scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Optical trapping and manipulation of nanostructures

01 Nov 2013-Nature Nanotechnology (Nature Publishing Group)-Vol. 8, Iss: 11, pp 807-819
TL;DR: The state-of-the-art in optical trapping at the nanoscale is reviewed, with an emphasis on some of the most promising advances, such as controlled manipulation and assembly of individual and multiple nanostructures, force measurement with femtonewton resolution, and biosensors.
Abstract: Optical trapping and manipulation of micrometre-sized particles was first reported in 1970. Since then, it has been successfully implemented in two size ranges: the subnanometre scale, where light-matter mechanical coupling enables cooling of atoms, ions and molecules, and the micrometre scale, where the momentum transfer resulting from light scattering allows manipulation of microscopic objects such as cells. But it has been difficult to apply these techniques to the intermediate - nanoscale - range that includes structures such as quantum dots, nanowires, nanotubes, graphene and two-dimensional crystals, all of crucial importance for nanomaterials-based applications. Recently, however, several new approaches have been developed and demonstrated for trapping plasmonic nanoparticles, semiconductor nanowires and carbon nanostructures. Here we review the state-of-the-art in optical trapping at the nanoscale, with an emphasis on some of the most promising advances, such as controlled manipulation and assembly of individual and multiple nanostructures, force measurement with femtonewton resolution, and biosensors.

Summary (1 min read)

Introduction

  • When the light was focused on a mirror attached to the balance, the radiation pressure moved the balance from its equilibrium position2,3.
  • These pioneering works have developed into two very successful research lines.
  • On one hand, early techniques for laser cooling of atoms7–10 paved the way to modern ultracold atom technology11.
  • But optical forces acting between ~1 and 100  nm, a range of primary interest for nanotechnology (Fig.  1), have not been widely exploited because of the challenges in scaling up the techniques optimized for atom cooling, or scaling down those used for microparticle trapping.

Optical forces on nanostructures

  • The authors review how optical forces arise.
  • All rights reserved NATURE NANOTECHNOLOGY | VOL 8 | NOVEMBER 2013 | www.nature.com/naturenanotechnology 809 symmetric in shape and composition, one can treat non-spherical objects by modelling them as clusters of small spheres29,41,89.
  • The force on each dipole is due to the incident field and the fields scattered by all other dipoles (equation (2)).
  • The plasmonic nature95 of MNPs can enhance the optical forces, so that stable trapping can be achieved at a much lower power (~2–3 mW; refs 26,29,71).
  • Temperature increases of hundreds of kelvin were observed by trapping AuNPs adjacent to fluorophore-containing lipid vesicles with permeability sensitive to temperature97.

Experimental designs and techniques

  • The authors review the most common experimental implementations relevant for optical tweezers, with an emphasis on those used for trapping and manipulation of NPs and nanostructures.
  • In the simplest configuration (Fig.  2a), optical tweezers can be generated by focusing a laser beam to a diffraction-limited spot using a high numerical aperture (NA) objective lens4,115,116.
  • It is also possible to use optical tweezers to trap and position single NPs: for example to manipulate, assemble and fuse different semiconductor nanowires43 (Fig. 3a).
  • A PFM can benefit from the ability to fabricate SERS-active nanoprobes, paving the way for local enhanced spectroscopy of biological surfaces.
  • Light scattered by the particle was monitored with photodiodes to infer the particle motion, then used in a feedback loop that modified the trapping light intensity76.

Perspective

  • In the context of the ongoing trend towards miniaturization of technology towards the nanoscale, optical trapping and manipulation of nanostructures can open new and exciting possibilities for assembly, characterization and optical control of nanodevices and biomolecules.
  • Optomechanical cooling schemes have already made progress towards the demonstration of quantized nanoparticle motional states76.
  • The realization of these goals will require the development of new techniques to manipulate nanoparticles beyond those currently available.
  • The ability to probe and control what happens in the trap is still missing.

Additional information

  • Reprints and permissions information is available online at www.nature.com/reprints.
  • Correspondence should be addressed to O.M.M. and A.C.F.

Did you find this useful? Give us your feedback

Figures (6)

Content maybe subject to copyright    Report

NATURE NANOTECHNOLOGY | VOL 8 | NOVEMBER 2013 | www.nature.com/naturenanotechnology 807
L
ight can exert a force on matter by means of momentum
exchange on scattering
1
. e existence of this force was rst
experimentally demonstrated by Lebedev
2
and Nichols and
Hull
3
in 1901 using thermal light sources (electric or arc lamps) and
a torsion balance. When the light was focused on a mirror attached
to the balance, the radiation pressure moved the balance from its
equilibrium position
2,3
. But the magnitude of these eects was con-
sidered insignicant for any practical use: to quote J.H. Poynting’s
presidential address to the British Physical Society in 1905 (reported
in ref.4), “A very short experience in attempting to measure these
forces is sucient to make one realize their extreme minuteness —
a minuteness which appears to put them beyond consideration in
terrestrial aairs.” It was not until 1970, and because of the advent
of lasers, that Arthur Ashkin showed that the use of optical forces to
alter the motion of micrometre-sized particles
5
and neutral atoms
6
could have applications in the manipulation of microscopic parti-
cles and of single atoms
4
.
ese pioneering works have developed into two very successful
research lines. On one hand, early techniques for laser cooling of
atoms
7–10
paved the way to modern ultracold atom technology
11
. On
the other hand, what is now commonly referred to as optical twee-
zers — that is, a tightly focused laser beam capable of conning par-
ticles in three dimensions
12
— has become a common tool for the
manipulation of micrometre-sized particles
13,14
and as a highly sen-
sitive force transducer
15
. But optical forces acting between ~1and
100nm, a range of primary interest for nanotechnology (Fig. 1),
have not been widely exploited because of the challenges in scal-
ing up the techniques optimized for atom cooling, or scaling down
those used for microparticle trapping. Indeed, ecient laser cooling
of atoms relies on light scattering close to a narrow spectral line,
without radiative losses, to reduce the atomic velocity distribution
11
.
Nanostructures lack these features, limiting both the cooling rate
and the minimum achievable temperature
11
. e techniques used
for manipulating microparticles rely on the electric dipole interac-
tion energy
16,17
. Because this scales down approximately with the
particle volume, thermal uctuations are large enough to over-
whelm the trapping forces at the nanoscale
18
.
New approaches were thus developed to stably trap and manip-
ulate nanoparticles. Over the past few years, these techniques
Optical trapping and manipulation of nanostructures
Onofrio M. Maragò
1
*, Philip H. Jones
2
, Pietro G. Gucciardi
1
, Giovanni Volpe
3
and Andrea C. Ferrari
4
*
Optical trapping and manipulation of micrometre-sized particles was first reported in 1970. Since then, it has been successfully
implemented in two size ranges: the subnanometre scale, where light–matter mechanical coupling enables cooling of atoms,
ions and molecules, and the micrometre scale, where the momentum transfer resulting from light scattering allows manipula-
tion of microscopic objects such as cells. But it has been dicult to apply these techniques to the intermediate— nanoscale—
range that includes structures such as quantum dots, nanowires, nanotubes, graphene and two-dimensional crystals, all of
crucial importance for nanomaterials-based applications. Recently, however, several new approaches have been developed and
demonstrated for trapping plasmonic nanoparticles, semiconductor nanowires and carbon nanostructures. Here we review the
state-of-the-art in optical trapping at the nanoscale, with an emphasis on some of the most promising advances, such as con-
trolled manipulation and assembly of individual and multiple nanostructures, force measurement with femtonewton resolution,
and biosensors.
have been successfully applied to a variety of objects, for exam-
ple metal nanoparticles (MNPs)
19–21
, plasmonic nanoparticles
(NPs)
22–30
, quantum dots
31,32
, carbon nanotubes (CNTs)
33–37
, gra-
phene akes
38,39
, nanodiamonds
40
, polymer nanobres
41
and semi-
conductor nanowires
42–49
. Typically these techniques rely either
on special properties of the trapped objects themselves — for
example force enhancement related to plasmonic resonances sup-
ported by the trapped particles
22–30
, or highly anisotropic geom-
etries, such as in CNTs and nanowires
35,38,43,44,46,47
— or on new
approaches to optical manipulation, such as exploiting the eld
enhancement due to plasmons supported by nanostructures on a
substrate
50–57
, or the feedback on the optical forces of the trapped
object
58
. Optical manipulation has been used to build compos-
ite nanoassemblies
32,42,43,59
. Optical tweezers have been developed
to measure forces with femtonewton resolution, enabling the
study of interactions between nanoobjects
34,59–64
. ey have also
been integrated with spectroscopic techniques, such as Raman
spectroscopy
33,36,38,65–71
and photoluminescence
40,44,45,48,49,72
, pav-
ing the way to the selection and manipulation of NPs aer their
individual characterization
36,38
. Optically levitated nanoparti-
cles have been laser-cooled towards their quantum-mechanical
ground state
73–76
.
Here, we review the state-of-the-art, open questions and future
directions in optical trapping and manipulation of nanostructures,
and show how the development of these techniques can aect nano-
science and nanotechnology.
Optical forces on nanostructures
In this section, we review how optical forces arise. We rst consider
the case of particles much smaller than the trapping wavelength
where one can make use of the Rayleigh approximation. We then
address the case of larger particles, where the full electromagnetic
scattering theory must be employed. We nally discuss plasmon-
enhanced forces and optical binding, particularly relevant for opti-
cal trapping and manipulation of nanostructures.
Forces in the dipole approximation. e optical response of
a nanostructure can be oen modelled as that of a dipole
16
or a
collection of dipoles
17
. e dipolar polarizability determines the
1
CNR-IPCF, Istituto per i Processi Chimico-Fisici, I-98158 Messina, Italy,
2
Department of Physics and Astronomy, University College London, London
WC1E 6BT, UK,
3
Department of Physics, Bilkent University, Cankaya, Ankara 06800, Turkey,
4
Cambridge Graphene Centre, University of Cambridge,
9JJ Thomson Avenue, Cambridge CB3 0FA, UK.
*e-mail: marago@me.cnr.it; acf26@hermes.cam.ac.uk
REVIEW ARTICLE
PUBLISHED ONLINE: 7 NOVEMBER 2013 | DOI: 10.1038/NNANO.2013.208
© 2013 Macmillan Publishers Limited. All rights reserved

808 NATURE NANOTECHNOLOGY | VOL 8 | NOVEMBER 2013 | www.nature.com/naturenanotechnology
strength of interaction with an optical eld
16
. For a sphere of
radius a and relative permittivity ε, this can be written as
77
:
α
=
α
0
ε
0
ik
3
α
0
1−
(1)
where α
0
is the point-like particle polarizability given by the
Clausius–Mossotti relation
77
α
0
= ε
0
a
3
(ε – 1)/(ε + 2); k is the
eld wavevector; and ε
0
is the vacuum dielectric permittivity. e
denominator in equation (1) acts as a correction to the Clausius–
Mossotti relation to account for the reaction of a nite-sized dipole
to the scattered eld at its own location
77
. e time-averaged force
acting on such a dipole is
16
:
=
F
j = x,y,z
αE
j
E
j
Re
2
1
(2)
where E
j
are the electric eld components. Equation (2) can be
recast into the more explicit form
78
:
=
Re(α)
Intensity gradientRadiation pressure Polarization gradient
2
+FE
Re(E × H )
+
× E × E
2c
σ
4ωi
σcε
0
4
1
(3)
where σ is the extinction cross-section, E the electric eld, H the
magnetic eld, c the speed of light in vacuum, and ω the angular
frequency of the optical eld. e rst term in equation(3)is the
force due to the gradient of the electric eld intensity, which per-
mits three-dimensional connement in optical tweezers
12
as long as
it dominates the second and third terms. e second term, responsi-
ble for the radiation pressure, corresponds to a force in the propaga-
tion direction
5
. e third term is a force arising from the presence of
spatial polarization gradients
78
.
Forces beyond the dipole approximation. When a particle cannot
be approximated as a dipole, for example in the case of CNTs,
nanowires, graphene and other two-dimensional material akes,
the time-averaged radiation force F
rad
on the centre of mass due to
scattering of an electromagnetic eld is equal in magnitude, and
opposite in sign, to the rate of change of momentum of the electro-
magnetic eld itself
79–84
. erefore, F
rad
can be calculated by inte-
grating the optical momentum ux over a closed orientable surface
S surrounding the object
81,83
:
F
rad
= ∫
S
T
M
·dS (4)
where T
M
is the Maxwell stress tensor, accounting for the interaction
between electromagnetic forces and mechanical momentum
79,80
,
which can be calculated from the scattered elds, and dS is an out-
ward-directed element of surface area. e time-averaged radiation
torque Γ
rad
on the centre of mass can be calculated in an analogous
way as
85
:
Γ
rad
= −∫
S
T
M
× r·dS (5)
where r is the position of the element of surface area.
e scattered electromagnetic elds in equations (4) and (5)can
be calculated using Maxwells equations. Oen, however, this turns
out to be a cumbersome procedure
79
. Various algorithms have
therefore been developed to handle this
79
. In the transition-matrix
(T-matrix) method
82–89
, the total electromagnetic eld — that is, the
sum of incident and scattered eld outside the particle and the eld
internal to the particle — is calculated by expanding all elds in a
common orthogonal basis set of functions and imposing boundary
conditions on the object surface
82,84,86,87
. Most oen, the T-matrix
method uses vector spherical wavefunctions
79
to take advantage of
the spherical symmetry of the scatterer, for example Au or poly-
mer NPs
84,88
. Because the T-matrix works best with objects highly
Figure 1 | The three size ranges of optical trapping. Objects of dierent sizes can be trapped within three main regimes (from left to right): atom trapping
(a few ångstrÖms to a few nanometres), nanotweezers (a few nanometres to a few hundred nanometres) and optical tweezers (from a fraction of a
micrometre up). The horizontal scale bar shows the average object size and the corresponding light wavelength. NV, nitrogen vacancy. Image of layered
material reproduced from ref. 169, © 2011 NPG.
Atom trapping Optical tweezers
Nanotweezers
nm10
−1
11010
2
10
3
10
4
Visible Infrared
Ultraviolet
Atoms
Molecules
Cells
Fullerenes
Nanowires and
nanotubes Graphene
Plasmonic nanoparticles
Synthetic colloids
Quantum dotsNV centres
Layered
materials
REVIEW ARTICLE
NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2013.208
© 2013 Macmillan Publishers Limited. All rights reserved

NATURE NANOTECHNOLOGY | VOL 8 | NOVEMBER 2013 | www.nature.com/naturenanotechnology 809
symmetric in shape and composition, one can treat non-spherical
objects by modelling them as clusters of small spheres
29,41,89
.
Another method to calculate scattered electromagnetic elds
is the discrete dipole approximation (DDA)
90
, also referred to as
the coupled dipole model (CDM)
17
, where the particle is modelled
as a collection of dipoles. e force on each dipole is due to the
incident eld and the elds scattered by all other dipoles (equa-
tion (2)). e force acting on the particle is given by the sum of
the forces acting on each dipole. e torque on the particle can be
calculated in an analogous way. e DDA, although more compu-
tationally intensive than T-matrix
91
, can be directly applied to par-
ticles of any shape and composition. Hybrid methods
92,93
have also
been developed that make use of the T-matrix obtained by point-
matching the near-elds calculated with DDA to get the radiation
force andtorque.
Plasmon-enhanced forces. Two main approaches can be exploited
to use plasmons to enhance optical forces on nanoparticles. e rst,
discussed in this section, is to use the plasmons supported by trapped
MNPs to enhance their mechanical reaction to the elds
22–30
. e
second, covered in the section ‘Plasmonic tweezers, uses plasmons
supported by nanostructures on a substrate, for example pads
52,53
,
nanoantennas
54,55
and nanoholes
56,58
, to generate enhanced elds, in
which nanoparticles can be more eectively trapped
50
.
e optical gradient forces (the rst term in equation(3)) expe-
rienced by nanoparticles are typically very weak (some femtonew-
tons or less), because the dipolar polarizability given by equation(1)
scales with the third power of the particle size
77
. e volume-scaling
of the maximum trapping force was evaluated explicitly in ref.94 for
polystyrene spheres, showing a decrease of three orders of magni-
tude in the maximum trapping force as the sphere radius decreased
from 100to 10nm. erefore, to conne nanoparticles against the
destabilizing eects of thermal uctuations, a signicantly higher
optical power is required: whereas a micrometre-sized polystyrene
sphere can be stably trapped with a fraction of a milliwatt in a stand-
ard optical tweezers set-up (Fig.2a,b), a 100-nm sphere requires
15mW (ref.12). is implies that for a 10-nm sphere ~1.5W would
be needed. e plasmonic nature
95
of MNPs can enhance the opti-
cal forces, so that stable trapping can be achieved at a much lower
power (~2–3mW; refs 26,29,71). On the one hand, far from plas-
mon resonances the optical response of small (<100-nm) spherical
MNPs is (mainly) the optical response of the free-electron plasma
95
yielding a large near-infrared (NIR) polarizability
95
. On the other
hand, MNPs are resonant systems
95
and their optical properties
(polarizability, cross-sections) are regulated by plasmon resonances
that can be tuned by changing size, shapes or aggregation
95
.
Svoboda and Block
19
compared 36.2-nm Au spheres with 38-nm
polystyrene ones, nding a maximum trapping force nearly seven
times as great for Au spheres, as a result of the (seven times) greater
polarizability at the 1,064-nm trapping wavelength
19
. Both Au nan-
oparticles (AuNPs, diameters 9.5–254nm)
20
and Ag nanoparticles
(AgNPs, diameters 20–275 nm)
21
have been optically trapped in
three dimensions. In both cases a maximum trapping force pro-
portional to the third power of the particle radius was observed for
diameters <100nm, with a crossover to a lower exponent for larger
radii
20,21
. is size-scaling behaviour was interpreted by accounting
for local heating
96,97
of the surroundings, and modelling the MNPs
as enclosed in a small steam bubble
88,29
.
Non-spherical MNPs, including Au nanorods (NRs)
22,24
(that
is, nanocylinders with an aspect ratio <10), Ag nanowires
98
and
aggregates of AuNPs
29
, can sustain plasmon resonances in a broad
spectral region in the visible/NIR. ese play a crucial role in
the enhancement of radiation forces and torques in optical twee-
zers
22–24,26–29
. More specically, elongated plasmonic nanostructures
(such as nanowires and NRs) are usually trapped with their axis par-
allel to the electric eld vector of the trapping laser, and orthogonal
to the propagation axis
22,24,84
. e strength of this aligning torque is
increased by tuning the laser close to the plasmon resonance
22
. is
provides a means to control their orientation by rotating the laser
polarization
26
. Plasmonic nanostructures with lengths from tens of
nanometres to several micrometres were aligned and rotated using
a single beam of linearly polarized NIR light
27
. Dienerowitz et al.
25
drew on elements of atom trapping
11
, changing the sign of the gradi-
ent force by blue-detuning the laser wavelength with respect to the
MNP plasmon resonance. us, particle connement was achieved
in the dark spot of an optical vortex beam. e frequency depend-
ence of the plasmon-enhanced radiation force was also used in a
system of two counterpropagating evanescent waves at dierent
wavelengths to selectively guide MNPs of dierent sizes in opposite
directions
30
. Cylindrical vector beams with radial polarization were
also suggested to trap plasmonic NPs, because for these structured
beams the second term in equation (3) (that is, the radiation pres-
sure that pushes particles out of the trap) is zero on the beam axis
99
.
Such structured beams trapped both dielectric microparticles
100,101
and single-walled nanotubes (SWNTs)
102
. Further analysis of the
optical forces
103
revealed, however, that in this case the polarization
gradient contribution to the optical force (the third term in equa-
tion (3)) can be signicant
103
and may eliminate the advantage of
such structured beams.
Resonant illumination of plasmonic NPs gives rise to strong
heating eects because of light absorption
104
. Temperature increases
of hundreds of kelvin were observed by trapping AuNPs adjacent
to uorophore-containing lipid vesicles with permeability sensitive
to temperature
97
. When heated above the gel-transition tempera-
ture, uorophores diused out of the vesicle
97
. Further experiments
made use of the diering longitudinal and transverse plasmon reso-
nances of AuNRs to control the local heating through the orienta-
tion of the AuNRs with respect to the electric eld vector of the
trapping laser
105
. It was suggested
105
that this would make AuNRs
sensitive and switchable remote-controlled heat transducers to
small-volumesamples
105
.
Optical binding forces. Optical binding forces emerge from
multiple scattering between several objects, and can result in the
formation of regular, ordered structures
106–108
. is oers a path
towards large-scale NP assembly and organization in one
109
, two
110
and three dimensions
111
. For example, one-dimensional chains of
MNPs were suggested as an ‘optical sail’
112
to achieve a high driv-
ing force on an attached nanoscopic object, taking advantage of
the huge extinction cross-section of the collective plasmon reso-
nance. Pairs of 200-nm AuNPs were optically bound perpendic-
ular to the direction of light propagation in an optical ‘line trap
formed by reection of a line-shaped focused beam, with particle
separations multiples of the optical wavelength
109
, consistent with
predictions based on light scattering from Rayleigh (dipolar) parti-
cles
113
. Yan et al.
110
used 40-nm-diameter AgNPs, a size well within
the dipole approximation, with both a line trap and a cylindrically
symmetric Bessel beam trap, and observed dimers, chains and
photonicclusters’
110
.
Optical binding interactions can also trap and organize one-
dimensional carbon nanostructures. In Fig. 2c, we show SWNT
bundles illuminated in aqueous suspension
114
by counterpropagat-
ing evanescent elds formed by total internal reection at a glass/
water interface. We observe their self-organization into optically
bound chains, where the bundle axes align parallel to the chain axis
and to the direction of propagation of the incident beams. ese
chains break as soon as the evanescent eld is switched o.
Experimental designs and techniques
In this section, we review the most common experimental imple-
mentations relevant for optical tweezers, with an emphasis on those
used for trapping and manipulation of NPs and nanostructures.
REVIEW ARTICLE
NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2013.208
© 2013 Macmillan Publishers Limited. All rights reserved

810 NATURE NANOTECHNOLOGY | VOL 8 | NOVEMBER 2013 | www.nature.com/naturenanotechnology
Optical tweezers. In the simplest conguration (Fig.2a), optical
tweezers can be generated by focusing a laser beam to a dirac-
tion-limited spot using a high numerical aperture (NA) objective
lens
4,115,116
. is serves the dual purpose of focusing the trapping
light and imaging the trapped object. Samples are oen placed in
small (few microlitres) microuidic chambers held on a motorized
or piezo-driven microscope stage with nanometre position resolu-
tion
116
. Generally, optical tweezers require little power (down to a
few milliwatts
115,116
): carbon and silicon nanostructures have been
trapped with as little as 1–2mW NIR light
34,35,38,46,117
. e optical
tweezer position can be controlled using two steerable mirrors
118,119
.
It is also possible to generate multiple optical tweezers by deecting
a single beam in various positions using, for example, an acousto-
optic deector — that is, a device where intensity and frequency of
an acoustic wave spatially controls the optical beam
115,120
.
Holographic optical tweezers. e range of optical tweezer
applications has been greatly expanded by the use of advanced
beam-shaping techniques, where the shape of a light beam is altered
by diractive optical elements (DOEs) to produce multiple optical
traps at denite positions
13,14
. Figure2d shows a schematic of a holo-
graphic optical tweezer (HOT) set-up, where the DOE is placed in a
plane conjugate to the objective focal plane so that the complex eld
distribution in the trapping plane is the Fourier transform of that in
the DOE plane
121,122
. Oen the DOE is a liquid-crystal spatial light
modulator (SLM) used to modulate the phase of the incoming beam,
because any modulation of the amplitude of the beam would entail a
loss of optical power
121,122
. erefore, various techniques have been
developed to determine the optimal phase modulation, for example
the Gerchberg–Saxton algorithm, based on iterative optimization of
the phase prole at the SLM in order to obtain the desired optical
L1
L2
DM
OBJ
Beam-
steering
mirror
Laser
beam
Illumination
DM
QPD
L3
C
To camera
Laser o
Free
Brownian motion
Laser on
Stable trap
5 μm
BM
−250
−250
−250
0
0
0
250
250
250
x (nm)
y (nm)
z (nm)
M
DM
Evanescent field on
Evanescent field o
Prism
Microscope
objective
SLM
Illumination
5 μm
5 μm
L2L1
DM
To camera
OBJ
C
a
a
m-
a
bc
d
Figure 2 | Basic experimental designs. a, Optical tweezers are obtained by focusing a laser beam to a diraction-limited spot, using a high-numerical-
aperture objective lens (OBJ). Additional optics is needed to steer the optical tweezer position (beam-steering mirror and telescope formed by lenses L1,
L2), to image the sample (illumination, dichroic mirrors DM, and camera) and to track it (condenser, lens L3 and quadrant photodiode QPD). The resulting
traces allow tracking of the Brownian motion (BM) and the calibration of the optical tweezer stiness. Inset: Bright-field image of (top) optically trapped
and (bottom) free SWNT bundle. b, Evanescent optical waves can be excited by total internal reflection at an interface between a high- and a low-refractive-
index medium, often a glass–water interface. The excited evanescent waves can be used to manipulate dielectric and metallic particles. A microscope
objective images the sample. c, The optical forces resulting from a standing evanescent wave created by the interference of two counterpropagating
evanescent fields (arrows) align SWNT bundles end-to-end (top). When the evanescent wave is switched o, the bundles are released from the locked
position and undergo thermal motion (bottom). Scale bar, 5 μm. d, HOTs rely on a programmable diractive optical element (SLM) for the creation, shaping
and control of multiple independent optical tweezers. Inset: (top) dark-field and (bottom) scanning electron microscope images of a 5×5 pattern of 80-nm
AuNPs deposited by HOTs on a glass substrate. Figure reproduced with permission from: a, ref. 35, © 2008 ACS: inset in d, ref.144, © 2011 ACS.
REVIEW ARTICLE
NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2013.208
© 2013 Macmillan Publishers Limited. All rights reserved

NATURE NANOTECHNOLOGY | VOL 8 | NOVEMBER 2013 | www.nature.com/naturenanotechnology 811
tweezer conguration at the trapping plane
123–125
. Over the past few
years, HOTs have been used to manipulate and assemble nanostruc-
tures. For example, semiconducting nanowires have been translated,
rotated, cut and fused into complex structures (Fig.3)
42,126
.
Plasmonic tweezers. In the 1990s and early 2000s, various groups
theoretically suggested harnessing the enhancement associated with
a plasmonic resonance to realize nanoscopic optical tweezers, for
example by using the extremity of a sharp metallic tip
127,128
, the light
transmitted through a nanohole in a metallic lm
129
or metallic pat-
terns to create multiple trapping positions at the nanoscale
50
.
In 2006, Volpe et al.
51
showed that surface plasmon polaritons at
a glass/Au/water interface produce a 40-times increase in the optical
forces on micrometre-sized dielectric particles (Fig.4a), so that col-
lections of such particles could self-organize in large crystals
130
. But
a at metal lm features a homogeneous optical potential
51
, whereas
controlled trapping of single nano-objects requires patterning of the
surface to create three-dimensional conning optical potentials
50
.
is is achieved by using properly designed metal nanostructures
such as pads
52,53
, antennas
54,55
or nanoapertures
56,58
. A typical optical
set-up to excite plasmons is based on the Kretschmann congura-
tion shown in Fig.2b. With similar schemes it is also possible to
arrange microscopic particles in complex congurations corre-
sponding to the locations of metallic micropads, where a plasmonic
resonance can be excited (Fig.4d), and also to integrate such plas-
monic traps with a microuidic environment
131
. Other congura-
tions based on nanoantennas allow one to localize the eld intensity
in hotspots
54,55
(Fig.4c). Fractal plasmonic structures
132
can allow
tight foci below the diraction limit far away (hundreds of nanome-
tres) from the metallic structures (Fig.4d).
Plasmonic interactions can also be harnessed by using the active
feedback from the interaction between the optical tweezer beam
and the trapped particle. It is possible to overcome the scaling of the
optical forces with the third power of the object size, as well as the
increase in Brownian uctuation, by making use of an optical trap
realized with a nanoaperture in a metal lm (Fig.4e) in which the
particle itself has a strong inuence on the local electric eld. e
particle thus has an active role in the trapping mechanism, increas-
ing the stiness of the trap only when the particle tries to escape
58
.
Plasmonic double nanoapertures were also used for optical trapping
of single proteins, paving the way to direct optical manipulation of
smaller objects
56
. Note that whenever plasmonic nanostructures are
involved, the problem of heating must be faced. Wang et al.
57
have
described a method of reducing heat in a plasmonic trap by using a
heat sink integrated with the optical structure.
Photonic force microscopy. A photonic force microscope
(PFM)
133–135
is a scanning probe technique based on optical tweezers
(Fig.5). is concept was originally developed when scanning a die-
lectric particle trapped on a surface and observing how its Brownian
motion in the trap was modied by the probe–sample interaction
133
.
In this way, it was possible to measure extremely small forces down
to femtonewtons, as well as image surface features below the trap-
ping light diraction limit
133,134,136,137
.
e motion of a trapped particle subject to thermal uc-
tuations can be modelled in one dimension by the overdamped
Langevin equation
115
:
=
dx(t)
x(t)
+
2DW(t)
γ
K
x
d(t)
(6)
where x(t) is the particle position, K
x
the stiness of the optical trap,
γ the friction coecient, D the Stokes–Einstein diusion coecient
and W(t) a white noise. When dealing with quantitative force meas-
urements, it is crucial to calibrate the optical trap stiness, K
x
. is
calibration can be obtained by measuring the Brownian trajectory
of the optically trapped particle using the deection of the trapping
beam onto a quadrant photodiode
136
, a device that allows one to map
the particle trajectory
135
. ese trajectories are typically analysed by
tting the autocorrelation function of x(t) to an exponential
138
: the
characteristic decay relaxation time of the autocorrelation function
is τ=γ/K
x
. Deriving the value of γ from hydrodynamics, it is then
possible to measure K
x
. Alternatively, it is possible to perform this
analysis in the frequency domain using the power spectral density
of x(t) (ref. 136), which can be tted to a Lorentzian lineshape
139
.
When dealing with force measurements in the presence of diusion
gradients on the probe, such as close to boundaries or objects, some
correction terms are necessary
18
, and these are more signicant for
a nanometre-sized probe
140
.
Using an optically trapped particle as a PFM probe may be
advantageous in imaging of so structures
135
, because the trap sti-
ness is low (10
–3
to 1pNnm
–1
)
15
compared with that of an atomic
force microscope cantilever (10 to 10
5
pNnm
–1
)
15
, and in volumet-
ric imaging
137
at high temporal resolution (tens of kilohertz sam-
pling rates
137
), which can be achieved by three-dimensional particle
b
5 µm
SnO
2
GaN
100 nm
10 µm
c
10 µm
10
µ
m
Position of focused laser beam
and trapped nanowire
a
Figure 3 | Optical manipulation and placement of nanowires. a, Semiconductor nanowires can be manipulated and assembled with optical forces.
The image shows a GaN nanowire laser-fused to a SnO
2
nanoribbon after manipulation and deposition with optical tweezers. The inset is a scanning
electron micrograph of the fused junction. b, Assembly of a rhombus constructed from semiconductor CdS-nanowires using HOTs. This entails nanowire
translation, cutting and fusion with the substrate. c, Optical tweezing of a In
2
O
3
nanowire (top) and placement by scanning optical tweezers, to connect
two branches of a circuit (bottom). Figure reproduced with permission from: a, ref.43, © 2006 NPG; b, ref.42, © 2005 OSA; c, ref.119, © 2009 OSA.
REVIEW ARTICLE
NATURE NANOTECHNOLOGY DOI: 10.1038/NNANO.2013.208
© 2013 Macmillan Publishers Limited. All rights reserved

Citations
More filters
Journal ArticleDOI
TL;DR: An overview of the key aspects of graphene and related materials, ranging from fundamental research challenges to a variety of applications in a large number of sectors, highlighting the steps necessary to take GRMs from a state of raw potential to a point where they might revolutionize multiple industries are provided.
Abstract: We present the science and technology roadmap for graphene, related two-dimensional crystals, and hybrid systems, targeting an evolution in technology, that might lead to impacts and benefits reaching into most areas of society. This roadmap was developed within the framework of the European Graphene Flagship and outlines the main targets and research areas as best understood at the start of this ambitious project. We provide an overview of the key aspects of graphene and related materials (GRMs), ranging from fundamental research challenges to a variety of applications in a large number of sectors, highlighting the steps necessary to take GRMs from a state of raw potential to a point where they might revolutionize multiple industries. We also define an extensive list of acronyms in an effort to standardize the nomenclature in this emerging field.

2,560 citations

Journal ArticleDOI
TL;DR: By utilizing dual excitation of plasmons at metal-fluid interface, this work creates interacting assemblies of metal nanoparticles, which may be further harnessed in dynamic lithography of dispersed nanostructures and have implications in realizing optically addressable, plasmofluidic, single-molecule detection platforms.
Abstract: Single-molecule surface-enhanced Raman scattering (SM-SERS) is one of the vital applications of plasmonic nanoparticles. The SM-SERS sensitivity critically depends on plasmonic hot-spots created at the vicinity of such nanoparticles. In conventional fluid-phase SM-SERS experiments, plasmonic hot-spots are facilitated by chemical aggregation of nanoparticles. Such aggregation is usually irreversible, and hence, nanoparticles cannot be re-dispersed in the fluid for further use. Here, we show how to combine SM-SERS with plasmon polariton-assisted, reversible assembly of plasmonic nanoparticles at an unstructured metal–fluid interface. One of the unique features of our method is that we use a single evanescent-wave optical excitation for nanoparticle assembly, manipulation and SM-SERS measurements. Furthermore, by utilizing dual excitation of plasmons at metal–fluid interface, we create interacting assemblies of metal nanoparticles, which may be further harnessed in dynamic lithography of dispersed nanostructures. Our work will have implications in realizing optically addressable, plasmofluidic, single-molecule detection platforms. Plasmonic hot-spot generation in solution is not reversible for single-molecule surface-enhanced Raman scattering, which limits its applications. Patra et al.tackle this problem by integrating this technique with thermo-plasmon-assisted reconfiguration of nanoparticles at a metal–fluid interface.

1,705 citations

Journal ArticleDOI
TL;DR: The general overview of the field and the background for appropriate modelling of the physical phenomena are provided and the current state of the art and most recent applications of plasmon resonance in Au NPs are reported.
Abstract: In the last two decades, plasmon resonance in gold nanoparticles (Au NPs) has been the subject of intense research efforts. Plasmon physics is intriguing and its precise modelling proved to be challenging. In fact, plasmons are highly responsive to a multitude of factors, either intrinsic to the Au NPs or from the environment, and recently the need emerged for the correction of standard electromagnetic approaches with quantum effects. Applications related to plasmon absorption and scattering in Au NPs are impressively numerous, ranging from sensing to photothermal effects to cell imaging. Also, plasmon-enhanced phenomena are highly interesting for multiple purposes, including, for instance, Raman spectroscopy of nearby analytes, catalysis, or sunlight energy conversion. In addition, plasmon excitation is involved in a series of advanced physical processes such as non-linear optics, optical trapping, magneto-plasmonics, and optical activity. Here, we provide the general overview of the field and the background for appropriate modelling of the physical phenomena. Then, we report on the current state of the art and most recent applications of plasmon resonance in Au NPs.

1,205 citations

Journal ArticleDOI
TL;DR: This Review provides an account of the recent advancements in reversible photocontrol of the structures and functions of photochromic nanosystems and their applications and outlines the challenges that need to be addressed and the opportunities that can be tapped into.
Abstract: The ability to manipulate the structure and function of promising nanosystems via energy input and external stimuli is emerging as an attractive paradigm for developing reconfigurable and programmable nanomaterials and multifunctional devices. Light stimulus manifestly represents a preferred external physical and chemical tool for in situ remote command of the functional attributes of nanomaterials and nanosystems due to its unique advantages of high spatial and temporal resolution and digital controllability. Photochromic moieties are known to undergo reversible photochemical transformations between different states with distinct properties, which have been extensively introduced into various functional nanosystems such as nanomachines, nanoparticles, nanoelectronics, supramolecular nanoassemblies, and biological nanosystems. The integration of photochromism into these nanosystems has endowed the resultant nanostructures or advanced materials with intriguing photoresponsive behaviors and more sophisticated functions. In this Review, we provide an account of the recent advancements in reversible photocontrol of the structures and functions of photochromic nanosystems and their applications. The important design concepts of such truly advanced materials are discussed, their fabrication methods are emphasized, and their applications are highlighted. The Review is concluded by briefly outlining the challenges that need to be addressed and the opportunities that can be tapped into. We hope that the review of the flourishing and vibrant topic with myriad possibilities would shine light on exploring the future nanoworld by encouraging and opening the windows to meaningful multidisciplinary cooperation of engineers from different backgrounds and scientists from the fields such as chemistry, physics, engineering, biology, nanotechnology and materials science.

463 citations

Journal ArticleDOI
TL;DR: This review highlights the latest optical trapping configurations and their applications in bioscience, as well as recent advances down to the nanoscale, and discusses the future prospects of nanomanipulation.
Abstract: Since the invention of optical tweezers, optical manipulation has advanced significantly in scientific areas such as atomic physics, optics and biological science. Especially in the past decade, numerous optical beams and nanoscale devices have been proposed to mechanically act on nanoparticles in increasingly precise, stable and flexible ways. Both the linear and angular momenta of light can be exploited to produce optical tractor beams, tweezers and optical torque from the microscale to the nanoscale. Research on optical forces helps to reveal the nature of light-matter interactions and to resolve the fundamental aspects, which require an appropriate description of momenta and the forces on objects in matter. In this review, starting from basic theories and computational approaches, we highlight the latest optical trapping configurations and their applications in bioscience, as well as recent advances down to the nanoscale. Finally, we discuss the future prospects of nanomanipulation, which has considerable potential applications in a variety of scientific fields and everyday life.

424 citations

References
More filters
Journal ArticleDOI
TL;DR: This work shows that graphene's electronic structure is captured in its Raman spectrum that clearly evolves with the number of layers, and allows unambiguous, high-throughput, nondestructive identification of graphene layers, which is critically lacking in this emerging research area.
Abstract: Graphene is the two-dimensional building block for carbon allotropes of every other dimensionality We show that its electronic structure is captured in its Raman spectrum that clearly evolves with the number of layers The D peak second order changes in shape, width, and position for an increasing number of layers, reflecting the change in the electron bands via a double resonant Raman process The G peak slightly down-shifts This allows unambiguous, high-throughput, nondestructive identification of graphene layers, which is critically lacking in this emerging research area

13,474 citations

Journal ArticleDOI
TL;DR: Because monolayer MoS(2) has a direct bandgap, it can be used to construct interband tunnel FETs, which offer lower power consumption than classical transistors, and could also complement graphene in applications that require thin transparent semiconductors, such as optoelectronics and energy harvesting.
Abstract: Two-dimensional materials are attractive for use in next-generation nanoelectronic devices because, compared to one-dimensional materials, it is relatively easy to fabricate complex structures from them. The most widely studied two-dimensional material is graphene, both because of its rich physics and its high mobility. However, pristine graphene does not have a bandgap, a property that is essential for many applications, including transistors. Engineering a graphene bandgap increases fabrication complexity and either reduces mobilities to the level of strained silicon films or requires high voltages. Although single layers of MoS(2) have a large intrinsic bandgap of 1.8 eV (ref. 16), previously reported mobilities in the 0.5-3 cm(2) V(-1) s(-1) range are too low for practical devices. Here, we use a halfnium oxide gate dielectric to demonstrate a room-temperature single-layer MoS(2) mobility of at least 200 cm(2) V(-1) s(-1), similar to that of graphene nanoribbons, and demonstrate transistors with room-temperature current on/off ratios of 1 × 10(8) and ultralow standby power dissipation. Because monolayer MoS(2) has a direct bandgap, it can be used to construct interband tunnel FETs, which offer lower power consumption than classical transistors. Monolayer MoS(2) could also complement graphene in applications that require thin transparent semiconductors, such as optoelectronics and energy harvesting.

12,477 citations

Book
15 May 2007
TL;DR: In this paper, the authors discuss the role of surface plasmon polaritons at metal/insulator interfaces and their application in the propagation of surfaceplasmon waveguides.
Abstract: Fundamentals of Plasmonics.- Electromagnetics of Metals.- Surface Plasmon Polaritons at Metal / Insulator Interfaces.- Excitation of Surface Plasmon Polaritons at Planar Interfaces.- Imaging Surface Plasmon Polariton Propagation.- Localized Surface Plasmons.- Electromagnetic Surface Modes at Low Frequencies.- Applications.- Plasmon Waveguides.- Transmission of Radiation Through Apertures and Films.- Enhancement of Emissive Processes and Nonlinearities.- Spectroscopy and Sensing.- Metamaterials and Imaging with Surface Plasmon Polaritons.- Concluding Remarks.

7,238 citations

Journal ArticleDOI
TL;DR: Optical trapping of dielectric particles by a single-beam gradient force trap was demonstrated for the first reported time, confirming the concept of negative light pressure due to the gradient force.
Abstract: Optical trapping of dielectric particles by a single-beam gradient force trap was demonstrated for the first reported time. This confirms the concept of negative light pressure due to the gradient force. Trapping was observed over the entire range of particle size from 10 μm to ~25 nm in water. Use of the new trap extends the size range of macroscopic particles accessible to optical trapping and manipulation well into the Rayleigh size regime. Application of this trapping principle to atom trapping is considered.

6,434 citations

Journal ArticleDOI
TL;DR: The state of the art, future directions and open questions in Raman spectroscopy of graphene are reviewed, and essential physical processes whose importance has only recently been recognized are described.
Abstract: Raman spectroscopy is an integral part of graphene research. It is used to determine the number and orientation of layers, the quality and types of edge, and the effects of perturbations, such as electric and magnetic fields, strain, doping, disorder and functional groups. This, in turn, provides insight into all sp(2)-bonded carbon allotropes, because graphene is their fundamental building block. Here we review the state of the art, future directions and open questions in Raman spectroscopy of graphene. We describe essential physical processes whose importance has only recently been recognized, such as the various types of resonance at play, and the role of quantum interference. We update all basic concepts and notations, and propose a terminology that is able to describe any result in literature. We finally highlight the potential of Raman spectroscopy for layered materials other than graphene.

5,673 citations

Frequently Asked Questions (1)
Q1. What have the authors contributed in "Optical trapping and manipulation of nanostructures" ?

But the magnitude of these effects was considered insignificant for any practical use: to quote J. H. Poynting ’ s presidential address to the British Physical Society in 1905 ( reported in ref. 4 ), “ A very short experience in attempting to measure these forces is sufficient to make one realize their extreme minuteness — a minuteness which appears to put them beyond consideration in terrestrial affairs. ” It was not until 1970, and because of the advent of lasers, that Arthur Ashkin showed that the use of optical forces to alter the motion of micrometre-sized particles5 and neutral atoms6 could have applications in the manipulation of microscopic particles and of single atoms4. On the other hand, what is now commonly referred to as optical tweezers — that is, a tightly focused laser beam capable of confining particles in three dimensions12 — has become a common tool for the manipulation of micrometre-sized particles13,14 and as a highly sensitive force transducer15. But optical forces acting between ~1 and 100 nm, a range of primary interest for nanotechnology ( Fig. 1 ), have not been widely exploited because of the challenges in scaling up the techniques optimized for atom cooling, or scaling down those used for microparticle trapping. The techniques used for manipulating microparticles rely on the electric dipole interaction energy16,17. Because this scales down approximately with the particle volume, thermal fluctuations are large enough to overwhelm the trapping forces at the nanoscale18.