scispace - formally typeset
Open AccessJournal ArticleDOI

Influence of global rotation and Reynolds number on the large-scale features of a turbulent Taylor–Couette flow

Florent Ravelet, +2 more
- 07 May 2010 - 
- Vol. 22, Iss: 5, pp 055103
Reads0
Chats0
TLDR
In this paper, the scaling of the torque with Reynolds numbers at various angular velocity ratios (Rotation numbers) and the behavior of the wall shear stress when varying the Rotation number at high Reynolds numbers were investigated.
Abstract
We experimentally study the turbulent flow between two coaxial and independently rotating cylinders. We determined the scaling of the torque with Reynolds numbers at various angular velocity ratios (Rotation numbers) and the behavior of the wall shear stress when varying the Rotation number at high Reynolds numbers. We compare the curves with particle image velocimetry analysis of the mean flow and show the peculiar role of perfect counter-rotation for the emergence of organized large scale structures in the mean part of this very turbulent flow that appear in a smooth and continuous way: the transition resembles a supercritical bifurcation of the secondary mean flow.

read more

Content maybe subject to copyright    Report

Science Arts & Métiers (SAM)
is an open access repository that collects the work of Arts et Métiers Institute of
Technology researchers and makes it freely available over the web where possible.
This is an author-deposited version published in: https://sam.ensam.eu
Handle ID: .http://hdl.handle.net/10985/6784
To cite this version :
Florent RAVELET, René DELFOS, Jerry WESTERWEEL - Influence of global rotation and
Reynolds number on the large-scale features of a turbulent Taylor–Couette flow - Physics of
Fluids p.1-8 - 2010
Any correspondence concerning this service should be sent to the repository
Administrator : archiveouverte@ensam.eu

Influence of global rotation and Reynolds number on the large-scale
features of a turbulent Taylor–Couette flow
F. Ravelet,
1,a
R. Delfos,
2
and J. Westerweel
2
1
DynFluid, Arts et Métiers-ParisTech, 151 Blvd. de l’Hôpital, 75013 Paris, France
2
Laboratory for Aero and Hydrodynamics, Mekelweg 2, 2628 CD Delft, The Netherlands
Received 25 January 2010; accepted 27 February 2010; published online 7 May 2010
We experimentally study the turbulent flow between two coaxial and independently rotating
cylinders. We determined the scaling of the torque with Reynolds numbers at various angular
velocity ratios Rotation numbers and the behavior of the wall shear stress when varying the
Rotation number at high Reynolds numbers. We compare the curves with particle image velocimetry
analysis of the mean flow and show the peculiar role of perfect counter-rotation for the emergence
of organized large scale structures in the mean part of this very turbulent flow that appear in a
smooth and continuous way: the transition resembles a supercritical bifurcation of the secondary
mean flow. © 2010 American Institute of Physics. doi:10.1063/1.3392773
I. INTRODUCTION
Turbulent shear flows are present in many applied and
fundamental problems, ranging from small scales such as in
the cardiovascular system to very large scales such as in
meteorology. One of the several open questions is the emer-
gence of coherent large-scale structures in turbulent flows.
1
Another interesting problem concerns bifurcations, i.e., tran-
sitions in large-scale flow patterns under parametric influ-
ence, such as laminar-turbulent flow transition in pipes, or
flow pattern change within the turbulent regime, such as the
dynamo instability of a magnetic field in a conducting fluid,
2
or multistability of the mean flow in von Kármán or free-
surface Taylor–Couette flows,
3,4
leading to hysteresis or non-
trivial dynamics at large scale. In flow simulation of homo-
geneous turbulent shear flow it is observed that there is an
important role for what is called the background rotation,
which is the rotation of the frame of reference in which the
shear flow occurs. This background rotation can both sup-
press or enhance the turbulence.
5,6
We will further explicit
this in Sec. III.
A flow geometry that can generate both motions—shear
and background rotation—at the same time is the Taylor–
Couette flow which is the flow produced between differen-
tially rotating coaxial cylinders.
7
When only the inner cylin-
der rotates, the first instability, i.e., deviation from laminar
flow with circular streamlines, takes the form of toroidal
Taylor vortices. With two independently rotating cylinders,
there is a host of interesting secondary bifurcations, exten-
sively studied at intermediate Reynolds numbers, following
the work of Coles
8
and Andereck et al.
9
Moreover, it shares
strong analogies with Rayleigh–Bénard convection,
10,11
which are useful to explain different torque scalings at high
Reynolds numbers.
12,13
Finally, for some parameters relevant
in astrophysical problems, the basic flow is linearly stable
and can directly transit to turbulence at a sufficiently high
Reynolds number.
14
The structure of the Taylor–Couette flow, while it is in a
turbulent state, is not so well known and only few measure-
ments are available.
15
The flow measurements reported in
Ref. 15 and other torque scaling studies only deal with the
case where only the inner cylinder rotates.
12,13
In that precise
case, recent direct numerical simulations suggest that
vortexlike structures still exist at high Reynolds number
Re10
4
,
16,17
whereas for counter-rotating cylinders, the
flows at Reynolds numbers around 5000 are identified as
“featureless states.”
9
The structure of the flow is exemplified
with a flow visualization in Fig. 1 in our experimental setup
for a flow with only the inner cylinder rotating, counter-
rotating cylinders, and only the outer cylinder rotating,
respectively.
In the present paper, we extend the study of torques and
flow field for independently rotating cylinders to higher
Reynolds numbers up to 10
5
and address the question of
the transition process between a turbulent flow with Taylor
vortices, and this “featureless” turbulent flow when varying
the global rotation while maintaining a constant mean shear
rate.
In Sec. II, we present the experimental device and the
measured quantities. In Sec. III, we introduce the specific set
of parameters we use to take into account the global rotation
through a “Rotation number” and the imposed shear through
a shear-Reynolds number. We then present torque scalings
and typical velocity profiles in turbulent regimes for three
particular Rotation numbers in Sec. IV. We explore the tran-
sition between these regimes at high Reynolds number vary-
ing the Rotation number in Sec. V and discuss the results in
Sec. VI.
II. EXPERIMENTAL SETUP AND MEASUREMENT
TECHNIQUES
The flow is generated between two coaxial cylinders
Fig. 2. The inner cylinder has a radius of r
i
=110
0.05 mm and the outer cylinder of r
o
=1200.05 mm.
The gap between the cylinders is thus d = r
o
r
i
=10 mm and
a
Electronic mail: florent.ravelet@ensta.org.
PHYSICS OF FLUIDS 22, 055103 2010
1070-6631/2010/225/055103/8/$30.00 © 2010 American Institute of Physics22, 055103-1
Downloaded 09 May 2010 to 145.94.113.34. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

the gap ratio is
=r
i
/ r
o
=0.917. The system is closed at both
ends, with top and bottom lids rotating with the outer cylin-
der. The length of the inner cylinder is L = 220 mm axial
aspect ratio is L/ d=22. Both cylinders can rotate indepen-
dently with the use of two dc motors Maxon, 250 W. The
motors are driven by a homemade regulation device, ensur-
ing a rotation rate up to 10 Hz, with an absolute precision of
0.02 Hz and a good stability. A
LABVIEW program is used
to control the experiment: the two cylinders are simulta-
neously accelerated or decelerated to the desired rotation
rates, keeping their ratio constant. This ratio can also be
changed while the cylinders rotate, maintaining a constant
differential velocity.
The torque T on the inner cylinder is measured with a
corotating torque meter HBM T20WN, 2 N m. The signal
is recorded with a 12 bit data acquisition board at a sample
rate of 2 kHz for 180 s. The absolute precision on the torque
measurements is 0.01 N m, and values below 0.05 N m are
rejected. We also use the encoder on the shaft of the torque
meter to record the rotation rate of the inner cylinder. Since
that matches excellently with the demanded rate of rotation,
we assume that the outer cylinder rotates at the demanded
rate as well.
Since the torque meter is mounted in the shaft between
driving motor and cylinder, it also records besides the in-
tended torque on the wall bounding the gap between the two
cylinders the contribution of mechanical friction such as in
the two bearings, and the fluid friction in the horizontal
Kármán gaps between tank bottom and tank top. While the
bearing friction is considered to be marginal and measured
so in an empty, i.e., air filled system, the Kármán-gap con-
tribution is much bigger: during laminar flow, we calculated
and measured this to be of the order of 80% of the gap
torque. Therefore, all measured torques were divided by a
factor of 2, and we should consider the scaling of torque with
the parameters defined in Sec. III as more accurate than the
exact numerical values of torque.
A constructionally more difficult, but also more accurate,
solution for the torque measurement is to work with three
stacked inner cylinders and only measure the torque on the
central section, such as is done in the Maryland Taylor–
Couette setup,
12
and under development in the Twente
Turbulent Taylor–Couette setup.
18
We measure the three components of the velocity by
stereoscopic particle image velocimetry PIV
19
in a plane
illuminated by a double-pulsed Nd:yttrium aluminum garnet
laser. The plane is vertical Fig. 2, i.e., normal to the mean
flow: the in-plane components are the radial u and axial v
velocities, while the out-of-plane component is the azimuthal
component w. It is observed from both sides with an angle
of 60° in air using two double-frame charge coupled device
cameras on Scheimpflug mounts. The light-sheet thickness is
0.5 mm. The tracer particles are 20
m fluorescent
rhodamine B spheres. The field of view is 1125 mm
2
,
corresponding to a resolution of 300 1024 pixels. Special
care has been taken concerning the calibration procedure, on
which especially the evaluation of the plane-normal azi-
muthal component heavily relies. As a calibration target we
use a thin polyester sheet with lithographically printed
crosses on it, stably attached to a rotating and translating
microtraverse. It is first put into the light sheet and traversed
perpendicularly to it. Typically, five calibration images are
taken with intervals of 0.5 mm. The raw PIV images are
processed using
DAVIS 7.2 by Lavision. They are first mapped
to world coordinates, then they are filtered with a min-max
filter, then PIV processed using a multipass algorithm, with a
last interrogation area of 3232 pixels with 50% overlap,
and normalized using median filtering as postprocessing.
Then, the three components are reconstructed from the two
camera views. The mapping function is a third-order polyno-
mial, and the interpolations are bilinear. The PIV data acqui-
sition is triggered with the outer cylinder when it rotates in
order to take the pictures at the same angular position as used
during the calibration.
FIG. 1. Photographs of the flow at Re=3.610
3
. Left, A: only the inner cylinder rotating. Middle, B: counter-rotating cylinders. Right, C: only the outer
cylinder rotating. The flow structure is visualized using microscopic mica platelets Pearlessence.
ω
ω
r = 120 mm
o
i
o
r = 110 mm
i
h = 220 mm
PIV plane
1
.5 mm
(b)(a)
FIG. 2. Picture and sketch with dimensions of the experimental setup. One
can see the rotating torque meter upper part of the picture, the calibration
grid displacement device on top of the upper plate, one of the two cameras
left side, and the light sheet arrangement right side. The second camera is
further to the right.
055103-2 Ravelet, Delfos, and Westerweel Phys. Fluids 22, 055103 2010
Downloaded 09 May 2010 to 145.94.113.34. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

To check the reliability of the stereoscopic velocity
measurement method, we performed a measurement for a
laminar flow when only the inner cylinder rotates at a
Reynolds number as low as Re
S
=90 using an 86% glycerol-
water mixture. In that case, the analytical velocity field is
known: the radial and axial velocities are zero, and the azi-
muthal velocity w should be axisymmetric with no axial de-
pendence, and a radial profile in the form wr=r
i
r
o
/ r
r/ r
o
/ 1−
2
.
8
The results are plotted in Fig. 3. The mea-
sured profile solid line hardly differs from the theoretical
profile dashed line in the bulk of the flow 0.1r r
i
/ d
0.7. The discrepancy is, however, quite strong close to the
outer cylinder 关共r r
i
/ d=1, which may be due to the refrac-
tion close to the curved wall that causes measurement errors.
The in-plane components, which should be zero, do not ex-
ceed 1% of the inner cylinder velocity everywhere. In con-
clusion, the measurements are very satisfying in the bulk.
Further improvements to the technique have been made since
this first PIV test, in particular a new outer cylinder of im-
proved roundness, and the measurements performed in water
for turbulent cases are reliable in the range 0.1r r
i
/ d
0.85. The PIV measurements close to the wall could be
further improved by using an outer cylinder that is machined
in a block with flat outer interfaces normal to the cameras.
III. PARAMETER SPACE
The two traditional parameters to describe the flow are
the inner respectively, outer Reynolds numbers Re
i
=r
i
i
d/
兲关respectively, Re
o
=r
o
o
d/
兲兴, with the inner
respectively, outer cylinder rotating at rotation rates
i
re-
spectively,
o
, and
the kinematic viscosity.
We choose to use the set of parameters defined by Du-
brulle et al.:
20
a shear Reynolds number Re
S
and a Rotation
number Ro,
Re
S
=
2
Re
o
−Re
i
1+
Ro = 1−
Re
i
+Re
o
Re
o
−Re
i
. 1
With this choice, Re
S
is based on the laminar shear rate
S:Re
S
=Sd
2
/
. For instance, with a 20 Hz velocity difference
in counter-rotation, the shear rate is around 1400 s
−1
and
Re
S
1.410
5
for water at 20 °C. A constant shear
Reynolds number corresponds to a line of slope
in the
Re
o
;Re
i
coordinate system see Fig. 4.
The Rotation number Ro compares the mean rotation
with the shear and is the inverse of a Rossby number. Its sign
defines cyclonic Ro0 or anticyclonic Ro0 flows. The
Rotation number is zero in case of perfect counter-rotation
r
i
i
=−r
o
o
. Two other relevant values of the Rotation
number are Ro
i
=
−1−0.083 and Ro
o
=1−
/
0.091
for, respectively, inner and outer cylinders rotating alone.
Finally, a further choice that we made in our experiment was
the value of
=r
i
/ r
o
, which we have chosen as relatively
close to unity, i.e.,
=110/ 1200.91, which is considered a
narrow gap, and is the most common in reported experi-
ments, such as Refs. 8, 9, 21, and 22, although a value as low
as 0.128 is described as well.
4
A high
, i.e., 1−
1, is
special in the sense that for
1 the flow is equivalent to a
plane Couette flow with background rotation; at high
, the
flow is linearly unstable for 1 Ro Ro
o
.
5,20,23
In the present study we experimentally explore regions
of the parameter space that, to our knowledge, have not been
reported before. We present in Fig. 4 the parameter space in
Re
o
;Re
i
coordinates with a sketch of the flow states iden-
tified by Andereck et al.
9
and the location of the data dis-
cussed in the present paper. One can notice that the present
range of Reynolds numbers is far beyond that of Andereck,
and that with the PIV data we mainly explore the zone be-
tween perfect counter-rotation and only the inner cylinder
rotating.
IV. STUDY OF THREE PARTICULAR
ROTATION NUMBERS
In the experiments reported in this section, we maintain
the Rotation number at constant values and vary the shear
Reynolds number. We compare three particular Rotation
numbers, Ro
i
,Ro
c
, and Ro
o
, corresponding to rotation of the
inner cylinder only, exact counter-rotation, and rotation of
the outer cylinder rotating only, respectively. In Sec. IV A,
we report torque scaling measurements for a wide range of
Reynolds numbers—from base laminar flow to highly turbu-
lent flows—and in Sec. IV B, we present typical velocity
profiles in turbulent conditions.
A. Torque scaling measurements
We present in Fig. 5 the friction factor c
f
=T/ 2
␲␳
r
i
2
LU
2
G/ Re
2
, with U = Sd and G = T /
L
2
,asa
function of Re
S
for the three Rotation numbers. A common
definition for the scaling exponent
of the dimensionless
torque is based on G: G Re
S
. We keep this definition and
present the local exponent
in the inset in Fig. 5.We
compute
by means of a logarithmic derivative,
=2
+d logc
f
/ d logRe
S
.
At low Re, the three curves collapse on a Re
−1
curve.
This characterizes the laminar regime where the torque is
proportional to the shear rate on which the Reynolds number
is based.
0 .2 .4 .6 .8
1
0
0.2
0.4
0.6
0.8
1
(r−r
i
)/d
Velocity
(
dimless
)
FIG. 3. Dimensionless azimuthal velocity profile w / r
i
i
vs r r
i
/ d for
Ro=Ro
i
at Re
S
=90 see Sec. III for the definition of the parameters. Solid
line: measured mean azimuthal velocity. Dashed line: theoretical profile.
The radial component u, which should be zero, is also shown as a thin solid
line.
055103-3 Influence of global rotation and Reynolds number Phys. Fluids 22, 055103 2010
Downloaded 09 May 2010 to 145.94.113.34. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

FIG. 4. Color online Parameter space in Re
o
;Re
i
coordinates. The vertical axis Re
o
=0 corresponds to Ro = Ro
i
=−0.083 and has been widely studied Refs.
12, 13, 16, and 17. The horizontal axis Re
i
=0 corresponds to Ro=Ro
o
=0.091. The line Re
i
=−Re
o
corresponds to counter-rotation, i.e., Ro=Ro
c
=0. The PIV
data taken at a constant shear Reynolds number of Re
S
=1.410
4
are plotted with . The torque data with varying Ro at constant shear for various Re
S
ranging from Re
S
=310
3
to Re
S
=4.710
4
are plotted as gray blue online lines. They are discussed in Fig. 8. We also plot the states identified at much
lower Re
S
by Andereck et al. Ref. 9 as patches: black corresponds to laminar Couette flow, gray green online to “spiral turbulence,” dotted zone to
“featureless turbulence,” and vertical stripes to an “unexplored” zone.
10
1
10
2
10
3
10
4
10
5
10
6
10
−3
10
−2
10
−1
R
e
C
f
10
1
10
2
10
3
10
4
10
5
10
6
1
1.5
2
Re
α
FIG. 5. Color online Friction factor c
f
vs Re
S
for Ro
i
=
−1 black ,Ro
c
=0 blue , and Ro
o
=1−
/
red . Relative error on Re
S
: 5%; absolute
error on torque: 0.01 N m. Dashed green online line: Lewis’ data Ref. 13,Eq.3兲兴, for Ro
i
and
=0.724. Dash-dotted magenta online line: Racina’s data
Ref. 22,Eq.10兲兴. Solid thin black line: laminar friction factor c
f
=1/
Re. Inset: local exponent
such that C
f
Re
S
−2
, computed as 2
+d logC
f
/ d logRe
S
, for Ro
i
=
−1 black ,Ro
c
=0 blue , and Ro
o
=1−
/
red . Dashed green online line: Lewis’ data Ref. 13, Eq. 3兲兴 for
Ro
i
and
=0.724.
055103-4 Ravelet, Delfos, and Westerweel Phys. Fluids 22, 055103 2010
Downloaded 09 May 2010 to 145.94.113.34. Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jsp

Figures
Citations
More filters
Journal ArticleDOI

High-Reynolds number Taylor-Couette turbulence.

TL;DR: In this article, the authors review the recent progress in understanding of fully developed Taylor-Couette turbulence from the experimental, numerical, and theoretical points of view, focusing on the parameter dependence of the global torque and on the local flow organization, including velocity profiles and boundary layers.
Journal ArticleDOI

High Reynolds number Taylor-Couette turbulence

TL;DR: In this article, the authors review the recent progress in understanding of fully developed Taylor-Couette turbulence from the experimental, numerical, and theoretical point of view, focusing on the parameter dependence of the global torque and on the local flow organisation, including velocity profiles and boundary layers.
Journal ArticleDOI

Multiple states in highly turbulent Taylor–Couette flow

TL;DR: This result verifies the notion that bifurcations can occur in high-dimensional flows (that is, very large Re) and questions Kolmogorov's paradigm.
Journal ArticleDOI

Exploring the phase diagram of fully turbulent Taylor–Couette flow

TL;DR: In this article, the effect of the aspect ratio on the effective torque versus Taylor number scaling is analyzed and it is shown that different branches of the torque-versus-Taylor relationship associated to different aspect ratios are found to cross within 15 % of the Reynolds number associated to the transition to the ultimate regime.
Journal ArticleDOI

Optimal Taylor-Couette flow: direct numerical simulations

TL;DR: In this paper, the authors numerically simulate turbulent Taylor-Couette flow for independently rotating inner and outer cylinders, focusing on the analogy with turbulent Rayleigh-B\'enard flow.
References
More filters
Book

Turbulence, Coherent Structures, Dynamical Systems and Symmetry

TL;DR: In this article, the authors present a review of rigor properties of low-dimensional models and their applications in the field of fluid mechanics. But they do not consider the effects of random perturbation on models.
Journal ArticleDOI

Flow regimes in a circular Couette system with independently rotating cylinders

TL;DR: In this paper, a flow visualization and spectral studies of flow between concentric independently rotating cylinders have revealed a surprisingly large variety of different flow states, including Taylor vortices, wavy vortice, modulated wavy vectors, outflow boundaries and internal waves.
Journal ArticleDOI

Transition in circular couette flow

TL;DR: In this article, two distinct kinds of transition have been identified in Couette flow between rotating cylinders: the Taylor motion (periodic in the axial direction) and a pattern of travelling waves in the circumferential direction.
Journal ArticleDOI

Stereoscopic particle image velocimetry

TL;DR: The principle of stereoscopic PIV, the different stereoscopic configurations that have been used, the relative error in the out-of-plane to the in-plane measurement, and the relative merits of calibration-based methods for reconstructing the three-dimensional displacement vector in comparison to geometric reconstruction are discussed.
Related Papers (5)
Frequently Asked Questions (13)
Q1. What are the contributions mentioned in the paper "Influence of global rotation and reynolds number on the large-scale features of a turbulent taylor–couette flow" ?

In this paper, the authors use the classical formalism of bifurcations and instabilities to study the transition between featureless turbulence and turbulent Taylor-vortex flow at constant ReS, which seems to be supercritical. 

It is very tempting to use the classical formalism of bifurcations and instabilities to study the transition between featureless turbulence and turbulent Taylor-vortex flow at constant ReS, which seems to be supercritical ; the threshold for the onset of coherent structures in the mean flow is Roc. In a considerable range of ReS, counter-rotation Roc is also close or equal to an inflexion point in the torque curve ; this may be related to the crossover point, where the role of the correlated fluctuations is taken over by the large scale vortical structures. The role of turbulent versus large-scale transport of angular momentum should be further investigated from existing numerical or PIV velocity data. Ro as measured at much higher ReS that is used for PIV does qualitatively not change, these measurements suggest that the large scale vortices are not only persistent in the flow at higher ReS, but that they also dominate the dynamics of the flow. 

Hz is taken, and 20 consecutive PIV images, i.e., approximately 11 cylinder revolutions, are sufficient to obtain a reliable estimate of the mean flow. 

In a considerable range of ReS, counter-rotation Roc is also close or equal to an inflexion point in the torque curve; this may be related to the crossover point, where the role of the correlated fluctuations is taken over by the large scale vortical structures. 

Hz in 20 s, the vortices grow very fast, reach a value with a velocity amplitude of 0.08 ms−1, and then decay to become stabilized at aDownloaded 09 May 2010 to 145.94.113.34. 

The two traditional parameters to describe the flow are the inner respectively, outer Reynolds numbers Rei = ri id / respectively, Reo= ro od / , with the inner respectively, outer cylinder rotating at rotation rates i respectively, o , and the kinematic viscosity. 

Two other relevant values of the Rotation number are Roi= −1 −0.083 and Roo= 1− / 0.091 for, respectively, inner and outer cylinders rotating alone. 

Redistribution subject to AIP license or copyright; see http://pof.aip.org/pof/copyright.jspTo check the reliability of the stereoscopic velocity measurement method, the authors performed a measurement for a laminar flow when only the inner cylinder rotates at a Reynolds number as low as ReS=90 using an 86% glycerolwater mixture. 

While the bearing friction is considered to be marginal and measuredso in an empty, i.e., air filled system , the Kármán-gap contribution is much bigger: during laminar flow, the authors calculated and measured this to be of the order of 80% of the gap torque. 

The authors present in Fig. 5 the friction factor cf =T / 2 ri2LU2 G /Re2, with U=Sd and G=T / L 2 , as a function of ReS for the three Rotation numbers. 

Ro as measured at much higher ReS that is used for PIV does qualitatively not change, these measurements suggest that the large scale vortices are not only persistent in the flow at higher ReS, but that they also dominate the dynamics of the flow. 

The authors observe the experimental flow to be still laminar up to high Re; then, in a rather short range of Re numbers, the flow transits to a turbulent state at 4000 Reto 5000. 

Note that the dimensional values of the torque at Roo are very small and difficult to measure accurately, and that these may become smaller than the contributions by the two Kármán layers end effects that the authors simply take into account by dividing by 2, as described in Sec. II.