Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2015 Oct 9;112(43):13190–13194. doi: 10.1073/pnas.1516704112

Lateral chirality-sorting optical forces

Amaury Hayat a,b,1, J P Balthasar Mueller a,1,2, Federico Capasso a,2
PMCID: PMC4629360  PMID: 26453555

Significance

In light of the difficulty often associated with sorting and characterizing materials by chirality, new research aimed toward the development of passive optical methods has stirred considerable excitement at the interface of analytical chemistry and physical optics. We describe here a mechanism through which chirality-sorting optical forces emerge through the interaction with the spin-angular momentum of light, a property that the community has recently learned to control with great sophistication using modern nanophotonics. In particular, the forces may be oriented perpendicularly to the propagation direction of evanescent waves, leading to a helicity-dependent lateral deflection of chiral particles in opposite directions. The highly unusual transverse optical force described herein is the strongest of its kind discovered so far to our knowledge.

Keywords: optical forces, chirality, optical spin-momentum locking, optical spin–orbit interaction, optical spin

Abstract

The transverse component of the spin angular momentum of evanescent waves gives rise to lateral optical forces on chiral particles, which have the unusual property of acting in a direction in which there is neither a field gradient nor wave propagation. Because their direction and strength depends on the chiral polarizability of the particle, they act as chirality-sorting and may offer a mechanism for passive chirality spectroscopy. The absolute strength of the forces also substantially exceeds that of other recently predicted sideways optical forces.


Chirality [from Greek χϵιρ (kheir), “hand”] is a type of asymmetry where an object is geometrically distinct from its mirror image, no matter how it is held or rotated (1). Such objects abound in nature, human hands being the classic example. Even though they share all properties other than their helicity, a chiral object and its mirror image differ in their interaction with chiral environments, such as, for example, biological systems. Analyzing and separating substances by chirality consequently represents an important problem in research and industry, affecting especially pharmaceuticals and agrochemicals (25). Whereas the sorting of substances by chirality normally has to be addressed through the introduction of a specific chiral resolving agent (6, 7), the manifestation of chirality in the electromagnetic response of materials has raised the question of passive sorting using optical forces (1016). In ref. 16 Wang and Chan recently predicted an electromagnetic plane wave to exert a lateral optical force on a chiral particle above a reflective surface, which emerges as the particle interacts with the reflection of its scattered field. This highly unusual force acts in a direction in which there is neither wave propagation nor an intensity gradient, and deflects particles with opposite helicities toward opposite sides. Apart from its fundamental interest, such a force may in theory be useful for all-optical enantiomer sorting with a single, unstructured beam. In ref. 17, Bliokh et al. almost simultaneously predicted another lateral force that is exerted on nonchiral particles in an evanescent wave. This force is a consequence of the linear momentum of the wave that is associated with its extraordinary spin angular momentum (SAM), which was first described by F. J. Belifante in the context of quantum field theory, and which vanishes for a propagating plane wave (18, 19).

In this paper we show that lateral forces can be achieved using a third mechanism, which is also inherently chirality-sorting. The forces emerge through a direct interaction of the optical SAM with chiral particles, and push either in the direction of the SAM vector or opposite to it, depending on the helicity of the particle. In general, the strength and direction of optical SAM may be estimated by considering that SAM manifests itself as a helical polarization of the electromagnetic field. For example, the SAM density of a right-circularly polarized plane wave points in the direction of propagation, and the SAM density of a left-circularly polarized wave in the direction opposite to the direction of propagation. Among plane waves, a circularly polarized wave has the highest optical SAM density, whereas a linearly polarized wave carries none. Elliptically polarized waves represent intermediate cases between these two extremes—the mathematical expression for the optical SAM density of an electromagnetic wave can be found in Supporting Information. The optical forces arising through the interaction of light with chiral matter may be intuited by considering that an electromagnetic wave with helical polarization will induce corresponding helical dipole (and multipole) moments in a scatterer. If the scatterer is chiral (one may picture a subwavelength metal helix), then the strength of the induced moments depends on the matching of the helicity of the electromagnetic wave to the helicity of the scatterer. The forces described in this paper, as well as other chirality-sorting optical forces, emerge because the induced moments in the scatterer result in radiation in a helicity-dependent preferred direction, which causes corresponding recoil.

The effect discussed in this paper enables the tailoring of chirality-sorting optical forces by engineering the local SAM density of optical fields, such that fields with transverse SAM can be used to achieve lateral forces. Fields with transverse SAM arise, for example, when light is totally internally reflected at an interface (2022). Fig. 1, Inset shows how totally internally reflected transverse-electric (TE) or transverse-magnetic (TM) polarized waves give rise to evanescent waves with elliptically polarized magnetic and electric fields, respectively. The polarization ellipses of the latter lie in the plane of incidence, implying optical SAM transverse to the direction of wave propagation.

Fig. 1.

Fig. 1.

Chirality-dependent lateral forces in an evanescent field. An evanescent field arises as light in a high index medium (n1) is totally internally reflected at the interface with a low index medium (n2) at an angle θ beyond the critical angle. Particles in an evanescent field with transverse spin angular momentum experience lateral forces depending on their chiral polarizability χ, with particles with opposite helicities experiencing lateral forces in oppose directions. (Inset) Totally internally reflected TE and TM waves give rise to evanescent waves that have transverse spin due to their elliptically polarized magnetic (TE) and electric (TM) fields.

We investigate the lateral chirality-sorting forces by calculating the forces on a small chiral particle in an evanescent field and show that the magnitude of the lateral component of the force substantially exceeds those of the previously predicted lateral forces. To this end, we discuss the optical forces exerted by an electromagnetic wave on a small chiral particle in general and then treat the specific case of an evanescent wave. We then use a nanosphere made of a chiral material to model a small chiral particle and compare the emerging lateral force to the strength of the lateral force that may arise due to the other two effects discussed by Wang and Chan (16) and Bliokh et al. (17). The optical forces exerted on an object are rigorously calculated by first finding the distribution of electromagnetic fields, and then integrating the Maxwell stress tensor over a surface enclosing the object (23). The scattering problem associated with finding the field distributions is tremendously simplified in the dipole approximation, which holds in the so-called Rayleigh limit that applies to particles much smaller than the wavelength. The dipole solution is sufficiently accurate for most molecular scattering problems and is furthermore the basis for efficient numerical methods that are used to find solutions for complex larger objects (24). We consider the optical forces on a small particle in a source-free, lossless, nondispersive, and isotropic medium with relative electric permittivity ε and relative magnetic permeability μ. In the dipole approximation, and in the absence of a permanent electric or magnetic dipole moment of the particle, the time-averaged total optical force F exerted on the particle by a monochromatic electromagnetic wave is given by (25):

F=12Re{d(E)+m(H)}F0k43Re{1μϵ(d×m)}Fint, [1]

where E and H are the electric and magnetic field vectors of the incident electromagnetic wave at the location of the particle, d and m are the electric and magnetic dipole moments of the particle that are induced by the incident field, k=nω/c is the wavenumber of the electromagnetic wave, ω is the frequency, c is the speed of light in vacuum, and n=ϵμ is the refractive index of the medium surrounding the dipole. The fields are written in complex phasor notation throughout this paper, where the factor exp(iωt) giving the time dependence is implied, and the superscript asterisk denotes the complex conjugate. Vector quantities are indicated by bold letters, and all expressions are given in Gaussian units unless stated otherwise. x^,y^, and z^ are unit vectors along the corresponding coordinate axes. The symbol denotes the dyadic product, so that the terms of the form W(V) in Eq. 1 have elements [W(V)]i=jWjiVj for i,j{x,y,z} (25) and can be written as W(V)=(W)V+W×(×V) in terms of more commonly used vector operators. Eq. 1 shows that the optical force has a component F0 corresponding to the force exerted by the incident field on the particle’s electric and magnetic dipole moments, and a component Fint that results from a direct interaction of the two dipole moments. In case of the evanescent field, it is the Fint term that gives rise to the lateral electromagnetic spin force. Chirality manifests itself in the electromagnetic response of a material through a cross-coupling of the induced electric and magnetic polarizations, such that an electric field also gives rise to a magnetic polarization, and vice versa (26, 27). Assuming isotropic polarizabilities and no permanent dipole moment, the electric and magnetic dipole moments induced by fields incident on a chiral particle are d=αeE+iχH and m=αmHiχE, where αe and αm are the electric and magnetic polarizabilities (16). The chiral polarizability of the particle χ captures the chiral nature of the dipole, such that setting χ=0 recovers the case of an achiral dipole. It is worth mentioning that, although the chiral polarizability is an inherently dynamic quantity, the dynamic dipole polarizabilities αe and αm differ from the more frequently listed corresponding static polarizabilities αe(0) and αm(0) by a radiation correction as αe=[1(i2k3/3ϵ)αe(0)]1αe(0) and αm=[1(i2k3/3μ)αm(0)]1αm(0) (28, 29). Although the static polarizability turns out to be sufficiently accurate in the particular case of the small metallic sphere considered in ref. 17, omission of the radiation correction can in general lead to a substantial underestimation of the optical forces (Supporting Information). The full expression for the force exerted on a chiral dipole by an electromagnetic field is obtained by inserting the expressions for the electric and magnetic dipole moments into Eq. 1:

F0=1g(ϵ1Re{αe}ue+μ1Re{αm}um)ωgRe{χ}hGradientForce+2ωg(μIm{αe}peo+ϵIm{αm}pmo)cgIm{χ}[×p2k2s]RadiationPressure [2]
Fint=2ωgk33(Re{αeαm*}pIm{αeαm*}p+|χ|2p)2ωg2k43n(ϵRe{χαm*}sm+μRe{χαe*}se). [3]

Here we marked the terms corresponding to the familiar gradient force and radiation pressure in the first-order term F0 (Eq. 2), both of which are modified when the dipole is chiral (i.e., when χ0). The lateral forces are contained in the Fint term, where an additional force term proportional to k4 emerges for chiral dipoles (the final term of Eq. 3). This term shows that the spin momentum density of the field gives rise to a linear momentum transfer on chiral particles, which occurs in opposite directions for opposite signs of χ (i.e., opposite helicities). The prefactor g=(4π)1 arises from the use of Gaussian units, ue=(gϵ/4)|E|2 and um=(gμ/4)|H|2 are the energy densities of the electric and the magnetic field, and h=(g/2ω)Im{EH} is the time-averaged optical helicity density (17, 18). The time-averaged Poynting momentum density p=(g/2c)Re{E×H} can be decomposed as p=po+ps into a component related to the orbital angular momentum, po, and the Belifante momentum ps related to the spin momentum density s. These quantities can be individually further separated into their electric and magnetic components, in particular po=peo+pmo and s=se+sm (Supporting Information contains a full list of the definitions of the field quantities, which are discussed in detail in ref. 17). Invoking a fluid mechanics analogy, the term ×p may be interpreted as the vorticity of the Poynting momentum density. p=(g/2c)Im{E×H} is the imaginary Poynting momentum (17, 30). For a monochromatic wave, the momentum and spin densities are (17):

ps=g8ω×[(iμ)1E×E+(iϵ)1H×H] (4)
po=g4ωIm{μ1E(E)+ϵ1H(H)} [5]
s=g4ω((μi)1E×E+(ϵi)1H×H), [6]

where the magnetic and electric contributions correspond to the terms proportional to the electric and magnetic fields. Evaluating Eqs. 2 and 3 for a plane wave E=μE0exp(ikz) propagating in the z^ direction shows that the forces are entirely in the direction of propagation:

F0=[(Im{μαe}+Im{ϵαm})/2+Im{χ}nσ]Ikz^ [7]
Fint=[(ϵRe{χαm*}+μRe{χαe*})σ/n+Re{αeαm*}+|χ|2]Ik4/3z^, [8]

where I=n1|E×H|=|E0|2. The factor σ=2Im{(Ex^)(Ey^)}/(|Ex^|2+|Ey^|2) is a measure for the degree of circular polarization (ellipticity) of the wave in the (x,y) plane, such that σ=+1 for right circular polarization and σ=1 for left circular polarization. In the calculations that follow, we will furthermore use the parameters τ=(|Ex^|2|Ey^|2)/(|Ex^|2+|Ey^|2) and ξ=2Re{(Ex^)(Ey^)}/(|Ex^|2+|Ey^|2) as measures for the linear polarization in the (x,y) plane, such that τ={+1,1} correspond to linear polarization along {x,y} and ξ={+1,1} correspond to linear polarization at angles of {+45°,45°} with respect to the x axis. The helicity-dependent change of the radiation pressure due to the chiral polarizability χ that shows up in Eq. 8 has previously been used for sorting highly chiral liquid crystal droplets in optical lattices (11). We now may consider an evanescent wave created by the total internal reflection of light at an interface in the (z,y) plane, which has the form E=μE0exp(κx)exp(ikzz). If the evanescent field is propagating in a medium with index n2=ϵμ and was created through the total internal reflection of a beam in a medium with index n1 incident on the interface with angle θ in the (x,z) plane (Fig. 1), then the wave vector components are given by kz=(n1/n2)sin(θ)k, kx=k2kz2, and κ=kz2k2=ikx, where k=n2ω/c. The energy and momentum densities of an evanescent field relevant for the calculation of the forces with Eqs. 2 and 3 are listed in Supporting Information. Notably, evanescent fields have longitudinally polarized field components (Supporting Information), which give rise to transverse spin angular momentum (17, 31). As mentioned, in case of a p-polarized wave (τ=1) it is the electric field that is elliptically polarized in the (x,z) plane, and in case of an s-polarized wave (τ=1) it is the magnetic field. Correspondingly, the out of plane electric and magnetic spin components are given by (y^se)=(1+τ)g4ωκkzIe and (y^sm)=(1τ)g4ωκkzIe. Using Eqs. 26 the forces exerted on a chiral dipole by an evanescent field are given by

F0=κIe[Re{χ}n2σ12(Re{μαe}(1+τκ2kz2)+Re{ϵαm}(1τκ2kz2))]x^Ie2kz[Im{μαe}(1+τκ2kz2)+Im{ϵαm}(1τκ2kz2)+2Im{χ}σn2]z^ [9]
Fint=Iek43κkz[Re{αeαm*}σ+Im{αe*αm}ξ+σ|χ|2+ϵn2(1τ)Re{χαm*}+μn2(1+τ)Re{χαe*}ChiralitysortingLateralForce]y^Iek43kkz[Re{αeαm*}+|χ|2+σn2(ϵRe{χαm*}+μRe{χαe*})]z^Iek43κkkz2(ξn2(Re{μχαe*}Re{ϵχαm*})+Im{αe*αm}τ)x^ [10]

with Ie=|E0|21+τ(κ/kz)2exp(2κx). The lateral forces experienced by the dipole are due to the Fint term and given by the first line of Eq. 10. Adjusting the polarization to τ=±1 renders the lateral forces directly proportional to χ, and their direction is therefore helicity-dependent. Unfortunately, the chirality-dependent lateral forces caused by ×p and s in the F0 force term exactly cancel each other (32). Small perturbations of the force that may arise due to the reflection of the field scattered by the particle at the interface at which the evanescent field is generated are neglected.

Fig. 2 compares the strength of the force to those of the lateral forces predicted by Wang and Chan (16) and Bliokh et al. (17) as it depends on the chirality and the size of a spherical nanoparticle in an electromagnetic field. In each case, the lateral force emerging through the interaction of the transverse SAM with the chiral particle is stronger by about 1.75 and 4 orders of magnitude over the considered range. Here material chirality is parametrized using the chirality parameter K=cχ[1,1] assuming real values to mirror what was done in ref. 16. The derivation of the expression of the force on such a sphere is listed in Supporting Information, as well as a discussion of the force on helicene molecules and gold nanohelices, which represent other types of chiral particles frequently considered in the literature, and which helps to further illustrate the characteristic magnitude of the force. We furthermore include calculations of the nonlateral force terms acting on these objects.

Fig. 2.

Fig. 2.

Lateral optical forces. The previously predicted lateral optical forces in comparison with the lateral force on a chiral particle in an evanescent field. The absolute value of the lateral force Fy is shown on a logarithmic scale in base 10. (A) The lateral optical force on a chiral nanosphere in a plane wave above a reflective surface according to Wang and Chan (16) and due to an evanescent field with τ=1 as a function of the chirality parameter K for positive helicity (20, 33) (Supporting Information). The sphere is nonmagnetic, situated 60 nm above the surface, and has a dielectric constant of ϵs=2 and a radius of 30 nm. In case of the force in the evanescent field, the intensity refers to the intensity of a beam that is totally internally reflected at a flint glass/air interface at an angle of θ=36.4°. (B) The lateral force on an achiral sphere (K=0) due to the Belifante spin momentum density according to Bliokh et al. (17) and on an equivalent chiral sphere (K=1) as a function of ka, where a is the radius of the sphere and k=n2ω/c is the wavenumber of the evanescent field. In both cases the sphere is nonmagnetic and situated at a height of 30 nm, has dielectric constant ϵs=2, and experiences the evanescent field generated at a flint glass/air interface at an incident angle of θ=36.4°. In the former case the polarization of the evanescent field has σ=1 and in the latter τ=1.

In summary, we predict lateral forces on materials with chiral optical response in evanescent fields, which push particles with opposite helicities in opposite directions with strength dependent on the chiral polarizability. The forces result from the direct interaction of the evanescent field’s transverse optical SAM density of the wave with the chiral electromagnetic response of the particle, and are particularly strong in comparison with previously predicted lateral optical forces. Transverse SAM may arise whenever light is laterally confined, so that the effect described in this paper represents a natural choice for the optical sorting by material chirality in an integrated system. The effect can be produced by a single beam, which avoids standing wave patterns that limit the separation of enantiomers to the width of interference fringes. However, the use of multiple beams may enable the cancellation of the longitudinal force component or, given the possibility of generating transverse spin, even the generation of spin-optical lateral forces in free space beams (34). The fact that the optical spin force described here represents by far the strongest force acting in the lateral direction renders the effect in a sense background-free, which is useful given the generally weak nature of chiral optical effects. Although we limited the discussion to the simplest case of an evanescent field created by the total internal reflection of light at a single interface, there is significant potential for improving the strength of the force by further engineering the light field. Natural choices for this are the intense and inherently evanescent fields of plasmonic excitations, or optical waveguides designed to have regions with high field enhancement, such as slot waveguides (35). We furthermore limited the discussion to the Rayleigh limit, where compact closed-form expressions exist, although the optical manipulation of objects smaller than a few 100 nm in liquid suspension is in general very challenging due to thermal agitation. Other than by tailoring the light fields, this may in particular be overcome in a low-pressure environment, for example in deflecting molecular beams (36). Larger particles may take advantage of resonances (such as Mie resonances) and are bound to experience much stronger forces. An experimental verification of the optical spin force on chiral media is very likely possible using a remarkable recently demonstrated method for measuring lateral optical forces using a cantilever with extremely low compliance (37). By immersing the cantilever into an evanescent field, Antognozzi et al. (37) were able to measure the substantially weaker lateral spin force on achiral materials with femtonewton resolution. In an experiment measuring the force described in this paper, a chiral particle could either be attached to the cantilever, or the cantilever may be manufactured from a chiral material itself (TeO2 being a standard material).

Forces Depending on Dynamic and Static Polarizabilities

Evanescent Field

Snell–Descartes Law.

When a propagating wave in a medium of index n1 arrives at the interface with a medium n with an incident angle θ the angle θt of the transmitted beam is given by Snell–Descartes law as:

n2sin(θt)=n1sin(θ). [S1]

The amplitude of the transmitted wave is subject to the incident polarization. If the incident electric field is oriented perpendicularly to the plane of incidence the wave is called s-polarized (or TE polarized). If the incident electric field is oriented parallel to the plane of incidence the wave is called p-polarized (or TM polarized). The transmission coefficients giving the amplitude of the transmitted wave for an incident s-wave and an incident a p-wave are given respectively by

ts=|EtEinc|swave=2n1cos(θ)n1cos(θ)+n2cos(θt) [S2]
tp=|EtEinc|pwave=2n1cos(θ)n2cos(θ)+n1cos(θt), [S3]

where Et stands for the transmitted field and Einc stands for the incident field in both case.

Considering Eq. S1 in case of n2<n1 there exists a critical angle of the incidence θc such that θ>θc implies sin(θt)>1. This corresponds to a complex angle of transmission θt. In this regime the incident beam is totally internally reflected, giving rise to an evanescent field.

Structure of the EM Field.

We can write the incident wave as

Einc=μ1Ainc1+|ηinc|2(ep+ηinces)exp(ikinc.r). [S4]

Here es and ep are the unit vectors corresponding, respectively, to s and p polarization. Ainc corresponds to the amplitude of the wave such that μ1|Ainc|2=Einc2. kinc=n1ωc, where n1 is the index of the medium as defined previously, μ1 the magnetic permittivity of this medium, and ϵ1 its permeability such that n1=ϵ1μ1. A time dependency of exp(iωt) is assumed. It is also obvious in Eq. S4 that ηinc characterizes the polarization, such that if ηinc=±i the wave is circularly polarized. In the same way ηinc=0 is a purely p polarized wave and ηinc=± is a purely s polarized wave.

Then, after transmission through the interface when θ>θc we have the evanescent field

E=μE0exp(κx)exp(ikzz)=μA21+|η2|2(1η2αiκkα)exp(κx)exp(ikzz). [S5]

Here we defined α=sin(θt)=kzk and k=n2ωc and n2 is the medium above the prism, where the evanescent wave is located. As previously A2 corresponds to the amplitude of the wave and can be derived with A2=αμ1μ|tp|2+|ηincts|21+|ηinc|2Ainc and we define the intensity as Ie=|E0|21+τ(κ/kz)2exp(2κx)=|A2|2e(2κx). Also, μ is the magnetic permittivity of this medium and ε its permeability such that n2=ϵμ. Also κ=kα21=ikx, as found in the main text. The polarization is described by η2=ηinc(ts/tp). Eqs. S2 and S3 give in this case

ts=2n1cos(θ)n1cos(θ)+in2α21 [S6]
tp=2n1cos(θ)n2cos(θ)+in1α21. [S7]

Defining E0=A21+|η2|2(1η2αiκkα) we obtain the form that appears in the main text.

Using this notation we can define the degree of linear, diagonal, and circular polarization in the {x,y} plane as in the main text with

σ=2Im{(Ex^)(Ey^)}(|Ex^|2+|Ey^|2)=2Im(η2)1+|η2|2 [S8]
ξ=2Re{(Ex^)(Ey^)}(|Ex^|2+|Ey^|2)=2Re(η2)1+|η2|2 [S9]
τ=(|Ex^|2|Ey^|2)(|Ex^|2+|Ey^|2)=1|η2|21+|η2|2. [S10]

So, in terms of the parameter ηinc, which defines the incident polarization,

σ=2Im(ηinctstp)1+|ηinctstp|2,ξ=2Re(ηinctstp)1+|ηinctstp|2,τ=1|ηinctstp|21+|ηinctstp|2. [S11]

These parameters are defined such that τ=1 corresponds to a p-wave, τ=1 corresponds to an s-wave, σ=1 to a left circularly polarized plane wave in the x–y plane, σ=1 to a right circularly polarized wave in the x–y plane, and ξ=±1 corresponds to opposite diagonal polarizations in the x–y plane.

Derivation of the Force Eqs. 2 and 3

Derivation of Fint.

Considering Eq. 1 in the main text:

Fint=k43Re{1μϵ[αeαm*(E×H*)+iαeχ*(E×E*)+iαm*χ(H×H*)+|χ|2(E×H)]}. [S12]

Using that Re{E×H}=2cgp=2ωgpμϵk and Im{E×E}=4ωμgse and Im{H×H}=4ωϵgsm and Re{E×E}=Re{H×H}=0 and using p=g2cIm{E×H*} we obtain

Fint=2ωgk33(Re{αeαm*}pIm{αeαm*}p+|χ|2p)2ωg2k43n2(ϵRe{χαm*}sm+μRe{χαe*}se) [S13]

as it appears in the main text in Eq. 3.

Derivation of F0.

From Eq. 1 in the main text we have

F0=12Re{αeE(E)+αmH(H)+iχ[H(E)E(H)]}. [S14]

Using W(V)=(W)V+W×(×V) and the definitions in Table S1 we can write the first term as (38)

12Re{αeE(E)}=2ωgμIm{αe}peo+12Re{αe}12|E|2 [S15]

with po defined in Table S1 using that 12|E|2=2ϵgue. Similarly rewriting the magnetic term, it follows that

12Re{αeE(E)+αmH(H)}=1g(ϵ1Re{αe}ue+μ1Re{αm}um)+2ωg(μIm{αe}peo+ϵIm{αm}pmo). [S16]

This is the achiral part of the force. Now considering the third and fourth terms and using again the identity W(V)=(W)V+W×(×V) we obtain

12Re{iχ[H(E)E(H)]}=12Re{iχ[(H)E(E)H+H×(×E)E×(×H)]}=12Im{χ[(H)E(E)H+H×(×E*)E×(×H)]}=12Re{χ}Im{(EH)}12Im{χ}Re{(H)E(E)H}12Im{χ}Re{H×(×E)E×(×H)}12Re{χ}Im{H×(×E)E×(×H)}. [S17]

Then, using Maxwell’s equations in Gaussian units (17) ×E=in2kϵ1H and ×H=ikn2μ1E and the definition of helicity, h=g2ωIm{EH}, we obtain

12Re{iχ[H(E)E(H)]}=ωgRe{χ}h12Im{χ}×Re{E×H}+12Im{χ}Re{ikn2[ϵ1H×(H)+μ1E×(E)]}+12Re{χ}Im{ikn2[ϵ1H×(H)+μ1E×(E)]}. [S18]

Also, using that 2cgp=Re{E×H*} and Re{[ϵ1H×(H)+μ1E×(E)]}=0 we have

12Re{iχ[H(E)E(H)]}=ωgRe{χ}hcgIm{χ}×p12Im{χ}Im{kn2[ϵ1H×H+μ1E×E]}. [S19]

Finally, using that n2Im{ϵ1H×(H)+μ1E×(E)}=4n2ωgs=4cgks we obtain

12Re{iχ[H(E)E(H)]}=ωgRe{χ}hcgIm{χ}(×p2k2s). [S20]

This is the chiral part of the force. So, combining the four terms of F0 we finally have

F0=1g(ϵ1Re{αe}ue+μ1Re{αm}um)+2ωg(μIm{αe}peo+ϵIm{αm}pmo)ωgRe{χ}hcgIm{χ}[×p2k2s], [S21]

as it appears in the main text in Eq. 2.

Table S1.

Definition of the field quantities relevant for the emergence of the optical forces on small chiral particles

Definition of the quantity Description
p=(g/2c)Re{E×H} Poynting momentum density
p=(g/2c)Im{E×H} Value of the imaginary adjoint to the Poynting momentum density
po=g4ωIm{μ1E(E)+ϵ1H(H)} Orbital (or canonical) Poynting momentum density
peo=g4ωIm{μ1E(E)} Electric contribution to the orbital Poynting momentum density
pmo=g4ωIm{ϵ1H(H)} Magnetic contribution to the orbital Poynting momentum density
ps=g8ω×[(iμ)1E×E+(iϵ)1H×H] Spin part of the Poynting momentum density (Belinfante spin momentum)
pes=g8ω×[(iμ)1E×E] Electric contribution to the spin momentum density
pms=g8ω×[(iϵ)1H×H] Magnetic contribution to the spin momentum density
s=g4ω((μi)1E×E+(ϵi)1H×H) SAM density
se=g4ω((μi)1E×E) Electric contribution to the SAM density
sm=g4ω((ϵi)1H×H) Magnetic contribution to the SAM
h=(g/2ω)Im{EH} Helicity
×p=g2c×Re{E×H} Vorticity of the photon flow
ue=g4ϵ|E|2 Electric contribution to the energy density
um=g4μ|H|2 Magnetic contribution to the energy density
I=n1|E×H*| Intensity

A derivation and discussion of many quantities can be found in Bliokh et al. (17).

Table S2.

Field quantities of a propagating wave and an evanescent wave

Symbol For a propagating wave E=μE0exp(ikz)I=|E0|2 For an evanescent field E=μE0exp(κx)exp(ikzz)Ie=|E0|21+τ(κ/kz)2exp(2κx) Quantity
p g2ωIkzz^ g2ωIe(σκkkzy^+k2kzz^) Poynting momentum density
p 0 g2ωIe(τκk2kz2x^+ξκkkzy^) Value of the imaginary adjoint to the Poynting momentum density
po g2ωIkzz^ g2ωIekzz^ Orbital (or canonical) Poynting momentum density
peo g4ωIkzz^ g4ωIe(1+κ2kz2)kzz^ Electric contribution to the orbital Poynting momentum density
pmo g4ωIkzz^ g4ωIe(1κ2kz2)kzz^ Magnetic contribution to the orbital Poynting momentum density
ps 0 g2ωIe(κ2kz2z^+σκkkzy^) Spin part of the Poynting momentum density [Belinfante’s momentum (17, 19)]
pes 0 g4ωIe((1+τ)κ2kz2z^+σκkkzy^) Electric contribution to the spin momentum density
pms 0 g4ωIe((1τ)κ2kz2z^+σκkkzy^) Magnetic contribution to the spin momentum density
s g2ωIσz^ g2ωIe(σkkzz^+κkzy^) SAM density
se g4ωIσz^ g4ωIe(σkkzz^+(1+τ)κkzy^+ξκkkz2x^) Electric contribution to the SAM density
sm g4ωIσz^ g4ωIe(σkkzz^+(1τ)κkzy^ξκkkz2x^) Magnetic contribution to the SAM
h g2ωInσ g2ωIen2σ Helicity
×p 0 g2ωIe(2κk2kzy^2σκ2kkzz^) Vorticity of the photon flow
ue gn24I gn224Ie(1+τκ2kz2) Electric contribution to energy density
um gn24I gn224Ie(1τκ2kz2) Magnetic contribution to energy density

Summary of the physical quantities involved in the force for a propagating plane wave and for an evanescent field. The magnetic and electric contributions to the orbital and spin angular momenta are listed explicitly.

Lateral Forces on a Molecule, a Nanosphere, and a Gold Nanohelix

In this section we briefly discuss the lateral forces that arise on three types of small chiral particles that commonly arise in the literature and leave the details of the associated calculations to the next section.

Helicene Molecules.

The optical manipulation of objects smaller than a few hundred nanometers in liquid suspension is in general very challenging due to thermal agitation. However, the optical sorting of individual atoms or molecules by chirality may nevertheless be possible in low-pressure environments or by tailoring the fields in modern metamaterials. A molecule of hexahelicene has a real chiral polarizability of χ=6.2102Å3 at a wavelength of 589.3 nm, which lies well outside of the molecule’s absorption band (39). At the same wavelength, the static electric polarizability of hexahelicene is αe=10.4Å3 with the magnetic polarizability |αm|<105|αe|. We consider a beam with a given intensity that is totally internally reflected at the interface of heavy flint glass (n1=1.74) and air (n2=1) at an angle of θ=36.4°, which is close to the critical angle and gives the strongest force given the angle-dependent intensity and wave-vector of the evanescent field. Situated 60 nm above the interface, the hexahelicene molecule experiences a lateral force of Fy2.041019pN/(mW/μm2) if σ=1 [circularly polarized in the (x,y) plane] and Fy4.041019pN/(mW/μm2) if τ=1 (p-polarized). Stronger optical forces are possible within the absorption band, where the chiral polarizability is complex and highly dependent on the wavelength. Coronene is a close relative of hexahelicene that is achiral but has the same size and electric polarizability as helicene. The lateral force on this achiral molecule due to the Belifante momentum density ps as predicted by Bliokh et al. (17) is Fy1.7.1021pN/(mW/μm2) for σ=1 and vanishes for σ=0 [linearly polarized in (x,y) plane]. The strength of the nonlateral forces are F0x^=3.13109pN/(mW/μm2), F0z^=1.941016pN/(mW/μm2), and Fintz^=1.021020pN/(mW/μm2).

Spherical Nanoparticles.

The polarizability of a chiral particle may be calculated using bi-isotropic constitutive relations for the material of the particle, where material chirality is parametrized with a chirality parameter K[1,1] (16). The derivation and the somewhat bulky expressions for the polarizabilities of a small chiral sphere are given in the following section. Examples of spherical particles with highly chiral electromagnetic response are cholesteric liquid crystal droplets (11). Assuming the refractive index of 5CB (4-cyano-4′-pentylbiphenyl) of ns=1.597 and K=1 for a 30-nm sphere in water with index n2=1.33, letting an incident beam with 589-nm wavelength totally internally reflect at θ=52° in heavy flint glass (n1=1.74) causes a lateral force of Fy2.6105pN/(mW/μm2) if the sphere is floating 60 nm above the flint glass surface and τ=1. The strength of the nonlateral forces are F0x^=4.73103pN/(mW/μm2), F0z^=2.30104pN/(mW/μm2), and Fintz^=1.87104pN/(mW/μm2).

The lateral force on a chiral sphere above a reflective surface in a propagating field as predicted by Wang and Chan (16) corresponds to Fy=106pN/(mW/μm2) for a sphere with 30-nm radius in vacuum with a dielectric constant of ϵ=2 and chirality parameter K=1 when it is located 60 nm above a metal surface and illuminated with a plane wave that has τ=1 and a wavelength of 600 nm. In comparison, the same sphere would experience chirality-sorting lateral force of Fy=4.6105pN/(mW/μm2) at the same height above the surface in an evanescent field that is created by total internal reflection at θ=36.4° in heavy flint glass (n1=1.74).

Nanohelices.

Increasing interest in chiral optical materials has recently driven the development of artificial nanostructures with extreme chiral responses (4043), including artificially chiral effective media (27). Metallic nanohelices in particular enable the analytical estimation of their very strong chiral polarizability and have recently been fabricated at size scales small enough to reach plasmonic resonances in the visible range (41). A perfectly conducting 50-nm helix with 25-nm diameter and one loop has a chiral polarizability of χ=2.72107Å3 and a static electric polarizability of αe=2.56108Å3 (26). Situated 30 nm above the surface in an evanescent field with wavelength of 589 nm and τ=1, a lateral force of Fy=4.25103pN/(mW/μm2) results if the beam is totally internally reflected at a flint glass/air interface at a θ=36.4° angle. The strength of the nonlateral forces are F0x^=1,16101pN/(mW/μm2), F0z^=0pN/(mW/μm2), and Fintz^=1.82103pN/(mW/μm2).

Optical Forces on a Sphere

As stated in the main text, if the sphere and the surrounding medium are assumed nonmagnetic the static polarizabilities are given at the first nonzero order by

αe0=ϵa3(ϵpϵ)ϵp+2ϵ [S22]
αm0=ϵpϵ30ϵk2a5. [S23]

Here a refers to the radius of the sphere and ϵp and ε are the relative permittivity of the sphere and the surrounding medium, respectively. Using Draine’s correction we obtain the dynamic polarizabilities as αe=[1(i2k3/3ϵ)αe(0)]1αe(0) and αm=[1(i2k3/3)αm(0)]1αm(0). The forces on an achiral sphere are then

F0=κIe2a3ϵRe{ϵpϵϵp+2ϵ}(1+τκ2kz2)x^ [S24]
Ie2kza3(ϵIm{ϵpϵϵp+2ϵ}+2ϵ3(ka)3Re{(ϵpϵϵp+2ϵ)2})(1+τκ2kz2)z^ [S25]
Fint=Iek43κkz|ϵpϵ2ϵ+ϵp|2[130Re{ϵp+2}k2a8σ+(130Im{ϵp}k2a8290k5a11Re{(ϵpϵ)(ϵp+2ϵ)*ϵp+2ϵ})ξ]y^ [S26]
Iek43kkz[130|ϵpϵ2ϵ+ϵp|2Re{ϵp+2}k2a8]z^ [S27]
Iek43κkkz2|ϵpϵ2ϵ+ϵp|2[(130Im{ϵp}k2a8290k5a11Re{(ϵpϵ)(ϵp+2ϵ)*ϵp+2ϵ})τ]x^. [S28]

For a chiral sphere the polarizabilities depend on the chirality of the sphere and of the surrounding medium. Assuming a nonchiral, nonmagnetic surrounding medium and μp=1 Lakthakia et al. (26) showed that we can derive:

αe0=a3ϵ[2K2ϵϵp+3(ϵpϵ)]4K2ϵϵp3(2ϵ+ϵp) [S29]
αm0=a3[2K2ϵϵp]4K2ϵϵp3(2ϵ+ϵp) [S30]
χ=3a3Kϵ4K2ϵϵp3(2ϵ+ϵp), [S31]

where a is the radius of the sphere and K the chiral parameter such that the usual dispersion relation for two opposite circularly polarized plane waves in the medium of the sphere is (16) k±=(ϵp±K)ωc.

Chiral Polarizability of a Helix

We know that we have the electric and magnetic dipole moments d and m induced by fields incident on a chiral particle with isotropic polarizabilities and no permanent dipole moment given by (16)

d=αeE+iχHm=αmHiχE, [S32]

where αe and αm are the dynamic electric and magnetic polarizabilities. The chiral polarizability of the particle χ captures the chiral nature of the dipole, such that setting χ=0 recovers the behavior for an achiral dipole. For a perfectly conducting helix in the Rayleigh regime and a perfect metal we know from ref. 33 that the polarizabilities are in Gaussian units by

χ=αe(πa2k2l) [S33]
αe=(l)24πk2a(L/μ0a), [S34]

where a is the radius of the helix, l its length taken, and L its inductance in SI units and here L/μ0a is taken in SI units. For a helix with one loop we have Lμ0a2πl(116a3πl+2a2l24a4l4)(SI) and we obtain

χl24πk(116a3πl+2a2l24a4l4)1 [S35]
αel34π2k2a2(116a3πl+2a2l24a4l4)1. [S36]

Using l = 50 nm and a = 12.5 nm so that the diameter is 2a = l/2 = 25 nm we obtain 1116a3πl+2a2l24a4l41.46.

Fig. S1.

Fig. S1.

Impact of Draine’s correction on the optical force. The optical force on an achiral sphere with radius a and dielectric constant ϵp in an evanescent field with wavenumber k=n2ωc as a function of ka using the static polarizability (red, dashed line) and the dynamic polarizability after Draine’s correction (green, continuous line). A and B correspond to the force on a dielectric sphere with ϵp=2.56 and an incident polarization corresponding to ξ=1 (linear diagonal polarization), respectively, along the z direction (parallel to po, A) and the y direction (lateral optical force, B). C and D correspond to the lateral optical force using a gold sphere with ϵp=12.2+3i with an incident polarization of the beam corresponding, respectively, to ξ=1 (linear diagonal polarization, C) and τ=1 (linear horizontal polarization, D). The force is normalized by F0=(a2/4π)Iinc, where Iinc=|Einc|2 is the square modulus of the incident electric field. The sphere is suspended in water (index n2=1.33) and the evanescent field is generated by total internal reflection in a prism of index n1=1.74.Taking an incident intensity of 1.0mW/μm2 we have F0/a2=Iinc/4π3.3104Ncm2. Here the incident angle of total internal reflection is θ51 ° (so, above the critical angle, which is around θ=49.5 °). Note that in B the force without correction is zero.

Acknowledgments

The authors thank X. Yin for helpful discussions. This research was supported by Air Force Office of Scientific Research Grant FA9550-12-1-0289.

Footnotes

The authors declare no conflict of interest.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1516704112/-/DCSupplemental.

References

  • 1.Kelvin Lord. 1904. Baltimore Lectures (C. J. Clay and Sons, London), pp 436, 619.
  • 2.Smith SW. Chiral toxicology: It’s the same thing...only different. Toxicol Sci. 2009;110(1):4–30. doi: 10.1093/toxsci/kfp097. [DOI] [PubMed] [Google Scholar]
  • 3.Sekhon BS. Chiral pesticides. J Pest Sci. 2009;34(1):1–12. [Google Scholar]
  • 4.US Food and Drug Administration FDA’s policy statement for the development of new stereoisomeric drugs. Chirality. 1992;4(5):338–340. doi: 10.1002/chir.530040513. [DOI] [PubMed] [Google Scholar]
  • 5.Caner H, Groner E, Levy L, Agranat I. Trends in the development of chiral drugs. Drug Discov Today. 2004;9(3):105–110. doi: 10.1016/s1359-6446(03)02904-0. [DOI] [PubMed] [Google Scholar]
  • 6.Nguyen LA, He H, Pham-Huy C. Chiral drugs: An overview. Int J Biomed Sci. 2006;2(2):85–100. [PMC free article] [PubMed] [Google Scholar]
  • 7.McKendry R, Theoclitou ME, Rayment T, Abell C. Chiral discrimination by chemical force microscopy. Nature. 1998;391(6667):566–568. [Google Scholar]
  • 8.Tang Y, Cohen AE. Optical chirality and its interaction with matter. Phys Rev Lett. 2010;104(16):163901. doi: 10.1103/PhysRevLett.104.163901. [DOI] [PubMed] [Google Scholar]
  • 9.Tang Y, Cohen AE. Enhanced enantioselectivity in excitation of chiral molecules by superchiral light. Science. 2011;332(6027):333–336. doi: 10.1126/science.1202817. [DOI] [PubMed] [Google Scholar]
  • 10.Canaguier-Durand A, Hutchison JA, Genet C, Ebbesen TW. 2013. Mechanical separation of chiral dipoles by chiral light. arXiv:1306.3708. [DOI] [PubMed]
  • 11.Tkachenko G, Brasselet E. Optofluidic sorting of material chirality by chiral light. Nat Commun. 2014;5:3577. doi: 10.1038/ncomms4577. [DOI] [PubMed] [Google Scholar]
  • 12.Tkachenko G, Brasselet E. Helicity-dependent three-dimensional optical trapping of chiral microparticles. Nat Commun. 2014;5:4491. doi: 10.1038/ncomms5491. [DOI] [PubMed] [Google Scholar]
  • 13.Cameron RP, Barnett SM, Yao AM. Discriminatory optical force for chiral molecules. New J Phys. 2014;16(1):013020. [Google Scholar]
  • 14.Smith D, Woods C, Seddon A, Hoerber H. Photophoretic separation of single-walled carbon nanotubes: A novel approach to selective chiral sorting. Phys Chem Chem Phys. 2014;16(11):5221–5228. doi: 10.1039/c3cp54812k. [DOI] [PubMed] [Google Scholar]
  • 15.Patterson D, Schnell M, Doyle JM. Enantiomer-specific detection of chiral molecules via microwave spectroscopy. Nature. 2013;497(7450):475–477. doi: 10.1038/nature12150. [DOI] [PubMed] [Google Scholar]
  • 16.Wang SB, Chan CT. Lateral optical force on chiral particles near a surface. Nat Commun. 2014;5:3307. doi: 10.1038/ncomms4307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Bliokh KY, Bekshaev AY, Nori F. Extraordinary momentum and spin in evanescent waves. Nat Commun. 2014;5:3300. doi: 10.1038/ncomms4300. [DOI] [PubMed] [Google Scholar]
  • 18.Bliokh KY, Nori F. Characterizing optical chirality. Phys Rev A. 2011;83(2):021803. [Google Scholar]
  • 19.Belinfante FJ. On the current and the density of the electric charge, the energy, the linear momentum and the angular momentum of arbitrary fields. Physica. 1940;7(5):449–474. [Google Scholar]
  • 20.Van Mechelen T, Jacob Z. 2015. Universal spin-momentum locking of evanescent waves. arXiv:1504.06361.
  • 21.Bliokh KY, Nori F. 2015. Transverse and longitudinal angular momenta of light. arXiv:1504.03113.
  • 22.Nakamura YO. Spin quantum number of surface plasmon. Solid State Commun. 1981;39(6):763–765. [Google Scholar]
  • 23. Landau LD, et al. (1984) Electrodynamics of Continuous Media (Elsevier, Oxford)
  • 24.Purcell EM, Pennypacker CR. Scattering and absorption of light by nonspherical dielectric grains. Astrophys J. 1973;186:705–714. [Google Scholar]
  • 25.Nieto-Vesperinas M, Sáenz JJ, Gómez-Medina R, Chantada L. Optical forces on small magnetodielectric particles. Opt Express. 2010;18(11):11428–11443. doi: 10.1364/OE.18.011428. [DOI] [PubMed] [Google Scholar]
  • 26.Jaggard DL, Mickelson AR, Papas CH. On electromagnetic waves in chiral media. Appl Phys. 1979;18(2):211–216. [Google Scholar]
  • 27.Yin X, Schäferling M, Metzger B, Giessen H. Interpreting chiral nanophotonic spectra: The plasmonic Born-Kuhn model. Nano Lett. 2013;13(12):6238–6243. doi: 10.1021/nl403705k. [DOI] [PubMed] [Google Scholar]
  • 28.Draine BT. The discrete-dipole approximation and its application to interstellar graphite grains. Astrophys J. 1988;333:848–872. [Google Scholar]
  • 29.Ho JS, Stell G. Theory of the refractive index of fluids. J Chem Phys. 1982;77(10):5173–5183. [Google Scholar]
  • 30.Jackson JD. Classical Electrodynamics. Wiley; New York: 1962. [Google Scholar]
  • 31.Bliokh KY, Nori F. Transverse spin of a surface polariton. Phys Rev A. 2012;85(6):061801. [Google Scholar]
  • 32.Bliokh KY, Kivshar YS, Nori F. Magnetoelectric effects in local light-matter interactions. Phys Rev Lett. 2014;113(3):033601. doi: 10.1103/PhysRevLett.113.033601. [DOI] [PubMed] [Google Scholar]
  • 33.Lakhtakia A, Varadan VK, Varadan VV. Effective properties of a sparse random distribution of non-interacting small chiral spheres in a chiral host medium. J Phys D Appl Phys. 1991;24(1) Available at dx.doi.org/10.1088/0022-3727/24/1/001. [Google Scholar]
  • 34.Bekshaev AY, Bliokh KY, Nori F. 2014. Transverse spin and momentum in two-wave interference. arXiv:1407.6786.
  • 35.Yang AH, et al. Optical manipulation of nanoparticles and biomolecules in sub-wavelength slot waveguides. Nature. 2009;457(7225):71–75. doi: 10.1038/nature07593. [DOI] [PubMed] [Google Scholar]
  • 36.Sakai H, et al. Optical deflection of molecules. Phys Rev A. 1998;57(4):2794. [Google Scholar]
  • 37.Antognozzi M, et al. 2015. Direct measurement of the extraordinary optical momentum using a nano-cantilever. arXiv:1506.04248.
  • 38.Chaumet PC, Nieto-Vesperinas M. Time-averaged total force on a dipolar sphere in an electromagnetic field. Opt Lett. 2000;25(15):1065–1067. doi: 10.1364/ol.25.001065. [DOI] [PubMed] [Google Scholar]
  • 39.Barron LD. 2004. Molecular Light Scattering and Optical Activity (Cambridge Univ Press, Cambridge, UK), 2nd Ed.
  • 40.Gansel JK, et al. Gold helix photonic metamaterial as broadband circular polarizer. Science. 2009;325(5947):1513–1515. doi: 10.1126/science.1177031. [DOI] [PubMed] [Google Scholar]
  • 41.Gibbs JG, et al. Nanohelices by shadow growth. Nanoscale. 2014;6(16):9457–9466. doi: 10.1039/c4nr00403e. [DOI] [PubMed] [Google Scholar]
  • 42.Rogacheva AV, Fedotov VA, Schwanecke AS, Zheludev NI. Giant gyrotropy due to electromagnetic-field coupling in a bilayered chiral structure. Phys Rev Lett. 2006;97(17):177401. doi: 10.1103/PhysRevLett.97.177401. [DOI] [PubMed] [Google Scholar]
  • 43.McPeak KM, et al. Complex chiral colloids and surfaces via high-index off-cut silicon. Nano Lett. 2014;14(5):2934–2940. doi: 10.1021/nl501032j. [DOI] [PubMed] [Google Scholar]

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES