A Revolution in Optical Manipulation
A Revolution in Optical Manipulation
Optical tweezers use the forces exerted by a strongly focused beam of light to trap and move objects ranging in
size from tens of nanometres to tens of micrometres. Since their introduction in 1986, the optical tweezer has
become an important tool for research in the fields of biology, physical chemistry and soft condensed matter
physics. Recent advances promise to take optical tweezers out of the laboratory and into the mainstream of
manufacturing and diagnostics; they may even become consumer products. The next generation of single-beam
optical traps offers revolutionary new opportunities for fundamental and applied research.
A
new generation of techniques that use the
forces exerted by carefully sculpted wavefronts Optical
axis
of light offers precisely the level of access and Radiation
control needed for rapid progress at the pressure
dominates
frontiers of several branches of science and
engineering. In particular, optical forces are ideally suited Wave front
to manipulating mesoscopic systems, which are
characterized by length scales ranging from tens of λ
nanometres to hundreds of micrometres, forces ranging
from femtonewtons to nanonewtons and time scales
ranging upward from a microsecond. In biology, this Colloidal
range covers many of the inter- and intracellular processes particle
responsible for respiration, reproduction and signalling.
In physics and chemistry, it corresponds to the still- Gradient
force
puzzling interface between classical and quantum dominates
mechanical behaviour, which is made all the more
perplexing by the general inapplicability of statistical
many-body theory in this realm. Fulfilment of the
promise of mesoscopic engineering has been held back by
the need for tiny motors to drive micromachines and for
robust human-scale interfaces with atomic-scale
Laser beam
nanotechnology. Until quite recently, the options for
manipulating, analysing and organizing mesoscopically Figure 1 Optical tweezers use a strongly focused beam of light to trap
textured matter have been limited. The advent of flexible objects. Intensity gradients in the converging beam draw small objects,
multifunctional optical traps meets this need. such as a colloidal particle, toward the focus, whereas the radiation
Many of the most powerful optical manipulation tech- pressure of the beam tends to blow them down the optical axis. Under
niques are derived from single-beam optical traps known as conditions where the gradient force dominates, a particle can be trapped,
optical tweezers (see Fig. 1), which were introduced by in three dimensions, near the focal point.
Arthur Ashkin, Steven Chu and their coworkers at AT&T
Bell Laboratories1,2. An optical tweezer uses forces exerted
by a strongly focused beam of light to trap small objects. usually constructed around microscope objective lenses,
Although the theory behind optical tweezers is still being whose high numerical apertures and well corrected aberra-
developed, the basic principles are straightforward for tions focus light as tightly as possible.
objects either much smaller than the wavelength of light or Optical tweezers can trap objects as small as 5 nm
much larger. Small objects develop an electric dipole (refs 4,5) and can exert forces exceeding 100 pN (refs 6–8)
moment in response to the light’s electric field, which, gen- with resolutions as fine as 100 aN (refs 9–11). This is the
erally speaking, is drawn up intensity gradients in the elec- ideal range for exerting forces on biological and macromol-
tric field toward the focus. Larger objects act as lenses, ecular systems and for measuring their responses. Biologi-
refracting the rays of light and redirecting the momentum cal and medical applications of optical tweezers have been
of their photons. The resulting recoil draws them toward reviewed extensively2,12,13, and so just a few examples of their
the higher flux of photons near the focus3. This recoil is all uses will be outlined. Optical tweezers have been used to
but imperceptible for a macroscopic lens but can have a sub- probe the viscoelastic properties of single biopolymers
stantial influence on mesoscopic objects. (such as DNA), cell membranes, aggregated protein fibres
Optical gradient forces compete with radiation pressure (such as actin), gels of such fibres in the cytoskeleton, and
resulting from the momentum absorbed or otherwise composite structures (such as chromatin and chromo-
transferred from the photons in the beam, which acts like a somes). They have also been used to characterize the forces
fire hose to blow particles down the optical axis. Stable trap- exerted by molecular motors such as myosin, kinesin,
ping requires the axial gradient force to dominate, and is processive enzymes and ribosomes. These measurements
achieved when the beam diverges rapidly enough away have revealed that cells use mechanical forces not only
from the focal point. For this reason, optical tweezers are for mobility, motility and chromosome sorting during
810 © 2003 Nature Publishing Group NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature
insight review articles
reproduction but also for regulating gene transcription, inter- and
intracellular signalling and respiration. As a natural extension of Sample
these studies, optical tweezers offer great promise for intracellular Illuminator volume
surgery, for instance, in modifying the chromosomes of living cells14.
On a larger scale, optical tweezers are useful for selecting individual Diffraction
microbes from heterogeneous populations. In addition, their ability grating
to transport and modify cells precisely has led to clinical applications exp (i ϕ( ρ) )
Objective lens
in such areas as in vitro fertilization15. A
Telescope
Optically
trapped
In the physical sciences, the unique ability of optical tweezers to Input
particles
organize matter non-invasively has led to a burst of activity in the field of pupil
classical statistical mechanics, including the first direct measurements La Mirror
s
of macromolecular interactions in solution16. Each new round of mea- (TE er b Eyepiece
M eam
00 )
surements has led to surprises, including the discovery of anomalous
Video camera
attractions between like-charged colloidal particles17, oscillatory col-
loidal interactions mediated by the entropy of smaller entities in solu-
tion18–21 and hydrodynamic fluctuations that may be interpreted as 2π
ρ θ
transient violations of the second law of thermodynamics22.
In all of these cases, and a great many more, fundamental insights
emerged from manipulating specially chosen systems at just one or ϕ 10 µ m
two discrete points. New frontiers of science and engineering would
present themselves if optical traps could interrogate more general
and more complex systems at many points at once, if they could
induce chemical as well as physical transformations and if they could 0
exert torques as well as forces. Recent advances in physical optics Wavefront phase ϕ (ρ)
20 × 20 array
reveal that such multifunctional optical traps can be crafted from sin- optical traps
gle beams of light by subtly modifying their wavefronts. The resulting
optical micromanipulators provide unprecedented access to the Figure 2 Creation of a large number of optical tweezers by using a computer-
microscopic world. generated holograms. Projecting a collimated TEM00 laser beam through the input pupil
of a strongly converging lens such as a microscope objective creates a single optical
Manipulating the microscopic world tweezer. The telescope in this implementation creates an image of the objective’s input
Figure 2 schematically depicts an optical tweezer system in which a pupil, centred at point A. Multiple beams passing through point A therefore pass into
strongly converging objective lens focuses beams of laser light into the objective lens to create multiple optical traps. A single TEM00 laser beam can be
optical traps. A collimated TEM00 beam passing straight into the spilt into an arbitrary fan-out of beams all emanating from point A by an appropriate
input pupil of the lens comes to a focus in the middle of the focal plane computer-designed diffraction grating centred there. The example phase grating w(r)
of the objective lens, where it forms a trap. Sweeping the angle of inci- creates the 20 2 20 array of traps shown in the video micrograph. These are shown
dence translates the trap across the field of view. If the beam diverges, trapping 800 nm diameter polystyrene spheres dispersed in water. This figure was
it focuses downstream of the focal plane, whereas if it converges, it adapted with permission from ref. 31 © Elsevier Science Ltd. Bar, 10 mm.
focuses upstream.
Translating an optical trap, creating multiple optical traps and
converting these into multifunctional optical traps are greatly facili- and a polarization vector e describing the field’s orientation. A multi-
tated by first forming an image of the input pupil of the lens using the beam interference hologram generally would modify both the ampli-
telescope in Fig. 2. Any beam passing through the pupil’s image, tude and the phase of the input beam, with the amplitude modifica-
which is centred at point A in Fig. 2, also passes through the actual tions diverting power away from the optical traps. Fortunately, a vari-
pupil and forms a trap. Tilting the beam as it passes through the ety of iterative optimization algorithms have been developed29–31 to
image scans the optical trap. A single rapidly scanned optical tweezer create equivalent holographic beamsplitters that modify only the
can trap multiple particles by dwelling briefly on each one before phase of the input beam. Such a phase-only diffractive optical element
moving on to the next23,24. The extent and complexity of such multi- (DOE), also known as a kinoform, was used to create the 20 2 20
particle patterns are limited by the time required to reposition each of optical traps shown in Fig. 2.
the multiple wandering objects. Scanned optical tweezers, further- Holographic optical tweezers really hit their stride when a com-
more, are restricted to the focal plane of the lens. Even so, scanned puter-addressed spatial light modulator (SLM) was used to project
optical tweezers are extremely useful for organizing planar assem- sequences of trap-forming kinoforms in real time30–32. An SLM
blies of colloidal particles25, for testing new ideas in statistical imposes a prescribed amount of phase shift at each pixel in an array
mechanics26 and for measuring macromolecular interactions27. by varying the local optical path length. Typically, this is accom-
Placing a diffractive beamsplitter at the pupil’s image converts a plished by controlling the local orientation of molecules in a layer of
single input beam into several beams, each of which forms a separate liquid crystal, although arrays of microelectromechanical (MEMS)
optical trap. Such a beamsplitter can be a computer-generated holo- mirrors are also becoming available for SLM applications. Slightly
gram, and the resulting trapping patterns are known as holographic displacing the traps from one pattern to the next transfers particles
optical tweezers (HOTs)28,29. To see how this works, consider multiple along arbitrary three-dimensional (3D) trajectories30–32, animating
beams all passing simultaneously through point A on their way to matter with light in much the same way that cartoons animate light
being focused into optical traps. Their superposition creates a dis- with matter. Figure 3 shows this principle in action.
tinctive interference pattern centred at point A. Imprinting this pat- In a variation on this theme, the generalized phase contrast (GPC)
tern onto the wavefronts of a single input laser beam transforms the technique converts a pattern of phase modulation across an SLM’s
one beam into the desired fan-out of beams and thus forms the same face directly into the corresponding intensity modulation in the
pattern of optical traps. focal plane of the objective lens33 and thus creates arbitrary planar
The input beam’s electric field, E(r) exp(iw(r)) e, around point A trapping patterns. The conversion involves an annular phase plate
is characterized by a real-valued amplitude E(r) and phase w(r), similar to that used in phase contrast microscopy. This approach
both of which are functions of position transverse to the optical axis avoids the need to calculate holograms and thus is extremely efficient.
NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature © 2003 Nature Publishing Group 811
insight review articles
Figure 3 Polysterene and silica spheres in two- and three-dimensional
a b c
configurations of holographic optical tweezers created from a single laser
beam with a computer-designed hologram of a single beam’s wavefront.
a–c, Thirty-six water-borne polystyrene spheres, 800 nm in diameter, are
trapped in a plane and reconfigured with dynamic trapping patterns.
Reproduced with permission from ref. 31 © Elsevier Science Ltd. Bar, 10 mm.
d,e, Two 1 mm diameter silica spheres being moved past each other in two
different planes. Reproduced with permission from ref. 30 © Elsevier
10 µm
Science Ltd. The different appearance of the spheres results from their
different heights relative to the microscope’s focal plane. f–h, Seven 1 mm d f In focus
diameter silica spheres being moved up and down through seven different
planes. Reproduced with permission from ref. 31 © Elsevier Science Ltd.
–5 µm +5 µm
g
In focus
h
In focus
+5 µm –5 µm
The spatial resolution of existing SLMs currently limits the GPC potentials is a classic problem in statistical physics, and colloids in
technique to creating lateral traps rather than 3D optical tweezers, modulated optical fields constitute a rare model system where micro-
but GPC has still proved useful for rapidly organizing small objects scopic interactions can be measured and controlled while macro-
in thin samples34. scopic thermodynamic properties unfold37,38. Insights obtained from
Even static arrays of optical traps have exciting and surprising studying optically modulated colloids are relevant to analogous sys-
applications. For example, an array of traps can continuously sort tems such as atoms adsorbed on crystal surfaces, electrons passing
fluid-borne particles, acting much like a sieve. The array sorts particles through charge density waves and 2D electron gases, magnetic flux
on the basis of their different affinities for optical traps and for their quanta passing through defects in type II superconductors and
affinity for the externally applied driving force. Inclining a regular motor proteins translating along filaments in living cells. Early stud-
array with respect to the driving force deflects the selected fraction so ies demonstrated that modulation along even one direction can
that it can be collected separately from the other, undeflected freeze a 2D colloidal fluid39,40. Deeper modulation actually melts the
fraction35. Unlike most sorting techniques that operate on discrete substrate-induced crystal41 by suppressing inter-row coupling42.
batches of sample, optical fractionation works continuously and can More recent studies have demonstrated other intriguing behaviours,
be dynamically optimized by adjusting the wavelength, intensity and such as rotational melting in an array of multiply occupied traps38,43,
geometry of the trap array. Moreover, because optical fractionation and have shed new light on the mechanisms by which magnetic flux
relies on the object’s ability to hop from potential well to potential lines invade superconductors37,38.
well, it is exponentially sensitive to particle size and so promises Time-varying potential energy landscapes created with dynamic
unparalleled size resolution36. optical traps promise new insights into the operation of molecular
An array of traps may also be viewed as a tailor-made potential motors by providing a powerful experimental system within which to
energy landscape for interacting colloidal particles. Determining study thermal ratchets26 and related models in non-equilibrium sta-
how strongly interacting systems evolve on modulated substrate tistical mechanics44. Once perfected, such ratchet potentials will also
812 © 2003 Nature Publishing Group NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature
insight review articles
a b
a Helical phase profile
TEM00
Laser beam Helical mode
0 ϕ 2π
b c
be useful for dynamically sorting mesoscopic objects and transport- Figure 6 Optical vortices and optical spanners created from helical modes of light.
ing them through tiny integrated laboratories for processing and a, The helical phase profile w(r)4,u converts a TEM00 laser beam into a helical
testing45. mode whose wavefronts resemble an ,-fold corkscrew. b, Rather than focusing to a
All such studies provide valuable insights into how nature creates point, a helical mode focuses to an optical vortex whose radius R, is proportional to
and exploits hierarchically organized structures. Once these principles its pitch, ,. c, A single colloidal particle trapped in the optical vortex travels around
are understood, they will be extraordinarily useful for creating new its circumference, driven by the orbital angular momentum of the helical beam. This
materials and devices to order. Until then, many of the most interest- multiple exposure shows 11 stages, at 1/6 s intervals, in one 800 nm particle’s
ing three-dimensionally structured functional systems can be transit. Reproduced with permission from ref. 67 © American Physical Society.
assembled using many optical tweezers operating in concert. Indeed,
optical tweezers have an essentially unique ability to construct 3D
heterostructures with features ranging in size from a few nanometres Beyond creating structures de novo, spatially resolved photo-
to a few millimetres. The real power of this approach only becomes chemistry can be used to modify pre-existing structures. The 3D
apparent when optical trapping is combined with other techniques to optical waveguide structure in Fig. 4e demonstrates this principle.
create permanent structures with embedded functionality. Here, a self-assembled crystal of colloidal silica spheres was perfused
with a photosensitive precursor and selectively patterned with an
Nanofabrication with optical tweezers optical tweezer to create the embedded polymer structure shown in
Once assembled, tweezer-organized structures can be fixed in place, Fig. 4e (ref. 52). Filling the gaps with a high-index material and then
for instance, by sintering or gelling. The tweezers themselves can be dissolving away the spheres and polymer would leave a tweezer-
used in this process. In particular, the intense illumination at an drawn waveguide pattern embedded in the otherwise self-assembled
optical tweezer’s focus is ideal for driving photochemical reactions in photonic crystal53. This hybrid approach to creating hierarchically
restricted volumes. If the reaction rate depends strongly on intensity, structured materials could sweep away many of the practical hurdles
the resulting spatially resolved photochemistry can yield features that have prevented self-assembled systems from making bigger
smaller than the wavelength of light. inroads in photonics and electro-optical systems54.
The first such application of optical tweezers involved the spatially Using many optical tweezers to simultaneously organize prefab-
resolved photo-oxidation of biological materials such as chromo- ricated nanometre-scale parts and to stitch them together with spa-
somes46,47. Essentially, a scalpel was created from light. Optical scalpels tially resolved photochemistry would yield a whole new category of
and scissors have been used for surgery on living cells15,48 as well as for hierarchically structured materials and devices. Such scale-span-
ablating subwavelength structures into microscopic substrates49. ning heterostructures would provide the building blocks for sen-
Spatially resolved photochemistry using optical tweezers has been sors, photonic devices and a host of other technologies. Hierarchi-
used to fabricate small complex 3D structures such as the examples cally structured micromechanical systems hold similar promise for
shown in Fig. 4. Figure 4a–c shows 3D plastic structures created by optomechanical and microfluidic applications. In this case, optical
multiphoton photopolymerization in scanned optical tweezers. The trapping also solves the outstanding problem of actuating such
smallest features in Fig. 4a are about 100 nm across. The tiny turbine small devices. Some aspects of this solution involve the unusual and
in Fig. 4b, c not only was created in this way but also was trapped and counterintuitive properties of traps created with newly discovered
spun on its axis with an optical tweezer 50. Arrays of interlocking tur- modes of light.
bines and gears assembled and driven by light have already been
demonstrated50. Other photochemical transformations provide Optical actuators
opportunities for optical tweezers to create 3D electronic and pho- Using conventional optical tweezers as actuators for microma-
tonic structures. The fine lines of MoS2 in Fig. 4d were patterned on chines is likely to speed up the adoption of lab-on-a-chip and relat-
glass by photoreduction of aqueous salts, and similar results have ed technologies for medical diagnostics, environmental testing and
been obtained in silver, gold and oxidized copper51. point-of-use microfabrication. Dynamic optical tweezers can both
NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature © 2003 Nature Publishing Group 813
insight review articles
organize and drive devices, such as the micrometre-scale hydraulic
pump shown in Fig. 5a. The microscopic valve flap in Fig. 5b is an a b c
example of a photopolymerized colloidal heterostructure that is
assembled and actuated with optical tweezers55.
Modifying the wavefronts of the optical tweezers transforms them
into whole new classes of optical traps, some of which have already
found applications as actuators for unconventional micromachines.
Some of the most useful of these are based on exotic modes of light
whose properties have only recently been elucidated.
Figure 6a shows how the deceptively simple phase profile Figure 7 Generalizations of the optical vortex principle. a, The holographic optical
w(r)=,u transforms the parallel wavefronts of a TEM00 laser mode tweezer technique can create arrays of optical vortices, each with individually specified
into the corkscrew topology of a helical mode56. Here, u is the intensity and topological charge. This 323 array of ,=30 optical matrices is shown
azimuthal angle around the optical axis and , is an integer winding trapping 800 nm diameter polystyrene spheres. Reproduced with permission from
number (also known as the topological charge). The modified beam ref. 31 © Elsevier Science Ltd. b, Trapped objects can be rotated within the
no longer focuses to a point, because the helical topology fosters interference pattern of an optical vortex and a plane wave. Such an optical rotator was
destructive interference along the optical axis. Instead, it converges to created by interference of an ,=3 helical mode with a plane wave. Reproduced with
a ring of light, as shown in Fig. 6b. The dark focus is suitable for trap- permission from ref. 75 © American Association for the Advancement of Science. c,
ping reflecting57, absorbing58 or low-dielectric-constant59,60 objects An optical rotator created with a five-fold modulation of the helicity of an ,=60 optical
that would be damaged or repelled by conventional optical tweezers. vortex. Reproduced with permission from ref. 67 © American Physical Society.
Because such traps lack the radiation pressure from axial rays, they
make more efficient traps for large dielectric objects than conven-
tional optical tweezers do61–63. Smaller dielectric particles are drawn orientated intensity pattern that is useful for orientating asymmetric
to the ring’s circumference, as shown in Fig. 6c. objects75. Superposing instead a helical mode on its mirror-image
What really distinguishes these ring-like optical traps is their ability counterpart creates 3D arrays of discrete traps that can be rotated
to exert torques as well as forces57,58,64. Just a decade ago, Allen demon- arbitrarily in three dimensions by varying the beams’ relative phase76.
strated that each photon in a helical mode carries an orbital angular Modulating the helical pitch of an optical vortex results in another
momentum ,ù, where ù is the quantum unit of angular momentum, class of optical rotators77, an example of which appears in Fig. 7c. Fur-
in addition to its intrinsic spin angular momentum56. This orbital ther generalizations create intensity patterns related to the caustics
angular momentum takes the form of a tangential component to the seen at the bottom of swimming pools and which can move objects
beam’s linear momentum density that can be transferred to illuminat- along complex trajectories transverse to the optical axis, all with static
ed objects65–67. A single colloidal microsphere is shown circulating holograms and no moving parts. Other superpositions can focus on
around such a topological ring-trap under the influence of the optical micrometre-scale dark regions surrounded by light on all sides
angular momentum flux in the time-lapse photograph of Fig. 6c. Such known as optical bottles78. These are useful for trapping very small
toroidal torque-exerting traps have come to be known as optical vor- dark-seeking objects, including clouds of ultracold atoms78. Holo-
tices68 or optical spanners69, and they have potentially widespread graphic arrays of optical bottles therefore should be useful for manip-
technological applications. Studying the motion of objects in optical ulating atoms79, perhaps for quantum computing applications, and
vortices has also provided valuable insights into the interplay of pho- will help to extend pioneering efforts to apply optical tweezers to
ton spin and orbital angular momentum57,64–67,70, which have been use- atomic physics80,81.
ful in elucidating the quantum mechanical nature of helical beams. Whereas azimuthal phase modulations extend optical tweezers
An optical vortex’s radius, R,, increases with topological into micromanipulators that are transverse to the optical axis, radial
charge67,71; therefore, the intensity pattern can be tailored to different modulations create axial devices with an intriguing twist. The sim-
applications. For example, a properly scaled ring of light can be pro- plest non-trivial radial phase profile modification, w(r)4gr, trans-
jected onto the teeth of a microfabricated gear, thereby creating a reli- forms a TEM00 beam into an approximation of a Bessel mode, a beam
able micrometre-scale motor. The distributed drive made possible by that propagates without diffracting even when focused to a wave-
projecting multiple optical vortices should also help to alleviate length-scale cross-section. The associated optical trap can extend for
problems associated with friction in micromechanical systems. The millimetres along the optical axis, as demonstrated in Fig. 8, and can
spin angular momentum carried by circularly polarized optical precisely push particles over very large distances83. The extended
tweezers similarly has been used to apply torques to birefringent range of Bessel-beam arrays should increase the throughput of optical
components72–74. Using an SLM to control the polarization of multi- fractionation by orders of magnitude. Still more remarkably, Bessel
ple optical tweezers opens up even more possibilities for developing beams are impervious to distortions by intervening particles and sur-
extensive micromachines assembled and driven by light. faces84—they can reconstruct their wavefronts as they propagate
Some micromechanical applications may require no microfabri- away from disturbances. Combining the robustness of Bessel beams
cation at all. Rapidly circulating particles entrain flows that can mix with the orbital angular momentum of helical modes yields optical
and pump extremely small volumes of fluid. This solves a problem in devices that can reach deeply into complex systems to apply forces
microfluidic systems, whose laminar flows are ideal for transporting and torques where needed.
minuscule quantities of reagents but do not promote mixing when
needed . Furthermore, the holographic optical tweezer technique can Future prospects
project multiple optical vortices, such as the 3 2 3 array in Fig. 7a, The ability to reach into the microscopic world dextrously and non-
each with an individually specified intensity and topological invasively at many points at once, to cut, assemble and transform with
charge31. Cooperative flows in such arrays can be reconfigured nanometre precision and submicrometre resolution and to do all these
dynamically by modifying the trap-forming hologram, opening up things with a single instrument promises revolutionary advances in
the possibility of developing adaptive microfluidics on length scales many disciplines. The sections above highlight just a few of these
ranging all the way down to tens of nanometres. advances. In particular, wavefront engineering provides a straightfor-
Other variations on this theme yield a family of distinct optical ward means to create large many optical traps in arbitrary 3D configu-
micromanipulators, each with its own applications. For example, rations, to move them freely and independently in three dimensions
superposing a helical beam on a conventional beam not only shows and to transform them into optical vortices, optical bottles, Bessel
the helical wavefronts’ structure, as in Fig. 7b, but also creates an traps and a host of other all-optical tools.
814 © 2003 Nature Publishing Group NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature
insight review articles
8. Litvinov, R. I., Shuman, H., Bennett, J. S. & Weisel, J. W. Binding strength and activation state of single
fibrinogen-integrin pairs on living cells. Proc. Natl Acad. Sci. 99, 7426–7431 (2002).
9. Gittes, F. & Schmidt, C. F. Signals and noise in micromechanical measurements. Methods Cell Biol. 55,
129–156 (1998).
10. Gittes, F. & Schmidt, C. F. Interference model for back-focal-plane displacement detection in optical
tweezers. Opt. Lett. 23, 7–9 (1998).
11. Pralle, A., Prummer, M., Florin, E. L., Stelzer, E. H. K. & Horber, J. K. H. Three-dimensional high-
II resolution particle tracking for optical tweezers by forward scattered light. Microsc. Res. Tech. 44,
378–386 (1999).
12. Svoboda, K., Mitra, P. P. & Block, S. M. Fluctuation analysis of motor protein movement and single
enzyme-kinetics. Proc. Natl Acad. Sci. 91, 11782–11786 (1994).
13. Bustamante, C., Smith, S. B., Liphardt, J. & Smith, D. Single-molecule studies of DNA mechanics.
~3 mm Curr. Opin. Struct. Biol. 10, 279–285 (2000).
14. Berns, M. W., Tadir, Y., Liang, H. & Tromberg, B. Laser scissors and tweezers. Methods Cell Biol. 55,
71–98 (1998).
15. Wright, G., Tucker, M. J., Morton, P. C., Sweitzer-Yoder, C. L. & Smith, S. E. Micromanipulation in
assisted reproduction: A review of current technology. Curr. Opin. Obstet. Gyn. 10, 221–226 (1998).
16. Crocker, J. C. & Grier, D. G. Microscopic measurement of the pair interaction potential of charge-
I stabilized colloid. Phys. Rev. Lett. 73, 352–355 (1994).
17. Crocker, J. C. & Grier, D. G. When like charges attract: The effects of geometrical confinement on
long-range colloidal interactions. Phys. Rev. Lett. 77, 1897–1900 (1996).
18. Ohshima, Y. N. Direct measurement of infinitesimal depletion force in a colloid-polymer mixture by
laser radiation pressure. Phys. Rev. Lett. 78, 3963–3966 (1997).
19. Crocker, J. C., Matteo, J. A., Dinsmore, A. D. & Yodh, A. G. Entropic attraction and repulsion in
Figure 8 The radial phase profile w(r)4gr creates a diffractionless Bessel beam binary colloids probed with a line optical tweezer. Phys. Rev. Lett. 82, 4352–4355 (1999).
20. Verma, R., Crocker, J. C., Lubensky, T. C. & Yodh, A. G. Attractions between hard colloidal spheres in
that focuses to a long axial trap that can extend for millimetres. Here, the same beam semiflexible polymer solutions. Macromolecules 33, 177–186 (2000).
is shown trapping multiple colloidal particles in two separate sample chambers 21. Yodh, A. G. et al. Entropically driven self-assembly and interaction in suspension. Phil. Trans. R. Soc.
separated by a distance of 3 mm. Bessel beams are impervious to distortions by A 359, 921–937 (2001).
22. Wang, G. M., Sevick, E. M., Mittag, E., Searles, D. J. & Evans, D. J. Experimental demonstration of
intervening particles because wavefronts are reconstructed as they propagate away
violations of the second law of thermodynamics for small systems and short time scales. Phys. Rev.
from disturbances. Left, a sphere is trapped between the central spot and the first ring Lett. 89, 050601 (2002).
of the Bessel beam in cell 1. A little way above, the beam is distorted, but further 23. Sasaki, K., Koshio, M., Misawa, H., Kitamura, N. & Masuhara, H. Pattern formation and flow control
above, the beam is no longer distorted. The process is repeated for the same beam in of fine particles by laser-scanning micromanipulation. Opt. Lett. 16, 1463–1465 (1991).
24. Sasaki, K., Fujiwara, H. & Masuhara, H. Optical manipulation of a lasing microparticle and its
cell II in the planes on the right. Reproduced from ref. 83 © Macmillan Magazines Ltd. application to near-field microspectroscopy. J. Vacuum Sci. Tech. B 15, 2786–2790 (1997).
25. Mio, C., Gong, T., Terray, A. & Marr, D. W. M. Morphological control of mesoscale colloidal models.
Fluid Phase Equilibria 185, 157–163 (2001).
As tools for biology, multifunctional optical traps will facilitate 26. Faucheux, L. P., Bourdieu, L. S., Kaplan, P. D. & Libchaber, A. J. Optical thermal ratchet. Phys. Rev.
Lett. 74, 1504–1507 (1995).
new approaches to cell sorting, macromolecular purification, intra- 27. Verma, R., Crocker, J. C., Lubensky, T. C. & Yodh, A. G. Entropic colloidal interactions in
cellular surgery, embryonic testing and highly parallel drug screening concentrated DNA solutions. Phys. Rev. Lett. 81, 4004–4007 (1998).
as well as great many other possibilities. The same tools have immediate 28. Dufresne, E. R. & Grier, D. G. Optical tweezer arrays and optical substrates created with diffractive
optics. Rev. Sci. Instr. 69, 1974–1977 (1998).
applications for organizing mesoscopic matter into heterogeneous,
29. Dufresne, E. R., Spalding, G. C., Dearing, M. T., Sheets, S. A. & Grier, D. G. Computer-generated
hierarchically structured 3D functional systems, such as photonic cir- holographic optical tweezer arrays. Rev. Sci. Instr. 72, 1810–1816 (2001).
cuit elements, integrated sensor arrays and high-density data storage 30. Liesener, J., Reicherter, M., Haist, T. & Tiziani, H. J. Multi-functional optical tweezers using
devices. Combining this organizational capability with optical- computer-generated holograms. Opt. Commun. 185, 77–82 (2000).
31. Curtis, J. E., Koss, B. A. & Grier, D. G. Dynamic holographic optical tweezers. Opt. Commun. 207,
tweezer-based spatially resolved photochemistry suggests bright 169–175 (2002).
prospects for optically assembling new materials and devices with 32. Reicherter, M., Haist, T., Wagemann, E. U. & Tiziani, H. J. Optical particle trapping with computer-
features ranging in size from nanometres to millimetres and beyond. generated holograms written on a liquid-crystal display. Opt. Lett. 24, 608–610 (1999).
In micromechanics and microfluidics, appropriately sculpted 33. Mogensen, P. C. & Glückstad, J. Dynamic array generation and pattern formation for optical
tweezers. Opt. Comm. 175, 75–81 (2000).
wavefronts of light can easily control motions and flows on a length 34. Rodrigo, P. J., Eriksen, R. L., Daria, V. R. & Glückstad, J. Interactive light-driven and parallel
scale that has challenged other technologies. In so doing, optical manipulation of inhomogeneous particles. Opt. Exp. 10, 1550–1556 (2002).
micromachines should hasten the adoption of lab-on-a-chip devices 35. Korda, P. T., Taylor, M. B. & Grier, D. G. Kinetically locked-in colloidal transport in an array of optical
tweezers. Phys. Rev. Lett. 89, 128301 (2002).
for diagnostics, sensing, testing, pathology and drug discovery. The
36. Ladavac, K., Kasza, K. & Grier, D. G. Optical fractionation. Phys. Rev. Lett. (in the press).
same wavefront-shaping techniques can sort and purify materials in 37. Korda, P. T., Spalding, G. C. & Grier, D. G. Evolution of a colloidal critical state in an optical pinning
these tiny flows and direct them towards further stages of purification potential. Phys. Rev. B 66, 024504 (2002).
and analysis. Optical testing and manufacturing, thus, could be high- 38. Mangold, K., Leiderer, P. & Bechinger, C. Phase transitions of colloidal monolayers in periodic
pinning arrays. Phys. Rev. Lett. 90, 158302 (2003).
ly integrated, with a single instrument providing flow, sorting, orga- 39. Chowdhury, A., Ackerson, B. J. & Clark, N. A. Laser-induced freezing. Phys. Rev. Lett. 55, 833–836
nization, synthesis and assembly. In all of these areas, the emerging (1985).
generation of optical manipulation tools should help to bridge the 40. Loudiyi, K. & Ackerson, B. J. Direct observation of laser-induced freezing. Physica A 184, 1–25
chasm between our macroscopic world and applications based on the (1992).
41. Wei, X. et al. T cell activation studied by optical trap. Biophys. J. 74, A378 (1998).
physics, chemistry and biology of microscopic systems. ■ 42. Bechinger, C., Brunner, M. & Leiderer, P. Phase behavior of two-dimensional colloidal systems in the
doi:10.1038/nature01935 presence of periodic light fields. Phys. Rev. Lett. 86, 930–933 (2001).
43. Brunner, M. & Bechinger, C. Phase behavior of colloidal molecular crystals on triangular light
1. Ashkin, A., Dziedzic, J. M., Bjorkholm, J. E. & Chu, S. Observation of a single-beam gradient force lattices. Phys. Rev. Lett. 88, 248302 (2002).
optical trap for dielectric particles. Opt. Lett. 11, 288–290 (1986). 44. Reimann, P. Brownian motors: Noisy transport far from equilibrium. Phys. Rep. 361, 57–265 (2002).
2. Ashkin, A. History of optical trapping and manipulation of small-neutral particle, atoms, and 45. Koss, B. A. & Grier, D. G. Optical peristalsis. Appl. Phys. Lett. 82, 3985–3987 (2003).
molecules. IEEE J. Sel. Top. Quantum Elec. 6, 841–856 (2000). 46. Ashkin, A., Dziedzic, J. M. & Yamane, T. Optical trapping and manipulation of single cells using
3. Ashkin, A. Forces of a single-beam gradient laser trap on a dielectric sphere in the ray optics regime. infrared-laser beams. Nature 330, 608–609 (1987).
Methods Cell Biol. 55, 1–27 (1998). 47. Liang, H., Wright, W. H., Cheng, S., He, W. & Berns, M. W. Micromanipulation of chromosomes in
4. Svoboda, K. & Block, S. M. Optical trapping of metallic Rayleigh particles. Opt. Lett. 19, 930–932 PtK2 cells using laser microsurgery (optical scalpel) in combination with laser-induced optical force
(1994). (optical tweezers). Exp. Cell Res. 204, 110–120 (1992).
5. Ke, P. C. & Gu, M. Characterization of trapping force on metallic Mie particles. Appl. Opt. 38, 160–167 48. Bayles, C. J., Aist, J. R. & Berns, M. W. The mechanics of anaphase-B in A basidomycete as revealed by
(1999). laser microbeam microsurgery. Exp. Mycology 17, 191–199 (1993).
6. Ghislain, L. P., Switz, N. A. & Webb, W. W. Measurement of small forces using an optical trap. Rev. Sci. 49. Fuhr, G. et al. Processing of micro-particles by UV laser irradiation in a field cage. Appl. Phys. A 69,
Instr. 65, 2762–2768 (1994). 611–616 (1999).
7. Rohrbach, A. & Stelzer, E. H. K. Trapping forces, force constants, and potential depths for dielectric 50. Galajda, P. & Ormos, P. Complex micromachines produced and driven by light. Appl. Phys. Lett. 78,
spheres in the presence of spherical aberrations. Appl. Opt. 41, 2494–2507 (2002). 249–251 (2001).
NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature © 2003 Nature Publishing Group 815
insight review articles
51. Lachish-Zalait, A., Zbaida, D., Klein, E. & Elbaum, M. Direct surface patterning from solutions: 69. Simpson, N. B., Allen, L. & Padgett, M. J. Optical tweezers and optical spanners with Laguerre-
Localized microchemistry using a focused laser. Adv. Funct. Mat. 11, 218–223 (2001). Gaussian modes. J. Mod. Opt. 43, 2485–2491 (1996).
52. Lee, W. M., Pruzinsky, S. A. & Braun, P. V. Multi-photon polymerization of waveguide structures 70. Simpson, N. B., Dholakia, K., Allen, L. & Padgett, M. J. Mechanical equivalence of spin and orbital
within three-dimensional photonic crystals. Adv. Mat. 14, 271–274 (2002). angular momentum of light: An optical spanner. Opt. Lett. 22, 52–54 (1997).
53. Joannopoulos, J. D., Meade, R. D. & Winn, J. N. Photonic Crystals (Princeton Univ., Princeton, 1995). 71. Padgett, M. J. & Allen, L. The Poynting vector in Laguerre-Gaussian modes. Opt. Commun. 121,
54. Taton, T. A. & Norris, D. J. Device physics: Defective promise in photonics. Nature 416, 685–686 36–40 (1995).
(2002). 72. Higurashi, E., Sawada, R. & Ito, T. Optically induced angular alignment of trapped birefringent
55. Terray, A., Oakey, J. & Marr, D. W. M. Fabrication of linear colloidal structures for microfluidic micro-objects by linearly polarized light. Phys. Rev. E 59, 3676–3681 (1999).
applications. Appl. Phys. Lett. 81, 1555–1557 (2002). 73. Higurashi, E., Sawada, R. & Ito, T. Optically driven angular alignment of microcomponents made of
56. Allen, L., Beijersbergen, M. W., Spreeuw, R. J. C. & Woerdman, J. P. Orbital angular-momentum of in-plane birefringent polyimide film based on optical angular momentum transfer. J. Micromech.
light and the transformation of Laguerre-Gaussian laser modes. Phys. Rev. A 45, 8185–8189 (1992). Microeng. 11, 140–145 (2001).
57. O’Neil, A. T. & Padgett, M. J. Three-dimensional optical confinement of micron-sized metal particles 74. Friese, M. E. J., Rubinsztein-Dunlop, H., Gold, J., Hagberg, P. & Hanstorp, D. Optically driven
and the decoupling of the spin and orbital angular momentum within an optical spanner. Opt. micromachine elements. Appl. Phys. Lett. 78, 547–549 (2001).
Commun. 185, 139–143 (2000). 75. Paterson, L. et al. Controlled rotation of optically trapped microscopic particles. Science 292, 912–914
58. Rubinsztein-Dunlop, H., Nieminen, T. A., Friese, M. E. J. & Heckenberg, N. R. Optical trapping of (2001).
absorbing particles. Adv. Quantum Chem. 30, 469–492 (1998). 76. MacDonald, M. P. et al. Creation and manipulation of three-dimensional optically trapped
59. Gahagan, K. T. & Swartzlander, G. A. Trapping of low-index microparticles in an optical vortex. J. structures. Science 296, 1101–1103 (2002).
Opt. Soc. Am. B 15, 524–534 (1998). 77. Curtis, J. E. & Grier, D. G. Modulated optical vortices. Opt. Lett. 28, 872–874 (2003).
60. Gahagan, K. T. & Swartzlander, G. A. Simultaneous trapping of low-index and high-index 78. Arlt, J. & Padgett, M. J. Generation of a beam with a dark focus surrounded by regions of higher
microparticles observed with an optical-vortex trap. J. Opt. Soc. Am. B 16, 533–537 (1999). intensity: The optical bottle beam. Opt. Lett. 25, 191–193 (2000).
61. Ashkin, A. Forces of a single-beam gradient laser trap on a dielectric sphere in the ray optics regime. 79. McGloin, D., Spalding, G. C., Melville, H., Sibbet, W. & Dholakia, K. Applications of spatial light
Biophys. J. 61, 569–582 (1992). modulators in atom optics. Opt. Exp. 11, 158–166 (2003).
62. Simpson, N. B., McGloin, D., Dholakia, K., Allen, L. & Padgett, M. J. Optical tweezers with increased 80. Gustavson, T. L. et al. Transport of Bose-Einstein condensates with optical tweezers. Phys. Rev. Lett.
axial trapping efficiency. J. Mod. Opt. 45, 1943–1949 (1998). 88, 020401 (2002).
63. O’Neil, A. T. & Padgett, M. J. Axial and lateral trapping efficiency of Laguerre-Gaussian modes in 81. Chikkatur, A. P. et al. A continuous source of Bose-Einstein condensed atoms. Science 296, 2193–2195
inverted optical tweezers. Opt. Commun. 193, 45–50 (2001). (2002).
64. He, H., Friese, M. E. J., Heckenberg, N. R. & Rubinsztein-Dunlop, H. Direct observation of transfer of 82. Arlt, J., Garces-Chavez, V., Sibbett, W. & Dholakia, K. Optical micromanipulation using a Bessel light
angular momentum to absorptive particles from a laser beam with a phase singularity. Phys. Rev. Lett. beam. Opt. Commun. 197, 239–245 (2001).
75, 826–829 (1995). 83. Garces-Chavez, V., McGloin, D., Melville, H., Sibbett, W. & Dholakia, K. Simultaneous
65. Allen, L., Padgett, M. J. & Babiker, M. The orbital angular momentum of light. Prog. Opt. 39, 291–372 micromanipulation in multiple planes using a self-reconstructing light beam. Nature 419, 145–147
(1999). (2002).
66. O’Neil, A. T., MacVicar, I., Allen, L. & Padgett, M. J. Intrinsic and extrinsic nature of the orbital 84. Kawata, S., Sun, H.-B., Tanaka, T. & Takada, K. Finer features for functional microdevicecs. Nature
angular momentum of a light beam. Phys. Rev. Lett. 88, 053601 (2002). 412, 697–698 (2001).
67. Curtis, J. E. & Grier, D. G. Structure of optical vortices. Phys. Rev. Lett. 90, 133901 (2003). 85. Terray, A., Oakey, J. & Marr, D. W. M. Microfluidic control using colloidal devices. Science 296,
68. Gahagan, K. T. & Swartzlander, G. A. Optical vortex trapping of particles. Opt. Lett. 21, 827–829 (1996). 1841–1844 (2002).
816 © 2003 Nature Publishing Group NATURE | VOL 424 | 14 AUGUST 2003 | www.nature.com/nature