Вы находитесь на странице: 1из 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260069190

Theory of trapping forces in optical tweezers

Article  in  Proceedings of The Royal Society A Mathematical Physical and Engineering Sciences · December 2003
DOI: 10.1098/rspa.2003.1164

CITATIONS READS
116 678

3 authors, including:

Paulo A. Maia Neto Herch Nussenzveig


Federal University of Rio de Janeiro Federal University of Rio de Janeiro
107 PUBLICATIONS   2,592 CITATIONS    98 PUBLICATIONS   3,876 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Enantioselection of the chiral molecules in optical tweezers View project

Dynamical Casimir effect View project

All content following this page was uploaded by Herch Nussenzveig on 02 April 2015.

The user has requested enhancement of the downloaded file.


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers


A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig
Proc. R. Soc. Lond. A 2003 459, 3021-3041
doi: 10.1098/rspa.2003.1164

References Article cited in:


http://rspa.royalsocietypublishing.org/content/459/2040/3021#related-urls

Email alerting service Receive free email alerts when new articles cite this article - sign up in the box at the top right-hand
corner of the article or click here

To subscribe to Proc. R. Soc. Lond. A go to: http://rspa.royalsocietypublishing.org/subscriptions

This journal is © 2003 The Royal Society


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

10.1098/rspa.2003.1164

Theory of trapping forces in optical tweezers


By A. M a z o l l i1 , P. A. M a i a N e t o1 a n d H. M. Nussenzveig1,2
1
Instituto de Fı́sica, Universidade Federal do Rio de Janeiro,
Caixa Postal 68528, 21945-970 Rio de Janeiro, RJ, Brazil
2
NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA

Received 31 May 2002; revised 10 December 2002; accepted 31 March 2003;


published 24 September 2003

Starting from a Debye-type integral representation valid for a laser beam focused
through a high numerical aperture objective, we derive an explicit partial-wave (Mie)
representation for the force exerted on a dielectric sphere of arbitrary radius, position
and refractive index. In the semi-classical limit, the ray-optics result is shown to
follow from the Mie expansion, holding in the sense of a size average. The equilibrium
position and trap stiffness oscillate as functions of the circumference-to-wavelength
ratio, a signature of interference, not predicted by previous theories. We also present
comparisons with experimental results.
Keywords: optical traps; semiclassical limit; Mie scattering

1. Introduction
Single-beam optical traps, commonly known as optical tweezers, have become a pow-
erful tool, with many applications in physics and biology (Ashkin 1997). They have
been used to measure forces in the piconewton range, opening the way for quan-
titative investigations in some fundamental fields of cell biology, such as molecular
motors (Wang et al . 1998; Mehta et al . 1999). The trapped particles are usually
dielectric microspheres, employed as handles in most quantitative applications (Svo-
boda & Block 1994).
In order to achieve trapping, a microscope objective is used to bring the laser
beam to a diffraction-limited focal spot. Typical values of numerical aperture (NA)
are greater than 1, corresponding to beam-opening angles θ0 > 60◦ . Gaussian beam
models are reasonable approximations for near-paraxial conditions (sufficiently small
opening angles; beam waist much larger than the wavelength). However, diffraction-
limited beams cannot be represented by such models: theoretical treatments based
on near-paraxial approximations (Barton et al . 1989; Gussgard et al . 1992; Ren et
al . 1996) are not valid descriptions of optical tweezers (Svoboda & Block 1994).
Alternative approaches, based on geometrical-optics (GO) approximations (Ashkin
1992; Gu et al . 1997), should not be applicable to the microspheres usually employed
as handles, since their sizes are not much larger than the laser wavelength.
Most measurements employ indirect force calibrations against the Stokes law under
complicated boundary conditions for the flow of surrounding fluid past the micro-
sphere, leading to discrepancies among results by factors of two or more. A reliable
derivation of the trapping force is sorely needed, so as to disentangle the effects of

Proc. R. Soc. Lond. A (2003) 459, 3021–3041 


c 2003 The Royal Society
3021
Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3022 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

aberrations, boundary perturbations (e.g. back reflection of light at the glass–water


sample cell interface) and radiometric forces, opening the way for more accurate
measurements.
In this paper we present a first-principles derivation of the optical force. We start
from the Debye-type integral representation of the laser beam as a superposition of
plane electromagnetic waves, developed by Richards & Wolf (1959). In contrast with
paraxial models that represent the strongly convergent light as a Gaussian beam, this
representation has the crucial advantage of correctly describing an optical beam with
large opening angle, also taking into account diffraction at the entrance aperture of
the lens objective.
The field scattered by the sphere is a superposition of Mie-type fields, each one
corresponding to a given plane-wave component of the incident beam. By making
use of suitable rotation operators, we derive analytical results resembling Debye’s
formula for the radiation pressure of a plane wave (Nussenzveig 1992). In particular,
we obtain a single-variable integral representation for the multipole coefficients.
The derivation of analytical results has two important consequences. First, it allows
us to explore a broad range of experimental parameters, since the numerical work
is limited to the evaluation of well-known Bessel and Legendre functions and of
integrals over a single variable, the angle θ defining the direction of a given plane
wave (θ runs from zero to θ0 ). Second, it allows for a first-principles derivation of
the GO approximation, starting from the partial-wave result, by taking the short-
wavelength (semi-classical) limit. This derivation clarifies the connection between
wave and GO, showing that the latter holds as an average over the size parameter
β = ka (a = microsphere radius, k = 2πn1 /λ0 = wavenumber in the surrounding
medium, of refractive index n1 , λ0 = laser vacuum wavelength), when β  1. Results
for the axial force in the simpler particular case in which the microsphere is aligned
along the optical axis were reported by two of us (Maia Neto & Nussenzveig 2000).
In § 2 we obtain the expressions for the force, from which we derive the GO result
in § 3. Numerical results and comparisons with the few available experimental data
are found in § 4. Section 5 contains the conclusions.

2. Optical force as a partial-wave series


We take circular polarization, so that the optical force field is symmetrical under rota-
tions around the beam axis. In order to simplify the notation, we consider left-handed
polarization σ+. Polarization σ− may be considered within the same formalism, and
the final results obtained from those presented here, as discussed below.
Thus the electric field of the strongly focused beam has the Debye-type (Richards
& Wolf 1959) integral representation (we omit the time factor e−iωt ),
 2π  θ0 √
dθ sin θ cos θ e−γ sin θ eik·(r+R) ˆ(θ, φ),
2 2
Ein (r) = E0 dφ (2.1)
0 0

where k = |k(θ, φ)| = n1 ω/c, ˆ(θ, φ) = x̂ + iŷ  and the unit vectors x̂ and ŷ 
are obtained from x̂ and ŷ, respectively, by rotation with Euler angles (φ, θ, −φ).
According to (2.1), the incident electric field is a superposition of circularly polar-
ized plane waves. The focus is at r = −R, the point where all plane waves are in
phase, yielding constructive interference and maximum intensity. As the dephasing

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3023

among the plane waves increases, the intensity at the focal plane decreases, becom-
ing localized to a transverse dimension of the order of the wavelength. The factor
exp(−γ 2 sin2 θ) results from the Gaussian intensity profile of the laser beam at the
entrance pupil of the objective, I = Im exp(−2ρ2 /w2 ) (where w is the waist, Im is
the intensity at the beam axis, and ρ is the distance to the beam axis), and from the
Abbe sine condition ρ = f sin θ, where f is the focal length, so that γ = f /w. Note
that we take a Gaussian beam before the passage through the objective. The actual
trapping beam, given by (2.1), is not Gaussian, and arbitrary values of θ0 may be
considered, in a consistent way.
The scattered electric and magnetic fields Es and Hs are computed in appendix A.
We denote by n2 the absolute refractive index of the sphere. We neglect absorption,
assuming that both n1 and n2 are real. We employ a spherical coordinate system,
with origin at the centre of the sphere. For each plane-wave component of the input
field (2.1), the scattered field is a corresponding Mie partial-wave series, and the
resulting total scattered field is a linear superposition of Mie components. They are
given in terms of the Mie coefficients aj and bj , which correspond (when multiplied
by i/k) to the scattering amplitudes for electric and magnetic multipole waves. The
Mie coefficients are functions of the size parameter β, and of the relative refractive
index n = n2 /n1 (Bohren & Huffman 1983).
The optical force is computed through the Maxwell stress tensor. Because of
momentum conservation, we can take a spherical surface at infinity:
  
F = lim − 12 r dΩ r(εE 2 + µ0 H 2 ) . (2.2)
r→∞ S(r)

Following Ashkin (1992), we define the dimensionless force through a vector efficiency
factor,
F
Q= , (2.3)
n1 P/c
where P is the laser power at the sample and c is the speed of light. Q quantifies
how efficiently the available field momentum is transferred to the sphere. Except for
possible nonlinear optical effects (not considered in this paper), Q does not depend
on the laser power P . For normal incidence of a plane wave on a totally reflecting
mirror, Q = 2. In optical tweezers, the maximum value of Q is usually smaller by at
least one order of magnitude.
We assume that the transmission loss through the objective is uniformly dis-
tributed within the focused beam, so that the intensity distribution is not affected
by this process. We may then ignore losses when considering the ratio in (2.3), and
replace P by the power filling the objective aperture,
P = 12 πw2 AIm , (2.4)

where A = 1 − exp(−2γ 2 sin2 θ0 ) is the fraction of available beam power that fills the
objective aperture.
It is convenient to express the force in terms of cylindrical coordinates, with the
origin at the focus and the centre of the sphere at the point R = (ρR , ϕR , zR ). Since
we take circular polarization, the cylindrical components do not depend on ϕR (by
axial symmetry). The final expressions are sums over the variables j, for the total

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3024 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

angular momentum J 2 (eigenvalues j(j + 1)) and m for its z component Jz , of the
form
  ∞  j
= .
j,m j=1 m=−j

The electric and magnetic fields in (2.2) can be written as


E = Ein + Es , H = Hin + Hs . (2.5)
Hence the squared fields in (2.2) contain terms of the form Ein · Es and
2
Ein , Es2
(like-
wise for H). The first term does not contribute, whereas the last two yield separate
contributions to the force, denoted, respectively, Qe and Qs . Qe represents the rate
at which momentum is removed from the input beam (normalized as in (2.3)). It is
closely related to the effect of extinction (energy removal), resulting from interfer-
ence between incident and scattered fields. Part of this momentum is carried away
by the scattered field, at the rate −Qs , and part is transferred to the microsphere.
Thus the force is given by
Q = Qs + Qe . (2.6)
For the axial components, we find
 
8γ 2 j(j + 2)(j − m + 1)(j + m + 1)
Qsz = − Re
A j+1
j,m 
∗ ∗ ∗ 2j + 1 ∗
× (aj aj+1 + bj bj+1 )Gj,m Gj+1,m + maj bj |Gj,m | ,
2
j(j + 1)
(2.7)
4γ 2 
Qez = Re (2j + 1)(aj + bj )Gj,m (Gj,m )∗ , (2.8)
A j,m

where Gj,m and Gj,m are multipole coefficients of the focused incident beam:
 θ0 √
dθ sin θ cos θ e−γ sin θ djm,1 (θ)Jm−1 (kρR sin θ)eikzR cos θ , (2.9)
2 2
Gj,m (ρR , zR ) =
0
∂Gjm
Gj,m (ρR , zR ) = −i , (2.10)
∂(kzR )
where djm,m (θ) are the matrix elements of finite rotations (Edmonds 1957) and
Jm are Bessel functions (Abramowitz & Stegun 1972). When the sphere is aligned
along the optical axis (ρR = 0), we may use the result Jm−1 (0) = δm,1 to sim-
plify (2.7)–(2.10). The resulting expressions, as expected, agree with those obtained
by Maia Neto & Nussenzveig (2000).
For the transverse components, we find

γ2  j(j + 2)(j + m + 1)(j + m + 2)
Qsρ = 4 Im
A j,m
j+1

× [(aj a∗j+1 + bj b∗j+1 )(Gj,m G∗j+1,m+1 + Gj,−m (Gj+1,−m−1 )∗ )]



2j + 1  ∗ ∗
−2 (j − m)(j + m + 1) Re(aj bj )Gj,m Gj,m+1 ,
j(j + 1)
(2.11)

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3025



γ2  j(j + 2)(j + m + 1)(j + m + 2)
Qsϕ = −4 Re
A j,m
j+1

× [(aj a∗j+1 + bj b∗j+1 )(Gj,m G∗j+1,m+1 − Gj,−m (Gj+1,−m−1 )∗ )]



2j + 1  ∗ ∗
−2 (j − m)(j + m + 1) Re(aj bj )Gj,m Gj,m+1
j(j + 1)
(2.12)
and
γ2 
Qeρ = 2 Im (2j + 1)[(aj + bj )Gj,m (G− + ∗
j,m+1 − Gj,m−1 ) ], (2.13)
A j,m

γ2 
Qeϕ = −2 Re (2j + 1)[(aj + bj )Gj,m (G− + ∗
j,m+1 + Gj,m−1 ) ], (2.14)
A j,m

where
 θ0 √
G± dθ sin2 θ cos θ e−γ sin θ djm±1,1 (θ)Jm−1 (kρR sin θ)eikzR cos θ .
2 2
jm (ρR , zR ) =
0
(2.15)
The transverse force contains products of multipole coefficients for consecutive values
of m.† Hence, as was remarked following (2.10), it vanishes when ρR = 0, as expected.
We have also considered polarization σ−. The force in this case may also be com-
puted from (2.7)–(2.15), provided that we change the sign of the azimuthal compo-
nent, Qϕ [σ−] = −Qϕ [σ+].
An important check of (2.7)–(2.15), performed next, is to consider their semi-
classical limits, which must be related to the GO approximation.

3. GO limit
In this section, we obtain a closed analytical representation for the force in the GO
approximation, starting from the exact results (2.7)–(2.15), in the form of an integral
over θ and φ of elementary functions. Previous approaches relied on time-consuming
numerical vector sums of the forces exerted by each ray (Ashkin 1992; Gu et al .
1997). Moreover, our approach takes into account the correct pre-factors describing
the intensity distribution in the focused beam, verifying the Abbe sine condition.
The derivation of the GO approximation from the exact partial-wave series is
a textbook example of the subtle nature of semi-classical approximations (Berry
& Mount 1972). It involves the following stages (Nussenzveig 1992; Nussenzveig &
Wiscombe 1980).

† Equations (2.7), (2.11) and (2.12), giving the scattering component Qs , agree with the results
of Barton et al. (1989) (see also Farsund & Felderhof 1996) if we replace Gjm by paraxial multipole
coefficients. In the same sense, the extinction component Qe as given by (2.8), (2.13) and (2.14) is also
formally equivalent to the result of Barton et al. To make the connection for the axial component Qez ,
we write cos θdjm,1 (θ) as a linear combination of dj+1 j j−1
m,1 (θ), dm,1 (θ) and dm,1 (θ), allowing us to express
Gjm in terms of Gj+1,m , Gjm and Gj−1,m . Likewise, for the transverse components Qeρ and Qeϕ , we
expand G± jm as a linear combination of Gj  m with different j .


Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3026 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

(i) As all ‘quantum numbers’ in eigenfunction expansions must be large, summa-


tions over them are replaced by integrals.
(ii) The WKB asymptotic approximation is employed for all special functions.
(iii) All integrals involving rapidly varying phase factors are evaluated by the sta-
tionary phase method.
(iv) Finally, to suppress rapidly varying interference oscillations, averages over the
oscillations are taken. These stages are performed in Appendices B–E.
The final results are conveniently expressed in terms of the projection of the sphere
position along the direction perpendicular to a given optical ray defined by the unit
vector r̂(θ, φ):
R⊥ = R − (r̂ · R)r̂. (3.1)
The corresponding length is given by (without loss of generality, we assume that the
centre of the sphere is in the (y, z)-plane)

R⊥ (θ, φ) = (1 − sin2 θ sin2 φ)ρ2R + sin2 θzR
2 − sin 2θ sin φρ z .
R R (3.2)

The force turns out to be the sum of two terms, Q = Q + Q⊥ , each one given by
an integral over the rays forming the incident focused beam. The axial components
are given by (Θ is the Heaviside step function)
  2π
 2γ 2 θ0 −2γ 2 sin2 θ
Qz = 2
dθ sin θ cos θe dφ F (R⊥ )Θ(a − R⊥ ), (3.3)
πA 0 0

2γ 2 θ0
Q⊥ dθ sin θ cos θe−2γ sin θ
2 2
z =
πA 0
 2π
F ⊥ (R⊥ )
× dφ [zR sin2 θ − 12 ρR sin(2θ) sin φ]Θ(a − R⊥ ), (3.4)
0 R⊥
,⊥ ,⊥
where F ,⊥ = 12 (FTE + FTM ), and
r cos(2θ1 ) + cos[2(θ1 − θ2 )]
F = 1 + r cos(2θ1 ) − (1 − r )2 , (3.5)
1 + r2 + 2r cos(2θ2 )
r sin(2θ1 ) + sin[2(θ1 − θ2 )]
F⊥ = r sin(2θ1 ) − (1 − r )2 . (3.6)
1 + r2 + 2r cos(2θ2 )
The r ( = TE, TM) are the Fresnel reflection coefficients for a plane interface, and
θ1 represents the angle of incidence of a given ray at the surface of the sphere (angle
between the ray direction and the normal at the point of incidence). It is given by
R⊥
sin θ1 = . (3.7)
a
θ2 is the angle of refraction inside the sphere,
sin θ1
sin θ2 = . (3.8)
n
Proc. R. Soc. Lond. A (2003)
Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3027

For the transverse components, we find Qϕ = 0, and


 θ0  2π
2γ 2
Qρ = dθ sin2 θ cos θe−2γ dφ sin φF  (R⊥ )Θ(a − R⊥ ),
2
sin2 θ
(3.9)
πA 0 0
 θ0
2γ 2
Q⊥ dθ sin θ cos θe−2γ
2
sin2 θ
ρ =
πA 0
 2π
F ⊥ (R⊥ )
× dφ [ρR (1 − sin2 θ sin2 φ) − 12 zR sin(2θ) sin φ]Θ(a − R⊥ ).
0 R⊥
(3.10)

In the axial case, ρR = 0 and equations (3.2) and (3.7) yield R⊥ = |zR | sin θ and
sin θ1 = |zR | sin θ/a. Hence, in this particular case, the results given by (3.3) and (3.4)
become equivalent to those of Maia Neto & Nussenzveig (2000), whereas the radial
component vanishes.

The factors F and F⊥ in the above expressions can be identified with the longi-
tudinal and transverse components of the force exerted by a given ray with respect
to its direction of propagation r̂(θ, φ). Thus we may also derive (3.3)–(3.10) in the
framework of ray-optics theory. Note that the set of rays that hit the sphere and
thereby contribute to the force is defined by the inequality

R⊥  a, (3.11)

which explains the step function in (3.3), (3.4) and (3.9), (3.10). The remaining
pre-factors, not accounted for previously,† describe the intensity distribution among
different rays in the focused beam, as implied by the Abbe sine condition and the
intensity profile at the entrance aperture of the objective. Indeed, the power dP
contained in a solid angle dΩ = sin θ dθ dφ is given by

−2ρ2
dP = Im exp ρ dρ dφ, (3.12)
w2

where ρ = f sin θ is the distance between the corresponding cylindrical sector of


the beam before the objective and the optical axis. From (3.12) and the Abbe sine
condition, we derive

dP = Im exp(−2γ 2 sin2 θ)f 2 cos θ sin θ dθ dφ. (3.13)

The force exerted by a ray of given power dP and direction r̂(θ, φ) was computed
by Roosen (1979) (note that R⊥ determines the direction perpendicular to the ray
in the plane of incidence, which contains the centre of the sphere):

 ⊥ R⊥ dP
dQ = F r̂ + F . (3.14)
R⊥ P

† Q⊥z , which provides the restoring force along the axial direction, is overestimated by Ashkin (1992),
as a consequence of neglecting the sine condition (cf. Gu et al. 1997) and the corresponding factor cos θ
in (3.4). This diminishes the contribution of rays at large angles.

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3028 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

1.5

1.0

zeq / a
0.5

0
0 5 10 15 20 25
ka
Figure 1. Equilibrium position as a function of β = ka, in units of the sphere radius.
The dashed horizontal line represents the GO value zeq /a = 0.286.

Decomposing the unit vectors in the right-hand side of (3.14) into cylindrical com-
ponents, using (3.13) and integrating over dΩ, we recover, now in the context of
ray-optics theory, the results (3.3), (3.4) and (3.9), (3.10).
The integrals yielding the optical force in the GO approximation can be easily
performed by standard numerical techniques. As discussed in the next section, they
provide a reliable check of the partial-wave numerical results.

4. Numerical results
A remarkable qualitative feature of the exact results, not present in the GO theory,
is the oscillatory behaviour of the force as a function of the size parameter β = ka.
When β  1 and |n − 1|  1, the force, with the sphere centre at the focus, is a
nearly sinusoidal function of the phase ∆ = 4n2 ωa/c associated with radial round-
trip propagation inside the sphere (Maia Neto & Nussenzveig 2000),
Q(R = 0) ≈ 8rcos θ sin2 ( 12 ∆),
where · denotes an average over the intensity distribution of the focused beam
and r = 12 (rTE + rT M )  1. In this approximation, the force is proportional to the
reflectivity of a parallel plate interferometer of optical length 2n2 a. In particular, it
vanishes when the fields back-reflected at the two interfaces interfere destructively.
In this case, the equilibrium position zeq is very close to the focus. More generally,
zeq as a function of β shows oscillatory behaviour, which results from the sinusoidal
variation of the force at R = 0. In figure 1, we plot zeq as a function of β, for a
fixed value of the wavenumber k. The parameters correspond to the experiment of
Ghislain et al . (1994): λ0 = 1.064 µm, A = 0.91, θ0 = 70◦ (angle in water), n1 = 1.32
and n2 = 1.57.†
Very small spheres correspond to the Rayleigh limit, β  1. The force is then
2
proportional to the gradient of Ein , with the dominant contribution coming from the
electric dipole term in (2.8) and (2.13) (Qs is negligible). Hence zeq /a → 0 in this
limit. We see that zeq /a has a peak near β = 1; in this intermediate range, the typical

† We have employed Cauchy’s dispersion formula (Bateman et al. 1959) for computing the refractive
indices of water and polystyrene.

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3029

0.5

0.4

0.3

z−/ a
0.2

0.1

0
0 0.2 0.4 0.6
ρR / a

Figure 2. Equilibrium position with a lateral external force. The plots represent the root z̄ of
Qz (ρR , z̄(ρR )) = 0 as a function of ρR . All lengths are expressed in units of the sphere radius.
Solid line, a = 3.05 µm (β = 23.8); dashed line, a = 3.09 µm (β = 24.1); dotted line, a = 3.15 µm
(β = 24.2); dot-dashed line, GO.

−0.1
−0.035

−0.2 −0.040

−0.045
0.13 0.14 0.15 0.16
−0.3 ρR / a

0 0.2 0.4 0.6


ρR / a

Figure 3. Transverse force Qρ , calculated at the point (ρR , z̄(ρR )) defined in (4.1), as a function
of ρR /a. Qρ is negative, corresponding to a force pointing towards the optical axis. Solid line,
a = 3.05 µm; dashed line, a = 3.09 µm; dotted line, a = 3.15 µm; dot-dashed line, GO.

length-scale of the optical potential well is still partly governed by the wavelength
λ = λ0 /n1 . For larger values of β, we find the oscillatory behaviour arising from
interference. We also show in figure 1 the value computed in the GO approximation
(horizontal dashed line).
Usually, in optical tweezers experiments, a lateral viscous drag force is applied
to the sphere. The sphere is then displaced from the position (ρR = 0, zR = zeq )
and a new equilibrium position is reached at a given distance ρR from the axis. If
the external force has no axial component, the new equilibrium position (ρR , z̄(ρR ))
satisfies
Qz (ρR , z̄(ρR )) = 0. (4.1)

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3030 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig


0.3

0.2

Qmax
0.1

0
0 5 10 15 20 25 30
ka
Figure 4. Maximum transverse force Qmax as a function of β = ka. Theoretical results: par-
tial-wave (‘guide-the-eye’ solid line) and GO value 0.275 (horizontal dashed line). Experimental
results of Ghislain et al . (1994) with reported values of radius (squares), and with rescaled radius
(circles).

250 60

50
κ (pN µm−1)

200 40

30
κ (pN µm−1)

150 20

10

100 0 2 4 6 8 10 12 14
k (µm−1)

50

0 5 10 15 20 25 30
ka
Figure 5. Transverse trap stiffness κ as a function of β = ka, for a fixed wavenumber
k = 7.8 µm−1 and power P = 60 mW. Theoretical results: partial-wave (solid line) and GO
result κ = 541/β pN µm−1 (dashed line). Experimental results of Ghislain et al . (1994) with
reported values of radius (squares), and with rescaled radius (circles). Insert: transverse trap
stiffness as a function of k, for a fixed radius a = 2.0 µm. Solid line, partial-wave results; dashed
horizontal line, GO result κ = 34.8 pN µm−1 .

This equation is solved numerically for z̄ as a function of ρR , for different values


of the size parameter (we define z̄ as the smallest root of (4.1)). As shown in fig-
ure 2, the interference oscillations have only a relatively small effect on the values
of z̄ for values of ρR /a larger than 0.2. For the size parameters of the order of 24
used in the numerical calculation, GO provides a better estimate when consider-
ing the transverse force itself. In figure 3, we plot Qρ (ρR , z̄(ρR )) as a function of
ρR /a, for the same size parameters as figure 2. The force is approximately linear,
and it is mildly affected by the interference effect, as illustrated in the insert of
figure 3.

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3031

There are no roots of (4.1) beyond a specific value of ρR , close to 0.7a in the numer-
ical example considered in figure 2. We define ρmax as the maximum value of ρR .
It corresponds to the maximum external transverse force that can be equilibrated
by the optical force. In figure 4, we plot Qmax ≡ −Qρ (ρmax , z̄(ρmax )) as a function
of the radius,† as well as the GO value Qmax = 0.275. No fitting is employed when
computing the theoretical values, but we apply a global rescaling of the parameter
β to the experimental values reported by Ghislain et al . (1994). Good agreement
is obtained for both Qmax and the transverse trap stiffness (see figure 5), provided
that we divide by a factor of 2 the values of β that result from the data of Ghislain
et al . (1994). We plot the experimental points with (circles) and without (squares)
rescaling in figures 4 and 5. A possible source of this discrepancy (see § 5 for further
discussion) may be the effect of spherical aberration (Ghislain et al . 1994), which
was theoretically analysed in the Rayleigh limit by Yao et al . (2001).
The transverse trap stiffness is given by
n1 P ∂Qρ
κ=− , ρ = 0, z = zeq . (4.2)
c ∂ρ
To calculate κ, we first derive from (2.11) and (2.13) the partial-wave expansion for
∂ρ Q, and then replace it into (4.2). In the GO approximation, Qρ depends on ρ only
through ρ/a. Hence the derivative in (4.2) is proportional to 1/a,
n1 P ∂Qρ (ρ/a = 0) k
κ=− . (4.3)
c ∂(ρ/a) β
In figure 5, we plot the results, for the same parameters employed above, against
the size parameter β, for a fixed wavelength. As before, we take the parameters of
Ghislain et al . (1994), with a power P = 60 mW. There is good agreement with the
GO result κ = (541/β) pN µm−1 for large β, except for a small-amplitude oscilla-
tion,‡ which is related to the interference effect discussed in connection with figure 1.
For very small spheres, on the other hand, the optical force is proportional to the
static polarizability of the dielectric sphere (Rayleigh regime), and hence κ vanishes
as β 3 . The crossover between these two regimes is associated with a peak near β = 3.
The experimental points of Ghislain et al . (1994) also feature a peak, which agrees
with the theoretical prediction, provided that we take, as in figure 4, one half of the
diameter values reported by Ghislain et al . (1994).
It is probably easier to verify experimentally the oscillations displayed in figures 1
and 5, which result from interference, by scanning the laser frequency. In this case,
the expected variation of the trap stiffness is shown in the insert of figure 5, for a
microsphere with a diameter of 2 µm, as a function of wavenumber k. The oscillations

† The numerical integration in (2.9), as well as in the calculation of Gjm and G±


jm , was performed with
the help of a Kronrod–Patterson adaptative Gaussian-type quadrature method. The cut-off value jmax
in the Mie summation was chosen at the point where the relative contribution from additional terms
would be less than 10−10 (taking care to avoid biasing in the addition of terms). In the computation
of the equilibrium position by (4.1), jmax was 46 for β = 15.6 and 64 for β = 31.2. For computation of
transverse force and stiffness, the values of jmax were 37 for β = 15.6 and 55 for β = 31.2. Contributions
decreased rapidly beyond the edge terms j ≈ β. The value of jmax did not depend on ρR in the range of
physical interest (ρR  ρmax ).
‡ The interference oscillations are more important for the axial stiffness, as discussed by Maia Neto
& Nussenzveig (2000).

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3032 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

around the GO value have the expected period δk = π/(2na) = 0.66 µm−1 corre-
sponding to a 2π variation of the round-trip phase ∆. In this example, five periods of
oscillation would be scanned by tuning the vacuum wavelength from 970 to 700 nm,
which is feasible with a Ti:Sapphire laser (trapping with a tunable CW Ti:sapphire
laser has been reported by Neuman et al . (1999)).

5. Conclusion
Partial (qualitative) agreement between wave and GO is found for size parameters
β > 5. However, GO clearly overestimates the maximum transverse force in the
range shown in figure 4, whereas, for the equilibrium position (figure 1) and the trap
stiffness (figure 5), the GO values tend to approach more closely the average of the
partial-wave results over a period of oscillation.
We have not taken into account the spherical aberration of the incident focused
beam, produced by the objective itself (which is corrected for the visible and not
for the infrared wavelength employed by Ghislain et al . (1994)) and by the glass–
water planar interface. Rays at the edge of the focused beam intersect the axis in
the interval between the paraxial (Gaussian) focus and the interface.
Because of spherical aberration, the laser spot size increases with increasing depth
into the specimen chamber, so that both axial and transverse intensity gradients
decrease. Ghislain et al . (1994) reported a decrease of axial stiffness with depth and
Felgner et al . (1995) found a reduction of both axial and transverse maximal trapping
forces. This is in line with figure 4, since the experimental values of the maximum
transverse force lie below the theoretical curve.
The stretching of the focal region by spherical aberration may help us to under-
stand why agreement between theory and experiment was improved by rescaling the
size parameter in figures 4 and 5 (no additional fitting was employed).
Inclusion of spherical aberration in the theoretical treatment thus appears to be
required. We plan to address this question in the context of a comprehensive com-
parison with experimental results based on a new technique for measuring the trap
stiffness (Viana et al . 2002a) and the power available at the sample (Viana et al .
2002b).
We thank Oscar Mesquita, Nathan Viana, Carlos Cesar and Adriana Fontes for valuable dis-
cussions, and Warren Wiscombe for suggestions on numerical integration and Mie scattering
calculations. This work was completed while H.M.N. held a National Research Council Research
Associateship Award at the NASA Goddard Space Flight Center. P.A.M.N. acknowledges sup-
port by CNPq, PRONEX, FAPERJ and the Millenium Institute of Quantum Information. We
thank the José Bonifácio Foundation for its support of the COPEA Optical Tweezers Labora-
tory.

Appendix A. Scattered field


For spherically symmetrical scatterers, it is convenient to work with the Debye poten-
tials Π E and Π M , associated with electric and magnetic multipole fields, respectively
(Born & Wolf 1980). For electric multipole fields, we have

ε
E E = −i∇ × LΠ E , HE = − kLΠ E , (A 1)
µ0

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3033

where L = −ir × ∇ is the differential operator representing the orbital angular


momentum. The Debye potential for a plane wave with wave vector k = k k̂(θk , φk ),
amplitude E0 and polarization σ+ is derived from the corresponding expression for
a plane wave along the ẑ-direction, with the help of the matrix elements of finite
rotations dJM,1 ,


E0  J−1 4π(2J + 1)  −i(M −1)φk J
J
E
Πk(r,θ,φ) = i jJ (kr) e dM,1 (θk )YJM (θ, φ),
k J(J + 1)
J=1 M =−J
(A 2)
where jJ and YJM are the spherical Bessel functions and the spherical harmonics.
According to (2.1), the incident field is a linear superposition of plane waves with k
varying continuously inside a solid angle bounded by θ0 , and with suitable coefficients
corresponding to the intensity distribution and position of the focus. From (2.1)
and (A 2), we derive the Debye potential describing the focused laser beam:


θ0  ∞

E0 4π(2J + 1)
cos θk e−γ
2 2
E sin θk
Πin (r, θ, φ) = dθk sin θk iJ−1 jJ (kr)
k 0 J(J + 1)
J=1


J  2π
× dJM,1 (θk )YJM (θ, φ) dφk eik·R e−i(M −1)φk .
M =−J 0
(A 3)

The integral over φk above is calculated in terms of the cylindrical Bessel function
JM −1 :
 2π
dφk eik·R e−i(M −1)φk = 2π(−i)M −1 eikzR cos θk JM −1 (kρR sin θk )e−i(M −1)ϕR .
0
(A 4)
Once the Debye potentials for the incident field are known, it is a simple matter
to obtain the scattered field by considering the boundary conditions at the surface of
the sphere. ΠsE , representing the electric multipole potential for the scattered field,
is written in terms of the Mie coefficients aJ :

E0  J−M
ΠsE (r, θ, φ) = −2π i GJM (ρR , zR )e−i(M −1)ϕR
k
J,M
4π(2J + 1) (1)
× aJ hJ (kr)YJM (θ, φ), (A 5)
J(J + 1)
(1)
where hJ is the spherical Hankel function and GJM is defined in (2.9). The result
for the magnetic Debye potential is computed in a similar way, in terms of the Mie
coefficients bJ .
From the multipole expansions of the scattered field, we calculate the optical force
using (2.2) (a detailed derivation of the optical force from the Debye potentials was
presented by Farsund & Felderhof (1996)).

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3034 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

Appendix B. WKB approximation for the rotation matrices


The WKB approximation to the rotation matrix djm,n (θ), for j  1, is given by
(Brussaard & Tolhoek 1957)
 1/2
2
djm,n (θ) ≈ [(1 − µ2 )(1 − ν 2 ) − (z − µν)2 ]−1/4
πj
  z 
2 −1
× cos (j + 2 )
1
(1 − x ) [(1 − µ )(1 − ν ) − (z − µν) ] dx ,
2 2 2 1/2

(B 1)
where
m n
z = cos θ, µ= , ν= , (z − µν)2 < (1 − µ2 )(1 − ν 2 ) (B 2)
j j
and the inequality defines the classically allowed (oscillatory) region. The lower limit
of integration was not determined by Brussaard & Tolhoek (1957), except for n = 0,
when the rotation matrix reduces to an associated Legendre function.
Employing this reduction, as well as results for the asymptotic expansion of Legen-
dre functions of large degree and order (Thorne 1957), we find that (B 1), for n = 1,
becomes
1/2
2
j
dm,1 (θ) ≈ (λ2 sin2 θ − m2 )−1/4 cos[fλ,m (θ) + ξλ,m ], (B 3)
π
where
λ≡j+ 1
2 (B 4)
and

λ cos θ m cot θ
fλ,m (θ) ≡ λ arccos √ − m arccos √ . (B 5)
λ 2 − m2 λ 2 − m2
The domain of validity in (B 2) becomes
m+1
|sin( 12 θ)| < . (B 6)
2j
The phase constant ξλ,m in (B 3) results from the unknown lower limit of integra-
tion in (B 1). In order to determine it, we apply the relationship (Gel’fand & Shapiro
1956) 
j(j + 1) sin θ[djm,1 (θ) + djm,−1 (θ)] = −2mdjm,0 (θ), (B 7)
together with the symmetry relation (Brink & Satchler 1968)
djm,n (θ) = (−1)j−m djm,−n (π − θ). (B 8)
Combining these relations for θ = 12 π, we get

djm,1 ( 12 π) = −m[j(j + 1)]−1/2 djm,0 (π/2)



1/2
m 2
≈ (−1)m+1 (λ2 − m2 )−1/4 cos[ 12 π(j − m)], (B 9)
λ π
where we have employed the WKB approximation to djm,0 .

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3035

Comparing (B 9) with (B 3), we get the value of the phase constant:


m
ξλ,m = (m + 34 )π + arccos . (B 10)
λ
A similar procedure can be applied to dj−m,1 (θ). The final WKB results are

1/2 

j m+1 2 2 −1/4 m
dm,1 (θ) ≈ (−1) (λ sin θ − m )
2 2
cos fλ,m (θ) + arccos − 4π ,
1
π λ
(B 11)

1/2 

2 m
dj−m,1 (θ) ≈ (λ2 sin2 θ − m2 )−1/4 cos fλ,m (θ) − arccos − 14 π . (B 12)
π λ
These results appear to be new. They have been numerically tested in several
sample cases, showing very good agreement and consistency with expected WKB
typical features.

Appendix C. WKB approximations to the multipole coefficients


By applying Bessel’s integral representation of Jm−1 , equation (2.9) becomes
Gj,m (ρR , zR )
 2π  θ0
−1
= (2π) dφ dθ sin θH(θ)djm,1 (θ)
0 0
× exp{i[k(ρR sin θ sin φ + zR cos θ) − (m − 1)φ]},
(C 1)
where √
cos θ e−γ
2
sin2 θ
H(θ) ≡ . (C 2)
Substituting djm,1 by its WKB approximation (B 3), we get

Gj,m (ρR , zR )
 2π  θ0
−3/2
= (2π) dφ dθ sin θ(λ2 sin2 θ − m2 )−1/4 H(θ)
0 0
× {exp[ikψλ,m
+
(ρR , zR ; θ, φ)]

+ exp[ikψλ,m (ρR , zR ; θ, φ)]}, (C 3)
where
±
ψλ,m (ρR , zR ; θ, φ) ≡ zR cos θ + ρR sin θ sin φ − ηφ ± [fλ,m (θ) + ξλ,m ]/k, (C 4)

with (m  1)
m−1 m
η≡ ≈ . (C 5)
k k
The asymptotic behaviour of (C 3) in the GO limit, k → ∞, is obtained by apply-
ing the method of stationary phase for double integrals (Born & Wolf 1980). The

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3036 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

stationary phase point (θ̄, φ̄) corresponding to ψ ± is determined by the equations


η m
cos φ̄ = ≈ , (C 6)
ρR sin θ̄ kρR sin θ̄

λ2 sin2 θ̄ − m2
ρR cos θ̄ sin φ̄ − zR sin θ̄ = ± . (C 7)
k sin θ̄
Solving these equations with respect to λ, we find
λ = kR⊥ (θ̄, φ̄), (C 8)
where R⊥ is defined in (3.2). We recognize R⊥ as the impact parameter of a ray
incident along the (θ̄, φ̄) direction, and (C 8) as the well-known semi-classical rela-
tionship between angular momentum and impact parameter (localization principle)
(van de Hulst 1957; Nussenzveig 1992).
The stationary phase points associated with the ± signs in (C 6), (C 7) differ from
each other, for zR = 0, by the transformation zR → −zR , θ̄ → π − θ̄. However,
in (C 1), θ0 < 12 π, so that, for a given sign of zR , only one of the two integrals
in (C 3) has a stationary phase point. The special case zR = 0, which we disregard
for brevity, can be similarly treated.
We also need to evaluate (Born & Wolf 1980) αβ − γ 2 at the stationary phase
point, where
∂2ψ ∂2ψ ∂2ψ
α≡ , β ≡ , γ ≡ ,
∂θ2 ∂φ2 ∂θ∂φ
±
and ψ stands for ψλ,m as defined in (C 4). From now on, we omit the bars in (θ̄, φ̄).
We find
ρR sin θ[ 12 sin(2θ) sin φ(ρ2R − zR2
) + (cos2 θ − sin2 θ sin2 φ)ρR zR ]
αβ − γ 2 = . (C 9)
ρR cos θ sin φ − zR sin θ
Finally, the stationary phase method yields
±i sin θH(θ) exp(ikψ)
Gj,m (ρR , zR ) = √  , (C 10)
2π αβ − γ 2 k(λ 2 sin2
θ − m2 )1/4
where all quantities must be evaluated at the stationary phase point. The double
signs are irrelevant to evaluate the combinations that appear in the partial-wave
series.

Appendix D. Evaluation of Mie coefficient averages


We evaluate here averages of Mie coefficients and their products, which appear in
the partial-wave expansions of Q, taken over a size parameter range sufficient to
eliminate rapid oscillations. We use the generic notation γλ−1/2 to represent either
aj or bj , where λ, defined by (B 4), is related to the angles of incidence θ1 and
refraction θ2 of an incident ray by
λ = β sin θ1 = nβ sin θ2 . (D 1)
The dominant contribution comes from values λ  β, which, jointly with (C 8),
implies
R⊥  a. (D 2)

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3037

The relationship between a given ray direction and θ1 , as given by (3.7), follows
from (C 8) and (D 1).
The WKB approximation to the Mie coefficients is obtained by substituting all
cylindrical functions by their Debye asymptotic expansions. Introducing the notation

χ ≡ β[cos θ1 − ( 12 π − θ1 ) sin θ1 ] − 14 π,
(D 3)
ζ ≡ nβ[cos θ2 − ( 21 π − θ2 ) sin θ2 ] − 14 π,

one finds (Nussenzveig 1992)


 
1 −2iχ T21 T12 e2i(ζ−χ)
γλ−1/2 = 1
2 [1 − S(λ, β)] = 1 − R22 e − , (D 4)
2 1 − R11 e2iζ
where S is the S-function and Rij , Tij are spherical reflection and transmission co-
efficients. In the Debye asymptotic approximation, these coefficients are determined
by the Fresnel reflection amplitude R at a plane interface, for the angle of incidence
θ1 and the corresponding polarization , with  = TE for the magnetic multipole
coefficient bj and  = TM for aj ,

R22 = R = −R11 , T21 = 1+R , T12 = 1−R , T12 T12 = 1−R2 = 1−r ,
(D 5)
where r is the reflectivity.
We also need γλ+1/2 . In order to relate it with γλ−1/2 , we employ recursion relations
among cylindrical functions, with the following result, to dominant order:
 
1 −2i(χ+θ1 ) (1 − r )e2i(θ2 −θ1 ) e2i(ζ−χ)
γλ+1/2 = 1 + R e − . (D 6)
2 1 − R e2i(ζ+θ2 )
While, according to (D 1), θ1 and θ2 are slowly varying functions of β, χ and ζ
are rapidly varying, by (D 3). Accordingly, by defining · to be an average over one
period 2π in ζ (corresponding to a small range in β), we get, to leading order,
γλ−1/2  = 12 . (D 7)
On the other hand, equations (D 4) and (D 6) yield

γλ−1/2 γλ+1/2 
  
1 1
= 1 − r e2iθ1 + (1 − r )2 e2i(θ1 −θ2 ) .
4 1 − r e−2iθ2 + R e2iζ − R e−2iθ2 e−2iζ
(D 8)
The evaluation of all such averages follows from the formula
  

α + β cos(2ζ) + γ sin(2ζ) βb + γc βb + γc 1
= 2 + α− a √ , (D 9)
a + b cos(2ζ) + c sin(2ζ) b + c2 b2 + c2 a − b2 − c2
2

where the average is over a period 2π in the variable 2ζ. The result is
 
∗ 1 (1 − r )2 e2i(θ1 −θ2 )
γλ−1/2 γλ+1/2  = 1 − r e2iθ1 + . (D 10)
4 1 + r e2iθ2

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3038 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

In terms of the definitions (3.5) and (3.6), it follows from (D 10) that
aj a∗j+1 + bj b∗j+1  = 1 − 12 (F  + iF ⊥ ). (D 11)
The ‘cross-polarization’ average a∗λ−1/2 bλ−1/2  will not be needed, because the cor-
responding term in the partial-wave series in (2.7), (2.11) and (2.12) is negligible in
the GO limit.

Appendix E. Evaluation of the GO limit


In all partial-wave sums, the discrete indices (j, m) are converted to continuous vari-
ables (λ = j + 12 , m), converting the sums into integrals. We then change the inte-
gration variables to (θ, φ) by (C 6) and (C 8), so that (cf. (D 2))
  θ0  2π
(·) → dθ dφ |J|Θ(a − R⊥ )(·), (E 1)
j,m 0 0

where the Jacobian J is computed from (3.2), (C 6) and (C 8):


∂(λ, m) ρR
J≡ = k2 sin θ[ 21 (ρ2R − zR2
) sin(2θ) sin φ + ρR zR (cos2 θ − sin2 θ sin2 φ)].
∂(θ, φ) R⊥
(E 2)
Taking into account (C 9), we find
 
 J  k2
 
 αβ − γ 2  = R⊥ |ρR cos θ sin φ − zR sin θ|, (E 3)

and (C 6)–(C 8) also yield



λ2 sin2 θ − m2 = k sin θ|ρR cos θ sin φ − zR sin θ|, (E 4)
so that, finally,
|J| 1
 = . (E 5)
k 2 |αβ
− γ2|
λ sin θ − m
2 2 2 kR ⊥ sin θ

From (C 10) and (2.10), we get, taking (E 5) into account,


∗ cos θ sin θ|H(θ)|2
|J|Gj,m Gj,m = cos θ|JGj,m |2 = , (E 6)
2πkR⊥
so that, with 2j + 1 = 2λ = 2kR⊥ , equation (2.8) leads to
  2π
8γ 2 θ0 dφ
Qez  = dθ cos θ sin θ|H(θ)|2 Θ(a − R⊥ ). (E 7)
A 0 0 2π
Similarly, from (C 10),
sin θ|H(θ)|2
|J|Gj,m G∗j+1,m = exp [ik(ψλ,m − ψλ+1,m )]
2πkR⊥

sin θ|H(θ)|2 ∂
≈ exp −ik ψλ,m
2πkR⊥ ∂λ

sin θ|H(θ)| (λ cos θ − i λ2 sin2 θ − m2 )
2
= √ (E 8)
2πkR⊥ λ 2 − m2

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3039

and 
j(j + 2)(j − m + 1)(j + m + 1)  2
≈ λ − m2 . (E 9)
j+1
The last, cross-polarization term in (2.7) is of order λ−1 times smaller and can be
neglected.
Combining these results with (D 11), we derive, from (2.7),
 θ0
4γ 2
Qsz  = − dθ cos θ sin θ|H(θ)|2
A 
0
 2π  
dφ  λ2 sin2 θ − m2 ⊥
× Θ(a − R⊥ ) 2 − F + F . (E 10)
0 2π kR⊥ cos θ
Adding up (E 7) and (E 10), and taking (E 4) into account, we obtain the final GO
result for Qz as given by (3.3) and (3.4).
By (2.15), G+j,m−1 differs from Gj,m by having an extra sin θ under the integral
sign, as well as by the substitution m → m − 1 in the order of the Bessel function,
which leads to an extra factor e−iφ when replaced in (2.13) and (2.14) (cf. (C 1)),
and similarly for G−j,m+1 . This leads to

sin2 θ
|J|[Gj,m (G− ∗ ∗
j,m+1 ) ∓ Gj,m (Gj,m−1 ) ] =
+
|H(θ)|2 (eiφ ∓ e−iφ ). (E 11)
2πkR⊥
Since  2π
dφ cos φg(sin φ) = 0 (E 12)
0

for any function g, we find (by (3.2), R⊥ depends on φ only through sin φ)
Qeϕ  = 0. (E 13)
We also derive from (E 11)
  2π
8γ 2 θ0 dφ
Qeρ  = dθ sin2 θ|H(θ)|2 Θ(a − R⊥ ) sin φ. (E 14)
A 0 0 2π
On the other hand, similarly to (E 8), we find


∗ sin θ|H(θ)|2 ∂ ∂
|J|Gj,m Gj+1,m+1 = exp −ik + ψλ·m + iφ , (E 15)
2πkR⊥ ∂λ ∂m
and, taking (B 12) into account,
Gj,−m (Gj+1,−m−1 )∗ = Gj,m (Gj+1,m+1 )∗ . (E 16)
Thus, neglecting the cross-polarization terms in (2.12), we find
Qsϕ  = 0. (E 17)
With 
j(j + 2)(j + m + 1)(j + m + 2)
≈ λ + m, (E 18)
j+1

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

3040 A. Mazolli, P. A. Maia Neto and H. M. Nussenzveig

we get, from (2.11), (D 11), (E 15) and (E 16),


 θ0  2π
2γ 2 |H(θ)|2
Qsρ  = − Im dθ dφ Θ(a − R⊥ ) sin θ [2 − (F  + iF ⊥ )]
πA 0 0 R⊥
× {[ρR (1 − sin2 θ sin2 φ)
− 12 zR sin(2θ) sin φ + R⊥ sin θ cos φ]
+ i[R⊥ sin θ sin φ + ρR sin2 θ sin φ cos φ
+ 21 zR sin(2θ) cos φ]}.
(E 19)
Several terms within square brackets in (E 19) are of the form of (E 12) and hence
do not contribute. Thus, adding the results of (E 14) and (E 19), we obtain the
GO result for the transverse force as given by (3.9) and (3.10). This completes the
derivation of the GO approximation from the partial-wave result.

References
Abramowitz, M. & Stegun, I. 1972 Handbook of mathematical functions. New York: Dover.
Ashkin, A. 1992 Forces of a single beam gradient laser trap on a dielectric sphere in the ray
optics approximation. Biophys. J. 61, 569–582.
Ashkin, A. 1997 Optical trapping and manipulation of neutral particles using lasers. Proc. Natl
Acad. Sci. USA 94, 4853–4860.
Barton, J. P., Alexander, D. R. & Schaub, S. A. 1989 Theoretical determination of the net
radiation force and torque for a spherical particle illuminated by a focused laser beam. J.
Appl. Phys. 66, 4594–4602.
Bateman, J. B., Weneck, E. J. & Eshler, D. C. 1959 Determination of particle size and concen-
tration from spectrophotometric transmission. J. Colloid Sci. 14, 308–329.
Berry, M. V. & Mount, K. E. 1972 Semiclassical approximations in wave mechanics. Rep. Prog.
Phys. 35, 315–397.
Bohren, C. F. & Huffman, D. R. 1983 Absorption and scattering of light by small particles.
Wiley.
Born, M. & Wolf, E. 1980 Principles of optics, 6th edn. Oxford: Pergamon.
Brink, D. M. & Satchler, G. R. 1968 Angular momentum, 2nd edn, p. 147. Oxford: Clarendon.
Brussaard, P. J. & Tolhoek, H. A. 1957 Classical limits of Clebsch–Gordan coefficients, Racah
l
coefficients and Dmn (φ, θ, ψ) functions. Physica 23, 955–971.
Edmonds, A. R. 1957 Angular momentum in quantum mechanics. Princeton University Press.
Farsund, Ö. & Felderhof, B. U. 1996 Force, torque and absorbed energy for a body of arbitrary
shape and constitution in an electromagnetic radiation field. Physica A 227, 108–130.
Felgner, H., Müller, O. & Schliwa, M. 1995 Calibration of light forces in optical tweezers. Appl.
Optics 34, 977–982.
Gel’fand, I. M. & Shapiro, Z. Ya. 1956 Representations of the group of rotations of 3-dimensional
space and their applications. Am. Math. Soc. Transl. 2 2, 207–316.
Ghislain, L. P., Switz, N. A. & Webb, W. W. 1994 Measurement of small forces using an optical
trap. Rev. Scient. Instrum. 65, 2762–2768.
Gu, M., Ke, P. C. & Gan, X. S. 1997 Trapping force by a high numerical-aperture microscope
objective obeying the sine condition. Rev. Scient. Instrum. 68, 3666–3668.
Gussgard, R., Lindmo, T. & Brevik, I. 1992 Calculation of the trapping force in a strongly
focused laser beam. J. Opt. Soc. Am. B 9, 1922–1930.

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

Theory of trapping forces in optical tweezers 3041

Maia Neto, P. A. & Nussenzveig, H. M. 2000 Theory of optical tweezers. Europhys. Lett. 50,
702–708.
Mehta, A. D., Rief, M., Spudich, J. A., Smith, D. A. & Simmons, R. M. 1999 Single-molecule
biomechanics with optical methods. Science 283, 1689–1695.
Neuman, K. C., Chadd, E. H., Liou, G. F., Bergman, K. & Block, S. M. 1999 Characterization
of photodamage to Escherichia coli in optical traps. Biophys. J. 77, 2856–2863.
Nussenzveig, H. M. 1992 Diffraction effects in semiclassical scattering. Cambridge University
Press.
Nussenzveig, H. M. & Wiscombe, W. J. 1980 Efficiency factors in Mie scattering. Phys. Rev.
Lett. 45, 1490–1494.
Ren, K. F., Grehan, G. & Gouesbet, G. 1996 Prediction of reverse radiation pressure by a
generalized Lorenz–Mie theory. Appl. Optics 35, 2702–2710.
Richards, B. & Wolf, E. 1959 Electromagnetic diffraction in optical systems. II. Structure of the
image field in an aplanatic system. Proc. R. Soc. Lond. A 253, 358–379.
Roosen, G. 1979 La lévitation optique de sphères. Can. J. Phys. 57, 1260–1279.
Svoboda, K. & Block, S. M. 1994 Biological applications of optical forces. A. Rev. Biophys.
Biomol. Struct. 23, 247–285.
Thorne, R. C. 1957 The asymptotic expansion of Legendre functions of large degree and order.
Phil. Trans. R. Soc. Lond. 249, 597–620.
van de Hulst, H. C. 1957 Light scattering by small particles. Wiley.
Viana, N. B., Freire, R. T. S. & Mesquita, O. N. 2002a Dynamic light scattering from an optically
trapped microsphere. Phys. Rev. E 65, 1–11.
Viana, N. B., Mesquita, O. N. & Mazolli, A. 2002b In situ measurement of laser power at the
focus of a high numerical aperture objective using a microbolometer. Appl. Phys. Lett. 81,
1765–1767.
Wang, M. D., Schnitzer, M. J., Yin, H., Landick, R., Gelles, J. & Block, S. M. 1998 Force and
velocity measured for single molecules of RNA polymerase. Science 282, 902–907.
Yao, X.-C., Li, Z.-L., Guo, H.-L., Cheng, B.-Y. & Zhang, D.-Z. 2001 Effects of spherical aber-
ration on optical trapping forces for Rayleigh particles. Chin. Phys. Lett. 18, 432–434.

Proc. R. Soc. Lond. A (2003)


Downloaded from rspa.royalsocietypublishing.org on May 24, 2011

View publication stats

Вам также может понравиться