Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Article

www.acsnano.org

Fully Steerable Symmetric Thermoplasmonic


Microswimmers
Martin Fränzl, Santiago Muiños-Landin, Viktor Holubec, and Frank Cichos*
Cite This: ACS Nano 2021, 15, 3434−3440 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIV OF BRITISH COLUMBIA on August 28, 2023 at 06:37:59 (UTC).

ABSTRACT: A cornerstone of the directed motion of micro-


scopic self-propelling particles is an asymmetric particle
structure defining a polarity axis along which these tiny
machines move. This structural asymmetry ties the orienta-
tional Brownian motion to the microswimmers directional
motion, limiting their persistence and making the long time
motion effectively diffusive. Here, we demonstrate a completely
symmetric thermoplasmonic microswimmer, which is propelled
by laser-induced self-thermophoresis. The propulsion direction
is imprinted externally to the particle by the heating laser
position. The orientational Brownian motion, thus, becomes
irrelevant for the propulsion, allowing enhanced control over
the particles dynamics with almost arbitrary steering capability.
We characterize the particle motion in experiments and simulations and also theoretically. The analysis reveals additional
noise appearing in these systems, which is conjectured to be relevant for biological systems. Our experimental results show
that even very small particles can be precisely controlled, enabling more advanced applications of these micromachines.
KEYWORDS: self-thermophoresis, microswimmers, active particles, thermoplasmonics, Janus particles, feedback control

example, when using thermophoresis5,21−23 or other light-

A rtificial self-propelled micro- and nanoparticles (NPs)


are synthetic active matter objects that are capable of
using energy to move persistently in liquid environ-
ments. They are often recognized as simple building blocks
that allow to explore emergent collective states of non-
controlled mechanisms such as temperature-induced diffusio-
phoresis24 or photocatalytic reactions.25 For example, a self-
thermophoretic microswimmer contains a heat source that
generates an asymmetric temperature distribution on the
equilibrium biological systems and the corresponding non- microswimmer surface. In many cases, they are realized as
equilibrium structures.1 With a variety of different propulsion metal-capped Janus-particles, where one-half of a polymer
mechanisms,2−8 a wealth of different self-propelled particles is sphere is covered with a thin metal film.5 The metal cap is then
available that allows to construct complex self-assembled heated, for example, by optical absorption,26 and generates a
autonomous micromachines, giving valuable physical insights9 temperature gradient along the polymer surface which in turn
and also offering environmental10,11 and medical12 applica- propels the particle by thermo-osmotic boundary flows.27−29
tions. All of these systems share a structural asymmetry With such switchable propulsion, the technique of photon
allowing them to break the temporal symmetry of hydro- nudging uses a feedback loop30,31 to control self-thermopho-
dynamics at low Reynolds numbers.6−8,13−16 This asymmetry retic Janus particles by harnessing the rotational diffusion as an
binds the propulsion direction to the particles symmetry axis, efficient reorientation process for micron-sized particles.22,32
and hence, its random reorientation due to the rotational Other approaches use large particles with slow rotational
Brownian motion randomizes the microswimmers motion to diffusion and active orientation steering to circumvent the
orientational randomization.33 Here, we introduce a symmetric
an effectively diffusive one on long time scales. To regain
control over the active motion, steering of the propulsion
direction has been achieved by applying external electric or Received: December 18, 2020
magnetic fields.7,17−19 Yet, precise and independent control of Accepted: February 3, 2021
multiple microswimmers remains a challenging task for these Published: February 8, 2021
external forcings.20
Control of active particles on the individual level can be
obtained when the propulsion mechanism is switchable, for

© 2021 American Chemical Society https://dx.doi.org/10.1021/acsnano.0c10598


3434 ACS Nano 2021, 15, 3434−3440
ACS Nano www.acsnano.org Article

Figure 1. Actuation of symmetric self-thermophoretic microswimmers. (a) Sketch of a symmetric microswimmer asymmetrically heated with
a focused laser. The microswimmer is composed of a MF microparticle (R = 1.1 μm). 10 % of its surface is uniformly covered with gold NPs
of about 8 nm diameter. When asymmetrically heated with a focused laser, an inhomogeneous surface temperature is generated, resulting in
a self-thermophoretic motion away from the laser focus. (b) A dark-field microscopy image of a symmetric swimmer detected in our
experimental setup. To control the propulsion direction of the swimmer, the laser focus is placed with an offset δ from the particle center at
rh, opposite to the desired target location rt. (c) A bright-field microscopy image of the liquid crystal phase transition around an immobilized
symmetric swimmer centrally heated with a laser power of P0 = 0.2 mW. (d) Maximum temperature increment ΔTmax as a function of the
incident laser power estimated from the radius of the phase transition. (e) Temperature distribution on the particle surface for δ = −0.8 μm
estimated from the projection of the intensity distribution of a Gaussian beam (beam waist w0 = 0.81 μm) on the surface of a sphere with
radius R. (f) Trajectories for driving a R = 1.1 μm symmetric swimmer between two target positions for P = 0.04 mW and δ = −0.5 μm
(Supplementary Video 1). The acceptance radius of the targets is depicted in blue and was set to 0.5 μm.

self-thermophoretic microswimmer for which an asymmetric the sample preparation are available in the Supporting
temperature distribution is induced by the heating laser itself. Information.
The actuation scheme does not depend on the time scale of Figure 1b depicts the actuation principle of the symmetric
the rotational diffusion providing a substantially better control microswimmer. To direct the particle to a target position rt, the
than techniques used for Janus-type particles. The symmetric directional vector rpt = rt − rp between the particle center rp
thermoplasmonic microswimmers are already of great interest and the target position rt is calculated, and the focus of the
in the field of active matter. They have been recently employed heating laser is placed at rh = rp + δrpt/|rpt|, where δ is the radial
to study structures emerging from information exchange34 and displacement from the particle center and typically a negative
machine learning techniques at the microscopic scale35−37 with number. Due to the self-propulsion of the particle as well as its
the ultimate goal to reach autonomous exploratory behavior. diffusive motion, the position of the heating laser needs to be
Here, we investigate their self-thermophoretic propulsion and constantly adapted in a feedback loop with the detection of the
confirm a delay-induced directional uncertainty which may particle center to control its motion direction or position.
actively contribute to the steering behavior of biological Propulsion Velocity. The self-thermophoretic propulsion
species. velocity v of the swimmer is determined by the temperature
distribution along the swimmers surface:27,30
RESULTS AND DISCUSSION χ
v=− ∫
T0S S
∇ T dS
(1)
Our symmetric microswimmer is composed of a melamine
formaldehyde (MF) microparticle (2.2 μm diameter) that is where χ is the thermophoretic mobility coefficient,28 T0 the
uniformly covered with gold NPs of about 10 nm diameter surrounding temperature, S the surface area of the particle, and
with a surface coverage of about 10% (Figure 1a and ∇∥ = (1−êrêr)·∇ the surface gradient operator, where êr is the
Supporting Information Figure S1). The gold NPs are surface normal vector.
thermally isolated from each other and act as point-like heat Experimental insight into the surface temperature rise of the
sources when heated with a laser beam with a wavelength close swimmer can be obtained with the help of the nematic/
to their plasmon resonance. When illuminated asymmetrically isotropic phase transition of a liquid crystal (see Supporting
with a highly focused laser beam, the gold NPs generate an Information for details). The heating of the gold NP at the
inhomogeneous surface temperature distribution, resulting in a colloid surface yields a phase transition in the liquid crystal
self-thermophoretic propulsion away from the laser focus surrounding the immobilized swimmer (Figure 1c). Measuring
(Figure 1a).22 All experiments are carried out in an inverted the size of the isotropic liquid crystal bubble as a function of
microscope using an acousto-optic deflector (AOD) controlled absorbed laser power and ambient temperature allows to
steerable focused laser beam at 532 nm wavelength. A sample extract the temperature at the swimmers surface. Figure 1d
consists of two glass coverslips confining a 3 μm thin liquid depicts the maximum temperature increment ΔTmax as a
film of the diluted particle suspension. The particles are function of the laser power for a 2.2 μm diameter particle. The
observed using dark-field illumination with an oil-immersion temperature values have been corrected to reflect the
dark-field condenser (NA 1.2) and a 100× oil-immersion corresponding values in water. The temperature increases
objective set to NA 0.6. Details of the experimental setup and linearly with the incident heating power and for the maximum
3435 https://dx.doi.org/10.1021/acsnano.0c10598
ACS Nano 2021, 15, 3434−3440
ACS Nano www.acsnano.org Article

Figure 2. Propulsion velocity. (a) The propulsion velocity as a function of the laser displacement δ for a laser power of 0.1 mW and an
inverse frame rate τ = 30 ms (blue squares). The blue curve represents a fit to the numerical result (see Supporting Information for details),
where the dashed blue line depicts the optimal displacement δmax = −0.5 μm. The gray curve represents the numerical result for τ = 0 with an
optimal displacement of δmax = −0.84 μm depicted as dashed gray line. The radius of the particle (R = 1.1 μm) is depicted as solid blue line.
(b) The detected propulsion velocity ⟨v⟩ as a function of the incident laser power P0 and the inverse frame rate τ (symbols) for δ = −0.5 μm.
The solid curves represent the numerical results according to eq 4. The dashed lines depict the maximum velocity ⟨v⟩max = ⟨v⟩(P0 = Pmax).
(c) Dependence of the maximum velocity ⟨v⟩max as a function of the inverse frame rate τ.

power used in our experiments (Pmax = 0.4 mW) yields ΔTmax The dependence of the swimmer velocity on the laser
= 12 K. displacement δ and the laser power P0 can be obtained from
The temperature distribution on the surface of the particle measurements driving the particle between two target positions
can be calculated assuming the illumination by a Gaussian (Figure 1f). The velocity is extracted from the time required to
beam with the intensity distribution: travel the distance between the two targets. For each set of

i w y
I(x , y , z) = I0jjjj 0 zzzz e−2((x − δ) + y )/ w(z)
2
parameters, five trajectories (from target 1 to target 2 and

k w(z) {
2 2 2 back) have been recorded and averaged. The resulting velocity
(2)
as a function of the laser displacement, plotted in Figure 2a,
shows a clear maximum. For the power of P0 = 0.1 mW, the
where the center of the Gaussian beam is displaced by δ with maximum velocity of vmax = 9 μm s−1 is observed for a
respect to the particle center (Figure 1a), displacement δmax = −0.5 μm (blue squares in Figure 2a). For
this laser power and the corresponding temperature increment
w(z) = w0 1 + (z /z R )2 is the radius of the beam at z, w0
ΔTmax = 2 K (Figure 1b), the numerical evaluation of eq 1
is the beam waist, zR = πw20/λ is the Rayleigh range, and λ is the yields a maximum velocity vmax = 8 μm s−1 in agreement with
wavelength. In the presented experiments, we used a laser the experimental result. We have used a thermometric mobility
wavelength of λ = 532 nm and a beam waist of w0 = 0.81 μm. coefficient χ = 10 × 10−10 m2 s−1 previously reported in ref 28
If we assume that the temperature increment on the for surfaces covered with the block copolymer Pluronic F127.
swimmer surface ΔT = T − T0 is proportional to the incident In contrast to the experimental result, the evaluation of eq 1
intensity ΔT ∝ I(x, y, z), the surface temperature distribution yields an optimal displacement of δmax = −0.84 μm (see
is obtained from projecting eq 2 on the surface of a sphere with Supporting Information for details).
radius R (see Supporting Information for details). The The deviation of δmax is a consequence of the experimental
resulting surface temperature distribution ΔT(θ, φ) is shown feedback delay. The position of the laser focus can only be
in Figure 1e for R = 1.1 μm and δ = 0.8 μm. adapted with a certain delay due to the finite frame rate τ−1 of
The projection of the laser intensity yields only the relative the image acquisition. During the time τ, the position of the
surface temperature increment ΔT/Tmax. To find a numerical laser is fixed, whereas the particle is self-propelled away from
estimate for the absolute temperature increment ΔTmax, we use the laser focus. Since the propulsion velocity depends on the
a superposition of the temperature field of N point-like heat distance to the laser focus (Figure 2a), the detected velocity
sources, which are randomly distributed over the surface of a will always be smaller than the instantaneous velocity. The
sphere with radius R: detected, that is, the averaged velocity reads:
I(r )σabs τ
ΔT (r ) = ∑ 1
N
4πκr (3)
⟨v⟩ =
τ
∫0 v(δ(t ))dt
(4)
where I(r) is the intensity of the Gaussian beam (eq 2), σabs is This equation can be integrated numerically for different initial
the absorption cross-section of the gold NPs and κ = (κH2O + displacements (see Supporting Information for details),
κMF)/2 is the averaged thermal conductivity around the gold leading to the instantaneous velocity profile depicted in Figure
NPs. For a NP radius of RNP = 4 nm and a surface coverage of 1a by the gray line. With increasing τ, the optimal initial
10%, the approximate number of gold NPs on the surface is N displacement δmax is indeed decreasing. When the laser focus is
= 0.1 × 4πR2/(πR2NP) ≈ 30,000. For a laser power of P0 = 0.4 placed closer to the particle center, the particle will pass
mW, the numerical evaluation of eq 3 yields a temperature through the velocity maximum during the time τ instead of
increment of ΔTmax = 12 K (see Supporting Information for moving away from it, leading to a higher averaged velocity. For
details) which perfectly matches the experimental result. τ = 30 ms, the numerical evaluation of eq 4 yields an optimal
3436 https://dx.doi.org/10.1021/acsnano.0c10598
ACS Nano 2021, 15, 3434−3440
ACS Nano www.acsnano.org Article

displacement of δmax = −0.69 μm. We attribute the even indicated by the dashed dotted curve in Figure 3. For a large
smaller experimental value (δmax = −0.5 μm, Figure 2a) to particle velocity, the positioning error is mainly defined by an
additional delays due to the time required for the image overshooting of the particle over the target position. This is
processing and laser adjustment not considered in eq 4. again due to the finite sampling of the particle position with
To find the dependence of the particle velocity on the the inverse frame rate τ. The overshooting distance is twice the
incident laser power, the experiment depicted in Figure 1f is distance traveled within the inverse frame rate ρ2 = 2vτ and
repeated for laser powers varying from 0.04 to 0.4 mW with a increases with increasing power, as shown by the dotted curve
fixed displacement of δmax = −0.5 μm. We find a nonlinear in Figure 3a. The sum of both contributions ρ = ρ12 + ρ22 is
scaling of the velocity with the laser power (Figure 2b). For an
inverse frame rate τ = 30 ms, a maximum velocity vmax = 14 μm depicted as solid curve in Figure 3a with no free parameters.
s−1 is found. The nonlinear power dependence of the velocity We have used D = 0.20 μm2 s−1 and τ = 30 ms. The diffusion
is again the result of the finite exposure time. Within the time coefficient was calculated from D = kBT0/(6πηR) with T0 = 25
τ, the particle moves and the instantaneous velocity changes, °C, η = 0.001 Pa s and R = 1.1 μm. The minimum ρmin = 240
while the laser is spatially fixed. If we assume that vmax ∝ P0, the nm in the positioning error is found for vmin = 3.0 μm s−1. The
detected velocity as a function of the power can be obtained positioning error decreases with the inverse frame rate τ. The
from the evaluation of eq 4 for different vmax (see Supporting minimal positioning error is, nevertheless, limited by the
Brownian motion and, thus, the diffusion coefficient D of the
Information for details). The result is in agreement with our
microswimmer. As compared to optical tweezers that trap
experimental findings (Figure 2b, solid lines) and predicts that
similar-sized dielectric particles with powers of several 10 mW
the velocity at high laser powers is proportional to τ−1. This
or more, our particle control is achieved with a few 100 μW
dependence is also confirmed by our experimental results
only. A major difference is thereby also the fact that the self-
depicted in Figure 1c.
propulsion works without an external force, that is, due to a
Positioning Accuracy. Based on this understanding of the
force balance in the sample, while optical tweezers exert
propulsion mechanism and its dependency on the power,
external forces on the particle. Therefore, the hydrodynamics
delay, and displacement, we are now able to explore in detail
involved in the active particle motion is short ranged and
the accuracy to control the particle. The control precision is
different from the long-ranged flow fields generated by external
best deduced from an experiment localizing the particle at a
body forces.
fixed (target) position in the sample by continuous actuation
Noise Infuence. Besides influencing the localization
(Figure 3, Supplementary Video 2).
accuracy, Brownian motion also impacts the directed motion
of the swimmer. Analyzing the trajectory in Figure 1f for τ = 30
ms, we extract the displacements Δr between subsequent
frames and project them on the desired direction to the target
(e∥) and a direction perpendicular to it (e⊥). Figure 4a shows
the experimental result for a laser power of P0 = 0.3 mW,
shifting the starting point of all displacements to the origin.
The end points of the displacements of the motion in the
horizontal direction are distributed along the horizontal and
vertical direction with an opening angle of about 53° (Figure
4a). Measuring the occurrences of horizontal and vertical
displacements yields the histograms in Figure 4b. Horizontal
displacements have a nonzero mean value, which measures the
propulsion velocity v∥ along the horizontal direction. The
Figure 3. Positioning error. (a) The symbols show the positioning displacements along the vertical direction have a zero mean,
error as a function of the laser power derived from an experiment indicating no effective propulsion along this direction. Yet, the
confining a single particle for a certain time at a target position observed variance of the horizontal and vertical distribution
(Supplementary Video 2). The dashed dotted and dotted curves differ considerably for the specific delay of τ = 30 ms (Figure
are the contributions from the two-dimensional sedimentation 4a). The variance in the horizontal direction is observed to
(dashed dotted) and the particle overshooting (dotted). The sum change only weakly with the propulsion velocity. However, the
of both contributions is represented by the solid curve. (b) An vertical displacement variance appears to grow superlinearly
example of a position distribution around a target for a laser power (Figure 4c).
of P0 = 0.01 mW. (c) Marginal probability distribution of the x- To understand the observed effects, we model the particle
position corresponding to (b).
motion with the help of a Langevin equation. In the model, the
directional uncertainty is again a result of the delayed response
The localization is then characterized by the positioning of the feedback loop. During the delay time τ, the particle
error ρ = ⟨(r − rt)2 ⟩ , where r and rt are the coordinates of diffuses away from the detected spot in the camera image,
the particle and the target, respectively. The positioning error giving rise to a deviating particle position and, thus, a
has two regimes. If the particle displacement due to the self- propulsion direction different than the intended one. This
propulsion during the inverse frame rate τ is smaller than the results in an angular distribution of the propulsion velocities
displacement due to the diffusion, the positioning error follows around the target direction p(θ) that can be approximated by
from a simple sedimentation model. The particle is radially 2
/(2σθ2)
p(θ ) ≈ (2πσθ2)−1/2 e−θ (5)
driven toward the target with a velocity v against its diffusive
motion with the diffusion coefficient D. In the steady state, the where σ2θ = 2Dτ/R2 (see Supporting Information for details).
corresponding characteristic length scale is ρ1 = 6 D/v , as The important consequence is that the particle motion not
3437 https://dx.doi.org/10.1021/acsnano.0c10598
ACS Nano 2021, 15, 3434−3440
ACS Nano www.acsnano.org Article

Figure 4. Noise influence. (a) The displacement vectors projected tangential and perpendicular to the target direction for a laser power of P0
= 0.3 mW and an inverse frame rate of τ = 30 ms. (b) The probability distribution of the projected displacement vectors in (a). The mean of
the tangential step size distribution (blue dashed line) corresponds to the velocity v∥ = ⟨Δr · e∥⟩/τ. (c) The experimental variances of the
tangential (red squares) and perpendicular (blue circles) step size distributions as functions of the propulsion velocity compared to the
theoretical model (solid and dashed lines).

Figure 5. Control of multiple swimmers. Experimental feedback control of nine individual symmetric swimmers that are arranged on a grid
and reconfigured to a circle (Supplementary Video 3).

only possesses an active component along the target direction Control of Multiple Swimmers. The actuation scheme,
but also perpendicular to it. The spread of positions so far discussed only for a single microswimmer, can be
perpendicular to the motion toward the target grows with an extended to control a larger number of swimmers by quickly
active component and is not purely diffusive. multiplexing the laser focus between different particle positions
Using the Langevin model detailed in the Supporting within the time of the inverse frame rate τ. Figure 5 and
Information, we find for the variances in tangential (σ∥) and Supplementary Video 3 demonstrate the control of nine
perpendicular direction (σ⊥): individual symmetric swimmers that are arranged on a grid
2
pattern and reconfigured to a circle. Initially, the particles are
σ 2 = (1 − σy2θ)vθ2τ 2 − e−σθ vθ2τ 2 + 2Dτ (6) randomly distributed in the field of view. When the feedback
control is enabled, each particle is driven toward its nearest
σ⊥2 = σy2θvθ2τ 2 + 2Dτ target and eventually confined there. As the resulting
(7) arrangement is constantly actuated, the structure is dynamic,
2 2 as can be seen from Supplementary Video 3. A fully
where σ2yθ = (1 − e−2σθ)/2 and vθ = eσθ/2v∥. As can be seen from simultaneous feedback control of multiple particles might be
eq 7, the perpendicular component of the displacement σ⊥ achieved using a spatial light modular or a digital micromirror
now has a term proportional to the velocity (first term) and device.
the expected diffusive term. Thus, one expects that for small
delays and small velocities, the uncertainty σ⊥ is governed by
the diffusive motion (second term), while with larger velocities
CONCLUSIONS
and delays, the first term dominates. We have presented a fully steerable thermoplasmonic
Figure 4c shows that the experimental results for σ2∥ (blue symmetric microswimmer that is self-propelled by laser-
circles) and σ2⊥ (red squares) as functions of the velocity v∥ are induced self-thermophoresis. The asymmetry, inherently
in good agreement with eqs 6 and 7 for the experimental required for a self-propulsion at low Reynolds numbers, is
opening angle 53° (blue and red lines). induced by the heating laser itself. In this way, the active
This active directional uncertainty due to the delayed particle motion becomes independent of the rotational
response might have important consequences also for the diffusion, providing a substantially higher level of control
motion of microscopic living systems such as bacteria. It than previously possible for microswimmers with an
suggests that a large propulsion speed is actually not beneficial asymmetric geometry, for example, Janus-type particles. We
if the object responds with a delay to a specific environmental find a dependence of the active particle velocity on the
signal.38 The delay, active propulsion, and Brownian noise will feedback cycle time as well as on the displacement of the
induce a directional uncertainty that grows with increasing control laser, which can be completely understood with our
speed and lets the propelled object fail to reach the target, theoretical model. This model also suggests that microscopic
while it has on average the right propulsion direction. objects subjected to positional noise with a delayed response
3438 https://dx.doi.org/10.1021/acsnano.0c10598
ACS Nano 2021, 15, 3434−3440
ACS Nano www.acsnano.org Article

develop an active propulsion uncertainty. This uncertainty ASSOCIATED CONTENT


could set, for example, a speed limit to bacteria targeting the *
sı Supporting Information
propulsion toward a specific goal. The Supporting Information is available free of charge at
The design of the symmetric microswimmer enables an https://pubs.acs.org/doi/10.1021/acsnano.0c10598.
actuation with high precision at the micrometer scale. The
Details on the symmetric thermoplasmonic particle,
precise control of active particles will thereby allow to create,
temperature distribution on particle surface, numerical
for example, active particle systems in well-controlled initial
temperature estimation, 5CB temperature measurement,
conditions, to explore their mutual interactions and collective
temperature gradient and propulsion velocity, frame rate
behaviors. Our approach can prepare dense ensembles with a
dependence of the optimal displacement, power depend-
well-defined number of members to lead self-assembly of
ence of the particle velocity, noise influence model,
nonequilibrium structures due to phoretic and hydrodynamic
experimental setup; supporting Figures S1−S19 (PDF)
interactions.39−42 The feedback control allows the incorpo-
Video 1: A symmetric swimmer driven between two
ration of nonreciprocal and delayed interactions with other
target positions (MP4)
active and passive real world objects, where collective
Video 2: A symmetric swimmer confined at a target
phenomena can emerge in an ensemble. We anticipate that
position (MP4)
the presented results will be of relevance for the development
Video 3: Control of nine individual symmetric swimmers
of more complex self-propelled particle systems and nano-
that are arranged on a grid and reconfigured to a circle
machines as well as to study synthetic active particles systems.
(MP4)
MATERIALS AND METHODS AUTHOR INFORMATION
Sample Preparation. The sample consists of two glass coverslips Corresponding Author
(22 mm × 22 mm) confining a thin liquid film. First, the coverslips Frank Cichos − Peter Debye Institute for Soft Matter Physics,
were thoroughly cleaned by rinsing successively with acetone,
isopropyl, and Milli-Q water and dried with a nitrogen gun. To
Molecular Nanophotonics Group, Universität Leipzig, 04103
prevent sticking of the microparticles, the glass surfaces were Leipzig, Germany; orcid.org/0000-0002-9803-4975;
passivated with Pluronic F-127 (Sigma-Aldrich). Therefore, the Email: cichos@physik.uni-leipzig.de
coverslips are immersed in 1% Pluronic F-127 solution for 10 min.
Thereafter, the coverslips were briefly dipped in Milli-Q water and Authors
dried with a nitrogen gun. Subsequently, the edges of one coverslip Martin Fränzl − Peter Debye Institute for Soft Matter Physics,
were covered with a thin layer of polydimethylsiloxane (PDMS) for Molecular Nanophotonics Group, Universität Leipzig, 04103
sealing. The particle solution used for the experiments was prepared Leipzig, Germany; orcid.org/0000-0001-6754-8554
by dispersing 2.2 μm diameter gold-coated melamine formaldehyde Santiago Muiños-Landin − Peter Debye Institute for Soft
(MF) particles (microParticles GmbH) in 0.1% Pluronic F-127 Matter Physics, Molecular Nanophotonics Group, Universität
solution. The surface of the MF particles is uniformly coated with Leipzig, 04103 Leipzig, Germany; Smart Systems and Smart
gold nanoparticles of about 8 nm diameter with a surface coverage of Manufacturing, Artificial Intelligence and Data Analytics
about 10% (Figure S1). To passivate the surface of the particles, they Laboratory, Polígono Industrial de Cataboi, AIMEN
have been washed in a 1% Pluronic F-127 solution before mixing.
Finally, 0.5 μL of the mixed particle suspension is pipetted in the
Technology Centre, 36418 Pontevedra, Spain
middle of one of the coverslips, and the other is placed on top. Viktor Holubec − Theory of Condensed Matter, Institute for
Depending on the area covered by the liquid, typically about (18 mm Theoretical Physics, Universität Leipzig, 04103 Leipzig,
× 18 mm), the resulting liquid film height is about 3 μm. Germany; Department of Macromolecular Physics, Faculty of
Experimental Setup. The experimental setup (Figure S19) Mathematics and Physics, Charles University, 18000 Prague,
consists of an inverted microscope (Olympus, IX71) with a mounted Czech Republic
piezo translation stage (Physik Instrumente, P-733.3). The micro-
particles are heated by a focused, continuous wave laser at a Complete contact information is available at:
wavelength of 532 nm (CNI, MGL-III-532). The beam diameter is https://pubs.acs.org/10.1021/acsnano.0c10598
increased by a beam expander and sent to an acousto-optic deflector
(AA Opto-Electronic, DTSXY-400-532) and a lens system to steer the Author Contributions
laser focus in the sample plane. The deflected beam is focused by an M.F., S.M.-L., and F.C. designed the experiments. M.F.
oil-immersion objective (Olympus, UPlanApo ×100/1.35, Oil, Iris, performed the experiments. M.F., S.M.-L., and F.C. analyzed
NA 0.5−1.35) to the sample plane (w0 ≈ 0.8 μm beam waist in the the experimental data. M.F. and F.C. implemented and
sample plane). The sample is illuminated with an oil-immersion dark- evaluated the numerical calculations. V. H. contributed the
field condenser (Olympus, U-DCW, NA 1.2−1.4) and a white-light noise influence model. M.F. and F.C. wrote the manuscript. All
LED (Thorlabs, SOLIS-3C). The scattered light is imaged by the authors discussed the results and commented on the
objective and a tube lens (250 mm) to an electron-multiplying charge-
coupled device (EMCCD) camera (Andor, iXon DV885LC). The manuscript.
variable numerical aperture of the objective was set to a value below Notes
the minimum aperture of the dark-field condenser. The dichroic beam The authors declare no competing financial interest.
splitter (D) was selected to reflect the laser wavelength (Omega
Optical, 560DRLP) and a notch filter (F) is used to block any ACKNOWLEDGMENTS
remaining back reflections from the laser (Thorlabs, NF533-17). The
The authors acknowledge financial support by the German
acousto-optic deflector (AOD) as well as the piezo stage, are driven
by an analog-digital/digital-analog (AD/DA) converter (Jäger Research Foundation (Deutsche Forschungsgemeinschaft,
Messtechnik, ADwin-Gold II). A LabVIEW program running on a DFG) through project no. 254960539 within the Priority
desktop PC (Intel Core i7 2600 4 × 3.40 GHz CPU) is used to record Program 1726 “Microswimmers” and project no. 432421051
and process the images as well as to control the AOD feedback via the (DFG-GACR cooperation). V.H. is supported by the Czech
AD/DA converter. Science Foundation (project no. 20-02955J).
3439 https://dx.doi.org/10.1021/acsnano.0c10598
ACS Nano 2021, 15, 3434−3440
ACS Nano www.acsnano.org Article

REFERENCES (25) Palacci, J.; Sacanna, S.; Steinberg, A. P.; Pine, D. J.; Chaikin, P.
M. Living Crystals of Light-Activated Colloidal Surfers. Science 2013,
(1) Bechinger, C.; Di Leonardo, R.; Löwen, H.; Reichhardt, C.; 339, 936−940.
Volpe, G.; Volpe, G. Active Particles in Complex and Crowded (26) Baffou, G.; Cichos, F.; Quidant, R. Applications and Challenges
Environments. Rev. Mod. Phys. 2016, 88, 045006. of Thermoplasmonics. Nat. Mater. 2020, 19, 946−958.
(2) Paxton, W. F.; Kistler, K. C.; Olmeda, C. C.; Sen, A.; St. Angelo, (27) Bickel, T.; Majee, A.; Würger, A. Flow Pattern in the Vicinity of
S. K.; Cao, Y.; Mallouk, T. E.; Lammert, P. E.; Crespi, V. H. Catalytic Self-Propelling Hot Janus Particles. Phys. Rev. E 2013, 88, 012301.
Nanomotors: Autonomous Movement of Striped Nanorods. J. Am. (28) Bregulla, A. P.; Würger, A.; Günther, K.; Mertig, M.; Cichos, F.
Chem. Soc. 2004, 126, 13424−13431. Thermo-Osmotic Flow in Thin Films. Phys. Rev. Lett. 2016, 116,
(3) Dreyfus, R.; Baudry, J.; Roper, M. L.; Fermigier, M.; Stone, H. 188303.
A.; Bibette, J. Microscopic Artificial Swimmers. Nature 2005, 437, (29) Bregulla, A. P.; Cichos, F. Flow Fields around Pinned Self-
862−865. Thermophoretic Microswimmers under Confinement. J. Chem. Phys.
(4) Howse, J. R.; Jones, R. A. L.; Ryan, A. J.; Gough, T.; Vafabakhsh, 2019, 151, 044706.
R.; Golestanian, R. Self-Motile Colloidal Particles: From Directed (30) Selmke, M.; Khadka, U.; Bregulla, A. P.; Cichos, F.; Yang, H.
Propulsion to Random Walk. Phys. Rev. Lett. 2007, 99, 048102. Theory for Controlling Individual Self-Propelled Micro-Swimmers by
(5) Jiang, H.-R.; Yoshinaga, N.; Sano, M. Active Motion of a Janus Photon Nudging I: Directed Transport. Phys. Chem. Chem. Phys.
Particle by Self-Thermophoresis in a Defocused Laser Beam. Phys. 2018, 20, 10502−10520.
Rev. Lett. 2010, 105, 268302. (31) Selmke, M.; Khadka, U.; Bregulla, A. P.; Cichos, F.; Yang, H.
(6) Ebbens, S. J.; Howse, J. R. Pursuit of Propulsion at the Theory for Controlling Individual Self-Propelled Micro-Swimmers by
Nanoscale. Soft Matter 2010, 6, 726. Photon Nudging II: Confinement. Phys. Chem. Chem. Phys. 2018, 20,
(7) Shields, C. W.; Velev, O. D. The Evolution of Active Particles: 10521−10532.
Toward Externally Powered Self-Propelling and Self-Reconfiguring (32) Qian, B.; Montiel, D.; Bregulla, A.; Cichos, F.; Yang, H.
Particle Systems. Chem 2017, 3, 539−559. Harnessing Thermal Fluctuations for Purposeful Activities: The
(8) Tsang, A. C. H.; Demir, E.; Ding, Y.; Pak, O. S. Roads to Smart Manipulation of Single Micro-Swimmers by Adaptive Photon
Artificial Microswimmers. Adv. Intell. Syst. 2020, 2, 1900137. Nudging. Chem. Sci. 2013, 4, 1420.
(9) Aubret, A.; Youssef, M.; Sacanna, S.; Palacci, J. Targeted (33) Bäuerle, T.; Löffler, R. C.; Bechinger, C. Formation of Stable
Assembly and Synchronization of Self-Spinning Microgears. Nat. Phys. and Responsive Collective States in Suspensions of Active Colloids.
2018, 14, 1114−1118. Nat. Commun. 2020, 11, 2547.
(10) Zarei, M.; Zarei, M. Self-Propelled Micro/Nanomotors for (34) Khadka, U.; Holubec, V.; Yang, H.; Cichos, F. Active Particles
Bound by Information Flows. Nat. Commun. 2018, 9, 3864.
Sensing and Environmental Remediation. Small 2018, 14, 1800912.
(35) Muiños-Landin, S.; Ghazi-Zahedi, K.; Cichos, F. Reinforcement
(11) Wang, L.; Kaeppler, A.; Fischer, D.; Simmchen, J. Photo-
Learning of Artificial Microswimmers. arXiv (Sof t Condensed Matter),
catalytic TiO2 Micromotors for Removal of Microplastics and
April 6, 2018, 1803.06425, ver. 2. https://arxiv.org/abs/1803.06425
Suspended Matter. ACS Appl. Mater. Interfaces 2019, 11, 32937− (accessed 2020-12-10)
32944. (36) Cichos, F.; Gustavsson, K.; Mehlig, B.; Volpe, G. Machine
(12) Sonntag, L.; Simmchen, J.; Magdanz, V. Nano-and Micro- Learning for Active Matter. Nat. Mach. Intell. 2020, 2, 94−103.
motors Designed for Cancer Therapy. Molecules 2019, 24, 3410. (37) Fränzl, M.; Cichos, F. Active Particle Feedback Control with a
(13) Purcell, E. M. Life at Low Reynolds Number. Am. J. Phys. 1977, Single-Shot Detection Convolutional Neural Network. Sci. Rep. 2020,
45, 3−11. 10, 12571.
(14) Golestanian, R.; Liverpool, T. B.; Ajdari, A. Designing Phoretic (38) Romanczuk, P.; Salbreux, G. Optimal Chemotaxis in
Micro- and Nano-Swimmers. New J. Phys. 2007, 9, 126. Intermittent Migration of Animal Cells. Phys. Rev. E 2015, 91,
(15) Moran, J. L.; Posner, J. D. Phoretic Self-Propulsion. Annu. Rev. 042720.
Fluid Mech. 2017, 49, 511−540. (39) Niu, R.; Fischer, A.; Palberg, T.; Speck, T. Dynamics of Binary
(16) Guix, M.; Weiz, S. M.; Schmidt, O. G.; Medina-Sánchez, M. Active Clusters Driven by Ion-Exchange Particles. ACS Nano 2018,
Self-Propelled Micro/Nanoparticle Motors. Part. Part. Syst. Charact. 12, 10932−10938.
2018, 35, 1700382. (40) Schmidt, F.; Liebchen, B.; Löwen, H.; Volpe, G. Light-
(17) Baraban, L.; Makarov, D.; Streubel, R.; Mönch, I.; Grimm, D.; Controlled Assembly of Active Colloidal Molecules. J. Chem. Phys.
Sanchez, S.; Schmidt, O. G. Catalytic Janus Motors on Microfluidic 2019, 150, 094905.
Chip: Deterministic Motion for Targeted Cargo Delivery. ACS Nano (41) Choudhury, U.; Singh, D. P.; Qiu, T.; Fischer, P. Chemical
2012, 6, 3383−3389. Nanomotors at the Gram Scale Form a Dense Active Optorheological
(18) Yoshizumi, Y.; Honegger, T.; Berton, K.; Suzuki, H.; Peyrade, Medium. Adv. Mater. 2019, 31, 1807382.
D. Trajectory Control of Self-Propelled Micromotors Using AC (42) Heckel, S.; Grauer, J.; Semmler, M.; Gemming, T.; Löwen, H.;
Electrokinetics. Small 2015, 11, 5630−5635. Liebchen, B.; Simmchen, J. Active Assembly of Spheroidal Photo-
(19) Guo, J.; Gallegos, J. J.; Tom, A. R.; Fan, D. Electric-Field- catalytic BiVO4 Microswimmers. Langmuir 2020, 36, 12473−12480.
Guided Precision Manipulation of Catalytic Nanomotors for Cargo
Delivery and Powering Nanoelectromechanical Devices. ACS Nano
2018, 12, 1179−1187.
(20) Ilic, O.; Kaminer, I.; Lahini, Y.; Buljan, H.; Soljačić, M.
Exploiting Optical Asymmetry for Controlled Guiding of Particles
with Light. ACS Photonics 2016, 3, 197−202.
(21) Brenner, H. Self-Thermophoresis and Thermal Self-Diffusion in
Liquids and Gases. Phys. Rev. E 2010, 82, 036325.
(22) Bregulla, A. P.; Yang, H.; Cichos, F. Stochastic Localization of
Microswimmers by Photon Nudging. ACS Nano 2014, 8, 6542−6550.
(23) Kroy, K.; Chakraborty, D.; Cichos, F. Hot Microswimmers. Eur.
Phys. J.: Spec. Top. 2016, 225, 2207−2225.
(24) Gomez-Solano, J. R.; Blokhuis, A.; Bechinger, C. Dynamics of
Self-Propelled Janus Particles in Viscoelastic Fluids. Phys. Rev. Lett.
2016, 116, 138301.

3440 https://dx.doi.org/10.1021/acsnano.0c10598
ACS Nano 2021, 15, 3434−3440

You might also like