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

Theory and design of two-dimensional high-contrast-grating phased arrays

Open Access Open Access

Abstract

Optical properties of two-dimensional (2D) high-contrast gratings are investigated. We analyze the mechanisms for high-contrast gratings to function as various high-performance optical components. Our top-down design procedure allows us to efficiently obtain initial structural parameters and engineer them for a wide range of applications, such as reflectors, filters, resonators, waveplates, and even 2D phase plates. Simulation results of our designed structures show ultra-high power efficiency, and excellent agreement with our predicted functionalities.

© 2015 Optical Society of America

1. Introduction

Diffractive gratings have been crucial components in the broad area of optics for centuries [1]. In the recent decade, researchers have shown extensive interest in an operation regime where the grating period Λ is comparable or slightly smaller than the wavelength λ [2–6], where strong field interaction among the near-wavelength structures can provide extraordinary optical behaviors. One interesting type of grating, known as the high-contrast grating (HCGs), is made of periodic high-index materials surrounded by materials with much smaller indices [6]. HCGs have been used as ultrabroadband reflectors [7] and high-Q resonators [8]. They can be integrated as compact, high-performance tunable mirrors in optoelectronic devices, such as vertical-cavity surface-emitting lasers (VCSELs) [3, 9].

HCGs are referred as one-dimensional (1D) if the periodicity is only in one direction (e.g. periodic bars, stripes, or grooves). In this case the fields can be decoupled into the transverse-electric (TE) and transverse-magnetic (TM) polarizations. An incident wave with one polarization will not excite an HCG mode with an orthogonal polarization. For 2D HCGs, structures are periodic in two directions and the modes supported by the grating have hybrid polarizations, that is, they are not separable into TE and TM. The modes in 2D HCGs are also more closely-packed spectrally. In other words, with a given incident wave, many more modes can be excited in a 2D HCG than in a 1D HCG. The diffracted modes, also known as the Floquet modes, have one extra degree of freedom for the propagation direction, which makes it very difficult to predict the optical properties of a 2D HCG. Moreover, there are many more structural parameters in 2D HCGs, which largely complicate the design and optimization procedure. Up to now, majority of the gratings in use are still one-dimensional.

However, there are applications which require beam-steering or focusing in both directions, or generating optical vortices [10, 11]. In these cases, 2D HCGs are more advantageous. Extensive experimental work has been done in demonstrating the functionalities of 2D gratings [2, 10]. In this work, we examine the physics of 2D HCGs and the mechanisms which allow high-performance operation. The dual-mode analysis we propose can largely simplify the searching of the initial design and the subsequent optimization. A top-down design procedure is provided. To model 2D HCGs, the analytical mode-matching method [6,12], which was previously used in 1D HCGs, is rather inefficient and requires complex root-searching for coupled nonlinear equations. An in-house developed rigorous coupled-wave analysis (RCWA) [13, 14] program is shown to be very convenient for design purposes, and both the convergence and accuracy are verified.

One important application for 2D HCGs is the 2D phase plate, which normally requires non-periodic designs. However, the high-index material can largely confine the field around the local position, where effective-medium approximation is applicable [11, 15, 16]. Based on our optimized design for periodic 2D HCGs, we can apply the grating parameters as the local parameters for the non-periodic grating, according to the design requirement at the local position. Our designed non-periodic phase plate is then verified through the finite-difference time-domain (FDTD) simulation. The results show excellent agreement with our designs.

There is much recent progress on beam shaping using plasmonic metastructures [2]. The localized resonance behavior resulting from the plasmonic effect also enables a wide range of beam engineering. However, the metallic structures supporting the plasmonic resonance also introduce much optical loss. HCGs, on the other hand, have intrinsically much lower loss. Using our designs we demonstrate excellent power efficiency (above 90%).

2. Optical properties of 2D high-contrast gratings

We first study the underlying physics of 2D high-contrast gratings. Figure 1(a) shows a general 2D grating on a substrate (optional for suspended grating membranes). The incident plane wave is characterized by the polar and azimuthal angles of its wave vector ki = (kix,kiy,kiz) = (k0 sinθ cosϕ, k0 sinθ sinϕ, k0 cosθ). The polarization of the incident electric field should also be characterized by two angles, that is, Ei = (E0 sinα1 cosα2, E0 sinα1 sinα2, E0 cosα1). Yet the electric field is transverse to the wave vector for propagating waves, thus one angle α is sufficient to describe the polarization. For oblique incidence, it is convenient to decompose the fields into s- and p-polarizations, which will be discussed in later sections.

 figure: Fig. 1

Fig. 1 (a) Schematic of a 2D high-contrast grating on a rectangular lattice with a substrate. The plane-wave incident polar and azimuthal angles are θ and ϕ, respectively. Parameters: Λx and Λy (periods in x^ and ŷ), Dx and Dy (grating widths in x^ and ŷ), ng and ns (grating and substrate indices), and tg and ts (grating and substrate thicknesses). (b) Schematic of a 2D high-contrast grating on a hexagonal lattice with a substrate. Parameters: period Λ, rod diameter d, grating and substrate indices ng and ns, grating and substrate thicknesses tg and ts.

Download Full Size | PDF

Design parameters include grating periods and widths in both x^ and ŷ directions (Λx, Λy, Dx, and Dy), as well as thicknesses and refractive indices of both the high-index material and the substrate (ng, ns, tg, and ts), as shown in Fig. 1(a). We define duty cycles for the high-index material as ηx = Dxx and ηy = Dyy.

For applications where optical fields express rotational symmetry, 2D circular gratings on a hexagonal lattice, as shown in Fig. 1(b), are more advantageous. Design parameters are simplified to grating period (Λ), rod diameter (d), thicknesses and refractive indices of both the grating and substrate (tg, ts, ng and ns). The duty cycle is defined as η = d/Λ. Gratings on a hexagonal lattice, being more closely packed, require Λ<2λ0/3 in order to have only the main harmonic being propagating, as compared to Λx,y < λ0 for gratings on a rectangular lattice.

We first look at the incident and transmitted regions, which are typically homogeneous. According to the Floquet theorem, the scattered field consists of Floquet modes with different orders of tangential wave vectors, as a result of the phase matching between the homogeneous region and the 2D grating with the translational symmetry. In rectangular lattice, the (m,n)-th order Floquet mode is a plane wave with a wave vector being

k(m,n)=x^kx+y^ky+z^kz,kx=kix+Gx,m,ky=kiy+Gy,n,Gx,m=mπΛx,m=0,±1,±2,,±Nx,Gy,n=nπΛy,n=0,±1,±2,,±Ny,kz=±4π2nr2λ2kx2ky2
where nr is the index of the incident or transmitted region. Thus the total number of modes is N = (2Nx + 1)(2Ny + 1). Using the RCWA formulation, we can further expand the transverse fields of the i-th eigenmode in the grating layer in terms of the same Floquet basis,
Et(i)=eiKz(i)z(m,n(x^E˜x,(m,n)(i)+y^E˜y,(m,n)(i))eiGx,mx+iGy,ny)eikixx+ikiyy
where the tilde symbols indicate the spectral coefficients. Here, Kz(i) is the propagation constant for the i-th eigenmode. Substituting Eq. (2) into the vector wave equation ××(Et(i)+z^Ez(i))=ω2c2nr2(Et(i)+z^Ez(i)), and eliminating Ez(i), we have the eigen-equation in spectral domain [14],
M¯E˜t(i)=(Kz(i))2N¯E˜t(i)
where M¯ and N¯ are 2N × 2N matrices dependent on the frequency ω and refractive index nr(x,y),(Kz(i))2 is the eigenvalue, and Et(i) is a 2N-dimensional vector consisting of spectral coefficients,
E˜t(i)=(E˜x,(Nx,Ny)(i),,E˜x,(m,n)(i),,E˜x,(Nx,Ny)(i),E˜y,(Nx,Ny)(i),,E˜y,(m,n)(i),,E˜y,(Nx,Ny)(i))T

Then the total transverse field in the grating layer can be expressed in spectral domain as

E˜t=iAiE˜t(i)=¯tA
where ¯t is a 2N × 2N matrix with columns being eigenvectors Et(i), and A contains 2N-dimensional eigenmode expansion coefficients. From the eigen-equations in Eq. (3) one can obtain forward and backward propagation constants for each eigenmode ( ±Kz(i)).

Similar to Eq. (5) we can expand the transverse magnetic field in terms of the magnetic components H˜t(i) of the eigenmodes ( H˜t(i)obtained from E˜t(i)). We can then express z-dependent transverse electric and magnetic fields including forward and backward propagation,

[E˜t(z)H˜t(z)]=[¯t¯t¯t¯t][A+(z)A(z)]=[¯t¯t¯t¯t][eiK¯zz00eiK¯zz][A+A]
where ¯t=[H˜t(t),,H˜t(2N)], and A±=[A1±,,A2N±]. Here, e±iK¯zz are diagonal matrices with elements being exp(±iKz(i)z). Matching the boundary conditions at the interface between two regions (e.g. I for air and II for grating), we have
[AII+AII]=[¯tII¯tII¯tII¯tII]1[¯tI¯tI¯tI¯tI][AI+AI]=[T¯III]4N×4N[AI+AI]
where TIII is the transfer matrix at the I-II interface. For a given incident Ainc, we can solve for the reflected Aref and transmitted Atrans once the total transfer matrix is found. Using Eq. (5) we can find the coefficients E˜t and H˜t for each spectral order.

Figure 2(a) shows the (0,0)-th order reflection spectra for a 2D HCG under normal incidence. Results are calculated with the finite-element method (FEM) using COMSOL Multiphysics, finite-difference time-domain method (FDTD) using Lumerical Solutions, as well as an in-house developed 2D rigorous coupled-wave analysis (RCWA) program. Good agreement is shown among three methods and good convergence of our RCWA program upon total number of 2D spectral orders (N) is observed. Moreover, calculating each frequency point using RCWA with N = 625 is only 90 seconds on a personal computer with an Intel Core i7-3520M processor, whereas the computation cost is much heavier for FDTD and FEM due to the full discretization of the 3D solution domain. When λ < max{Λxy}, higher spectral orders become propagating, and extracting each spectral order accurately from the total field solution in FDTD or FEM becomes more challenging.

 figure: Fig. 2

Fig. 2 (a) Comparison among the (0,0)-th order reflectivity spectra of a 2D high-contrast grating (HCG) under normal incidence calculated using finite-element method (blue star), finite-difference time-domain (red dots), and rigorous coupled-wave analysis (RCWA) using N = 289 (magenta), N = 441 (green), and N = 625 (blue). HCG parameters are Λx = 1μm, Λy = 0.5μm, Dx = 0.6μm, Dy = 0.3μm, and tg = 0.5μm. (b) Spectra of normalized scattered power in the z^-direction for (0,0), (+1,0), and (1,0) spectral orders (blue, red, and black, respectively), and the sum of the three spectra (green).

Download Full Size | PDF

We also verify that our RCWA simulation is energy-conserving. We calculate the scattered power flux (reflection and transmission) normalized by the incident power flux for the (0,0), (+1,0), (1,0) modes, as shown in Fig. 2(b). The power scattered into the (0,0) mode drops from unity when λ < 1μm, indicating power transfering into higher diffraction orders. Nonetheless, the total scattered power remains unity, satisfying the conservation of energy.

We can also study the 2D HCG properties under oblique incidence. The field polarization is defined relative to the incidence plane, which is formed by the surface normal and the incidence wave vector. We use s- and p-polarization for electric field perpendicular and parallel to the incidence plane, respectively. Figure 3(a) shows the (0,0)-th order reflectivity and transmissivity as functions of the incidence polar angle θ with wavelength λ = 2μm and incidence azimuthal angle ϕ = 0. In this case, all higher spectral orders remain evanescent for the whole 90-degree range of θ. Thus the sum of reflectivity and transmissivity remains unity. Results from our 2D RCWA program and FEM software agree very well.

 figure: Fig. 3

Fig. 3 (a) Incident polar angle dependent reflectivity (black), transmissivity (magenta), and their sum (green) for the (0,0) fundamental order under p-polarized incidence with λ = 2μm calculated using rigorous coupled-wave analysis, and comparison with finite-element method (red and blue crosses). The same grating structure is solved under (b) s-polarized and (c) p-polarized incidence with wavelength being λ = 0.6μm. Normalized scattered power fluxes in the z^-direction for the (0,0)-th, (+1,0)-th, (1,0)-th, (2,0)-th, and (3,0)-th spectral orders are shown as the blue, red, black, green, and cyan lines, respectively. The magenta line indicates the total normalized scattered power flux in the z^-direction. Arrows indicate the cutoff angles for spectral orders to appear or disappear.

Download Full Size | PDF

As the wavelength decreases, the higher spectral orders may be propagating. The critical condition for the (m,n)-th spectral order to transition between propagating and evanescent is

(k0sinθcosϕ+mπΛx)2+(k0sinθsinϕ+nπΛy)2=k02
where 0 < θ < 90°. If we consider the incidence plane being the xz-plane (i.e. ϕ = 0), wavelength being fixed, and λ > Λy, then the cutoff angles from the above critical condition become
θc=sin1(1|m|λΛx)for(+|m|,0)to disappearθc=sin1(|m|λΛx1)for(|m|,0)to appear

Figures 3(b) and 3(c) show the normalized scattered power in the z^-direction for different spectral orders under s- and p-polarized oblique plane wave incidence, respectively. The cutoff angles for the (2,0), (+1,0), and (3,0) modes are calculated from Eq. (9) to be 11.54°, 23.58°, and 53.13°, which agree with the results indicated by the arrows in Figs. 3(b) and 3(c).

We also observe resonance behavior around the cutoff angles for p-polarized but not the s-polarized incidence. Such resonance is not a numerical artifact since it remains as the numerical accuracy is varied (N in RCWA). It is indeed the Wood’s anomaly which has been studied extensively since 1902 [17–22]. Hessel and Oliner [21] classified the anomaly into two types: a Rayleigh type where a diffraction order appears or disappears, and a resonance type which is related to the coupling between the diffracted wave and the leaky wave inside the grating. The two types may occur very closely or separately. The second type is dependent on the incident field polarization, which requires the surface reactance of the grating to be capacitive or inductive for resonance to occur. The surface reactance is affected by grating thickness, which explains why for a range of thicknesses, the anomaly only happens to one polarization. Previous investigation on Wood’s anomaly in 1D gratings is well applicable to 2D gratings, though rigorous theoretical analysis could be rather involved. In this case, our 2D RCWA program can be a convenient tool to study this effect. One thing to note is that, many studies [21, 22] on 1D grating defined s- and p-polarizations relative to the grooves of the grating, while in this work on 2D grating, polarizations are defined relative to the incidence plane. Thus the analysis on “s-anomalies” in [21] is applicable to the p-polarized case in Fig. 3(c).

Regardless of the anomalies, the energy conservation is still satisfied, as confirmed by our RCWA results. The total scattered power flux in the z^-direction remains a smooth function and has a cosine dependence on the incident polar angle when normalized by the total incident power flux, as shown by the magenta lines in Figs. 3(b) and 3(b).

Even using only the fundamental order of reflection or transmission with normal plane wave incidence, the 2D grating can be designed as various optical components, such as polarizers and waveplates. Figure 4(a) shows the magnitude of the reflected, transmitted, and total scattered wave under normal plane wave incidence in the subwavelength regime (λ > max {Λx, Λy}). The scattered wave is decomposed into two polarizations, one being parallel and the other being perpendicular to the incidence. As the incident polarization rotates relative to the grating about the z-axis, the magnitude of the parallel component varies relative to the perpendicular component. This provides the possibility of using 2D gratings as polarizers. Moreover, the phase difference between the two components is also dependent on the azimuthal angle of the incident electric field, as shown in Fig. 4(b). This enables us to design quarter-wave plates. In this example, the parallel and perpendicular reflected waves have a 90° phase difference when the incident polarization angle is 27.7°. If such angle is 52.7°, the two transmitted components have a 90° phase difference.

 figure: Fig. 4

Fig. 4 (a) Reflectivity (blue), transmissivity (red), and the normalized scattered power (green) of a normal incident plane wave with λ = 1.55μm as functions of the azimuthal polarization angle. The arrows indicate the components parallel and perpendicular to the incident wave. (b) Phase differences between the parallel and perpendicular components for reflection (red) and transmission (blue) as functions of the azimuthal polarization angle. The 90° phase differences correspond to polarization angles of 27.7° and 52.7° for reflection and transmission types, respectively.

Download Full Size | PDF

3. Design principles of 2D high-contrast gratings

We have verified our 2D RCWA program as a design tool and studied the optical properties of 2D HCGs, yet the design of 2D HCGs remains a challenge. The structural parameters include ng, tg, Λx, Λy, ηx, and ηy. The incidence condition includes wavelength (λ), propagation direction (ϕ, θ), and polarization (α). Here we present a design procedure which largely simplifies the searching of initial parameters and the optimization process. As shown in Fig. 5, we start from the eigenmodes in the 2D HCG, which is considered as an infinitely long waveguide along the z^-direction and periodic in the xy-plane. Based on the dispersion relations of the eigenmodes, we can obtain initial designs for the grating structure in the xy-plane. Then using the dual-mode analysis we can find the resonance condition, and thus the values of grating thickness to provide high and possibly perfect reflection or transmission. With the initial structural parameters, we can then optimize the design for various applications. For reflectors, filters and resonators, we are concerned with locations of the passband, stopband, and transition band. By checking the frequency spectra and adjusting the grating thickness, we can shift these bands to our desired frequency range. For polarizers and waveplates, we want to obtain desired amplitude selectivity and relatively phase shift between two orthogonal polarizations. Both can be achieved by rotating the incident polarization relative to the 2D HCG, as shown in Fig. 4. For 2D phase plate, we need to obtain a full 2π phase tuning range by varying the HCG transverse structure (e.g. Λ or η). At the same time, we need to maintain high reflection or transmission for designs to be practical. More details about the phase plate are in the next section.

 figure: Fig. 5

Fig. 5 Design procedures of 2D high-contrast gratings.

Download Full Size | PDF

3.1. Eigenmodes in 2D high-contrast gratings

Eigenmodes in 2D dielectric gratings are hybrid modes, which cannot be separated into transverse electric (TE) and transverse magnetic (TM), as in the case of 1D gratings. Furthermore, eigenmodes are more closely spaced in the frequency domain due to one extra degree of freedom (one more mode number), which largely complicates our analysis and prediction. However, the majority of the modes with strong confinement have dominant transverse field components, namely Ex/Hy-dominant (EH-like) and Ey/Hx-dominant (HE-like). These modes in periodic gratings resemble the EH and HE modes in the rectangle dielectric waveguides, using Marcatili’s approximation [23]. We can analyze the two separately under the approximation that grating modes with one dominant polarization only couple strongly with the free-space Floquet modes with the same polarization. Thus we can reduce the number of modes to analyze by almost half. Moreover, the field distribution of the incident plane wave possesses even symmetry in the xy-plane across the unit cell under normal incidence. Therefore eigenmodes with odd symmetry will not be excited and we can further narrow down our analysis. Figure 6(a) shows the dispersion curves of the eigenmodes with even symmetry and Hx-component dominant over Hy-component. In fact, the four modes shown in Fig. 6(a) are very much like the HE modes in rectangular dielectric waveguides when they are well-confined between the light lines (kz = ω/c and kz = ngω/c). Similarly, we can separate out Ex/Hy-dominant eigenmodes with even symmetry, as shown in Fig. 6(b). The Hy field profiles for EH00-like and EH20-like modes at ω=0.82πcΛx indeed are very close to those in dielectric waveguides, as shown in Fig. 6(c).

 figure: Fig. 6

Fig. 6 (a) HE-like even eigenmodes in a 2D rectangular grating on a rectangular lattice. The two dashed lines indicate the kz = ω/c and kz = ngω/c light lines. (b) EH-like even eigenmodes in a 2D rectangular grating on a rectangular lattice. (c) ℜe[Hy] for the EH00-like eigenmode and EH20-like eigenmode at ω=0.82πcΛx, as indicated by the green dots in (b).

Download Full Size | PDF

Now we are able to focus on the modes which have major contribution to the HCG behavior. For instance, between the cutoff frequencies ωc20 and ωc02 (for the EH20-like and EH02-like modes, respectively), we have a spectral region where only two key modes exist. This means within this region we can engineer the interference between the two modes to produce desired functionalities.

In other cases, 2D gratings on hexagonal lattice are preferred. Our RCWA program can be easily modified for this case. As shown in Fig. 7(a), the unit cell can be chosen as the yellow rectangle with a fixed aspect ratio of 3:1. We can still separate eigenmodes with Hx-dominant and Hy-dominant polarizations. The dispersion curves of the first few Hx- and Hy-dominant eigenmodes are shown in Figs. 7(b) and 7(c), respectively. We further identify the eigenmodes which are symmetric with the translation along 12x^+32y^ from those which are anti-symmetric, as indicated by the red and blue lines in the dispersion curves. The Hy components of the first symmetric, first anti-symmetric, and second symmetric Hy-dominant modes are shown in Figs. 7(d)–7(f), respectively.

 figure: Fig. 7

Fig. 7 (a) Circular high-contrast grating on hexagonal lattice. The yellow rectangle indicates a choice of the unit cell. Parameters: Λ = 1μm, η = 0.6. Dispersion curves of (b) Hx-dominant and (c) Hy-dominant eigenmodes possessing symmetry (red) and anti-symmetry (blue). ℜe[Hy] for (d) the first symmetric, (e) the first anti-symmetric, and (f) the second symmetric modes at frequency ω=0.52πcΛ, as indicated by the green dots in (c).

Download Full Size | PDF

3.2. Dual-mode analysis for perfect reflection and transmission

In many cases, the incident wave only strongly couples to a few eigenmodes due to mismatch of the field polarization or symmetry. We are particularly interested in the case when only two eigenmodes are strongly excited. Then we can engineer the structure to produce nearly perfect interference between them. In such case we can simplify our study to a dual-mode analysis. The transverse field in Eq. (5) is rewritten as

E˜t¯tDMADM
where the “DM” indicates dual-mode. Here, ¯tDM is a 2N×2 matrix containing only two eigenvectors. And ADM = [Ap,Aq]T contains two elements which are the expansion coefficients of the two modes. The superposition of eigenmodes in the HCG layer (Region II) forms a supermode. At certain wavelength λ and HCG thickness tg, such a supermode can satisfy the Fabry-Pérot resonance condition in the z^-direction. Assume the field is scaled by a complex number Ω = |Ω|e after a round-trip, we have
|Ω|eiφAII+(0)=R¯IIIAII(0)=R¯IIIeiK¯z.(tg)AII(tg)=R¯IIIeiK¯ztgR¯IIIIIAII+(tg)=R¯IIIeiK¯ztgR¯IIIIIeiK¯ztgAII+(0)=M¯(λ,tg)AII+(0)
where R¯III and R¯IIIII are the reflection matrices for the Region II eigenmodes bounced back by Region I (incident region) and Region III (transmitted region), respectively. Both reflection matrices can be obtained from Eq. (7). Equation (11) is an eigen-equation for the round-trip propagation matrix M¯(λ,tg). When the phase φ of the eigenvalue Ω is even or odd multiples of π, we have perfect constructive or destructive interference, respectively.

Using our dual-mode analysis, the dimensions of matrices and vectors in Eq. (11) are reduced to contain only two modes,

[M¯DM]2×2(λ,tg)[ApAq]=|ΩDM|eiφDM[ApAq]
where the phase ϕDM can be adjusted by the wavelength and the grating structure.

In Fig. 8(a) we see the reflection and transmission spectra of a 2D circular HCG on a hexagonal lattice under Ex/Hy-polarized normal incidence at λ = 1.55μm. We have perfect transmission (zero reflection) at two wavelengths and ultra-high transmission in between. Using our dual-mode analysis, we can easily extract the contribution to the reflected (0,0) mode from the back-coupling of each of the two strongly excited eigenmodes, as well as the contribution from all other modes, as shown in Fig. 8(b). Furthermore, we can calculate the phase difference of the contribution from the two modes. We can see that at zero reflection, indicated by the dashed lines, the coupling magnitudes of the two modes are almost the same, and the phase difference is almost ±π. This indicates the interference of the two modes is very close to but not yet perfectly destructive. We still see some coupling from other eigenmodes. This is because the eigenmodes in 2D HCG are hybrid modes and we have no perfect selectivity of polarizations. The result is a slight difference between the wavelengths for zero-reflection and the dual-mode perfect interference. Nonetheless, this analysis provides a very good estimation of the zero-reflection condition. The zero-transmission case can be analyzed similarly when we look at the forward-coupling to the transmitted region.

 figure: Fig. 8

Fig. 8 (a) Reflectivity (red) and transmissivity (blue) of a 2D circular high-contrast grating on a hexagonal lattice with Λ = 750nm, λ = 1.55μm, and tg = 713nm. Blue dashed lines indicate the wavelengths for perfect transmission. (b) Back-coupling magnitude from the first (red), second (black) symmetric eigenmodes and all other eigenmodes (green) in the grating to the (0,0) mode in the incident region. The dashed lines in (b) are at the same wavelengths as in (a). Blue solid line is the phase difference between the reflected waves back-coupled from the two eigenmodes.

Download Full Size | PDF

3.3. Resonance conditions in 2D high-contrast gratings

For a given desired wavelength, we can also adjust the HCG thickness tg to study the resonance condition. We first look at a rectangular HCG with the same structure as in Fig. 2 under Ex/Hy-polarized normal incidence. The dual-mode eigen-equation in Eq. (12) yields two eigen-solutions which we will name as supermode 1 and supermode 2. The phases of the eigenvalues (φ1,DM and φ2,DM), which are the round-trip phases for the two supermodes, are functions of the grating thickness. Figure 9(a) shows the half-trip phases ( φ1,DMhalf and φ2,DMhalf), and we record the HCG thicknesses at which the half-trip phases are even and odd multiples of π, indicated by the empty boxes and solid circles, respectively. At each given wavelength, we can solve this dual-mode eigen-problem and obtain the thicknesses for resonance to occur. Figure 9(b) shows the contour plot for the resonance conditions. Red and black lines indicate the half-trip phases φ1,DMhalf and φ2,DMhalf being odd multiples of π, respectively. Blue and green lines indicate the half-trip phases φ1,DMhalf and φ2,DMhalf being even multiples of π, respectively.

Figure 10(a) shows the reflectivity contour plot as a function of the grating thickness and wavelength. From Fig. 6(b) we can find the cutoff frequencies for the EH20 and EH02 to be ωcEH20=0.3922πcΛx(λcEH20=2.55Λx) and ωcEH02=0.8272πcΛx(λcEH02=1.21Λx). Between these two frequencies we see a dual-mode spectral window from the dispersion curves. These two frequencies are also indicated by the black lines in Fig. 10. We can see the ultra-high-reflection regions are mostly within the dual-mode window. Figure 10(b) shows the overlap between the resonance lines in Fig. 9(b) and the reflectivity contour in Fig. 10(a). Excellent agreement is shown which indicates that our dual-mode analysis can successfully predict the resonance and high-reflection conditions in a 2D HCG.

 figure: Fig. 9

Fig. 9 (a) Half-trip phases of the supermodes 1 (blue) and 2 (red) in the same grating as in Fig. 2 under λ = 2.1μm normal incidence. Empty boxes and solid circles indicate half-trip phases being odd and even multiples of π, respectively. (b) Resonance conditions for the grating thicknesses at given wavelengths. Red and blue lines indicate the half-trip phase of supermode 1 being odd and even multiples of π, respectively. Black and green lines indicate the half-trip phase of supermode 1 being odd and even multiples of π, respectively.

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 (a) Reflectivity contour plot of a rectangular high-contrast grating (same as in Fig. 2) as a function of the thickness tg and wavelength λ. The solid lines indicate λ = 2.55Λx and λ = 1.21Λx, corresponding to the cutoff frequencies ωcEH20=0.3922πcΛx and ωcEH02=0.8272πcΛx for the EH20- and EH02-like modes, as shown in Fig. 6(b). (b) Overlap between the resonance contour plot (white) and the reflectivity contour plot.

Download Full Size | PDF

For a circular HCG on a hexagonal lattice, we can also identify a dual-mode spectral window where only two symmetric modes with the same dominant polarization are strongly excited, as shown in Fig. 11(a). From the transmissivity contour plot of our designed transmission-type HCG in Fig. 11(b), we see wide high-transmission regions inside the dual-mode window. These high-transmission regions are mostly bounded by or along the resonance lines, as shown in Fig. 11(c). We again see an excellent overlap between the dual-mode resonance lines and the contour plot. From the resonance lines, we can easily identify the design regions for broadband or sharp-transition applications, such as reflectors, filters, and resonators.

 figure: Fig. 11

Fig. 11 (a) Dual-mode dispersion curves in a hexagonal-lattice grating and the dualmode window (same as in Fig. 7) is indicated by the blue dashed lines at ω=0.442πcΛ and ω=0.62πcΛ. (b) Transmission contour plot as a function of the grating thickness tg and wavelength λ. The dual-mode window indicated by the two black lines. (c) Overlap between the resonance lines (white) and the transmission contour plot.

Download Full Size | PDF

4. Engineering of 2D phased arrays using high-contrast gratings

With the knowledge of the resonance conditions and the mechanisms for ultra-high reflection or transmission, we can also apply the initial design to 2D phase plates using HCGs. Now we fix the wavelength and grating thickness, but vary the transverse structure to find a full 2π reflection (or transmission) phase tuning range while maintaining the power efficiency. Figure 12 shows an example of a transmission-type hexagonal-lattice HCG operating at λ = 1.55μm with a thickness of 0.8μm. We are able to find an optimum design range by fixing Λ = 0.75μm, and tuning η from 10% to 63%, as indicated by the black lines. Thus we have a mapping between the transmission phase and the transverse geometry. The design of the phase plate is essentially aimed at generating a transverse position-dependent phase alternation profile, that is, ΔΦ(x,y). This can be translated into the design of the η(x,y) profile. In many applications, we would need substrates for the HCGs. Figure 13(a) shows the magnitude and phase tuning by the HCG duty cycle with different substrate thicknesses (ts = 0, λ/4, and λ/2). We can see the full 2π phase tuning range is still obtainable except the phase discontinuity occurs at different locations. The magnitude remains almost unchanged for thickness being multiples of half-wavelength, but is strongly perturbed at quarter-wavelength. In the situation when the grating transmission is not close enough to unity, we can use the substrate to further improve the transmission, and find the optimum value of ts, as shown in Fig. 13(b).

 figure: Fig. 12

Fig. 12 Contour plots as functions of the grating period Λ and duty cycle η for the (a) magnitude and (b) phase of the transmission through a hexagonal-lattice grating under λ = 1.55μm normal incidence. The black lines indicates a 2π phase range at Λ = 750nm. The white lines indicate the contour for 90% transmission.

Download Full Size | PDF

 figure: Fig. 13

Fig. 13 (a) Tuning of the transmission magnitude (solid) and phase (dots) by grating duty cycle with thickness being 0 (blue), 258nm (black), and 517nm (red). Same structure as in Fig. 7 and λ = 1.55μm. (b) Transmission magnitude (blue) and phase (green) as functions of the substrate thickness with 48% duty cycle.

Download Full Size | PDF

By introducing a phase alternation ΔΦ based on the transverse spatial location (x,y), we can have 2D HCGs behaving like many conventional optical components. For example, having ΔΦ(x,y)=2πλ(xcosϕsinθ+ysinϕsinθ), we can steer the normal-incident plane wave into an oblique direction of x^cosϕsinθ+y^sinϕsinθ+z^cosθ, where θ and ϕ are the polar and azimuthal angles of the transmitted wave vector. For a lens with a focusing length f, we need ΔΦ(x,y)=2πλ(ff2+x2+y2). For an axicon [24] with an opening angle αaxi, an index nr and a radius R, we have ΔΦ(x,y)=(Rx2+y2)(nr1)tanαaxi. For a focusing lens that generates orbital angular momentum (OAM) of mh¯(m=0,±1,±2), we have ΔΦ(x,y)=2πλ(ff2+x2+y2+mϕ). Using Fig. 13 we can design the local duty cycle η(x, y) for the 2D HCG based on the desired phase alternation ΔΦ(x, y).

Once we have the design for 2D HCGs, we use the 3D FDTD method to verify the performance. Figure 14(a) shows a normal-incident x-polarized plane wave toward the +z^-direction at λ = 1.55μm passing through our designed HCG beam steering plate with a 15° steering angle. The two dashed lines indicate the normal and 15° from normal directions. The transmitted wave is indeed traveling along the desired direction with high power efficiency. Figure 14(b) shows a normal-incident x-polarized Gaussian beam with a beam waist w0 = 7μm transmitting through a 10-μm-radius HCG phase plate, which is designed to have a focal length f = 15μm. The magenta dashed lines indicate the focusing behavior of an equivalent lens with a numerical aperture of 0.5547. Figure 14(c) shows that the same Gaussian beam passes through a HCG phase plate, which acts like an axicon [25], and becomes a Bessel beam. The magenta dashed lines indicate an Bessel beam opening angle of 2θBS = 0.4rad, which is the result of an equivalent axicon with a radius R = 10μm, an index nr = 1.5, and an axicon opening angle αaxi = θBS/(nr 1) = 0.4rad. Such a Bessel beam has a maximum propagation distance zmax ≈ w0BS = 35μm, within which the beam will have no diffraction. This maximum distance can be increased by reducing the axicon opening angle αaxi.

 figure: Fig. 14

Fig. 14 e[Ex] in the xz-plane for an x-polarized normal-incident wave toward +z^ at λ = 1.55μm transmitting through HCG phase plates designed for Gaussian beams to: (a) be deflected by 15°; (b) focus at a 15-μm distance; (c) be converted to Bessel beams. Black solid lines indicate the grating layer. Blue dashed lines indicate the source locations. Other dashed lines are for visual aids. All field intensities are normalized by the incident field intensity.

Download Full Size | PDF

We can allow our 2D HCG phase plate to focus and simultaneously generate orbital angular momentum as a beam passes through. Such modification can be easily done by adding a spatial-dependent, or more specifically, an azimuthal angle-dependent phase delay. Using the phase information from Fig. 13(a) we arrive at a 2D HCG design in Fig. 15(a) on top of a glass substrate, with an operation wavelength at λ = 1.55μm. The transmitted field intensity profiles at z = f − 4λ, z = f, and z = f + 4λ are shown in Figs. 15(b)–15(d), respectively, where the focal length is designed to be 20μm. We can see a clear intensity null at the center, which is necessary for nonzero OAM. The incident beam has a Gaussian distribution with a beam waist of 7μm and zero OAM, as shown in Fig. 15(e). The focusing behavior is observed by comparing the beam waists and the peak intensities 4λ below, 4λ above, and exactly at the focal plane, as well as those of the source. Our design provides high transmissivity with a power efficiency above 90%. Figures 15(f)–15(h) show the phase distribution profiles at z = f −λ/3, z = f, and z = f + λ/3. The spatial-dependent phase distribution already indicates nonzero OAM. The phase of the transmitted wave increases from −π to π as the azimuthal angle increases in a 2π range, indicating the OAM is +1h¯ per photon. Alternatively, we see the phase profile rotates clockwise by 2π/3 as the beam propagates by Δz = λ/3 in the +z^-direction, showing a left-hand helical phase front. Therefore the OAM is indeed +1h¯ per photon.

 figure: Fig. 15

Fig. 15 (a) Structure of the 2D high-contrast grating phase plate for focusing at f = 20μm and generating +1h¯ orbital angular momentum. Field intensities at z = f − 4λ, z = f, and z = f +4λ are shown in (b), (c), and (d), respectively. (e) Gaussian source intensity with a beam waist of 12μm. Phase distributions (in π) of the transmitted wave at z = f − λ/3, z = f, and z = f +λ/3. are shown in (f), (g), and (h), respectively.

Download Full Size | PDF

5. Conclusion

We have investigated the physics of 2D high-contrast gratings (HCGs). Our in-house developed 2D rigorous coupled-wave analysis (RCWA) program is shown to be an efficient and accurate tool for understanding the optical behavior of 2D HCGs with various structural and incidence parameters. We further demonstrate the design rules for various optical applications using 2D HCGs, based on the HCG mode properties we obtained from RCWA. Using our dual-mode analysis, the design process can be largely simplified. Once our high-performance initial design is obtained, it can be optimized according to the desired functionalities, such as broadband reflection, high-Q resonance, filtering, polarizing, etc. At last, we discuss the design of phase plates using 2D HCGs. Our designed HCG phase plates can function as conventional optical components, such as lenses, deflectors, axicons, spiral phase shifters, with excellent agreement and ultra-high performance.

Acknowledgments

This work is supported by the DARPA E-PHI program.

References and links

1. M. Born and E. Wolf, Principles of Optics7th ed. (Cambridge University Press, 1999). [CrossRef]  

2. N. Yu and F. Capasso, “Flat optics with designer metasurfaces,” Nat. Mater. 13, 139–150 (2014). [CrossRef]   [PubMed]  

3. M. C. Y. Huang, Y. Zhou, and C. J. Chang-Hasnain, “A surface-emitting laser incorporating a high-index-contrast subwavelength grating,” Nat. Photonics 1, 119–122 (2007). [CrossRef]  

4. H. Yang, D. Zhao, S. Chuwongin, J.-H. Seo, W. Yang, Y. Shuai, J. Berggren, M. Hammar, Z. Ma, and W. Zhou, “Transfer-printed stacked nanomembrane lasers on silicon,” Nat. Photonics 6, 615–620 (2012). [CrossRef]  

5. B. Zhang, Z. Wang, S. Brodbeck, C. Schneider, M. Kamp, S. Höfling, and H. Deng, “Zero-dimensional polariton laser in a subwavelength grating-based vertical microcavity,” Light Sci. Appl. 3, e135 (2014). [CrossRef]  

6. C. J. Chang-Hasnain and W. Yang, “High-contrast gratings for integrated optoelectronics,” Adv. Opt. Photonics 4, 379–440 (2012). [CrossRef]  

7. C. F. R. Mateus, M. C. Y. Huang, Y. Deng, A. R. Neureuther, and C. J. Chang-Hasnain, “Ultrabroadband mirror using low-index cladded subwavelength grating,” IEEE Photonics Technol. Lett. 16, 518–520 (2004). [CrossRef]  

8. Y. Zhou, M. Moewe, J. Kern, M. C. Huang, and C. J. Chang-Hasnain, “Surface-normal emission of a high-Q resonator using a subwavelength high-contrast grating,” Opt. Express 16, 17282–17287 (2008). [CrossRef]   [PubMed]  

9. J. Ferrara, W. Yang, L. Zhu, P. Qiao, and C. J. Chang-Hasnain, “Heterogeneously integrated long-wavelength VCSEL using silicon high contrast grating on an SOI substrate,” Opt. Express 23, 2512–2523 (2015). [CrossRef]   [PubMed]  

10. D. Fattal, J. Li, Z. Peng, M. Fiorentino, and R. G. Beausoleil, “Flat dielectric grating reflectors with focusing abilities,” Nat. Photonics 4, 466–470 (2010). [CrossRef]  

11. S. Vo, D. Fattal, W. V. Sorin, Z. Peng, T. Tran, M. Fiorentino, and R. G. Beausoleil, “Sub-wavelength grating lenses with a twist,” IEEE Photonics Technol. Lett. 26, 1375–1378 (2014). [CrossRef]  

12. V. Karagodsky, F. G. Sedgwick, and C. J. Chang-Hasnain, “Theoretical analysis of subwavelength high contrast grating reflectors,” Opt. Express 18, 16973–16988 (2010). [CrossRef]   [PubMed]  

13. M. G. Moharam and T. K. Gaylord, “Rigorous coupled-wave analysis of planar-grating diffraction,” J. Opt. Soc. Am. 71, 811–818 (1981). [CrossRef]  

14. S. G. Tikhodeev, A. L. Yablonskii, E. A. Muljarov, N. A. Gippius, and T. Ishihara, “Quasiguided modes and optical properties of photonic crystal slabs,” Phys. Rev. B 66, 045102 (2002). [CrossRef]  

15. J. N. Mait, D. W. Prather, and M. S. Mirotznik, “Design of binary subwavelength diffractive lenses by use of zeroth-order effective-medium theory,” J. Opt. Soc. Am. A 16, 1157–1167 (1999). [CrossRef]  

16. M. A. Golub and A. A. Friesem, “Effective grating theory for resonance domain surface-relief diffraction gratings,” J. Opt. Soc. Am. A 22, 1115–1126 (2005). [CrossRef]  

17. R. W. Wood, “On a remarkable case of uneven distribution of light in a diffraction grating spectrum,” Phylos. Mag. 4, 396–402 (1902). [CrossRef]  

18. L. Rayleigh, “Note on the remarkable case of diffraction spectra described by Prof. Wood,” Philos. Mag. 14, 60 (1907) [CrossRef]  

19. U. Fano, “The theory of anomalous diffraction gratings and of quasi-stationary waves on metallic surfaces (Sommerfeld’s waves),” J. Opt. Soc. Am. 31, 213–222 (1941). [CrossRef]  

20. C. H. Palmer, “Parallel diffraction grating anomalies,” J. Opt. Soc. Am. 42, 269–273 (1952). [CrossRef]  

21. A. Hessel and A. A. Oliner, “A new theory of Wood’s anomalies on optical gratings,” Appl. Opt. 4, 1275 (1965). [CrossRef]  

22. A. Maurel, S. Felix, J.-F. Mercier, A. Ourir, and Z. E. Djeffal, “Wood’s anomalies for arrays of dielectric scatterers,” J. Europ. Opt. Soc. Rap. Public. 9, 14001 (2014). [CrossRef]  

23. E. A. J. Marcatili, “Dielectric rectangular waveguide and directional coupler for integrated optics,” Bell Syst. Tech. J. 48, 2071–2102 (1969). [CrossRef]  

24. J. H. Mcleod, “The axicon: a new type of optical element,” J. Opt. Soc. Am. 44, 592 (1954). [CrossRef]  

25. G. Indebetouw, “Nondiffracting optical fields: some remarks on their analysis and synthesis,” J. Opt. Soc. Am. A 6, 150–152 (1989). [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 (15)

Fig. 1
Fig. 1 (a) Schematic of a 2D high-contrast grating on a rectangular lattice with a substrate. The plane-wave incident polar and azimuthal angles are θ and ϕ, respectively. Parameters: Λ x and Λ y (periods in x ^ and ŷ), Dx and Dy (grating widths in x ^ and ŷ), ng and ns (grating and substrate indices), and tg and ts (grating and substrate thicknesses). (b) Schematic of a 2D high-contrast grating on a hexagonal lattice with a substrate. Parameters: period Λ, rod diameter d, grating and substrate indices ng and ns, grating and substrate thicknesses tg and ts.
Fig. 2
Fig. 2 (a) Comparison among the (0,0)-th order reflectivity spectra of a 2D high-contrast grating (HCG) under normal incidence calculated using finite-element method (blue star), finite-difference time-domain (red dots), and rigorous coupled-wave analysis (RCWA) using N = 289 (magenta), N = 441 (green), and N = 625 (blue). HCG parameters are Λ x = 1μm, Λ y = 0.5μm, Dx = 0.6μm, Dy = 0.3μm, and tg = 0.5μm. (b) Spectra of normalized scattered power in the z ^ -direction for (0,0), (+1,0), and (1,0) spectral orders (blue, red, and black, respectively), and the sum of the three spectra (green).
Fig. 3
Fig. 3 (a) Incident polar angle dependent reflectivity (black), transmissivity (magenta), and their sum (green) for the (0,0) fundamental order under p-polarized incidence with λ = 2μm calculated using rigorous coupled-wave analysis, and comparison with finite-element method (red and blue crosses). The same grating structure is solved under (b) s-polarized and (c) p-polarized incidence with wavelength being λ = 0.6μm. Normalized scattered power fluxes in the z ^ -direction for the (0,0)-th, (+1,0)-th, (1,0)-th, (2,0)-th, and (3,0)-th spectral orders are shown as the blue, red, black, green, and cyan lines, respectively. The magenta line indicates the total normalized scattered power flux in the z ^ -direction. Arrows indicate the cutoff angles for spectral orders to appear or disappear.
Fig. 4
Fig. 4 (a) Reflectivity (blue), transmissivity (red), and the normalized scattered power (green) of a normal incident plane wave with λ = 1.55μm as functions of the azimuthal polarization angle. The arrows indicate the components parallel and perpendicular to the incident wave. (b) Phase differences between the parallel and perpendicular components for reflection (red) and transmission (blue) as functions of the azimuthal polarization angle. The 90° phase differences correspond to polarization angles of 27.7° and 52.7° for reflection and transmission types, respectively.
Fig. 5
Fig. 5 Design procedures of 2D high-contrast gratings.
Fig. 6
Fig. 6 (a) HE-like even eigenmodes in a 2D rectangular grating on a rectangular lattice. The two dashed lines indicate the kz = ω/c and kz = ngω/c light lines. (b) EH-like even eigenmodes in a 2D rectangular grating on a rectangular lattice. (c) ℜe[Hy] for the EH00-like eigenmode and EH20-like eigenmode at ω = 0.8 2 π c Λ x , as indicated by the green dots in (b).
Fig. 7
Fig. 7 (a) Circular high-contrast grating on hexagonal lattice. The yellow rectangle indicates a choice of the unit cell. Parameters: Λ = 1μm, η = 0.6. Dispersion curves of (b) Hx-dominant and (c) Hy-dominant eigenmodes possessing symmetry (red) and anti-symmetry (blue). ℜe[Hy] for (d) the first symmetric, (e) the first anti-symmetric, and (f) the second symmetric modes at frequency ω = 0.5 2 π c Λ , as indicated by the green dots in (c).
Fig. 8
Fig. 8 (a) Reflectivity (red) and transmissivity (blue) of a 2D circular high-contrast grating on a hexagonal lattice with Λ = 750nm, λ = 1.55μm, and tg = 713nm. Blue dashed lines indicate the wavelengths for perfect transmission. (b) Back-coupling magnitude from the first (red), second (black) symmetric eigenmodes and all other eigenmodes (green) in the grating to the (0,0) mode in the incident region. The dashed lines in (b) are at the same wavelengths as in (a). Blue solid line is the phase difference between the reflected waves back-coupled from the two eigenmodes.
Fig. 9
Fig. 9 (a) Half-trip phases of the supermodes 1 (blue) and 2 (red) in the same grating as in Fig. 2 under λ = 2.1μm normal incidence. Empty boxes and solid circles indicate half-trip phases being odd and even multiples of π, respectively. (b) Resonance conditions for the grating thicknesses at given wavelengths. Red and blue lines indicate the half-trip phase of supermode 1 being odd and even multiples of π, respectively. Black and green lines indicate the half-trip phase of supermode 1 being odd and even multiples of π, respectively.
Fig. 10
Fig. 10 (a) Reflectivity contour plot of a rectangular high-contrast grating (same as in Fig. 2) as a function of the thickness tg and wavelength λ. The solid lines indicate λ = 2.55Λ x and λ = 1.21Λ x , corresponding to the cutoff frequencies ω c EH 20 = 0.392 2 π c Λ x and ω c EH 02 = 0.827 2 π c Λ x for the EH20- and EH02-like modes, as shown in Fig. 6(b). (b) Overlap between the resonance contour plot (white) and the reflectivity contour plot.
Fig. 11
Fig. 11 (a) Dual-mode dispersion curves in a hexagonal-lattice grating and the dualmode window (same as in Fig. 7) is indicated by the blue dashed lines at ω = 0.44 2 π c Λ and ω = 0.6 2 π c Λ . (b) Transmission contour plot as a function of the grating thickness tg and wavelength λ. The dual-mode window indicated by the two black lines. (c) Overlap between the resonance lines (white) and the transmission contour plot.
Fig. 12
Fig. 12 Contour plots as functions of the grating period Λ and duty cycle η for the (a) magnitude and (b) phase of the transmission through a hexagonal-lattice grating under λ = 1.55μm normal incidence. The black lines indicates a 2π phase range at Λ = 750nm. The white lines indicate the contour for 90% transmission.
Fig. 13
Fig. 13 (a) Tuning of the transmission magnitude (solid) and phase (dots) by grating duty cycle with thickness being 0 (blue), 258nm (black), and 517nm (red). Same structure as in Fig. 7 and λ = 1.55μm. (b) Transmission magnitude (blue) and phase (green) as functions of the substrate thickness with 48% duty cycle.
Fig. 14
Fig. 14e[Ex] in the xz-plane for an x-polarized normal-incident wave toward + z ^ at λ = 1.55μm transmitting through HCG phase plates designed for Gaussian beams to: (a) be deflected by 15°; (b) focus at a 15-μm distance; (c) be converted to Bessel beams. Black solid lines indicate the grating layer. Blue dashed lines indicate the source locations. Other dashed lines are for visual aids. All field intensities are normalized by the incident field intensity.
Fig. 15
Fig. 15 (a) Structure of the 2D high-contrast grating phase plate for focusing at f = 20μm and generating + 1 h ¯ orbital angular momentum. Field intensities at z = f − 4λ, z = f, and z = f +4λ are shown in (b), (c), and (d), respectively. (e) Gaussian source intensity with a beam waist of 12μm. Phase distributions (in π) of the transmitted wave at z = f − λ/3, z = f, and z = f +λ/3. are shown in (f), (g), and (h), respectively.

Equations (12)

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

k ( m , n ) = x ^ k x + y ^ k y + z ^ k z , k x = k i x + G x , m , k y = k i y + G y , n , G x , m = m π Λ x , m = 0 , ± 1 , ± 2 , , ± N x , G y , n = n π Λ y , n = 0 , ± 1 , ± 2 , , ± N y , k z = ± 4 π 2 n r 2 λ 2 k x 2 k y 2
E t ( i ) = e i K z ( i ) z ( m , n ( x ^ E ˜ x , ( m , n ) ( i ) + y ^ E ˜ y , ( m , n ) ( i ) ) e i G x , m x + i G y , n y ) e i k i x x + i k i y y
M ¯ E ˜ t ( i ) = ( K z ( i ) ) 2 N ¯ E ˜ t ( i )
E ˜ t ( i ) = ( E ˜ x , ( N x , N y ) ( i ) , , E ˜ x , ( m , n ) ( i ) , , E ˜ x , ( N x , N y ) ( i ) , E ˜ y , ( N x , N y ) ( i ) , , E ˜ y , ( m , n ) ( i ) , , E ˜ y , ( N x , N y ) ( i ) ) T
E ˜ t = i A i E ˜ t ( i ) = ¯ t A
[ E ˜ t ( z ) H ˜ t ( z ) ] = [ ¯ t ¯ t ¯ t ¯ t ] [ A + ( z ) A ( z ) ] = [ ¯ t ¯ t ¯ t ¯ t ] [ e i K ¯ z z 0 0 e i K ¯ z z ] [ A + A ]
[ A II + A II ] = [ ¯ t II ¯ t II ¯ t II ¯ t II ] 1 [ ¯ t I ¯ t I ¯ t I ¯ t I ] [ A I + A I ] = [ T ¯ I II ] 4 N × 4 N [ A I + A I ]
( k 0 sin θ cos ϕ + m π Λ x ) 2 + ( k 0 sin θ sin ϕ + n π Λ y ) 2 = k 0 2
θ c = sin 1 ( 1 | m | λ Λ x ) for ( + | m | , 0 ) to disappear θ c = sin 1 ( | m | λ Λ x 1 ) for ( | m | , 0 ) to appear
E ˜ t ¯ t DM A DM
| Ω | e i φ A II + ( 0 ) = R ¯ II I A II ( 0 ) = R ¯ II I e i K ¯ z . ( t g ) A II ( t g ) = R ¯ II I e i K ¯ z t g R ¯ II III A II + ( t g ) = R ¯ II I e i K ¯ z t g R ¯ II III e i K ¯ z t g A II + ( 0 ) = M ¯ ( λ , t g ) A II + ( 0 )
[ M ¯ DM ] 2 × 2 ( λ , t g ) [ A p A q ] = | Ω DM | e i φ DM [ A p A q ]
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.