ArticlePDF Available

Computational toolbox for optical tweezers in geometrical optics

Optica Publishing Group
Journal of the Optical Society of America B
Authors:

Abstract and Figures

Optical tweezers have found widespread application in many fields, from physics to biology. Here, we explain in detail how optical forces and torques can be described within the geometrical optics approximation and we show that this approximation provides reliable results in agreement with experiments for particles whose characteristic dimensions are larger than the wavelength of the trapping light. Furthermore, we provide an object-oriented software package implemented in MatLab for the calculation of optical forces and torques in the geometrical optics regime: \texttt{OTGO - Optical Tweezers in Geometrical Optics}. We provide all source codes for \texttt{OTGO} as well as the documentation and code examples -- e.g., standard optical tweezers, optical tweezers with elongated particle, windmill effect, Kramers transitions between two optical traps -- necessary to enable users to effectively employ it in their research and teaching.
Content may be subject to copyright.
Computational toolbox for optical tweezers in geometrical optics
Agnese Callegari,Mite Mijalkov, A. Burak G¨ok¨oz, and Giovanni Volpe
Soft Matter Lab, Department of Physics, Bilkent University, Cankaya, Ankara 06800, Turkey.
(Dated: March 4, 2014)
Optical tweezers have found widespread application in many fields, from physics to biology. Here,
we explain in detail how optical forces and torques can be described within the geometrical op-
tics approximation and we show that this approximation provides reliable results in agreement
with experiments for particles whose characteristic dimensions are larger than the wavelength of
the trapping light. Furthermore, we provide an object-oriented software package implemented in
MatLab for the calculation of optical forces and torques in the geometrical optics regime: OTGO -
Optical Tweezers in Geometrical Optics. We provide all source codes for OTGO as well as the
documentation and code examples – e.g., standard optical tweezers, optical tweezers with elongated
particle, windmill effect, Kramers transitions between two optical traps – necessary to enable users
to effectively employ it in their research and teaching.
INTRODUCTION
Optical tweezers are tightly focused laser beams capa-
ble of holding and manipulating microscopic particles in
three dimensions. Since their invention in 1986 [1], opti-
cal tweezers have been increasing and consolidating their
importance in several fields, from physics to biology [2–
7]. In the last fifteen years, thanks to the development
of relatively simple and cheap setups, optical tweezers
have also started to be employed in undergraduate and
graduate laboratories as a tool to introduce students to
advanced experimental techniques [8–11].
Part of the reason for the success of optical tweez-
ers lies in that the forces they can exert – from tens of
piconewtons down to tens of femtonewtons – are just
in the correct order of magnitude for a gentle but ef-
fective manipulation of colloidal particles and biologi-
cal samples [2–7]. An accurate mathematical descrip-
tion of these forces requires the use of electromagnetic
theory in order to model the interaction between an in-
coming electromagnetic wave and a microscopic particle
[12–14]. However, this can be a daunting task. There-
fore, it comes handy that simpler theoretical approaches
have been shown to deliver accurate results in the lim-
its where the particle characteristic dimensions are much
smaller or much larger than the wavelength of the trap-
ping light [15], which is typically between 532 nm and
1064 nm for optical tweezing applications. For particles
much smaller than the wavelength, one can make use of
the dipole approximation, which has already been ex-
tensively described and employed to describe the trap-
ping of nanoparticles [7]. For particles much larger than
the wavelength, such as cells and large colloidal particles,
whose size is typically significantly larger than one mi-
crometer, one can make use of geometrical optics for the
calculation of optical forces [16]. This approach has been
successfully employed, for example, to describe optical
forces acting on cells [17], the deformation of microscopic
bubbles in a optical field [18], the optical lift effect [19]
and the emerging of negative optical forces [20].
In this paper, we explain in detail how geometrical
optics can be employed in order to study the optical
forces and torques arising in an optical tweezers. We
will first introduce how optical tweezers can be modeled
in geometrical optics. Then, we will study in detail the
forces associated to the scattering of a ray and of an
optical beam by a spherical particle, distinguishing be-
tween scattering and gradient forces. Finally, we will
explore some more complex situations, such as the aris-
ing of torque on non-spherical objects and the emergence
of Kramers transitions between two optical tweezers. As
an integral part of this article, we provide a complete
MatLab software package – OTGO - Optical Tweezers
in Geometrical Optics – to perform the calculation of
optical forces and torques within the geometrical optics
approach [21]. OTGO is fully documented, accompanied by
code examples and ready to be employed to explore more
complex situations, both in research and in teaching. In
fact, we have implemented OTGO using an object-oriented
approach so that it can be easily extended and adapted
to the specific needs of users; for example, it is possible
to create more complex optically trappable particles by
extending the objects provided for spherical, cylindrical
and ellipsoidal particles. In particular, we have used OTGO
to obtain all the results presented in this article.
GEOMETRICAL OPTICS MODEL OF OPTICAL
TWEEZERS
A schematic of a typical optical tweezers is shown in
Fig. 1(a) and in supplementary movies 1, 2 and 3 [21].
A laser beam is focused by a high-NA objective (O1) in
order to create a high-intensity focal spot where a micro-
scopic particle can be trapped. Typically, the particle is
a dielectric sphere with refractive index npimmersed in a
liquid medium with refractive index nm. The scattering
of the focused beam on the particle generates some opti-
cal restoring forces that keep the particle near the focus.
The sum of the incoming and scattered electromagnetic
arXiv:1402.5439v1 [physics.optics] 21 Feb 2014
2
fields can be collected by a second objective (O2) and
projected onto a screen placed in the back-focal plane.
The position of the optically trapped particle can be de-
tected by using the image on the screen [22], as shown in
Figs. 1(b) and 1(c). Note that Fig. 1 is not to scale by
a factor 100 because, in an actual setup, the objective
focal length is 170 µm and the particle size is typically
2µm.
In the geometrical optics approach [16], the incoming
laser beam, whose intensity profile is shown on the left
of Figs. 1(a)-1(c), is decomposed into a set of optical
rays, which are then focused by the objective O1. As the
rays reach the particle, they get partially reflected and
partially transmitted. The direction of the reflected and
transmitted rays are different from those of the incoming
rays. This change of direction entails a change of mo-
mentum and, because of the action-reaction law, a force
acting on the sphere. As we will see, if np> nm, these
optical forces tend to pull the sphere towards the equilib-
rium position near the focal point. As the scattered rays
reach the objective O2, they are collected and projected
onto the back-focal plane.
FORCES BY A RAY ON A PLANAR SURFACE
The energy flux transported by a monochromatic elec-
tromagnetic field, such as the one of a laser beam, is given
by its Poynting vector
S=1
2µRe {E×B},(1)
where Eand Bare the complex electric and magnetic
fields. In order to describe how this energy is transported,
a series of rays can be associated with the electromagnetic
field [23]. These rays are lines perpendicular to the elec-
tromagnetic wavefronts and pointing in the direction of
the electromagnetic energy flow.
When a light ray impinges on a flat surface between
two media with different refractive indices, it is partly
reflected and partly transmitted. Given an incidence an-
gle θi, i.e., the angle between the incoming ray riand
the normal nto the surface at the incidence point, the
reflection angle θris given by the reflection law
θr=θi(2)
and the transmission angle θtis given by Snell’s law
θt= asin ni
nt
sin θi,(3)
where niis the refractive index of the medium of the
incident ray rrand ntis the one of the medium of the
transmitted ray rt. Both rrand rtlie in the plane of
incidence, i.e., the plane that contains riand n. Because
of energy conservation, the power Piof rimust be equal
(a)
O1O2
nm= 1.33
np= 1.50
NA = 1.00
F
(b)
O1O2
F
(c)
O1O2
F
FIG. 1. (color online). Schematic of an optical tweezers setup
(not to scale). A Gaussian laser beam, whose intensity is
shown on the left, is divided into a set of optical rays (lines).
The rays that can cross an aperture stop, whose radius is
equal to the beam waist in this case, are then focused by an
objective (O1, NA = 1.00 in water). Near the focal point, a
dielectric spherical particle (refractive index np= 1.50) im-
mersed in a fluid (refractive index nm= 1.33) scatters the
rays (for clarity, the reflection of the incoming beam and the
internally scattered rays are omitted) and, therefore, experi-
ences a restoring recoil optical force F(black arrow). The
scattered rays are collected by a second objective (O2) and
projected onto a screen placed in the back-focal plane. The
position of the particle can be tracked by monitoring the back-
focal plane image, which sensitively depends on the position
of the particle, e.g., (a) at the focal point, (b) displaced in
the transverse plane and (c) displaced along the longitudinal
direction. The distance between the objectives and the size
of the particle are not to scale by a factor 100. See also
supplementary movies 1, 2 and 3 [21].
to the sum of the power Prof rrand the power Ptof rt,
3
i.e.,
Pi=Pr+Pt.(4)
How the power is split can be calculated by using
Maxwell’s equations with the appropriate boundary con-
ditions [24]. The result is expressed by Fresnel’s equa-
tions and depends on the polarization of the incoming
ray, as we must distinguish the case when the electric
field of the ray oscillates in the plane of incidence (p-
polarization) from the one when it oscillates in a plane
perpendicular to the plane of incidence (s-polarization).
The Fresnel’s reflection and transmission coefficients for
p-polarized light are
Rp=
nicos θtntcos θi
nicos θt+ntcos θi
2
,(5)
Tp=4nintcos θicos θt
|nicos θt+ntcos θi|2(6)
and for s-polarized light
Rs=
nicos θintcos θt
nicos θi+ntcos θt
2
,(7)
Ts=4nintcos θicos θt
|nicos θi+ntcos θt|2.(8)
For unpolarized and circularly polarized light, one can
use the average of the previous coefficients, i.e.,
R=Rp+Rs
2,(9)
T=Tp+Ts
2.(10)
The recoil optical forces are equal and opposite to the
rate of change of linear momentum of the light. Since for
a ray of power Pin a medium of refractive index n, the
momentum flux is nP/c, where cis the speed of light in
vacuum, the optical force is [16]
Fpl =niPi
cˆ
uiniPr
cˆ
urntPt
cˆ
ut,(11)
where ˆ
uiis the unit vector of ri,ˆ
uris the unit vec-
tor of rrand ˆ
utis the unit vector of rt. We must
note that the definition of the momentum of light in a
medium is a thorny issue, which is often referred to as
the Abraham-Minkowski dilemma after the works of Her-
mann Minkowski [25] and Max Abraham [26]. This issue
is discussed in detail, e.g., in Refs. [27, 28]. Since most re-
sults in optical trapping and manipulation do not depend
qualitatively on the momentum definition, in this work
we employ the Minkowski momentum definition which in
fact is the most often employed in optical tweezers stud-
ies [16, 29]. However, we remark that all results can be
easily adapted to the Abraham momentum definition by
changing the definition of the force in Eq. (11) [27].
i
r
i
θ
(1 )
r
r
(2 )
t
r
(3 )
t
r
(4 )
t
r
(5 )
t
r
(6 )
t
r
(1 )
t
r
(2 )
r
r
(3 )
r
r
(4 )
r
r
Qray
Qray,g
Qray,s
306090
0
0.2
0.4
0.6
θi
Q(θi)
(a)
(b)
FIG. 2. (color online). (a) Scattering of a ray impinging on a
sphere. The incident ray riimpinges on a glass spherical parti-
cle (np= 1.50) immersed in water (nm= 1.33). The reflected
(r(j)
r) and transmitted (r(j)
t) rays for the first seven scattering
events are represented. Because of the spherical symmetry of
the particle, all rays lie in the plane of incidence. See also
supplementary movies 4 and 5 [21]. (b) Corresponding trap-
ping efficiencies as a function of the incidence angle θi. See
also supplementary movies 6 and 7 [21].
FORCES BY A RAY ON A SPHERE
We now consider a ray riof power Piimpinging from
a medium with refractive index nmon a dielectric sphere
with refractive index npat an incidence angle θi, as shown
in Fig. 2(a) and in supplementary movies 4 and 5 [21]. As
soon as rihits the sphere, a small amount of its power,
P(1)
r, is diverted into the reflected ray r(1)
r, while most
power, P(1)
t, goes into the transmitted ray r(1)
t. The ray
r(1)
tcrosses the sphere until it reaches the opposite sur-
face, where again a large portion of its power, P(2)
t, is
transmitted outside the sphere into the ray r(2)
t, while a
small amount of its power, P(2)
r, is reflected inside the
sphere into the ray r(2)
r. The ray r(2)
rundergoes another
scattering event as soon as it reaches the sphere bound-
ary, and the process continues until all light has escaped
from the sphere. The force Fray produced on the sphere
by this series of scattering events can be calculated by
4
using repeatedly Eq. (11), i.e.,
Fray =nmPi
cˆ
uinmP(1)
r
cˆ
u(1)
r
X
j=2
nmP(j)
t
cˆ
u(j)
t,(12)
where ˆ
ui,ˆ
u(1)
rand ˆ
u(j)
tare the unit vectors of the incident
ray, the first reflected ray and the j-th transmitted ray,
respectively. We note that the dependence of Eq. (12)
on npis hidden in the dependence of the quantities P(1)
r
and P(j)
ton the Fresnel’s coefficients [Eqs. (7), (8), (5)
and (6)]. Furthermore, we can notice that the absolute
value of the force does not depend on the dimension of
the particle.
Since all the reflected and transmitted rays are con-
tained in the plane of incidence, as can be seen in
Fig. 2(a) and in the supplementary movies 4 and 5 [21],
also the force Fray in Eq. (12) has components only
within the incidence plane. We can, therefore, split Fray
into a component along the direction of the incoming ray,
i.e., the scattering force Fray,s= (Fray ·ˆ
ui)ˆ
ui=Fray,sˆ
ui,
and a component perpendicular to the direction of the in-
coming ray, i.e., the gradient force Fray,g=Fray (Fray ·
ˆ
ui)ˆ
ui=Fray,gˆ
u:
Fray =Fray,s+Fray,g=Fray,sˆ
ui+Fray,gˆ
u,(13)
where ˆ
uis the unit vector perpendicular to ˆ
uiand con-
tained in the incidence plane. Interestingly, the gradient
force is a conservative force, while the scattering force is
nonconservative. If np> nm, the particle is attracted
towards the ray [supplementary movie 4 [21]], while, if
np< nm, the particle is pushed away from the ray [sup-
plementary movie 5 [21]].
In order to quantify the effectiveness of the transfer of
momentum from the ray to the particle, we can introduce
the trapping efficiency, i.e., the ratio between the mod-
ulus of the optical force and the momentum per second
of the incoming ray in a medium with refraction index
ni. The trapping efficiency is bound to lie between 0,
corresponding to a ray that is not deflected, and 2, cor-
responding to a ray that is reflected back on its path [16].
For example, for a 1 mW ray, the maximum optical force
is 7 ·1012 N, i.e., 7 piconewtons. Albeit small, this force
is comparable to the forces that are relevant in the mi-
croscopic and nanoscopic world, e.g., the forces generated
by molecular motors [30], and gives us a first impression
of the potential of optical manipulation. In particular,
we can define the scattering trapping efficiency
Qray,s=c
niPi
Fray,s,(14)
the gradient trapping efficiency
Qray,g=c
niPi
Fray,g(15)
and the total scattering efficiency
Qray =qQ2
ray,g+Q2
ray,s.(16)
Fig. 2(b) and supplementary movie 6 [21] show the trap-
ping efficiencies as a function of θifor a circularly po-
larized ray impinging on a glass sphere (np= 1.50) im-
mersed in water (nm= 1.33); supplementary movie 7 [21]
shows the trapping efficiencies for a circularly polarized
ray impinging on an air bubble (np= 1.00) immersed in
water (nm= 1.33). In both cases, the major contribu-
tion to the total trapping efficiency is given by Qray,g,
while only for very large incidence angles Qray,sbecomes
appreciable.
FORCES BY A FOCUSED BEAM ON A SPHERE
It is not possible to achieve a stable trapping using
a single ray because the particle is permanently pushed
by the scattering force in the direction of the incom-
ing ray, as we have seen in Fig. 2(b). A possible ap-
z
x
x
y
z
Qbe am
Qbe am,g
Qbe am,s-.3
-.2
-.1
.1
.2
.3
R
x
-.3
-.2
-.1
.1
.2
.3
R
(a) (b)
(c) (d)
FIG. 3. (color online). Optical force fields in an optical tweez-
ers. The arrows represent the direction and magnitude of the
force exerted on a glass spherical particle (np= 1.50) in water
(nm= 1.33) illuminated by a highly focused Gaussian beam
(NA = 1.30, beam waist equal to the aperture stop radius)
propagating along the z-direction as a function of the parti-
cle position (a) in the longitudinal (zx) plane and (b) in the
transverse (xy) plane. The shaded area represents the dimen-
sion of the particle. The corresponding trapping efficiencies
are shown in (c) and (d) for particle displacements along the
z-axis and x-axis, respectively. Note that for displacements
along the z-axis both the scattering and the gradient force are
directed along z, while for displacements along the x-axis the
gradient force is directed along x, but the scattering force is
directed along z.
5
(a)
z
x
x
y
(b)
z
x
x
y
(c)
z
x
x
y
FIG. 4. (color online). Force fields in an optical tweezers
generated using various kinds of beams. (a) A Gaussian beam
with a waist much smaller than the radius of the aperture stop
can trap a particle in the transverse plane but cannot confine
the particle along the longitudinal direction. (b) A Laguerre-
Gaussian beam, or doughnut beam, improves the trapping
along the longitudinal direction because of the presence of
more power at large angles. (c) A Hermite-Gaussian beam
clearly breaks the cylindrical symmetry of the trap, as can be
seen from the force field in the transverse plane. In all cases,
the force fields are calculated for a glass spherical particle
(np= 1.50) in water (nm= 1.33). The circle on the left
corresponds to NA = 1.30.
proach to achieve a stable trap is to use a second counter-
propagating light ray. In fact, such a configuration using
two laser beams was amongst the first ones to be em-
ployed in order to trap and manipulate microscopic par-
ticles [31] and a modern version has been obtained us-
ing the light emerging from two optical fibers facing each
other [32]. This approach works also if the two beams are
not perfectly counter-propagating, but they are arranged
with a sufficiently large angle.
A more convenient alternative to using several counter-
propagating light beams is to use a single highly-focused
light beam. In fact, rays originating from diametrically
opposite points of a high-NA focusing lens produce in
practice a set of rays that converge at a very large angle,
as can be seen in Fig. 1.
The most commonly employed laser beam is a Gaus-
sian beam. Its intensity profile at the waist is given by
IG(ρ) = I0eρ2
2w2
0,(17)
where ρis the radial coordinate, w0is the beam waist,
I0=1
20nmE2
0is the beam intensity at ρ= 0, ε0is the
dielectric permittivity of vacuum, and E0is the modulus
of the electric field magnitude at ρ= 0. Such a beam
can be approximated by a set of rays parallel to the op-
tical axis (z) each endowed with a power proportional to
the local intensity of the beam. The resulting rays are
then focused by an objective lens, which has the effect of
bending the light rays towards the focal point, as shown
in Fig. 1 and supplementary movies 1, 2 and 3 [21]. Each
one of these rays produces a force F(m)
ray on the sphere
given by Eq. (12). The total optical force exerted by the
focused beam on the sphere is then the sum of all the
rays’ contributions, i.e.,
Fbeam =X
m
F(m)
ray (18)
In Fig. 3(a) and 3(b), the force fields in the longitudinal
(zx) and transverse (xy) planes are represented as a func-
tion of the distance of the high-refractive index spherical
particle (np= 1.50, nm= 1.33) from the focal point, in
the case of an objective with NA = 1.30 and a circularly
polarized beam.
To have a comparison of the force obtained from our
simulations with the forces usually found in experiments,
we can compare our prediction with the results in Ref.
[30]. We take, for comparison, the measured trapping
stiffnesses for a polystyrene sphere (np= 1.57) with a
1.66 µm diameter in a trap generated using a Gaussian
beam with power P= 10 mW focused by an objective
with NA = 1.20. Performing a calculation with the cited
parameters, we obtain a trapping stiffness along the lon-
gitudinal direction (z) equal to kOTGO
z= 6.45 pNm, and
a trapping stiffness along the transversal direction (x)
equal to kOTGO
x= 12.66 pNm, in reasonable agreement
with the experimental values found in Ref. [30], which are
respectively kexp
z= 3.85 pNm and kexp
x= 11.0 pNm.
The optical force field is cylindrically symmetric
around the z-axis. The equilibrium position lies on z-
axis, i.e., (x, y) = (0,0), and is slightly displaced towards
positive zbecause of the presence of scattering forces,
as is commonly observed in experiments [33]. In fact, a
Brownian particle in an optical trap is in dynamic equi-
librium with the thermal noise pushing it out of the trap
and the optical forces driving it toward the center of the
trap [11], as can be seen from the Brownian motion that
the particles experience in supplementary movies 1, 2 and
3 [21]. The maximum value of the force is achieved when
the particle displacement is about equal to the particle
radius R. We can notice again that, like in the case of a
single ray, the value of the force does not depend on R;
however, the trap stiffness, which is given by the force
divided by the displacement is inversely proportional to
R.
The force Fbeam in Eq. (17) can be split into a scat-
tering force Fbeam,s, given by the sum of the scattering
6
forces of each ray, and a gradient force Fbeam,g, given by
the sum of the gradient forces of each ray [16], i.e.,
Fbeam,s=X
m
F(m)
ray,s(19)
and
Fbeam,g=X
m
F(m)
ray,g.(20)
Like in the case of a single ray, while the gradient force
Fbeam,gis conservative, the scattering force Fbeam,scan
give rise to nonconservative effects, as has been shown in
various experiments [33]. These nonconservative effects
are, however, small [34] as can be seen from the small
displacement along the z-axis of the equilibrium position.
It is now possible to define the scattering efficiencies
for the focused beam as
Qbeam,s=c
niPbeam
Fbeam,s,(21)
Qbeam,g=c
niPbeam
Fbeam,g(22)
and
Qbeam =c
niPbeam
Fbeam,(23)
where Pbeam is the power of beam that contributes to the
focal fields, i.e., after the aperture stop. The scattering
coefficients are shown in Fig. 3(c) and 3(d) for a sphere
displaced along the longitudinal and transverse direction,
respectively. If the sphere is on the z-axis, i.e., the prop-
agation axis of the beam, both the scattering force and
the gradient force act only along the z-direction because
of symmetry. For displacements of the particle along the
x-direction, the gradient force is along the x-direction
and the scattering force along the z-direction.
Other kinds of beams can also be used in optical trap-
ping experiments. In particular, Laguerre-Gaussian and
Hermite-Gaussian beams [24] have been widely exploited.
An accurate description of these beams requires one to
take into account their orbital angular momentum [35],
which can have major effects on their trapping proper-
ties. However, the features related to the presence of spin
angular momentum and of a non-uniform phase profile in
the beam, which lead, e.g., to the presence of orbital an-
gular momentum, cannot be accurately modeled within
the geometrical optics approach. Nevertheless, some fea-
tures connected to the different intensity distribution in
Laguerre-Gaussian and Hermite-Gaussian beams can be
explored, as shown in Fig. 4.
FURTHER NUMERICAL EXPERIMENTS
This section provides some guidelines and examples
on how readers can use geometrical optics and OTGO to
explore more complex situations both in the lab and in
the classroom, going beyond the basic optical tweezers
case of a microscopic sphere optically trapped in a highly
focused laser beam.
FIG. 5. (color online). Simulation of the motion of a pro-
late ellipsoidal Brownian particle (np= 1.50, short semiaxes
2.00 µm, long semiaxis 3.33 µm) in water (nm= 1.33) un-
der the action both of Brownian motion and of the optical
forces and torques arising from a highly focused Gaussian
beam (NA = 1.30, Pi= 1 mW), whose rays are coming from
the bottom. Because of the presence of optical torque, after
70 ms the particle’s long axis gets aligned along the longi-
tudinal direction. See also supplementary video 8 [21].
We will first consider the case of a non-spherical parti-
cle. If the particle is convex, one can still use the formula
in Eq. (13) to calculate the forces due to a single ray.
However, in general, the scattered rays and the force will
not lie all on incidence plane and, therefore, apart from
the optical force also an optical torque can arise:
T= (P1C)×nmPi
cˆ
ui(P1C)×nmP(1)
r
cˆ
u(1)
r
X
j=2
(PjC)×nmP(j)
t
cˆ
u(j)
t(24)
where Cis center of mass of the particle and Pjis the po-
sition where the j-th scattering event takes place. This
is different from the case of a spherical particle, such
as the one shown in Fig. 2(a), where the torque is null
[16]. The typical order of magnitude of the torque on a
particle with characteristic dimension of 1µm is ap-
proximatively 1018 Nm to 1021 Nm for a ray of power
1mW, as shown in experiments [30, 36–38]. For ex-
ample, we can consider the case of an elongated parti-
cle, which can be modelled as a prolate ellipsoidal glass
particle (short semiaxes 2.00 µm, long semiaxis 3.33 µm,
np= 1.50, nm= 1.33), as shown in Fig. 5. Elongated
particles are known to get aligned with their longer axis
along the longitudinal direction because of the presence
of an optical torque [39]. We simulated the optical forces
using OTGO and the Brownian motion using the approach
described in Ref. [40] using the freeware HYDRO++ to cal-
culate the diffusion tensor [41], assuming the particle to
be a room temperature (T= 300 K). We start from a
7
Tto t
ri
rr
F
Mirror 1
Mirror 2
Mirror 3
Mirro r 4
FIG. 6. (color online). The windmill effect. An asymmetric
object illuminated by a plane wave, i.e., by a set of paral-
lel rays (coming from the top), undergoes an optical torque
that can set it into rotation. The object in the illustration is
characterized by a rotational symmetry that permits the can-
cellation of the optical forces in the transverse plane, while
an optical torque is still present.
configuration where the particle center of mass is at the
focal point, but the longer semiaxis lies in the transverse
plane, as shown in Fig. 5(a). Because of the presence of
the optical torque due to a focused Gaussian laser beam
of power 1 mW, the particle gets aligned with the long
semiaxis along the longitudinal direction in about 70 ms,
as shown in Figs. 5(b) and 5(c), which is a result compa-
rable to experiments [42].
Closely related to the optical torque, the windmill ef-
fect [43], where an asymmetric object illuminated by a
plane wave, i.e., a series of parallel rays, can start rotating
around its axis, can also be reproduced using OTGO. In the
simulation, we took a set of parallel rays and shone them
onto a perfectly reflecting object reproducing the shape
of a windmill wheel, i.e., four circular mirrors oriented
as shown in Fig. 6. In the presence of an illuminating
electromagnetic field, this object starts rotating. Simi-
lar structures have been indeed experimentally realized
[43, 44].
Another interesting effect that can be reproduced using
OTGO is the emergence of Kramers’ transitions [45, 46].
We simulated the motion of a Brownian spherical par-
ticle with radius R= 1 µm in the presence of a dou-
ble trap obtained by focalizing two Gaussian beams each
with power 0.25 mW so that their focal points laid at a
distance d= 1.7µm in the transverse plane, as shown in
Fig. 7(a). Letting the system free to evolve under the
(a)
(b)
0.85
0.85
0
x[µm]
t[s]
20 40 60 80 100 120 140 16 0
FIG. 7. (color online). (a) Kramers’ transitions of a spherical
particle of radius R= 1 µm (np= 1.50) in water (nm= 1.33)
held in the optical potential generated by a double optical
tweezers, i.e., two highly focused Gaussian beams (NA=1.30,
Pi= 0.2 mW) whose focal points are separated in the trans-
verse plane by a distance d= 1.7µm and whose rays are
coming from the bottom. The position of the particle is il-
lustrated by the solid line and shows that the particle jumps
between two equilibrium positions. (b) Particle position along
the transverse axis that joins the two traps centers as a func-
tion of time. See also supplementary movie 8 [21].
action of the Brownian motion and the optical forces,
the particle jumps from one potential well to the other
one, as shown by the trajectory in Fig. 7(b). The rel-
atively low value of the power of the trapping beams,
necessary in order to be able to observe the transitions
at room temperature T= 300 K within a relatively short
time frame, is comparable with the one in actual experi-
ments [46]. Changing the parameters of the system, e.g.,
distance between the focal spots, beam power, temper-
ature of the system, one can alter the transition rates
and, moreover, an additional local minimum may arise
between the two traps (e.g., for d= 1.5R), which has
8
ri
r(1c )
r
r(1c )
t
r(1n )
r
r(1n )
t
r(2c )
t
r(2n )
r
r(2n )
tr(3c )
t
r(3n )
t
r(4c )
t
FIG. 8. (color online). A biological cell can be modeled by
using two spherical particles with different refractive indices
placed one inside the other to represent the cytoplasm and
the nucleus. As the cytoplasm is a non-convex shape, the
scattering process is more complex than in the case of a sphere
and for a given ray multiple scattering events may need to be
taken into consideration.
indeed been observed in experiments [47].
It is also possible to extend the computational capabili-
ties of OTGO to non-convex and/or non-simply-connected
shapes. For example, in Fig. 8 we show the case of a
simple optical model for a biological cell [17]: in first ap-
proximation, a cell containing a nucleus can be modeled
by a sphere (the cytoplasm) containing a smaller sphere
of different refractive index (the nucleus). It is interest-
ing to notice that in a scattering event a ray can now be
split into multiple rays that may not necessarily be able
to escape the particle; this is typical of all non-convex
shapes and can lead to a steep increase of the number of
rays to be taken into account.
Acknowledgements
We would like to thank Sevgin Sakıcı for her help in the
early stages of the development of the code, and Onofrio
M. Marag´o, Philip H. Jones and Rosalba Saija for useful
discussions and suggestions. This work has been partially
financially supported by the Scientific and Technological
Research Council of Turkey (TUBITAK) under Grants
111T758 and 112T235, Marie Curie Career Integration
Grant (MC-CIG) under Grant PCIG11 GA-2012-321726,
and COST Actions MP-1205 and IP-1208.
agnese.callegari@fen.bilkent.edu.tr
giovanni.volpe@fen.bilkent.edu.tr;
http://www.softmatter.bilkent.edu.tr
[1] A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and Steven
Chu. Observation of a single-beam gradient force optical
trap for dielectric particles. Opt. Lett., 11:288–290, 1986.
[2] A. Ashkin. Optical trapping and manipulation of neu-
tral particles using lasers. Proc. Natl. Acad. Sci. USA,
94:4853–4860, 1997.
[3] K.C. Neuman and A. Nagy. Single-molecule force spec-
troscopy: optical tweezers, magnetic tweezers and atomic
force microscopy. Nature Methods, 5:491–505, 2008.
[4] K. Dholakia and T. ˇ
Ciˇzm´ar. Shaping the future of ma-
nipulation. Nature Photon., 5:335–342, 2011.
[5] M. Padgett and R. Bowman. Tweezers with a twist.
Nature Photon., 5:343–348, 2011.
[6] M.L. Juan, M. Righini, and R. Quidant. Plasmon nano-
optical tweezers. Nature Photon., 5:349–356, 2011.
[7] O.M. Marag`o, P.H. Jones, P.G. Gucciardi, G. Volpe,
and A.C. Ferrari. Optical trapping and manipulation
of nanostructures. Nature Nanotech., 8:807–819, 2013.
[8] S.P. Smith, S.R. Bhalotra, A.L. Brody, B.L. Brown, E.K.
Boyda, and M. Prentiss. Inexpensive optical tweezers
for undergraduate laboratories. Am. J. Phys., 67:26–35,
1999.
[9] J. Bechhoefer and S. Wilson. Faster, cheaper, safer op-
tical tweezers for the undergraduate laboratory. Am. J.
Phys., 70:393–400, 2002.
[10] D.C. Appleyard, K.Y. Vandermeulen, H. Lee, and M.J.
Lang. Optical trapping for undergraduates. Am. J.
Phys., 75:5–14, 2007.
[11] G. Volpe and G. Volpe. Simulation of a brownian particle
in an optical trap. Am. J. Phys., 81:224–230, 2013.
[12] F. Borghese, P. Denti, and R. Saija. Scattering from
model nonspherical particles: theory and applications to
environmental physics. Springer, Heidelberg, 2003.
[13] T.A. Nieminen, V.L.Y. Loke, A.B. Stilgoe, G. Kn¨oner,
A.M. Bran´czyk, N.R. Heckenberg, and H. Rubinsztein-
Dunlop. Optical tweezers computational toolbox. J. Opt.
A: Pure Appl. Opt., 10:1464–4258, 2007.
[14] S.H. Simpson and S. Hanna. Application of the discrete
dipole approximation to optical trapping calculations of
inhomogeneous and anisotropic particles. Opt. Express,
19:16526–16541, 2011.
[15] P.A.M. Neto and H.M. Nussenzveig. Theory of optical
tweezers. EPL (Europhys. Lett.), 50:702–708, 2000.
[16] A. Ashkin. Forces of a single-beam gradient laser trap on
a dielectric sphere in the ray optics regime. Biophys. J.,
61:569–582, 1992.
[17] Y.-R. Chang, L. Hsu, and S. Chi. Optical trapping of a
spherically symmetric sphere in the ray-optics regime:
a model for optical tweezers upon cells. Appl. Opt.,
45:3885–3892, 2006.
[18] S.E. Skelton, M. Sergides, G. Memoli, O.M. Marag´o, and
P.H. Jones. Trapping and deformation of microbubbles
in a dual-beam fibre-optic trap. J. Opt., 14:075706–1–10,
2012.
[19] G.A. Swartzlander Jr., T.J. Peterson, A.B. Artusio-
Glimpse, and A.D. Raisanen. Stable optical lift. Nature
Photonics, 5:48–51, 2010.
9
[20] V. Kajorndejnukul, W. Ding, S. Sukhov, C.-W. Qiu, and
A. Dogariu. Linear momentum increase and negative op-
tical forces at dielectric interface. Nature Photon., 7:787–
790, 2013.
[21] See supplemental material at xxx for the Matlab software
package OTGO, some code examples, a brief introduction
on how to use OTGO, and a series of animations.
[22] L.P. Ghislain, N.A. Switz, and W.W. Webb. Measure-
ment of small forces using an optical trap. Rev. Sci.
Instrum., 65:2762–2768, 1994.
[23] D.S. Goodman. General principles of geometric optics.
In Handbook of Optics 1. McGraw-Hill, New York, 1995.
[24] C.G. Someda. Electromagnetic Waves, Second Edition.
Electrical engineering textbook series. Taylor & Francis,
London, 2006.
[25] H. Minkowski. Die grundgleichungen f¨ur die elektromag-
netischen vorg¨ange in bewegten k¨orpern. Nachr. Ges.
Wiss. G¨ottingen, pages 53–111, 1908.
[26] M. Abraham. Zur elektrodynamik bewegter k¨orper. R.
C. Circ. Mat. Palermo, 28:1–28, 1909.
[27] R.N.C. Pfeifer, T.A. Nieminem, N.R. Heckenberg, and
H. Rubinsztein-Dunlop. Momentum of an electromag-
netic wave in dielectric media. Rev. Mod. Phys., 74:1197–
1216, 2007.
[28] S.M. Barnett. Resolution of the abraham-minkowski
dilemma. Phys. Rev. Lett., 104(7):070401, 2010.
[29] R.N.C. Pfeifer, T.A. Nieminem, N.R. Heckenberg, and
H. Rubinsztein-Dunlop. Constraining validity of the
minkowski energy-momentum tensor. Phys. Rev. A,
79:023813–1–7, 2009.
[30] A. Rohrbach. Stiffness of optical traps: Quantitative
agreement between experiment and electromagnetic the-
ory. Phys. Rev. Lett., 95:168102–1–4, 2005.
[31] A. Ashkin, J. M. Dziedzic, J. E. Bjorkholm, and Steven
Chu. Acceleration and trapping of particles by radiation
pressure. Phys. Rev. Lett., 24:156–159, 1970.
[32] P. R. T. Jess, V. Garc´es-Chav´ez, D. Smith, M. Mazilu,
L. Paterson, A. Riches, C. S. Herrington, W. Sibbett,
and K. Dholakia. Dual beam fibre trap for raman micro-
spectroscopy of single cells. Opt. Express, 14:5779–5791,
2006.
[33] F. Merenda, G. Boer, J. Rohner, G. Delacr´etaz, and R.-
P. Salath´e. Escape trajectories of single-beam optically
trapped micro-particles in a transverse fluid flow. Opt.
Express, 14:1685–1699, 2006.
[34] G. Pesce, G. Volpe, A.C. De Luca, G. Rusciano, and
G. Volpe. Quantitative assessment of non-conservative
radiation forces in an optical trap. EPL (Europhys.
Lett.), 86:38002–1–6, 2009.
[35] D.L. Andrews and M. Babiker. The Angular Momentum
of Light. Cambridge University Press, Cambridge, 2012.
[36] L. Oroszi, P. Galajda, H. Kirei, S. Bottka, and P. Or-
mos. Direct measurement of torque in an optical trap
and its application to double-strand DNA. Phys. Rev.
Lett., 97:058301–1–4, 2006.
[37] G. Volpe and P. Petrov. Torque detection using brownian
fluctuations. Phys. Rev. Lett., 97:210603–1–4, 2006.
[38] K.C. Neuman and S.M. Block. Optical trapping. Rev.
Sci. Instrum., 75:2787–2809, 2004.
[39] F. Borghese, P. Denti, R. Saija, M. A. Iat´ı, and O. M.
Marag´o. Radiation torque and force on optically trapped
linear nanostructures. Phys. Rev. Lett., 100:163903–1–4,
2008.
[40] M. X. Fernandes and J. Garci´ıa de la Torre. Brownian
dynamics simulation of rigid particles of arbitrary shape
in external fields. Biophys. J., 83:3039–3048, 2002.
[41] AA. VV. Hydro++. http://leonardo.inf.um.es/
macromol/programs/hydro++/hydro++.htm.
[42] S. Bayoudh, T. A. Nieminen, N. R. Heckenberg, and
H. Rubinsztein-Dunlop. Orientation of biological cells
using plane-polarized gaussian beam optical tweezers. J.
Mod. Opt., 50:1581–1590, 2003.
[43] T.A. Nieminen, J. Higuet, G.G. Kn¨oner, V.L.Y.
Loke, S. Parkin, W. Singer, N.R. Heckenberg, and
H. Rubinsztein-Dunlop. Optically driven micromachines:
Progress and prospects. Proc. SPIE 6038, Photonics:
Design, Technology, and Packaging, 2:603813–1–9, 2006.
[44] P. Galajda and P. Ormos. Complex micromachines pro-
duced and driven by light. Appl. Phys. Lett., 78:249–251,
2001.
[45] H. A. Kramers. Brownian motion in a field of force and
the diffusion model of chemical reactions. Physica, 7:284–
304, 1940.
[46] L.I. McCann, M. Dykman, and B. Golding. Thermally
activated transitions in a bistable three-dimensional op-
tical trap. Nature, 402:785–787, 1999.
[47] A. B. Stilgoe, N. R. Heckenberg, T. A. Nieminen, and
H. Rubinsztein-Dunlop. Phase-transition-like proper-
ties of double-beam optical tweezers. Phys. Rev. Lett.,
107:248101–1–4, 2011.
... For particles larger than the ligth wavelength, such as cells 9,10 , micro-bubbles 11 , micro-plastics 12 , or metal-coated Janus micro-particles 13 , these forces can be described using the geometrical optics (GO) approximation. In this approximation, the light field is described as a collection of rays and the momentum exchange between the rays and the particle is calculated via the laws of reflection and refraction 14 . ...
... The force acting on the particle is equal and opposite to the change in momentum of the light, i.e., the momentum of the incident light minus the momenta of the reflected ray in the first scattering event (directionû r,1 ) and the transmitted rays in all subsequent scattering events (directionû t,s where s > 1 is the number of the scattering event). Therefore, the force of a single ray on a particle 14,26,27 is: ...
... We have employed fully connected NNs because they have already proved successful in similar situations 20 . The NNs have been trained using data generated with GO using the toolbox OTGO 14 . Even though the training data comes with artifacts due to the finite number of rays, both the NNs architecture and the training process are designed to obtain NNs predictions that get rid of these artifacts. ...
Preprint
Optical forces are often calculated by discretizing the trapping light beam into a set of rays and using geometrical optics to compute the exchange of momentum. However, the number of rays sets a trade-off between calculation speed and accuracy. Here, we show that using neural networks permits one to overcome this limitation, obtaining not only faster but also more accurate simulations. We demonstrate this using an optically trapped spherical particle for which we obtain an analytical solution to use as ground truth. Then, we take advantage of the acceleration provided by neural networks to study the dynamics of an ellipsoidal particle in a double trap, which would be computationally impossible otherwise.
... As RBCs are significantly larger than the wavelength of the incident light, the optical forces acting on them can be calculated with GO. We perform this calculation with the specialized software OTGO [11]. For biological samples, such as RBCs, that have a low refractive index contrast with the typical suspending medium, the fraction of power that is reflected after a scattering event is very low (< 0.001) [22], therefore in our ray tracing calculations, only the first two scattering (refraction) events are considered. ...
... 90% of these data points are used as a training data set while the remaining 10% is kept as a testing data set to evaluate the accuracy of the NN. The training data are generated via GO calculations made in OTGO [11]. The cell is placed in uniformly distributed positions in a cube of side 8µm centred at the origin of the Cartesian coordinates system (i.e. ...
Preprint
Full-text available
Optically trapping red blood cells allows to explore their biophysical properties, which are affected in many diseases. However, because of their nonspherical shape, the numerical calculation of the optical forces is slow, limiting the range of situations that can be explored. Here we train a neural network that improves both the accuracy and the speed of the calculation and we employ it to simulate the motion of a red blood cell under different beam configurations. We found that by fixing two beams and controlling the position of a third, it is possible to control the tilting of the cell. We anticipate this work to be a promising approach to study the trapping of complex shaped and inhomogeneous biological materials, where the possible photodamage imposes restrictions in the beam power.
... The method of the positional calibration of the particle's trajectory using the microscope camera was published in the TweezPal toolbox in 2010 [7]. The 2014 toolbox [8] describes the geometrical optics to study the force acting on the captured bead in OT. The Tweezpy toolbox [9], published in 2021, reviewed the techniques for force calibration based on analysis of high-frequency camera footage of a particle captured in an OT, also using a PSD detector, and programmed in the Python language. ...
Article
Full-text available
Optical tweezers (OT), or optical traps, are a device for manipulating microscopic objects through a focused laser beam. They are used in various fields of physical and biophysical chemistry to identify the interactions between individual molecules and measure single-molecule forces. In this work, we describe the development of a homemade optical tweezers device based on a cost-effective IR diode laser, the hardware, and, in particular, the software controlling it. It allows us to control the instrument, calibrate it, and record and process the measured data. It includes the user interface design, peripherals control, recording, A/D conversion of the detector signals, evaluation of the calibration constants, and visualization of the results. Particular stress is put on the signal filtration from noise, where several methods were tested. The calibration experiments indicate a good sensitivity of the instrument that is thus ready to be used for various single-molecule measurements.
... Numerical simulations via ray optics model were used to verify and extend experimental results. [28][29][30] The laser beams are decomposed into a set of rays described by a direction vectorki, propagating along the +z-axis with the total power distributed to each rays following the intensity profile of the beam. Incoming rays interact with the surface of dielectric material (polystyrene bead) by means of momentum transfer. ...
Article
Thermodynamics of far-from-equilibrium systems often require measurement of effective parameters such as temperature. Whether such approach is valid for the general case of resetting protocols, active systems, or of confined systems under time-varying fields is still under investigation. We report on the effect of switching ON-OFF of an asymmetric bistable potential to the mean first passage time (MFPT) of a probed particle to go from one potential minima to the other. Experimental results coupled with numerical simulations shows the potential becoming more symmetric at slow switching. Moreover, the MFPT deviates from equilibrium condition with an effective temperature, Teff < T, at slow switching but approaches room temperature, T, at fast switching. For each switching rate, we quantify how far the system is from equilibrium by measuring deviation from a detailed balance like relation and the net circulation of flux present in phase-space. Both analysis suggest equilibrium condition are met at high switching.
... The axial component of gradient force must dominate the scattering force in order to get a stable trap. Usually, the laser beam has a Gaussian intensity profile and under tight focus it leads to a 3D optical trap which, in the geometric optics limit, is explained based on bending of light rays [2][3]. However, such a 'light ray' model does not capture the Gaussian intensity distribution for which a 'light cone' model was proposed [4]. ...
Conference Paper
We present experimental studies to measure optical trapping efficiencies for annular beams of variable diameters but of fixed widths, with aid of laser beam-shaping, to test the ‘light cone’ model in the geometric optics limit.
... The model, by considering the geometrical optics approximation, 46 computes both the exchange of momentum between light and particle (generating optical forces 21 ) and the absorption and consequent heating of the gold cap (generating thermal forces 47 ). While the optical force draws the particle toward the center, the thermal force, caused by the difference in temperature between the gold (inner part) and silica (outer part), pushes the particle away. ...
Preprint
Full-text available
Microengines have shown promise for a variety of applications in nanotechnology, microfluidics, and nanomedicine, including targeted drug delivery, microscale pumping, and environmental remediation. However, achieving precise control over their dynamics remains a significant challenge. In this study, we introduce a microengine that exploits both optical and thermal effects to achieve a high degree of controllability. We find that in the presence of a strongly focused light beam, a gold-silica Janus particle becomes confined at the equilibrium point between optical and thermal forces. By using circularly polarized light, we can transfer angular momentum to the particle breaking the symmetry between the two forces and resulting in a tangential force that drives directed orbital motion. We can simultaneously control the velocity and direction of rotation of the particle changing the ellipticity of the incoming light beam, while tuning the radius of the orbit with laser power. Our experimental results are validated using a geometrical optics model that considers the optical force, the absorption of optical power, and the resulting heating of the particle. The demonstrated enhanced flexibility in the control of microengines opens up new possibilities for their utilization in a wide range of applications, encompassing microscale transport, sensing, and actuation.
... Given the generality of the theoretical framework introduced in [26], our model is well fitted within this context. A freely available software namely Optical Tweezers in Geometrical (OTGO) implements (in MatLab, version R2019a) the theory reported in [25] in a modular object oriented software [28]. In the available distribution of OTGO only simple geometrical shapes, such as spheres or ellipsoids, are present, but the modularity of the software makes it easy for researchers to implement new objects describing more complex shapes. ...
Article
Full-text available
Red blood cells (RBCs) or erythrocytes are essential for oxygenating the peripherical tissue in the human body. Impairment of their physical properties may lead to severe diseases. Optical tweezers have in experiments been shown to be a powerful tool for assessing the biochemical and biophysical properties of RBCs. Despite this success there has been little theoretical work investigating of the stability of erythrocytes in optical tweezers. In this paper we report a numerical study of the trapping of RBCs in the healthy, native biconcave disk conformation in optical tweezers using the ray optics approximation. We study trapping using both single- and dual-beam optical tweezers and show that the complex biconcave shape of the RBC is a significant factor in determining the optical forces and torques on the cell, and ultimately the equilibrium configuration of the RBC within the trap. We also numerically demonstrate how the addition of a third or even fourth trapping laser beam can be used to control the cell orientation in the optical trap. The present investigation sheds light on the trapping mechanism of healthy erythrocytes and can be exploited by experimentalist to envisage new experiments.
Article
Full-text available
Optical tweezers are tools made of light that enable contactless pushing, trapping, and manipulation of objects ranging from atoms to space light sails. Since the pioneering work by Arthur Ashkin in the 1970s, optical tweezers have evolved into sophisticated instruments and have been employed in a broad range of applications in life sciences, physics, and engineering. These include accurate force and torque measurement at the femtonewton level, microrheology of complex fluids, single micro- and nanoparticle spectroscopy, single-cell analysis, and statistical-physics experiments. This roadmap provides insights into current investigations involving optical forces and optical tweezers from their theoretical foundations to designs and setups. It also offers perspectives for applications to a wide range of research fields, from biophysics to space exploration.
Article
Full-text available
Optical tweezers are tools made of light that enable contactless pushing, trapping, and manipulation of objects ranging from atoms to space light sails. Since the pioneering work by Arthur Ashkin in the 1970s, optical tweezers have evolved into sophisticated instruments and have been employed in a broad range of applications in life sciences, physics, and engineering. These include accurate force and torque measurement at the femtonewton level, microrheology of complex fluids, single micro- and nanoparticle spectroscopy, single-cell analysis, and statistical-physics experiments. This roadmap provides insights into current investigations involving optical forces and optical tweezers from their theoretical foundations to designs and setups. It also offers perspectives for applications to a wide range of research fields, from biophysics to space exploration.
Preprint
Full-text available
Optical tweezers are tools made of light that enable contactless pushing, trapping, and manipulation of objects ranging from atoms to space light sails. Since the pioneering work by Arthur Ashkin in the 1970s, optical tweezers have evolved into sophisticated instruments and have been employed in a broad range of applications in life sciences, physics, and engineering. These include accurate force and torque measurement at the femtonewton level, microrheology of complex fluids, single micro- and nanoparticle spectroscopy, single-cell analysis, and statistical-physics experiments. This roadmap provides insights into current investigations involving optical forces and optical tweezers from their theoretical foundations to designs and setups. It also offers perspectives for applications to a wide range of research fields, from biophysics to space exploration.
Article
Full-text available
Optical tweezers are widely used for the manipulation of cells and their internal structures. However, the degree of manipulation possible is limited by poor control over the orientation of the trapped cells. We show that it is possible to controllably align or rotate disc-shaped cells—chloroplasts of Spinacia oleracea—in a plane-polarized Gaussian beam trap, using optical torques resulting predominantly from circular polarization induced in the transmitted beam by the non-spherical shape of the cells.
Article
Full-text available
Current optical manipulation techniques rely on carefully engineered setups and samples. Although similar conditions are routinely met in research laboratories, it is still a challenge to manipulate microparticles when the environment is not well controlled and known a priori, since optical imperfections and scattering limit the applicability of this technique to real-life situations, such as in biomedical or microfluidic applications. Nonetheless, scattering of coherent light by disordered structures gives rise to speckles, random diffraction patterns with well-defined statistical properties. Here, we experimentally demonstrate how speckle fields can become a versatile tool to efficiently perform fundamental optical manipulation tasks such as trapping, guiding and sorting. We anticipate that the simplicity of these “speckle optical tweezers” will greatly broaden the perspectives of optical manipulation for real-life applications.
Article
Full-text available
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.
Article
Full-text available
We extend a previous proposal for absolute calibration of optical tweezers by including optical setup aberrations into the first-principles theory, with no fitting parameters. Astigmatism, the dominant term, is determined from images of the focused laser spot. Correcting it can substantially increase stiffness. Comparison with experimental results yields agreement within error bars for a broad range of bead sizes and trap heights, as well as different polarizations. Absolute calibration is established as a reliable and practical method for applications and design of optical tweezers systems.
Article
Full-text available
Light carries momenta that can be transferred to objects. Relying on gradient forces created by structured light, one can trap and move microscopic particles. Aside from the conservative action of gradient forces, light always pushes an object along its direction of propagation. Here, we demonstrate that gradientless light fields can exert pulling forces on arbitrary objects in a purely passive dielectric environment and without resorting to non-paraxial illumination, interference of multiple beams, gain or other exotic materials. The forces acting against the flow of light arise naturally due to the appropriate amplification of the photon linear momentum when light is scattered from one dielectric medium into another with higher refractive index. This situation opens up a number of intriguing prospects for optical forces and their effects on surface-bound objects. Here, we demonstrate that this new mechanism can be used to manipulate objects over macroscopic distances along dielectric interfaces.
Article
Full-text available
The forces acting on an optically trapped particle are usually assumed to be conser-vative. However, the presence of a non-conservative component has recently been demonstrated. Here, we propose a technique that permits one to quantify the contribution of such a non-conservative component. This is an extension of a standard calibration technique for optical tweezers and, therefore, can easily become a standard test to verify the conservative optical force assumption. Using this technique, we have analyzed optically trapped particles of different size under different trapping conditions. We conclude that the non-conservative effects are effectively negligible and do not affect the standard calibration procedure, unless for extremely low-power trapping, far away from the trapping regimes usually used in experiments. Introduction. – The detection and measurement of forces and torques in microscopic systems is an important goal in many areas such as biophysics, colloidal physics, and hydrodynamics of small systems. Since 1993, the photonic force microscope (PFM) has become a standard tool to probe such forces [1–3]. A typical PFM setup comprises an optical trap —a highly focused Gaussian light beam— that holds a probe —a dielectric or metallic particle of micrometer size— and a position sensing system. Using a PFM it has been possible to measure forces as small as 25 fN [4] and torques as small as 4000 fN nm [5]. In order to assess the mechanical properties of micro-scopic systems, the first step is always to have an accu-rately calibrated optical probe. Modelling the interaction between the light of a focused laser beam and an extended dielectric or metallic object can be a complicated task [6].
Article
Full-text available
An optically trapped Brownian particle is a sensitive probe of molecular and nanoscopic forces. An understanding of its motion, which is caused by the interplay of random and deterministic contributions, can lead to greater physical insight into the behavior of stochastic phenomena. The modeling of realistic stochastic processes typically requires advanced mathematical tools. We discuss a finite difference algorithm to compute the motion of an optically trapped particle and the numerical treatment of the white noise term. We then treat the transition from the ballistic to the diffusive regime due to the presence of inertial effects on short time scales and examine the effect of an optical trap on the motion of the particle. We also outline how to use simulations of optically trapped Brownian particles to gain understanding of nanoscale force and torque measurements, and of more complex phenomena, such as Kramers transitions, stochastic resonant damping, and stochastic resonance.
Article
Activated escape from a metastable state underlies many physical, chemical and biological processes: examples include diffusion in solids, switching in superconducting junctions, chemical reactions and protein folding. Kramers presented the first quantitative calculation of thermally driven transition rates in 1940. Despite widespread acceptance of Kramers' theory, there have been few opportunities to test it quantitatively as a comprehensive knowledge of the system dynamics is required. A trapped brownian particle (relevant to our understanding of the kinetics, transport and mechanics of biological matter) represents an ideal test system. Here we report a detailed experimental analysis of the brownian dynamics of a sub-micrometre sized dielectric particle confined in a double-well optical trap. We show how these dynamics can be used to directly measure the full three-dimensional confining potential-a technique that can also be applied to other optically trapped objects. Excellent agreement is obtained between the predictions of Kramers' theory and the measured transition rates, with no adjustable or free parameters over a substantial range of barrier heights.
Article
Preface D. L. Andrews and M. Babiker; 1. Light beams carrying orbital angular momentum J. B. Götte and S. M. Barnett; 2. Vortex transformation and vortex dynamics in optical fields G. Molina-Terriza; 3. Vector beams in free space E. J. Galvez; 4. Optical beams with orbital angular momentum in nonlinear media A. S. Desyatnikov and Y. S. Kivshar; 5. Ray optics, wave optics and quantum mechanics G. Nienhuis; 6. Quantum formulation of angle and orbital angular momentum J. B. Götte and S. M. Barnett; 7. Dynamic rotational frequency shift I. Bialynicki-Birula and Z. Bialynicka-Birula; 8. Spin-orbit interactions of light in isotropic media K. Y. Bliokh, A. Aiello and M. A. Alonso; 9. Quantum electrodynamics, angular momentum and chirality D. L. Andrews and M. Babiker; 10. Trapping of charged particles by Bessel beams I. Bialynicki-Birula, Z. Bialynicka-Birula and N. Drozd; 11. Theory of atoms in twisted light M. Babiker, D. L. Andrews and V. E. Lembessis; 12. An experimentalist's introduction to orbital angular momentum for quantum optics J. Romero, D. Giovannini, S. Franke-Arnold and M. J. Padgett; 13. Measurement of light's orbital angular momentum M. P. J. Lavery, J. Courtial and M. J. Padgett; 14. Efficient generation of optical twisters using helico-conical beams V. R. Daria, D. Palima and J. Glückstad; 15. Self similar modes of coherent diffusion with orbital angular momentum O. Firstenberg, M. Shuker, R. Pugatch and N. Davidson; 16. Dimensionality of azimuthal entanglement M. van Exter, E. Eliel and H. Woerdman; Index.
Article
We present results of numerical calculations to evaluate the performance of a dual-beam fibre-optic trap for low refractive index particles such as ultrasound contrast agent microbubbles. Using a geometrical optics approach, we determine the range of parameters of microbubble size and beam dimensions over which the optical trap is stable and evaluate the trapping forces and spring constants. Additionally, we calculate the optically induced stress profile over the surface of the microbubble and evaluate the resulting deformation of the microbubble using elastic membrane theory. Our results suggest that such an experiment could be a useful tool for quantifying the mechanical properties (elastic modulus) of the shell material of an ultrasound contrast agent microbubble.