You are on page 1of 10

Optical binding via surface plasmon polariton interference

Natalia Kostina1 , Aliaksandra Ivinskaya1 , Sergey Sukhov2 , Andrey Bogdanov1 , Ivan Toftul1,3 ,
Manuel Nieto-Vesperinas4 , Pavel Ginzburg1,5 , Mihail Petrov1,3 , and Alexander Shalin1
1
ITMO University, Department of Nanophotonics and Metamaterials, Saint-Petersburg, 199034, Russia
2
CREOL, University of Central Florida, Tampa, 32816-2700, USA
3
Saint-Petersburg Academic University, Saint-Petersburg, 194021, Russia
4
Instituto de Ciencia de Materiales de Madrid, Consejo Superior de Investigaciones Cientificas,
Campus de Cantoblanco, Madrid E-28049, Spain and
5
School of Electrical Engineering, Tel Aviv University, Tel Aviv, 6997801, Israel
(Dated: August 21, 2017)
Optical binding allows creation of mechanically stable nanoparticle configurations owing to for-
arXiv:1708.05471v1 [physics.optics] 18 Aug 2017

mation of self-consistent optical trapping potentials. While the classical di↵raction limit prevents
achieving deeply subwavelength arrangements, auxiliary nanostructures enable tailoring optical
forces via additional interaction channels. Here, a dimer configuration next to metal surface was
analyzed in details and the contribution of surface plasmon polariton waves was found to govern
the interaction dynamics. It was shown that the interaction channel, mediated by resonant sur-
face waves, enables achieving subwavelength stable dimers. Furthermore, the vectorial structure of
surface modes allows binding between two dipole nanoparticles along the direction of their dipole
moments, contrary to vacuum binding, where a stable configuration is formed in the direction ori-
ented perpendicularly to the polarization of dipole moments. In addition, the enhancement of optical
binding sti↵ness for one order of magnitude was predicted owing to the surface plasmon polariton
interaction channel. These phenomena pave a way for developing new flexible optical manipulators,
allowing for the control over a nanoparticle trajectory on subwavelength scales and opens a room of
opportunities for optical-induced anisotropic, i.e. with di↵erent periods along the field polarization
as well as perpendicular to it, organization of particles on a plasmonic substrate.

INTRODUCTION cal binding can be enhanced by localized plasmon reso-


nance in the nanoparticles8,17,18 . The localized plasmons
can improve trapping efficiency at hot spots of corru-
Light carries momentum which can influence matter gated metal19,20 , or provide particle acceleration against
through optical forces enabling manipulation of micro beam direction in plasmonic V-grooves21 . Three dimen-
and nanoscale objects1 and even atom ensembles2 . The sional structures of plasmonic particles, such as meta-
methods of optical tweezing3,4 rely on attraction of small materials, can be also employed to trap or manipulate
objects to the regions of high field intensity. Spatially nanoparticles, e.g., for realization of optical pulling forces
non-uniform intensity distributions used for positioning attracting nanoparticles to a light source22,23 . In the
microobjects at predefined pattern can be achieved with context of optical binding flat metal surfaces also may
a nanostructured environment or by interference of sev- be very relevant. The excitation of propagating sur-
eral beams. Yet, since early years of optical tweezing ex- face plasmon polartions (SPPs) and induced optical ther-
periments, it was discovered that several particles tend mal forces are responsible for self-organization of micron
to self-organize under homogeneous illumination5,6 . This size nanoparticles24 . Moreover, the direct momentum
e↵ect is referred to as transverse optical binding. The transfer from SPP to micron size particles25,26 can be
interference between incident and scattered light, ow- used for enhancing the optical forces near planar metal-
ing to interaction with particles, results in the forma- lic surface, which can be used for sorting and ordering
tion of a set of potential wells defining stable positions of nanoparticles15,27,28 . Recently, it was suggested that
of particles. Optical binding has been intensively stud- SPP modes can open a way for manipulating the optical
ied both theoretically7–11 and experimentally12–16 also forces acting on nanosize particles by the directional ex-
as a prospective method for self-organization of parti- citation of the propagating SPP modes29–31 .
cles. However, the strength of optical binding drops Here, we propose a new mechanism of transverse opti-
rapidly with nanoparticle size as the scattering efficiency cal binding via excitation of SPP modes (SPP binding)
decreases as ⇠ R6 , where R is the nanoparticle radius. near a metallic planar interface. This mechanism is based
On the other hand, the viscous damping is also reduced on far-field interaction through the interference of SPP
for smaller particles, which makes the fluctuations and waves and is di↵erent to formation of resonant nanoparti-
stochastic processes in liquids to be more influential. As cle molecules due to their near field interaction17,18,32,33 .
a result, for reliable optical control of subwavelength Comparing to common transverse binding in a free space
nanoparticles, strong optical fields are required. In or- (photon binding), the proposed approach has several ad-
der to achieve this without strong heating of the sur- vantages: i) it can enhance the binding e↵ect for small
rounding media, it was suggested to use plasmonic struc- nanoparticles due to resonant excitation of SPP modes;
tures, which can enhance optical fields locally. The opti- ii) the distance between the bounded nanoparticles is de-
2

fined by the SPP e↵ective wavelength and, thus, can be (a) 0.44 Rxx,eff
significantly smaller surpassing the di↵raction limit; iii) sxx,eff
0.42

Re(eff), norm.
the binding occurs in the direction of dipole polarization 
in accordance with the directivity of SPP emission, which 0.4
di↵ers from the case of a free space binding, where stable
configurations are formed in the direction perpendicular 0.38
to the dipole moments. In this paper, we theoretically Im(kSPP)/k0
0.36 0 0.2 0.4 0.6 0.8 1
show how SPP-based transverse optical binding can bring
new features to nanoparticle trapping and manipulation. 700

Vacuum wavelength, nm
0.34
(b)
500

0.06
I. NANOPARTICLES POLARIZATION NEAR A

Im(eff), norm.
SUBSTRATE 300

0.04
100 200 300 400 500 600 700

We consider two identical nanoparticles placed close to 0.02


2/kSPP, nm

a planar metallic interface at coordinates r1 , and r2 in


the field of a normally incident plane wave (see Fig. 1). 0
We assume that nanoparticles have radius R and are 300 400 500 600 700
made of dielectric material with permittivity ". In the Wavelength, nm
dipole approximation the radius of nanoparticles R is
much smaller than the typical scale of electric field varia- FIG. 2. The spectrum of real (a) and imaginary (b) parts of
s
tions. In this limit, the optical force acting on a nanopar- the xx-components of e↵ective polarizability tensor ↵ˆ ef f (blue
R
ticle is given by the expression34 dashed line) and ↵ˆ ef f (red solid line) shown along with vac-
uum polarizability ↵ (black dash-dot line). The polarizability
1 X ⇤ is normalized over 4⇡"0 R3 , and calculated for a nanoparticle
F= Re pi rEi (r, !), (1) of radius R = 15 nm with " = 3 whose center is located above
2 i the surface at z = 20 nm. The inset shows the dependence of
the SPP typical period denoted as LSP P = 2⇡/kSP P (upper
where Ei (r, !) is the i-th component of a local field. x-axis), and imaginary part of SPP wavevector (lower x-axis)
on the vacuum wavlength.
Einc(r)
G0(r2,r1) particles via free-space and substrate channels, and self-
k z
induced contribution of each particle through the reflec-
  tion from the substrate. The local field is given by

1=1 r1 y
r2 Gs(r2,r2)
k02 b k2 b
E(r) = E0 (r) + G(r, r1 )p1 + 0 G(r, r2 )p2 . (3)
"0 "0
0 x
Gs(r2,r1) Here, the first term in the right hand side is the am-
plitude of the external field, the second and third terms
s correspond to the field generated by the first and second
nanoparticles respectively. The total Green’s function
b r0 ) = G
G(r, b 0 (r, r0 ) + G b s (r, r0 ) is a sum of the scattered
FIG. 1. The scheme of the problem. Nanoparticles with
permittivity " are positioned at equal distances from a surface b b
Gs and vacuum G0 components respectively35 . One can
z1 = z2 = z. We assume that the permittivity of the upper simplify the consideration if renormalizes the polarizabil-
half-space equals to "1 = 1. ity tensor with respect to the self-action Green’s function
component G b s (ri , ri ) term.
The dipole moment of a nanoparticle p(r) is defined ✓ ◆
s k02 b
as p(r) = ↵(!)E(r), where ↵(!) is the vacuum dipole pi = ↵ bi,ef f E0 (ri ) + G(ri , rj )pj , (4)
polarizability corrected with account for retardation ef- "0
fects: i = 1, 2 j = 2, 1.

1 1 ik03 " "1 Here, we have introduced the e↵ective polarizability ten-
= , ↵0 = 4⇡"0 R3 , (2) sor ↵ s
bi,ef f as follows:
↵ ↵0 6⇡"0 " + 2"1
✓ ◆ 1
where k0 is the wavevector in a free space, and "0 is the s k02 b
bi,ef
↵ (r
f i , !) = ↵(!) 1 ↵(!) Gs (ri , ri , !) , i=1,2.
vacuum permittivity. The local electric field includes the "0
incident plane wave, multiply rescattered field between (5)
3

This tensor gives a correction of a vacuum polarizability nanoparticle using the expression (1) (see the details in
↵(!) with accounting for nanoparticle self-action through the Appendix Eq. (A13)). In the following, we will re-
the substrate. This tensor is diagonal as G b s (ri , ri ) is fer to the optical force acting on the second nanoparticle
diagonal in the case of a flat isotropic substrate35 . only, fixing the first nanoparticle in the coordinate origin.
For the sake of simplicity, in the following we fix the In order to find the equilibrium positions of nanoparti-
position of the first particle in the origin of the coordinate cle, we plot the dependence of the x-component of the
system at x1 = 0 and y1 = 0, and will consider the force optical force as a function of interparticle distance along
acting on the second particle only. Computing the field x-axis as shown in Fig. 3 a). The force is normalized over
at the point of the dipole according to the expression in the optical pressure force acting on the same nanoparti-
Eq. (3), one can achieve a system of equations for dipole cle in vacuum F0 = 1/2k|E0 |2 Im(↵(!)), where |E0 | is
moments p1 and p2 (see Appendix A), and, in the special the amplitude of the incident plane wave. One can see
case of normal incidence of the plane wave, the expression that the force changes at particular point on the x-axis
for the dipole moments can be simplified even further: denoting the equilibrium positions points. These points
can be stable along x if the force is restoring (shown
R
pi = ↵ bi,ef f E0 (ri ), (6) with solid circles, i.e. point 1), and unstable otherwise
✓ (shown with white filled circles). One should also note
R k2 b
bi,ef
↵ f (ri , !) = ↵(!) 1 ↵(!) G s (ri , ri , !) that when nanoparticles are close to each other the force
"0 goes to minus infinity, until the nanoparticles touch each
◆ 1
k4 b i , rj )b s b other. However this case is out of the scope of the present
↵(!) G(r ↵ (r
j,ef f j , !) G(r ,
j ir ) ⇥ (7) paper.
"20
✓ ◆
k2 b s To identify the role of plasmons in the interaction force
1 + G(ri , rj )b ↵j,ef f (rj , !) ,
"0 we have excluded SPP contribution from the Green’s
i = 1, 2 j = 2, 1. function by integrating over the free space modes only in
the spectral representation (see Appendix B). One can
Now, the polarizability ↵ R
bef see that in the absence of SPPs the interaction force be-
f (see the Appendix A for
the details) includes all the interaction channels: (i) the comes one order of magnitude weaker, and the period
self-action of nanoparticle through the substrate, and (ii) between stable positions is significantly enlarged being
the cross-action of two nanoparticle via vacuum and sub- defined by the vacuum wavelength and photons interfer-
strate. Moreover, it is worth mentioning that though the ence. Moreover, the equilibrium points shown with blue
e↵ective polarizability tensor ↵ s
bef circles are stable both along x and y directions making
f is diagonal, the ten-
R them globally stable, which does not happen in case of
sor ↵bef f is non-diagonal as the presence of the second
photon binding. To illustrate this, we plotted the Fy
nanoparticle does not preserve translational symmetry
force (see Fig. 3 b)) as function of the transverse angle
of the system.
(see the inset in Fig. 3) in the vicinity of points of stable
The excitation of SPP modes a↵ects both the e↵ective
equilibrium positions. In Fig. 3 (c) one can see the depen-
polarizability due to the substrate mediated self-action,
dence of the binding length on the vacuum wavelength
and cross action of nanoparticles. The spectra of real and
R shown along with the period of SPP wave, which is equal
imaginary parts of xx - components of ↵ bef f (solid line)
s to LSP P . One can see that the binding distance is fully
and ↵ bef f (dashed line) are plotted in Fig. 2 for the case defined by the period of the SPP wave when the excita-
of silver substrate. The vacuum polarizability ↵ is also tion condition is fulfilled, thus, providing the binding at
shown in the figure with dash-dot line. One can see that distances significantly shorter than vacuum wavelength.
the e↵ective polarisabilities have resonance at around 350 This also strongly di↵ers from work of Salary et. al.17 ,
nm. From the inset of Fig. 2, one can see that this wave- where the optical forces between two nanoparticles over
length corresponds to SPP resonant excitation for sil- metallic substrate were considered in the regime, when
ver/vacuum interface, which is defined by the condition the interaction force is mainly defined by the near-field
Re("s (!)) + 1 = 0 and also corresponds to maximal
p value components.
of real part of SPP wavevector kSP P = k0 "s /("s + 1).
In the inset the e↵ective wavelength of SPP mode defined In order to support the results discussed above, we
as LSP P = 2⇡/kSP P is also shown. The strong enhance- have performed numerical simulations in COMSOL Mul-
ment of the imaginary part of the e↵ective polarizability typhysics package (see Fig. 3 (d), scatter line) and showed
is a sign of strong rescattering of light into the SPP mode. good correspondence with the obtained solution based
on Green’s function approach (see Fig. 3 (d) solid line).
The Green’s function approach shows good agreement
II. DETERMINING THE STABLE with numerical results also at the distances comparable
CONFIGURATIONS to nanoparticle size. Moreover, basing on Green’s func-
tion formalism we have derived the approximate expres-
By determining the dipole moments of nanoparti- sion for the contribution of SPP mode into the optical
cles, one can calculate the optical force acting on each force (see Appendix C for the details):
4

E
(b)
(a) 6 1 k y

1

x-component of force Fx/F0


4

y-component of force Fy/F0


0.6
2 x
1 x20
0.2
0

-2 2
With SPP -0.2
Without SPP
2
-4
Stable equilibrium -0.6
-6
x10 Unstable equilibrium
-8 -1
0 100 200 300 400 500 600 700 800 -20 -10 0 10 20
x - coordinate, nm Angle , deg.

(c) 500 (d)


LSPP

x-component of force Fx/F0


Equilibrium 0
400
distance
Distance, nm

300 -20 Green’s function approach

Exact numerical simulation

200 -40 Analytical expression

100 -60
300 330 360 390 100 200 300 400
Wavelength, nm x - coordinate, nm

FIG. 3. (a) The x-component of optical force plotted along the x-axis coinciding with the direction of electric field polarization.
The blue solid line denotes the force with account for all interaction channels. The red dashed line is for the interaction
through the free space photons only. (b) The y-component of optical force in the direction perpendicular to x-axis, showing the
stability of the binding position in the direction transverse to the x-axis. The blue solid line and red dashed line correspond
to force calculated with or without account for SPP interaction channel for the equilibrium positions labeled 1 and 2 in a)
correspondingly. The results are shown for the wavelength = 350 nm. (c) The distance between the stable equilibrium
positions obtained from (a) compared with the distance L = 2⇡/kSP P . (d) The comparison of the optical force calculated
within Green’s function approach (same as in Fig.3 (a)) and calculated numerically with COMSOL Multyphysics package. The
approximate analytical expression given by the Eq. (8) for SPP induced force is also shown with dashed line. All the results
shown in the figure are computed for R = 15 nm and z = 25 nm.

along x-axis36 . This di↵erence of SPP and photon bind-


3 2
 ing can be understood through the di↵erence in the scat-
kSP P k1z k2z tering diagrams of SPPs and photons. This is illustrated
Fx ⇡ ⇡|px |2 Re 2 ⇥
k0 (1 "s ) in Fig.4 where two-dimensional maps of x and y force
i
(1)
H1 (kSP P x) exp( Im(k1z )z) . (8) components are plotted. The photon binding is well-
known to have stable configuration perpendicular to the
p p field polarization direction in accordance with the dipole
Here k1z = k02 kSP 2
P , and k2z = "s k02 kSP
2
P emission pattern (see Fig. 4 (a, b)). The SPP binding, on
are z-components of SPP wavevector in the upper half- the contrary, has stable configurations along the polar-
(1)
space and in the substrate correspondingly, H1 (q) is the ization direction, in which preferable excitation of SPP
first order Hankel function of the first kind. The derived modes occurs (see Fig. 4 (c, d)). It is also worth not-
expression very illustratively shows the origin of the SPP ing that the amplitudes of lateral forces are several times
mode: the Hankel function describes the SPP mode ex- higher when SPP modes a↵ect binding.
cited by a dipole and propagating over a flat surface. Its
zeros define the equilibrium positions of the nanoparti- We illustrate the character of SPP binding by calcu-
cle. The z-components of the wavevector is complex since lating the dynamics of the second nanoparticle motion
SPP is a localized wave, thus, the exponent in Eq. (8) in the force field of the first nanoparticle, which is fixed
shows that strength of dipole-SPP coupling decays. at the origin of the coordinates. We consider only two-
One needs to stress, that the transverse binding in vac- dimensional motion of the nanoparticle, keeping the z-
uum does not not provide stable equilibrium positions coordinate to be constant. The dynamics is obtained
5

(a) (b) 0.8 Attraction force, a.u.


0.6
300 60
0.4 y
Fy 0.4
Without SPP

0.2 Fx E
200 50
0 0
0 x
-0.2 0
-0.4 100 40
-0.4

y, nm
-0.6 -0.8 30
0

5 (c) (d) 2
-100 20
3
Fx Fy 1
With SPP

1 -200 10
0
-1
-300 0
-1 -300 -200 -100 0 100 200 300
-3 Time, ms
x, nm
-5 -2
-800 -400 0 400 800 (nm) FIG. 5. The dynamics of the second nanoparticle motion.
The first nanoparticle is fixed at the origin. The colour of
FIG. 4. The two-dimensional maps showing the x and y the trajectory line denotes the time elapsed since the begin-
forces for SPP and photon binding. The di↵erent direction- ning of motion. The arrows show the force field: the darker
ality of the scattering pattern is responsible for di↵erent ge- are the arrows the stronger is the optical force. The colour
ometry of stable equilibrium positions. The results are shown map at the background shows the force which attracts or re-
for wavelength = 350 nm and z = 25 nm. pulses nanoparticle to/from the equilibrium positions at y = 0
and x = ±175 nm. The intensity of the red colour gives the
strength of nanoparticle attraction, while blue shows repul-
sion of the nanoparticle. The parameters of computation are
through direct solution of equations of motion under the R = 15 nm, z = 25 nm, " = 3. The laser intensity was taken
external optical force with the account for viscous damp- 5·105 W/m2 , and the dimensionless damping factor = 0.015
ing. The details are discussed in Appendix D. Two typi- (see Appendix D for the details of the simulation method).
cal trajectories are shown in Fig. 5 for two di↵erent sets
of initial coordinates of the second nanoparticle. The
color map shows the intensity of nanoparticle attraction is enabled. We have plotted (see Fig. 6) the spectral
to the equilibrium positions along x-axis. The arrows dependence of sti↵ness parameter x calculated at the
show the force field, while the lines show the trajectories first equilibrium point, labeled by point 1 in Fig. 3 (a).
with colour changing from blue to red while time elapses. To avoid the dependence of sti↵ness on the illumina-
One can see that the nanoparticle actively tends to set tion intensity, we have normalized it to the magnitude
the position along the x-axis where the binding force is 0 = F0 /R, which is the sti↵ness of a system where the
the strongest. vacuum pressure force F0 can be restored when nanopar-
ticle is displaced for one radius from its equilibrium posi-
The important parameter, which characterizes the sta-
tion. One can see that the sti↵ness has a strong resonant
bility of the equilibrium states, is the sti↵ness of the
behaviour, which corresponds to the excitation of SPP
trap. At the equilibrium positions the total optical force
modes at wavelengths longer than 350 nm. With the
is zero, but when shifted from the stable positions the
increase of the distance from nanoparticle center to the
nanoparticles undergo action of a restoring force, which
surface the sti↵ness rapidly drops, as the coupling with
is locally proportional to the amplitude of the displace-
the SPP mode decreases.
ment Fr = x x, with the parameter x characterizing
From Fig. 6 one can see that the spectral maximum
the sti↵ness of the system along the x-direction. How-
of sti↵ness depends on the height from the surface. This
ever, this approximation of the restoring force only ap-
spectral dependence can be understood better by ana-
plies to the gradient component of the optical force. In-
lyzing the analytical expression Eq. (8). In the case of
deed, we consider the nanoparticles significantly smaller
neglible losses in the substrate, one can get a simple ex-
than the wavelength, that results in low and non-resonant
pression for the sti↵ness at the first stable equilibrium
at the wavelengths imaginary part of the polarizabil-
point (see Appendix C for more details):
ity Im(↵ef f ) ⌧ Re(↵ef f ), as Im(↵ef f ) ' (R6 / 3 ), and
R << (see Fig. 2). Thus, the radiation force, which
is proportional to imaginary part of the polarizability, (kSP P )3 |k1z |2 |k2z |
can be neglected (see Appendix E). The sti↵ness in the x ⇡ ⇡|px |2 Y2 (q1 ) exp( |k1z |z). (9)
k02 (1 "s )
considered system strongly depends on the mechanism of
the nanoparticles interaction, and, as can be seen from Here Y2 (q) is the cylindrical Webber function of the sec-
Fig. 3, it is much higher when the plasmon interaction ond order, and q1 is the first positive root of Y1 (q1 ) = 0.
6

Note that in the regime of SPP excitation without ohmic III. CONCLUSION
losses the SPP wavevector can be in a range from k0 to
+1 when 1+"s ! 0. With that the expression Eq. (C12) In this work we consider transverse optical binding
goes to zero in both limiting cases: based on surface plasmon polariton interference. We
show that two nanoparticles placed in the vicinity of a
plasmonic interface can form a stable bound dimer with
x ! 0, binding length defined by the SPP wavelength. This al-
kSP P !0,1 lows formation of the dimers with interparticle distance
significantly shorter than the free-space wavelength sup-
pressing the di↵ractional limit. The binding states are
which implies that the sti↵ness reaches its maximum at formed along the direction of the incident field polariza-
some particular wavelength. This wavelength can be de- tion, which on the contrary tothe photon binding, where
fined for each given distance over substrate z. The max- the stable bound states are formed perpendicular to the
imal sti↵ness can be achieved close to SPP resonance polarization direction. The excitation of SPP modes also
when SPP wavevector equals to k̃ ⇡ 6/z k0 . Then, enhances the amplitude of the binding forces, resulting
the maximal sti↵ness at the n-th equilibrium position de- in resonant enhancement of the trap sti↵ness.
creases with distance to substrate as z 6 (see Eq. (C13)): Appendix A: Calculation of a binding force

The force F2 (r2 ) acting on the second nanoparticle is


✓ ◆6 calculated as
2 1 1
x,n ⇠ |px | Y2 (q2n+1 ). (10)
z k02 1 hX ⇤ i
F2 (r2 ) = Re p2i (r2 )rEi (r2 ) . (A1)
2
The introduction of the e↵ective polarizability parame-
ters significantly simplifies the formula for the electric
field in the center of the second nanoparticle E2 :
4
z=15 nm k02 b k2 b
3.5 E(r2 ) = E0 (r2 )+ Gs (r2 , r2 )p2 + 0 G(r 2 , r1 )p1 . (A2)
"0 "0
3
Stiffness, norm.

2.5 The dipole moments then can be expressed as:


2 s
z=20 nm b1,ef
p1 = ↵ f E(r) =
1.5 z=25 nm ✓ ◆
s k02 b
1
z=35 nm b1,ef f E0 (r1 ) + G(r1 , r2 )p2 ,
↵ (A3)
z=45 nm "0
✓ ◆
0.5
s k02 b
p2 = ↵b2,ef f E0 (r2 ) + G(r2 , r1 )p1 , (A4)
0
340 360 380 400 420 440 "0
Wavelength, nm ✓ ◆ 1
s k02 b
bi,ef
↵ (r
f i , !) = ↵(!) 1 ↵(!) G (r ,
s i i r , !) ,
FIG. 6. The sti↵ness x in units of 0 = F0 /R as a function "0
of the excitation wavelength. The spectra are shown for dif- (A5)
ferent distances z from the nanoparticle center to the surface. i = 1, 2.
Nanoparticle radius R = 15 nm.
Let us solve this equation system. After some manipula-
tions one can get

✓ ✓  ◆◆
s k02 b s k02 b
p1 = b1,ef
↵ E0 (r1 ) + G(r1 , r2 ) ↵
f b2,ef f E0 (r2 ) + G(r2 , r1 )p1 , (A6)
"0 "0
✓ ◆
s k02 b s k04 s b 1 , r2 )b s b
b1,ef
p1 = ↵ f E0 (r1 ) + G(r1 , r2 )b
↵2,ef f E0 (r2 ) + 2 ↵ b G(r ↵2,ef f G(r2 , r1 )p1 , (A7)
"0 "0 1,ef f

The last expression in (A6) can be simplified even fur- ther, if one renormalizes the e↵ective polarizability tensor
with account for nanoparticle cross action:
7
✓ ◆ 1
r k2 b k4 b s b
bi,ef
↵ f (ri , !) = ↵i (!) 1 ↵i (!) G s (ri , ri , !) ↵i G(ri , rj )b
↵j,ef f G(rj , ri ) , i = 1, 2 j = 2, 1. (A8)
"0 "20

Here the self-action Green’s function G b s (ri , ri ) contains


the scattered part only, whereas the cross-action part Z 1
b i , rj ) = G
G(r b 0 (ri , rj ) + G
b s (ri , rj ) include both vacuum b ', z > 0) = ik1
G(⇢, c(s, ⇢, ') exp (isz1 z) ds,
M
and scattered parts determining the cross-interaction 8⇡ 0

through vacuum and via substrate respectively. The final (B1)


expression for the dipole moment will be as follows: where k1 is the wavevector in the upper space, and
s = kr /k0 and sz1 = kz1 /k0 are the radial and z-
✓ ◆ components of the dimensionless wavevector normalized
r k2 b s
pi = bi,ef
↵ f E0 (ri ) + G(ri , rj )b↵j,ef f E0 (rj ) , (A9) over the wavevector in the free space.
"0
i = 1, 2 j = 2, 1 (A10) 0 1
mxx mxy mxz
c(s, ⇢, ') = @ myx
M myy myz A (B2)
The case of normal plane wave incidence on a planar
substrate, when the nanoparticles are located at the same mzx mzy mzz
height above the surface (see Fig.1) is of a particular s
mxx = rs f (s, ⇢, ') ssz1 rp g(s, ⇢, '),
interest. In this case the external electric field E0 is equal sz1
in the centres of both nanoparticles, and, thus, the dipole s
myy = rs g(s, ⇢, ') ssz1 rp f (s, ⇢, '),
moment has very simple form: sz1
s3
mzz = 2⇡J0 (s⇢)rp ,
sz1
R
bi,ef
pi = ↵ f E0 (ri ), (A11) (rs + s2z1 rp )
✓ 2
◆ mxy = myx = h(s, ⇢, '),
R r k b s ssz1
bi,ef
↵ f bi,ef f 1 + G(ri , rj )b
=↵ ↵j,ef f , (A12)
"0 mxz = mzx = srp t(s, ⇢, '),
i = 1, 2 j = 2, 1. myz = mzy = srp w(s, ⇢, '),

The optical force component, then, can be calculated where the functions f (s, ⇢, '), g(s, ⇢, '), h(s, ⇢, '),
as t(s, ⇢, '), w(s, ⇢, ') can be expressed:
" # ✓ ◆
1 X J1 (s⇢)
F2x (r2 ) = Re ⇤
p2n (r2 )@x2 En (r2 ) = f (s, ⇢, ') = 2⇡ sin2 (')J0 (s⇢) + cos(2') ,
2 s⇢
n=x,y,z ✓ ◆
" J1 (s⇢)
X ⇣ g(s, ⇢, ') = 2⇡ cos2 (')J0 (s⇢) cos(2') ,
1 s⇢
Re p⇤2n (r2 ) @x2 E0n (r2 ) +
2 n=x,y,z h(s, ⇢, ') = ⇡s2 J2 (s⇢) sin(2'), (B3)
k02 X t(s, ⇢, ') = 2⇡i sJ1 (s⇢) cos('),
@x02 Gs,nm (r02 , r2 )p2m
"0 w(s, ⇢, ') = 2⇡i sJ1 (s⇢) sin(').
m=x,y,z
!#
k2 X Here Jn (z) is the first kind Bessel function of the order
+ 0 @x2 Gnm (r2 , r1 )p1m . (A13) n.
"0 m=x,y,z

The y and z components can be calculated with the Appendix C: Analytical expression
same expression (A13) by substituting the partial deriva-
tive with @y and @z correspondingly. Here we analyze the x-component of the optical force
acting on nanoparticles when normal incident light is po-
ralized along the x-axis. This is the case considered in
Appendix B: Green’s function Fig. 3. According to Eq.(1) the expression for the force
will be as follows:
The Green’s function tensor of a two half-spaces with
permittivities "1 (for z > 0) and "2 (for z  0) can be ex- !
1 X
pressed in the cylindrical coordinates through the recip- Fx = Re pi (r)⇤ @x Ei (ri ) . (C1)
rocal representation in wave-vector space35 (for z > 0): 2 i
8

In order to get a simple analytical result showing all the which gives us
key features of the SPP-assisted force, we will take into Z 1
account that the e↵ective ↵ˆ R tensor has diagonal domi- ik1
@x Gs,xx (⇢, 0, z) = k0 s2 s1z rp (s)⇥ (C7)
R
nation, which implies that ↵ii R
↵ij , i 6= j. 4⇡
✓ ◆0
The the expression (C1) can be simplified: J2 (s⇢)
J1 (s⇢) exp(is1z z)ds. (C8)
1 (s⇢)2
Fx = Re(p⇤x @x Exs ). (C2)
2 In order to compute the integral with help of com-
The electrical field generated by the dipole at the distance plex analysis, we first continue the integral bounds to
r from the first nanoparticle can be expressed through the 1, +1 using of the identity:
Green’s function Exs (r) = 4⇡k02 Gs,xx (r, 0)px . 1 (1)
Then we have an expression for the lateral component Jn (q) = (H (q) ( 1)n Hn(1) ( q)).
2 n
of the optical force written in a very simple form:

Fx = 2⇡k02 |px |2 Re (@x Gs,xx (r, 0)) (C3) ik1


k0 ⇥
@x Gs,xx (⇢, 0, z) =
The Greens function is expressed through the integral 8⇡ !
Z 1 (1)
(1) H2 (s⇢)
s2 s1z rp (s) H1 (s⇢) exp(is1z z) ds.
ik1 1 (s⇢)
Gs,xx (x, y, z) = Gs,xx (⇢, , z) = 2
⇥ (C4) | {z }
Z 1 8⇡ I(s)

mxx (⇢, s) exp(is1z z)ds, (C9)


0
a1 (⇢, ) Now, using Cauchy theorem we finally evaluate this
mxx = srs (s) ss1z rp (s)a2 (⇢, ), integral:
s1z
✓ ◆
2 J1 (s⇢) ik1
a1 = 2⇡ sin( 0 ) J0 (s⇢) + cos(2 0 ) , @x Gs,xx (⇢, 0, z) = k0 2⇡iRes(I(s))|s=s̃ =
s⇢ 8⇡ !
✓ ◆ (1)
J1 (s⇢) k1 (1) H2 (s̃⇢)
a2 = 2⇡ cos( 0 )2 J0 (s⇢) cos(2 0 ) , k0 (s̃)2 s̃1z H1 (s̃⇢) ⇥ (C10)
s⇢ 4 (s̃⇢)
p p
s1z = 1 s2 s2z = "2 s2 . exp(is̃1z z)Res(rp (s))|s=s̃ ,
p
Here we use the same notation as in Appendix B. We where s̃ = "2 "1 /("1 + "2 ) is dimensionless wavevector
are interested only in the component containing rp term of SPP mode.
as only it gives rise to SPP response, and also we put Finally, computing the explicit expression for the
= 0‘. Then, residue and substituting the obtained results into
Eq.(C3) one can get:
Z 1 "
ik1
Gs,xx (⇢, , z) = m0xx (⇢, s) exp(is1z z)ds, (C5) (k̃)3 (k̃1z )2 k̃2z ⇣ (1)
8⇡ 2 0 Fx = ⇡|px |2 Re H1 (k̃x)
0 k02 ("1 "2 )
mxx = ss1z rp (s)a2 (s, ⇢), ! #
✓ ◆ (1)
J1 (s⇢) H2 (k̃x)
a2 = 2⇡ J0 (s⇢) . exp(ik̃1z z) .
s⇢ (k̃x)2

Next, we have Here we use dimension variables denoting k̃ = s̃k0 ,


and k̃z1,z2 = s̃z1,z2 k0 . We can go even further taking
(1)
ik1 into account that |H1 (k̃x)| ⌧ |J2 (k̃x)/(k̃x)2 |:
Gs,xx (⇢, , z) = ⇥
" #
Z 1 ✓ ◆4⇡ (k̃)3 (k̃1z )2 k̃2z (1)
J1 (s⇢) 2
Fx ⇡ ⇡|px | Re H (k̃x) exp(ik̃1z z) .
ss1z rp (s) J0 (s⇢) exp(is1z z)ds. (C6) k02 ("1 "2 ) 1
0 s⇢
(C11)
With this we need to compute @x Gs,xx :
The case of low losses is of special interest. Then, the
@x J0 (s⇢) = k0 @⇢ J0 (s⇢) = k0 sJ1 (s⇢) final expression for the force can be reduced to:

J1 (s⇢) J1 (s⇢) J2 (s⇢) (k ⇤ )3 |k1z


⇤ 2 ⇤
| |k2z |
@x = k0 s@s⇢ = k0 s Fx = ⇡|px |2 2 Y1 (k ⇤ r) exp( k1z

z),
s⇢ s⇢ (s⇢) k0 ("1 "2 )
9

where Y1 (q) is the cylindrical Webber function. By ex- After such substitutions we have
panding this expression around the zeros qn of the Web-
ber function Y1 (q) ⇡ Y2 (qn )(q qn ), one can find the
d2 X ⇢ ✓ X
⇤ @ e0i + e 2 e s (⇠2 , ⇠2 )e
expression for the sti↵ness at the n-th equilibrium posi- ⇠ 2 = Re e
p 2i E k G ij p2j +
tion along x-axis of the system (see Fig. 3 a)): d⌧ 2 i
@⇠ j
X ◆
+k e2 e
Gij (⇠2 , ⇠1 )e
p1j . (D3)
(k̃)3 |k̃1z |2 |k̃2z |
n ⇡ ⇡|px |2 Y2 (q2n 1 ) exp( |k̃1z |z). j
k02 ("1 "2 )
(C12) We also include the viscosity of the environment by
Note that in the regime of SPP excitation without ohmic adding the damping factor :
losses the SPP wavevector can be in the range from k0
to +1 when "1 + "2 ! 0. With that the expression
Eq. (C12) goes to zero in both limiting cases: d2 e 2 (⇠1 , ⇠2 ) d
⇠2 = F ⇠2 , (D4)
d⌧ 2 d⌧
n ! 0,
k̃!0,1
where F e 2 is given by the r.h.s. of Eq. (D3). Expression
which implies that the sti↵ness reaches its maximum
(D4) was a target for the numerical simulation. As a
at some particular wavelength. This wavelength can be
good compromise between stability and computational
defined for each given distance over substrate z. The
complexity the Runge-Kutta of fourth-order method was
maximal sti↵ness can be achieved close to SPP resonance
applied. Due to the fact that the motion along z axis is
when SPP wavevector equals to k̃ ⇡ 6/z. Close to the
fixed, we have plane symmetry, which simplifies the force
frequency of SPP resonance when k̃ ! 1, the SPP be- e 2 (⇠1 , ⇠2 ) = F
e 2 (⇠1 ⇠2 ).
function to F
comes highly localized close to the interface |k̃z | k0 .
Then, the maximal sti↵ness can be expressed as:
✓ ◆6
2 6 1
n ⇠ |px | Y2 (q2n+1 ) exp( 6). (C13) Appendix E: Conservative vs non-conservative force
z k02 "1
components

Appendix D: Dynamics simulation Here we present the results of calculation of total op-
tical force and the conservative component only. By ex-
We write down the equation of the Newton’s equation cluding imaginary part of polarizability one can isolate
for the second particle the conservative force only35 . The result are shown in
Fig. 7. One can see that for the considered set of pa-
d2 rameters the conservative force strongly dominates over
m r2 = F2 ,
dt2 non-conservative, which is the the di↵erence between the
where F2 is given by (A1) and (A2). One can rewrite it total and conservative forces.
as:
⇢ ✓ 8 Total force
d2 1 X ⇤ k2 X Gradient component
r 2 = Re p 2i r E 0i + Gs,ij (r2 , r2 )p2j +
dt2 2m i "0 j 6
x-component of force Fx/F0

4
(D1)

k2 X 2
+ Gij (r2 , r1 )p1j , i, j = x, y, z
"0 j 0

-2
where Gb=G b0 + Gb s . In order to decrease the numerical
error during numerical simulations, we apply the follow- -4

ing natural scaling: -6


100 150 200 250 300 350 400
x-coordinate, nm
r t e= E, G eb b
⇠= , ⌧= , E = aG, (D2)
a T E0 FIG. 7. The total force (blue solid line) and the conservative
e = ak, p p ↵
k e= , ↵e= , (red circles) components of the optical force are shown for
4⇡"0 a3 E0 4⇡"0 a3 di↵erent wavelengths. The parameters of the calculation are
r
m the same as in Fig. 2.
T = .
2⇡"0 aE02
10

1 20
O. M. Maragò, P. H. Jones, P. G. Gucciardi, G. Volpe, R. Quidant and C. Girard, Laser and Photonics Reviews
and A. C. Ferrari, Nature Nanotechnology 8, 807 (2013). 2, 47 (2008).
2 21
I. Bloch, J. Dalibard, and S. Nascimbène, Nature Physics A. S. Shalin and S. V. Sukhov, Plasmonics 8, 625 (2013).
22
8, 267 (2012). A. S. Shalin, S. V. Sukhov, A. A. Bogdanov, P. A. Belov,
3
A. Ashkin, Physical Review Letters 24, 156 (1970). and P. Ginzburg, Physical Review A 91, 063830 (2015).
4 23
V. S. Letokhov, JETP Letters 7, 348 (1968). A. A. Bogdanov, A. S. Shalin, and P. Ginzburg, Scientific
5
M. M. Burns, J. M. Fournier, and J. A. Golovchenko, Reports 5, 15846 (2015), arXiv:arXiv:1506.01754v1.
24
Physical Review Letters 63, 1233 (1989). V. Garces-Chavez, R. Quidant, P. J. Reece, G. Badenes,
6
M. M. Burns, J. M. Fournier, and J. A. Golovchenko, L. Torner, and K. Dholakia, Physical Review B - Con-
Science (New York, N.Y.) 249, 749 (1990). densed Matter and Materials Physics 73, 1 (2006).
7 25
F. Depasse and J. M. Vigoureux, Journal of Physics D: G. Volpe, R. Quidant, G. Badenes, and D. Petrov, Phys-
Applied Physics 27, 914 (1994). ical Review Letters 96, 1 (2006).
8 26
P. C. Chaumet and M. Nieto-Vesperinas, Physical Review M. Yuan, L. Cheng, P. Cao, X. Li, X. He, and X. Zhang,
B 64, 035422 (2001), arXiv:0305045 [physics]. Plasmonics , 1 (2017).
9 27
J. Kesava, P. C. Chaumet, T. Langtry, and A. Rahmani, Z. Yan, Y. Bao, U. Manna, R. A. Shah, and N. F. Scherer,
Journal of Nanophotonics 4, 041583 (2010). Nano Letters 14, 2436 (2014).
10 28
M. Mazilu, A. Rudhall, E. M. Wright, and K. Dholakia, C. Min, Z. Shen, J. Shen, Y. Zhang, H. Fang, G. Yuan,
Journal of Physics: Condensed Matter 24, 464117 (2012). L. Du, S. Zhu, T. Lei, and X. Yuan, Nature communica-
11
S. Sukhov, A. Shalin, D. Haefner, and A. Dogariu, Optics tions 4, 2891 (2013).
29
express 23, 247 (2015). S. B. Wang and C. T. Chan, Nature communications 5,
12
R. W. Bowman and M. J. Padgett, Reports on progress 3307 (2014).
30
in physics. Physical Society (Great Britain) 76, 026401 M. I. Petrov, S. V. Sukhov, A. A. Bogdanov, A. S. Shalin,
(2013). and A. Dogariu, Laser and Photonics Reviews 10, 116
13
M.-T. Wei, J. Ng, C. T. Chan, and H. D. Ou-Yang, Sci- (2016), arXiv:1601.01863.
31
entific Reports 6, 38883 (2016). A. Ivinskaya, M. I. Petrov, A. A. Bogdanov, I. Shishkin,
14
L. Chvatal, O. Brzobohaty, and P. Zemanek, Optical Re- P. Ginzburg, and A. S. Shalin, Light: Science & Applica-
view 22, 157 (2015). tions 6, e16258 (2017).
15 32
V. Demergis and E. L. Florin, Nano Letters 12, 5756 F. J. Valdivia-Valero and M. Nieto-Vesperinas, Optics ex-
(2012). press 20, 13368 (2012).
16 33
S. H. Simpson, P. Zemánek, O. M. Maragò, P. H. Jones, J. M. Auñón, F. J. Valdivia-Valero, and M. Nieto-
and S. Hanna, Nano Letters 17, 3485 (2017). Vesperinas, Journal of the Optical Society of America. A,
17
M. M. Salary and H. Mosallaei, Physical Review B - Con- Optics, image science, and vision 31, 206 (2014).
34
densed Matter and Materials Physics 94, 1 (2016). P. C. Chaumet and M. Nieto-Vesperinas, Optics Letters
18
C. Van Vlack, P. Yao, and S. Hughes, Physical Review B 25, 1065 (2003), arXiv:0305043 [physics].
35
- Condensed Matter and Materials Physics 83, 1 (2011), L. Novotny and B. Hecht, Principles of Nano-Optics
arXiv:1102.0794. (Cambridge University Press, 2012) p. 578.
19 36
M. L. Juan, M. Righini, and R. Quidant, Nature Photonics K. Dholakia and P. Zemánek, Reviews of Modern Physics
5, 349 (2011). 82, 1767 (2010).

You might also like