Arrested Dimer's Diffusion by Self-Induced Back-Action Optical Forces

Jun 3, 2016 - The diffusion of a dimer made out of two resonant dipolar scatters in an optical lattice is theoretically analyzed. When a small particl...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Article

Arrested dimer's diffusion by self-induced back-action optical forces Jorge Luis-Hita, Juan Jose Saenz, and Manuel I Marqués ACS Photonics, Just Accepted Manuscript • DOI: 10.1021/acsphotonics.6b00259 • Publication Date (Web): 03 Jun 2016 Downloaded from http://pubs.acs.org on June 7, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Photonics is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Arrested dimer’s diffusion by self-induced back-action optical forces Jorge Luis-Hita,∗,†,‡ Juan José Sáenz,∗,‡,¶ and Manuel I. Marqués∗,§ †Departamento de Física de la Materia Condensada, Universidad Autónoma de Madrid, 28049 Madrid, Spain ‡Donostia International Physics Center (DIPC), Paseo Manuel de Lardizabal 4, 20018 Donostia-San Sebastian, Spain ¶IKERBASQUE, Basque Foundation for Science, 48013 Bilbao, Spain. §Departamento de Física de Materiales, Condensed Matter Physics Center (IFIMAC) and Instituto “Nicolás Cabrera”, Universidad Autónoma de Madrid, 28049 Madrid, Spain E-mail: [email protected]; [email protected]; [email protected]

Abstract The diffusion of a dimer made out of two resonant dipolar scatters in an optical lattice is theoretically analyzed. When a small particle diffuses through an optically induced potential landscape, its Brownian motion can be strongly suppressed by gradient forces, proportional to the particle’s polarizability. For a single lossless monomer at resonance, the gradient force vanishes and the particle diffuses as in absence of external fields. However, we show that when two monomers link in a dimer, the multiple scattering among the monomers induces both a torque and a net force on the dimer’s center of mass. This “self-induced back-action” force leads to an effective potential energy landscape, entirely dominated by the mutual interaction between monomers, which strongly influences the dynamics of the dimer. Under appropriate illumination,

1

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

single monomers in a colloidal suspension freely diffuse while dimers become trapped. Our theoretical predictions are tested against extensive Langevin molecular dynamics simulations.

Keywords Resonant dimer, Self-induced back-action optical forces, Arrested diffusion, Potential energy landscapes, Langevin Molecular Dynamics

Understanding Brownian motion in spatially periodic and random landscapes has long been of great interest from a fundamental and practical point of view. (1 –5 ) The ability of light to exert significant forces on small particles (6 ) offers the opportunity to sculpt potential energy profiles enabling the study of Brownian dynamics in complex landscapes. Extended periodic landscapes, known as optical lattices, can be generated by the periodic intensity maxima arising in the interference pattern of several crossed laser beams (7 –9 ) or by an array of optical tweezers (10 ,11 ) generated by holographic techniques. (12 ) Tracking the Brownian motion of particles in a single optical trap can be used to study non-equilibrium diffusion processes, (13 –15 ) to detect acoustic vibrations (16 ) or to map the accessible space by Photonic Force Microscopy. (17 ) Optical forces have been widely exploited to understand, manipulate and control the statistical properties (18 ,19 ) and dynamics (20 ,21 ) of colloidal particles by optically induced potential energy landscapes (PEL). (22 ) Randomly modulated intensity patterns, so-called speckle patterns, can also be used to create a random PEL. (23 ,24 ) Counter intuitive phenomena like the giant diffusion induced by an oscillating periodic potential (25 ) or kinetically locked-in states in driven diffusive transport (26 ) have been realized by using optically induced PEL, (27 ,28 ) demonstrating optical guiding and sorting of particles in microfluidic flows. (29 –31 ) The intriguing properties of Brownian motion under 2

ACS Paragon Plus Environment

Page 2 of 23

Page 3 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

periodic landscapes of non-conservative “curl forces” (optical vortex lattices (32 –34 ) ) is also a subject of increasing interest (35 –40 ) . Although the Brownian motion of non-interacting particles through PEL is relatively well understood, the influence of multiple scattering effects on the Brownian dynamics has been much less studied. Active superdiffusion can be realized in colloidal suspensions where the particles move in a time-dependent, self-generated, speckle intensity pattern induced by multiple scattering of light. (41 ) Multiple scattering effects can induce significant interactions between particles leading to optical binding (8 ,42 ) with peculiar non-conservative effects. (43 ,44 ) Attractive or repulsive Casimir-like forces can be controlled by using rapidly fluctuating random light fields (45 ) . Optical binding manifest itself in the extent of Brownian fluctuations of particles optically linked. (46 ) Multiple scattering not only leads to the formation of aggregates but also modifies the Brownian motion of the aggregates themselves. Multiple scattering effects are behind the complex behavior of the driven diffusion of dimers and other optically bound structures experimentally observed in a “tractor beam” set-up. (47 ) Our main goal here is to understand the role of multiple scattering in the diffusion of aggregates in an optical lattice. In order to understand the interplay between the optically induced PEL and multiple scattering in the diffusion of dimers or other aggregates we have analysed a simple system consisting of a dimer made out of two resonant particles diffusing on a two-dimensional optical lattice. Among other possible physical realisations, direct fabrication of metal nanoparticle dimers of a given length could be feasible by using DNA-origami and related techniques. (48 –50 ) In an optical potential energy landscape, the electromagnetic forces on small particles are conservative gradient interactions (33 ,51 ) proportional to the real part of the particle’s polarizability. At resonance, the real part of the polarizability is negligible and, as a consequence, a single resonant nanoparticle does not see the underlying optical lattice and undergoes free thermal Brownian motion. However, as we will show, when two resonant monomers link into a dimer, multiple scattering among them induces both a torque and a

3

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

Page 4 of 23

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

as sketched in Fig. 1. The external illuminating electric field is given by E (0) (x, y, z) = i2E0 (sin kx + sin ky) eˆz = iE0 Ψ(x, y) eˆz

(1)

where k = nω/c = 2π/λ is the wave number, λ is the wavelength in water, n the refractive index, E0 is the amplitude of the electric field, Ψ is a real function and eˆz is the unitary vector in the z-direction. For a single electric dipolar particle moving through a field E = Ez eˆz , linearly polarized along z, the time averaged optical force is given by (51 ) ǫǫ0 Re {α∗ Ez∗ ∇Ez } 2   ǫǫ0 α′ 2 ′′ ∗ ∇|Ez | + α Im {Ez ∇Ez } . = 2 2

F =

(2) (3)

At resonance we have α = i 6π , i.e. α′ = 0, and the total force on a single monomer moving k3 through an external field given by Eq. (1) is identically zero: there are no forces coming from the gradient force term (proportional to α′ ) and the “scattering force” (proportional to o n (0)∗ (0) = 0) vanishes. Im Ez ∇Ez

However, when a dimer is formed, the actual field that polarises each monomer comes

not only from the external field but also from the radiation scattered by the other dipole . This scattered light contributes to the optical force and may alter some dynamical properties of the system such as the diffusion. The optical polarising fields on particles 1 and 2 are expressed as E(r1 ) = E (0) (r1 ) + k 2 G(r1 , r2 )αE(r2 )

(4)

E(r2 ) = E (0) (r2 ) + k 2 G(r2 , r1 )αE(r1 ) where G(ri , rj ) is the Green dyadic function and E (0) the external field. Since the motion of the dimer in our system is restricted to the xy-plane and the external electric field is along 5

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 23

the z-axis, the only relevant component of the Green dyadic function is eikL Gzz (ri , rj ) = Gzz (L) = 4πL being L = |r1 − r2 | =

p



1 1 1+i − 2 2 kL k L



(5)

(x1 − x2 )2 + (y1 − y2 )2 the distance between the particles. Defining,

G12 ≡ Gzz (|r1 − r2 |) = G21 , the system of equations (4) admits a simple analytic solution given by (0)

(0)

Ez (r1 ) =

Ez (r1 ) + ik 2 α′′ G12 Ez (r2 ) 1 + [k 2 G12 α′′ ]2

Ez (r2 ) =

Ez (r2 ) + ik 2 α′′ G21 Ez (r1 ) , 1 + [k 2 G12 α′′ ]2

(0)

(6)

(0)

where we considered the resonant condition α′ = 0. Then, the field polarising the particles always points along z and, from Eq. (2), the force on the resonant monomers “1” and “2” is given by ǫǫ0 ′′ α Im {Ez∗ (r1 )∇1 Ez (r1 )} 2 ǫǫ0 ′′ α Im {Ez∗ (r2 )∇2 Ez (r2 )} F2 = 2 F1 =

(7)

where Ez∗ (r1 ) and Ez∗ (r2 ) are given by Eq. (6), ∇1 ≡ {∂/∂x1 , ∂/∂y1 } and ∇1 Ez (r1 ) = ∇1 Ez(0) (r1 ) + iα′′ k 2 ∇1 G12 Ez (r2 )

(8)

∇2 Ez (r2 ) = ∇2 Ez(0) (r2 ) − iα′′ k 2 ∇1 G12 Ez (r1 ) (where we took into account that ∇2 G12 = −∇1 G12 ). These expressions will be used to perform molecular dynamics simulations discussed below. Notice that they are exact only within the dipolar approximation which, for resonant nanoparticles, is known to be valid for L larger than approximately three times the monomer radius. (57 ) At smaller inter-particle separations, higher order multipoles become relevant in the computation of the forces. 6

ACS Paragon Plus Environment

Page 7 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

Analytical approach In order to get some insight into the problem, let us consider Eqs. (6) and (8) in the limit k 2 G12 α′′ = 3λG12 ≪ 1 ( a good approximation for L & λ/2), Ez (r1 ) ≈ Ez(0) (r1 ) + ik 2 G12 α′′ Ez(0) (r2 ) + · · · ∇1 Ez (r1 ) ≈ ∇1 Ez(0) (r1 ) + iα′′ k 2 ∇1 G12 Ez(0) (r2 ) + · · · , and analogous expressions for monomer “2”. At lowest order in [k 2 G12 α′′ ], the forces [Eq. (7)] are given by n o ǫǫ0 2 ′′ 2 k (α ) |E0 |2 [Ψ(r1 )Ψ(r2 )] ∇1 Re {G12 } − [Ψ(r2 )∇1 Ψ(r1 )] Re {G12 } + · · · (9) 2 n o ǫǫ0 2 ′′ 2 k (α ) |E0 |2 − [Ψ(r1 )Ψ(r2 )] ∇1 Re {G12 } − [Ψ(r1 )∇2 Ψ(r2 )] Re {G12 } + · · · F2 = 2

F1 =

and we can readily calculate the force on the center of mass:

Fcm = F1 + F2 n o ǫǫ0 2 ′′ 2 2 ∼ − k (α ) |E0 | [Ψ(r2 )∇1 Ψ(r1 )] + [Ψ(r1 )∇2 Ψ(r2 )] Re {G12 } + · · · 2

(10)

It is interesting to consider Fcm when the external field is identical on both nanoparticles, (0)

(0)

i.e. Ez (r1 ) = Ez (r2 ). In this particular case, Eqs. (6) become (0)

Ez (r1 ) = Ez (r2 ) =

Ez (r1 ) , 1 − ik 2 G12 α′′

(11)

and it is possible to obtain a simple, and exact, closed expression for Fcm :

Fcm = −

n o ǫǫ0 ∗ |E0 |2 Re {αeff } Ψ(r1 ) ∇1 Ψ(r1 ) + ∇2 Ψ(r2 ) 2

7

ACS Paragon Plus Environment

(12)

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 23

where

αeff ≡

iα′′ 1 − ik 2 G12 α′′

(13)

can be understood as an effective dimer’s polarizability. The force could then be understood as a result of a shift in the dimer resonance that leads to a real part in the effective polarizability. However, this description is not strictly valid to describe the actual forces at arbitrary dimer positions and orientation. We shall then use Eq. (10) to obtain information about the dynamical behavior of the dimer: First, the force on the center of mass is equal to zero if Re {G12 } vanishes. For this particular situation, the diffusion of the dimer is purely Brownian and the existence of an external electromagnetic field is not going to modify the dynamics. The real part of the Green function is written as     1 1 1 cos kL 1 − 2 2 − sin kL . Re{G12 } = 4πL k L kL

(14)

First two zeros of equation (14) are given by L = 0.72λ and L = 1.2λ. So, for dimers of those lengths, we expect to find free Brownian dynamics. In second place, we can also obtain information about the behavior of the dimer when the force on the center of mass is different from zero. We will describe the dimer configuration in terms of its center-of-mass coordinates (x, y) and an angular coordinate θ describing the angle of the dimer with respect to the x-axis. The relationship between the monomer’s coordinates and the center of mass is given by

r1 = r + ∆rθ

,

r2 = r − ∆rθ

∆rθ = (

L L cos θ, sin θ). 2 2

if we consider now a fixed orientation of the dimer with respect to the electromagnetic field 8

ACS Paragon Plus Environment

Page 9 of 23

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 23

Langevin molecular dynamics simulations. First, we must write the equation of motion for the dimer by transforming the coordinates of the monomers(x1 , y1 , x2 , y2 ) into the dimer coordinates (x, y, θ). Then, the equations of motion are given by d2 x dx = −2γ + ξx1 + ξx2 + Fx d2 t dt d2 y dy 2m 2 = −2γ + ξy1 + ξy2 + Fy dt dt 2 d θ dθ mL2 2 = −γ L2 + dt dt 2m

+ L(ξx2 + Fx2 − ξx1 − Fx1 ) sin(θ) + L(ξy1 + Fy1 − ξy2 − Fy2 ) cos(θ) being γ the friction coefficient (γ = 6πaη, η is the viscosity, η = 0.89 × 10−3 kg m−1 s−1 for water at T = 298K and we consider particles with a = 50nm radius) and ξ(t) a thermal force with zero mean and variance given by the fluctuation-dissipation theorem

hξi (t)ξj (t′ )i = 2γkB T δij δ(t − t′ )

(18)

with i and j equal to (x1 , y1 , x2 , y2 ) and T being the system’s temperature. F , F1 and F2 are the optical forces on the center of mass and on each particle respectively obtained through the exact expressions for the fields and their derivatives [with λ = 387.585nm wavelength in water. The refractive index is n = 1.8]. Notice that for the parameters used in the Langevin simulations the condition that the dimer length is larger than three times the particle radius is fulfilled (dipolar limit condition). When no optical force is applied the dimer follows diffusive dynamics with mean squared displacement given by < r2 (t) >= 4D0 t

10

ACS Paragon Plus Environment

(19)

Page 11 of 23

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 23

dimer’s dimension. Note that, since each isolated monomer feels no optical force at all, the decrease in the value of the diffusion constant is due only to the non symmetric scattering interaction among the monomers. That is, the interaction force exerted on particle 1 by 2 is not reciprocal to the interaction force exerted on particle 2 by 1. We already found in the literature a situation in which action-reaction in optical forces is not fulfilled (44 ) for a dimer made up of two different particles. In our case both particles are identical and the asymmetry in the interactions comes from the external field.

(a)

(b)

Figure 4: (a) Typical trajectories of the dimer’s center of mass in x, y ∈ [0, λ) obtained from Langevin molecular dynamics simulations for L = 0.6λ and P = 0.3 × 105 W/cm2 . Polar plots of the dimer orientation probability distribution at specific points “A” to “E” are also shown. (b) The same as in (a) but for for L = 0.84λ and P = 0.8 × 105 W/cm2 . The mean square displacement, the reduction of the diffusion constant and the strong dependence of the main statistical features of the dimer’s dynamics on the dimer’s length L, can be understood in terms of the potential energy landscape obtained in Eq. (17). First, we have performed statistics of the data obtained from the Langevin molecular dynamics simulations to obtain the distribution function for the dimer orientation with respect to the field (θ). Figure 4 shows typical Brownian trajectories of the dimer’s center of mass together with a polar plot of the dimer orientation at specific points for two different dimer lengths. 12

ACS Paragon Plus Environment

Page 13 of 23

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(a)

Page 14 of 23

(b)

Figure 6: (a) Probability distribution for the dimer’s center of mass in x, y ∈ [0, λ) obtained from Langevin molecular dynamics simulations for L = 0.6λ and P = 0.3 × 105 W/cm2 . Darker regions correspond to a higher probability (the integral of the distribution over the area of the figure is normalized to 1). Typical Brownian trajectories from “A” to “E” are shown in Fig. 4a. The contour map shows the underlying field intensity landscape [contour lines 2 correspond to 0.15, 0.30, 0,45, 0.60, 0.75 and [red line] 0.90 E (0) max ]. (b) Potential energy landscape −Uθ (x, y)/(kB T ) from Eq. (17) for a fixed angle θ = 0 or θ = π/2 corresponding to the preferential angles observed in Fig. 5a. Potential minima are located in the darker regions.

14

ACS Paragon Plus Environment

Page 15 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

It is interesting to note that the relative positions of the dimer, i.e. the potential minima, are not trivially linked to the underlying field intensity map (see Fig. 6a). We should emphasise that this potential landscape is felt by the system only when the two particles are linked together in a dimer. For the isolated dipoles there is no energy landscape at all.

(a)

(b)

Figure 7: (a) Probability distribution for the dimer’s center of mass in x, y ∈ [0, 2λ) obtained from Langevin molecular dynamics simulations for L = 0.84λ and P = 0.8 × 105 W/cm2 . Darker regions correspond to a higher probability. Typical Brownian trajectories from “A” to “E” [or “A” to “C ”] are shown in Fig. (4b). The contour lines show the underlying field intensity landscape as in Fig. 6a. (b) Potential energy landscape −Uθ (x, y)/(kB T ) from Eq. (17) for fixed angles θ = π/4 and θ = π/4 + π/2 corresponding to the preferential angles observed in Fig. 5b. Although the translational Brownian motion can be understood as a diffusive motion in an effective potential energy landscape, this simplified description does not capture the interplay between forces and exerted torques in the Brownian motion of the dimer. The coupling between rotational and translational diffusion processes can be understood by analysing the orientation distributions along the typical paths between trapping positions [indicated by the A,B,C,D,E spots in Figs. 4a and 6a]. Our results show that rotational and translational diffusion processes are clearly locked since dimer’s orientation changes as it moves along the path as shown by the orientation distributions in B,C and D. The results for a different length L = 0.84λ are sumarized in Figs. 4b and 7. Again, 15

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the effective potential landscape minima, with θ = π/4 + nπ/2, capture the most probable locations of the center of mass. However, in contrast with the shorter dimer, each trapping position corresponds now to a single preferred orientation (which is not captured by the simplified effective potential). In this case, the rotational and translational locking is clear since the dimer has to change its orientation as it moves from a trapping position to the next one (see trajectories A-B-C and A-D-E in Fig. 4b).

Conclusions In conclusion, two resonant lossless particles, moving free on an optical lattice, become arrested when they couple to form a dimer. Although our discussion focused on dimers made out of resonant lossless monomers, we expect a similar behaviour for actual dimers of plasmonic nanoparticles at resonance, where the real part of the polarizability can be neglected. The predicted dramatic change on the dynamical properties is due to the self induced back action forces (SIBA forces) arising as a consequence of the non symmetric multiple scattering between the monomers. In first approximation, this interaction results in a potential energy landscape that only prevails when the two particles are linked. The properties of this landscape, including wells depth and position, strongly depend on dimer’s length, making it suitable as an active sorting tool.

Acknowledgement The authors thank financial support from Spanish Ministerio de Economía y Competitividad ( MICINN) project FIS2012-36113-C03 and FIS2015-69295-C3-3-P and the “María de Maeztu” Program Ref: MDM-2014-0377.

16

ACS Paragon Plus Environment

Page 16 of 23

Page 17 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

References (1) Risken, H. The Fokker-Planck Equation; Springer Berlin Heidelberg, 1984. (2) Bouchaud, J.-P.; Georges, A. Anomalous diffusion in disordered media: statistical mechanisms, models and physical applications. Phys. Rep. 1990, 195, 127–293. (3) Sahimi, M. Flow Phenomena in Rocks - from Continuum Models to Fractals, Percolation, Cellular-Automata, and Simulated Annealing. Rev. Mod. Phys. 1993, 65, 1393–1534. (4) Reimann, P. Brownian motors, noisy transport far from equilibrium. Phys. Rep. 2002, 361, 57–265. (5) Hanggi, P.; Marchesoni, F. Artificial Brownian motors: Controlling transport on the nanoscale. Rev. Mod. Phys. 2008, 81, 387–442. (6) Ashkin, A. Acceleration and Trapping of Particles by Radiation Pressure. Phys. Rev. Lett. 1970, 24, 156–159. (7) Chowdhury, A.; Ackerson, B. J.; Clark, N. A. Laser-induced freezing. Phys. Rev. Lett. 1985, 55, 833–836. (8) Burns, M. M.; Fournier, J. M.; Golovchenko, J. a. Optical matter: crystallization and binding in intense optical fields. Science 1990, 249, 749–754. (9) Hemmerich, A.; Hänsch, T. Two-dimesional atomic crystal bound by light. Phys. Rev. Lett. 1993, 70, 410–413. (10) Ashkin, A.; Dziedzic, J.; Bjorkholm, J.; Chu, S. Observation of a single-beam gradient force optical trap for dielectric particles. Opt. Lett. 1986, 11, 288–290. (11) Jones, P.; Maragó, O.; Volpe, G. Optical tweezers: Principles and applications; Cambridge University Press, 2015. 17

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(12) Polin, M.; Ladavac, K.; Lee, S.-H.; Roichman, Y.; Grier, D. Optimized holographic optical traps. Opt. Express 2005, 13, 5831–5845. (13) Roichman, Y.; Sun, B.; Roichman, Y.; Amato-Grill, J.; Grier, D. G. Optical forces arising from phase gradients. Phys. Rev. Lett. 2008, 100, 013602. (14) Wu, P.; Huang, R.; Tischer, C.; Jonáš, A.; Florin, E. L. Direct measurement of the nonconservative force field generated by optical tweezers. Phys. Rev. Lett. 2009, 103, 108101. (15) Gieseler, J.; Quidant, R.; Dellago, C.; Novotny, L. Dynamic Relaxation of a Levitated Nanoparticle from a non-Equilibrium Steady State. Nat. Nanotechnol. 2014, 9, 358– 364. (16) Ohlinger, A.; Deak, A.; Lutich, A. A.; Feldmann, J. Optically Traped Gold Nanoparticle Enables Listening at the Microscale. Phys. Rev. Lett. 2012, 108, 018101. (17) Rohrbach, A.; Tischer, C.; Neumayer, D.; Florin, E.-L.; Stelzer, E. H. Trapping and tracking a local probe with a photonic force microscope. Rev. Sci. Instrum. 2004, 75, 2197–2210. (18) Bechinger, C.; Brunner, M.; Leiderer, P. Phase behavior of two-dimensional colloidal systems in the presence of periodic light fields. Phys. Rev. Lett. 2001, 86, 930–933. (19) Mikhael, J.; Roth, J.; Helden, L.; Bechinger, C. Archimedean-like tiling on decagonal quasicrystalline surfaces. Nature 2008, 454, 501–504. (20) Zurita Sánchez, J. R.; Greffet, J. J.; Novotny, L. Friction Forces Arising from Fluctuating Thermal Fields. Phys. Rev. A 2004, 69, 022902. (21) Šiler, M.; Zemánek, P. Particle jumps between optical traps in a one-dimensional (1D) optical lattice. New J. Phys. 2010, 12, 083001.

18

ACS Paragon Plus Environment

Page 18 of 23

Page 19 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(22) Evers, F.; Hanes, R. D. L.; Zunke, C.; Capellmann, R. F.; Bewerunge, J.; DalleFerrier, C.; Jenkins, M. C.; Ladadwa, I.; Heuer, A.; Castañeda Priego, R.; Egelhaaf, S. U. Colloids in light fields: Particle dynamics in random and periodic energy landscapes. Eur. Phys. J 2013, 222, 2995–3009. (23) Volpe, G.; Kurz, L.; Callegari, A.; Volpe, G.; Gigan, S. Speckle optical tweezers: micromanipulation with random light fields. Opt. Express 2014, 22, 18159–18167. (24) Bewerunge, J.; Egelhaaf, S. U. Experimental creation and characterization of random potential-energy landscapes exploiting speckle patterns. Phys. Rev. A 2016, 93, 013806. (25) Reimann, P.; Van den Broeck, C.; Linke, H.; Hänggi, P.; Rubi, J. M.; Pérez-Madrid, A. Giant acceleration of free diffusion by use of tilted periodic potentials. Phys. Rev. Lett. 2001, 87, 010602. (26) Reichhardt, C.; Nori, F. Phase Locking, Devil’s Staircases, Farey Trees, and Arnold Tongues in Driven Vortex Lattices with Periodic Pinning. Phys. Rev. Lett. 1999, 82, 414–417. (27) Korda, P. T.; Taylor, M. B.; Grier, D. G. Kinetically locked-in colloidal transport in an array of optical tweezers. Phys. Rev. Lett. 2002, 89, 128301. (28) Lee, S.-H.; Grier, D. G. Giant colloidal diffusivity on corrugated optical vortices. Phys. Rev. Lett. 2006, 96, 190601. (29) MacDonald, M. P.; Spalding, G. C.; Dholakia, K. Microfluidic sorting in an optical lattice. Nature 2003, 426, 421–424. (30) Ladavac, K.; Kasza, K.; Grier, D. G. Sorting mesoscopic objects with periodic potential landscapes: optical fractionation. Phys. Rev. E 2004, 70, 010901. (31) Xiao, K.; Grier, D. G. Sorting colloidal particles into multiple channels with optical forces: Prismatic optical fractionation. Phys. Rev. E 2010, 82, 051407. 19

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(32) Hemmerich, A.; Hänsch, T. Radiation pressure vortices in two crossed standing waves. Phys. Rev. Lett. 1992, 68, 1492–1495. (33) Albaladejo, S.; Marqués, M. I.; Laroche, M.; Sáenz, J. J. Scattering Forces from the Curl of the Spin Angular Momentum of a Light Field. Phys. Rev. Lett. 2009, 102, 113602. (34) Gómez-Medina, R.; Nieto-Vesperinas, M.; Sáenz, J. J. Nonconservative electric and magnetic optical forces on submicron dielectric particles. Phys. Rev. A 2011, 83, 033825. (35) Albaladejo, S.; Marqués, M. I.; Scheffold, F.; Sáenz, J. J. Giant Enhanced Diffusion of Gold Nanoparticles in Optical Vortex Fields. Nano Lett. 2009, 9, 3527–3531. (36) Zapata, I.; Albaladejo, S.; Parrondo, J. M. R.; Sáenz, J. J.; Sols, F. Deterministic ratchet from stationary light fields. Phys. Rev. Lett. 2009, 103, 130601. (37) Albaladejo, S.; Marqués, M. I.; Sáenz, J. J. Light Control of Silver Nanoparicle’s Diffusion. Opt. Express 2011, 19, 11471–11478. (38) Berry, M. V.; Shukla, P. Physical Curl Forces: Dipole Dynamics near Optical Vortices. J. Phys. A 2013, 46, 422001. (39) Berry, M. V.; Shukla, P. Hamiltonian Curl Forces. P. Roy. Soc. Lond. A 2015, 471, 20150002. (40) Zapata, I.; Delgado-Buscalioni, R.; Sáenz, J. J. Control of diffusion of nano-particles in an optical vortex lattice. arXiv preprint arXiv:1510.08664 2015, (41) Douglass, K. M.; Sukhov, S.; Dogariu, A. Superdiffusion in optically controlled active media. Nat. Photonics 2012, 6, 834–837. (42) Dholakia, K.; Zemánek, P. Gripped by light: Optical binding. Rev. Mod. Phys. 2010, 82, 1767–1791. 20

ACS Paragon Plus Environment

Page 20 of 23

Page 21 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Photonics

(43) Haefner, D.; Sukhov, S.; Dogariu, A. Conservative and Nonconservative Torques in Optical Binding. Phys. Rev. Lett. 2009, 103, 173602–173605. (44) Sukhov, S.; Shalin, A.; Haefner, D.; Dogariu, A. Actio et reactio in optical binding. Opt. Express 2015, 23, 247–252. (45) Brügger, G.; Froufe Pérez, L. S.; Scheffold, F.; Sáenz, J. J. Controlling dispersion forces between small particles with artificially created random light fields. Nat. Commun. 2015, 6, 7460. (46) Han, X.; Jones, P. H. Evanescent wave optical binding forces on spherical microparticles. Opt. Lett. 2015, 40, 4042–4045. (47) Brzobohat` y, O.; Karásek, V.; Šiler, M.; Chvátal, L.; Čižmár, T.; Zemánek, P. Experimental demonstration of optical transport, sorting and self-arrangement using a ’tractor beam’. Nat. Photonics 2013, 7, 123–127. (48) Pal, S.; Deng, Z.; Ding, B.; Yan, H.; Liu, Y. DNA-Origami-Directed Self-Assembly of Discrete Silver-Nanoparticle Architectures. Angew. Chem. 2010, 122, 2760–2764. (49) Busson, M. P.; Rolly, B.; Stout, B.; Bonod, N.; Larquet, E.; Polman, A.; Bidault, S. Optical and topological characterization of gold nanoparticle dimers linked by a single DNA double strand. Nano Lett. 2011, 11, 5060–5065. (50) Steinhauer, C.; Jungmann, R.; Sobey, T. L.; Simmel, F. C.; Tinnefeld, P. DNA Origami as a Nanoscopic Ruler for Super-Resolution Microscopy. Angew. Chem. 2009, 48, 8870– 8873. (51) Chaumet, P.; Nieto Vesperinas, M. Time-averaged total force on dipolar sphere in an electromagnetic field. Opt. Lett. 2000, 25, 1065–1067. (52) Gómez-Medina, R.; San José, P.; García-Martín, A.; Lester, M.; Nieto-Vesperinas, M.;

21

ACS Paragon Plus Environment

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Sáenz, J. J. Resonant Radiation Pressure on Neutral Particles in a Waveguide. Phys. Rev. Lett. 2001, 86, 4275–4277. (53) Gómez-Medina, R.; Sáenz, J. J. Unusually strong optical interactions between particles in quasi-one-dimensional geometries. Phys. Rev. Lett. 2004, 93, 243602. (54) Juan, M. L.; Gordon, R.; Pang, Y.; Eftekhari, F.; Quidant, R. Self-induced back-action optical trapping of dielectric nanoparticles. Nat. Phys. 2009, 5, 915–919. (55) Descharmes, N.; Dharanipathy, U. P.; Diao, Z.; Tonin, M.; Houdré, R. Observation of backaction and self-induced trapping in a planar hollow photonic crystal cavity. Phys. Rev. Lett. 2013, 110, 123601. (56) Witten Jr, T.; Sander, L. M. Diffusion-limited aggregation, a kinetic critical phenomenon. Phys. Rev. Lett. 1981, 47, 1400–1403. (57) Miljković, V. D.; Pakizeh, T.; Sepulveda, B.; Johansson, P.; Käll, M. Optical Forces in Plasmonic Nanoparticle Dimers. J. Phys. Chem. C 2010, 114, 7472–7479.

22

ACS Paragon Plus Environment

Page 22 of 23

Page 23 of 23

ACS Photonics

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 ACS Paragon Plus Environment