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

Multipolar analysis of the second harmonic generated by dielectric particles

Open Access Open Access

Abstract

We studied the second harmonic generation (SHG) by two-dimensional dielectric particles made of a centrosymmetric high-index material. The calculated scattered fields at the fundamental and harmonic frequencies are decomposed on a multipolar basis, allowing the evaluation of the relative strengths of the multipolar resonances excited at the particle. With these tools, we studied the strength of the multipoles that produce the second harmonic field and the role played by those excited at the fundamental frequency.

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

1. Introduction

The absorption and scattering of light by small particles is a subject of relevance in several fields of science and technology and constitutes a problem of long-standing interest [1]. With the recent development of nano-optics, the importance of understanding and manipulating the optical response of nano-scale structures has become even more evident. Since it is well-known that metallic nanoparticles support electric-type resonances, a very common strategy for achieving this manipulation is based on the exploitation of the localized surface plasmon resonances (LSPR) of metallic nanoparticles [2,3]. These nanoresonators, which can be spectrally tuned by modifying the shape of the particle, provide some of the building blocks for the design of optical metamaterials [4].

A key ingredient in the search of metamaterials with novel optical properties is the ability to excite resonances of a magnetic character. In this respect, the split ring resonators (SRRs) proposed by Pendry et al. [5] constitute an important development in the field, and have motivated much of the work on the subject. When combined with structures that support electric-type resonances, SRRs can be used in the design of artificial materials with unusual optical properties. More recently, however, the realization that high-index dielectric nanoparticles can support strong magnetic-type resonances [6–12], has opened new alternatives for the design optical metamaterials. Interestingly, this kind of excitation occurs even for relatively simple structures like spheres or infinite cylinders, for which analytic solution to the scattering problem exist and have been known for over a century [13]. In addition, these structures present lower losses that their metallic counterparts.

Given the resonant nature of the metamaterial designs, it is clear that the internal fields can be quite strong and it is then not surprising that the possibility of exciting nonlinear optical effects in their constituents have also attracted some attention. Of particular interest to us here is the nonlinear generation of second harmonic radiation. The case of metallic nanoparticles, in which the excitations are mainly of dipolar electric origin, has been a well studied subject [14–17]. The role of dipolar magnetic resonances in the SHG by metallic structures and, in particular, by arrays of U-shaped nanoparticles (the optical version of the SRRs) has also been investigated by several groups, and it has been reported that such arrays can enhance the SHG efficiency with respect to those of other noncentrosymmetric particles. Nevertheless, the physical origin of the enhancement has been a point of some discussion in the literature. The initial studies seemed to indicate that the enhancement was linked to resonances of a magnetic-dipole type [18], but later studies that compare the response of arrays of magnetic split-ring resonators and arrays of complementary split-ring resonators (electric response) seem to indicate that in general there was no correlation between SHG efficiency and the excitation of the magnetic-dipole moment [19]. More recently, it was shown that the reported enhancement of the second harmonic in split-ring resonator based media is driven by the electric rather than the magnetic properties of the structure [20].

The SHG by dielectric particles that support magnetic-type resonances has received less attention, but is also a subject of interest and, as could be expected, the role played by the magnetic-type resonances has been one of the issues of interest. Most of these works deal with particles made on non-centrosymmetric materials and, in particular, AlGaAs and GaAs [21–25]. In a recent publication Kruk et al. [24] studied the role played by the excitation of multipoles in the SHG by nanodisks made from noncentrosymmetric materials, and found that both, electric and magnetic resonances were important for the nonlinear response. More recently, Smirnova at al. [26] reported studies of SHG by high-index dielectric nanoparticles made of centrosymmetric materials, and discussed the contributions of the different nonlinear sources and multipolar excitations in the generated field.

In this paper, we study the SHG by small two-dimensional particles made of centrosymmetric high index dielectrics in a spectral region in which strong magnetic dipole resonances can be excited. Since the method of calculation has been explained in some detail in previous work [27,28], we only present enough material to make the paper self-contained and facilitate the discussion. The calculation of the surface sources and the scattered fields at ω and 2ω are quite rigorous, and the nonlinear response of the medium is modeled on the basis the dipollium model [29]. Both, the linear and second harmonic fields are decomposed into their multipolar components, allowing the visualization of the elementary contributions to the second harmonic field, as well as those that drive the nonlinear processes.

2. Theoretical methods

The physical situation considered is illustrated in Fig. 1. A two-dimensional particle defined by the boundary between vacuum (medium I) and a nonlinear material (II) is illuminated by a p-polarized time-harmonic plane wave of frequency ω propagating in the direction of the x1 axis. Medium II is assumed to be isotropic, homogeneous, and nonmagnetic and is characterized by its permittivity ϵ(ω). Apart from the condition of invariance along x2, the shape of the particle can be rather arbitrary, and the curve describing its profile can be conveniently represented in terms of two functions of a parameter t. These two functions represent the x1 and x3 coordinates of the vector-valued function

rs(t)=[ξ(t),η(t)].

 figure: Fig. 1

Fig. 1 Schematic diagram of the geometry considered.

Download Full Size | PDF

The particle shapes employed in the calculations were generated using Gielis superformula [30,31].

The solution to the problem of scattering in the SHG depends on the solution of the linear problem and on the non-linearities of the medium, represented by the non-linear bulk polarization and the nonlinear surface susceptibilities. Since the medium composing the particle is assumed isotropic, most of the contribution to SHG will arise from nonlinear sources located on the surface of the particle. The procedure employed to calculate the scattered fields is fairly rigorous and based on the numerical solution of coupled surface integral techniques that determine the source functions for the fundamental and second harmonic fields [27]. As we have already mentioned, we only present the results that are needed for the discussion, focusing on the aspects that differ from the treatment presented in [27], which was formulated to deal with the case of metallic particles.

2.1. The nonlinear sources

For isotropic media the nonlinear polarization takes the general form [32]

PNL(r|2ω)=α[E(r|ω)]E(r|ω)+βE(r|ω)[E(r|ω)]+γ[E(r|ω)E(r|ω)],
where E(r|ω) represents the electric field at the fundamental frequency, and α, β, and γ are frequency-dependent parameters that characterize the bulk nonlinear response of the medium.

Due to the invariance of the system along x2, the scattering problem is essentially of a scalar nature and, at both, the ω and 2ω frequencies, there are two independent modes of polarization (s and p). With our assumptions, illumination with purely s or p waves produces only p polarization at the second harmonic frequency and, in order to produce an s-polarized field at 2ω, a combination of s and p waves would be required [27]. To keep the paper to a manageable size, we consider only the case of p-polarized illumination. The second harmonic signal produced under s-polarized illumination is at least two orders of magnitude smaller than that the one obtained with p polarization.

We will denote by ψ(I)(r, θ|Ω) the x2 component of the magnetic field H2(I)(r,θ|Ω) in medium I, where, Ω stands for the fundamental (ω) or second-harmonic (2ω) frequencies. The fields are assumed to have a time dependence of the form exp(−iΩt), but the explicit reference to it will be suppressed.

The nonlinear boundary conditions satisfied by the (p-polarized) harmonic field can be written in the general form [27]

ψp(I)(t|2ω)ψp(II)(t|2ω)=Ap(t|ω),
1ϵI(2ω)ϒp(I)(t|2ω)1ϵII(2ω)ϒp(II)(t|2ω)=Bp(t|ω),
where ψp(I,II)(t|2ω) and ϒp(I,II)(t|2ω) are the x2-components of, respectively, the magnetic field and its non-normalized normal derivative, both evaluated at the surface. The nonlinear sources Ap(t|ω) and Bp(t|ω) that appear on the right hand side of Eqs. (3a) and (3b), are responsible for the generation of the second harmonic field, and can be determined from the solutions of the electromagnetic problem at frequency ω and the nonlinear properties of the material.

Starting from the expressions found in [27], and after some algebra, we obtain the following expressions for the nonlinear sources,

Ap(t|ω)=8πicωχ ϵI(ω)1ϕ2(t)ϒp(I)(t|ω)dψp(I)(t|ω)dt,
Bp(t|ω)=8πicω[χddt(1ϕ(t)dψp(I)(t|ω)dt)2+χϵI2(ω)ddt(ϒp(I)(t|ω)ϕ(t))2+χbd[ψp(I)(t|ω)]2dtηd[ψs(I)(t|ω)]2dt],
where ϕ(t)=[ξ(t)]2+[η(t)]2 and c represents the speed of light in vacuum. The second order susceptibilities that appear in these equations are related to the bulk nonlinear constants that appear in Eq. (2), and to the elements of the second-order effective surface susceptibility tensor χijk, defined as follows [27,33]
χ =χttzs,
χ=χzzzs+α/2+γϵII(2ω)ϵII2(ω),
χ=χztts+α/2+γϵII(2ω),
η=χztts+γϵII(2ω),
χb=α2ϵII(ω)ϵII(2ω)(ωc)2,
where the terms containing α and/or γ are bulk contributions and the pure surface susceptibilities χijks are given below in Eq. (7).

As we have already mentioned, we are interested in the case of high refractive index dielectric particles and, in particular, silicon. The required nonlinear susceptibilities and the bulk respond are calculated with the so-called dipollium model [29], which models the medium as a collection of polarizable entities with a single resonant frequency that are driven by the field. The permittivity of such a medium may be written as ϵ(ω) = 1 + 4πnBχ(ω), where nB is the density of the polarizable entities, and the linear susceptibility is of the form

χ(ω)=e2/mω02ω2iΓω,
where ω0 is the resonant frequency of the oscillator and Γ is a damping factor. From a practical point of view, the parameters nB, ω0 and Γ, may be viewed as free parameters that can be used to fit the experimental data of the permittivity of the material in the region of interest. Setting the oscillator strength F = 4πnBe2/m = 3.65 × 1032 s−2, the frequency of resonance ω0 = 5.9 × 1015 s−1, and the damping constant Γ = 3.0 × 1014 s−1 one obtains a fairly good approximation to the dielectric function of silicon in the region 0.4 − 3 µm. From this fit, we estimate nB ≈ 1.15 × 1023 cm−3.

With this model, the frequency-dependent parameters α(ω), β(ω), and γ(ω) that enter into the general expression for the bulk nonlinear polarization [27] can be written in terms of the linear permittivity as [29]

α(ω)=ϵ(ω)132π2nBe[4ϵ(2ω)ϵ(ω)3],
β(ω)=ϵ(ω)132π2nBe,
γ(ω)=ϵ(ω)132π2nBe[ϵ(2ω)1],
while the only three distinct components of the second order surface susceptibility tensor χijks are given by [29]
χzzzs(ω)=(ϵ(ω)1)232π2nBeϵ(ω)2{2ϵ(ω)ϵ(2ω)ϵ(ω)ϵ(2ω)[ϵ(2ω)ϵ(ω)]+[1ϵ(2ω)][ϵ(2ω)ϵ(ω)]2ln(ϵ(ω)ϵ(2ω))},
χttzs(ω)=χtzts(ω)=132π2nBe[ϵ(ω)1]2ϵ(ω),
χztts(ω)=0.

The main difference between the free-electron model (FEM) adopted in [27] and the dipollium model employed here, is that the FEM assumes a linear susceptibility of the form χ(ω) = −e2/2, and that the number density is that of the free electrons instead of that of the dipole oscillators in the material. It is then clear that the dipollium model reduces to the FEM when Γ and ω0ω.

With Eqs. (5), (7) and (8), we can determine the nonlinear response of the material in terms of the linear permittivity ϵ(Ω). We point out that the constant factor (nBe)1 is common to all the nonlinear bulk and surface contributions [Eqs. (7) and (8)], and the second harmonic scattered field ψsc(I)(r,θ|2ω) will be scaled with this factor.

At this stage, it is worth saying a few words about the relative strengths of the second harmonic contributions arising from the bulk and the surface. For plasmonic structures, it is commonly assumed that the bulk contribution contribution is much smaller than that due to the surface. The case of dielectrics has only been addressed recently [34], with the conclusion that the surface and bulk effects are comparable. As Eqs. (3)(5) show, the nonlinear sources have contributions from both, the bulk and the surface, and by switching off selectively one or the other, it is possible to assess their relative importance for the results. In the calculations presented here, the bulk contribution was three orders of magnitude lower than the surface contribution.

Inspecting the nonlinear sources of Eq. (4), we observe that the last term of Eq. (4b) represents the generation of a p-polarized signal with an s-polarized excitation. As we have already mentioned, in our calculations this p-polarized signal is two orders of magnitude smaller than the one obtained under p-polarized illumination.

2.2. The scattered field and the multipolar decomposition

For large values of r, the scattered field can be written as [27]

ψsc(I)(r,θ|Ω)=eiπ4eikΩr[8πkΩr]1/2S(θ|Ω),
where kΩ = Ω/c, and the scattering amplitude is
S(θ|Ω)=S[ikΩ[η(t)cosθξ(t)sinθ]ψp(I)(t|Ω)ϒp(I)(t|Ω)]×exp{ikΩ[ξ(t)cosθ+η(t)sinθ]}dt.

The source functions ψp(I)(t|Ω) and ϒp(I)(t|Ω) are calculated by solving numerically a pair of coupled integral equations and imposing the linear and nonlinear boundary conditions [27].

Outside a virtual cylinder that contains entirely the scattering particle, the field can be expressed in terms of cylindrical harmonics [35]

ψsc(I)(r,θ|Ω)=m=cmimHm(1)(kΩr)eimθ,
where Hm(1)(kΩr) are the Hankel functions of the first kind and cm are the expansion coefficients. The series represented by Eq. (11) is the natural basis for solving analytically the problem of scattering by a normally-illuminated infinite cylinder (Mie-type solution) but, since the basis is complete and orthogonal, it provides a formal representation for the field scattered by an arbitrarily-shaped two-dimensional particle.

Following [36], we find it convenient to rewrite Eq. (11) as

ψsc(I)(r,θ|Ω)=m=0cm+ψm+(r,θ|Ω)+m=1cmψm(r,θ|Ω),
where, for m ≠ 0,
ψm±(r,θ,|Ω)=imHm(1)(kΩr){cosmθisinmθ},
cm±=(cm±cm),
and for m = 0
ψ0+(r,θ,|Ω)=H0(1)(kΩr),
c0+=c0.

When considering the expansion Eq. (12) at the fundamental frequency, we will denote the coefficients cm± by am±, while those at the second harmonic frequency will be denoted by bm±.

We note that the base ψm+(r,θ,|Ω) contains functions that are symmetric about the direction of incidence, whereas the base ψm(r,θ,|Ω) contains functions that are antisymmetric about that direction. In other words, Eq. (12) is a representation of the field as the sum of an even and an odd function and, in general, the two contributions are needed to describe it completely [36]. It can also be observed that, for any order m, the ± contributions are rotated by an angle π/2m between them. The first few terms of the even and odd bases are illustrated in Fig. 2. It is also worth pointing out that, for the case of p polarization, the first three coefficients represent the magnetic dipole (MD, term with c0), the electric dipole (ED, terms with c1±), and the electric quadrupole contributions (EQ, terms with c2±).

 figure: Fig. 2

Fig. 2 Illustration of the lowest order functions of the symmetric and antisymmetric bases. In the symmetric base the lowest order terms represent the magnetic dipole (MD), the electric dipole (ED) and the electric quadrupole (EQ) contributions. In the antisymmetric base they are the electric dipole (ED) and the electric quadrupole (EQ) contributions.

Download Full Size | PDF

The strength of the multipolar contributions to the field scattered by a particle with non-circular cross sections, can be determined by evaluating its overlap with the base functions. Thus, the coefficients of the expansion can be calculated from the expression

cm±=02πψsc(I)(r0,θ|Ω)ψm±(r0,θ|Ω)dθ02π|ψm±(r0,θ|Ω)|2dθ,
where r0 is the radius of an imaginary circle centered at the origin that completely contains the particle.

Based on the far field approximation Eq. (9), for m ≠ 0 the overlap integral reduces to the simple expression

c2±=14π02πS(θ|Ω){cosmθisinmθ}dθ,
and for m = 0,
c0=18π02πS(θ|Ω)dθ.

It is important to mention, however, that the higher order multipoles decay more rapidly than the lower order ones and, therefore, that their strength can only be determined from near field data. In consequence, the coefficients of the high order multipoles cannot be estimated from far field data.

2.3. The scattering efficiency

The total power scattered by a two-dimensional particle can be evaluated by constructing an imaginary cylinder of length L and radius r0 that encloses the particle (see Fig. 1). Then, at the fundamental frequency the scattering efficiency of the particle can be expressed as [27]

Qsc(ω)=Psc(ω)Pinc(ω)=02πqsc(θ|ω)dθ,
where Pinc(ω) represents the power intersecting a particle with geometrical cross section σg = DL, and the scattering efficiency is given by
qsc(θ|ω)=18πD(ω/c)|S(θ|ω)|2|ψ0|2,
and where ψ0 represents the amplitude of the incident field.

The scattering efficiency at 2ω can be defined as

Qsc(2ω)=σgPsc(2ω)[Pinc(ω)]2=02πqsc(θ|2π)dθ,
where we have used the fact that with our assumptions the scattered light is originated from p-type waves will always be p-polarized. The differential scattering efficiency for the second-harmonic radiation is given by
qsc(θ|2ω)=12ωD|S(θ|2ω)|2|ψ0|4.

It is worth pointing out that in contrast with the linear case, in which the scattering efficiency has no dimensions, by its definition, the second-harmonic scattering efficiency has units of area over power (e. g. cm2/Watt).

It is straightforward to show that in terms of the expansion coefficients of Eq. (12), the scattering efficiency can be written as [1,28]

Qsc(Ω)=[|c0|2+12m=1[|cm+|2+|cm|2]]×{2kD,Ω=ω,8πωD|ψ0|4,Ω=2ω,
and the normalized scattered intensity as
qsc(θ|Ω)=[c0+m=1[cm+cosmθ+icmsinmθ]]2×{1πkD,Ω=ω,16ωD|ψ0|4,Ω=2ω.

3. SHG by symmetric particles

With reference to Fig. 1, we shall say that a particle is symmetric if it is symmetric about the plane x1x2. It is clear that, for such particles, the linear source functions and the linearly scattered field must also be symmetric about the x1x2 plane; that is to say that ψsc(I)(r,θ|ω)=ψsc(I)(r,2πθ|ω). A necessary implication of this symmetry is that the coefficients am=0 and therefore, that only the base ψm+(r,θ|ω) is needed for the representation of the field.

Similarly, at the second harmonic frequency, the field will be antisymmetric about the direction of incidence. This can be deduced from the expressions for the source functions given in Eq. (4) (the derivative of an even function is an odd function) or by invoking some simple physical arguments compatible with the assumed symmetry. With reference to Fig. 3(a), let us assume momentarily that the field scattered at 2ω is symmetric about the x1x2 plane. We allow now to pass the time and consider the situation after half a cycle for frequency ω (a full cycle at 2ω). The situation is illustrated on the left panel of Fig. 3(a); we see that the sign of the field at ω has changed, but that of the field at 2ω is unchanged. If we now flip the lower part of Fig. 3(a) about the x1x2 plane, we would end up with a situation that is different from the original one (the sign of the 2ω field would be different from the one assumed). This means that the assumption of a symmetric scattered field at 2ω led us to some unphysical or contradictory situation. On the other hand, is one assumes that the scattered field at 2ω is antisymmetric, as on Fig. 3(b), the same arguments lead to a physically compatible situation with the one assumed originally. Then, we have that ψsc(I)(r,θ|2ω)=ψsc(I)(r,2πθ|2ω), and the coefficients bm+=0, so that only the base ψm(r,θ|2ω) is needed to represent the scattered field. An immediate conclusion that can be obtained from this observation is that, for symmetric particles, b0+=0 and, thus, that even when the particles support a strong magnetic resonance at 2ω, this is not excited through the nonlinear process.

 figure: Fig. 3

Fig. 3 Illustration of situations in which the second harmonic scattered field is assumed to be symmetric (a) and antisymmetric (b).

Download Full Size | PDF

To substantiate these arguments, we consider the specific example of SHG from a silicon cylinder (2D particle of circular cross section) of radius a = 150 nm. The case of the cylinder is particularly interesting, as it also allows us to test the approach by comparing the numerical results with those obtained with the analytical solution [27]. The wavelength of incidence spans the range from 800 to 3000 nm, so that the second harmonic covers the visible and near infrared range, from 400 to 1500 nm. As in all cases, the incident and second-harmonic waves are p-polarized.

The linear scattering cross section is shown in Fig. 4, together with the first three multipolar contributions which, due to the symmetry of the system are from the symmetric base ψm+(r,θ|2ω). The magnetic dipole contribution, denoted by the green dashed curve that peaks around λ = 1403 nm, is quite strong. The electric dipole contribution, shown with the dot-dashed red curve is also quite strong and peaks around λ = 977 nm. The quadrupole contribution is small in the range considered.

 figure: Fig. 4

Fig. 4 Scattering efficiency at the fundamental frequency ω for an infinite cylinder (2D particle with a circular cross section) of radius a = 150 nm. The first three multipolar contributions are also shown. Note that due to the symmetry of the system, only multipoles of the base ψm+ are excited in the linear case.

Download Full Size | PDF

Results for the second harmonic efficiency are shown in Fig. 5(a), together with the first three contributions from the multipole expansion. To help in the interpretation of the results, in Fig. 5(b), we present the results of the linear scattering efficiency in the same spectral region, together with the curves for the first 4 multipolar contributions. This figure is helpful to identify the kind of multipoles supported by the particle in this wavelength range and the position of their resonances. It is clear that the first three components of the even base are excited quite strongly and, thus, that the system supports well the magnetic dipole and the electric dipole and quadrupole modes. The a3+ term is weakly excited, but has a resonance around λ = 600 nm that somehow drives a relatively strong response in the second harmonic results (term proportional to b3).

 figure: Fig. 5

Fig. 5 Nonlinear and linear scattering efficiency of the cylinder considered in Fig. 4. (a) Generated second harmonic and the first three multipolar contributions when the fundamental wavelength λ spans the range shown in Fig. 4. Note that due to the symmetry of the system only multipoles of the base ψm are excited. (b) Linear scattering efficiency in the same spectral region as in (a). In this case, only multipoles of the base ψm+ are excited. In both cases, λ denotes the wavelength of incidence.

Download Full Size | PDF

The second harmonic efficiency has, as expected, no contributions from the magnetic dipole term b0 and the excited multipoles belong to the antisymmetric base ψm(r,θ|2ω). Although the system has a strong magnetic dipole resonance around λ = 1403 nm, the second harmonic sources are not able to excite it. The particle has also a strong electric dipole-type resonance around λ = 977 nm. However, this resonance (term proportional to b1) is not excited strongly in the second harmonic. Clearly, the strongest contribution to the second harmonic efficiency is that corresponding to the electric quadrupole (coefficient b2, denoted by the short-dashed blue curve). The scattering patterns at the wavelengths of resonance of the three contributions (shown as insets) confirm the predominance of the quadrupole term of the antisymmetric base (see Fig. 2). Also, since the contributions to the field are of the form bm sin , no radiation is produced in the exact forward and backscattering directions (θ = 0°and 180°).

A near field map corresponding to |ψsc(I)(r,θ|2ω)| at the wavelength λ/2 = 977 nm, which coincides with the peak of the electric dipole resonance, is presented in Fig. 6. It is interesting to note that even at that wavelength, the quadrupole term seems to dominate the near field pattern.

 figure: Fig. 6

Fig. 6 Modulus of the magnetic field in the near field of the cylinder considered in Fig. 4 at λ/2 = 977 nm, which coincides with the maximum of the electric dipole resonance in the second harmonic field. Note that even at the peak of the dipolar resonance, the quadrupole contribution is predominant.

Download Full Size | PDF

3.1. The effect of the fundamental multipoles in SHG

The effect of exciting multipoles at the fundamental on the SHG can be studied from the analytical expression of the scattering efficiency of the cylinder at 2ω [28]. The strength of the multipoles excited at the second harmonic, represented by the coefficients bm, depend on the strength of the multipolar contributions to the fundamental frequency, represented by am+. The effect produced in the second harmonic by each elementary excitation at the fundamental frequency can be visualized by artificially turning on and off particular multipolar contributions.

To illustrate the procedure and study these effects we continue with our example of a cylindrical silicon particle of radius a = 150 nm. In Fig. 7, we show the second harmonic efficiency, together with curves obtained when only one or two of the multipoles excited at the fundamental frequency are considered. The curve with the green dashed line corresponds to the case in which only the dipolar term (a1+) is kept at the fundamental, to drive the nonlinear processes that give rise to the second harmonic field. One can see that the strong quadrupolar peak around 726 nm is already present at the second harmonic with this fundamental dipolar excitation. On the other hand, if we keep all the excitations, except the dipolar one (grey curve in the figure), practically all the second harmonic response is lost. Interestingly, the nonlinear interaction between the dipolar and the quadrupolar excitations produces the second harmonic peaks that correspond to the b1 and b3 contributions.

 figure: Fig. 7

Fig. 7 Scattering efficiency in the SHG for the cylinder considered in Fig. 4, considering the effect of the fundamental multipolar excitations. The full curve includes all the multipolar contributions at the fundamental. The green dashed curve considers only the dipolar electric contribution, the gray continuous line considers all the contributions but the dipolar electric one, and the short-dashed blue curve considers the dipolar and quadrupolar electric contributions.

Download Full Size | PDF

From the results presented in these figures we conclude that, for cylindrical particles, the magnetic dipole excitation at ω plays no role in the generation of second harmonic light, and that the magnetic dipole resonance supported by the particle at 2ω is not excited by the nonlinear processes involved. Since the MD mode is associated with a circulation of the electric field that is tangential to the surface of the cylinder, this observation is a consequence of the fact that, from the assumed model of the nonlinear polarization, the surface susceptibility χztts=0.

4. Particles with non-circular cross sections

The results presented up to this point correspond to particles with a circular cross section, for which one can proceed by means of the analytical expressions for the coefficients of the multipolar expansions [28]. For particles with more general cross sections, the scattered fields are calculated numerically [27] and, from these results, the coefficients of the multipolar expansion are estimated from Eqs. (15)(17). In this section, we present results for particles with non-circular cross sections and discuss the consequences of the lack of symmetry of the particle about the direction of incidence.

4.1. Particles with a rectangular cross section

Let us consider now a particle with a smoothed-square cross section whose side D = 300 nm. Results for the second harmonic efficiency are shown in Fig. 8. As expected, since the particle is symmetric about the x1x2 plane, only the components of the antisymmetric base are excited at the second harmonic frequency. In comparison with the case of the circular cylinder, we observe that with the square cross section particle the electric dipole excitation becomes relatively stronger. From the figure we observe that the scattering efficiency around the wavelength of the electric dipole resonance (∼ 1052 nm) is larger than in the other case, and that the scattering pattern seems to have a clear dipolar shape. On the other hand, in the position of the quadrupolar resonance, the other contributions are negligible and, in consequence, the scattering distribution is clearly quadrupolar in character. A map of |ψsc(r, θ)| at λ/2 = 1052 nm, which coincides with the maximum of the dipolar excitation is shown in Fig. 9. As expected, the near field pattern is dominated by the dipolar contribution.

 figure: Fig. 8

Fig. 8 Second harmonic scattering efficiency for a dielectric particle with a 300 nm per side square cross section. Also shown are the first three multipolar contributions to the efficiency. The insets illustrate the form of the far field scattering distributions at the wavelengths of the three multipolar resonances.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 Modulus of the magnetic field in the near field of the particle with square cross section considered in Fig. 8 at λ/2 = 1052 nm, which coincides with the maximum of the electric dipole resonance at the second harmonic wavelength.

Download Full Size | PDF

4.2. Non-symmetric particles

When the particle is not symmetric about the x1x2 plane it is possible to excite the two multipolar bases. We now consider particle whose cross section has the shape of an equilateral triangle, illuminated under different orientations. The sides of the triangle have a length of 285 nm and the illumination has a wavelength of λ = 1251 nm, which was chosen because the second harmonic wavelength (λ/2 = 626 nm) coincides with the quadrupolar resonance of the particle.

To illustrate the situation, in Fig. 10 we present the near field intensity maps corresponding to the four orientations considered. The angle α denotes the angle of rotation from the original orientation, shown in (a), the upper left image of the figure. When α = 0, the figure is symmetric about the x1x2 plane and it is expected that only the antisymmetric base will be excited. Since the field is antisymmetric, its squared modulus has to be symmetric about the x1x2 plane, and this property can be clearly observed in the figure. Further verification of the antisymmetric property of the field is provided by the calculated values of the expansion coefficients, which are shown in Table 1. We see that for α = 0°, the coefficients of the symmetric base are identically zero.

 figure: Fig. 10

Fig. 10 Modulus of the magnetic field in the near field of the particle with triangular cross section with different orientations at λ/2 = 626 nm, which coincides with the maximum of the electric quadrupole resonance at the second harmonic wavelength. The symmetric orientation of the particle is shown in (a), while (b), (c) and (d) correspond to rotations by α = 10°, 20° and 30°.

Download Full Size | PDF

Tables Icon

Table 1. Values of the multipolar expansion coefficients of the second harmonic field for different orientations of the particle of triangular cross section (see Fig. 10) at λ/2 = 626 nm. For simplicity, the common factor 1/(nBe) has been omitted from the values shown.

Continuing with the discussion of Fig. 10(a), we observe that the generated near field pattern appear to be fairly close to what one would expect for a field dominated by a quadrupolar resonance, which coincides with the fact that the coefficient b2 shown in the table for that case is the largest. Since the second harmonic wavelength coincides with the position of the quadrupolar resonance, this is not surprising. As the triangular particle is rotated, the near field pattern changes and we observe on the table that the contributions from the symmetric base become non-zero and grow with the rotation angle. Similarly the field along the x3 = 0 line is no longer zero.

5. Summary and conclusions

We have studied the scattering and SHG of light by dielectric particles made of centrosymmetric a high index material, focusing on the case of silicon. The solution of the scattering problem at both, ω and 2ω is based on the numerical solution of coupled integral equations for the surface sources and the nonlinear response of the surface is modeled with the dipollium model. The scattered fields at ω and 2ω are written in terms of a multipolar basis that permits the visualization of the kind of excitations that take place at the particle.

Using simple symmetry arguments, we show that for the case of particles that are symmetric about an axis in the direction of incidence, the magnetic dipole contribution cannot be excited at the second harmonic frequency. This result is illustrated by calculations with particles with circular, square and triangular cross sections. Similarly, for the case of infinite cylinders (circular cross section), we also find that the magnetic dipole excitation at the fundamental frequency plays no role in the second harmonic generated by the particle, which is mainly driven by the electric dipole excited at the fundamental frequency. At the harmonic frequency, the response is normally dominated by the quadrupolar excitation.

Funding

Consejo Nacional de Ciencia y TecnologÍa (CONACYT), under grants CB-2016-285419 and FC-2016-2221 and scholarship for M. A. G. Mandujano.

Acknowledgments

M. A. G. Mandujano is grateful for the financial support of CONACYT.

References

1. C. F. Bohren and D. R. Huffman, Absorption and Scattering of Light by Small Particles(Wiley, 1998). [CrossRef]  

2. A. V. Zayats, I. I. Smolyaninov, and A. A. Maradudin, “Nano-optics of surface plasmon polaritons,” Phys. Rep. 408, 131–314 (2005). [CrossRef]  

3. S. A. Maier, Plasmonics: Fundamentals and Applications(Springer, 2007).

4. W. Cai and V. Shalaev, Optical Metamaterials: Fundamentals and Applications(Springer, 2010). [CrossRef]  

5. J. B. Pendry, A. J. Holden, D. J. Robbins, and W. J. Stewart, “Magnetism from conductors and enhanced nonlinear phenomena,” IEEE Trans. Microwave Theory Tech. 47, 2075–2084 (1999). [CrossRef]  

6. Q. Zhao, J. Zhou, F. Zhang, and D. Lippens, “Mie resonance-based dielectric metamaterials,” Mat. Today 12, 60–69 (2009). [CrossRef]  

7. J. A. Schuller and M. L. Brongersma, “General properties of dielectric optical antennas,” Opt. Express 17, 24084–24095 (2009). [CrossRef]  

8. A. B. Evlyukhin, C. Reinhardt, A. Seidel, B. S. Luk’yanchuk, and B. N. Chichkov, “Optical response features of Si-nanoparticle arrays,” Phys. Rev. B 82, 045404 (2010). [CrossRef]  

9. A. García-Etxarri, R. Gómez-Medina, L. S. Froufe-Pérez, C. López, L. Chantada, F. Scheffold, J. Aizpurua, M. Nieto-Vesperinas, and J. J. Sáenz, “Strong magnetic response of submicron Silicon particles in the infrared,” Opt. Express 19, 4815–4826 (2011). [CrossRef]   [PubMed]  

10. A. I. Kuznetsov, A. E. Miroshnichenko, Y. H. Fu, J. B. Zhang, and B. Luk’yanchuk, “Magnetic light,” Sci. Rep. 2, 492 (2012). [CrossRef]  

11. I. Staude, A. E. Miroshnichenko, M. Decker, N. T. Fofang, S. Liu, E. Gonzales, J. Dominguez, T. S. Luk, D. N. Neshev, I. Brener, and Y. Kivshar, “Tailoring directional scattering through magnetic and electric resonances in subwavelength silicon nanodisks,” ACS Nano 7, 7824–7832 (2013). [CrossRef]  

12. A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma, Y. S. Kivshar, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science 354, aag2472 (2016). [CrossRef]  

13. G. Mie, “Beiträge zur optik truüber medien, speziell kolloidaler metallösungen,” Ann. Phys. (Leipzig) 25, 377–445 (1908). [CrossRef]  

14. J. I. Dadap, J. Shan, K. B. Eisenthal, and T. F. Heinz, “Second-harmonic Rayleigh scattering from a sphere of centrosymmetric material,” Phys. Rev. Lett. 83, 4045–4048 (1999). [CrossRef]  

15. V. L. Brudny, B. S. Mendoza, and W. L. Mochan, “Second-harmonic generation from spherical particles,” Phys. Rev. B 62, 11152 (2000). [CrossRef]  

16. J. I. Dadap, J. Shan, and T. F. Heinz, “Theory of optical second-harmonic generation from a sphere of centrosymmetric material: small-particle limit,” J. Opt. Soc. Am. B 21, 1328–1347 (2004). [CrossRef]  

17. Y. Pavlyukh and W. Hübner, “Nonlinear Mie scattering from spherical particles,” Phys. Rev. B 70, 245434 (2004). [CrossRef]  

18. M. W. Klein, C. Enkrich, M. Wegener, and S. Linden, “Second-harmonic generation from magnetic metamaterials,” Science 313, 502–504 (2006) [CrossRef]   [PubMed]  

19. N. Feth, S. Linden, M. W. Klein, M. Decker, F. B. P. Niesler, Y. Zeng, W. Hoyer, J. Liu, S. W. Koch, J. V. Moloney, and M. Wegener, “Second-harmonic generation from complementary split-ring resonators,” Opt. Lett. 33, 1975–1977 (2008). [CrossRef]  

20. C. Ciracì, E. Poutrina, M. Scalora, and D. R. Smith, “Origin of second-harmonic generation enhancement in optical split-ring resonators,” Phys. Rev. B 85, 201403 (2012). [CrossRef]  

21. L. Carletti, A. Locatelli, O. Stepanenko, G. Leo, and C. De Angelis, “Enhanced second-harmonic generation from magnetic resonance in AlGaAs nanoantennas,” Opt. Express 23, 26544–26550 (2015). [CrossRef]  

22. V. F. Gili, L. Carletti, A. Locatelli, D. Rocco, M. Finazzi, L. Ghirardini, I. Favero, C. Gomez, A. Lemaître, M. Celebrano, C. De Angelis, and G. Leo, “Monolithic AlGaAs second-harmonic nanoantennas,” Opt. Express 2415965–15971 (2016). [CrossRef]   [PubMed]  

23. R. Camacho-Morales, M. Rahmani, S. Kruk, L. Wang, L. Xu, D. A. Smirnova, A. Solntsev, A. E. Miroshnichenko, H. H. Tan, F. Karouta, S. Naureen, K. Vora, L. Carletti, C. de Angelis, C. Jagadish, Y. S. Kivshar, and D. N. Neshev, “Nonlinear generation of vector beams from AlGaAs nanoantennas,” Nano Lett. 16, 7191–7197 (2016). [CrossRef]  

24. S. S. Kruk, R. Camacho-Morales, L. Xu, M. Rahmani, D. A. Smirnova, L. Wang, H. H. Tan, C. Jagadish, D. N. Neshev, and Y. S. Kivshar, “Nonlinear optical magnetism revealed by second-harmonic generation in nanoantennas,” Nano Lett. 17, 3914–3918 (2017). [CrossRef]  

25. S. Liu, M. B. Sinclair, S. Saravi, G. A. Keeler, Y. Yang, J. Reno, G. M. Peake, F. Setzpfandt, I. Staude, T. Pertsch, and I. Brener, “Resonantly enhanced second-harmonic generation using III-V semiconductor all-dielectric metasurfaces,” Nano Lett. 16, 5426–5432 (2016). [CrossRef]  

26. D. Smirnova, A. I. Smirnov, and Y. S. Kivshar, “Multipolar second-harmonic generation by Mie-resonant dielectric nanoparticles,” Phys. Rev. A 97, 013807 (2018). [CrossRef]  

27. C. I. Valencia, E. R. Méndez, and B. S. Mendoza, “Second-harmonic generation in the scattering of light by two-dimensional particles,” J. Opt. Soc. Am. B 20, 2150–2161 (2003). [CrossRef]  

28. C. I. Valencia, E. R. Méndez, and B. S. Mendoza, “Second-harmonic generation in the scattering of light by an infinite cylinder,” J. Opt. Soc. Am. B 21, 36–44 (2004). [CrossRef]  

29. B. S. Mendoza and W. L. Mochán, “Exactly solvable model of surface second-harmonic generation,” Phys. Rev. B 53, 4999–5006 (1996). [CrossRef]  

30. J. Gielis, “A generic geometric transformation that unifies a wide range of natural and abstract shapes,” Am. J. Bot. 90, 333–338 (2003). [CrossRef]   [PubMed]  

31. D. Macías, P.-M. Adam, V. Ruiz-Cortés, R. Rodríguez-Oliveros, and J. A. Sánchez-Gil, “Heuristic optimization for the design of plasmonic nanowires with specific resonant and scattering properties,” Opt. Express 20, 13146–13163 (2012). [CrossRef]  

32. N. Bloembergen, R. K. Chang, S. S. Jha, and C. H. Lee, “Optical second-harmonic generation in reflection from media with inversion symmetry,” Phys. Rev. B 174, 813–822 (1968). [CrossRef]  

33. M. E. Inchaussandague, M. L. Gigli, K. A. O’Donnell, E. R. Méndez, R. Torre, and C. I. Valencia, “Second-harmonic generation from plasmon polariton excitation on silver diffraction gratings: comparisons of theory and experiment,” J. Opt. Soc. Am. B 34, 27–37 (2017). [CrossRef]  

34. D. Timbrell, J. W. You, Y. S. Kivshar, and N. C. Panoiu, “A comparative analysis of surface and bulk contributions to second-harmonic generation in centrosymmetric nanoparticles,” Sci. Rep. 8, 3586 (2018). [CrossRef]   [PubMed]  

35. W. K. H. Panofsky and M. Phillips, Sec. 13–7 in Classical Electricity and Magnetism, 2nd Ed. (Dover, 2005).

36. J. Petschulat, J. Yang, C. Menzel, C. Rockstuhl, A. Chipouline, P. Lalanne, A. Tünnermann, F. Lederer, and T. Pertsch, “Understanding the electric and magnetic response of isolated metaatoms by means of a multipolar field decomposition,” Opt. Express 18, 14454–14466 (2010). [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 (10)

Fig. 1
Fig. 1 Schematic diagram of the geometry considered.
Fig. 2
Fig. 2 Illustration of the lowest order functions of the symmetric and antisymmetric bases. In the symmetric base the lowest order terms represent the magnetic dipole (MD), the electric dipole (ED) and the electric quadrupole (EQ) contributions. In the antisymmetric base they are the electric dipole (ED) and the electric quadrupole (EQ) contributions.
Fig. 3
Fig. 3 Illustration of situations in which the second harmonic scattered field is assumed to be symmetric (a) and antisymmetric (b).
Fig. 4
Fig. 4 Scattering efficiency at the fundamental frequency ω for an infinite cylinder (2D particle with a circular cross section) of radius a = 150 nm. The first three multipolar contributions are also shown. Note that due to the symmetry of the system, only multipoles of the base ψ m + are excited in the linear case.
Fig. 5
Fig. 5 Nonlinear and linear scattering efficiency of the cylinder considered in Fig. 4. (a) Generated second harmonic and the first three multipolar contributions when the fundamental wavelength λ spans the range shown in Fig. 4. Note that due to the symmetry of the system only multipoles of the base ψ m are excited. (b) Linear scattering efficiency in the same spectral region as in (a). In this case, only multipoles of the base ψ m + are excited. In both cases, λ denotes the wavelength of incidence.
Fig. 6
Fig. 6 Modulus of the magnetic field in the near field of the cylinder considered in Fig. 4 at λ/2 = 977 nm, which coincides with the maximum of the electric dipole resonance in the second harmonic field. Note that even at the peak of the dipolar resonance, the quadrupole contribution is predominant.
Fig. 7
Fig. 7 Scattering efficiency in the SHG for the cylinder considered in Fig. 4, considering the effect of the fundamental multipolar excitations. The full curve includes all the multipolar contributions at the fundamental. The green dashed curve considers only the dipolar electric contribution, the gray continuous line considers all the contributions but the dipolar electric one, and the short-dashed blue curve considers the dipolar and quadrupolar electric contributions.
Fig. 8
Fig. 8 Second harmonic scattering efficiency for a dielectric particle with a 300 nm per side square cross section. Also shown are the first three multipolar contributions to the efficiency. The insets illustrate the form of the far field scattering distributions at the wavelengths of the three multipolar resonances.
Fig. 9
Fig. 9 Modulus of the magnetic field in the near field of the particle with square cross section considered in Fig. 8 at λ/2 = 1052 nm, which coincides with the maximum of the electric dipole resonance at the second harmonic wavelength.
Fig. 10
Fig. 10 Modulus of the magnetic field in the near field of the particle with triangular cross section with different orientations at λ/2 = 626 nm, which coincides with the maximum of the electric quadrupole resonance at the second harmonic wavelength. The symmetric orientation of the particle is shown in (a), while (b), (c) and (d) correspond to rotations by α = 10°, 20° and 30°.

Tables (1)

Tables Icon

Table 1 Values of the multipolar expansion coefficients of the second harmonic field for different orientations of the particle of triangular cross section (see Fig. 10) at λ/2 = 626 nm. For simplicity, the common factor 1/(nBe) has been omitted from the values shown.

Equations (35)

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

r s ( t ) = [ ξ ( t ) , η ( t ) ] .
P N L ( r | 2 ω ) = α [ E ( r | ω ) ] E ( r | ω ) + β E ( r | ω ) [ E ( r | ω ) ] + γ [ E ( r | ω ) E ( r | ω ) ] ,
ψ p ( I ) ( t | 2 ω ) ψ p ( I I ) ( t | 2 ω ) = A p ( t | ω ) ,
1 ϵ I ( 2 ω ) ϒ p ( I ) ( t | 2 ω ) 1 ϵ I I ( 2 ω ) ϒ p ( I I ) ( t | 2 ω ) = B p ( t | ω ) ,
A p ( t | ω ) = 8 π i c ω χ   ϵ I ( ω ) 1 ϕ 2 ( t ) ϒ p ( I ) ( t | ω ) d ψ p ( I ) ( t | ω ) d t ,
B p ( t | ω ) = 8 π i c ω [ χ d d t ( 1 ϕ ( t ) d ψ p ( I ) ( t | ω ) d t ) 2 + χ ϵ I 2 ( ω ) d d t ( ϒ p ( I ) ( t | ω ) ϕ ( t ) ) 2 + χ b d [ ψ p ( I ) ( t | ω ) ] 2 d t η d [ ψ s ( I ) ( t | ω ) ] 2 d t ] ,
χ   = χ t t z s ,
χ = χ z z z s + α / 2 + γ ϵ I I ( 2 ω ) ϵ I I 2 ( ω ) ,
χ = χ z t t s + α / 2 + γ ϵ I I ( 2 ω ) ,
η = χ z t t s + γ ϵ I I ( 2 ω ) ,
χ b = α 2 ϵ I I ( ω ) ϵ I I ( 2 ω ) ( ω c ) 2 ,
χ ( ω ) = e 2 / m ω 0 2 ω 2 i Γ ω ,
α ( ω ) = ϵ ( ω ) 1 32 π 2 n B e [ 4 ϵ ( 2 ω ) ϵ ( ω ) 3 ] ,
β ( ω ) = ϵ ( ω ) 1 32 π 2 n B e ,
γ ( ω ) = ϵ ( ω ) 1 32 π 2 n B e [ ϵ ( 2 ω ) 1 ] ,
χ z z z s ( ω ) = ( ϵ ( ω ) 1 ) 2 32 π 2 n B e ϵ ( ω ) 2 { 2 ϵ ( ω ) ϵ ( 2 ω ) ϵ ( ω ) ϵ ( 2 ω ) [ ϵ ( 2 ω ) ϵ ( ω ) ] + [ 1 ϵ ( 2 ω ) ] [ ϵ ( 2 ω ) ϵ ( ω ) ] 2 ln ( ϵ ( ω ) ϵ ( 2 ω ) ) } ,
χ t t z s ( ω ) = χ t z t s ( ω ) = 1 32 π 2 n B e [ ϵ ( ω ) 1 ] 2 ϵ ( ω ) ,
χ z t t s ( ω ) = 0 .
ψ s c ( I ) ( r , θ | Ω ) = e i π 4 e i k Ω r [ 8 π k Ω r ] 1 / 2 S ( θ | Ω ) ,
S ( θ | Ω ) = S [ i k Ω [ η ( t ) cos θ ξ ( t ) sin θ ] ψ p ( I ) ( t | Ω ) ϒ p ( I ) ( t | Ω ) ] × exp { i k Ω [ ξ ( t ) cos θ + η ( t ) sin θ ] } d t .
ψ s c ( I ) ( r , θ | Ω ) = m = c m i m H m ( 1 ) ( k Ω r ) e i m θ ,
ψ s c ( I ) ( r , θ | Ω ) = m = 0 c m + ψ m + ( r , θ | Ω ) + m = 1 c m ψ m ( r , θ | Ω ) ,
ψ m ± ( r , θ , | Ω ) = i m H m ( 1 ) ( k Ω r ) { cos m θ i sin m θ } ,
c m ± = ( c m ± c m ) ,
ψ 0 + ( r , θ , | Ω ) = H 0 ( 1 ) ( k Ω r ) ,
c 0 + = c 0 .
c m ± = 0 2 π ψ s c ( I ) ( r 0 , θ | Ω ) ψ m ± ( r 0 , θ | Ω ) d θ 0 2 π | ψ m ± ( r 0 , θ | Ω ) | 2 d θ ,
c 2 ± = 1 4 π 0 2 π S ( θ | Ω ) { cos m θ i sin m θ } d θ ,
c 0 = 1 8 π 0 2 π S ( θ | Ω ) d θ .
Q s c ( ω ) = P s c ( ω ) P i n c ( ω ) = 0 2 π q s c ( θ | ω ) d θ ,
q s c ( θ | ω ) = 1 8 π D ( ω / c ) | S ( θ | ω ) | 2 | ψ 0 | 2 ,
Q s c ( 2 ω ) = σ g P s c ( 2 ω ) [ P i n c ( ω ) ] 2 = 0 2 π q s c ( θ | 2 π ) d θ ,
q s c ( θ | 2 ω ) = 1 2 ω D | S ( θ | 2 ω ) | 2 | ψ 0 | 4 .
Q s c ( Ω ) = [ | c 0 | 2 + 1 2 m = 1 [ | c m + | 2 + | c m | 2 ] ] × { 2 k D , Ω = ω , 8 π ω D | ψ 0 | 4 , Ω = 2 ω ,
q s c ( θ | Ω ) = [ c 0 + m = 1 [ c m + cos m θ + i c m sin m θ ] ] 2 × { 1 π k D , Ω = ω , 16 ω D | ψ 0 | 4 , Ω = 2 ω .
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.