You are on page 1of 12

Article

Turbulence suppression by cardiac-cycle-


inspired driving of pipe flow

https://doi.org/10.1038/s41586-023-06399-5 D. Scarselli1, J. M. Lopez1,2, A. Varshney1,3 & B. Hof1 ✉

Received: 13 December 2022

Accepted: 4 July 2023 Flows through pipes and channels are, in practice, almost always turbulent, and the
Published online: 6 September 2023 multiscale eddying motion is responsible for a major part of the encountered friction
losses and pumping costs1. Conversely, for pulsatile flows, in particular for aortic
Check for updates
blood flow, turbulence levels remain low despite relatively large peak velocities. For
aortic blood flow, high turbulence levels are intolerable as they would damage the
shear-sensitive endothelial cell layer2–5. Here we show that turbulence in ordinary
pipe flow is diminished if the flow is driven in a pulsatile mode that incorporates all
the key features of the cardiac waveform. At Reynolds numbers comparable to those
of aortic blood flow, turbulence is largely inhibited, whereas at much higher speeds,
the turbulent drag is reduced by more than 25%. This specific operation mode is more
efficient when compared with steady driving, which is the present situation for
virtually all fluid transport processes ranging from heating circuits to water, gas and
oil pipelines.

Turbulent flows are ubiquitous in nature and applications and are associ- In these studies, an initial steady turbulent flow undergoes a periodic
ated with large friction levels and high pumping costs when compared change in fluid speed, and the statistical properties of the evolving flow
with laminar conditions. Available estimates show that around 10% of are investigated. Flow acceleration typically delays the production
global electric power is consumed for pumping fluids6. In this context, of turbulent kinetic energy and decreases the wall shear stress with
turbulence not only is encountered at large scales, such as in oil or gas respect to the quasi-steady value. Deceleration, on the other hand,
pipelines, but also equally dominates flows in domestic settings (for enhances friction, although at higher deceleration rates, there is evi-
example, in heating pipes or the flow from a faucet). Even aortic blood dence of friction reduction27. Recent numerical studies28,29 identified
flow in humans and large mammals periodically exceeds transition unsteady driving conditions that can result in considerable drag reduc-
thresholds. Compared with a transition Reynolds number of Rec ≈ 2,040 tion, although the proposed numerical strategies are not necessarily
(ref. 7), aortic peak Reynolds numbers in humans reach more than twice8,9 straightforward to implement in practice.
this value, whereas in equine aortas peak values of 10,000 are common10. In this study, we present an alternative approach to turbulence con-
In the cardiovascular context, high turbulence levels constitute a severe trol in which drag reduction is achieved by means of unsteady, pulsatile
health hazard, as intense fluctuations and varying shear stresses are driving, specifically mimicking the cardiac cycle and extending this
attributed to endothelial cell dysfunction and arteriosclerosis2–5. method to large Reynolds numbers.
In engineering applications, in addition to the excessive drag levels, Experiments are carried out in a 1.2-m-long pipe (inner diameter
fluctuations and alternating shear stresses can equally have adverse D = 10 mm), and water is driven through the set-up by a piston. The
effects, and much effort has been dedicated to developing means to piston speed is accurately controlled by a servo motor that enables us
control turbulence. However, despite extensive efforts11, a broadly to alter the flow rate and in particular to realize a pulsatile flow of the
applicable method has not been discovered thus far. Active control desired waveform. For further details of the experimental set-up, see
techniques12–14 require complex actuation devices15,16, and in experimen- section ‘Experimental set-up’. Initial experiments were carried out at
tal realizations, the costs often far exceed the gains. Passive approaches moderate Reynolds numbers—values that are comparable to those in
equally suffer from high implementation costs and typically have a aortic blood flow. We compare three flows in the same pipe set-up at
limited operation range1,17. Additives, such as long-chain polymers, identical instantaneous Reynolds numbers (Re = UmD/ν, where Um is the
degrade over time18,19 and contaminate the liquids. Available control instantaneous bulk speed in the pipe and ν is the kinematic viscosity).
techniques are hence problem specific and intrusive, requiring either In the first case, the flow is driven steadily at Re = 2,800 (Fig. 1a) and the
manipulation of fluid properties or costly and often impractical imple- fluid motion is turbulent throughout. In the second experiment, the flow
mentations. Conversely, aortic flow provides an example in which a is driven periodically using the waveform (Fig. 1d) reported for cardio-
specific propulsion scheme, consisting of impulsive bursts separated vascular flow in the descending part of a human aorta30, choosing a peak
by quiescent intervals, seems to prevent turbulence despite relatively value close to the maximum values reported in the literature31. Although
large peak velocities. the pipe set-up is unchanged (including the inlet condition), the flow is
The effect of unsteady, pulsatile driving on turbulence has been fully laminar despite instantaneous Reynolds numbers being larger than
extensively investigated in experiments and numerical simulations20–26. 5,000 (Fig. 1b). In the third case, we tested a cycle in which the diastolic

1
Institute of Science and Technology Austria, Klosterneuburg, Austria. 2Universidad de Málaga, Málaga, Spain. 3School of Physical Sciences, National Institute of Science Education and Research,
HBNI, Bhubaneswar, India. ✉e-mail: bhof@ist.ac.at

Nature | Vol 621 | 7 September 2023 | 71


Article
a

30D
d e
6,000 6,000
5,000 5,000
4,000 4,000

Re
Re

3,000 3,000
2,000 2,000
1,000 1,000
0 0
0 0.5 1.0 1.5 2.0 0 0.2 0.4 0.6 0.8 1.0
t (s) t (s)

Fig. 1 | Decay of turbulence in aortic flow. a–c, Three instantaneous snapshots with (b,d) and without (c,e) a diastolic rest phase. d,e, For the periodic
of flows at Re = 2,800. Images capture flow structures in a 30D-long area waveforms, the minimum Reynolds numbers are Remin = 270 (d) and 180
located 60D downstream of the pipe inlet. In a, the flow is driven steadily and (e), whereas the maximum values are Remax = 5,300 (d) and 3,500 (e). The
the Reynolds number is constant. In b and c, the flows are driven periodically corresponding cycle-averaged values are Re = 1,730 (d) and 1,890 (e).

rest phase was removed (Fig. 1e). Compared with the cardiovascular equation (1). To ease the comparison of different cycles, τw is non-
case (Fig. 1d), we downscaled the peak velocity by a factor of 1.5 so that dimensionalized by the dynamic pressure at the minima of the cycles,
2
the average Reynolds numbers of the two cycles remain comparable. corresponding to (0.5ρU min ), where Umin is the bulk velocity at
2
In this case (Fig. 1c), the flow generally remains turbulent, which hints Re = 3,200. We define τ * = 2τw /(ρU min ) and in Fig. 2d the instantaneous
at the relevance of the diastolic rest phase for turbulence suppression. experimental values (blue circles) are compared with the quasi-steady
From the above experiments, it is apparent that unlike for steady reference case, τ*qs (black line), that is, the wall shear stress expected if
driving, the state of the flow—that is, laminar or turbulent—is not solely turbulence instantaneously adjusts to changes in Reynolds number.
determined by the instantaneous Reynolds number and that the wave- At the beginning of each cycle, τ*, although low, is considerably larger
form has a decisive role. How turbulence develops during a cycle also than τ*qs. Only as the flow acceleration proceeds, do the measured
depends on the initial fluctuation level at the beginning of the cycle, values eventually fall below τ*qs, in line with previous observations of
and hence on the history of the flow. In this respect, the diastole has a drag reduction during flow acceleration.
central role, as it effectively decouples the acceleration from the previ- Although the instantaneous wall shear stress values indicate an
ous deceleration, enabling turbulence and fluctuations to decay before overall drag reduction compared with the drag in quasi-steady case,
the Reynolds number increases again. In the following, we investigate this does not necessarily indicate drag reduction compared with a
the effect of the pulsatile operation mode on a fully turbulent flow at steadily driven flow of identical average Reynolds number. The drag
substantially larger time-averaged Reynolds numbers Re = U D /ν (where change with respect to the steadily driven flow is
U is the bulk speed averaged over one pulsation period, D is the pipe
diameter and ν is the kinematic viscosity of the fluid) to investigate the *
τ steady − τ*
R= , (2)
impact of pulsation on drag. For this, we pump water through a 7-m τ *steady
long pipe (inner diameter D = 30 mm) by means of the same syringe
set-up described above. The pressure drop Δp is measured across a where the overbar denotes an average over several cycles, and the
length L = 120D after a development length of 60D from the pipe inlet. steady-flow wall shear stress τ*steady is obtained from the Blasius33
Subsequently, the wall shear stress τw is reconstructed by using the friction-factor relation and normalized in the same manner as that for
force balance in the streamwise direction: the cyclic flow. For the cycle shown in Fig. 2a, the drag in experiments
is 4.4% larger (R = −0.044) than that of the steady flow. We note that
dUm ∆p 4τw the quasi-steady case τ*qs generally has a drag considerably larger than
ρ =− − , (1)
dt L D that of the actual steady flow. In the present case, it results in a 14%
increase in drag compared with τ*steady. Pulsation, therefore, does not
where ρ is the water density and Um is the instantaneous bulk flow necessarily lead to drag reduction, let alone energy saving. Inspired
velocity. Experiments are accompanied by direct numerical simula- by the diastolic phase found in aortic flow and the transition delay
tions (DNS) of the Navier–Stokes equations in which identical time obtained for the cardiac waveform, we designed a new cycle in which
variation of the Reynolds number is imposed. The DNS are performed a region of constant Reynolds number (rest phase) is inserted that
for a 5D-long pipe with periodic boundary conditions using a parallel effectively decouples the deceleration from the consecutive accelera-
solver32 (nsPipeFlow; Methods). tion phase (Fig. 2b). Remarkably, the flow now responds with consider-
In initial experiments and simulations, we tested a cycle consisting ably lower values of τ* during acceleration, as well as during part of the
of a series of linear flow-rate ramps smoothly joined together, cor- deceleration phase (Fig. 2e). The peak value of τ* is reduced by a factor
responding to Reynolds numbers oscillating between Remin = 3,200 of two, and in this case, we obtain a net drag reduction of 23% (R = 0.23).
and Remax = 18,800 with a period T = 4.5 s (Fig. 2a). Note that for steady The central role of the rest phase can be understood as follows. During
driving, flows at Re = 3,200 are fully turbulent in our pipe set-up. acceleration, turbulence remains initially frozen, that is, variations in
From measurements of the pressure drop Δp(t) and the imposed the mean velocity have a minimal impact on the turbulent stresses,
bulk velocity Um(t), we can determine the wall shear stress using leading to significant drag reduction. The amount of drag reduction

72 | Nature | Vol 621 | 7 September 2023


a b c
20,000 20,000 20,000

15,000 15,000 15,000


Re

Re
10,000 10,000

Re
10,000

5,000 5,000 5,000

0 0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08
Dimensionless time (t × 4Q/D2) Dimensionless time (t × 4Q/D2) Dimensionless time (t × 4Q/D2)
d e f
0.25 0.25 0.25

0.20 0.20 0.20

0.15 0.15 0.15


W*

W*
0.10 W* 0.10 0.10

0.05 0.05 0.05

0 0 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0 0.02 0.04 0.06 0.08 0 0.02 0.04 0.06 0.08
Dimensionless time (t × 4Q/D2) Dimensionless time (t × 4Q/D2) Dimensionless time (t × 4Q/D2)

Fig. 2 | Friction reduction in pulsating flow and the effect of three different dimensionless wall shear stress τ* for experiments (blue circles) and DNS
cycles on wall shear stress. In all cases, Remin = 3,200 and Remax = 18,800. (red line). For comparison, the friction associated with the quasi-steady
a–c, In a, Re = 11,000 and in b and c, Re = 8,600. The respective Reynolds number flow is provided by the black line. The quasi-steady values are given by
modulation is imposed in experiments and in DNS. d–f, The measured τqs(t ) = 0.079Re(t )−0.25 U m(t ) 2 /U 2min (where Blasius friction scaling is assumed).

sensitively depends on the turbulence level at the beginning of this where Psteady is the power dissipated by the steadily driven, that is,
acceleration phase. By contrast, during deceleration, an inflection constant–Reynolds number reference flow and P is the time-averaged
point emerges in the velocity profile, causing turbulence levels that power dissipation of the pulsatile flow. The instantaneous values of the
typically exceed those expected for the instantaneous Reynolds power dissipation are given by P(t) = Δp(t)Q(t), where Q(t) is the volume
number, especially by the end of this phase. The subsequent rest phase flow rate, determined from the measured instantaneous piston speed.
enables turbulence levels to die down and sets a favourable initial con- Computing the power savings for the cycle of Fig. 2b yields a loss
dition for the next acceleration phase. (S = −0.03), notwithstanding the large drag reduction. The power loss
From an energy perspective, in unsteady flows a reduction of the is caused by the additional energy input required to accelerate the flow,
mean friction τ* is not sufficient to ensure that the power dissipation because an increase in flow rate requires the pressure gradient and
per unit length (P = QΔp/L, where Q is the volume flow rate) is lower hence the power to grow (equation (1)). The waveform of Fig. 2b is hence
with respect to steady conditions. To quantify this aspect, we introduce advantageous (compared with steady driving) in situations in which
the power saving high shear stresses are detrimental, as is the case for the endothelium,
but counterproductive if energy efficiency is the main incentive.
Although we have considered waveforms with lower acceleration and
Psteady − P
S= , (3) higher deceleration rates so far (Fig. 2a,b), the opposite holds for veloc-
Psteady ity waveforms in the aorta30. Correcting for this, we chose the waveform

a b
0.014 0.034 10

Blasius 5
0.032
0.012
0

0.030
Ta × 4Q/D2

–27% drag –5
0.010 Tr
f

g
s in –10
0.028 ea
cr
De
–15
0.008
0.026
–20

0.006 0.024 –25


0 0.005 0.010 0.015 8,000 9,000 10,000 11,000
Tr × 4Q/D2 Re

Fig. 3 | Optimization of power savings. a, Percentage of the power savings S as and the star those of Fig. 2c. b, Corresponding values of S represented in the
2
a function of the duration of the acceleration Ta and rest phase Tr. The white, traditional f–Re plane, where f = 2D ∆p /(ρU L) is the Darcy friction factor.
dashed line separates the regions of positive and negative S values. The circle The grey dashed line is the friction level of a steady turbulent flow (Blasius
represents the parameters for the cycles of Fig. 2a, the square those of Fig. 2b correlation).

Nature | Vol 621 | 7 September 2023 | 73


Article
displayed in Fig. 2c, with a higher acceleration rate, while the rest phase acknowledgements, peer review information; details of author contri-
is left unchanged. During the more rapid increase in Reynolds number, butions and competing interests; and statements of data and code avail-
friction initially increases slightly faster than for the waveform in Fig. 2b; ability are available at https://doi.org/10.1038/s41586-023-06399-5.
subsequently, however, the friction drops at the beginning of the decel-
eration phase (Fig. 2f). Here friction reaches levels comparable to the 1. Kühnen, J. et al. Destabilizing turbulence in pipe flow. Nat. Phys. 14, 386–390 (2018).
ones assumed during the rest phase, although at very high Reynolds 2. Davies, P. F., Remuzzi, A., Gordon, E. J., Dewey, C. F.Jr & Gimbrone, M. A. Turbulent fluid
number. This effect further improves drag reduction, which now reaches shear stress induces vascular endothelial cell turnover in vitro. Proc. Natl Acad. Sci. USA
83, 2114–2117 (1986).
27% (R = 0.27). Computing the power balance, in this case, we obtain a 3. DePaola, N., Gimbrone, M. A. Jr, Davies, P. F.Jr & Dewey, C. F. Vascular endothelium
net savings of 9% (S = 0.09) compared with steadily driven pipe flow. responds to fluid shear stress gradients. Arterioscler. Thromb. J. Vasc. Biol. 12, 1254–1257
(1992).
Equally for the DNS, the cycle shown in Fig. 2c is the only one that results
4. Davies, P. F. Hemodynamic shear stress and the endothelium in cardiovascular
in drag reduction as well as energy saving. The amount of drag reduc- pathophysiology. Nat. Clin. Pract. Cardiovasc. Med. 6, 16–26 (2009).
tion in the DNS (R = 0.28) almost precisely matches those in experi- 5. Gimbrone, M. A.Jr & García-Cardeña, G. Endothelial cell dysfunction and the pathobiology
of atherosclerosis. Circ. Res. 118, 620–636 (2016).
ments, while the energy saving (S = 0.07) is slightly smaller. Taking into
6. Frenning, L. Pump Life Cycle Costs: A Guide to LCC Analysis for Pumping Systems
account that the DNS results were averaged over a much smaller number (Hydraulic Institute, Europump, and the US Department of Energy’s Office of Industrial
of cycles because of computational cost, small deviations from the Technologies, 2001).
7. Avila, K. et al. The onset of turbulence in pipe flow. Science 333, 192–196 (2011).
experimental result are not surprising (see Extended Data Table 2 and
8. Ku, D. N. et al. Blood flow in arteries. Annu. Rev. Fluid Mech. 29, 399–434 (1997).
Extended Data Fig. 3 for a comparison of experiments and DNS). 9. Stalder, A. F. et al. Assessment of flow instabilities in the healthy aorta using flow-sensitive
Finally, we investigate how changing the acceleration and rest phase MRI. J. Magn. Reson. Imaging 33, 839–846 (2011).
10. Nerem, R. M., Rumbergerr, J. A. Jr, Gross, D. R., Hamlin, R. L. & Geiger, G. L. Hot-film
affects drag reduction and power savings. Therefore, we carried out 225
anemometer velocity measurements of arterial blood flow in horses. Circ. Res. 34,
experiments spanning different rest phase and acceleration durations 193–203 (1974).
(denoted, respectively, by Tr and Ta), while keeping the minimum and 11. Brunton, S. L. & Noack, B. R. Closed-loop turbulence control: progress and challenges.
Appl. Mech. Rev. 67, 050801 (2015).
maximum Reynolds number and the combined duration of the accel-
12. Karniadakis, G. & Choi, K.-S. Mechanisms on transverse motions in turbulent wall flows.
eration and deceleration phases (T = 0.024 × ν/D2) constant (Extended Annu. Rev. Fluid Mech. 35, 45–62 (2003).
Data Fig. 4). The resulting map of power savings S is shown in Fig. 3a. 13. 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).
In Fig. 3a, the white dashed line separates the regions of positive and 14. Nakanishi, R., Mamori, H. & Fukagata, K. Relaminarization of turbulent channel flow using
negative S values and the cycles of Fig. 2a–c are denoted by a circle, traveling wave-like wall deformation. Int. J. Heat Fluid Flow 35, 152–159 (2012).
square and star, respectively. Shorter acceleration times consistently 15. Auteri, F., Baron, A., Belan, M., Campanardi, G. & Quadrio, M. Experimental assessment of
drag reduction by traveling waves in a turbulent pipe flow. Phys. Fluids 22, 115103 (2010).
lead to higher power savings, suggesting the importance of a brief, 16. Kasagi, N., Suzuki, Y. & Fukagata, K. Microelectromechanical systems-based feedback
intense acceleration followed by a longer, more gentle deceleration. control of turbulence for skin friction reduction. Annu. Rev. Fluid Mech. 41, 231–251
Specifically, for power saving, we find that the acceleration phase has (2009).
17. García-Mayoral, R. & Jiménez, J. Drag reduction by riblets. Philos. Trans. R. Soc. Math.
to be much shorter (≲1%) than the viscous time scales of the flow. This Phys. Eng. Sci. 369, 1412–1427 (2011).
abrupt change prevents the adjustment of the flow profile to its 18. Virk, P. S., Mickley, H. S. & Smith, K. A. The ultimate asymptote and mean flow structure in
(high-drag) quasi-steady shape. Strictly for the parameter regime Toms’ phenomenon. J. Appl. Mech. 37, 488–493 (1970).
19. Choueiri, G. H., Lopez, J. M. & Hof, B. Exceeding the asymptotic limit of polymer drag
investigated, a non-zero rest phase is required to save power. However, reduction. Phys. Rev. Lett. 120, 124501 (2018).
there is an optimal rest phase, and longer rest phases are counterpro- 20. He, S. & Jackson, J. D. A study of turbulence under conditions of transient flow in a
ductive. The optimal value of Tr depends weakly on Ta, and it is approx- pipe. J. Fluid Mech. 408, 1–38 (2000).
21. Greenblatt, D. & Moss, E. A. Rapid temporal acceleration of a turbulent pipe flow. J. Fluid
imately equal to half the duration of the unsteady part of the cycle Mech. 514, 65–75 (2004).
(t × 4ν/D2 ≈ 0.01). Remarkably, withT heart
r × 4ν /D 2 ≈ 0.012, the rest phase 22. He, S., Ariyaratne, C. & Vardy, A. E. Wall shear stress in accelerating turbulent pipe
observed for the aortic cycle in humans is close to this value. flow. J. Fluid Mech. 685, 440–460 (2011).
23. He, S. & Seddighi, M. Turbulence in transient channel flow. J. Fluid Mech. 715, 60–102
The same parameter space can be mapped to the usual f–Re plane, (2013).
2
where f = 2D∆p /(ρU¯ L) is the Darcy friction factor (Fig. 3b), to high- 24. He, K., Seddighi, M. & He, S. DNS study of a pipe flow following a step increase in flow
light the effect of the cycles on the drag reduction R and the dependence rate. Int. J. Heat Fluid Flow 57, 130–141 (2016).
25. Mathur, A. et al. Temporal acceleration of a turbulent channel flow. J. Fluid Mech. 835,
on Re. For comparison, we plot the Blasius relation for turbulent fric- 471–490 (2018).
tion. The largest reduction in f (27% drag reduction) is found for 26. Cheng, Z., Jelly, T. O., Illingworth, S. J., Marusic, I. & Ooi, A. S. H. Forcing frequency
Re ≈ 8,600 and is close to the region of maximum S. effects on turbulence dynamics in pulsatile pipe flow. Int. J. Heat Fluid Flow 82, 108538
(2020).
The circulatory system manages to combine flow speeds, signifi- 27. Ariyaratne, C., He, S. & Vardy, A. E. Wall friction and turbulence dynamics in decelerating
cantly exceeding the onset values of turbulence, with low shear-stress pipe flows. J. Hydraul. Res. 48, 810–821 (2010).
levels. Sufficient flow rates are crucial for a functioning organism, and 28. Kobayashi, W., Shimura, T., Mitsuishi, A., Iwamoto, K. & Murata, A. Prediction of the drag
reduction effect of pulsating pipe flow based on machine learning. Int. J. Heat Fluid Flow
the stress levels have to remain tolerable for the endothelial cell layer of 88, 108783 (2021).
the blood vessels. As we have shown, the waveform of the cardiac cycle 29. Foggi Rota, G., Monti, A., Rosti, M. E. & Quadrio, M. Saving energy in turbulent flows with
unsteady pumping. Sci. Rep. 13, 1299 (2023).
is close to the optimal value to achieve both of these objectives. A rest
30. Bürk, J. et al. Evaluation of 3D blood flow patterns and wall shear stress in the normal and
phase during the cycle is crucial to diminish wall shear stress, and at dilated thoracic aorta using flow-sensitive 4D CMR. J Cardiovasc. Magn. Reson. 14, 84
the same time, this rest phase has to be optimally timed and combined (2012).
31. Stein, P. D. & Sabbah, H. N. Turbulent blood flow in the ascending aorta of humans with
with a subsequent rapid flow acceleration to not only reduce the flow
normal and diseased aortic valves. Circ. Res. 39, 58–65 (1976).
drag but also optimize efficiency and minimize power consumption. 32. López, J. M. et al. nsCouette—a high-performance code for direct numerical simulations
Fluid transport is one of the largest sources of energy consumption of turbulent Taylor–Couette flow. SoftwareX 11, 100395 (2020).
33. Blasius, H. in Mitteilungen über Forschungsarbeiten auf dem Gebiete des Ingenieurwesens
in present-day societies, and a notable part of pumping costs can be
Vol. 131, 1–41 (Springer, 1913); https://doi.org/10.1007/978-3-662-02239-9_1.
attributed to turbulence. Although pipeline flows are commonly run at
a steady flow rate, our study demonstrates that, from an energy point Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
of view, this is not necessarily the optimal operation mode.
Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this
article under a publishing agreement with the author(s) or other rightsholder(s); author
Online content self-archiving of the accepted manuscript version of this article is solely governed by the
terms of such publishing agreement and applicable law.
Any methods, additional references, Nature Portfolio reporting sum-
maries, source data, extended data, supplementary information, © The Author(s), under exclusive licence to Springer Nature Limited 2023

74 | Nature | Vol 621 | 7 September 2023


Methods a linear actuator driven by a servo motor (Festo, ESBF-BS-80-1500-15P,
and Festo, EMMS-AS-70-M-LS-RS, not shown in Extended Data Fig. 1).
Direct numerical simulations A computer is used to control the motor and thus the plunger speed
We solve the incompressible Navier–Stokes equations in cylindrical within an accuracy of ±0.01 mm s−1. The syringe has an internal diam-
coordinates in a pipe of length 5D with periodic boundary conditions eter of Dc = 125 ± 0.11 mm and a total length Lc = 1,500 ± 0.1 mm, cor-
at the extremities. The equations are written in non-dimensional units responding to a maximum run time of about 870 advective time units
by using the pipe radius, D/2, as the length scale, the viscous time, D2/4ν, (D/U) for the chosen pipe diameter. The development of turbulence is
as the time scale, and 2ν/D as the velocity scale. These equations take ensured by perturbing the flow at the pipe inlet with a pin and letting
the following form: the flow develop for 60 D. Differential pressure is measured over the
subsequent 120 D with a carefully calibrated pressure transducer, full
∂t u + (u ⋅ ∇u) = − ∇p + ∇2 u, (4) scale 2.5 kPa. The wall taps (diameter d = 0.5 mm) are drilled through the
poly(methyl methacrylate) flanges and have been polished to remove
∇ ⋅ u = 0, (5) any burr. Water temperature is monitored at the outlet of the pipe with
a Pt-100 probe (indicated as ‘T’ in Extended Data Fig. 1) and is typically
To impose a time-varying Reynolds number, the mean velocity is held constant within ±0.05 °C. In a typical measurement run, the desired
updated at every time step, namely, flow rate waveform is repeated cyclically while traversing the available
stroke length. Temperature is measured in real time to compute the
Um,new = Um,old + α(t)δt , (6) correct motor speed and impose the correct Reynolds number. The
control and acquisition frequencies are set to 50 Hz. Depending on
where δt is the time step and α(t) is a prescribed acceleration rate. An the period duration, single runs consist of between 10 and 12 cycles. To
axial forcing term is then added to equation (4) to enforce that the inte- ensure a proper statistical representation of the unsteady friction, each
gral of the instantaneous velocity profile yields the mean flow, that is, cycle is then repeated several times (100 times for the results of Fig. 2
and 50 times for the parametric study of Fig. 3). The pressure signal
1
Um = ∫ 2u(r )r dr . (7) has been filtered to attenuate oscillations due to set-up vibrations by
0 using a cut-off frequency of 5 Hz. The first cycle was found to be sys-
tematically different from the others and has been excluded from the
Simulations were carried out using the custom, highly scalable, averaging process. Overall, the drag reduction R and the power savings
pseudo-spectral solver nsPipeFlow. The codes use Fourier–Galerkin S are estimated with a 2σ accuracy of ±1.8% and ±2.2%, respectively.
expansions along the axial and azimuthal directions, and eighth-order,
finite central differences for the radial dimension, collocated on a Calibration
Gauss–Lobatto–Chebyshev grid. The equations evolve in time with a To ensure repeatable and accurate differential pressure measurements,
second-order, predictor–corrector algorithm and a time step dynami- we calibrate the pressure sensor immediately before starting a batch
cally adjusted to satisfy the Courant–Friedrich–Lewy condition. Typi- of measurements (a typical batch consists of 10 runs with a minimum
cal values of the time-step size in our simulations range from 10−8 to of 10 cycles each). For calibration, we measure the pressure drop along
10−10 viscous units. Further details about the code implementation are the test section Δp for five values of the Reynolds number. Each steady
available in a previously published work32. As the Reynolds number measurement is repeated five times and the values of Δp, the piston
changes over time by more than one order of magnitude, the code speed in the cylinder Uc and the water temperature T are recorded.
can adaptively change the grid spacing to match the required spatial The reference value of Δp is computed for each Reynolds number by
resolution. More specifically, the adaptive-grid method ensures that using the Blasius formula,
the spatial resolution in the axial and azimuthal directions is consist-
ently below 7.5 wall units, whereas the maximum and minimum spac-
1 L
∆p = 0.316Re−0.25 ρU 2 , (8)
ing in the radial direction is below 3 and 0.1 wall units, respectively. 2 D
This spatial resolution is more stringent than that customarily used in
DNS studies of pipe flow at steady Reynolds numbers. Typical values where U = UcDc2/D 2 is the mean flow velocity in the pipe, Dc is the dia­
found at the minimum and maximum Reynolds numbers are given in meter of the cylinder, L is the length of the test section and ρ is the
Extended Data Table 1. density of water derived from the temperature T by following the pro-
The number of cycles needed to achieve statistical convergence cedure described previously34. As a result, a calibration curve is
depends on the pulsation waveform. For the case shown in Fig. 2d, obtained to convert the sensor output (in volts) to a differential pres-
cycles have little variation, and convergence occurs quickly. Statistics sure (in pascals). To assess the validity of the calibration, we compute
were computed in this case using four cycles. For the cases shown in the residuals of the linear fit and take the maximum value. In the case
Fig. 2e,f, there is more variability among cycles, and it is necessary of the optimal cycle (Fig. 2c,f), we find a maximum deviation of about
to average over more cycles to obtain converged statistics. For the 10 Pa, which is well representative of the values found throughout the
waveform shown in Fig. 2e, statistics were obtained averaging over experi­mental set-up.
nine cycles, whereas for the waveform shown in Fig. 2f, 14 cycles
were used. Standard deviation of mean pressure and flow-rate measurements
The mean pressure drop during a cycle is estimated by taking the sample
Experimental set-up average of the signal Δp(t), namely,
We use a large-scale, customized syringe pump (Extended Data Fig. 1) to
N
control the flow rate precisely and impose an arbitrary modulation of 1
∆p =
N
∑ ∆pi , (9)
the Reynolds number. The test section consists of a 7-m-long precision i =1
bore glass pipe (Duran, KPG; internal diameter D = 30 ± 0.01 mm) made
by joining 1-m-long segments with custom poly(methyl methacrylate) where Δpi is the ith sample Δp(ti) and N is the number of samples per
flanges (in the experiments of Fig. 1, the test section consists of a single cycle. The pressure signal recorded for the optimal waveform is shown
pipe segment with D = 10 ± 0.01 mm and length 1.2 m). Water flows in Extended Data Fig. 2b. The blue curves correspond to 100 instanta-
through the pipe into a reservoir as the syringe pump is displaced by neous cycles and the phase average is shown in red. For comparison,

    
Article
we superimpose the computed pressure drop from DNS (dotted line). histograms of the 100 cycles measured experimentally are shown in
Extended Data Fig. 2a shows the corresponding time dependence of Extended Data Fig. 3a,b. Because (owing to computational costs) fewer
the Reynolds number based on the recording of the piston position. cycles were simulated, instead of histograms, we computed the mean
Using equation (9), we determined the mean pressure for all 100 values for the DNS for R and S (orange dashed lines).
cycles and found that the standard deviation between the cycles Finally, we report in Extended Data Table 2 the average values for R
amounts to 2.9% of the mean. This value provides an upper bound on and S for all three waveforms in Fig. 2 for experiments and DNS, and
the measurement error involved. We note that because of the unsteady, they are in close agreement.
chaotic nature of the flow, consecutive cycles start from different ini-
tial conditions. This leads to a natural variation of cycle-averaged quan- Waveforms
tities such as mean pressure. Therefore, the standard deviation between The waveforms considered are composed of linear ramps in Reynolds
cycles would be expected to be non-zero even in the absence of meas- number and periods of constant flow rate. Throughout the experi-
urement error. The error of the mean flow rate Q can be estimated from ments, the minimum and maximum Reynolds numbers are held con-
positioning measurements of the piston and the manufacturing accu- stant and equal to Remin = 3,200 and Remax = 18,800, respectively. The
racy of the piston cylinder to be 0.6%. For the power input curves combined duration of acceleration and deceleration T is always fixed
(Extended Data Fig. 2c), the standard deviation is 3.1%. to 4.5 s, whereas the duration of the acceleration and rest phase are,
respectively, varied in the intervals Ta ∈ [1.35, 3.15] s and Tr ∈ [0, 4] s.
Computing drag reduction from experimental measurement To avoid abrupt changes in the piston acceleration, the sudden slope
The drag reduction rate R can be computed by integrating equation (1) changes that occur at the transition points from acceleration to decel-
over integer multiples of the period T of the waveform. eration to rest phase have been locally smoothed with a moving average
filter of width 0.8 s (Fig. 2a–c).
mT mT
dU ( t )  ∆p(t) 4τw(t) 
∫0 ρ
dt
dt = − ∫ 
0  L
+
D 
dt , (10)
Data availability
where we assume incompressible flow and make use of the fact that The datasets generated and analysed during the current study are
the integral of the bulk flow over a period is zero freely available in the Zenodo repository (https://doi.org/10.5281/
zenodo.7828996)35.
mT mT
∆p(t) 4τw(t)
∫0 L
dt = − ∫
0 D
dt , (11)

Code availability
We can, therefore, rewrite equation (2) to express the drag reduction The numerical simulations were conducted with the open-source code
rate R in terms of the time-averaged pressure: nsPipeFlow, distributed under the terms of the GNU General Public
License v.3. A detailed description of the code and user guide are
∆psteady − ∆p provided in ref. 30. The code version used in this study and an initial
R= , (12)
∆psteady condition to start the simulations are openly available in the Zenodo
repository (https://doi.org/10.5281/zenodo.7828996)35.
where Δpsteady is computed for the mean Reynolds number Re of the
cycle by using equation (8).
34. Tanaka, M., Girard, G., Davis, R., Peuto, A. & Bignell, N. Recommended table for the
As a consequence, the standard deviation in ∆p (2.9% in the case of density of water between 0 °C and 40 °C based on recent experimental reports.
the optimal cycle) provides an upper bound for the uncertainty in R. Metrologia 38, 301 (2001).
35. Scarselli, D., Lopez, J. M., Varshney, A. & Hof, B. Turbulence suppression by cardiac cycle
inspired pulsatile driving of pipe flow: datasets and numerical code used to perform the
Estimation of power savings
simulations. Zenodo https://doi.org/10.5281/zenodo.7828996 (2023).
Estimating the power savings S is also straightforward, as it requires
taking time averages of the power P(t) = Δp(t)Q(t), where Q(t) is the
Acknowledgements We acknowledge the assistance of the Miba machine shop and the
volume flow rate, and using equation (3), reported here for clarity: team of the ISTA-HPC cluster. We thank M. Quadrio for the discussions. The work was
supported by the Simons Foundation (grant no. 662960) and by the Austrian Science Fund
Psteady − P (grant no. I4188-N30), within Deutsche Forschungsgemeinschaft research unit FOR 2688.
S= . (13)
Psteady
Author contributions B.H. supervised the project. D.S. and A.V. designed and performed
the experiment. D.S. analysed the experimental data. J.M.L. designed and performed the
The accuracy of S can be estimated in the same way we described for computer simulations of the Navier–Stokes equations and analysed the numerical results.
R. We compute the mean power P for each of the 100 cycles and then D.S., J.M.L., A.V. and B.H. wrote the paper.
estimate the standard deviation between the cycles. We find a value of
3.2%, which also represents the accuracy of S because in equation (13) Competing interests The authors declare no competing interests.

P is the only uncertain quantity.


Additional information
Comparison with DNS Correspondence and requests for materials should be addressed to B. Hof.
Peer review information Nature thanks Yongyun Hwang and the other, anonymous, reviewer(s)
Selecting the optimal waveform (Fig. 2c), we compare the values of for their contribution to the peer review of this work.
R and S obtained from experiments with those observed in DNS. The Reprints and permissions information is available at http://www.nature.com/reprints.
Extended Data Fig. 1 | Sketch of the experimental setup. Drawing not to scale.
Article

Extended Data Fig. 2 | Waveform, pressure and power signal. Signals from the
optimal cycle (Fig. 2c) measured in experiments. (a) Waveform based on the
linear piston speed and (b) pressure drop Δp measured over the test section.
The signal from all the 100 cycles measured is shown in blue, while the phase
average is represented in orange. The number of samples per cycles is N = 325.
For comparison, we report also the pressure drop computed with the DNS for
the same cycle (gray dotted line). In this case the signal is obtained by phase-
averaging the available 8 cycles. (c) The power input for the same waveform.
The values for the 100 cycles in experiments are shown in blue, the ensemble
average in red and the power input in DNS is given by the grey doted line.
Extended Data Fig. 3 | Comparison of the values of (a) R and (b) S between experiments (blue histogram) and DNS. For the optimal cycle (Fig. 2c, main text) the
histogram shows the distribution of the values obtained from 100 runs. The orange, dashed line shows the mean of the available corresponding DNS cycles.
Article

Extended Data Fig. 4 | Definition of the flow cycle used in the experiments of Fig. 2 and Fig. 3 .
Extended Data Table 1 | Parameters used in the direct
numerical simulations

From left to right: Reynolds number Re based on the mean velocity, minimum and maximum
radial resolution (in inner units), azimuthal resolution (in inner units), axial resolution (in inner
units) and average time step size δt.
Article
Extended Data Table 2 | Comparison between the values of
R and S in percentage for the waveforms of Fig. 2 obtained
from experiments and DNS

You might also like