ebook img

Mie scattering of Laguerre-Gaussian beams: photonic nanojets and near-field optical vortices PDF

0.34 MB·
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Mie scattering of Laguerre-Gaussian beams: photonic nanojets and near-field optical vortices

Mie scattering of Laguerre-Gaussian beams: photonic nanojets and near-field optical vortices Alexei D. Kiselev1,∗ and Dmytro O. Plutenko1,† 1Institute of Physics of National Academy of Sciences of Ukraine, prospekt Nauki 46, 03680 Ky¨ıv, Ukraine (Dated: January 30, 2014) We study Mie light scattering of Laguerre-Gaussian (LG) beams remodelled using the method of far-field matching. The theoretical results are applied to examine the optical field in the near-field region for purely azimuthal LG beams characterized 4 by the nonzero azimuthal mode number m . The mode number m is found to LG LG 1 have a profound effect on the morphology of photonic nanojets and the near-field 0 2 structure of optical vortices associated with the components of the electric field. n a PACS numbers: 42.25.Fx, 42.68.Mj, 42.25.Bs J Keywords: light scattering; Laguerre-Gaussianbeams; photonic nanojets; optical vortices; 9 2 ] s c I. INTRODUCTION i t p o The problem of light scattering by particles of one medium embedded in another has a . s long history, dating back more than a century to the classical exact solution due to Mie [1]. c The Mie solution applies to scattering by uniform spherical particles with isotropic dielectric i s properties. The analysis of a Mie–type theory uses a systematic expansion of the electro- y h magnetic field over vector spherical harmonics [2–7]. The specific form of the expansions p is also known as the T–matrix ansatz that has been widely used in the related problem [ of light scattering by nonspherical particles [5, 8, 9]. More recently this strategy has been 1 successfully applied to optically anisotropic particles [10–18]. v 5 In its original form the Mie theory assumes that the scatterer is illuminated with a plane 8 electromagnetic wave. For laser beams, it is generally necessary to go beyond the plane-wave 5 7 approximation that may severely break down when the beam width becomes of order of the . scatterer size. The problem of light scattering from arbitrary shaped laser beams has now 1 0 a more than two decade long history [19–23] and has been the key subject of the Mie–type 4 theory — the so-called generalized Lorenz–Mie theory (GLMT) [7, 24] — extended to the 1 : case of arbitrary incident-beam scattering. v Mathematically, in such generalization of the Mie theory, the central and the most im- i X portant task is to describe illuminating beams in terms of expansions over a set of basis r a wavefunctions (for the spherical coordinate system, it is the multipole expansion over the basis of vector spherical wavefunctions). In GLMT, a variety of formally exact (the quadra- ture and double quadrature formulas) and approximate (the finite series and localized ap- proximations) methods [25] were developed to evaluate the expansion coefficients that are referred to as the beam shape coefficients (for a recent review see Ref. [26] and references therein). ∗ Email address: [email protected] † Email address: [email protected] 2 The central problem with laser beams is due to the fact that in their standard mathemat- ical form these beams are not radiation fields which are solutions to Maxwell’s equations. Typically, the analytical treatment of laser beams is performed using the paraxial approxi- mation [27] and the beams are described as pseudo-fields which are only approximate solu- tions of the vector Helmholtz equation (higher order corrections can be used to improve the accuracy of the paraxial approximation [27, 28]). Unfortunately, multipole expansions do not exist for such approximate pseudo-fields. Therefore, some remodelling procedure must be invoked to obtain a real radiation field which can be regarded as an approximation to the original paraxial beam. The basic concept that might be calledmatching the fields on a surface lies at the heart of various traditional approaches to the laser beam remodelling and is based on the assumption that there is a surface where the actual incident field is equal to the paraxial field. Examples of physically reasonable and natural choice arescatterer-independent matching surfaces such as a far-field sphere [29], the focal plane (for beams with well-defined focal planes) [29, 30], andaGaussianreference sphere representing alens [31]. Giventheparaxialfielddistribution onthe matching surface, the beam shape coefficients can beevaluated using either numerical integration or the one-point matching method [29]. An alternative approach is to describe analytically propagation of a laser beam, which is known in the paraxial limit, without recourse to the paraxial approximation. In Refs. [32– 36] this strategy has been applied to the important case of Laguerre–Gaussian (LG) beams using different methods such as the vectorial Rayleigh–Sommerfeld formulas [33, 36], the vector angular spectrum method [35], approximating LG beams by nonparaxial beams with (near) cylindrical symmetry [32, 34]. The nonparaxial beams are solutions of Maxwell’s equations and the beam shape coeffi- cients can be computed using the methods of GLMT. In recent studies of light scattering by spherical and spheroidal particles illuminated with LG beams [37, 38], the analytical results of Ref. [34] were used to calculate the beam shape coefficients. It is now well known [39] that LG beams represent optical vortex beams that carry angu- lar momentum of two kinds: spin angular momentum associated with the polarization state of the beam and orbital angular momentum related to spatial variations of the field. These variations derive from the helical structure of the wavefronts comprising the beam or, equiv- alently, from a phase singularity at the beam axis. The topological charge characterizing the phase singularity andassociated orbital angular momentum gives rise to distinctive phenom- ena such as soliton generation [40], entanglement of photon quantum states, orbital angular momentum exchange with atoms and molecules (in addition to the collection of papers [39], see reviews in Ref. [41]), rotation and orbital motion of spherical particles illuminated with LG beams [42, 43]. In this paper the problem of light scattering from LG beams that represent laser beams exhibiting a helical phase front and carrying a phase singularity will be of our primary interest. Inour calculationswe shall followRef.[15]anduse theT–matrixapproachinwhich the far-field matching method is combined with the results for nonparaxial propagation of LG beams [35, 36]. Our goal is to examine the near-field structure of electromagnetic field depending on the parameters characterizing both the beam and the scatterer. This structure has recently attracted considerable attention that was stimulated by an upsurge of interest to the so-called photonic nanojets and their applications (for a review see Ref. [44]). These nanojets were originally identified in finite-difference-time-domain simulations [45, 46] as narrow, high-intensity electromagnetic beams that propagate into 3 background medium from the shadow-side surface of a plane-wave illuminated dielectric microcylinder [45] or microsphere [46] of diameter greater than the illuminating wavelength. In other words, a photonic nanojet can be regarded as a localized, subdiffractional, non- evanescent light focus propagating along the line of incidence. The bulk of theoretical studies devoted to nanojets [47–52] has been predominantly fo- cused on the case of plane-wave illumination. In this paper we intend to fill the gap. The layout of the paper is as follows. In Sec. II, we describe our theoretical approach and then, in Sec. III, we obtain the analytical results for the beam shape coefficients of LG beams. The numerical procedure and the results of numerical computations representing the near-field intensity distributions and phase maps of electric field components for purely azimuthal LG beams are presented in Sec. IV. Finally, in Sec. V, we present our results and make some concluding remarks. II. T–MATRIX FORMULATION OF LORENZ–MIE THEORY WeconsiderscatteringbyasphericalparticleofradiusR embeddedinauniformisotropic p dielectric medium with dielectric constant ǫ and magnetic permeability µ . The dielec- med med tric constant and magnetic permittivity of the particle are ǫ and µ , respectively. p p In this subsection we remind the reader about the relationship between Maxwell’s equa- tions in the region of a scatterer and the formulation of scattering properties in terms of the T–matrix [2, 5]. Our formulation is slightly non-standard and closely follows to the line of our presentation given in Ref. [15]. Weshall need to write theMaxwell equations for a harmonicelectromagnetic wave (time– dependent factor is exp iωt ) in the form: {− } µ ik−1∇ E = iH, (1a) − i × n i n med, r > R ik−1∇ H = iE, i = p (1b) i × µ p, r < R i ( p where n = √ǫ µ is the refractive index outside the scatterer (in the ambient med med med medium), where r > R (i = med) and k = k = n k (k = ω/c = 2π/λ is p i med med vac vac the free–space wavenumber); n = √ǫ µ is the refractive index for the region inside the p p p spherical particle (scatterer), where r < R (i = p) and k = k = n k . p i p p vac A. Vector spherical harmonics and Wigner D functions The electromagnetic field can always be expanded using the vector spherical harmonic basis, Y (φ,θ) Y (ˆr) (δ = 0, 1) [53], as follows: j+δjm j+δjm ≡ ± E = E = p(0)(r)Y(0)(ˆr)+p(e)(r)Y(e)(ˆr)+p(m)(r)Y(m)(ˆr) , (2a) jm jm jm jm jm jm jm Xjm Xjm h i H = H = q(0)(r)Y(0)(ˆr)+q(e)(r)Y(e)(ˆr)+q(m)(r)Y(m)(ˆr) , (2b) jm jm jm jm jm jm jm Xjm Xjm h i 4 where Y(m) = Y and Y(e) = [j/(2j + 1)]1/2Y + [(j + 1)/(2j + 1)]1/2Y are jm jjm jm j+1jm j−1jm electric and magnetic harmonics respectively, and Y(0) = [j/(2j + 1)]1/2Y [(j + jm j−1jm − 1)/(2j + 1)]1/2Y are longitudinal harmonics. In Ref. [15], it was shown that the j+1jm spherical harmonics can be conveniently expressed in terms of the Wigner D–functions [53, 54] as follows Y(m)(ˆr) = N /√2 Dj∗ (ˆr)e (ˆr) Dj∗ (ˆr)e (ˆr) , (3a) jm j m,−1 −1 − m,1 +1 Yj(em)(ˆr) = Nj/√2 (cid:8)Dmj∗,−1(ˆr)e−1(ˆr)+Dmj∗,1(ˆr)e+1(ˆr) (cid:9), (3b) Yj(0m)(ˆr) = NjDmj∗,0(cid:8)(ˆr)e0(ˆr) = Yjm(ˆr)ˆr, Nj = [(2j +(cid:9)1)/4π]1/2, (3c) where e (ˆr) = (e (ˆr) ie (ˆr))/√2; e (ˆr) ϑˆ = (cosθcosφ,cosθsinφ, sinθ), e (ˆr) ±1 x y x y ∓ ± ≡ − ≡ ϕˆ = ( sinφ,cosφ,0) are the unit vectors tangential to the sphere; φ (θ) is the azimuthal − (polar) angle of the unit vector ˆr = r/r = (sinθcosφ,sinθsinφ,cosθ) e (ˆr) e (ˆr). 0 z ≡ ≡ (Hats will denote unit vectors and an asterisk will indicate complex conjugation.) Note that, for the irreducible representation of the rotation group with the angular num- ber j, the D-functions, Dj (α,β,γ) = exp( imα)dj (β)exp( iµγ), give the elements of mν − mµ − the rotation matrix parametrized by the three Euler angles [53, 54]: α, β and γ. In formu- las (3) and throughout this paper, we assume that γ = 0 and Dj (ˆr) Dj (φ,θ,0). These mν ≡ mν D-functions meet the following orthogonality relations [53, 54] hDmj∗ν(ˆr)Dmj′′ν(ˆr)iˆr = 2j4+π 1 δjj′δmm′ , (4) 2π π where f ˆr dφ sinθdθf(ˆr) and f(ˆr) f(φ,θ). The orthogonality condition (4) h i ≡ ≡ Z0 Z0 and Eqs. (3) show that a set of vector spherical harmonics is orthonormal: hYj(αm)∗(ˆr)·Yj(β′m)′(ˆr)iˆr = δαβ δjj′δmm′ . (5) We can now use the relations [53] N Dj∗(ˆr) = Y (ˆr), (6) j m0 jm i N Dj∗ (ˆr) = n ∂ + ∂ Y (ˆr), n [j(j +1)]−1/2 (7) j m±1 j ∓ θ sinθ φ jm j ≡ (cid:20) (cid:21) where ∂ stands for a derivative with respect to x and Y (ˆr) is the normalized spherical x jm function (2j +1)(j m)! Y (φ,θ) = N exp(imφ)dj (θ) = ( 1)m − exp(imφ)P m(cosθ) (8) jm j m,0 − s 4π(j +m)! j expressed in terms of the associated Legendre polynomial of degree j and order m ( 1)m/(2jj!)(1 x2)m/2∂j+m(x2 1)j, m > 0 P m(x) = − − x − , (9) j ((−1)|m|(j −|m|)!/(j +|m|)!Pj|m|(x), m < 0 5 and derive the following expressions for the magnetic and electric vector spherical functions m Y(m)(ˆr) = in [∂ Y ]ϕˆ i Y ϑˆ = jm − j θ jm − sinθ jm n LY = iˆr hY(e), i (10) j jm − × jm m Y(e)(ˆr) = n [∂ Y ]ϑˆ +i Y ϕˆ = jm j θ jm sinθ jm n r∇Y = hiˆr Y(m), i (11) j jm − × jm where L is the operator of angular momentum given by iL = r ∇ = ϕˆ ∂ ϑˆ [sinθ]−1∂ . (12) θ φ × − Formulas (10) and (11) give the vector spherical harmonics (3) rewritten in the well-known standard form [55]. B. Wave functions and T–matrix (λ) The electric field (2a) is completely described by the coefficients p (r) and similarly { jm } (λ) the magnetic field (2b) is described by q (r) with λ = 0,e,m . In order to find the { jm } { } coefficient functions we can use separation of variables. This implies that the expansions (2) must be inserted into Maxwell’s equations (1). The coefficient functions then can be derived by solving the resulting system of equations. In the simplest case of an isotropic medium the coefficient functions can be expressed in terms of spherical Bessel functions, j (x) = j [π/(2x)]1/2J (x), and spherical Hankel functions [56], h(1,2)(x) = [π/(2x)]1/2H(1,2) (x), j+1/2 j j+1/2 and their derivatives. Alternatively, it is well-known (a discussion of the procedure can be found, e.g., in Ref. [57]) that solutions of the scalar Helmholtz equation, (∇2 + k2)ψ(r) = 0, taken in the form ψ(α) = n z(α)(ρ)Y(ˆr), n [j(j +1)]−1/2, (13) jm j j j ≡ (α) whereρ = kr andz (ρ)iseitherasphericalBessel orHankelfunction, canbeusedtoobtain j the following solenoidal solutions of the vector Helmholtz equation, ∇ [∇ Ψ] = k2Ψ: × × M(α)(ρ,ˆr) = Lψ(α) = z(α)(ρ)Y(m)(ˆr), (14) jm jm j jm j(j +1) N(α)(ρ,ˆr) = ik−1∇ M(α) = z(α)(ρ)Y(0)(ˆr)+Dz(α)(ρ)Y(e)(ˆr), (15) jm − × jm ρ j jm j jm p whereDf(x) x−1∂ (xf(x)). Thevectorwave functions, M(α) andN(α), arelinkedthrough ≡ x jm jm the identities i∇ M(α) = kN(α), i∇ N(α) = kM(α) (16) − × jm jm × jm jm and their linear combination represents the expansions (2) over the vector spherical harmon- ics. 6 There are three cases of these expansions that are of particular interest. They corre- spond to the incident wave, E ,H , the outgoing scattered wave, E ,H and the inc inc sca sca { } { } electromagnetic field inside the scatterer, E ,H : p p { } E = α(α)M(α)(ρ ,ˆr)+β(α)N(α)(ρ ,ˆr) , α inc,sca,p (17a) α jm jm i jm jm i ∈ { } jm X(cid:2) (cid:3) H = n /µ α(α)N(α)(ρ ,ˆr) β(α)M(α)(ρ ,ˆr) , (17b) α i i jm jm i − jm jm i jm X(cid:2) (cid:3) j (ρ), α = inc j med, α inc,sca i = ∈ { } , z(α)(ρ ) = h(1)(ρ), α = sca , (17c) p, α = p j i  j ( j (ρ ), α = p j p where ρ = k r ρ, ρ = k r nρ, and n = n /n is the ratio of refractive indexes med med p p p med ≡ ≡ also known as the optical contrast. Thus outside the scatterer the electromagnetic field is a sum of the incident wave field (inc) (sca) (1) with z (ρ) = j (ρ) and the scattered waves with z (ρ) = h (ρ) as required by the j j j j Sommerfeld radiation condition. In the far field region (ρ 1), the asymptotic behaviour of the spherical Bessel and ≫ Hankel functions is known [56]: ij+1h(1)(ρ),ijDh(1)(ρ) exp(iρ)/ρ, (18) j j ∼ ( i)j+1h(2)(ρ),( i)jDh(2)(ρ) exp( iρ)/ρ, (19) − j − j ∼ − ij+1j (ρ),ij+1Dj (ρ) exp(iρ) ( 1)jexp( iρ) /(2ρ). (20) j j+1 ∼ − − − (cid:2) (1) (cid:3) So, the spherical Hankel functions of the first kind, h (ρ), describe the outgoing waves, j (2) whereas those of the second kind, h (ρ), represent the incoming waves. j The incident field is the field that would exist without a scatterer and therefore includes both incoming and outgoing parts (see Eq. (20)) because, when no scattering, what comes in must go outwards again. As opposed to the spherical Hankel functions that are singular at the origin, the incident wave field should be finite everywhere and thus is described by the regular Bessel functions j (ρ). j (inc) (inc) Now the incident wave is characterized by amplitudes α , β and the scattered out- jm jm (sca) (sca) going waves are similarly characterized by amplitudes α , β . So long as the scattering jm jm (sca) (sca) problem is linear, the coefficients α and β can be written as linear combinations of jm jm (inc) (inc) α and β : jm jm α(sca) = T11 α(inc) +T12 β(inc) , jm jm,j′m′ j′m′ jm,j′m′ j′m′ jX′,m′h i β(sca) = T21 α(inc) +T 22 β(inc) . (21) jm jm,j′m′ j′m′ jm,j′m′ j′m′ jX′,m′h i These formulae define the elements of the T–matrix in the most general case. In general, the outgoing wave with angular momentum index j arises from ingoing waves of all other indices j′. In such cases we say that the scattering process mixes angular 7 momenta [8]. The light scattering from uniformly anisotropic scatterers [15, 58] provides an example of such a scattering process. In simpler scattering processes, by contrast, such angular momentum mixing does not take place. Many quantum scattering processes and classical Mie scattering belong to this category. For example, radial anisotropy keeps intact spherical symmetry of the scatterer [10, 15, 18]. The T–matrix of a spherically symmetric scatterer is diagonal over the angular momenta and the azimuthal numbers: Tnn′ = jj′,mm′ δjj′δmm′Tjnn′. In order to calculate the elements of T-matrix and the coefficients α(p) and β(p), we jm jm need to use continuity of the tangential components of the electric and magnetic fields as boundary conditions at r = R (ρ = k R x). p med p ≡ (p) (p) So, the coefficients of the expansion for the wave field inside the scatterer, α and α , jm jm are expressed in terms of the coefficients describing the incident light as follows (inc) α (p) jm iα = , µ = µ /µ , (22) jm µ−1v (x)u′(nx) n−1v′(x)u (nx) p med j j − j j (inc) β (p) jm iβ = , n = n /n , (23) jm n−1v (x)u′(nx) µ−1v′(x)u (nx) p med j j − j j (1) where x = k R , u (x) = xj (x) and v (x) = xh (x). The similar result relating the med p j j j j scattered wave and the incident wave n−1u′(x)u (nx) µ−1u (x)u′(nx) α(sca) = T11α(inc) = j j − j j α(inc), (24) jm j jm µ−1v (x)u′(nx) n−1v′(x)u (nx) jm j j − j j µ−1u (x)u′(nx) n−1u′(x)u (nx) β(sca) = T22β(inc) = j j − j j β(inc), (25) jm j jm n−1v (x)u′(nx) µ−1v′(x)u (nx) jm j j − j j defines the T-matrix for the simplest case of a spherically symmetric scatterer. In addition, since the parity of electric and magnetic harmonics with respect to the spatial inversion ˆr ˆr ( φ,θ φ+π,π θ ) is different → − { } → { − } Y(m)( ˆr) = ( 1)jY(m)(ˆr), Y(e)( ˆr) = ( 1)j+1Y(e)(ˆr), (26) jm − − jm jm − − jm where f(ˆr) f(φ,θ) and f( ˆr) f(φ + π,π θ), they do not mix provided the mirror ≡ − ≡ − symmetry has not been broken. In this case the T-matrix is diagonal and T12 = T21 = 0. j j The diagonal elements T11 and T22 are also called the Mie coefficients. j j III. INCIDENT WAVE BEAMS The formulas (22)- (25) are useful only if the expansion for the incident light beam is known. First we briefly review the most studied and fundamentally important case where the incident light is represented by a plane wave. 8 A. Plane waves The electric field of a transverse plane wave propagating along the direction specified by a unit vector kˆ is inc E = E(inc)exp(ik r), E(inc) = E(inc)e (kˆ ), k = kkˆ . (27) inc inc · ν ν inc inc inc ν=±1 X where the basis vectors e (kˆ ) are perpendicular to kˆ . Then the vector version of the ±1 inc inc well known Rayleigh expansion (see, for example, [2]) ∞ l exp(iρkˆ ˆr) = 4π ilj (ρ)Y (ˆr)Y∗ (kˆ), ρ kr (28) · l lm lm ≡ l=0 m=−l X X which is given by e (kˆ)exp[iρ kˆ ˆr ] = α Dj (kˆ) iνM (ρ,ˆr) N (ρ,ˆr) , ν = 1, (29) ν · j mν jm − jm ± (cid:0) (cid:1) Xjm n o where α = ij+1[2π(2j + 1)]1/2, immediately gives the expansion coefficients for the plane j wave α(inc) = iα Dj (kˆ )νE(inc), β(inc) = α Dj (kˆ )E(inc), (30) jm j mν inc ν jm − j mν inc ν ν=±1 ν=±1 X X where Dj is the Wigner D-function. mm′ In the far field region, the electric field of scattered wave is related to the polarization vector of the plane wave through the scattering amplitude matrix as follows [2, 8, 59] Eν(sca) ≡ (eν∗(kˆsca),Esca) = ρ−1exp(iρ) Aνν′(kˆsca,kˆinc)Eν(i′nc), ν = ±1 (31) ν′=±1 X where kˆ = ˆr. For a spherically symmetric scatterer, the expression for the scattering sca amplitude matrix in terms of T-matrix is given by Aνν′(kˆsca,kˆinc) = Ajνν′(kˆsca,kˆinc) = j X = i (j +1/2)D˜j (kˆ ,kˆ ) νν′T 11 iνT12 +iν′T21 +T22 , (32a) − νν′ sca inc j − j j j j X (cid:2) (cid:3) D˜j (kˆ ,kˆ ) = Dj∗(kˆ )Dj (kˆ ). (32b) νν′ sca inc mν sca mν′ inc m X Equation (32b) shows that the scattering amplitude matrix (32a) depends only on the angle between kˆ and kˆ . All far-field scattering characteristics of the system can be computed inc sca from the scattering amplitude matrix. 9 B. Far-field matching Now we consider a more general case where an incident electromagnetic wave is written as a superposition of propagating plane waves: E (r) E (ρ,ˆr) = exp(iρkˆ ˆr)E (kˆ) , E (kˆ) = E (kˆ)e (kˆ), (33a) inc inc inc kˆ inc ν ν ≡ h · i ν=±1 X n H (r) H (ρ,ˆr) = exp(iρkˆ ˆr) kˆ E (kˆ) , (33b) inc inc inc kˆ ≡ µ h · × i (cid:2) (cid:3) 2π π where f dφ sinθ dθ f. kˆ k k k h i ≡ Z0 Z0 Our first step is to examine asymptotic behavior of the wave field (33) in the far-field region, ρ 1. The results can be easily obtained by using the asymptotic formula for a ≫ plane wave (see, e.g., [5]) 2πi exp(iρkˆ ˆr) − exp(iρ)δ(kˆ ˆr) exp( iρ)δ(kˆ +ˆr) at ρ 1, (34) · ∼ ρ − − − ≫ (cid:2) (cid:3) where δ(kˆ ˆr) is the solid angle Dirac δ-function symbolically defined through the expansion ∓ ∞ l δ(kˆ ˆr) = Y ( ˆr)Y∗ (kˆ). (35) ∓ lm ± lm l=0 m=−l X X Applying the relation (34) to the plane wave superposition (33a) gives the electric field of the incident wave in the far-field region 1 E (ρ,ˆr) E(∞)(ρ,ˆr) = exp(iρ)E (ˆr)+exp( iρ)E (ˆr) , (36) inc ∼ inc ρ out − in E (ˆr) = E ( ˆr), (cid:2) (cid:3) (37) in out − − where E (ˆr) is the far-field angular distribution for the outgoing part of the electric field out of the incident wave: E (ˆr) = 2πiE (ˆr) = E(out)(ˆr)e (ˆr)+E(out)(ˆr)e (ˆr), (38) out − inc θ θ φ φ whereas the incoming part of the incident wave is described by the far-field angular distri- bution E (ˆr). in The result for the far-field distribution of the magnetic field (33b) can be written in the similar form: 1 H (ρ,ˆr) H(∞)(ρ,ˆr) = exp(iρ)H (ˆr)+exp( iρ)H (ˆr) , (39) inc ∼ inc ρ out − in H (ˆr) = H ( ˆr), (cid:2) (cid:3) (40) in out − − µ/nH (ˆr) = ˆr E (ˆr), µ/nH (ˆr) = ˆr E ( ˆr). (41) out out in out × × − Formulas (36)-(41) explicitly show that, in the far-field region, the incident wave field is defined by the angular distribution of the outgoing wave (38). In particular, from these 10 formulas, it is not difficult to obtain the far-field expression for the Poynting vector of the incident wave S = c/(8π)Re(E H∗ ) inc inc × inc S (ρ,ˆr) S(∞)(ρ,ˆr) = ρ−2 S (ˆr)+S (ˆr) , (42) inc ∼ inc in out S (ˆr) = S ( ˆr), µ/nS (ˆr) = c/(8π) E (ˆr) 2ˆr, (43) in out (cid:8)out (cid:9)out − − | | where E (ˆr) 2 = (E (ˆr) E∗ (ˆr)). From this expression it immediately follows that the | out | out · out flux of Poynting vector of the outgoing wave, S (ˆr), through a sphere of sufficiently large out radius is exactly balanced by the flux of Poynting vector of the incoming wave, S (ˆr). inc Alternatively, the far-field distribution of an incident light beam, E (ˆr), can be found out from the expansion over the vector spherical harmonics (17a). The far-field asymptotics for the vector wave functions that enter the expansion for the incident wave (17) ( i)j+1 M(inc)(ρ,ˆr) − exp(iρ)Y(m)(ˆr) exp( iρ)Y(m)( ˆr) , (44) jm ∼ 2ρ jm − − jm − ( i)j (cid:2) (cid:3) N(inc)(ρ,ˆr) − exp(iρ)Y(e)(ˆr) exp( iρ)Y(e)( ˆr) , (45) jm ∼ 2ρ jm − − jm − (cid:2) (cid:3) can be derived from Eqs. (14)-(15) with the help of the far-field relation (20). Substituting Eqs. (44) and (45) into the expansion (17a) gives the far-field distribution of the form (36) with E (ˆr) = 2−1 ( i)j+1α(inc)Y(m)(ˆr)+( i)jβ(inc)Y(e)(ˆr) . (46) out − jm jm − jm jm Xjmh i The coefficients of theincident wave cannow beeasily found astheFourier coefficients of the far-field angular distribution, E , expanded using the vector spherical harmonics basis (3). out The final result reads αj(imnc) = 2ij+1hYj(mm)∗(ˆr)·Eout(ˆr)iˆr = iαj νhDmjν(kˆ)Eν(kˆ)ikˆ, (47a) ν=±1 X βj(imnc) = 2ijhYj(em)∗(ˆr)·Eout(ˆr)iˆr = −αj hDmjν(kˆ)Eν(kˆ)ikˆ. (47b) ν=±1 X A comparison between the expressions on the right hand side of Eq. (47) and those for the plane wave (30) shows that, in agreement with the representation (33a), the result for plane waves represents the limiting case where the angular distribution is singular: E (kˆ) = ν E(inc)δ(kˆ kˆ ). ν inc − By using Eqs. (10) and (11) formulas (47) can be conveniently rewritten in the explicit form αj(imnc) = 2njij+1hYj∗m(ˆr)(L·Eout(ˆr))iˆr = 2π π 2n ij dφ dθY∗ (φ,θ) ∂ (sinθE(out)) ∂ E(out) , (48a) j jm θ φ − φ θ Z0 Z0 h i βj(imnc) = −2nj ijhYj∗m(ˆr)(r∇·Eout(ˆr))iˆr = 2π π 2n ij dφ dθY∗ (φ,θ) ∂ (sinθE(out))+∂ E(out) , (48b) − j jm θ θ φ φ Z0 Z0 h i

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.