Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Sectoral multipole focused beams

Open Access Open Access

Abstract

We discuss the properties of pure multipole beams with well-defined handedness or helicity, with the beam field a simultaneous eigenvector of the squared total angular momentum and its projection along the propagation axis. Under the condition of hemispherical illumination, we show that the only possible propagating multipole beams are “sectoral” multipoles. The sectoral dipole beam is shown to be equivalent to the non-singular time-reversed field of an electric and a magnetic point dipole Huygens’ source located at the beam focus. Higher order multipolar beams are vortex beams vanishing on the propagation axis. The simple analytical expressions of the electric field of sectoral multipole beams, exact solutions of Maxwell’s equations, and the peculiar behaviour of the Poynting vector and spin and orbital angular momenta in the focal volume could help to understand and model light-matter interactions under strongly focused beams.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

“Gaussian beams” are the most typical beams employed in optical manipulation [1] because most lasers emit light beams whose transverse electric fields and intensity distributions are well approximated by Gaussian functions. However, after focusing, when the half width of the beam waist is comparable to the wavelength, λ, the exact solution of Maxwell’s equations predicts significant deviations from the Gaussian shape, only valid in the paraxial scalar approximation (Gaussian approximation can be give accurate results only when the half width of the beam waist is greater than ∼ 10λ) [2].

A rigorous description of focal and scattered fields, usually based on the multipolar expansion of the field in terms of electric and magnetic multipoles, is needed for an accurate and quantitative discussion of field intensities, polarization and optical forces near the focal region of a focused beam [3–7]. A correct multipolar description of the light field is also relevant to analyze optical angular momentum (AM) phenomena [8,9], including the interplay between spin (SAM) and orbital (OAM) angular momenta in the focal volume of tightly focused beams [10,11] as well as to understand the spectral response and optical torques on small objects: Illumination with a tightly focused beam can only excite particle multipolar modes which are already present in the incident beam [12, 13]. As a recent example, the strong size selectivity in optical printing of Silicon particles has been associated to the dominant contribution of dipolar modes of the tightly focused laser beam [14].

In many optical applications, where one wishes to maximize the electric energy density and/or minimize the cross-polarization, a converging electric-dipole wave is the natural choice as the incident wave [15–19]. since higher order multipoles vanish at the focal point. In the so-called 4π illumination (when the incident light can cover the full solid angle), the highest energy density for a given incoming power for monochromatic beams is attained when the beam is a perfect converging dipolar beam [15]. However, for propagating beams where the incident angles are limited to a hemisphere (i.e. 2π illumination - 0 ≤ φ ≤ 2π and π/2 < θπ for a beam propagating along the z axis) the properties of the dipolar beam in the focal region can be very different, although the total energy density at the focus can still be more than half of the maximum possible [16]. The vector properties of the light strongly affect the field polarization and intensity distribution near the focus of tightly focused vector beams [20–22] Our main goal here is to explore the field properties of propagating spherical multipole beams within a framework based on helicity, angular momentum and symmetry [23].

To this end, we consider pure multipole beams (PMB) with well-defined handedness or helicity, σ = ±1 (we associate left polarized light with σ = +1 positive helicity -handness-), being simultaneous eigenvectors of the squared total angular momentum, J2, and its z-component, Jz, with eigenvalues l = 1, 2, . . . and m = −l, −l + 1, . . . l − 1, l, respectively. For a beam propagating along the z-axis, the radial component of the Poynting vector far from the focus is assumed to be always negative for incoming light (π/2 ≤ θπ) and always positive for outgoing (0 ≤ θ < π/2). As we will see, this condition of hemispherical illumination restricts the possible propagating multipole beams to the so-called “sectoral” multipoles with m = σl. Figure 1 illustrates two examples of sectoral and non-sectoral beams. Interestingly, for l > 1, sectoral PMBs are vortex beams whose field vanishes on the propagation axis with an integer vortex topological charge of σ(l − 1). Closely related vortex Hankel beams, described by outgoing waves, have been recently described [24]. The interesting case of fractional vortex beams [25–27] would require a superposition of integer vortices, i.e. a combination of multipole beams.

 figure: Fig. 1

Fig. 1 Poynting vector streamlines projected on the xz plane, perpendicular to the focal plane xy. The colormap illustrates the modulus of the Poynting vector, |S|, in the logarithm scale.

Download Full Size | PDF

In contrast with (l > 1) vortex beams, dipole beams with well defined helicity (l = 1, m = σ) concentrate the field at the focus with an energy density that is 2/3 times the Bassett upper bound of passive energy concentration [15]. Dipole beams are equivalent to mixed-dipole waves [16] and can be seen as the sum of outgoing and incoming waves radiated from an electric and a magnetic dipole located at the focus with spinning axes on the focal plane. We will see that, although the helicity and the (dimensionless) z-component of the total angular momentum per photon (m = σl) are fixed and well-defined in a PMB, the (z-component) SAM, sz, and OAM, lz = σlsz, densities present a non-trivial spatial distribution.

2. Beam expansion in vector spherical wavefunctions in the helicity representation

Let us assume that a monochromatic light beam (with an implicit time-varying harmonic component e−iωt) propagates through an homogeneous medium with real refractive index nh with a wave number k = nhω/c = 2πnh/λ0 (being λ0 the light wavelength in vacuum). To this end we will consider the expansion of the electric field, E, of the incident focused beam in vector spherical wavefunctions (VSWFs), Ψlmσ, with well defined helicity, σ = ±1 [28]:

E=σ=±1Eσ=σ=±1{l=0m=1+lClmσΨlmσ}
where Ψlmσ is defined as
Ψlmσ=12[Nlm+σMlm]
Mlmjl(kr)Xlm,Nlm1k×Mlm
Xlm1l(l+1)LYlm(θ,φ)
Here, Mlm and Nlm are Hansen’s multipoles [29], Xlm denotes the vector spherical harmonic [30], jl(kr) are the spherical (well-defined at r = 0) Bessel functions, Ylm are the spherical harmonics and L ≡ {−ir × } is the OAM operator. The expansion coefficients Clmσ in the helicity basis are equivalent to the so-called beam shape coefficients (BSCs) [4,31]. An explicit expression of the VSWFs in spherical polar coordinates (with unitary vectors êr,θ,φ) is given by
Ψlmσ=e^ril(l+1)2jl(kr)krYlm+e^θ12l(l+1){jl(kr)(σmsinθYlm)+ij˜l(kr)(Ylmθ)}+e^ϕ12l(l+1){j˜l(kr)(msinθYlm)iσjl(kr)(Ylmθ)},
with
j˜l(kr)jl1(kr)ljl(kr)kr=(l+1)jl(kr)krjl+1(kr).

Let us recall that the multipoles Ψlmσ can be built following the standard rules of angular momentum addition [29, 32] as simultaneous eigenvectors of the square of the total angular momentum, J2, and its z-component, Jz, with J = L + S⃡ given by the sum of the OAM, L, and SAM, S⃡, operators:

SiI×,L{ir×}
Si=e^iS=ie^i×,
where êi=x,y,z indicate unitary Cartesian vectors and I is the unit dyadic.

The multipoles Ψlmσ are then simultaneous eigenvectors of J2 and Jz [32] as well as of the helicity operator [23] Λ = (1/k)×, i.e. :

J2Ψlmσ=l(l+1)Ψlmσ
JzΨlmσ=mΨlmσ
ΛΨlmσ1k×Ψlmσ=σΨlmσ

3. Poynting vector for propagating pure multipole beams

In the helicity representation, the Poynting vector, P, for a monochromatic optical field, when calculated using either the electric or the magnetic field, separates into right-handed and left-handed contributions, with no cross-helicity contributions [33]:

P12Re{E*×H}=σ=±1σ2ZIm{Eσ*×Eσ}=12ZRe(E+*SE+E*SE).
where S⃡ is the spin tensor defined in Eq. (7) and
H=iZk×(E++E)=iZ(E+E)
with Z=1/(0nhc)=μ0/0h. This is an interesting result showing that for beams with well defined helicity σ, the z-component of the spin is simply proportional to the proyection of the Poynting vector on the propagation axis.

In the far-field, where

limkrjj~sin(krlπ/2)kr,
the Poynting vector for a pure multipole beam (PMB) (with l, m, σ) is given by
Plmσ~σm|Clmσ|24Z(kr)2l(l+1)[1sinθ|Ylm|2θ]e^r.
Integrating over all incoming angles (π/2 ≤ θπ), the total incoming power, PW, for a pure multipole beam is then given by
PW(l,m,σ)=r202πdφπ/2πsinθdθPlmσe^r=2πσm|Clmσ|24Zk2l(l+1)|Ylm|θ=π/22
which, from the parity relation of the spherical harmonics and associated Legendre functions, is zero for l + m odd. This implies that the total amount of power that is carried by beams with l + m odd is identical to zero, even though they may present incoming and outgoing Poynting vectors (notice that PW is the actual total power flowing through the focal plane). We shall consider incoming beams whose far-field Poynting vector points towards (outwards) the focus for all incoming angles π/2 < θπ (outgoing angles 0 ≤ θ < π/2), i.e. whose far-field Poynting vector only changes sign at θ = π/2:
e^rPlmσ<0,forπ/2θπ.ande^rPlmσ>0,for0θπ/2.

Since e^rPlmσ changes sign at the maxima and minima of |Ylm(θ,ϕ)|2, which has l − |m| distinct zeros in the interval (0 < θ < π), the only possible PMBs are the so-called ‘sectoral’ multipoles with m = σl. It is worth noticing that the integral of the Poynting vector over the whole 4π solid angle is zero since we considered a source free region. The analytical expression of the Poynting vector of a sectoral beam in closed form is given in (32). As an illustrative example, Fig. 2 shows the Poynting vector map of sectoral and a non-sectoral quadrupole beams with l = 2. From these plots it is straightforward to notice that the far-field Poynting streamlines of the sectoral beam, corresponding to Fig. 2(a), satisfies Eq. (16), while the non-sectoral, illustrated in Fig. 2(b), does not fulfil the propagating PMB criterium.

 figure: Fig. 2

Fig. 2 Poynting vector streamlines projected on the xz plane, perpendicular to the focal plane xy. The intensity color map corresponds to the field intensity normalised to the maximum field intensity for each case I = |E|2/|Emax|2. The maximum value is in both cases approximately given by |Emax|2=0.4|E1,σσ(r=0)|2, being |E1,σσ(r=0)|2 intensity at the focal plane in the (sectoral) dipolar case.

Download Full Size | PDF

4. Sectoral multipole beams with well-defined helicity

Taking into account that

Ylm=σl=(σ)l2ll!(2l+1)!4πsinlθeiσlφ,
it can be shown that the electric field of a sectoral PMB (with total incoming power PW) in spherical polar coordinates [Eβ with β = r, θ, φ] can be written as
El,σlσ=Cl,σlσΨl,σlσ=E0(σ)lsinl1θeiσlφ[e^ri(l+1)jlkrsinθ+e^θ{jl+ij˜lcosθ}+e^φ{σj˜liσjlcosθ}]
with
E0klZPWπ.
In the spherical basis
ξ^+1=e^x+ie^y2,ξ^0=e^z,ξ^1=e^xie^y2,
this reads as
El,σlσ=E0(σ)lsinl1θeiσlφ12[ξ^+1eiφ(ijl+1sin2θ+(σ+1)[jlcosθij˜l])+ξ^02(ijl+1cosθ+jl)sinθ+ξ^1eiφ(ijl+1sin2θ+(σ1)[jlcosθij˜l])].

The behaviour near the z-axis, i.e. for θπ and θ ≳ 0,

El,σlσ~E0(σ)lsinl1θ2σ2eiσ(l1)φ[z|z|jlij˜l]ξ^σ
limkrEl,σlσ~E0(σ)lsinl1θ2iσ2eiσ(l1)φeikzk|z|ξ^σ
shows that, for l > 1, the sectoral PMBs are vortex beams whose field vanishes on the propagation axis with a vortex topological charge of σ(l − 1). Figure 2(a) shows the Poynting vector lines of a sectoral quadrupolar beam (l = 2, m = 2, σ = +1) which present a “doughnut-like” field intensity pattern around the focus.

For pure dipolar beams with l = 1, the field given by Eq. (18) is identical to the first term of the expansion of a circularly polarized plane wave around the origin [1,14]. The field at the focus, r = 0, given by

E1,σσ(r=0)=i223E0ξ^σ,
is circularly polarized on the focal plane. The ratio between the electric energy density at the focal point and the total incoming power is then given by
UPW=04PW|E1,σσ(r=0)|2=23{k23πnhc}
i.e. 2/3 of the Bassett upper bound. The field intensities and Poynting vector lines for a dipole beam are shown in Fig. 3. As it can be seen the intensity is concentrated in a volume ∼ (λ/2)3, as expected for a diffraction limited beam with an interesting toroidal flow of the Poynting vector around the focus. It is worth mentioning that steady-state toroidal current flow of the Poynting vector of highly focused beams was already demonstrated experimentally and explained with an approximate field-model [34].

 figure: Fig. 3

Fig. 3 Projected flow lines of the Poynting vector, P1,1+1 and field intensities for a dipole beam. The intensity color map corresponds to the field intensity normalised to the maximum field intensity at the center I = |E|2/|Emax|2, where |Emax|2=8|E0|2/9=|C11+1|2/12π.

Download Full Size | PDF

4.1. Dipolar beams and time-reversal of a Huygens’s source

Dipole beams are equivalent to mixed-dipole waves [16] and can be seen as the sum of outgoing and incoming waves radiated from an electric dipole, p(σ), and an equivalent magnetic dipole, m(σ), located at the focus. These electric and magnetic point sources are known as Huygens’ sources and their direct, outgoing, emission present a peculiar asymmetric distribution of the radiation pattern [16, 35, 36]. As we show in the Appendix B, the field of a dipole beam (Eq. (18) with l = 1) can be rewritten as

E1,σσ=2ik20Im{Gee(r)}p(σ)*+2k2ZIm{Gem(r)}m(σ)*,
where Gee(r) is the outgoing dyadic Green function, Gem ≡ (1/k) × Gee and
p(σ)0h=p(σ)*0h=3πk3E1,σσ(r=0),withm(σ)=m(σ)*=3πk3H1,σσ(r=0).
Interestingly, for an incoming beam with helicity σ, Eq. (26) is exactly the time-reversed electric field radiated by the Huygens’ source (with helicity −σ) as it would be obtained in a time-reversal mirror cavity [37].

5. Linearly polarized dipolar beams

Focused linearly polarized dipolar beams can be built by combining two well-defined helicity beams with opposite signs from Eq. (18), i.e.

Ex(r)=12(E1,11(r)E1,11(r))andEy(r)=i2(E1,11(r)+E1,11(r)).
The field at the focus is linearly polarized on the focal plane
Ex,y(r=0)=i223E0e^x,y.
The Poynting vector, obtained by inserting Eq. (32) into Eq. (12) including both helicities in the summation, present the characteristic toroidal flow around the focus but, as expected, the intensity pattern at the focal plane is not longer axially symmetric (see Fig. 4).

 figure: Fig. 4

Fig. 4 Projected Poynting vector and colormap of the intensity for linarly polarized dipolar beams (a) on the xz plane, (b) on the yz plane and (c) on the xy focal plane. The streamlines of the Poynting vector are normal to the focal plane.

Download Full Size | PDF

6. Spin and orbital angular momenta of pure multipolar beams

Let us now consider the separation of the total angular momentum of the sectoral PMB field into SAM and OAM. Since the PMBs are eigenvectors of the z-component of the total angular momentum, we may define its (constant) density as

{El,σlσ}*JzEl,σlσ|El,σlσ|2=m=σl={El,σlσ}*(Lz+Sz)El,σlσ|El,σlσ|2=lz(r)+sz(r)
where we have introduced lz and sz as the OAM and SAM densities, respectively (although Lz and Sz are not proper angular-momentum operators [38]). For fields with well defined helicity σ, the z component of the SAM can be calculated from the projection of the Poynting vector on the beam axis, i.e. from Eqs. (12) and (32), or, alternatively, taking into account that the vectors ξμ are eigenfunctions of Sz with eigenvalue μ:
{El,σlσ}*SzEl,σlσ=2ZσPl,σlσe^z=μ=+1,0,1μ|El,σlσξ^μ|2=2σ|E0|2Zsin2l2θ{[|j˜l|2+|jl|2]cos2θ+l+1kr{jlj˜l}sin2θ}

Figure 5 illustrates the non-trivial behaviour of the SAM for a dipole beam (σ = +1) on two perpendicular planes in the focal volume. As shown in Fig. 5(a), the spin changes sign inside the torus defined by the Poynting vector lines (which is simply due to the change of sign of the z-component of the Poynting vector in the torus). In these (blue) regions, the z-component of the OAM lz is larger than the z-component of the total momentum l = 1 (since lz = lsz). This is often referred to in the literature as the super-momentum regime [39–41].

 figure: Fig. 5

Fig. 5 Projected Poynting vector, P1,1+1, flow lines and color map of the z-component of the SAM density for a dipole beam. The toroidal structure can be easily inferred from the sign of the spin density in both planes.

Download Full Size | PDF

7. Concluding remarks

Pure multipolar beams are exact analytical solutions of Maxwells equations with well defined angular momentum properties that can be extremely useful to understand and model different phenomena associated to light-matter interactions under strongly focused beams without requiring sophisticated and cumbersome numerical calculations. We have shown that dipole beams could be generated by time-reversal techniques [37] as the time reversal of a Huygens’ source. The steady-state toroidal Poynting vector flow in focused beams, which was already demonstrated experimentally and explained with approximated beam models [34], appears here in a natural way in the exact analytical solution for the dipole beam. The peculiar distribution of the SAM and OAM around the focus could lead to some interesting angular momentum transfer phenomena to small asymmetric and/or absorbing particles trapped by circularly polarized highly focused beams [42]. Higher sectoral multipole vortex beams, with well-defined total angular momentum and its projection on the propagation axis could also be useful to understand scattering and AM phenomena in vortex beams [43]. The peculiar properties of the fields in the focus of highly focused beams may also be relevant to understand the light-induced emergence of cooperative phenomena in colloidal suspensions of nanoparticles [44,45].

A. Poynting vector for sectoral multipole beams

Using Eq. (12) together with (18), the Poynting vector for a sectoral multipole beam is given by:

Pl,σlσ=k2PWlπsin2l2θ[e^r[|j˜l|2+|jl|2]cosθ+e^θ{l+1kr{jlj˜l}sinθ}+e^φ{(l+1)kr[|jl|2σsinθ]}]

B. Field of a sectoral dipole beam in terms of the Green tensors

The electric field radiated by an electric dipole, p(σ), and a magnetic dipole, m(σ), located at the origin of coordinates can be written as

Ed=k2Gee(r)p(σ)0h+k2GemiZm(σ)=k20h(Gee(r)p(σ)+inhcGemm(σ))
with
Gee(r)p(σ)=ik4π{h0(kr)(r×p(σ)×rr2+h1(kr)kr(3(rp(σ))rr2p(σ))},
Gem(r)m(σ)=ik4πh1(kr)r×m(σ)r.
In spherical polar coordinates,
Ed=(ErEθEφ)=ik34π0h(2h1(kr)kr000h˜1(kr)000h˜1(kr))(prpθpφ)+k34πZh1(kr)(000001010)(mrmθmφ)
where h1 are the outgoing Hankel functions, j1=(h1+h1*)/(2i), and
h˜1(kr)h0(kr)h1(kr)kr.

Assuming

p(σ)0h=3πk3E1,σσ(r=0)=i4πk3E02ξ^σ=iσ4πk3E02eiσφ(sinθcosθiσ)
m(σ)=3πk3H1,σσ(r=0)=3πk3iσZE1,σσ(r=0),
It is easy to show that
Ed=E1,σσ=2ik20Im{Gee(r)}p(σ)+k2ZIm{Gem(r)}m(σ)=2ik20Im{Gee(r)}p(σ)*+k2ZIm{Gem(r)}m(σ)*,
which corresponds to Eq. (26) and the extension of Eq. (8) of Carminati et al. [37] for electric and magnetic sources.

Funding

Spanish Ministerio de Economía y Competitividad (MICINN) and European Regional Development Fund (ERDF) (FIS2015-69295-C3-3-P); Basque Dep. de Educación (PI-2016-1-0041) and PhD Fellowship (PRE- 2018-2-0252).

References

1. P. H. Jones, O. M. Maragò, and G. Volpe, Optical tweezers: Principles and applications (Cambridge University, 2015). [CrossRef]  

2. P. Varga and P. Török, “The gaussian wave solution of maxwell’s equations and the validity of scalar wave approximation,” Opt. Commun. 152, 108–118 (1998). [CrossRef]  

3. C. Sheppard, “Electromagnetic field in the focal region of wide-angular annular lens and mirror systems,” IEE J. on Microwaves, Opt. Acoust. 2, 163–166 (1978). [CrossRef]  

4. G. Gouesbet, B. Maheu, and G. Gréhan, “Light scattering from a sphere arbitrarily located in a gaussian beam, using a bromwich formulation,” J. Opt. Soc. Am. A 5, 1427–1443 (1988). [CrossRef]  

5. T. Nieminen, H. Rubinsztein-Dunlop, and N. Heckenberg, “Multipole expansion of strongly focussed laser beams,” J. Quant. Spectrosc. Radiat. Transf. 79, 1005–1017 (2003). [CrossRef]  

6. J. A. Lock, “Calculation of the radiation trapping force for laser tweezers by use of generalized lorenz-mie theory. i. localized model description of an on-axis tightly focused laser beam with spherical aberration,” Appl. Opt. 43, 2532–2544 (2004). [CrossRef]   [PubMed]  

7. I. Iglesias and J. J. Sáenz, “Scattering forces in the focal volume of high numerical aperture microscope objectives,” Opt. Commun. 284, 2430–2436 (2011). [CrossRef]  

8. S. M. Barnett, L. Allen, and M. J. Padgett, Optical angular momentum (CRC Press, 2016).

9. S. M. Barnett, M. Babiker, and M. J. Padgett, “Optical orbital angular momentum,” Phil. Trans. R. Soc. A 375, 20150444 (2017). [CrossRef]   [PubMed]  

10. K. Y. Bliokh, E. A. Ostrovskaya, M. A. Alonso, O. G. Rodríguez-Herrera, D. Lara, and C. Dainty, “Spin-to-orbital angular momentum conversion in focusing, scattering, and imaging systems,” Opt. Express 19, 26132–26149 (2011). [CrossRef]  

11. K. Y. Bliokh, F. Rodríguez-Fortuño, F. Nori, and A. V. Zayats, “Spin–orbit interactions of light,” Nat. Photonics 9, 796 (2015). [CrossRef]  

12. N. M. Mojarad, V. Sandoghdar, and M. Agio, “Plasmon spectra of nanospheres under a tightly focused beam,” JOSA B 25, 651–658 (2008). [CrossRef]  

13. X. Zambrana-Puyalto, X. Vidal, and G. Molina-Terriza, “Excitation of single multipolar modes with engineered cylindrically symmetric fields,” Opt. Express 20, 24536–24544 (2012). [CrossRef]   [PubMed]  

14. C. Zaza, I. L. Violi, J. Gargiulo, G. Chiarelli, L. Langolf, J. Jakobi, J. Olmos-Trigo, E. Cortes, M. König, S. Barcikowski, S. Schlücker, J. J. Sáenz, S. Maier, and F. D. Stefani, “Size-selective optical printing of silicon nanoparticles through their dipolar magnetic resonance,” ACS Photonics, doc. ID 8b01619, (posted 13 March 2019, in press). [CrossRef]  

15. I. Bassett, “Limit to concentration by focusing,” Opt. Acta: Int. J. Opt. 33, 279–286 (1986). [CrossRef]  

16. C. Sheppard and K. Larkin, “Optimal concentration of electromagnetic radiation,” J. Mod. Opt. 41, 1495–1505 (1994). [CrossRef]  

17. V. Dhayalan and J. J. Stamnes, “Focusing of mixed-dipole waves,” Pure Appl. Opt. 6, 317 (1997). [CrossRef]  

18. G. Zumofen, N. Mojarad, V. Sandoghdar, and M. Agio, “Perfect reflection of light by an oscillating dipole,” Phys. Rev. Lett. 101, 180404 (2008). [CrossRef]   [PubMed]  

19. I. Gonoskov, A. Aiello, S. Heugel, and G. Leuchs, “Dipole pulse theory: Maximizing the field amplitude from 4 π focused laser pulses,” Phys. Rev. A 86, 053836 (2012). [CrossRef]  

20. R. Dorn, S. Quabis, and G. Leuchs, “Sharper focus for a radially polarized light beam,” Phys. Rev. Lett. 91, 233901 (2003). [CrossRef]   [PubMed]  

21. T. Bauer, S. Orlov, U. Peschel, P. Banzer, and G. Leuchs, “Nanointerferometric amplitude and phase reconstruction of tightly focused vector beams,” Nat. Photonics 8, 23 (2014). [CrossRef]  

22. S. Nechayev, J. S. Eismann, G. Leuchs, and P. Banzer, “Orbital-to-spin angular momentum conversion employing local helicity,” Phys. Rev. B 99, 075155 (2019). [CrossRef]  

23. I. Fernandez-Corbaton, X. Zambrana-Puyalto, and G. Molina-Terriza, “Helicity and angular momentum: A symmetry-based framework for the study of light-matter interactions,” Phys. Rev. A 86, 042103 (2012). [CrossRef]  

24. V. V. Kotlyar and A. A. Kovalev, “Circularly polarized hankel vortices,” Opt. Express 25, 7778–7790 (2017). [CrossRef]   [PubMed]  

25. M. Berry, “Optical vortices evolving from helicoidal integer and fractional phase steps,” J. Opt. A: Pure Appl. Opt. 6, 259 (2004). [CrossRef]  

26. J. Leach, E. Yao, and M. J. Padgett, “Observation of the vortex structure of a non-integer vortex beam,” New J. Phys. 6, 71 (2004). [CrossRef]  

27. P. Jones, M. Rashid, M. Makita, and O. Maragò, “Sagnac interferometer method for synthesis of fractional polarization vortices,” Opt. Lett. 34, 2560–2562 (2009). [CrossRef]   [PubMed]  

28. J. Olmos-Trigo, C. Sanz-Fernández, A. García-Etxarri, G. Molina-Terriza, F. S. Bergeret, and J. J. Sáenz, “Enhanced spin-orbit optical mirages from dual nanospheres,” Phys. Rev. A 99, 013852 (2019). [CrossRef]  

29. F. Borghese, P. Denti, and R. Saija, Scattering from model nonspherical particles: theory and applications to environmental physics (Springer Science & Business Media, 2007).

30. J. D. Jackson, Classical Electrodynamics (John Wiley & Sons, 1999).

31. A. A. R. Neves, A. Fontes, L. A. Padilha, E. Rodriguez, C. H. de Brito Cruz, L. C. Barbosa, and C. L. Cesar, “Exact partial wave expansion of optical beams with respect to an arbitrary origin,” Opt. Lett. 31, 2477–2479 (2006). [CrossRef]   [PubMed]  

32. A. R. Edmonds, Angular momentum in quantum mechanics (Princeton University, 1957). [CrossRef]  

33. A. Aiello and M. Berry, “Note on the helicity decomposition of spin and orbital optical currents,” J. Opt. 17, 062001 (2015). [CrossRef]  

34. Y. Roichman, B. Sun, A. Stolarski, and D. G. Grier, “Influence of nonconservative optical forces on the dynamics of optically trapped colloidal spheres: the fountain of probability,” Phys. Rev. Lett. 101, 128301 (2008). [CrossRef]   [PubMed]  

35. R. Gomez-Medina, B. Garcia-Camara, I. Suárez-Lacalle, F. González, F. Moreno, M. Nieto-Vesperinas, and J. J. Sáenz, “Electric and magnetic dipolar response of germanium nanospheres: interference effects, scattering anisotropy, and optical forces,” J. Nanophotonics 5, 053512 (2011). [CrossRef]  

36. J.-M. Geffrin, B. García-Cámara, R. Gómez-Medina, P. Albella, L. Froufe-Pérez, C. Eyraud, A. Litman, R. Vaillon, F. González, M. Nieto-Vesperinas, J. J. Sáenz, and F. Moreno, “Magnetic and electric coherence in forward-and back-scattered electromagnetic waves by a single dielectric subwavelength sphere,” Nat. Commun. 3, 1171 (2012). [CrossRef]  

37. R. Carminati, R. Pierrat, J. De Rosny, and M. Fink, “Theory of the time reversal cavity for electromagnetic fields,” Opt. Lett. 32, 3107–3109 (2007). [CrossRef]   [PubMed]  

38. S. Van Enk and G. Nienhuis, “Spin and orbital angular momentum of photons,” EPL (Europhysics Lett. 25, 497 (1994). [CrossRef]  

39. A. Y. Bekshaev, K. Y. Bliokh, and F. Nori, “Transverse spin and momentum in two-wave interference,” Phys. Rev. X 5, 011039 (2015).

40. G. Araneda, S. Walser, Y. Colombe, D. Higginbottom, J. Volz, R. Blatt, and A. Rauschenbeutel, “Wavelength-scale errors in optical localization due to spin–orbit coupling of light,” Nat. Phys. 15, 17 (2019). [CrossRef]   [PubMed]  

41. J. Olmos-Trigo, C. Sanz-Fernández, F. S. Bergeret, and J. J. Sáenz, “Asymmetry and spin-orbit coupling of light scattered from subwavelength particles,” Opt. Lett. 44, 1762–1765 (2019). [CrossRef]   [PubMed]  

42. D. B. Ruffner and D. G. Grier, “Optical forces and torques in nonuniform beams of light,” Phys. Rev. Lett. 108, 173602 (2012). [CrossRef]   [PubMed]  

43. M. Chen, M. Mazilu, Y. Arita, E. M. Wright, and K. Dholakia, “Dynamics of microparticles trapped in a perfect vortex beam,” Opt. Lett. 38, 4919–4922 (2013). [CrossRef]   [PubMed]  

44. T. Kudo, S.-J. Yang, and H. Masuhara, “A single large assembly with dynamically fluctuating swarms of gold nanoparticles formed by trapping laser,” Nano Lett. 18, 5846–5853 (2018). [CrossRef]   [PubMed]  

45. R. Delgado-Buscalioni, M. Meléndez, J. Luis-Hita, M. Marqués, and J. J. Sáenz, “Emergence of collective dynamics of gold nanoparticles in an optical vortex lattice,” Phys. Rev. E 98, 062614 (2018). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Poynting vector streamlines projected on the xz plane, perpendicular to the focal plane xy. The colormap illustrates the modulus of the Poynting vector, |S|, in the logarithm scale.
Fig. 2
Fig. 2 Poynting vector streamlines projected on the xz plane, perpendicular to the focal plane xy. The intensity color map corresponds to the field intensity normalised to the maximum field intensity for each case I = |E|2/|Emax|2. The maximum value is in both cases approximately given by | E max | 2 = 0.4 | E 1 , σ σ ( r = 0 ) | 2, being | E 1 , σ σ ( r = 0 ) | 2 intensity at the focal plane in the (sectoral) dipolar case.
Fig. 3
Fig. 3 Projected flow lines of the Poynting vector, P 1 , 1 + 1 and field intensities for a dipole beam. The intensity color map corresponds to the field intensity normalised to the maximum field intensity at the center I = |E|2/|Emax|2, where | E max | 2 = 8 | E 0 | 2 / 9 = | C 11 + 1 | 2 / 12 π.
Fig. 4
Fig. 4 Projected Poynting vector and colormap of the intensity for linarly polarized dipolar beams (a) on the xz plane, (b) on the yz plane and (c) on the xy focal plane. The streamlines of the Poynting vector are normal to the focal plane.
Fig. 5
Fig. 5 Projected Poynting vector, P 1 , 1 + 1, flow lines and color map of the z-component of the SAM density for a dipole beam. The toroidal structure can be easily inferred from the sign of the spin density in both planes.

Equations (41)

Equations on this page are rendered with MathJax. Learn more.

E = σ = ± 1 E σ = σ = ± 1 { l = 0 m = 1 + l C lm σ Ψ lm σ }
Ψ lm σ = 1 2 [ N lm + σ M lm ]
M lm j l ( k r ) X lm , N lm 1 k × M lm
X lm 1 l ( l + 1 ) L Y l m ( θ , φ )
Ψ lm σ = e ^ r i l ( l + 1 ) 2 j l ( k r ) kr Y l m + e ^ θ 1 2 l ( l + 1 ) { j l ( k r ) ( σ m sin θ Y l m ) + i j ˜ l ( kr ) ( Y l m θ ) } + e ^ ϕ 1 2 l ( l + 1 ) { j ˜ l ( k r ) ( m sin θ Y l m ) i σ j l ( k r ) ( Y l m θ ) } ,
j ˜ l ( kr ) j l 1 ( kr ) l j l ( kr ) kr = ( l + 1 ) j l ( kr ) kr j l + 1 ( kr ) .
S i I × , L { i r × }
S i = e ^ i S = i e ^ i × ,
J 2 Ψ lm σ = l ( l + 1 ) Ψ lm σ
J z Ψ lm σ = m Ψ lm σ
Λ Ψ lm σ 1 k × Ψ lm σ = σ Ψ lm σ
P 1 2 Re { E * × H } = σ = ± 1 σ 2 Z Im { E σ * × E σ } = 1 2 Z Re ( E + * S E + E * S E ) .
H = i Z k × ( E + + E ) = i Z ( E + E )
lim kr j j ~ sin ( k r l π / 2 ) kr ,
P lm σ ~ σ m | C lm σ | 2 4 Z ( kr ) 2 l ( l + 1 ) [ 1 sin θ | Y l m | 2 θ ] e ^ r .
P W ( l , m , σ ) = r 2 0 2 π d φ π / 2 π sin θ d θ P lm σ e ^ r = 2 π σ m | C lm σ | 2 4 Z k 2 l ( l + 1 ) | Y l m | θ = π / 2 2
e ^ r P lm σ < 0 , for π / 2 θ π . and e ^ r P lm σ > 0 , for 0 θ π / 2 .
Y l m = σ l = ( σ ) l 2 l l ! ( 2 l + 1 ) ! 4 π sin l θ e i σ l φ ,
E l , σ l σ = C l , σ l σ Ψ l , σ l σ = E 0 ( σ ) l sin l 1 θ e i σ l φ [ e ^ r i ( l + 1 ) j l kr sin θ + e ^ θ { j l + i j ˜ l cos θ } + e ^ φ { σ j ˜ l i σ j l cos θ } ]
E 0 k l Z P W π .
ξ ^ + 1 = e ^ x + i e ^ y 2 , ξ ^ 0 = e ^ z , ξ ^ 1 = e ^ x i e ^ y 2 ,
E l , σ l σ = E 0 ( σ ) l sin l 1 θ e i σ l φ 1 2 [ ξ ^ + 1 e i φ ( i j l + 1 sin 2 θ + ( σ + 1 ) [ j l cos θ i j ˜ l ] ) + ξ ^ 0 2 ( i j l + 1 cos θ + j l ) sin θ + ξ ^ 1 e i φ ( i j l + 1 sin 2 θ + ( σ 1 ) [ j l cos θ i j ˜ l ] ) ] .
E l , σ l σ ~ E 0 ( σ ) l sin l 1 θ 2 σ 2 e i σ ( l 1 ) φ [ z | z | j l i j ˜ l ] ξ ^ σ
lim kr E l , σ l σ ~ E 0 ( σ ) l sin l 1 θ 2 i σ 2 e i σ ( l 1 ) φ e i k z k | z | ξ ^ σ
E 1 , σ σ ( r = 0 ) = i 2 2 3 E 0 ξ ^ σ ,
U P W = 0 4 P W | E 1 , σ σ ( r = 0 ) | 2 = 2 3 { k 2 3 π n h c }
E 1 , σ σ = 2 i k 2 0 Im { G e e ( r ) } p ( σ ) * + 2 k 2 Z Im { G em ( r ) } m ( σ ) * ,
p ( σ ) 0 h = p ( σ ) * 0 h = 3 π k 3 E 1 , σ σ ( r = 0 ) , with m ( σ ) = m ( σ ) * = 3 π k 3 H 1 , σ σ ( r = 0 ) .
E x ( r ) = 1 2 ( E 1 , 1 1 ( r ) E 1 , 1 1 ( r ) ) and E y ( r ) = i 2 ( E 1 , 1 1 ( r ) + E 1 , 1 1 ( r ) ) .
E x , y ( r = 0 ) = i 2 2 3 E 0 e ^ x , y .
{ E l , σ l σ } * J z E l , σ l σ | E l , σ l σ | 2 = m = σ l = { E l , σ l σ } * ( L z + S z ) E l , σ l σ | E l , σ l σ | 2 = l z ( r ) + s z ( r )
{ E l , σ l σ } * S z E l , σ l σ = 2 Z σ P l , σ l σ e ^ z = μ = + 1 , 0 , 1 μ | E l , σ l σ ξ ^ μ | 2 = 2 σ | E 0 | 2 Z sin 2 l 2 θ { [ | j ˜ l | 2 + | j l | 2 ] cos 2 θ + l + 1 kr { j l j ˜ l } sin 2 θ }
P l , σ l σ = k 2 P W l π sin 2 l 2 θ [ e ^ r [ | j ˜ l | 2 + | j l | 2 ] cos θ + e ^ θ { l + 1 kr { j l j ˜ l } sin θ } + e ^ φ { ( l + 1 ) kr [ | j l | 2 σ sin θ ] } ]
E d = k 2 G e e ( r ) p ( σ ) 0 h + k 2 G e m i Z m ( σ ) = k 2 0 h ( G e e ( r ) p ( σ ) + i n h c G e m m ( σ ) )
G e e ( r ) p ( σ ) = i k 4 π { h 0 ( kr ) ( r × p ( σ ) × r r 2 + h 1 ( kr ) kr ( 3 ( r p ( σ ) ) r r 2 p ( σ ) ) } ,
G e m ( r ) m ( σ ) = i k 4 π h 1 ( k r ) r × m ( σ ) r .
E d = ( E r E θ E φ ) = i k 3 4 π 0 h ( 2 h 1 ( kr ) kr 0 0 0 h ˜ 1 ( kr ) 0 0 0 h ˜ 1 ( kr ) ) ( p r p θ p φ ) + k 3 4 π Z h 1 ( kr ) ( 0 0 0 0 0 1 0 1 0 ) ( m r m θ m φ )
h ˜ 1 ( kr ) h 0 ( kr ) h 1 ( k r ) kr .
p ( σ ) 0 h = 3 π k 3 E 1 , σ σ ( r = 0 ) = i 4 π k 3 E 0 2 ξ ^ σ = i σ 4 π k 3 E 0 2 e i σ φ ( sin θ cos θ i σ )
m ( σ ) = 3 π k 3 H 1 , σ σ ( r = 0 ) = 3 π k 3 i σ Z E 1 , σ σ ( r = 0 ) ,
E d = E 1 , σ σ = 2 i k 2 0 Im { G e e ( r ) } p ( σ ) + k 2 Z Im { G e m ( r ) } m ( σ ) = 2 i k 2 0 Im { G e e ( r ) } p ( σ ) * + k 2 Z Im { G e m ( r ) } m ( σ ) * ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.