Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Perspective

pubs.acs.org/JACS

Chemically Propelled Molecules and Machines


Krishna Kanti Dey*,† and Ayusman Sen*,‡

Department of Physics, Indian Institute of Technology Gandhinagar, Palaj, Gandhinagar 382355, Gujarat, India

Department of Chemistry, The Pennsylvania State University, University Park, Pennsylvania 16802, United States
responding to their environment (sensor applications) without
ABSTRACT: Self-propelled, synthetic active matters that being tethered to a single power source or location. In this
transduce chemical energy into mechanical motion are perspective, we focus on chemically powered motors and
examples of biomimetic nonequilibrium systems. They are pumps at submicroscopic length scales. Motors are objects that
of great current interest, with potential applications in can transduce chemical energy into mechanical motion through
Downloaded via INDIAN INST OF TECH GANDHINAGAR on March 20, 2024 at 09:06:26 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

nanomachinery, nanoscale assembly, fluidics, and chem- a variety of propulsion mechanisms. When these motors are
ical/biochemical sensing. Many of the physical challenges anchored onto a surface, their mechanical energy is transferred
associated with generating motility on the micro- and to the surrounding fluid, making them function as miniaturized
nanoscale have recently been overcome, leading to the first fluid pumps.


generation of autonomous motors and pumps on scales
ranging from microns to nanometers. This perspective NANOMOTOR DESIGNS: CHALLENGES AT LOW
focuses on catalytically powered motile systems, outlining REYNOLDS NUMBER
major advances to date in motor/pump design, propulsion
mechanisms and directional control, and intermotor In the natural world, controlled motion at the nano and micron
communications leading to collective behavior. We con- scale is ubiquitous. Motor proteins power the motion of
clude by discussing the possible future directions, from the bacterial flagella and cilia, facilitating the movement of these
fundamental questions that remain to be addressed to the tiny organisms through solutions.9 Some bacteria use short,
design principles required for useful applications. hair-like appendages known as pili for movement, that binds to
the host surface receptors and when retracted, pull the bacteria


along the surface.10 Active ion pumps inside the cellular interior
INTRODUCTION constantly transcribe chemical information by transporting
various ions against chemical potential gradients; maintaining
We live in the information age, driven by the ability to com- proper balance in intracellular osmolyte concentrations.11
municate and integrate ideas. However, this is also the age of Evolutions, however, had millions of years in optimizing the
miniaturization, with the ability to fabricate, manipulate, and extraordinary range of biological innovations and tools. Lacking
assimilate matter on the nanoscale facilitating the information the luxury of time, we must take advantage of our ability to
exchange.1−3 While we have become quite adept at fabricating intelligently design potential motors. In such a task, we must
materials on the small scale and manipulating the flow of consider the nature of physical laws operative at low Reynolds
energy through them, the ability to precisely control the motion number regimes and their consequences on fluid motion and
of the materials themselves is still in its infancy. Until recently, particle dynamics. Physics at submicroscopic length scales is
the work in this area has mostly focused on coaxing micro- and usually dominated by shear forces parallel to particles’ surface
nanostructures to deform, rotate, and move repeatedly over rather than by inertial forces which scale with particles’ volume
well-defined molecular-scale distances.4−7The pioneering work and which dictate their dynamics at the macroscale.12 Propul-
in this area has led to the award of the 2016 Nobel Prize in sion at micro- and nanometer length scales can therefore be
Chemistry to Savage, Stoddart, and Feringa.8 This perspective achieved effectively by generating various shear force gradients
focuses on a complementary, and arguably more challenging, across the particle surface. Additionally motions of tiny particles
avenue of research: How can we design populations of arti- in fluids are influenced by Brownian forces, which results in
ficial micro- and nanostructures that can move autonomously erratic random walk like trajectories at longer time scales.13 The
over long distances, while at the same time have the ability to biological world is dominated by random thermal fluctua-
organize themselves on command to perform complex tasks. tions but nature has developed machines that can rectify the
The potential applications of such synthetic interacting nano fluctuations to do useful work (e.g., Brownian ratchets such as
and micromachines would be almost limitless. Access to ratio- F0 motor in ATP synthase or actin).14 Molecules of kinesins
nally designed dynamic materials that are capable of remodeling and dyneins move along microtubules, and certain myosins are
themselves and transforming their environment will (i) mini- designed to move along actin filaments by overcoming the
mize waste (they will change their function and purpose rather strong Brownian fluctuations, driving the intracellular traffic of
than being of single-use), (ii) improve performance (they will vesicles and organelles with remarkable efficiency.15 Finally,
continuously evolve their structures to optimize performance), several physical forces, which are negligible across long dis-
and (iii) accomplish tasks collectively and emergently (like a tances, begin to dominate on the nanoscale. Electrostatic forces
colony of ants) that a single constituent element (like a single
ant) cannot perform. By making these dynamic materials Received: March 8, 2017
self-powered, they can be made capable of exploring and Published: May 11, 2017

© 2017 American Chemical Society 7666 DOI: 10.1021/jacs.7b02347


J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

between charged objects, for instance, grow proportional to metric force fields of some nature around each particles in an
ΔR2 as the distance between objects is decreased. The (usually assembly. In the presence of local force fields, the mobile ions
attractive) van der Waals interactions between an object and within the interfacial layer of each particle move and cause
its neighbors, on the other hand, grow by ΔR6.16 This is the liquid flow, making the particle drift in the opposite direction.
basis for the “sticky fingers” argument as to why it is physically Motions of this kind which arises due to forces operating within
impossible to build nano robots that reproduce themselves the interfacial layer of the particles, rather than on the bulk are
through atom by atom assembly.17 called phoresis. Molecules and microparticles, capable of
Nature of fluids at smaller length scales offers additional chemically interacting with their surroundings have recently
challenges in designing propulsion strategies for nanomotors. been shown to self-propel using different phoretic mecha-
The systems are characterized by very low Reynolds num- nisms.25−27 As described below, this offers a range of possi-
bers, and to achieve propulsion, it requires swimmers to bilities to maneuver micro- and nanoscale particles in liquid
deform in a way that is not invariant under time reversal. The independently, while simultaneously endowing them with
Reynolds number (Re) of the particle, which is usually mea- specific functionalities in an assembly.
sured as the ratio of inertial to the viscous force, can be
expressed as18
ρVL
■ ELECTROPHORESIS
Electrophoresis results when motion is induced in colloidal
Re = particles using electric fields. Dynamics of freely diffusing ions
μ (1) can also be influenced by external electric fields. For larger
Here, ρ is the density of the fluid, V is the velocity of the colloids, however, the mobile ions within the interfacial layer of
particle relative to the fluid, L is the characteristic length a particle interact with the electric field, resulting in fluid flow
scaleusually considered to be the dimension of the particle, around the particle surface. The fluid flow together with
and μ is the viscosity of the fluid. One far reaching consequence particle’s own electrostatic interaction with the field causes a
of low Re limit is kinematic reversibility, which facilitates lam- steady drift of the particle with respect to the fluid layer. The
inar flow of liquids within narrow confinementsa property electrophoretic speed of the particle can be calculated from the
that has been extensively exploited in microfluidic devices.19,20 magnitude of the electro-osmotic flow around it, using the
This has further been demonstrated by the famous Couette cell famous Smoluchowski equation:16
demonstration,21 where a drop of dye suspended in a high εζ E∞
viscosity liquid is sheared between two concentric cylinders and U=
stretched to wrap around the inner cylinder several times. The η (3)
motion is then reversed which brings back the drop of dye to its Here ε is the permittivity of the solution, ζ is the zeta potential
original shape. This is famous Scallop theorem of Purcell22 and of the particle, η is the viscosity of the solution and E∞ is the
microscopic swimmers overcome this hurdle by many fascinat- electric field. The electric field can be external or generated
ing swimming strategies that break the time reversal symmetry in situ through redox chemical reactions.
at low Reynolds number. This is very different from the macro- Self-electrophoresis has been demonstrated by several groups
scopic world where owing to the dominance of inertial motion, working on motors of different compositions and geometry.28−35
reciprocal swimming strategies can lead to effective propulsion The initial Pt/Au nanorod motor discovered by Paxton et al.
of objects in fluids.


functions based on self-electrophoresis (Figure 1), where
MECHANISMS OF MOTILITY AT MICRO- AND
NANOMETER LENGTH SCALES
Micro and nanoscale colloidal particles often have a few
molecular layers of fluid pinned over their surfaces by strong
van der Waal’s forces, resulting in what is known as the no-slip
condition.23 Beyond this pinned layer, charged colloids are Figure 1. Schematic of Pt/Au bimetallic nanomotor fabricated by
usually surrounded by a cloud of counterions forming what is Paxton et al.28,30 The rods self-electrophoretically move in hydrogen
peroxide solution by redox reactions occurring at their ends. Figure
known as the interfacial or double layer. The thickness of adapted from ref 30 with permission from the American Chemical
interfacial layer is decided by the particle surface charge and the Society.
local ionic strength of the medium, and is normally measured in
terms of the Debye length (κ−1 by convention) of the particle,
which in most systems lies within 1−100 nm.24 The Debye catalytic decomposition of hydrogen peroxide results in the
length can be calculated from eq 2,13 generation of an electric field, powering the nanorods.28,30
Mano et al. fabricated a bioelectrochemical motor by attaching
ε0εrkBT two electrochemically coupled enzymes to the opposite ends of
κ −1 = a carbon fiber.35 Several factors influence self-electrophoretic
2NAe 2I (2) motility of particles including the fluid medium, solution
where ε0 is the permittivity of free space, εr is the dielectric viscosity and ionic strength.36
constant of the electrolyte, kB is the Boltzmann constant, NA is
Avogadro’s number, e is the charge of an electron, and I is the
solution ionic strength.
■ DIFFUSIOPHORESIS
The generation of ionic products by catalytic particles in a
Each particle together with its interfacial layer is electrically liquid has additional effects on the particle double layer. First,
neutral and, without any external field, is in equilibrium. How- spontaneous electric field generated within the liquid by the
ever, they can be set into motion by a variety of external fields diffusing ions can result in electrophoresis of the particles.
and most motility mechanisms involves generation of asym- Moreover, as mentioned earlier, the thickness of double layer
7667 DOI: 10.1021/jacs.7b02347
J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

varies inversely with the local ionic strength and the gradient The parameter K is known as adsorption length, which physically
of an electrolyte can polarize the counterion cloud around a signifies the strength of particle-solute interaction.
particle. Such polarization causes ions from the higher solute Nonionic diffusiophoresis can also propel particles that
regions to migrate across the particle toward the other end, generate local solute gradients via chemical reactions. Possible
generating a fluid flow within the double layer. The particle nonionic self-diffusiophoretic motion was demonstrated by
in consequence, will drift steadily toward the higher solute Pavlick et al., where gold/silica Janus particles, coated with a
regions. Particle drift toward higher solute regions, arising only polymerization catalyst on the silica face, exhibited enhanced
as a result of double layer polarization is known as chem- diffusive motion in a solution of monomers.42 The polymeriza-
iphoresis. The combination of self-electrophoretic and chemi- tion reaction occurring on one face of the particle created a
phoretic effects leads to an overall diffusiophoretic interfacial concentration gradient of the monomer across the particle
flow, which powers the movement of particles.37−40 Depending surface. The gradient generates an interfacial fluid flow toward
on the nature of the solute added, diffusiophoresis can be higher concentration of monomers (Au face), eventually pro-
classified into two categories: (a) ionic diffusiophoresis, trig- pelling it with the catalyst end forward (Figure 2). Over longer
gered by an electrolyte38 and (b) nonionic diffusiophoresis,
triggered by a nonelectrolyte.37 For diffusiophoresis in a mono-
valent electrolyte, the propulsion speed can be approximated by
eq 4.38,39
⎧ ⎫
⎪ ⎪
⎪ εζ kBT ⎛ DC − DA ⎞ ε ⎛ kBT ⎞
2

U = ⎨− ⎜ ⎟+ ⎜ ⎟ ln(1 − ξ ) ⎬
2
4πη e ⎝ DC + DA ⎠ 
⎪  ⎝ e ⎠

2πη
  ⎪
⎪ ⎪
⎩ electrophoresis chemiphoresis ⎭
d ln C Figure 2. Schematic showing the propulsion mechanism of the
dx (4) polymerization powered motor. Fluid flow from low to high monomer
concentration regions causes the motor to move with the catalytic end
where U is the particle velocity, kB is the Boltzmann constant, T forward.42 Figure adapted from ref 42 with permission from John
is temperature, η is the viscosity of the solution, e is the charge Wiley and Sons.
of an electron, d ln C is the gradient of the electrolyte, DC is the
dx
diffusion coefficient of the cation, DA is the diffusion coefficient time scale, the motion however remains diffusive. Interestingly,
of the anion, ζ is the zeta potential of the particle surface and there are systems where the inverse mechanism has also been
realized. By asymmetric catalysis, the particles generate gradient
ξ = tanh ( )eζ
4kBT of product molecules across their surface, which propel them in
From the above discussion, it follows that gradient of an solution with their noncatalytic end forward.43−45 Powered
electrolyte imparts motion in colloidal particles independently motion of Pt/Si Janus particles is believed to follow this
through two distinct ways. Chemiphoresis arising out of mechanism. The catalytic Pt side of the motor decomposes
interfacial layer polarization always directs particles toward hydrogen peroxide into water and oxygen, which leads to the
higher solute concentrations.41 However, the direction of the creation of an oxygen gradient across the particle surface,
electrophoretic propulsion depends on the nature of particle’s setting it into motion.45
zeta potential as well as on the relative diffusivities of the ions One of the major advantages of powering nanomotors with
(DC and DA for cations and anions, respectively) produced as a nonionic self-diffusiophoresis is that it can propel particles in
result of the dissociation of the electrolyte. The resultant high ionic strength media. Electrolyte diffusiophoresis being
particle speed in ionic diffusiophoresis therefore, can be dependent on total ion concentration, is not effective in high
estimated by summing up these contributions. The diffusio- ionic strength media due to the collapse of interfacial layer on
phoresis propulsion mechanism can lead to many biomimetic the particle surface. However, in low ionic strength solutions,
collective emergent patterns, which will be discussed later. ionic diffusiophoresis is always more powerful and can result in
Gradient of neutral molecular solutes can also induce pro- much higher propulsion speeds than the nonionic counterpart.
pulsion in colloidal particles with specific directionality. This is due to the fact that both nonionic diffusiophoresis and
Molecules of a nonelectrolyte usually interact with the particle chemiphoretic part of ionic diffusiophoresis originates from
surface with short-range van der Waal’s forces. In nonionic similar sourcesasymmetries in solute concentrations across
diffusiophoresis, colloidal particles move toward regions of the particle surface. Their contributions toward the net propul-
higher or lower solute concentrations depending on the nature sion speed is therefore of the same order of magnitude. The
of solute-particle interaction. Velocity of a rigid spherical par- additional self-electrophoresis component in ionic diffusiopho-
ticle in the presence of a nonelectrolyte gradient, which is resis however augments the propulsion speed making it stronger
assumed locally parallel to the particle surface, was quantified propulsion mechanisms in low ionic strength conditions (eq 4).
by Anderson et al. as37,39
kT
U = B KL*
dC ■ COLLECTIVE DYNAMICS OF
SELF-DIFFUSIOPHORETIC MOTORS
η dx (5) Many microorganisms in nature use chemical signals derived
from each other to communicate and accomplish specific tasks
Here, dC is the magnitude of the unperturbed solute gradient, collectively. An example is chemotaxis which involves organisms
dx
L* is a measure of the range of particle-solute interaction. moving up or down a chemical gradient.46,47 The cells of
7668 DOI: 10.1021/jacs.7b02347
J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

Dictyostelium discoideum show robust chemotactic response for designing smart nanomachines that can carry out specific
upon sensing the signaling chemical 3′-5′-cyclic adenosine tasks collectively in complex environments. There are many
monophosphate (cAMP). cAMP is secreted when the cells are possible ways of designing ion-producing nano/microparticles,
starved and chemotaxis leads to their aggregation into a including particles with attached catalysts or enzymes that form
migrating slug, which then differentiates into a multicellular ions as products. These particles should then interact with each
fruiting body.48−50 The key aspect of this collective behavior is other or with inert nano/microparticles through their ionic
that these organisms both produce and respond to cAMP products. In addition, it allows coordinated and controlled
trigger for surviving harsh conditions of nutrient depletion. movement of particles that are not attached to each other,
Purely synthetic inorganic systems can also display similar col- which should find wide applicability in targeted material trans-
port and delivery.


lective dynamics. Micron-sized silver chloride (AgCl) particles
dispersed in deionized water collectively move under UV illu-
mination in response to self-generated ion gradients (Figure 3).51 TRANSITIONS BETWEEN COLLECTIVE BEHAVIORS
Apart from motility, colonies of microorganisms also display
novel transitions between different modes of collective behavior
in response to external stimuli. This provides motivation to
realize dynamic transitions in synthetic active systems, in order
to accomplish different tasks in succession with the same group
of multifunctional motors. Dynamic transition in active sys-
tems has been demonstrated in AgCl/UV system, when driven
by two different chemical reactions. Transition in collective
behavior occurs when H2O2 is added to the systemresulting
in periodic oscillations of particles.54 Two competing processes
are responsible behind the chemical clock. The reduction of
AgCl to Ag by UV light and the oxidation of Ag back to AgCl
by H2O2 produce and consume HCl, respectively, which leads
to periodic reversal of ion gradient. As a result, the AgCl par-
ticles exhibit an oscillatory attach−release motion with neigh-
boring silica spheres, as shown in Figure 6(a).
Figure 3. Schematic showing photocatalysis of AgCl particles resulting Using a reversible chemical reaction, transitions between two
in the formation of H+ and Cl− ions. Higher diffusivity of H+ than that modes of collective behavior has been demonstrated in system
of Cl− generates an electric field to which the nearby particles respond involving silver orthophosphate (Ag3PO4) microparticles dis-
and move with a diffusiophoretic speed Udp. The electric field also persed in dilute ammonia solution. The system reaches equi-
influences ions within the double layer of nearby surfaces creating fluid librium quickly and any subsequent addition or removal of
flow with diffusioosmotic speed Udo.51 Figure adapted from ref 51 with
permission from Royal Society of Chemistry. With sufficiently high
ammonia shifts the equilibrium, producing or consuming ions
number density, the particles respond to each other’s ion gradients, that include fast-diffusing OH−. The assembly then transitions
forming macroscopic schools (shown in Figure 4 below).52 between clustering and dispersion (Figure 6b).55 The transition
can also occur in response to UV light, and with two orthog-
Like Dictyostelium discoideum, each AgCl particle secretes onal stimuli, the behavior of the system can be used to design a
the chemoattractant in the form of H+ and Cl− ions as it colloidal logic (NOR) gate (Figure 6c).
moves, to which the other particles respond and modulate
their dynamics. This eventually forms “schools” of AgCl par-
ticles, mimicking the slime mold’s behavior in two dimensions
■ BUBBLE PROPELLED SYSTEMS
Propulsion induced in small scale machines by bubble recoil
(Figure 4). is another widely investigated mechanism.56−62 Systems with
An even more interesting dynamic behavior is observed with different geometries have been shown to display impressive
a mixture of photo activated AgCl particles and silica micro- propulsion speeds following chemically generated bubble recoil.
spheres dispersed in deionized water. Upon UV illumination, The motion can be controlled using external (e.g., magnetic)
silica spheres respond to the ions secreted by AgCl particles fields,63,64 which is ideal for certain applications where motors
and actively surround them, displaying predator−prey behavior, must reach a destination quickly or move against opposing
like that of neutrophils (Figure 5).53 Such long-range emergent fluid flows, e.g., blood, with varying speeds and forces. These
interaction in active particle assemblies offers new possibilities high speed motors are easy to design and can find useful

Figure 4. Schooling behavior of AgCl particles in deionized water (A) before UV illumination, (B) after 30 s of UV exposure, and (C) after 90 s of
UV exposure.52 Figure reproduced from ref 52 with permission from John Wiley and Sons.

7669 DOI: 10.1021/jacs.7b02347


J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

Figure 5. Predator−prey behavior displayed by photoactivated AgCl particles and silica microspheres dispersed in deionized water. Upon UV
illumination, AgCl particles are gradually surrounded by silica particles (b and c). Around each AgCl particle an exclusion zone is seen, which
disappears when the UV is turned off. The disappearance of exclusion zones is seen in (d). Times in second are indicated in the upper left corners of
the images while the status of UV illumination is mentioned in the bottom left.52 Figure reproduced from ref 52 with permission from John Wiley
and Sons.

Figure 6. Transition between collective behaviors. (a) Oscillatory attach-release interactions between active silver chloride particles and nearby
passive silica spheres due to competing redox reactions;54 (b) silver orthophosphate particles disperse and cluster reversibly with addition and
removal of ammonia; (c) The transition from dispersion to clustering in (b) can be halted by UV, and the two different responses of the particles
under two orthogonal inputs enable the design of the universal “NOR” logic gate.55 Figures reproduced from refs 54 and 55 with permission from
the American Chemical Society.

Figure 7. Catalytic micromotors propelled by asymmetric generation and release of bubbles, demonstrating model applications such as (A) water
purification60 and (B) drug delivery.57 Figures reproduced from refs 60 and 57, respectively, with permission from the American Chemical Society.

Figure 8. Diffusion of (A) urease67 and (B) catalase68 increased as a function of increased substrate concentrations (urea and H2O2 respectively).
(* indicates a significance value of p < 0.05). (C) Enzyme-functionalized microparticles behave as hybrid autonomous motors in substrate solution.78
Figures adapted from refs 67, 68, and 78, respectively, with permission from the American Chemical Society.

nonbiological applications such as cargo delivery57 and water


purification60 (Figure 7). However, their use in living systems
■ SELF-POWERED MOLECULAR MOTORS
Along with the development of all-inorganic and hybrid motile
can be limited because of harmful effects of bubbles. systems, a major incentive has been to design biocompatible
7670 DOI: 10.1021/jacs.7b02347
J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

Figure 9. (A) Schematic of enzyme chemotaxis in Y-shaped microfluidic channel. In the presence of an imposed substrate concentration gradient,
active enzymes collectively move toward areas of higher substrate concentrations.68,69 (B) Chemotactic separation of active enzymes from their
inactive forms in the presence of imposed substrate concentration gradient.89 Figures adapted from refs 68 and 88, respectively, with permission from
the American Chemical Society.

active systems powered by biological fuels. Molecules of active biological systems, the ideal way to power motion in nanoscale
enzymes, which catalyze various biomolecular reactions with systems is though catalysis of chemical reactionsa critical
extraordinary efficiency and specificity, offer excellent model element missing from most work on synthetic molecular sys-
systems to develop smart molecular motors.65 In a series of tems. Thus, it is important to investigate catalytically powered
experiments, we demonstrated that active enzymes, while turn- motion that is compatible with this length scale. Using diffu-
ing over substrates, generate enough mechanical force to sion NMR spectroscopy, we have examined the behavior of a
enhance their own diffusion in solution (Figure 8(A,B)).66−69 ruthenium-based Grubbs catalyst during its catalysis of the
The diffusive movement of the enzymes increases with ring-closing metathesis of diethyldiallylmalonate (DDM).86 As
increasing reaction rate, following Michaelis−Menten kinetics. with enzymes, the diffusion of the ∼6 Å radius organometallic
In order to account for the observed enhanced diffusion of catalysts was found to increase with increasing reaction rate.
enzymes during substrate turnover, a number of theoretical Additionally, the diffusion of the catalyst returned to the base
proposals have recently been put forward. Propulsion through value upon either the addition of an inhibitor or the completion
self-generated diffusiophoretic forces has been suggested for of the reaction. Density functional theory (DFT) calculations
molecules like urease, which generate ionic products during were performed to probe the effect of structural changes in the
reaction.70 Motility resulting from reaction-induced conforma- ruthenium complex as a source of the increased diffusion. The
tional changes in enzymes has also been proposed as a pos- conclusion was that the difference between the largest and the
sibility.71 Bustamante et al. attributed the enhanced diffusion of smallest form of the catalyst was not significant enough to
enzymes to the exothermicity of the chemical reactions account for the increased catalyst diffusivity at most substrate
involved, where motion is induced via a local thermophoretic concentrations.
effect.72 More recently, Golestanian estimated the magnitude of
diffusion enhancement originating from different possible
mechanisms and suggested that stochastic swimming due to
■ MOLECULAR CHEMOTAXIS
In the presence of an imposed substrate concentration gradient,
collective heating and conformational changes are more likely active enzymes collectively move toward areas of higher sub-
to account for the experimentally measured diffusion enhance- strate concentrations, displaying a novel example of molecular
ment of the enzymes studied thus far.73 Using Langevin/ chemotaxis (Figure 9A).68,69,87 The collective migration of
Brownian dynamics simulations, it was determined that a force enzymes toward specific targets promises a variety of applica-
of ∼10 pN per turnover was sufficient to cause the enhance- tions including sensing, targeted transport, and delivery. The
ment in diffusion of urease and catalase.66−68 These forces are exact mechanism behind enzyme chemotaxis, however, remains
comparable to that produced by myosin, kinesin, and dynein an open question. It may involve a Brownian ratchet mech-
motors (approximately 10 pN),74−76 and within the range anism arising from substrate-dependent enhanced diffusion.88
necessary to activate integrins,77 biological adhesion molecules As with any nonequilibrium system, a continuous energy input
responsible for mechanosensing by cells, making force produc- is required for the directional movement, in this case, to
tion by enzyme catalysis a potentially novel mechanobiology- maintain the substrate gradient. A second possible mechanism
relevant event. Moreover, by tethering active enzymes to larger involves the favorable enzyme−substrate interaction. If the
micronscale particles, Sanchez et al. and we have independently enzyme−substrate complex is more stable than the enzyme and
reported enhanced motion of enzyme-powered motors and the substrate separately, then the enzyme will diffuse to regions
demonstrated that biocatalytic reactions may be used to power of higher substrate concentrations because of the resulting
large microscopic objects in solution (Figure 8(C)).78−80 lowering of the chemical potential for the system. Irrespective
The interesting behavior of self-powered systems at the of the mechanism, however, molecular chemotaxis is stochastic
micro- and nanometer length scales encouraged us to extend in nature and is different from biological chemotaxis, which
our investigation to even smaller active molecular systems. requires organisms to have temporal memory of the concen-
There has been significant work on nanometer-scale synthetic tration gradient.48,49 The observation of chemotaxis suggests
molecular systems, including rotors synthesized by the Feringa,81 that, for directed motion, it is not necessary for the active
nanocars designed by Tour,82 fluctuation driven transport molecule or particle to be asymmetric; one simply needs a
proposed by Astumian,83 and the shuttle systems discovered gradient in the substrate concentration.
by Leigh,84 Stoddart85 and others that move molecules along The observed chemotactic behavior of single enzymes
chains. These systems are useful as switches, transport devices, suggests that for an enzyme that acts on the products of a
and even as subnanometer pumps. Again, taking a cue from the second, nearby enzymatic reaction might result in its collective
7671 DOI: 10.1021/jacs.7b02347
J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

movement toward the second enzyme; an example of collective surface. The negatively charged enzyme molecules bound
behavior at the molecular level. Thus, catalase was observed to selectively to the SAM-functionalized gold patterned surface via
migrate toward glucose oxidase in the presence of glucose since electrostatic assembly, resulting in an enzyme pattern on the
hydrogen peroxide is a product in the oxidation of glucose.68 surface (Figure 10(A)). Alternatively, enzymes can also be
Importantly, the novel observation of the universality of force attached to the gold patch using biotin−streptavidin linkage
production by enzymes and the ability to harness theses forces procedure.91,92 To monitor the fluid flow, sulfate-functionalized
for directed molecular chemotaxis up a substrate concentration polystyrene microspheres were used as tracers. Only in the
gradient allows the rational design of new technologies for the presence of the substrate, the tracer particles move away from
separation of active enzymes from a complex media. The (urease) or toward (catalase, lipase and glucose oxidase) the
chemotactic separation of active biomolecules from a mixture gold surface, indicating that the surrounding fluid is pumped
has been demonstrated to be sensitive enough to sort out outward or inward, depending on the specific enzyme pump.
molecules possessing identical physical properties, which The pumping velocity increases with increasing substrate con-
cannot otherwise be accomplished using currently known centration and reaction rate. The fluid flow is driven by a
separation techniques (Figure 9(B)).89 In principle, the tech- gradient in fluid density generated primarily through changes in
nique of chemotactic separation can also be used to separate the solutal composition, rather than due to reaction exother-
out other active catalysts from their less active or inactive micity.91,92 These rechargeable pumps can be triggered by the
counterparts in the presence of their respective substrates and presence of specific analytes, which enables the design of
should, therefore, find wide applicability. We also observed enzyme-based smart devices capable of acting both as sensors
controlled chemotactic propulsion of enzyme-functionalized and pumps. Similar pumping can occur in gels incorporating
microscopic particles, a promising step toward developing immobilized enzymes in the presence of substrates. For example,
smart biocompatible payload carriers in complex physiological gel-bound glucose oxidase pumps insulin out of gel particles
environments.78 when glucose is added to the surrounding solution.90 This


research addressed one of the major drawbacks of many
synthetic micropumps that are nonbiocompatible and do not
MOLECULAR PUMPS
interface well with the biological systems. An interesting
Enzyme molecules, when immobilized on a surface, transfer the application of enzyme pumps is in the area of toxin detection
mechanical energy derived from catalysis to the surrounding where the attenuation in pumping velocity can be related to the
fluid, resulting in directional fluid pumping. In effect, surface- concentration of toxins that inhibit the enzymatic reaction.94
bound enzymes function as chemically powered machines Using this principle, sensors for toxic substances, like mercury,
(Figure 10(A)).90−93 This was demonstrated using four differ- cadmium, cyanide and azide, were designed using urease and
ent enzymes as modelscatalase, urease, lipase and glucose catalase-powered pumps, respectively, with limits of detection
oxidase.90 Using e-beam evaporator, gold was patterned on a well below the concentrations permitted by the Environmental
polyethylene glycol (PEG)-coated glass surface. The patterned Protection Agency (EPA) (Figure 10(B)).
surface was functionalized with a quaternary ammonium thiol,
which formed a self-assembled monolayer (SAM) on the Au ■ LONG RANGE EFFECTS OF MOLECULAR MOTOR
SYSTEMS
An important aspect of the dynamics of active catalyst particles
is their effect on their immediate surroundings. Studies con-
ducted with micron-scale organisms and catalytic particles show
that such active assemblies exert considerable influence over the
behavior of their surroundings.95−101 Diffusion of inert tracer
particles dispersed in a suspension of micron-scale swimmers is
dependent upon the total activity of a system; the higher the
total activity of the system, the greater the diffusion enhance-
ment of the tracers. Qualitatively, the enhancement was found
to be independent of swimming patterns and mechanisms,
signifying the generic role of dynamic coupling among the
swimmers and their surroundings, which facilitates the
distribution of momentum. A key question is whether these
effects are also observable at the very low Reynolds number
regime (i.e., at the molecular scale). A recent theory predicts
that the advection effects induced by active molecular catalysts
should also result in significant enhancement of diffusion of
passive molecules present in solution.102 Accordingly, we mea-
sured the diffusion of molecular tracers, benzene and tetra-
Figure 10. (A) Schematic showing formation of enzyme patterns on methylsilane (TMS), in a solution of Grubbs catalyst using
PEG coated glass surface and fluid pumping by immobilized enzymes diffusion NMR spectroscopy.103 Interestingly, the diffusion of
in the presence of substrates.90−93 Figure adapted from ref 90 with
permission from the Nature Publishing Group. (B) Inhibitor assay
these inert molecules also increased with increasing reaction
based on enzyme powered micropumps. The presence of inhibitor is rate and returned to the respective base values upon the com-
detected by monitoring changes in the fluid flow speed, the attenua- pletion of the reaction. The experimental observations demon-
tion of which is directly correlated with the inhibitor concentration.94 strate transfer of momentum from the active catalyst molecules
Figure reproduced from ref 94 with permission from John Wiley and to the surrounding medium, the nature of which is surpris-
Sons. ingly similar to the behavior of microscopic active swimmers.
7672 DOI: 10.1021/jacs.7b02347
J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

Note that the enthalpy change for the ring-closing metathesis move faster or longer, little have been done on optimization of
of DDM is approximately +8 kcal/mol (endothermic)86 and the fuel or geometries. Such factors will significantly impact
therefore cannot contribute to the observed enhanced diffusion motor speed, fluid drag, fuel consumptions and other factors.
of the tracers. The most plausible mechanism involves reaction- In particular, motor systems need to be examined in greater
generated advection caused by catalyst molecules “stirring” the detail in terms of efficiency and compatibility with respect to
solution, which is of great importance especially in intracellular intended applications.112
mechanics.104−109 The optimization of motor geometries is critically important.
Energy transduction by enzymes at the single molecule and Currently, the majority of motors have cylindrical, spherical or
collective level opens up a new area in molecular biophysics. screw-like structures. Such geometries may not be best suited
Enzymes, in addition to their primary function as catalysts for for their future intended applications. Motors for in vivo
biochemical reactions, may have secondary functions as applications have limitations on their shape and size. They must
mediators of dynamic interaction pathways. In addition, the be large enough for directional propulsion while being small
possibility of enhanced diffusive transport of molecules by and flexible enough to squeeze through narrow capillaries. If the
controlled “stirring” in physiological environment has impor- interior of living cells is the intended target, the easiest way to
tant scientific and technological relevance. The phenomenon enter is through endocytosis which depends on the size and
could provide new insights into the dynamics of cytoplasmic shape of the motor.12
glass-gel transitions and enhanced mixing in bacterial cells, Thus, far, the focus of motor research has been on the
observed during various metabolic transformations.108 Finally, development of motors that move by one dominant propulsion
momentum transfer from freely diffusing active catalysts mechanism. However, incorporating orthogonal propulsion
suggests the possibility that membrane-bound enzymes exert mechanisms into a motor offers a greater degree of control.
force on cell-membranes, amplifying mechanically induced Examples include magneto-acoustic hybrid nanomotors,113 as
signaling mechanisms.109 well as motors that encompass opposing chemical and acoustic

■ FANTASTIC VOYAGE
External field-driven motion leads to “lock-step” ensemble
propulsion.114 These are the first steps toward the fabrication of
true multifunctional motors by incorporating different
components for propulsion (multiple fuels), triggered collective
behavior. To realize the dream of Asimov’s Fantastic Voyage, behavior (cargo capture/release, schooling/exclusion), and
the bots must be autonomous, harvesting energy from the targeting behavior (response to multiple stimuli).112
surroundings and reacting to information at the local level.110 An important goal is to use motors as micro/nano machines
Investigating systems involving molecules and micro/ to accomplish what our current tools and instruments cannot.
nanoparticles is inherently challenging due to the stochastic While there have been proof-of-concept applications in recent
nature of particle dynamics, the limitations of experimental and publications, we have yet to find a niche where motors are
modeling techniques in characterizing nonequilibrium systems, indispensable or the technique of choice for a given application.
and need for multidisciplinary expertise.12 As such, the field is expected to evolve into one with more
Despite the progress in the design of synthetic motors, they emphasis on application-based systems. Perhaps the most
cannot carry out complex tasks like their biological counter- attractive application for catalytic motors, given the scenario
parts. More integrated functionalities and “division of labor” are in Fantastic Voyage, is medicine. The immense advantage of
two key elements in future design of synthetic motors. The targeted motion over diffusion, allowing for the use of less
ultimate objective of research in this area is to create a new material, is a major incentive for the use of autonomous motors
paradigm for the design of active functional materials and for drug or cargo delivery.112,115−117 Biomedical applications of
systems by leveraging (a) precise molecular-level control of motors, however, remain relatively unexplored, in part due to
materials to create functional building blocks, (b) mobility the difficulty of motor navigation in vivo. Most studies on
resulting from biomimetic catalytic energy harvesting from the motors are conducted in water with low ion content. This is in
local environment, (c) rapid and reversible assembly capabil- contrast to biological fluids, which contain a multitude of ions
ities provided by emergent self-assembly, (d) intelligence and and other components. Blood, for example, is a non-Newtonian
communication capabilities found in interacting microorgan- viscous fluid that contains plasma, platelets, red and white
isms, and (e) ability to perform specific tasks in response to blood cells that may affect motors’ movement in many different
signals from each other and the environment. ways. Several additional hurdles remain before practical in vivo
From a fundamental standpoint, the scientific questions applications of motors become a reality. First, the nano/
that need to be addressed going forward are (a) What are the microtransporters must be derived from biocompatible mate-
possible mechanisms for momentum creation at the micro- rials with surfaces appropriately functionalized to promote cell
and nanometer length scales and their efficiencies in different viability. Second, and equally important, is to design self-
environments? (b) Are there optimal motor geometries for powered nano/microbots that can use fuels that are biocom-
sustaining directional motion? (c) How well can motion be patible, preferably fuels present in the body. Ideally, the nano/
directed, preferentially by chemical gradients (i.e., chemotaxis)? microtransporters will employ enzymes as catalysts and fuels
(d) What is the nature of interparticle interactions in driven (e.g., glucose) present in living systems.118−120 Third, the
systems? (e) How do the ensemble dynamics of active materials transporters need to be powerful enough to move against fluid
evolve in space-time? and (f) How to we effectively integrate flows, such as blood flow. Furthermore, the energy transduction
synthetic active molecular assemblies into the biological world? mechanism should operate in high ionic medium present in
The majority of the motors systems to date have not been biological fluids. Finally, the most “futuristic” scenario involves
optimized. There has only been one study so far on efficiency the design of populations of synthetic nano and micromotors
of artificial motor systems by Wang, Mallouk, and collabo- and pumps that have the ability to organize themselves intel-
rators.111 While the focus of nanomotors has been on the ligently, based on signals from each other and from their
development of new systems and how to make the motors environment, to perform complex tasks. Particularly attractive
7673 DOI: 10.1021/jacs.7b02347
J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

are designs that allow coordinated movement of particles with


different functionalities that are not attached to each other,
■ AUTHOR INFORMATION
Corresponding Authors
making it easier to transport and deliver cargo at specific areas
*k.dey@iitgn.ac.in
as per requirements. The discovery of particle assemblies that
*asen@psu.edu
exhibit chemotaxis and predator−prey behavior is a step in this
direction. For many future applications, it will be important ORCID
to have motors that can independently carry out operations Ayusman Sen: 0000-0002-0556-9509
such as sensing and reporting, with different populations of Notes
interacting motors performing different tasks. The authors declare no competing financial interest.


The observation that enzymes such as urease, catalase, lipase,
DNA polymerase, and others undergo powered movement and ACKNOWLEDGMENTS
pump fluids as they turn over their substrates amends the
paradigm that ATP-powered biomotors are a special class of We gratefully acknowledge Penn State MRSEC under NSF
enzymes. This also clearly suggests that (a) single enzyme grant DMR-1420620 and IIT, Gandhinagar for current financial
molecules generate sufficient mechanical force through sub- support and Anamika Dey for drawing some of the figures.
strate turnover to cause their own movement and that of the We thank the students and collaborators who have made
surrounding particles and fluid and (b) the movement becomes possible the results described in this perspective. A.S. also
thanks Dr. Suchismita Sen for insightful discussions.


directional through the imposition of a gradient in substrate
concentration. Indeed, other than the presence of “tracks” that
provide directionality, there may not be a significant difference REFERENCES
between traditional motor proteins and free swimming (1) Peercy, P. S. Nature 2000, 406, 1023−1026.
enzymes. In addition, it will be interesting to examine the (2) Madou, M. J. Fundamentals of Microfabrication: The Science of
role of membrane-bound enzymes in cell membrane fluctua- Miniaturization, 2nd ed.; CRC Press: Boca Raton, FL, 2002.
(3) Lee, M. L. J. Microcolumn Sep. 1989, 1, 1.
tions. The results described open up a new area of mech- (4) Schliwa, M.; Woehlke, G. Nature 2003, 422, 759−765.
anobiology: intrinsic force generation by enzymes and their role (5) Browne, W. R.; Feringa, B. L. Nat. Nanotechnol. 2006, 1, 25−35.
in biochemical regulation of cell function. These enzymes (6) Stoddart, J. F. Acc. Chem. Res. 2001, 34, 410−411.
can provide sufficient force for the stochastic motion of the (7) Erbas-Cakmak, S.; Leigh, D. A.; McTernan, C. T.; Nussbaumer,
cytoplasm and the convective transport of fluid in cells.121 A. L. Chem. Rev. 2015, 115, 10081−10206.
Further, they may be responsible for the observed glass to fluid (8) Astumian, R. D. Chem. Sci. 2017, 8, 840−845.
transition in active bacterial cells.108 (9) Cooper, G. M. The Cell, A Molecular Approach, 2nd ed.; Sinauer
The interaction between enzymes in living cells is an area of Associates: Sunderland, MA, 2000.
(10) Merz, A. J.; So, M.; Sheetz, M. P. Nature 2000, 407, 98−102.
active research. In many instances, enzymes that participate in (11) Gouaux, E.; MacKinnon, R. Science 2005, 310, 1461−1465.
reaction cascades have been shown to assemble into meta- (12) Dey, K. K.; Wong, F.; Altemose, A.; Sen, A. Curr. Opin. Colloid
bolons in response to the presence of the initial substrate to Interface Sci. 2016, 21, 4−13.
facilitate substrate channeling.122−124 The mechanism for (13) Sengupta, S.; Ibele, M. E.; Sen, A. Angew. Chem., Int. Ed. 2012,
metabolon formation, however, remains uncertain in many 51, 8434−8445.
cases. Metabolon formation through noncovalent interactions (14) Peskin, C. S.; Odell, G. M.; Oster, G. F. Biophys. J. 1993, 65,
has been suggested but rarely demonstrated.125,126 Our obser- 316−324.
vation of preferential migration of enzymes up the substrate (15) Hess, H.; Vogel, V. Rev. Mol. Biotechnol. 2001, 82, 67−85.
(16) Velegol, D. Colloidal Systems; Wild Scholars Media on Amazon,
gradient suggests that enzymes along a metabolic pathway in 2016.
which the product of one is the reactant for the next may (17) Smalley, R. E. Sci. Am. 2001, 285, 76−77.
associate through a process of sequential, directed chemotactic (18) Yadav, V.; Duan, W.; Butler, P. J.; Sen, A. Annu. Rev. Biophys.
movement. 2015, 44, 77−100.


(19) Stone, H. A.; Stroock, A. D.; Ajdari, A. Annu. Rev. Fluid Mech.
2004, 36, 381−411.
CONCLUSION
(20) Lee, C.-Y.; Chang, C.-L.; Wang, Y.-N.; Fu, L.-M. Int. J. Mol. Sci.
Chemically powered micro- and nanoscale motors offer immense 2011, 12, 3263−3287.
potential toward mimicking natural biomolecular systems (21) Popova, M.; Vorobieff, P.; Ingber, M. Computational Methods in
promising a myriad of novel technological innovations.127 Multiphase Flow IV; WIT Press: Southampton, U.K., 2007.
However, the fabrication of tiny, efficient synthetic motors is (22) Purcell, E. M. Am. J. Phys. 1977, 45, 3−11.
(23) Day, M. A. Erkenntnis. 1990, 33, 285−296.
often associated with scientific and engineering challenges, (24) Karnik, R.; Fan, R.; Yue, M.; Li, D.; Yang, P.; Majumdar, A.
which need to be tackled with complementary theoretical and Nano Lett. 2005, 5, 943−948.
experimental approaches. Since the discovery of the first (25) Wang, W.; Duan, W.; Ahmed, S.; Mallouk, T. E.; Sen, A. Nano
chemically powered system,128 many designs and propulsion Today 2013, 8, 531−554.
mechanisms for autonomous motors have been investigated (26) Golestanian, R.; Liverpool, T. B.; Ajdari, A. New J. Phys. 2007, 9,
demonstrating their applications in transport, assembly, and 126.
sensing. Freed of usual biological constraints, we now have the (27) Moran, J. L.; Posner, J. D. Annu. Rev. Fluid Mech. 2017, 49,
511−540.
unprecedented opportunity to probe the limits of self-organiza-
(28) Paxton, W. F.; Kistler, K. C.; Olmeda, C. C.; Sen, A.; St. Angelo,
tion in these synthetic systems that operate far from equilib- S. K.; Cao, Y.; Mallouk, T. E.; Lammert, P. E.; Crespi, V. H. J. Am.
rium.129 Thus, it is possible to imagine a day when intelligent Chem. Soc. 2004, 126, 13424−13431.
machines navigate through the body and perform critical tasks, (29) Fournier-Bidoz, S.; Arsenault, A. C.; Manners, I.; Ozin, G. A.
realizing the functioning of Asimov’s Proteus. Chem. Commun. 2005, 441−443.

7674 DOI: 10.1021/jacs.7b02347


J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

(30) Paxton, W. F.; Baker, P. T.; Kline, T. R.; Wang, Y.; Mallouk, T. (69) Sengupta, S.; Spiering, M. M.; Dey, K. K.; Duan, W.; Patra, D.;
E.; Sen, A. J. Am. Chem. Soc. 2006, 128, 14881−14888. Butler, P. J.; Astumian, R. D.; Benkovic, S. J.; Sen, A. ACS Nano 2014,
(31) Ibele, M. E.; Wang, Y.; Kline, T. R.; Mallouk, T. E.; Sen, A. J. 8, 2410−2418.
Am. Chem. Soc. 2007, 129, 7762−7763. (70) Colberg, P. H.; Kapral, R. Europhys. Lett. 2014, 106, 30004.
(32) Moran, J. L.; Posner, J. D. J. Fluid Mech. 2011, 680, 31−66. (71) Cressman, A.; Togashi, Y.; Mikhailov, A. S.; Kapral, R. Phys. Rev.
(33) Yariv, E. Proc. R. Soc. London, Ser. A 2011, 467, 1645−1664. E 2008, 77, 050901.
(34) Liu, R.; Sen, A. J. Am. Chem. Soc. 2011, 133, 20064−20067. (72) Riedel, C.; Gabizon, R.; Wilson, C. A. M.; Hamadani, K.;
(35) Mano, N.; Heller, A. J. Am. Chem. Soc. 2005, 127, 11574− Tsekouras, K.; Marqusee, S.; Pressé, S.; Bustamante, C. Nature 2015,
11575. 517, 227−230.
(36) Pumera, M. Nanoscale 2010, 2, 1643−1649. (73) Golestanian, R. Phys. Rev. Lett. 2015, 115, 108102.
(37) Anderson, J. L.; Lowell, M. E.; Prieve, D. C. J. Fluid Mech. 1982, (74) Mahadevan, L.; Matsudaira, P. Science 2000, 288, 95−99.
117, 107−121. (75) Mehta, A. D.; Rief, M.; Spudich, J. A.; Smith, D. A.; Simmons, R.
(38) Prieve, D. C.; Anderson, J. L.; Ebel, J. P.; Lowell, M. E. J. Fluid M. Science 1999, 283, 1689−1695.
Mech. 1984, 148, 247−269. (76) Yin, H.; Wang, M. D.; Svoboda, K.; Landick, R.; Block, S. M.;
(39) Anderson, J. L. Annu. Rev. Fluid Mech. 1989, 21, 61−99. Gelles, J. Science 1995, 270, 1653−1657.
(40) Keh, H. J.; Luo, S. C. Phys. Fluids 1995, 7, 2122−2131. (77) Wang, N.; Butler, J. P.; Ingber, D. E. Science 1993, 260, 1124−
(41) Lee, S. Y.; Yalcin, S. E.; Joo, S. W.; Baysal, O.; Qian, S. J. Phys. 1127.
Chem. B 2010, 114, 6437−6446. (78) Dey, K. K.; Zhao, X.; Tansi, B. M.; Méndez-Ortiz, W. J.;
(42) Pavlick, R. A.; Sengupta, S.; McFadden, T.; Zhang, H.; Sen, A. Córdova-Figueroa, U. M.; Golestanian, R.; Sen, A. Nano Lett. 2015, 15,
8311−8315.
Angew. Chem., Int. Ed. 2011, 50, 9374−9377.
(79) Ma, X.; Wang, X.; Hahn, K.; Sánchez, S. ACS Nano 2016, 10,
(43) Zhang, H.; Yeung, K.; Robbins, J. S.; Pavlick, R. A.; Wu, M.; Liu,
3597−3605.
R.; Sen, A.; Phillips, S. T. Angew. Chem., Int. Ed. 2012, 51, 2400−2404. (80) Ma, X.; Hortelao, A. C.; Miguel-López, A.; Sánchez, S. J. Am.
(44) Howse, J. R.; Jones, R. A. L.; Ryan, A. J.; Gough, T.; Vafabakhsh,
Chem. Soc. 2016, 138, 13782−13785.
R.; Golestanian, R. Phys. Rev. Lett. 2007, 99, 048102. (81) Fletcher, S. P.; Dumur, F.; Pollard, M. M.; Feringa, B. L. Science
(45) Ke, H.; Ye, S.; Carroll, R. L.; Showalter, K. J. Phys. Chem. A 2015, 310, 80−82.
2010, 114, 5462−5467. (82) Shirai, Y.; Osgood, A. J.; Zhao, Y.; Kelly, K. F.; Tour, J. M. Nano
(46) Adler, J. Science 1966, 153, 708−716. Lett. 2005, 5, 2330−2334.
(47) Webre, D. J.; Wolanin, P. M.; Stock, J. B. Curr. Biol. 2003, 13, (83) Astumian, R. D. Science 1997, 276, 917−922.
R47−R49. (84) Kay, E. R.; Leigh, D. A.; Zerbetto, F. Angew. Chem., Int. Ed.
(48) Schäfer, E.; Tarantola, M.; Polo, E.; Westendorf, C.; Oikawa, N.; 2006, 46, 72−191.
Bodenschatz, E.; Geil, B.; Janshoff, A. PLoS One 2013, 8, e54172. (85) Tseng, H.-R.; Vignon, S. A.; Stoddart, J. F. Angew. Chem. 2003,
(49) Cai, H.; Huang, C.-H.; Devreotes, P. N.; Iijima, M. Methods Mol. 115, 1529−1533.
Biol. 2011, 757, 451−468. (86) Pavlick, R. A.; Dey, K. K.; Sirjoosingh, A.; Benesi, A.; Sen, A.
(50) Gerisch, G. Annu. Rev. Physiol. 1982, 44, 535−552. Nanoscale 2013, 5, 1301−1304.
(51) Sen, A.; Ibele, M.; Hong, Y.; Velegol, D. Faraday Discuss. 2009, (87) Yu, H.; Jo, K.; Kounovsky, K. L.; de Pablo, J. J.; Schwartz, D. C.
143, 15−27. J. Am. Chem. Soc. 2009, 131, 5722−5723.
(52) Ibele, M.; Mallouk, T.; Sen, A. Angew. Chem., Int. Ed. 2009, 48, (88) Peskin, C. S.; Odell, G. M.; Oster, G. F. Biophys. J. 1993, 65,
3308−3312. 316−324.
(53) Billadeau, D. D. Nat. Immunol. 2008, 9, 716−718. (89) Dey, K. K.; Das, S.; Poyton, M. F.; Sengupta, S.; Butler, P. J.;
(54) Ibele, M. E.; Lammert, P. E.; Crespi, V. H.; Sen, A. ACS Nano Cremer, P. S.; Sen, A. ACS Nano 2014, 8, 11941−11949.
2010, 4, 4845−4851. (90) Sengupta, S.; Patra, D.; Ortiz-Rivera, I.; Agrawal, A.; Shklyaev,
(55) Duan, W.; Liu, R.; Sen, A. J. Am. Chem. Soc. 2013, 135, 1280− S.; Dey, K. K.; Córdova-Figueroa, U.; Mallouk, T. E.; Sen, A. Nat.
1283. Chem. 2014, 6, 415−422.
(56) Gao, W.; Pei, A.; Wang, A. ACS Nano 2012, 6, 8432−8438. (91) Ortiz-Rivera, I.; Shum, H.; Agrawal, A.; Sen, A.; Balazs, A. C.
(57) Mou, F.; Chen, C.; Zhong, Q.; Ying, Y.; Ma, H.; Guan, J. ACS Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 2585−2590.
Appl. Mater. Interfaces 2014, 6, 9897−9903. (92) Valdez, L.; Shum, H.; Ortiz-Rivera, I.; Balazs, A. C.; Sen, A. Soft
(58) Gao, W.; Feng, X.; Pei, A.; Gu, Y.; Li, J.; Wang, J. Nanoscale Matter 2017, 13, 2800−2807.
2013, 5, 4696−4700. (93) Zhou, C.; Zhang, H.; Li, Z.; Wang, W. Lab Chip 2016, 16,
(59) Manjare, M.; Yang, B.; Zhao, Y. P. J. Phys. Chem. C 2013, 117, 1797−1811.
4657−4665. (94) Ortiz-Rivera, I.; Courtney, T. M.; Sen, A. Adv. Funct. Mater.
(60) Soler, L.; Magdanz, V.; Fomin, V. M.; Sanchez, S.; Schmidt, O. 2016, 26, 2135−2142.
(95) Kasyap, T. V.; Koch, D. L.; Wu, M. Phys. Fluids 2014, 26,
G. ACS Nano 2013, 7, 9611−9620.
081901.
(61) Wang, H.; Moo, J. G. S.; Pumera, M. ACS Nano 2016, 10,
(96) Minõ, G. L.; Dunstan, J.; Rousselet, A.; Clément, E.; Soto, R. J.
5041−5050. Fluid Mech. 2013, 729, 423−444.
(62) Jiang, C.; Huang, G.; Ding, S.-J.; Dong, H.; Men, C.; Mei, Y. (97) Peng, Y.; Lai, L.; Tai, Y.-S.; Zhang, K.; Xu, X.; Cheng, X. Phys.
Nanoscale Res. Lett. 2016, 11, 289. Rev. Lett. 2016, 116, 068303.
(63) Zhao, G.; Sanchez, S.; Schmidt, O. G.; Pumera, M. Chem. (98) Leptos, K. C.; Guasto, J. S.; Gollub, J. P.; Pesci, A. I.; Goldstein,
Commun. 2012, 48, 10090−10092. R. E. Phys. Rev. Lett. 2009, 103, 198103.
(64) Zhao, G.; Pumera, M. Langmuir 2013, 29, 7411−7415. (99) Kurtuldu, H.; Guasto, J. S.; Johnson, K. A.; Gollub, J. P. Proc.
(65) Ma, X.; Hortelão, A. C.; Patiño, T.; Sánchez, S. ACS Nano 2016, Natl. Acad. Sci. U. S. A. 2011, 108, 10391−10395.
10, 9111−9122. (100) Minõ, G.; Mallouk, T. E.; Darnige, T.; Hoyos, M.; Dauchet, J.;
(66) Butler, P. J.; Dey, K. K.; Sen, A. Cell. Mol. Bioeng. 2015, 8, 106− Dunstan, J.; Soto, R.; Wang, Y.; Rousselet, A.; Clement, E. Phys. Rev.
118. Lett. 2011, 106, 048102.
(67) Muddana, H. S.; Sengupta, S.; Mallouk, T. E.; Sen, A.; Butler, P. (101) Orozco, J.; Jurado-Sánchez, B.; Wagner, G.; Gao, W.; Vazquez-
J. J. Am. Chem. Soc. 2010, 132, 2110−2111. Duhalt, R.; Sattayasamitsathit, S.; Galarnyk, M.; Cortés, A.; Saintillan,
(68) Sengupta, S.; Dey, K. K.; Muddana, H. S.; Tabouillot, T.; Ibele, D.; Wang, J. Langmuir 2014, 30, 5082−5087.
M. E.; Butler, P. J.; Sen, A. J. Am. Chem. Soc. 2013, 135, 1406−1414. (102) Kapral, R.; Mikhailov, A. S. Phys. D 2016, 318−319, 100−104.

7675 DOI: 10.1021/jacs.7b02347


J. Am. Chem. Soc. 2017, 139, 7666−7676
Journal of the American Chemical Society Perspective

(103) Dey, K. K.; Pong, F. Y.; Breffke, J.; Pavlick, R.; Hatzakis, E.;
Pacheco, C.; Sen, A. Angew. Chem., Int. Ed. 2016, 55, 1113−1117.
(104) Guo, M.; Ehrlicher, A. J.; Jensen, M. H.; Renz, M.; Moore, J.
R.; Goldman, R. D.; Lippincott-Schwartz, J.; Mackintosh, F. C.; Weitz,
D. A. Cell 2014, 158, 822−832.
(105) Brangwynne, C. P.; Koenderink, G. H.; MacKintosh, F. C.;
Weitz, D. A. J. Cell Biol. 2008, 183, 583−587.
(106) Hendricks, A. G.; Holzbaur, E. L. F.; Goldman, Y. E. Proc. Natl.
Acad. Sci. U. S. A. 2012, 109, 18447−18452.
(107) Rai, A. K.; Rai, A.; Ramaiya, A. J.; Jha, R.; Mallik, R. Cell 2013,
152, 172−182.
(108) Parry, B. R.; Surovtsev, I. V.; Cabeen, M. T.; O’Hern, C. S.;
Dufresne, E. R.; Jacobs-Wagner, C. Cell 2014, 156, 183−194.
(109) Lodish, H.; Berk, A.; Zipursky, S. L.; Matsudaira, P.; Baltimore,
D.; Darnell, J. Mol. Cell. Biol.; W. H. Freeman: New York, 2000.
(110) Ozin, G. A.; Manners, I.; Fourier-Bidoz, S.; Arsenault, A. Adv.
Mater. 2005, 17, 3011−3018.
(111) Wang, W.; Chiang, T.; Velegol, D.; Mallouk, T. E. J. Am. Chem.
Soc. 2013, 135, 10557−10565.
(112) Wong, F.; Dey, K. K.; Sen, A. Annu. Rev. Mater. Res. 2016, 46,
407−432.
(113) Li, J.; Li, T.; Xu, T.; Kiristi, M.; Liu, W.; Wu, Z.; Wang, J. Nano
Lett. 2015, 15, 4814−4821.
(114) Wang, W.; Duan, W.; Zhang, Z.; Sun, M.; Sen, A.; Mallouk, T.
E. Chem. Commun. 2015, 51, 1020−1023.
(115) Patra, D.; Sengupta, S.; Duan, W.; Zhang, H.; Pavlick, R.; Sen,
A. Nanoscale 2013, 5, 1273−1283.
(116) Guix, M.; Mayorga-Martinez, C. C.; Merkoçi, A. Chem. Rev.
2014, 114, 6285−6322.
(117) Gao, W.; Wang, J. Nanoscale 2014, 6, 10486−10494.
(118) Gáspár, S. Nanoscale 2014, 6, 7757−7763.
(119) Qin, W.-W.; Sun, L.-L.; Peng, T.-H.; Xu, Y.; Gao, Y.-J.; Wang,
W.-F.; Li, D. Chin. J. Anal. Chem. 2016, 44, 1133−1139.
(120) Nijemeisland, M.; Abdelmohsen, L. K. E. A.; Huck, W. T. S.;
Wilson, D. A.; van Hest, J. C. M. ACS Cent. Sci. 2016, 2, 843−849.
(121) Pickard, W. F. Plant Cell Environ. 2003, 26, 1−15.
(122) Wu, F.; Pelster, L. N.; Minteer, S. D. Chem. Commun. 2015, 51,
1244−1247.
(123) Wheeldon, I.; Minteer, S. D.; Banta, S.; Barton, S. C.;
Atanassov, P.; Sigman, M. Nat. Chem. 2016, 8, 299−309.
(124) An, S.; Kumar, R.; Sheets, E. D.; Benkovic, S. J. Science 2008,
320, 103−106.
(125) Graham, J. W. A.; Williams, T. C. R.; Morgan, M.; Fernie, A.
R.; Ratcliffe, R. G.; Sweetlove, L. J. Plant Cell 2007, 19, 3723−3738.
(126) Campanella, M. E.; Chu, H.; Low, P. S. Proc. Natl. Acad. Sci. U.
S. A. 2005, 102, 2402−2407.
(127) Sánchez, S.; Soler, L.; Katuri, J. Angew. Chem., Int. Ed. 2015, 54,
1414−1444.
(128) Ismagilov, R. F.; Schwartz, A.; Bowden, N.; Whitesides, G. M.
Angew. Chem., Int. Ed. 2002, 41, 652−654.
(129) Wang, J. ACS Nano 2009, 3, 4−9.

7676 DOI: 10.1021/jacs.7b02347


J. Am. Chem. Soc. 2017, 139, 7666−7676

You might also like