You are on page 1of 6

Article

www.acsnano.org

Highly Efficient Light-Driven TiO2−Au Janus


Micromotors
Renfeng Dong,† Qilu Zhang,† Wei Gao,‡,§ Allen Pei,‡,∥ and Biye Ren*,†

Research Institute of Materials Science, South China University of Technology, Guangzhou 510640, China

Department of Nanoengineering, University of California San Diego, La Jolla, California 92093, United States
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: A highly efficient light-driven photocatalytic TiO2−Au


Janus micromotor with wireless steering and velocity control is
described. Unlike chemically propelled micromotors which com-
Downloaded via SHENZHEN UNIV on December 26, 2020 at 16:21:30 (UTC).

monly require the addition of surfactants or toxic chemical fuels, the


fuel-free Janus micromotor (diameter ∼1.0 μm) can be powered in
pure water under an extremely low ultraviolet light intensity (2.5 ×
10−3 W/cm2), and with 40 × 10−3 W/cm2, they can reach a high
speed of 25 body length/s, which is comparable to common Pt-based
chemically induced self-electrophoretic Janus micromotors. The
photocatalytic propulsion can be switched on and off by incident
light modulation. In addition, the speed of the photocatalytic TiO2−
Au Janus micromotor can be accelerated by increasing the light
intensity or by adding low concentrations of chemical fuel H2O2 (i.e.,
0.1%). The attractive fuel-free propulsion performance, fast movement triggering response, low light energy requirement,
and precise motion control of the TiO2−Au Janus photocatalytic micromotor hold considerable promise for diverse
practical applications.
KEYWORDS: Janus micormotors, self-electrophoresis, TiO2, light driven, fuel free

T he autonomous motion of nano/microscale objects has


stimulated considerable research efforts over the past
decade.1−9 Particular attention has been given to
chemically powered micromotors which include bimetallic
catalytic nanowires,10,11 microtubular microrockets,12,13 and
propulsion of Janus micromotor can be conveniently triggered
even at extremely low ultraviolet (UV) light intensity (2.5 ×
10−3 W/cm2) with high repeatability and accelerated by
increasing the light intensity or by adding low concentrations
of chemical fuel H2O2. As illustrated in Figure 1A, the Janus
Janus microspheres.14,15 However, they mostly exhibit auton- micromotors consist of plain TiO2 particles (∼1.0 μm mean
omous self-propulsion in the presence of hydrogen peroxide
diameter) with one hemisphere coated with Au metal. On the
fuel10,11,14 or hydrazine15 which greatly hinders the practical
utility of chemically powered nanomotors. Therefore, efforts one hand, anatase TiO2 is widely studied due to its interesting
toward the fuel-free nanomotors powered by external stimuli properties, including photocatalytic activity, biocompatibility,
such as magnetic fields,16,17 electrical fields,18,19 ultrasound,20,21 high chemical stability, and low costs;28,29 it is commonly used
and light22,23 are particularly exciting and considerably for decontamination purposes due to the high photoactivity of
promising toward practical applications. In particular, light is TiO2.30,31 On the other hand, it is well-established that Au-
one of the most important and versatile physical stimuli to doped TiO2 or the TiO2−Au system has a greatly enhanced
facilitate and regulate the propulsion of micro/nanomotors.24,25 catalytic performance compared to that of bulk TiO2.32−34 In
However, the major challenge of light-driven micromotors is to light of these advantages, we designed catalytic TiO2−Au Janus
achieve the high propulsion speed and precise motion micromotors by coating a thin film of gold (gold film thickness:
control.26,27 40 nm) on TiO2 particles (Figure 1B). When the TiO2 surface
Here, we demonstrate light-driven TiO2−Au Janus micro-
is exposed to UV light, the self-electrophoretic propulsion
motors (∼1.0 μm diameter) based on light induced self-
electrophoresis mechanism which display efficient propulsion mechanism is immediately triggered and efficiently propels the
(over 25 body length/s) under an extremely low UV light micromotor in pure water.
intensity (40 × 10−3 W/cm2) in the presence of pure water
without adding any extra fuel. The light-driven Janus motors Received: September 21, 2015
reach a fast speed which is comparable to common Pt-based Accepted: November 23, 2015
chemically induced self-electrophoretic Janus micromotors. The Published: November 23, 2015

© 2015 American Chemical Society 839 DOI: 10.1021/acsnano.5b05940


ACS Nano 2016, 10, 839−844
ACS Nano Article

(Supporting Information Video S3). The micromotor stops


moving in the presence of 10.0 mM salt. These results further
confirm the self-electrophoresis mechanism. We also examined
the mixed potentials of Au and TiO2 electrodes with and
without illumination (Figure 2). It is clear from the Figure 2

Figure 1. (A) Schematic of catalytic TiO2−Au Janus micromotors Figure 2. A Tafel plot of Au, Ni, and TiO2 with and without
powered by UV light in water. (B) Scanning electron microscope illumination (0.5 × 10−3 W/cm2).
(SEM) image of a spherical TiO2−Au micromotor, and (C−E) the
corresponding Energy Dispersive X-ray (EDX) images for Ti, Au,
O, respectively. Scale bar, 0.5 μm. (F) Tracking lines (taken from that the potential difference between the Au electrode and
Supporting Information Video S1) illustrating the distances illuminated TiO2 (ΔETiO2(light)−Au = 307 mV) electrode is larger
traveled by three micromotors in pure water over 1 s. Scale bar,
10 μm. than that between Au electrode and TiO2 electrode without
illumination (ΔETiO2(dark)−Au = 237 mV). As reported previously
RESULTS AND DISCUSSION by Nakato and Tsubomura, the height of the Schottky barrier at
the Au/TiO2 contact is considerably increased by illumination,
The propulsion of the TiO2−Au micromotors dominantly and the photovoltage obtained is much larger than that
originates from the light induced self-electrophoresis.35 expected from the barrier height in the dark.37 These findings
Compared to Pt−Au Janus motors which require H2O2 for are consistent with the obtained results. The motion of such
self-electrophoresis, the light-driven TiO2−Au system can light-driven Janus micromotor is strongly dependent on the
generate a self-induced electric field in pure water and the coated layer due to the self-electrophoresis mechanism. We
motion can be controlled by light. Upon UV irradiation of the tried TiO2−nickel (Ni) Janus micromotors as control, as
Janus micromotors, charge separation occurs within the TiO2 illustrated in the Supporting Information Video S4, and the
and electrons are injected from the TiO2 conduction band into speed of TiO2−Ni Janus micromotor (8 μm/s under 40 × 10−3
the Au hemisphere. Protons are produced from the oxidation of W/cm2 UV light intensity) is much lower than that of TiO2−
water at TiO2 and the resulting electrons are consumed during Au Janus micromotors (25 μm/s). The Tafel plot of Au, Ni,
the reduction of protons at Au. The resulting flux of H+ and TiO2 with and without illumination (0.5 × 10−3 W/cm2)
generates a fluid flow toward the Au hemisphere, generating a (shown in Figure 2) clearly indicates that the potential
slip velocity and propelling the micromotors with the TiO2 difference between Ni and illuminated TiO2 electrode
hemisphere forward (Supporting Information Video S2 shows
(ΔETiO2(light)−Ni = 150 mV) is significantly smaller than that
that the TiO2−Au Janus motors are moving toward TiO2 side
under an extremely low UV intensity, 2.5 × 10−3 W/cm2). In between Au and illuminated TiO2 electrodes (ΔETiO2(light)−Au =
contrast, plain TiO2 microspheres without Au layer only display 307 mV). This finding is consistent with the propulsion
Brownian motion even upon exposure to 40 × 10−3 W/cm2 UV behavior of the corresponding micromotors and further
light (Supporting Information Video S1). Previous reports have confirms the self-electrophoresis mechanism of the light-driven
demonstrated thermophoretic micromotor propulsion from metal-TiO2 micromotors. Regarding the insulation layer,
infrared irradiation of Au.36 However, to validate which is the aluminum oxide (Al2O3)−TiO2 Janus microsphere was chosen
dominant mechanism resulting in the propulsion of the TiO2− as the control. As illustrated from Supporting Information
Au micromotors, photoinduced self-electrophoresis or thermo- Video S4, Al2O3−TiO2 Janus microsphere does not display
phoresis, control experiments were conducted using poly- directional motion under 40 × 10−3 W/cm2 UV light intensity.
styrene (PS)−Au Janus particles (1.0 μm diameter). Under 40 Such phenomenon can be attributed to the greatly hindered
× 10−3 W/cm2 intensity light, no obvious directional electrophoresis between the insulation layer and TiO2.
movement was observed for PS−Au Janus particles, while the As indicated from Figure 1A, the photoactivity of TiO2
TiO2−Au micromotors were moving pretty fast (Supporting comes from its hole−electron separation triggered by photons
Information Video S1). The control experiments of the of energy equal to or higher than its bandgap. The conduction
micromotors in salt rich environment (NaCl solution) have band of Au lies below that of TiO2, and as a result, the Au
been investigated. The speed of TiO2−Au Janus motor in 1.0 hemisphere acts as a sink for electrons, increasing electron−
mM NaCl solution is obviously smaller than that in pure water hole pair lifetime and recombination times which result in
840 DOI: 10.1021/acsnano.5b05940
ACS Nano 2016, 10, 839−844
ACS Nano Article

Figure 3. (A−D) Tracklines of micromotors (Supporting Information Video S5) with UV intensities of (A) 2.5, (B) 5, (C) 10, and (D) 40 ×
10−3 W/cm2, over 1 s. Scale bar, 10 μm. (E) The influence of the UV light intensity on the speed of Janus micromotors in pure water. (F)
Average MSD versus time interval (Δt) analyzed from tracking trajectories. (G) Diffusion coefficient of TiO2−Au Janus micromotor under
different UV light intensity, the values determined from the MSD plots (20 micromotors were analyzed).

enhanced reactivity on the micromotors surfaces.35 Therefore, photothermally propelled motors.26 Figure 3E illustrates the
such structure makes the TiO2−Au Janus micromotor much diagram of the influence of light intensity (I) on the speed of
more efficient than the reported plain TiO2 micromotor whose TiO2−Au Janus micromotor. To clearly indicate the directional
average speed is around 10 μm/s under 2.5 W/cm2 UV light movement of the Janus motors at extremely low UV light
intensity.27 Compositional analysis using SEM and EDX intensity, we experimentally estimated the enhanced diffusion
mapping indicates that the Au coating only covers half of the coefficients of the motors. The motors (number of motors n =
TiO2 spherical particle (Figure 1B−E, the relative EDX pattern 20) were tracked over 10 s, the mean squared displacement
in Supporting Information Figure S1 also illustrating the (MSD) was calculated at different UV light intensity (0, 2.5 ×
composition of the TiO2−Au Janus micromotor), generating 10−3, 5 × 10−3 W/cm2) and the diffusion coefficient (D) was
the asymmetry necessary for directional motion. The micro- calculated by the equation D = MSD/iΔt, where Δt is the time
motor tracklines in Figure 1F show that the new TiO2−Au interval. Here, for the case of two-dimensional analysis from the
Janus motors move immediately at a speed of around 25 μm/s recorded video, i is equal to 4 (Figure 3F,G, respectively). It
in pure water upon UV light irradiation. can be clearly observed that although the speed under 2.5 ×
As photocatalysis primarily drives the TiO2−Au micromotor 10−3 W/cm2 UV light is almost the same as that without UV
propulsion, the motion speed of the micromotors is able to be light, the MSD of Janus motor under 2.5 × 10−3 W/cm2 UV
precisely controlled by adjusting the UV irradiation intensity light is significantly higher than that without UV light. These
(I). The highly efficient TiO2−Au Janus micromotors can be results demonstrated that the UV light with extremely low
propelled by super low UV light intensity. Figure 3A reveals intensities still has obvious influence on the propulsion of the
pretty slow but directional motion when I is 2.5 × 10−3 W/cm2. Au−TiO2 micromotors. In general, the relationship of incident
When increasing I from 5- to 10 × 10−3 W/cm2, the speed of light (I) and the flux (Φ, the number of incident photons per
the micromotors increases from 5.6 to 11.2 μm/s (Figure unit area per second) can be described as equation 1.24
3B,C). The micromotors can move at a speed of over 25 μm/s hc
corresponding to a relative speed of 25 body lengths/s under a I=Φ
λ (1)
light intensity of 40 × 10−3 W/cm2 (Figure 3D). Such efficient
−34
speed is almost 2 times faster than reported speed of light Here, h, c, and λ represent Planck’s constant (6.626 × 10 J·
driven TiO2 micromotors in pure water;27 however, over 60 s), speed of light (3 × 108 m/s), and wavelength of the UV light
times lower UV light power is needed. Also, such efficient (330−380 nm in this case), respectively. As a result, the flux Φ
speed is comparable to that of the catalytic micro/nanomotors, of photons of wavelength λ increases with increasing I on the
and significantly larger than the speed of previously reported micromotor, and thus, the number of the photogenerated holes
841 DOI: 10.1021/acsnano.5b05940
ACS Nano 2016, 10, 839−844
ACS Nano Article

Figure 4. Cyclic “On” and “Off” UV light activation of the TiO2−Au Janus micromotor. (A) Schematic and time-lapse images illustration of
TiO2−Au Janus micromotor moving status in pure water with UV light and without UV light, respectively. (B) Corresponding speed/time
dependence illustrating the UV light triggered “On/Off” motion control of the TiO2−Au Janus micromotor in pure water without any
additional chemicals. The images were taken at 1-s intervals from Video S6. Scale Bar, 10 μm. UV light intensity, 40 × 10−3 W/cm2.

and electrons in TiO2 increases. Therefore, there will be a Video S7). Such directional control could offer more possibility
proton gradient enhancement and corresponding fluid shear for the light-driven micromotor applying in variety of practical
velocity resulting from increased irradiation intensity. Accord- applications.
ingly, the speed of the micromotors can be modulated by Interestingly, low concentrations of H2O2 can also improve
incident light intensity. the motion of the TiO2−Au micromotors. The micromotors
Compared to catalytic micromotors which consume chemical can be accelerated gradually by increasing the concentration of
fuels, the as-developed light-driven TiO2−Au micromotors H2O2 under 40 × 10−3 W/cm2 UV light. The TiO2−Au
have the impressive advantages of highly repeatable simple micromotors display Brownian motion (Figure 6A) in 0.1%
motion control. As Figure 4 illustrates, the repeated on/off
cycling of UV light illumination induces the activation and
inactivation of the movement of micromotors reflecting the fast
response rate of the micromotors upon the UV irradiation. The
speed of the micromotors is around 25 μm/s under 40 × 10−3
W/cm2 UV light exposure, but stop moving immediately when
the UV light is off. This “stop/go” propulsion behavior
indicates that high reversibility and controllable micromotor
motion can be achieved by switching the UV irradiation on or
off (as shown in Supporting Information Video S6). Even after
30 cycles of such on and off control, the micromotors still show
highly repeatability.
Directional control of micromotor is also a critical perform-
ance for diverse practical applications. The deposition of a
paramagnetic Ni layer between Au layer and TiO2 can be easily
realize magnetic control of the directionality of the light-driven
TiO2−Au Janus micromotor.38−40 Figure 5A illustrates the
structure of such magnetic guided Janus micromotors. With the
use of an external magnetic field, Au−Ni−TiO2 Janus Figure 6. Tracklines for TiO2−Au micromotors under three
micromotors can be precisely navigated following predeter- conditions: (A) with H2O2 without UV, (B) without H2O2 with
mined trajectories (Figure 5B and Supporting Information UV, and (C) with both H2O2 and UV, over 1 s. Scale bar, 10 μm
(taken from Supporting Information Video S8). (D) Schematic of
catalytic TiO2−Au Janus micromotors powered by UV light in low
concentration of H2O2.

H2O2 without UV light exposure; however, the speed of such


light-driven Janus motor can reach 42 μm/s (Figure 6C) with
additional 0.1% H2O2, which is almost 2 times compared to the
one in pure water (Figure 6B). Such fast speed of TiO2−Au
Janus motors in 0.1% H2O2 is almost 20 times faster than the
common Pt based Janus motors in 0.2% H2O2 at room
temperature,14,41 which is the lowest H2O2 concentration
Figure 5. (A) Schematic of the magnetic control of multilayer Au− requirement with high speed for Janus motors so far. The
Ni−TiO2 Janus micromotors. (B) Time-lapse images (taken from acceleration of TiO 2 −Au Janus motors in low H 2 O 2
Supporting Information Video S7) showing the magnetically concentrations with UV light is due to the enhanced self-
guided propulsion of an Au−Ni−TiO2 micromotor under 40 × electrophoretic effects in H2O2 compared to pure water (Figure
10−3 W/cm2 UV light. Scale bar, 10 μm. 6D).42,43
842 DOI: 10.1021/acsnano.5b05940
ACS Nano 2016, 10, 839−844
ACS Nano Article

CONCLUSIONS environment. Gold, nickel, and TiO2 films (all the films’ thicknesses,
100 nm) on ITO glass disc (diameter, 1.0 cm) were used as the
In conclusion, we described light-driven TiO2−Au Janus working electrode in the electrochemical potential measurements,
micromotors which can be powered efficiently by extremely respectively. We use the CH Instrument Model CHI600C to test the
low UV light energy in pure water. In addition, the light potential at a scan rate of 5 mV/s and over a potential range of −0.2 to
induced self-electrophoresis mechanism has been demonstrated 0.3 V (vs Ag/AgCl, 3 M KCl reference).
in detail also. This study showed that photocatalytic micro- Equipment. SEM and EDX pictures were obtained by Zeiss VEO
motors can be accelerated by increasing the light power 18; XRD pattern was obtained by X-ray Diffractomer (Panlytical, Inc.
intensity or addition of low concentrations of extra chemical X′ Pert Pro). UV light was generated by Mercury lamp sockets,
fuels H2O2. Furthermore, such light-driven, TiO2-based micro- dichroic mirror DM 400 and barrier filter BA420, intensity controlled
by ND filters (4×, 8×, 16×) (all from Nikon), and light intensity was
motors can be driven in a precisely controlled manner,
measured by UV radiometer UV-A (Videos were captured by an
involving controllable activation, acceleration, deceleration, inverted optical microscope (Nikon In-strument, Inc. Ti−S), coupled
stop and directional control. In addition, the gold hemisphere with 40× objectives, and a Zyla scmos digital camera (ANDOR) using
in TiO2−Au Janus micromotors can be easily modified with the NIS-Elements AR 4.3 software.
diverse functional groups offering various possibilities to
challenging complicate tasks. The fuel-free property of these ASSOCIATED CONTENT
light-driven micromotors combined with the decontaminative
capabilities of TiO2 should naturally lead to useful applications
*
S Supporting Information
The Supporting Information is available free of charge on the
in environmental remediation.31,44 These light-driven, precisely
ACS Publications website at DOI: 10.1021/acsnano.5b05940.
controllable, and highly efficient TiO2-based phtotocatalytic
Janus micromotors hold considerable promise for the design of Supporting videos description, supporting figures (PDF)
practical light-driven nanomachines toward a wide range of Supporting videos (ZIP
important future applications ranging from nanofabrication45 to
environmental remediation.3 AUTHOR INFORMATION
EXPERIMENTAL SECTION Corresponding Author
Synthesis of Janus Micromotors. TiO2 microspheres were *E-mail: mcbyren@scut.edu.cn.
prepared by the solvent extraction/evaporation method using Present Addresses
tetrabutyl titanate as a precursor.46 Briefly, 1.0 mL of tetrabutyl §
Wei Gao, Electrical Engineering and Computer Sciences,
titanate (Sigma #244112) was dissolved in 40.0 mL of ethanol and University of California, Berkeley, California 94720, United
incubatde at room temperature for 3 h; then, TiO2 microspheres were States.
collected by centrifugation at 7000 rpm for 5 min and washed ∥
Allen Pei, Department of Materials Science & Engineering,
repeatedly with ethanol (Guangzhou Chemical Reagent Co.) and
ultrapure water (18.2 MΩ·cm), three times each, then dried in air at Stanford University, Stanford, California 94305, United States.
room temperature. TiO2 (anatase) microsphere is obtained after Notes
annealing for 2 h at 400 °C. The X-ray diffraction (XRD) pattern The authors declare no competing financial interest.
(Supporting Information Figure S2) reveals that the TiO2 micrspheres
have a good anatase phase. For the TiO2−Au light-driven Janus ACKNOWLEDGMENTS
micromotor, TiO2 microspheres (1.0 μm mean diameter) are used as
the base particles. TiO2 particles (10.0 μg) were first dispersed in The financial support from the NSFC (21274047) and the
150.0 μL of ethanol. The sample was then spread onto glass slides and Specialized Research Fund for the Doctoral Program of the
dried uniformly to form particle monolayers. The particles were Education Ministry (20120172110005) is gratefully acknowl-
sputter coated with a thin gold and nickel layer using a Quorum Q edged.
150T ES Sputter Coater for 3 cycles with 60 s per cycle. The metal
layer thickness was found to be 40 nm, as measured by the Veeco REFERENCES
DEKTAK 150 Profilometer. For the TiO2−Ni−Au magnetic Janus
motors, TiO2 particle monolayers were prepared as in the method (1) Wang, J. Nanomachines: Fundamentals and Applications; John
above. A 40 nm layer of Au followed by a 10 nm layer of Ni were Wiley & Sons, 2013.
sequentially deposited on half of the particles by Quorum Q 150T ES (2) Mallouk, T. E.; Sen, A. Powering Nanorobots. Sci. Am. 2009, 300,
Sputter Coater. The TiO2 microspheres were coated with Al2O3 layer 72−77.
using ultrahigh Vacuum Magnetron Sputter Coater JPG 560. The (3) Moo, J. G.; Pumera, M. Chemical Energy Powered Nano/micro/
micromotors were subsequently released from the glass slides via pipet macromotors and the Environment. Chem. - Eur. J. 2015, 21, 58−72.
pumping and dispersed into double distilled water. The polystyrene− (4) Mei, Y.; Solovev, A. A.; Sanchez, S.; Schmidt, O. G. Rolled-Up
Au Janus microsphere as a control was fabricated with the same Nanotech on Polymers: from Basic Perception to Self-Propelled
method using polystyrene microsphere (Baseline #6-1-0100). Catalytic Microengines. Chem. Soc. Rev. 2011, 40, 2109−19.
Speed Calibration Experiments. To determine the relationship (5) Ozin, G. A.; Manners, I.; Fournier-Bidoz, S.; Arsenault, A. Dream
between the TiO2−Au, TiO2−Ni, and TiO2−Al2O3 motor speed and Nanomachines. Adv. Mater. 2005, 17, 3011−3018.
light intensity, the light intensity ranged from 2.5 × 10−3 to 40 × 10−3 (6) Purcell, E. M. Life at Low Reynolds Number. Am. J. Phys. 1977,
W/cm2. Sodium chloride (NaCl) solutions (0.2−200.0 mM) were 45, 3−11.
prepared for testing salt rich environment control experiment. A 0.2% (7) Sanchez, S.; Soler, L.; Katuri, J. Chemically Powered Micro- And
aqueous hydrogen peroxide (Alfa Aesar #33323) solution was Nanomotors. Angew. Chem., Int. Ed. 2015, 54, 1414−44.
prepared and directly mixed with the motor droplets. The propulsion (8) Wang, J. Can Man-Made Nanomachines Compete with Nature
calibration experiments were performed by mixing 1.0 μL of the motor Biomotors? ACS Nano 2009, 3, 4−9.
and hydrogen peroxide solutions each. (9) Guix, M.; Mayorga-Martinez, C. C.; Merkoçi, A. Nano/
Electrochemical Potential Measurements. Tafel plot is used to Micromotors in (Bio)chemical Science Applications. Chem. Rev.
obtain the potential established at different segment of different Janus 2014, 114, 6285−6322.
micromotors (Au, Ni, and TiO2) with and without illumination (0.5 × (10) Paxton, W. F.; Kistler, K. C.; Olmeda, C. C.; Sen, A.; St Angelo,
10−3 W/cm2 UV light intensity, λ = 330−380 nm) in a pure water S. K.; Cao, Y.; Mallouk, T. E.; Lammert, P. E.; Crespi, V. H. Catalytic

843 DOI: 10.1021/acsnano.5b05940


ACS Nano 2016, 10, 839−844
ACS Nano Article

Nanomotors: Autonomous Movement of Striped Nanorods. J. Am. (33) Green, I. X.; Tang, W. J.; Neurock, M.; Yates, J. T.
Chem. Soc. 2004, 126, 13424−31. Spectroscopic Observation of Dual Catalytic Sites during Oxidation
(11) Sattayasamitsathit, S.; Gao, W.; Calvo-Marzal, P.; Manesh, K. of CO on an Au/TiO2 Catalyst. Science 2011, 333, 736−739.
M.; Wang, J. Simplified Cost-Effective Preparation of High-Perform- (34) Murdoch, M.; Waterhouse, G. I. N.; Nadeem, M. A.; Metson, J.
ance Ag-Pt Nanowire Motors. ChemPhysChem 2010, 11, 2802−5. B.; Keane, M. A.; Howe, R. F.; Llorca, J.; Idriss, H. The Effect of Gold
(12) Gao, W.; Dong, R.; Thamphiwatana, S.; Li, J.; Gao, W.; Zhang, Loading and Particle Size on Photocatalytic Hydrogen Production
L.; Wang, J. Artificial Micromotors in the Mouse’s Stomach: A Step from Ethanol over Au/TiO2 Nanoparticles. Nat. Chem. 2011, 3, 489−
toward in Vivo Use of Synthetic Motors. ACS Nano 2015, 9, 117−123. 492.
(13) Wu, Z.; Wu, Y.; He, W.; Lin, X.; Sun, J.; He, Q. Self-Propelled (35) Ni, M.; Leung, M. K.; Leung, D. Y.; Sumathy, K. A Review and
Polymer-Based Multilayer Nanorockets for Transportation and Drug Recent Developments in Photocatalytic Water-Splitting Using TiO2
Release. Angew. Chem., Int. Ed. 2013, 52, 7000−3. for Hydrogen Production. Renewable Sustainable Energy Rev. 2007, 11,
(14) Ma, X.; Hahn, K.; Sanchez, S. Catalytic Mesoporous Janus 401−425.
Nanomotors for Active Cargo Delivery. J. Am. Chem. Soc. 2015, 137, (36) Xuan, M.; Shao, J.; Lin, X.; Dai, L.; He, Q. Light-Activated Janus
4976−4979. Self-Assembled Capsule Micromotors. Colloids Surf., A 2015, 482, 92−
(15) Gao, W.; Pei, A.; Dong, R.; Wang, J. Catalytic Iridium-Based 97.
Janus Micromotors Powered by Ultralow Levels of Chemical Fuels. J. (37) Nakato, Y.; Tsubomura, H. The Photoelectrochemical Behavior
Am. Chem. Soc. 2014, 136, 2276−2279. of an n-TiO2 Electrode Coated with a Thin Metal Film, as Revealed by
(16) Kathrin, E. P.; Soichiro, T.; Famin, Q.; Li, Z.; Bradley, J. N. Measurements of the Potential of the Metal Film. Isr. J. Chem. 1982,
Magnetic Helical Micromachines. Chem. - Eur. J. 2013, 19, 28−38. 22, 180−183.
(17) Fischer, P.; Ghosh, A. Magnetically Actuated Propulsion at Low (38) Burdick, J.; Laocharoensuk, R.; Wheat, P. M.; Posner, J. D.;
Wang, J. Synthetic Nanomotors in Microchannel Networks: Direc-
Reynolds Numbers: Towards Nanoscale Control. Nanoscale 2011, 3,
tional Microchip Motion and Controlled Manipulation of Cargo. J.
557−563.
Am. Chem. Soc. 2008, 130, 8164−8165.
(18) Loget, G.; Kuhn, A. Propulsion of Microobjects by Dynamic
(39) Baraban, L.; Makarov, D.; Streubel, R.; Mönch, I.; Grimm, D.;
Bipolar Self-Regeneration. J. Am. Chem. Soc. 2010, 132, 15918−15919.
Sanchez, S.; Schmidt, O. G. Catalytic Janus Motors on Microfluidic
(19) Chang, S. T.; Vesselin, N. P.; Dimiter, N. P.; Orlin, D. V.
Chip: Deterministic Motion for Targeted Cargo Delivery. ACS Nano
Remotely Powered Self-Propelling Particles and Micropumps Based 2012, 6, 3383−3389.
on Miniature Diodes. Nat. Mater. 2007, 6, 235−240. (40) Kline, T. R.; Paxton, W. F.; Mallouk, T. E.; Sen, A. Catalytic
(20) Kagan, D.; Michael, J. B.; Jonathan, C. C.; Erdembileg, C.-E.; Nanomotors: Remote-Controlled Autonomous Movement of Striped
Sadik, E.; Joseph, W. Acoustic Droplet Vaporization and Propulsion of Metallic Nanorods. Angew. Chem. 2005, 117, 754−756.
Perfluorocarbon-Loaded Microbullets for Targeted Tissue Penetration (41) Xuan, M.; Shao, J.; Lin, X.; Dai, L.; He, Q. Self-Propelled Janus
and Deformation. Angew. Chem. 2012, 124, 7637−7640. Mesoporous Silica Nanomotors with Sub-100 nm Diameters for Drug
(21) Wang, W.; Luz Angelica, C.; Mauricio, H.; Thomas, E. M. Encapsulation and Delivery. ChemPhysChem 2014, 15, 2255−60.
Autonomous Motion of Metallic Microrods Propelled by Ultrasound. (42) Ilisz, I.; Föglein, K.; Dombi, A. The Photochemical Behavior of
ACS Nano 2012, 6, 6122−6132. Hydrogen Peroxide in Near UV-Irradiated Aqueous TiO2 Suspensions.
(22) Liu, M.; Thomas, Z.; Yongmin, L.; Guy, B.; Xiang, Z. Light- J. Mol. Catal. A: Chem. 1998, 135, 55−61.
Driven Nanoscale Plasmonic Motors. Nat. Nanotechnol. 2010, 5, 570− (43) Salvador, P.; Decker, F. The Generation of Hydrogen Peroxide
573. during Water Photoelectrolysis at N-Titanium Dioxide. J. Phys. Chem.
(23) Wu, Z.; Xiankun, L.; Yingjie, W.; Tieyan, S.; Jianmin, S.; Qiang, 1984, 88, 6116−6120.
H. Near-Infrared Light-Triggered “On/Off” Motion of Polymer (44) Li, J.; Singh, V. V.; Sattayasamitsathit, S.; Orozco, J.; Kaufmann,
Multilayer Rockets. ACS Nano 2014, 8, 6097−6105. K.; Dong, R.; Gao, W.; Jurado-Sanchez, B.; Fedorak, Y.; Wang, J.
(24) Mou, F.; Li, Y.; Chen, C.; Li, W.; Yin, Y.; Ma, H.; Guan, J. Water-Driven Micromotors for Rapid Photocatalytic Degradation of
Single-Component TiO2 Tubular Microengines with Motion Con- Biological and Chemical Warfare Agents. ACS Nano 2014, 8, 11118−
trolled by Light-Induced Bubbles. Small 2015, 11, 2564−70. 25.
(25) Solovev, A. A.; Smith, E. J.; Bof’ Bufon, C. C.; Sanchez, S.; (45) Li, J.; Gao, W.; Dong, R.; Pei, A.; Sattayasamitsathit, S.; Wang, J.
Schmidt, O. G. Light-Controlled Propulsion of Catalytic Micro- Nanomotor Lithography. Nat. Commun. 2014, 5, 5026.
engines. Angew. Chem., Int. Ed. 2011, 50, 10875−8. (46) Freitas, S.; Merkle, H. P.; Gander, B. Microencapsulation by
(26) Jiang, H.-R.; Yoshinaga, N.; Sano, M. Active Motion of a Janus Solvent Extraction/Evaporation: Reviewing the State of the Art of
Particle by Self-Thermophoresis in a Defocused Laser Beam. Phys. Rev. Microsphere Preparation Process Technology. J. Controlled Release
Lett. 2010, 105, 268302. 2005, 102, 313−332.
(27) Hong, Y.; Diaz, M.; Córdova-Figueroa, U. M.; Sen, A. Light-
Driven Titanium-Dioxide-Based Reversible Microfireworks and Micro-
motor/Micropump Systems. Adv. Funct. Mater. 2010, 20, 1568−1576.
(28) Fujishima, A. Electrochemical Photolysis of Water at a
Semiconductor Electrode. Nature 1972, 238, 37−38.
(29) O’Regan, B.; Grätzel, M. A Low-Cost, High-Efficiency Solar Cell
Based on Dye-Sensitized Colloidal TiO2 Films. Nature 1991, 353,
737−740.
(30) Liu, Y. Q.; Deng, C. B.; Xian, P.; He, J. H.; Li, X. P.; Xu, Y. B.;
Tang, M. The Research of the Performance of Active Carbon Loaded
Nano-Sized TiO2 in Formaldehyde Decontamination in Fluidized Bed
Reactor. Adv. Mater. Res. 2012, 518−523, 2925−2929.
(31) Ayati, A.; Ahmadpour, A.; Bamoharram, F. F.; Tanhaei, B.;
Mänttäri, M.; Sillanpäa,̈ M. A Review on Catalytic Applications of Au/
TiO2 Nanoparticles in the Removal of Water Pollutant. Chemosphere
2014, 107, 163−174.
(32) Buso, D.; Post, M.; Cantalini, C. Gold Nanoparticle-Doped
TiO2 Semiconductor Thin Films: Gas Sensing Properties. Adv. Funct.
Mater. 2008, 18, 3843−3849.

844 DOI: 10.1021/acsnano.5b05940


ACS Nano 2016, 10, 839−844

You might also like