic method to derive displacement vector fields from
satellite image sequences. Int. Journ. of Remote Sensing 17 (2), 383–392. 4
B ÉR ÉZIAT, D., H ERLIN , I. & YOUNES , L. 2000 A generalized optical flow constraint and its physical interpretation. In Proc.
Conf. Comp. Vision Pattern Rec., , vol. 2, pp. 487–492. Hilton Head Island, South Carolina, USA. 4
13
B LACK , M. 1994 Recursive non-linear estimation of discontinuous flow fields. In Proc. Europ. Conf. Computer Vision, pp.
138–145. Stockholm, Sweden. 3
B LOOR , M. & G ERRARD , J. 1966 Measurements on turbulent vortices in a cylinder wake. Proc. R. Soc. Lond. 294, 319–342.
13
C OHEN , I. & H ERLIN , I. 1999 Non uniform multiresolution method for optical flow and phase portrait models : Environmental
applications. Int. J. Computer Vision 33 (1), 29–49. 3, 4
C ORPETTI , T., M ÉMIN , E. & P ÉREZ , P. 2002 Dense estimation of fluid flows. IEEE Trans on Pattern Analysis and Machine
Intelligence 24 (3), 365–380. 4, 5, 6
F ITZPATRICK , J. 1988 The existence of geometrical density-image transformations corresponding to object motion. Comput.
Vision, Graphics, Image Proc. 44 (2), 155–174. 4
G UPTA , S. & P RINCE , J. 1996 Stochastic models for div-curl optical flow methods. Signal Proc. Letters 3 (2), 32–34. 5
H EITZ , D. 1999 Etude expérimentale du sillage d’un barreau cylindrique se développant dans une couche de mélange plane
turbulente. PhD thesis, Université de Poiters. 5, 7, 8
H ORN , B. & S CHUNCK , B. 1981 Determining optical flow. Artificial Intelligence 17, 185–203. 2, 3
H UBER , P. 1981 Robust Statistics. John Wiley & Sons. 3
KORNPROBST, P., D ERICHE , R. & AUBERT, G. 1999 Image sequence analysis via partial differential equations. Journal of
Mathematical Imaging and Vision 11 (1), 5–26. 4
L ARSEN , R., C ONRADSEN , K. & E RSBOLL , B. 1998 Estimation of dense image flow fields in fluids. IEEE trans. on Geoscience
and Remote sensing 36 (1), 256–264. 4
L ECUONA , A., RUIZ -R IVAS , U. & RODRIGUEZ -AUMENTE , P. 2002 Near field vortex dynamics in axially forced, co-flowing
jets : quantitative description of a low-frequency configuration. Euro. J. of Mech. B/Fluids Vol. 21, pp. 701–720. 2
L OURENCO , L. & K ROTHAPALLI , A. 1995 On the accuracy of velocity and vorticity measurements with piv. Exp. in fluids Vol.
18, pp. 421–428. 2, 6
L OURENCO , L. & K ROTHAPALLI , A. 2000 True resolution piv : A mesh-free second-order accurate algoritm. In 10’rd International Symposium on Applications of Laser Techniques in Fluid Mechanics. Lisbon – Portugal. 2, 3
M C K ENNA , S. & M C G ILLIS , W. 2002 Performance of digital image velocimetry processing techniques. Exp. in fluids Vol. 32,
pp. 106–115. 2
M ÉMIN , E. & P ÉREZ , P. 1998 Dense estimation and object-based segmentation of the optical flow with robust techniques. IEEE
Trans. Image Processing 7 (5), 703–719. 3, 4
M ÉMIN , E. & P ÉREZ , P. 1999 Fluid motion recovery by coupling dense and parametric motion fields. In Proc. Int. Conf.
Computer Vision, , vol. 3, pp. 732–736. Corfou, Greece. 4
N OGUEIRA , J., L ECUONA , A. & RODRIGUEZ , P. 2001 Identification of a new source of peak locking, analysis and its removal
in conventional and super-resolution PIV techniques. Exp. in fluids 30, pp 309–316. 3
R AFFEL , M., W ILLERT, C. & KOMPENHANS , J. 2000 Particle Image Velocimetry. Springer. 2, 6
S CARANO , F. & R EITHMULLER , L. 2000 Advances in iterative multigrid piv image processing. Exp. in fluids [Suppl], pp.
S51–S60. 2
S CHUNK , B. 1984 The motion constraint equation for optical flow. In Proc. Int. Conf. Pattern Recognition, pp. 20–22. Montreal.
4
S ONG , S. & L EAHY, R. 1991 Computation of 3D velocity fields from 3D cine and CT images of human heart. IEEE trans. on
medical imaging 10 (3), 295–306. 4
S UTER , D. 1994 Motion estimation and vector splines. In Proc. Conf. Comp. Vision Pattern Rec., pp. 939–942. Seattle, USA. 5
WALLACE , J. & F OSS , J. 1995 The measurement of vorticity in turbulent flows. Annu. Rev. Fluid Mech. 27, 469–514. 4
W ERELEY, S. & M EINHART, C. 2001 Second-order accurate particule image velocimetry. Exp. in fluids Vol. 31, pp. 258–268.
2
W ILDES , R., A MABILE , M., L ANZILLOTTO , A. & L EU , T. 1997 Physically based fluid flow recovery from image sequences.
In Proc. Conf. Comp. Vision Pattern Rec., pp. 969–975. 4
W ILLERT, C. & G HARIB , M. 1991 Digital particle image velocimetry. Exp. in fluids Vol. 10, pp. 181–193. 2
Z HOU , L., K AMBHAMETTU , C. & G OLDGOF, D. 2000 Fluid structure and motion analysis from multi-spectrum 2D cloud
images sequences. In Proc. Conf. Comp. Vision Pattern Rec., , vol. 2, pp. 744–751. Hilton Head Island, South Carolina, USA.
4
14
Download