Vous êtes sur la page 1sur 6

Letters

https://doi.org/10.1038/s41567-017-0018-3

Destabilizing turbulence in pipe flow


Jakob Kühnen   1*, Baofang Song   1,2,3, Davide Scarselli1, Nazmi Burak Budanur1, Michael Riedl1,
Ashley P. Willis4, Marc Avila2 and Björn Hof1*

Turbulence is the major cause of friction losses in transport can be pushed outside its limit of existence and as a consequence
processes and it is responsible for a drastic drag increase in the entire flow relaminarizes. Disturbance schemes are developed
flows over bounding surfaces. While much effort is invested with the aid of direct numerical simulations (DNSs) of pipe flow
into developing ways to control and reduce turbulence intensi- and subsequently implemented and tested in experiments. In the
ties1–3, so far no methods exist to altogether eliminate turbu- DNS, a flow is simulated in a five-diameter (D)-long pipe and peri-
lence if velocities are sufficiently large. We demonstrate for odic boundary conditions are applied in the axial direction. Initially
pipe flow that appropriate distortions to the velocity profile we perturb laminar pipe flow by adding fluctuation levels of a fully
lead to a complete collapse of turbulence and subsequently fric- turbulent velocity field rescaled by a factor (k) to a laminar flow
tion losses are reduced by as much as 90%. Counterintuitively, field. As shown in Fig. 1a (dark blue curve), for small initial pertur-
the return to laminar motion is accomplished by initially bations (that is, small k), the disturbance eventually decays and the
increasing turbulence intensities or by transiently amplifying flow remains laminar. For sufficiently large amplitudes (k of order
wall shear. Since neither the Reynolds number nor the shear unity), turbulence is triggered (purple, red and cyan curves). So far,
stresses decrease (the latter often increase), these measures this is the familiar picture of the transition to turbulence in shear
are not indicative of turbulence collapse. Instead, an amplifica- flows, where turbulence is triggered only if perturbation amplitudes
tion mechanism4,5 measuring the interaction between eddies surpass a certain threshold. However, when increasing the turbulent
and the mean shear is found to set a threshold below which fluctuations well beyond their usual levels (k > 2.5), surprisingly the
turbulence is suppressed beyond recovery. highly turbulent flow almost immediately collapses and returns to
Flows through pipes and hydraulic networks are generally turbu- laminar (light and dark green curves). Here the initially strong vor-
lent and the friction losses encountered in these flows are responsi- tical motion leads to a redistribution of shear resulting in an unusu-
ble for approximately 10% of the global electric energy consumption. ally flat velocity profile (black profile in Fig. 1c).
Here turbulence causes a severe drag increase and consequently To achieve a similar effect in experiments, we increase the turbu-
much larger forces are needed to maintain desired flow rates. In lence level by vigorously stirring a fully turbulent pipe flow (Re = 3,500),
pipes, both laminar and turbulent states are stable (the former is employing four rotors located inside the pipe 50D downstream of the
believed to be linearly stable for all Reynolds number (Re) values; pipe inlet (see Supplementary Movie 1 and Supplementary Fig. 1). As
the latter is stable if Re > 2,040 (ref. 6)), but with increasing speed the highly turbulent flow proceeds further downstream, it surprisingly
the laminar state becomes more and more susceptible to small dis- does not return to the normal turbulence level but instead it quickly
turbances. Hence, in practice most flows are turbulent at sufficiently reduces in intensity until the entire flow is laminar (Fig. 1b top to bot-
large Re. While the stability of laminar flow has been studied in great tom and Supplementary Movie 1). Being linearly stable, the laminar
detail, little attention has been paid to the susceptibility of turbu- flow persists for the entire downstream pipe. In a second experiment,
lence, the general assumption being that once turbulence is estab- turbulent flow (Re = 3,100) is disturbed by injecting fluid through 25
lished it is stable. small holes (0.5 mm diameter) in the pipe wall (holes are distributed
Many turbulence control strategies have been put forward to across a pipe segment with a length of 25D, see Supplementary Fig. 3).
reduce the drag encountered in shear flows7–17. Recent strategies Each injected jet creates a pair of counter-rotating vortices, intensify-
employ feedback mechanisms to actively counter selected velocity ing the eddying motion beyond the levels of ordinary turbulence at
components or vortices. Such methods usually require knowledge of this Re. The additional vortices redistribute the flow and as a conse-
the full turbulent velocity field. In computer simulations7,8, it could quence the velocity profile is flattened (Fig.  1c, purple dotted line).
be demonstrated that under these ideal conditions, flows at a low When the perturbation is actuated downstream, fluctuation levels
Re number can even be relaminarized. In experiments, the required drop and the centre line velocity returns to its laminar value (Fig. 1d).
detailed manipulation of the time-dependent velocity field is, how- Laminar motion persists for the remainder of the pipe. In this case,
ever, currently impossible to achieve. Other studies employ passive the frictional drag is reduced by a factor of 2. Overall the injected fluid
(for example, riblets) or active (oscillations or excitation of travel- amounts to only ~1.5 % of the total flow rate in the pipe. With the
ling waves) methods to interfere with the near-wall turbulence cre- present actuation device, we achieved a net power-saving (taking all
ation. Typically here drag reduction of 10 to 40% has been reported, actuation losses into account) of 31% over the remainder of the pipe.
but often the control cost is substantially higher than the gain, or a On the other hand, the minimum actuation cost required to create the
net gain can be achieved only in a narrow Re number regime. necessary flow disturbance is substantially lower (~1%), so that the
Instead of attempting to control or counter certain components net saving potential at this Re is 45% (see Methods).
of the complex fluctuating flow fields, we will show in the follow- In another experiment, we attempted to disrupt turbulence
ing that by appropriately disturbing the mean profile, turbulence (Re = 5,000) by injecting fluid parallel to the wall in the streamwise

Nonlinear Dynamics and Turbulence Group, IST Austria, Klosterneuburg, Austria. 2Center of Applied Space Technology and Microgravity (ZARM),
1

University of Bremen, Bremen, Germany. 3Center for Applied Mathematics, Tianjin University, Tianjin, China. 4School of Mathematics and Statistics,
University of Sheffield, Sheffield, UK. Jakob Kühnen and Baofang Song contributed equally to this work. *e-mail: jakob.kuehnen@ist.ac.at; bhof@ist.ac.at

Nature Physics | www.nature.com/naturephysics

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letters NaTuRe PhySICS

a b
3.0
Moderate perturbation

Perturbation energy/Turb. kin. energy


Large perturbation Turbulence
2.5 t0
(upstream)
t0
t1
2.0 Intensified
t1
t2 turbulence
1.5 t2
Turbulence decay
1.0
(downstream)

0.5
Laminar
motion
0
0 20 40 60 80 100
tU/D

c d
1.4 0.08
114.5 Upstream
34.8 0.06

(|u′x|+|u′y|)/U
13 28 disturbance off
1.2 0.04 Upstream
0.02 disturbance on
1.0
4.8 0
0.8 -0.02
0 20 40 60 80 100
uz/U

Turbulent profile, Re=5000 (exp)


0.6 Wall-normal inject., Re=3100 (exp) 2.0
Streamwise inject., Re=5000 (exp) Upstream
1.8
0.4 Moving wall, Re=5000 (exp)
DNS initial condition, Re=4000 1.6 disturbance off
uzc/U
Upstream
0.2 1.4 disturbance on
Transient growth amplitude 1.2
0 1.0
-0.50 -0.25 0 0.25 0.50 0 20 40 60 80 100
r/D tU/D

Fig. 1 | Perturbing turbulence. a, Direct numerical simulations of pipe flow starting from turbulent initial conditions (taken from a run at Re = 10,000),
rescaled by a constant factor k and added to the laminar base flow at Re = 4,000, which was then integrated forward in time (at Re = 4,000). For small
initial energies, perturbations die out (dark blue curve). For sufficiently large energies (k≈1), transition to turbulence occurs (red, purple and cyan). For
even larger energies (k > 2.5), however, the initially turbulent flow is destabilized and collapses after a short time (light and dark green curves). The six
streamwise vorticity isosurface figures show ωz =  +/− (red/blue) 7.2, 2.0 and 1.6 U/D respectively at snapshot times t0 = 0, t1 = 5 and t2 = 10 (D/U).
b, Fully turbulent flow (top panel) at Re = 3,100 is perturbed by vigorously stirring the fluid with four rotors. The more strongly turbulent flow
(second panel) eventually relaminarizes as it proceeds downstream (third and fourth panel). c, Temporally and azimuthally averaged velocity fields of
modified/perturbed flow fields in simulations and experiments. uz is the streamwise velocity component; the cross-stream components are denoted by
ur and uθ. d, Relaminarization of fully turbulent flow in experiments at Re = 3,100. The flow is perturbed by injecting 25 jets of fluid radially through the pipe
wall. When actuated, the fluctuation levels in the flow drop (top panel) and the centre line velocity switches from the turbulent level to the laminar value
(2U, where U is the mean velocity in the pipe) (bottom panel).

direction (see Supplementary Figs. 2 and 4). Unlike for the previous it decelerates the flow in the central part of the pipe cross-section
case, this disturbance does not result in a magnification of cross-stream while it accelerates the flow in the near-wall region. The mass flux
fluctuations, but instead it directly increases the wall shear stress and through the pipe and hence Re remain unaffected (see Supplementary
hence also the friction Re number, Reτ. Directly downstream of the equation (17) and Supplementary Fig. 7). Unlike in the experiments
injection point, the latter is increased by about 15%. The acceleration where the disturbance is applied locally and persists in time, here
of the near-wall flow automatically causes deceleration of the flow in the forcing is applied globally. As shown in Fig.  2b, when turning
the pipe centre (the overall mass flux is held constant), hence again on the forcing with sufficient amplitude the initially fully turbulent
resulting in a flatter velocity profile (blue in Fig. 1c). Despite the local flow completely relaminarizes. Hence, a profile modification alone
increase in Reτ, further downstream the fluctuation levels begin to suffices to destabilize turbulence. Interestingly, the energy required
drop and the turbulent flow has been sufficiently destabilized that for the forcing is smaller than the energy gained due to drag reduc-
eventually (30D downstream) it decays and the flow returns to lami- tion (even for intermediate forcing amplitudes, see Supplementary
nar. As a result, friction losses drop by a factor of 2.9 (see Fig. 2a) and Fig. 8). In this case, we therefore obtain a net energy-saving already
the potential net power saving (not including actuation losses) is 55% in the presence of the forcing (in experiments, the saving is achieved
(see Methods). For this type of perturbation, we find that relaminar- downstream of the perturbation location). After removal of the forc-
ization occurs for an intermediate injection range (~15% of the flow ing (see Supplementary Fig. 12), turbulence fluctuation levels con-
rate in the pipe), while for smaller and larger rates, the flow remains tinue to drop exponentially and the flow remains laminar for all
turbulent. A property common to all above relaminarization mecha- times. This effect has been tested for fully turbulent flow for Re num-
nisms is their effect on the average turbulent velocity profile. bers between 3,000 and 100,000, and in all cases a sufficiently strong
To test a possible connection between the initial flat velocity pro- force was found to lead to a collapse of turbulence resulting in drag
file and the subsequent turbulence collapse, we carried out further decrease and hence energy-saving in the numerical simulations of up
computer simulations where, this time, a forcing term was added to to 95% (in practical situations, finite-amplitude perturbations may
the full Navier–Stokes equations. The force was formulated such that limit the persistence of laminar flow at such high Re).

Nature Physics | www.nature.com/naturephysics

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NaTuRe PhySICS Letters
a 3.5 b 20
Re = 5,000 Re = 50,000
3.0
15
2.5
∆p/∆plam

∆p/∆plam
2.0

disturbance off
Upstream 10
1.5 Upstream

Force off
disturbance on
Force on
1.0
5
0.5

0 0
0 50 100 150 200 250 0 50 100 150
tU/D tU/D

c d Injection (percentage of total flow rate)


0 6 12 18 24 30
12 350
Re = 24,000 Moving wall Total/partial relaminarization
10 300
Streamwise injection Total/partial relaminarization
250

Transient growth
8
∆p/∆plam

200
wall motion

6
150
4 Theoretical level (DNS)
100
2 50

0 0
-50 0 50 100 150 200 250 0 1 2 3 4 5
tU/D Uwall/U

Fig. 2 | Laminarization mechanisms. a, After the streamwise near-wall injection is actuated, the pressure drop reduces to its laminar value. b, A body force
term is added in the numerical simulations that leads to an on average flatter flow profile (the fluid close to the wall is accelerated while it is decelerated
in the near-wall region). Disturbing the flow profile in this manner leads to a collapse of turbulence, here shown for Re = 50,000 where consequently
friction losses drop by a factor of 10. c, In the experiment, the near-wall fluid is accelerated via a sliding pipe segment, which is impulsively moved in the
axial direction. Directly after the pipe segment is stopped, the flow has a much flatter velocity profile. Subsequently, turbulence collapses and the frictional
drag drops to the laminar value. d, Transient growth measures the efficiency of the lift-up mechanism (that is, how perturbations in the form of streamwise
vortices are amplified while growing into streaks (deviations of the streamwise velocity component)). All disturbance schemes used lead to a reduction in
transient growth. The threshold value below which relaminarization occurs in the numerical simulations (control via body force) is indicated by the orange
line. For comparison, the experimental flow disturbance mechanisms are shown in blue (streamwise injection) and red (moving wall). In agreement with the
numerical prediction, all disturbance amplitudes that lead to a collapse of turbulence (solid symbols) fall below the threshold value found in the simulations.

We next investigate whether a profile modification on its own In turbulent wall-bounded shear flows, energy has to be transferred
also relaminarizes turbulence in experiments. While body forces continuously from the mean shear into eddying motion, and a key
such as that used in the simulations are not available (at least not factor here is the interplay between streamwise vortices (that is, vor-
for electrically non-conducting fluids), profiles can nevertheless tices aligned with the mean flow direction) and streaks. The latter are
be flattened by a local change in the boundary conditions. For this essentially dents in the flow profile that have either markedly higher
purpose, one pipe segment is replaced by a pipe of slightly (4%) or lower velocities than their surroundings. Streamwise vortices ‘lift
larger diameter that is pushed over the ends of the original pipe and up’ low-velocity fluid from the wall and transport it towards the cen-
can be impulsively moved with respect to the rest of the pipe (see tre (see Supplementary Fig. 8). The low-velocity streaks created in the
Supplementary Movie 2 and Supplementary Fig. 5). The pipe seg- process give rise to (nonlinear) instabilities and the creation of further
ment is then impulsively accelerated in the streamwise direction and vortices. Key to the efficiency of this ‘lift-up mechanism’ is that weak
abruptly stopped, the peak velocity of the 300D-long movable pipe vortices suffice to create large-amplitude streaks. This amplification
segment is equal to or larger than (up to three times) the bulk flow process is rooted in the non-normality4 of the linear Navier–Stokes
speed in the pipe. The impulsive acceleration of the near-wall fluid operator and its magnitude is measured by the so-called transient
leads to a flattened velocity profile (red profile in Fig. 1c). Despite growth (see also Supplementary Figs. 10 and 11).
the fact that overall the fluid is accelerated and additional shear Computing transient growth for the forced flow profiles in
is introduced (Reτ is increased), after the wall motion is stopped the DNS, we indeed observe that transient growth monotonically
(abruptly, over the course of 0.2 s) turbulence also in this case decays decreases with forcing amplitude (see Supplementary Fig. 11) and
(see Fig. 2c and Supplementary Movie 2). If, on the other hand, the it assumes its minimum value directly before turbulence collapses.
wall acceleration is reduced, with wall velocities lower than 0.8U, Generally, the flatter the velocity profile the more the streak vortex
turbulence survives. The impulsive wall motion is found to relami- interaction is suppressed, and in the limiting case of a uniformly flat
narize turbulence very efficiently up to the highest Re number profile the lift-up mechanism breaks down entirely.
(Re = 40,000) that could be tested in the experiment (here the wall Revisiting the experiments, the velocity profiles of all the dis-
was moved at the bulk flow speed). turbed flows considered exhibit a substantially reduced transient

Nature Physics | www.nature.com/naturephysics

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letters NaTuRe PhySICS

a friction value before the profile modification corresponds to the


Turbulent friction characteristic Blasius law for turbulence (upper line) and after the

Rotors
disturbance it drops directly to the laminar Hagen–Poiseuille law.

Injection
Hence, the maximum drag reduction feasible in practice is reached

Moving wall
Radial jets

(Fig.  3b), and at the highest Re numbers studied in experiments,


Darcy friction factor

Volume forcing
10–2
90% reduction is obtained. Although the numerical and experimen-
tal relaminarization methods affect the flow in different ways, the
common feature is that the velocity profile is flattened.
The presented control schemes require manipulation of only a
Laminar friction
single velocity component and moreover they do not require any
information about the instantaneous turbulent velocity field. The
10–3 overall control strategy is far simpler compared with recently pro-
posed active and feedback control schemes, while at the same time
104 105 it offers the maximum possible drag reduction. The future challenge
Re
is to develop and optimize methods that lead to the desired profile
b 100 modifications in high-Re-number turbulent flows.

80
Methods
Methods, including statements of data availability and any asso-
Drag reduction (%)

ciated accession codes and references, are available at https://doi.


60 org/10.1038/s41567-017-0018-3.

40
Theoretical limit Received: 5 July 2017; Accepted: 3 November 2017;
Volume forcing Published: xx xx xxxx
Radial jets
Axial injection
20
Moving wall References
Rotors 1. Lumley, J. & Blossey, P. Control of turbulence. Ann. Rev. Fluid Mech. 30,
311–327 (1998).
0 2. Kasagi, N., Suzuki, Y. & Fukagata, K. Microelectromechanical systems-based
104 105
Re feedback control of turbulence for skin friction reduction. Ann. Rev. Fluid
Mech. 41, 231–251 (2009).
3. Kim, J. & Bewley, T. R. A linear systems approach to flow control. Ann. Rev.
Fig. 3 | Drag reduction. a, Friction factor, f, as a function of Re. Initially Fluid Mech. 39, 383–417 (2007).
all flows are fully turbulent and friction factors follow the Blasius–Prandtl 4. Trefethen, L. N., Trefethen, A. E., Reddy, S. C. & Driscoll, T. A.
scaling (f = 0.316 Re−0.25, red line). When the control is turned on, flows Hydrodynamic stability without eigenvalues. Science 261, 578–584 (1993).
relaminarize and the friction factors drop to the corresponding laminar 5. Brandt, L. The lift-up effect: The linear mechanism behind transition and
values (Hagen–Poiseuille law in blue, f = 64/Re). The rotors, radial jet turbulence in shear flows. Eur. J. Mech. B 47, 80–96 (2014).
6. Avila, K. et al. The onset of turbulence in pipe flow. Science 333, 192–196 (2011).
injection, axial injection and moving wall controls are carried out in 7. Bewley, T. R., Moin, P. & Temam, R. DNS-based predictive control of
laboratory experiments while the volume force cases are from direct turbulence: an optimal benchmark for feedback algorithms. J. Fluid Mech.
numerical simulations of the Navier–Stokes equations. For all cases, the Re 447, 179–225 (2001).
number is held constant throughout the experiment. b, Drag reduction as a 8. Högberg, M., Bewley, T. R. & Henningson, D. S. Relaminarization of
function of Re. For the injection perturbation, a maximum drag reduction of Reτ = 100 turbulence using gain scheduling and linear state-feedback control.
Phys. Fluids 15, 3572–3575 (2003).
~70% was reached, whereas for the moving wall and volume forcing, 90 and 9. Auteri, F., Baron, A., Belan, M., Campanardi, G. & Quadrio, M. Experimental
95% were achieved, respectively. All data points reach the drag reduction assessment of drag reduction by traveling waves in a turbulent pipe flow.
limit set by relaminarization except for the Re > 30,000 in experiments Phys. Fluids 22, 115103 (2010).
where values are slightly above. Although these flows are laminar (that is, 10. Lieu, B., Moarref, R. & Jovanović, M. Controlling the onset of turbulence by
fluctuations are zero), the profile shape is still developing and has not quite streamwise travelling waves. Part 2. Direct numerical simulation. J. Fluid
Mech. 663, 100–119 (2010).
reached the Hagen–Poiseuille profile yet (the development length required
11. Moarref, R. & Jovanović, M. Controlling the onset of turbulence by
to reach a fully parabolic profile increases linearly with Re). streamwise travelling waves. Part 1. Receptivity analysis. J. Fluid Mech. 663,
70–99 (2010).
12. Quadrio, M., Ricco, P. & Viotti, C. Streamwise-traveling waves of spanwise
growth (Fig. 1c). For the streamwise injection, amplitudes relami- wall velocity for turbulent drag reduction. J. Fluid Mech. 627, 161–178 (2009).
narizing the flow also show the minimum amplification (Fig.  2d) 13. Hof, B., De Lozar, A., Avila, M., Tu, X. Y. & Schneider, T. M. Eliminating
while at lower and higher injection rates where turbulence survives turbulence in spatially intermittent flows. Science 327, 1491–1494 (2010).
14. Rathnasingham, R. & Breuer, K. Active control of turbulent boundary layers.
the amplification factors are higher and above the threshold found
J. Fluid Mech. 495, 209–233 (2003).
in the simulations. Similarly for the moving wall at sufficiently large 15. Willis, A. P., Hwang, Y. & Cossu, C. Optimally amplified large-scale streaks
wall acceleration where relaminarization is achieved, the lift-up and drag reduction in turbulent pipe flow. Phys. Rev. E 82, 036321 (2010).
efficiency is reduced below threshold, while at lower wall speeds it 16. Du, Y. & Karniadakis, G. E. Suppressing wall turbulence by means of a
remains above. transverse traveling wave. Science 288, 1230–1234 (2000).
Some parallels between the present study and injection and suc- 17. Min, T., Kang, S. M., Speyer, J. L. & Kim, J. Sustained sub-laminar drag in a
fully developed channel flow. J. Fluid Mech. 558, 309–318 (2006).
tion control in channels and boundary layers18–20 can be drawn. 18. Park, J. & Choi, H. Effects of uniform blowing or suction from a spanwise
While for boundary layers during the injection phase the drag slot on a turbulent boundary layer flow. Phys. Fluids 11, 3095–3105 (1999).
downstream increases, during the suction it decreases. Suction 19. Sumitani, Y. & Kasagi, N. Direct numerical simulation of turbulent transport
applied to a laminar Blasius boundary layer leads to a reduction of with uniform wall injection and suction. AIAA J. 33, 1220–1228 (1995).
20. Fukagata, K., Iwamoto, K. & Kasagi, N. Contribution of Reynolds stress
the boundary layer thickness and this is well known to delay transi-
distribution to the skin friction in wall-bounded flows. Phys. Fluids 14,
tion21 and push the transition location downstream. 73–76 (2002).
The drag reduction achieved for the different methods used to 21. Fransson, J. H. M. & Alfredsson, P. H. On the disturbance growth in an
destabilize turbulence is summarized in Fig.  3. In each case, the asymptotic suction boundary layer. J. Fluid Mech. 482, 51–90 (2003).

Nature Physics | www.nature.com/naturephysics

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
NaTuRe PhySICS Letters
Acknowledgements B.S., M.A. and N.B.B. analysed the numerical results. M.A., A.P.W. and B.H. supervised
We acknowledge the European Research Council under the European Union’s the computer simulations. D.S., A.P.W., M.A. and B.S. performed the theoretical analysis.
Seventh Framework Programme (FP/2007-2013)/ERC Grant Agreement 306589, J.K., B.S., D.S., M.A. and B.H. wrote the paper.
the European Research Council (ERC) under the European Union’s Horizon 2020
research and innovation programme (grant agreement no. 737549) and the Deutsche Competing interests
Forschungsgemeinschaft (Project No. FOR 1182) for financial support. We thank our The authors declare no competing financial interests.
technician P. Maier for providing highly valuable ideas and greatly supporting us in all
technical aspects. We thank M. Schaner for technical drawings, construction and design.
We thank M. Schwegel for a Matlab code to post-process experimental data. Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/
s41567-017-0018-3.
Author contributions Reprints and permissions information is available at www.nature.com/reprints.
J.K. and B.H. designed the experiments. J.K. and D.S. carried out the experiments and
post-processed the data. M.R. carried out the rotor experiments. J.K., D.S. and B.H. Correspondence and requests for materials should be addressed to J.K. or B.H.
analysed the experimental results. J.K. and B.H. supervised the experimental work. Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
B.S. and N.B.B. performed the computer simulations of the Navier–Stokes equations. published maps and institutional affiliations.

Nature Physics | www.nature.com/naturephysics

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
Letters NaTuRe PhySICS

Methods U, such that uwall = Upipe/U. The flow rate and hence the Re number (Re = UD/ν,
Experimental set-up for the rotors. The facility consists of a glass pipe where U is the mean velocity, D is the diameter of the tube and ν is the kinematic
(poly(methyl methacrylate); PMMA) with inner diameter D = 54 ± 0.2 mm and a viscosity of the fluid) can be adjusted by means of a valve in the supply pipe. The
total length of 12 m (222D) made of 2 m sections (see Supplementary Fig. 1). The temperature of the water is continuously monitored at the pipe exit. The flow
flow in the pipe is gravity driven and the working fluid is water that enters the pipe rate is measured with an electromagnetic flowmeter (ProcessMaster FXE4000,
from a reservoir. The flow rate and hence the Re number (Re = UD/ν, where U is ABB). The accuracy of Re is within ±1%. The flow is always turbulent when
the mean velocity, D is the diameter of the tube and ν is the kinematic viscosity of entering the control area. The velocity field is measured ~50D upstream from the
the fluid) can be adjusted by means of a valve in the supply pipe. The temperature downstream steel pipe. The measurement plane is parallel to the streamwise flow
of the water is continuously monitored at the pipe exit. The flow rate is measured direction (pipe z axis) and located in the centreline of the pipe. The two velocity
with an electromagnetic flowmeter (ProcessMaster FXE4000, ABB). The accuracy components within a plane of ~3.5D length are measured using a high-speed 2D
of Re is within ±1%. To ensure fully turbulent flow, the flow is perturbed by a small particle image velocimetry system (LaVision) with a full resolution of 2,400 × 1,900
static obstacle (a 1-mm-thick, 20-mm-long needle located 10D after the inlet). px. The resulting spatial resolution is 56 vectors per D. The data rate is 100 hertz.
Two metres downstream from the inlet the turbulent flow is perturbed by four Hollow glass spheres with a mean diameter of 13 μm are used for seeding. Around
small rotors that are mounted on a support structure within the pipe as indicated the measurement plane, the pipe is encased by a small rectangular Perspex box
in Supplementary Fig. 1. The wiring of the motors is incorporated in the support (50 × 50 × 350 mm) filled with water, such that the optical axis of the camera is
structure of the motors. The rotors are small rectangular bars with even smaller perpendicular to the air/water interface to reduce refraction and distortion of the
rectangular bars at the tips. Their only purpose is to induce perturbations to the images. A differential pressure sensor (DP 45, Validyne, full range of 550 Pa with
flow but no propelling motion or thrust. The rotors are turned at a rate of seven an accuracy of ±0.5%) is mounted onto the movable Perspex pipe. Here, the
rotations per second. For the purpose of visual observations and video recordings pressure drop Δp in the Perspex pipe is measured between two pressure taps
of the flow field, the flow is seeded with neutrally buoyant anisotropic particles22. (axial spacing 260 mm).
The three locations where Supplementary Movie 1 was recorded are indicated in
Supplementary Fig. 1. Numerical method. We solve the incompressible Navier–Stokes equations

Experimental set-up for the wall-normal jet injection and the streamwise ∂u 1
+ u ⋅ ∇u = −∇p + Δu + F, ∇ ⋅ u = 0 (1)
injection through an annular gap. The facility consists of a glass pipe with inner ∂t Re
diameter D = 30 ± 0.01 mm and a total length of 12 m (400D) made of 1 m sections
(see Supplementary Fig. 2). The flow in the pipe is gravity driven and the working in a straight circular pipe in cylindrical coordinates (r, θ, z), with r, θ and z being
fluid is water that enters the pipe from a reservoir. The flow rate and hence the the radial, azimuthal and axial coordinate respectively. Throughout this study, the
Re number (Re = UD/ν, where U is the mean velocity, D is the diameter of the flow is driven by a constant mass flux. The velocity u is normalized by the mean
tube and ν is the kinematic viscosity of the fluid) can be adjusted by means of a velocity U and length by pipe diameter D. F is the external body force and p is
valve in the supply pipe. The temperature of the water is continuously monitored pressure. A Fourier–Fourier-finite difference code is used for the integration of the
at the pipe exit. The flow rate is measured with an electromagnetic flowmeter governing equations, with periodic boundary conditions in the axial and azimuthal
(ProcessMaster FXE4000, ABB). The accuracy of Re is within ±1%. To ensure directions. In the radial direction, a central finite-difference scheme with a nine-
fully turbulent flow, the flow is perturbed by a small static obstacle (a 1-mm-thick, point stencil is adopted. In this formulation, velocity can be expressed as
20-mm-long needle located 10D after the inlet). Two metres downstream from
the inlet, the turbulent flow can be perturbed in a controlled way by two different K M
devices that are mounted within the pipe (see Supplementary Figs. 3 and 4). The
velocity field is measured ~330D downstream from the disturbance (control) at
u(r , θ , z ) = ∑ ∑ uk, m(r , t )e (iαkz +imθ) (2)
k =−K m =−M
the position of the light sheet. The measurement plane is perpendicular to the
streamwise flow direction (pipe z axis). All three velocity components within the
plane are recorded using a high-speed stereo particle image velocimetry system where αk and m give wavenumbers of the modes in axial and azimuthal direction
(Lavision GmbH) consisting of a laser and two Phantom V10 high-speed cameras respectively, 2π/α gives the pipe length and uk, m is the complex Fourier coefficient
with a full resolution of 2,400 × 1,900 px. The resulting spatial resolution is of mode (k,m). The governing equations are integrated with a second-order semi-
77 vectors per D. The data rate is 100 hertz. Hollow glass spheres (mean diameter implicit time-stepping scheme (for details, see23). The code has been verified and
13 μm, ϱ = 1.1 g cm–1) are used as seeding particles. Around the measurement extensively used in many studies (for example,24,25,26).
plane, the pipe is encased by a water-filled prism such that the optical axes of In Supplementary Table 1, we list the Re numbers, pipe lengths and resolutions
the cameras are perpendicular to the air/water interface to reduce refraction and we considered in our simulations. To avoid significant domain size effect, the pipe
distortion of the images. Downstream of the perturbation, the pressure drop lengths are selected to contain a few low-speed streaks, whose streamwise length
Δp is measured between two pressure tabs with a differential pressure sensor is typically around 500 wall units in our normalization (see27). The pipe length was
(DP 45, Validyne, full range of 220 Pa with an accuracy of ±0.5%) separated by doubled for Re = 4,000 and 5,000 to verify that the pipe lengths here in the table
39.5D in the axial direction. As the difference in pressure drop between laminar are sufficient. The resolutions are set to be able to sufficiently resolve the near-
and turbulent flows is very distinct even at moderate Re numbers, the signal is wall structures (see reference grid sizes shown in Table 1 of24). Note that there is a
utilized to observe whether the flow is laminar or turbulent. difference of a factor of two in length scales between our normalization and theirs
(double ours to compare with theirs).
Experimental set-up for the moving pipe. A movable Perspex pipe with inner
diameter D = 26 ± 0.1 mm and a total length of 12 m (461D) is fitted to very thin- Data availability. The data that support the plots within this paper and other
walled stainless-steel pipes (MicroGroup) with outer diameter dst,o = 25.4 ± 0.13 mm findings of this study are available from the corresponding author upon
and a wall thickness of 0.4 ± 0.04 mm such that the Perspex pipe overlaps the steel reasonable request.
pipes at the upstream and downstream end (see Supplementary Fig. 5). The steel
pipes are stationary (mounted on fixed bearings). With respect to the Perspex pipe, References
they act as support and slide bearing, allowing the Perspex pipe to be moved back 22. Matisse, P. & Gorman, M. Neutrally buoyant anisotropic particles for flow
and forth in the axial direction. To prevent sagging, the Perspex pipe is supported visualization. Phys. Fluids 27, 759–760 (1984).
by six additional bushings (polymer sleeve bearings, Igus). To avoid leakage, a 23. Willis, A. P. The Openpipeflow Navier–Stokes solver. SoftwareX 6,
radial shaft seal is mounted at both ends of the Perspex pipe. The length of the 124–127 (2017).
control section between the stationary upstream and downstream stainless-steel 24. Avila, M., Willis, A. P. & Hof, B. On the transient nature of localized pipe
pipes (that is, the actual length where the wall of the Perspex pipe is in contact with flow turbulence. J. Fluid Mech. 646, 127–136 (2010).
the fluid and can be moved relative to the mean flow by moving the Perspex pipe) 25. Barkley, D. et al. The rise of fully turbulent flow. Nature 526,
is Lcontrol = 385D. The Perspex pipe is connected to a linear actuator (toothed belt 550–553 (2015).
axis with a roller guide driven by a servomotor, ELGA-TB-RF-70-1500-100H-P0, 26. Willis, A. & Kerswell, R. Turbulent dynamics of pipe flow captured in a
Festo; not shown in the figure). The linear actuator can move the Perspex pipe reduced model: puff relaminarization and localized ‘edge’ states. J. Fluid Mech.
for an adjustable distance (traverse path) s ≤ smax = 1.5 m at an adjustable velocity 619, 213–233 (2009).
Upipe ≤ Upipe,max = 5.5 m s–1. The maximum acceleration is a = 50 m s-2. The resulting 27. Jimenez, J. & Pinelli, A. The autonomous cycle of near-wall turbulence.
wall velocity of the Perspex pipe is specified as a ratio to the mean flow velocity J. Fluid Mech. 389, 335–359 (1999).

Nature Physics | www.nature.com/naturephysics

© 2018 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

Vous aimerez peut-être aussi