Sunteți pe pagina 1din 5

PAPER www.rsc.

org/softmatter | Soft Matter

A basic swimmer at low Reynolds number†


Marco Leoni,ab Jurij Kotar,b Bruno Bassetti,a Pietro Cicuta*b and Marco Cosentino Lagomarsino*a
Received 21st July 2008, Accepted 13th October 2008
First published as an Advance Article on the web 17th November 2008
DOI: 10.1039/b812393d

Swimming and pumping at low Reynolds numbers are subject to the ‘‘Scallop theorem’’, which states
that there is no net fluid flow for time-reversible motions. Microscale organisms such as bacteria and
cells are subject to this constraint, and so are existing and future artificial ‘‘nano-bots’’ or microfluidic
pumps. We study a very simple mechanism to induce fluid pumping, based on the forced motion of
three colloidal beads through a cycle that breaks time-reversal symmetry. Optical tweezers are used to
vary the inter-bead distance. This model is inspired by a theoretical swimmer proposed by Najafi and
Golestanian (A. Najafi and R. Golestanian, Phys. Rev. E, 2004, 69, 062901), but in this work the
relative softness of the optical trapping potential introduces a new control parameter. We show that
this system is able to generate flow in a controlled fashion, characterizing the model experimentally and
numerically.

1 Introduction Fig. 1 the trapping stiffness of the central bead in this work is
ktrap ¼ 4.1  1.3 pN/mm, whereas the external beads were held
Insight into swimming of microorganisms and bacteria can be more strongly, with ktrap, external ¼ 8.1  1.4 pN/mm. The errors
gained by studying the motion at low Reynolds numbers of quoted here are the standard deviation over several independent
experimental and theoretical model systems.1,2 Compared to the experiments, with the same beads and same tweezers configu-
macroscopic world, a striking feature of propulsion of these ration. They could also be a source for the slight discrepancy in
micron-scale objects, for which inertia is typically negligible, is relaxation times seen in Fig. 2(a) and (b). The sample is
that a body striving to move has to change its shape with time in illuminated with a halogen lamp and is observed in bright field
a non-reciprocal fashion.3 For example, swimming of a magnet- with a fast CMOS camera (Allied Vision Technologies, Marlin
ically driven semiflexible artificial filament was recently demon- F-131B). Silica beads of 3.0mm diameter (Bangs Labs) were
strated, by a wave-like motion similar to flagella.4 In this work we diluted to extremely low concentration to avoid any spurious
realize a much simpler system which uses only two effective beads falling into the laser trap. The set of three trapped beads
degrees of freedom, the distances between central and lateral was floated well above the glass slide surface (at about 10 times
spheres. This is inspired by a 3-bead linear chain theoretical the bead diameter) to minimize any hydrodynamic drag from
model5 which has never been realized experimentally. the solid surface. Except where specified differently, we used
a solution of glycerol (Fisher, Analysis Grade) 51% by weight in
2 Materials and methods water (Ultrapure grade, ELGA) which has a viscosity of
2.1 Optical tweezers 6.23 mPa s.6 Experiments were performed at 25  C.

The optical tweezers setup used in this work consists of a laser 2.2 Experimental protocol
(IPG Photonics, PYL-1-1064-LP, l ¼ 1064nm, Pmax ¼ 1.1W)
focused through a water immersion objective (Zeiss, Achroplan The experimental protocol is composed of two parts. In the first
IR 63x/0.90 W), trapping from below. The laser beam is steered calibration stage all the traps are kept at rest, and the beads
via a pair of acousto-optic deflectors (AA Opto-Electronic, undergo only Brownian motion confined by the traps. The driven
AA.DTS.XY-250@1064nm) controlled by custom built elec- dynamics occurs only during the second stage. Data is acquired
tronics, allowing multiple trap generation with sub-nanometre at 48.79 frames per second, with an exposure time 19  103s. A
position resolution. Instrument control and data acquisition are run lasts ten minutes, during which we collect about 30 000
performed by custom software. The trapping potential is locally frames, equally divided between calibration and dynamics. We
described by a harmonic spring, and the trap stiffness was collected 4 runs for each set of parameters. Images are analysed
calibrated by measuring the thermal displacements of the using a correlation filter with a kernel optimized to the bead
trapped beads. With reference to the bead arrangement seen in profile, followed by a 2-d least-square fit. This gives the center-
of-mass coordinates of the beads in each frame with an error of
a
Università degli Studi di Milano, Dip. Fisica and INFN, sez. Milano, Via the order of 103mm.
Celoria 16, 20100 Milano, Italy. E-mail: marco.cosentino@unimi.it
b
University of Cambridge, Cavendish Laboratory and Nanoscience Center,
JJ Thomson Avenue, CB3 0HE Cambridge, UK. E-mail: pc245@cam.ac.
2.3 Numerical simulations
uk
Numerical simulations were performed by integrating directly
† Electronic supplementary information (ESI) available: Additional
considerations on the model and the data analysis. See DOI: the equation of motion of the three beads with Taylor’s method.7
10.1039/b812393d We considered both simulations that included thermal motion

472 | Soft Matter, 2009, 5, 472–476 This journal is ª The Royal Society of Chemistry 2009
(i.e. adding a white noise term to the equations) and deterministic Despite its simplicity, the model encapsulates all the essential
ones, finding no relevant differences. The results in the body of characteristics of a swimmer. We study the flow generation on
the paper mainly refer to the deterministic model. the fluid, using the central bead as a probe. This corresponds to
studying the propulsion of a freely moving swimmer. By tracking
the position of the central bead it is possible to detect a left-right
3 Experimental symmetry breaking and quantify the net flow at varying
parameters.
We study a swimmer composed of three beads immersed in
a high viscosity Newtonian liquid. The beads are controlled by
an optical tweezer8,9 and interact with each other through the 3.1 Quantifying displacements
fluid. The laser traps are set-up following the scheme of
Fig. 1(a and b). The central trap is kept at rest, holding the Trapping lasers act with good approximation as harmonic
central bead in a soft harmonic potential of stiffness ktrap, while potentials on each bead. Thus, the average displacement hDxit of
the lateral ones switch between two positions. By forcing the the central bead from the position of the central trap can be
lateral beads to move, we implement the four time-reversal converted using Hooke’s law into the mean force exerted on the
symmetry breaking phases shown in Fig. 1(a). fluid by the whole system. However, a direct measurement of
Two of the parameters that characterize the system are purely hDxit is at the limit of experimental resolution. The limiting
geometric: 3, the maximum oscillation amplitude for the lateral factor is not the imaging resolution, but the presence of thermal
beads, and d, the distance between the laser positions at the fluctuations. This thermal noise can be compensated by longer
starting phase of the cycle. Their physical interpretation is that sampling, however to reduce this error down to the nanometre
3 sets the strength of the stroke, while d sets the magnitude of the scale would require experiments 102 times longer than ours,
hydrodynamic interaction. The additional parameter is t, the which is not practical.‡ Instead, we quantify the flow using
switching time of the laser, which corresponds to the temporal a different approach. The electronically controlled trap move-
length of each phase in the cycle. Finally, the viscosity h, which is ment temporization allows the phase within each cycle to be
set in the preparation of the sample, characterizes the fluid, the precisely kept track of. We can therefore average over the
hydrodynamic interactions, and the amplitude of thermal repeated realizations, and reconstruct the mean dynamic cycle of
motion. The former parameters, together with the radius R of the the beads. It is significant to focus on the central one, in the
beads, characterize completely the system. stationary trap: an example of the mean cycle for this bead is
shown in Fig. 2(a), where four peaks due to the lateral motion are
clearly visible. These peaks correspond to the maximum
displacement of the central bead in each phase of the motion. The
configurations of the three traps are indicated schematically in
the bottom panel of Fig. 2(a). Due to the different dispositions of
the outer beads, the drags the central bead is subject during each
phase are different and cause unequal displacements. Our
method characterizes the pumping by means of the asymmetry of
the peaks. On the mean cycle, we label p1, p2, p3, p4 (in order of
decreasing value) the peaks in the displacement, and we define
the following observables, devised to quantify the asymmetry of
the motion: (1) dinv :¼ (p1  p2)  (p4  p3), the difference
between the lower and the upper peak values, (2) d2 :¼ p1  p4,
the difference between the upper and the lower maximum peak
values. For an illustration we refer to the mid panel in Fig. 2(a).
The reason for using both definitions (1) and (2) to quantify
Fig. 1 Experimental realization of the low Reynolds number swimmer, asymmetry is the fact that the mean cycles have different shapes
and proof of force generation. (a) Scheme of the lasers’ disposition during for different values of t. For example, in the case of small values
a single phase of motion, in which the active traps (green) act on the beads of t, the intermediate peaks are not well distinguished, as shown
as harmonic potentials. The parameters 3 and d are sketched. (b) in the top panel of Fig. 2(a). Thus, we adopt the observable d2
Sequence of snapshots of the experiment showing four time-symmetry which makes use of the upper and the lower maximum peak
breaking steps in the basic cycle. Crosses indicate the active laser trap values only. In the opposite case of high values of t, see Fig. 2(a),
positions. By moving the lateral beads and keeping the central one at rest, the four peaks are well defined, and it is more effective to use the
it is possible to study the flow generated on the fluid, using the central observable dinv. The advantage is that dinv is determined inde-
bead as a probe. A left-right asymmetry in its position signals that flow is
pendently from the equilibrium position x0, which is needed for
generated. Top panel: starting configuration, where the distances between
d2 and cannot be measured during the active motion. Thus, it
beads take their maximum values. (c) Two-dimensional map of the
effective potential felt by the central bead. The plot is obtained by taking allows for more direct measurements. These observables lack
the logarithm of the relative frequencies of the positions occupied by the a simple physical meaning and should rather be thought of as
central bead during 500 repetitions of the basic cycle. The color scale order parameters useful in characterizing the asymmetry of the
indicates the most frequent position in a darker shade. The map shows
a left-right symmetry breaking (the left-hand minimum is deeper), which ‡ Already at times of the order of a few tens of minutes there can be
is a proof of the mean generated flow in the system. sources of mechanical noise leading to drift.

This journal is ª The Royal Society of Chemistry 2009 Soft Matter, 2009, 5, 472–476 | 473
Fig. 2 Reconstruction of the mean cycle. The experiment is a sequence of repetitions of the elementary cycle illustrated in Fig. 1. By treating each period
as a single realization and averaging over all realizations, one recovers the mean cycle of the dynamics, averaged over the thermal fluctuations due to
Brownian motion in the fluid. (a) Detail of the mean cycle for the central bead (> with errorbars) compared with simulations (solid lines) for different t.
The plot reports the longitudinal displacement coordinate of each bead as a function of the cycle phase in the interval (0, 2p). The peaks become more
defined as t increases. Top panel: t ¼ 40 ms; middle panel: t ¼ 80 ms; bottom panel: t ¼ 320 ms. The position of the four peaks and the traps’
configuration in the four phases of the cycle are illustrated in the middle and bottom panel respectively. (b) Mean cycle for the whole system (three
beads). The longitudinal displacement coordinate of each bead is plotted as a function of time. Dotted lines correspond to experimental values for the left
bead (black), the central bead (red) and the right bead (green). These are compared with the results of numerical simulation (solid lines with the same
color code). (c) Close up and complete cycle showing relaxation of a single bead in a switching trap, with no interacting partners.

peaks. Both these quantities change sign by reverting the the experiment for one trapped particle in water/glycerol, in
sequence of steps and allow verification as to whether the system a trap undergoing repeated displacements, see Fig. 2(c). Also
behaves as expected, pumping in the opposite direction. We also in this case, the simulation and experiment match very well. In
define a third observable (3) s :¼ |p1| + |p2| + |p3| + |p4|, the sum of conclusion, there is excellent agreement in the single particle
absolute values of the peaks. s is always positive, and quantifies regime, confirming experimental calibration and numerical
the amplitude of the swimmer’s motion. methods. Therefore the slight discrepancy in the three-bead data
must be due to the excessively simple theoretical description of
3.2 Comparison of experiments with simulations multiple interacting bodies.x In other words, the flow induced
by each bead is not perfectly represented by a linear superpo-
We compare the measured flow to simple numerical simulations sition of single particle Oseen propagators.10 This is not
(see methods). The hydrodynamic interaction between the surprising, considering that this is a long-distance, or far-field
spheres is described by the Oseen tensor,10 corresponding to the approximation. A common procedure to correct for this effect
limit of point force, or far-field. Along the xb axis of the swimmer, makes use of a perturbative expansion in the parameter R/d. We
the equation of motion for bead n is checked this by implementing the first and second perturbative
!
1 X 3R  corrections13,14 in our simulations, finding very similar quanti-
x_ n ¼ Fn þ Fm (1) tative results as with the Oseen tensor. The reason for this is
g nsm
2rnm that the parameter R/d is not small, being of the order of 1/3, so
that the real dynamics is not accessible perturbatively.{ Despite
with n, m ¼ 1, 2, 3, where g ¼ 6phR is the Stokes’ drag, Fi is the this limit, the simulations do allow us to compare the experi-
external force exerted on the ith particle, and rnm indicates the mental results with a prediction for the net force, or flow,
relative distance between n and m. The experimental values of h generated in the fluid by the swimmer. A mapping between each
and ktrap were used in the simulations. of the observables (d2, dinv) and the temporal average hDxit is
As can be seen by inspecting Fig. 2(b), the observed relaxa- obtained by means of simulation. Here sampling problems are
tion times are shorter in the experimental data of bead motions not present and hDxit is quantifiable directly by time averaging
than in the simulations at corresponding parameters. To
exclude possible errors, we performed the following tests. First,
we simulated a single particle in a stationary trap, including
x Note that we can exclude that inertia and the propagation time of
thermal fluctuations, and calculated the relaxation time in the hydrodynamic interaction play a role, as deviations were found to be at
time autocorrelation function. For an isolated particle, the the nano-second scale,11,12 well outside the time scales of our experiments.
relaxation time is the ratio ktrap/g. Using the experimental { While approaches that do not make use of the explicit form of the
solution of hydrodynamic equations15,16 may help in solving this
values in the simulations, we found excellent agreement for
problem, in general, to compensate for this small discrepancy one
beads both in pure water and water/glycerol. Subsequently, we would have to solve Stokes’ equation with the proper boundary
compared the simulation (with and without thermal noise) to conditions.17

474 | Soft Matter, 2009, 5, 472–476 This journal is ª The Royal Society of Chemistry 2009
with arbitrary accuracy; d2, dinv are obtained from the steady- this property holds. Fig. 3(a) shows a comparison between the
state mean cycle, similar to experiments. For both d2 and dinv, experimental and simulated observable dinv as a function of 3.
we verify that there exists a one to one mapping with hDxit. We The absolute value of the corresponding mean force is shown in
use this relation to associate as a function of 3 each point of the Fig. 3(b). The maximum mean forces reached in our experiments
experimental curve to the corresponding mean displacement, are of the order of 0.03pN, roughly corresponding to a net
obtaining hDxit as a function of 3 that is converted into a mean swimming/pumping speed of about 0.2mm/s. Note that this force
force via Hooke’s law. The procedure is illustrated in Fig. S1 of is exceedingly small to be measured with an optical tweezer using
the ESI.† conventional techniques, which has posed a barrier in past
investigations.18 The indirect technique used here enables this
barrier to be overcome, and can possibly be useful in different
4 Discussion
contexts.
Our first result is that the system is able to generate flow. This is It is interesting to study the influence of t, by fixing the value of
qualitatively visible in the distribution of the position of the 3, and in Fig. 3(c) we plot the propulsion force at varying t,
central bead (Fig. 1(c)). This effective potential is composed of compared with simulations. In both cases there is a good quan-
two contributions: the harmonic potential exerted by the twee- titative agreement between the model with Oseen interactions
zers, and the configurational bias induced by the interaction with and the experimental data. By increasing the distance d between
the lateral moving beads. The latter contribution causes the the beads, the asymmetry in the peaks becomes more difficult to
emergence of two effective asymmetric minima. This asymmetry detect. However it is still possible to characterize the dependence
can be quantified accurately using the above-defined observables of fluid pumping on the parameter d by using the amplitude
(Fig. 3). Simulations show that there exists a one-to-one map variable s. In Fig. 3(d), s is plotted versus the distance, for both
between dinv and d2, and the values for the mean force, so that it is simulations and data. In both cases this quantity decays as 1/d.
possible to convert dinv and d2 directly to forces. For fixed values Quantitatively, there is a small systematic deviation between
of t, we ran a series of experiments to characterize the depen- experiment and simulations. The 1/d decay directly depends on
dence of the mean position from the displacements 3 of the lateral hydrodynamic interactions, and can be understood with a simple
beads. Fig. 1(b) shows one of the two possible time-reversal argument on the maximum deviation of an initially resting
breaking sequences. By reverting the sequence into the specular trapped bead subject to the perturbation induced through the
one, the swimmer should move in the opposite direction. fluid by another bead at distance d relaxing for a stretch 3 in
Experimentally, we implemented both sequences, and found that another harmonic potential (see the ESI for further details†).

Fig. 3 Quantitative characterization of the swimmer. The main observable is the mean force felt by the central bead, measured by its mean displacement
in the optical trap. The mean force is studied here as a function of the parameters 3 and t. (a) shows dinv as a function of the lateral amplitude 3,
comparing experimental values (triangles) with simulation (continuous line). There are two possible time-reversal symmetry breaking cycles, corre-
sponding to the two opposite pumping directions. By temporally reversing the cycle shown in Fig. 1, dinv changes sign, showing that the direction of
pumping can be controlled. The following panels show experimental data (>) compared to simulations (continuous line). Each > represents the
average over 4 different experiments and the error bars correspond to the standard deviations of the four values. (b) shows the mean force (obtained from
the observable dinv and numerical simulations) as a function of 3, for d ¼ 6mm, t ¼ 80 ms, h ¼ 6.23 mPa s. (c) Mean force as a function of t. Here the other
experimental parameters are d ¼ 6mm, 3 ¼ 1mm, h ¼ 6.23 mPa s. (d) Amplitude of motion as a function of d. Increasing d the force becomes weaker and
difficult to detect, and we have to use the amplitude parameter s to compare with simulations. The measured and simulated values of s scale as d1
(as discussed further in the ESI†).

This journal is ª The Royal Society of Chemistry 2009 Soft Matter, 2009, 5, 472–476 | 475
4.1 Comparison with other model systems intrinsic variant being slightly faster and more efficient (ESI,
Fig. S3†).
It is instructive to compare our swimmer with the closely related
Finally, it may be of interest to compare this pump to other
theoretical model originally proposed in the literature5 by Najafi
engines. We can estimate the dissipation rate in our pump to be
and Golestanian, and the variants that have been recently
between 0.5  1017 and 4  1017J/s, which is comparable to an
explored.19,20 There are two main features characterizing our
E. coli bacterium swimming in water.24 The efficiency of this
model. Firstly, beads are subject to the compliance of the trap-
pump is of the order of 103% (calculated as the ratio of the mean
ping potentials. The original Najafi-Golestanian swimmer has
output to input power) which is lower than for typical biological
rigid links. More recently, the model has been solved for the case
systems such as E. coli.1,25 These results are important on
of chemical transitions between states19 and for general bead and
fundamental grounds and for applications in nano-scale
link sizes and imposed deformations.17,20 However, an actuated
machines.
movement by quadratic potentials has not been addressed
explicitly in the previous literature. While the behavior of the
simplest rigid Najafi-Golestanian swimmer is essentially char- Acknowledgements
acterized by the ratio 3/d,21 more general models can have more We thank Ray Goldstein, Christopher Lowe and Ignacio Pago-
complex dynamics, imposed by additional relevant time or nabarraga for useful comments and discussion. This work was
length-scales. Our system is also subject to an additional supported by the Korea Foundation for International Cooper-
t
parameter, the ratio , with t0 ¼ g/ktrap of the imposed ation of Science & Technology (KICOS) through a grant
t0
provided by the Korean MOST (No. 2007-00338) and by the UK
displacement time to the characteristic relaxation time of the
EPSRC.
bead. Hence the phenomenology of our system depends also on
the ratio of the two time scales. The most important consequence
is that our simulations show a power-law dependence of the References
speed, or mean force, on 3, with an exponent that increases 1 H. Berg, Physics Today, 2000, 53, 24.
monotonically with increasing t, slightly larger than the 2 M. Buchanan, Nature Phys., 2008, 4, 83.
quadratic law found for the Golestanian swimmer. Experimen- 3 E. Purcell, Am J Phys, 1977, 45, 3–11.
4 R. Dreyfus, J. Baudry, M. L. Roper, M. Fermigier, H. A. Stone and
tally, it is problematic to measure incontrovertibly this scaling, J. Bibette, Nature, 2005, 437(7060), 862–5.
due to the large errors in the fit for the exponents, whose average 5 A. Najafi and R. Golestanian, Phys Rev E Stat Nonlin Soft Matter
values, however, do exceed two. In the limiting case of small t, Phys, 2004, 69(6 Pt 1), 062901.
the two models behave in the same way. This can be understood 6 R. C. Weast and M. J. Astle, CRC Handbook of Chemistry and
Physics, CRC Press, Boca Raton, 60th edn, 1979.
in the following way. Assume for simplicity that the lateral 7 M. Cosentino Lagomarsino, P. Jona and B. Bassetti, Phys Rev E,
forcing beads can be described as free particles relaxing through 2003, 68(2 Pt 1), 021908.
the minimum of an harmonic potential. Their position follows 8 A. Ashkin, Proc Natl Acad Sci U S A, 1997, 94(10), 4853–60.
9 S. M. Block, Nature, 1992, 360(6403), 493–5.
a simple temporal law governed by an exponential decay x(t) z
10 M. Doi and S. Edwards, The Theory of Polymer Dynamics, Clarendon
x0 exp( t/t0), and accordingly the velocity v(t) z (x0/t0) exp( t/t0). Press, Oxford, 1986.
For small values of t, the ratio t/t0 is small ensuring that the 11 S. Henderson, S. Mitchell and P. Bartlett, Phys Rev Lett, 2002, 88(8),
velocity of the trapped bead does not decay significantly during 088302.
12 M. Atakhorrami, G. H. Koenderink, C. F. Schmidt and
each step in the cycle, being a constant proportional to t0 as in F. C. MacKintosh, Phys Rev Lett, 2005, 95(20), 208302.
the Najafi-Golestanian swimmer, where the beads have 13 J. Rotne and S. Prager, J Chem Phys, 1969, 50, 4831.
a constant displacement velocity.20,21 On the other hand, our 14 G. Batchelor, J Fluid Mech, 1982, 119, 379–408.
results confirm that in the small t limit the mean force scaling 15 D. Mazur and P. Bedeaux, Physica, 1974, 76, 235–246.
16 T. B. Liverpool and F. C. MacKintosh, Phys Rev Lett, 2005, 95(20),
with 3/d is compatible with a quadratic law. This regime is also 208303.
the one where the swimmer’s propulsion is optimal. In the 17 F. Alouges, A. DeSimone and A. Lefebvre, J Nonlinear Sci, 2007, 18,
numerical experiments, the fitted exponents we find are consis- 277–302.
18 D. Riveline, C. Wiggins, R. Goldstein and A. Ott, Phys Rev E, 1997,
tently larger than 2 (between 2.12 and 2.41). However, the pre-
56, R1330–R1333.
dicted trend with 3 cannot be observed experimentally because of 19 R. Golestanian and A. Ajdari, Phys Rev Lett, 2008, 100(3), 038101.
the large errors (ESI, Fig. S2†). The second difference with the 20 R. Golestanian and A. Ajdari, Phys. Rev. E, 2008, 77, 036308.
original Najafi-Golestanian swimmer is that our swimmer is not 21 D. J. Earl, C. M. Pooley, J. F. Ryder, I. Bredberg and J. M. Yeomans,
J. Chem. Phys., 2007, 6, 126.
intrinsic, but actuated by external forces.4,22 In particular, the 22 M. C. Cosentino Lagomarsino, F. Capuani and C. P. Lowe, J Theor
total actuating force is not null instantaneously, but only over Biol, 2003, 224(2), 215–24.
a cycle. This has been shown to lead to qualitatively distinct 23 E. Lauga, Phys Rev E Stat Nonlin Soft Matter Phys, 2007, 75(4 Pt 1),
behavior in other systems.23 In our case, using simulations of an 041916.
24 C. Bustamante, J. Liphardt and F. Ritort, Physics Today, 2005, 58, 43.
analogous intrinsic swimmer actuated by two-body two-state 25 S. Chattopadhyay, R. Moldovan, C. Yeung and X. L. Wu, Proc. Natl.
springs, we find that the differences are mostly quantitative, the Acad. Sci., 2006, 103, 13712.

476 | Soft Matter, 2009, 5, 472–476 This journal is ª The Royal Society of Chemistry 2009

S-ar putea să vă placă și