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

Computationally fast EM field propagation through axi-symmetric media using cylindrical harmonic decomposition

Open Access Open Access

Abstract

We describe and provide a systematic procedure for computationally fast propagation of arbitrary vector electromagnetic (EM) fields through an axially symmetric medium. A cylindrical harmonic field propagator is chosen for this purpose and in most cases, this is the best and the obvious choice. Firstly, we describe the cylindrical harmonic decomposition technique in terms of both scalar and vector basis for a given input excitation field. Then we formulate a generalized discrete Fourier-Hankel transform to achieve efficient vector basis decomposition. We allow a slower, pre-computation step, that finds a representation of the axi-symmetric medium as a transfer matrix in a discrete, cylindrical-harmonic basis. We find this matrix from a series of axi-symmetric (2D) finite element simulations (also known as the 2.5D technique). This transfer matrix approach significantly reduces the computational load when the transverse size or range exceeds about 30 wavelengths. This matrix is independent of the input excitation field for a given space-bandwidth product and hence makes it reusable for different excitation fields. We numerically validate the above approaches for different axi-symmetric EM scattering media which include a hemispherical gradient-index Maxwell’s fish-eye lens, a transformation optics designed spherical invisibility cloak, a thin aspheric lens, and a cylindrical perfect lens.

© 2016 Optical Society of America

1. Introduction

Large scale computational electromagnetic problems involving propagation through inhomogeneous media are often computationally very expensive. One of the major challenges is to reduce the computational cost of using methods like finite element or finite difference, in electrically large problems, without sacrificing the accuracy of the result. One approach for reducing the computational complexity is to downgrade the optics abstraction level from a vector EM model to a scalar optics (wave optics) model or even further down to ray tracing model. However, in some cases these approaches are unacceptable. The problem difficulty increases when complete 3D vector field results are required in real-time. Some of the motivating applications are: forward model calculation in computational imaging systems with an array of axi-symmetric lenses [1–3], radome design and distortion mitigation for radio telescopes and airborne vehicular antennas [4,5], specialized coaxial waveguide or conical feed antenna design [6,7], magnetic coil design or Fresnel plate design for wireless power transfer or super-lensing applications [8,9] and recently, in plasmonic applications such as excitonic laser [10,11] and metallic nanosphere design for plasmonic enhancement at optical frequencies [12–14]. The above mentioned electromagnetic problems are computationally hard due to the fact that the scattering structures involved are large compared to the operating wavelength (> 100λ), or a design optimization process requires repeated calculation of the scattering fields. However, in some cases the scattering structures might possess mirror/reflection, translational or rotational symmetry which can be exploited to reduce the computational cost [15]. For example, a simple circular cross-section waveguide possesses all of the above symmetries.

In this article, we focus on exploiting rotational/axial symmetry in scattering structures (which are bodies of revolution), as this symmetry reduces the dimensionality of the problem. In such structures the solutions (modes) can be represented by products of separate functions of radial, azimuthal and longitudinal variables. This in turn helps to reduce a 3D problem , ϕ, z) to a 2D problem , z) and thereby reduces the computational effort [16]. (The excitation source need not be axi-symmetric, to exploit this symmetry.) This happens due to the fact that each cylindrical harmonic propagates independently. Often, the fields of the arbitrary source are decomposed into a series of plane wave components using Fourier transform. The response from each of the obliquely incident (w.r.t z) plane wave is simulated in 2D , z) and these responses are revolved and summed to get the complete 3D solution (2.5D technique). The spectral decomposition happens in plane wave basis (Cartesian) and the evaluation of scattered fields happens in cylindrical basis.

Instead of performing spectral decomposition of the excitation source in plane wave basis, we wish to perform the decomposition in cylindrical harmonic basis (also a complete basis). The choice of cylindrical basis avoids the unnecessary coordinate transformation (which, in some cases, requires regridding or interpolation) that might be needed if plane-wave basis is used. We consider both a scalar and vector cylindrical-harmonic basis decomposition to accommodate the cases when either: the normal components of the electric and magnetic fields are known, or the transverse components of the electric field are known. We also formulate a discrete cylindrical harmonic decomposition using generalized discrete Fourier-Hankel transforms. This decomposition approach also provides a significant computational advantage in terms of constructing the overall response transfer matrix for the scattering structure involved. This sparse block diagonal transfer matrix is independent of the arbitrary source of excitation, provided, one pre-calculates the scattering response for all possible propagating modes with unit mode coefficients for a given space-bandwidth product. Our approach provides the most general methodology to exploit axial symmetry, however, some very specific numerical methods exist for cylindrical waveguides [17] and cylindrical rod scatterers [18].

In this article, we first formalize cylindrical harmonic decomposition of arbitrary fields (including both TMz and TEz polarizations, which are in further sections referred to as TM and TE respectively) by deriving the continuous integral equations for mode coefficients using appropriate orthogonality conditions. In Section 3, we discretize these continuous integrals using discrete Fourier-Hankel transforms (DFHT). For the discrete vector basis we formulate generalized DFHT. We also develop a composite indexing approach which facilitates the expression of these transforms in matrix form, contributing to the efficiency of the cylindrical decomposition process. In Section 4, we describe the 2.5D cylindrical harmonic simulations using the COMSOL RF module, whose scattered field results are used to compute the transfer matrix for a given scattering medium. We show four numerical validation results involving axi-symmetric scattering media namely: (a) gradient refractive index (GRIN) hemispherical Maxwell’s fish-eye lens, (b) Coordinate transformed spherical invisibility cloak with anisotropic material properties illuminated with offset vector Gaussian beam, (c) A simple ∼ 40λ diameter thin-aspheric-plano-convex lens illuminated with Gaussian vector beam, and (d) An isotropic negative index perfect lens with arbitrary excitation field. In each case, we performed a series of axi-symmetric 2D finite element simulations for all possible propagating modes to construct a transfer matrix of the scattering medium. Then this transfer matrix is multiplied by the cylindrical harmonic coefficients of the excitation field to derive the output mode coefficients at a prescribed plane. The results of this method are compared and agree well with analytical results. For the GRIN lens case (a), we also perform complete conventional 3D finite element simulations to test accuracy and compare the computational resource requirements. For the other numerical validation cases, there exist straightforward analytical results for comparison. The choice of using a cylindrical harmonic propagator in such problems is also well justified, and we compare the computational requirement of this method with an another approach, namely, the plane wave propagator. Even though we restrict our discussions to EM problems, some of the techniques and methods described in this article can also be extended to other areas of computational physics problems such as: acoustics, thermodynamics, and fluid dynamics.

2. Cylindrical harmonic decomposition

Cylindrical harmonic decomposition (CHD) is the process of decomposing an arbitrary function into basic patterns/harmonics that have simple radial and angular structures. Cylindrical harmonics, just like plane waves or spherical harmonics, form a complete orthonormal basis set. Any arbitrary field that is a solution to Helmholtz’s equation can be represented as an infinite weighted sum of these cylindrical harmonics. Often, for most of the real world signals of interest, the cylindrical harmonic expansion converges rapidly leading to good representation accuracy using a finite and relatively small number of cylindrical harmonic modes. Any arbitrary vector field can be decomposed into cylindrical harmonics as follows. Consider an arbitrary electromagnetic vector field including both TM and TE polarizations propagating in +z direction. We decompose either the z-components of both the electric and magnetic fields, or the transverse component of electric field, into cylindrical harmonic modes. We formulate the expansions so that both types use the same set of expansion coefficients. These expansion coefficients can then be propagated (from a plane z = z0 to another plane z = z1) by a transfer matrix and recomposed into the spatial field representation as follows.

Et(ρ,ϕ,z0)Et(ρ,ϕ,z1)oralmnTblmnorEz(ρ,ϕ,z0),Hz(ρ,ϕ,z0)Ez(ρ,ϕ,z1),Hz(ρ,ϕ,z1)
where, sub-index l refers to the mode type, l = 1 → TM mode and l = 2 → TE mode; m and n are the azimuthal and radial mode numbers respectively; almn and blmn are the input and output plane cylindrical harmonic mode coefficients related through the transfer matrix T. Field components are derived by expanding and summing all their respective basis functions.

For decomposing the field z-components we use the usual scalar basis functions employed in the analysis of cylindrical wave-guides, but the transverse electric field decomposition requires a less well-known set of vector basis functions. Below we show both the continuous and discretized versions of this change of basis. For the discrete case of the change to the scalar basis we use a Discrete Fourier-Hankel Transform (DFHT). For the discrete case of the change to the vector basis we use a generalization of the DFHT.

In what follows we assume a maximum radius over which the fields are non-negligible is given by R, and a maximum radial phase constant (or maximum radial spatial bandwidth) is given by B. The latter, might be for example, the maximum radial phase constant which permits propagating fields, i.e. the plane wave propagation constant in the medium.

2.1. Scalar basis

The expansion of arbitrary z-components of the fields at z = z0 is given by [19],

Ez(ρ,ϕ,z0)=m,na1mnψ1mn(ρ,ϕ)
η0Hz(ρ,ϕ,z0)=m,na2mnψ2mn(ρ,ϕ)
and at z = z1 is given by,
Ez(ρ,ϕ,z1)=m,nb1mnψ1mn(ρ,ϕ)
η0Hz(ρ,ϕ,z1)=m,nb2mnψ2mn(ρ,ϕ)
The Eqs. (2a) and (3a) represents the TM modes and Eqs. (2b) and (3b) represents the TE modes. ψlmn, almn, and blmn are the basis functions, input mode coefficients, and output mode coefficients respectively, with mode type, l, azimuthal mode number, m, and radial mode number, n. The normalized basis functions are given by,
ψlmn(ρ,ϕ)=ClmnJm(βlmnρ)exp(jmϕ)
The radial phase constants β1mn and β2mn are found from the roots of the first kind Bessel functions and their derivatives respectively.
Jm(β1mnR)=0andJm(β2mnR)=0
The roots are frequently defined with respect to unit radius as,
Jm(χ1mn)=0andJm(χ2mn)=0
The Eq. (6) arises from PEC (perfect electric conductor) boundary conditions. The normalizing constants are given by (see Appendix A),
C1mn1=πRJm+1(χ1mn)
C2mn1=πR[Jm2(χ2mn)Jm1(χ2mn)Jm+1(χ2mn)]1/2

The basis functions and the normalizing coefficients are defined separately for convenience in representing the discrete transforms below. Within each mode type (TM or TE) the basis functions are ortho-normal,

ψlmn|ψlmn=0Rππψlmn*(ρ,ϕ)ψlmn(ρ,ϕ)dϕρdρ=δmmδnn
where δij is the Kronecker delta function. Using Eq. (15) below, and the fact that Ez is zero for TE modes, one can see that, for the full vector electric field, TM and TE fields are orthogonal. However, the scalar basis functions of different mode types are not necessarily orthogonal to each other. Having access to Ez and Hz means that the TM and TE fields are already separated, and need not be decomposed. l = 1 basis functions need only, and should only, be used with Ez, and the l = 2 basis functions need only, and should only, be used with Hz.

Using the orthogonality condition and the expansions above, we find the expressions for the cylindrical harmonic mode coefficients,

almn=ψlmn|Ezl=1η0Hzl=2=0Rππψlmn*(ρ,ϕ){Ez(ρ,ϕ,z0)l=1η0Hz(ρ,ϕ,z0)l=2}dϕρdρ

2.2. Vector basis

We expand the transverse electric field,

Et=EEzz^
in terms of a vector basis set at z = z0
Et(ρ,ϕ,z0)=l,m,nalmnϒlmnΨlmn(ρ,ϕ)
and at z = z1,
Et(ρ,ϕ,z1)=l,m,nblmnϒlmnΨlmn(ρ,ϕ)
where we can use the same expansion coefficients as the scalar basis and the customary normalization of these vector basis functions, if we include an additional factor,
ϒlmn=jω20μ0βlmn2δl1
The normalized vector basis functions are given by [20],
Ψ1mn(ρ,ϕ)=C1mn[Jm(β1mnρ)ρ^+jmJm(β1mnρ)β1mnρϕ^]exp(jmϕ)
Ψ2mn(ρ,ϕ)=C2mn[jmJm(β2mnρ)β2mnρρ^+Jm(β2mnρ)ϕ^]exp(jmϕ)
using the same radial phase constants βlmn from Eq. (5) and normalizing constants Clmn from Eq. (7) as were found for the scalar basis functions above. The orthogonality condition (including cross-mode-type) is given by,
Ψlmn|Ψlmn=0RππΨlmn*(ρ,ϕ)Ψlmn(ρ,ϕ)dϕρdρ=δllδmmδnn
where the primed superscript indices (l′, m′, n′) just refer to independent indices of same type as (l, m, n).

Using this condition and the field expansion above, we find the expression for the expansion coefficients as,

almnϒlmn=Ψlmn|Et=0RππΨlmn*(ρ,ϕ)Et(ρ,ϕ,z0)dϕρdρ
One could use scalar or vector basis expansions depending on the type of fields specified. For both expansions, the z = z1 plane coefficients are related to z = z0 plane coefficients through the transfer matrix, T, as:
blmn=l,m,nTlmnlmnalmn
For example, the transfer function of a uniform medium is,
Tlmnlmn=δllδmmδnnexp(jβzlmn(z1z0))where,βzlmn=ω2μβlmn2
since all the modes propagate independently. The construction of the transfer matrix is discussed in detail in Section 4.

3. Discrete Fourier-Hankel transform

The discretized form of transforming to the cylindrical harmonic basis from the position basis requires a combined discrete Fourier transform of the angle, ϕ, and a discrete Hankel transform of the radial variable, ρ. This is seldom done because, on the surface, the Discrete Fourier-Hankel Transform (DFHT) appears to be more complex and confusing than the applying the Discrete Fourier Transform (DFT) in the Cartesian basis. A relatively simple, but useful, scheme for understanding and implementing the DFHT incorporates three composite indices: one for the cylindrical harmonic components, l, one for the input samples, i, and one for the output samples, α. Here we use bold face to denote these composite indices.

We wish to compute a set of cylindrical harmonic coefficients that span a specified range of in-plane spatial bandwidth, B. For a given domain radius, R, this implies a specific space-bandwidth product, BR. Our composite index should include all modes that lie in this space-bandwidth product, and thus satisfy βlmn R = χlmnBR.

We can start from an initial set of azimuthal and radial mode indices, m and n, with maximum values given by,

M=max{m|χ1m1BRorχ2m1BR}N=max{n|χ10nBRorχ21nBR}
and choose the index triples, lmn, that represent modes in the specified space-bandwidth product range,
{l}={l,m,n|χlmnBR}{l=1,2}{MmM1}{1nN}
Such a composite index is determined solely by the space-bandwidth product and can be computed and stored once for repeated use with different field patterns (bound by the space-bandwidth product).

The best accuracy for cylindrical harmonic coefficients results by using the roots of the radial basis functions (i.e. the first-kind Bessel functions) for selecting the radial sample points. These samples are not uniformly spaced and their values depend on both the mode type, l, and azimuthal mode number, m, of the cylindrical harmonic coefficient being computed, that is to say that the computation of each cylindrical harmonic coefficient has its own unique set of radial sample points. The density of the samples points in the spatial domain is controlled by the desired spatial bandwidth. Just as with the DFT, a certain amount of over-sampling provides better accuracy. We define the sampling bandwidth, Bs, to be larger than the spatial bandwidth of the desired cylindrical harmonics, Bs > B.

The index ranges for the azimuthal angle and radial samples are controlled by this sampling bandwidth,

J=max{j|χ1j1BsRorχ2j1BsR}BsBMK=max{k|χ10kBsRorχ21kBsR}BsBN
The azimuthal angle is transformed using a standard DFT that uses uniform sampling. The sample points are given by,
ρlmk=χlmkBsandϕj=jπj
The composite index should include the full range of azimuthal angle samples for each radial coordinate sample as required for a DFT. Finally, the composite input sample index should include a sub-index that specifies whether the field component that is sampled. For the scalar basis decomposition, this sub-index specifies either the electric or magnetic field z-component. For the vector basis decomposition, this sub-index specifies either the ρ or ϕ component of the electric field. The complete composite index is thus given by,
{i}={i,j,k}={i=1,2}{JjJ1}{1kK}

To form a proper inverse, the output sample coordinates, that result from the inverse DFHT, should be the same as the input sample coordinates. However, this leads to distinct sets of sample coordinates for each cylindrical harmonic mode. Typically, one wants the total field, due to all modes, on some convenient grid of points. One could perform the inverse DFHT and then interpolate the fields from the different modes onto a single grid, or one could accomplish both operations together, by choosing a set of output sample coordinates that are mode independent, and in effect using the basis functions for the interpolation. Though any suitable grid can be used, here we show a Cartesian grid,

xα=αδandyγ=γδ
with grid indices, α and γ, and uniform grid spacing, δ. The corresponding cylindrical coordinates are,
ραγ=δα2+γ2andϕαγ=atan2(γ,α)
and the composite index with points selected to lie within the domain radius, R, is given by,
{α}={i,α,γ|ραγR}{i=1,2}{AαA}{AγA}
where, A=floor(Rδ).

An example of such composite indices are shown in Table 1 on page 9, with unit domain radius, spatial bandwidth of B = 2π, sampling bandwidth of Bs = 1.1B, and grid spacing of δ = 0.5. The representative form of the transfer matrix T for this example is shown in Eq. (46) in Appendix C. We also present a pseudo-algorithm for one of the composite indicies, l, in Appendix C.

Tables Icon

Table 1. Composite indices l, i, and α constructed for an example case: R = 1, B = 2π, Bs = 1.1B and δ = 0.5. Highlighted (grey colored rows) in l index indicate that the harmonics have space-bandwidth product less than or equal to RB. Highlighted (grey colored rows) in α index indicate that the spatial coordinate is within the domain radius R.

We can use these composite indices to form matrices for efficient computation of DFHT as;

a=diag(HE0)
the transfer function operation,
b=Ta
and the inverse DFHT,
E1=H¯b
The above operations can be composed together and written compactly as,
E1=H¯Tdiag(HE0)
Note that the diag operation returns a vector. The superscripts 0 and 1 on E denote input (incoming fields at z = z0) and output (out going fields at z = z1) respectively. The correspondence between matrix elements indexed by the original indices and the composite indices are as follows:
E0=Eil0=Eijklmn0b=bl=blmnH=Hli=HlmnijkT=Tll=TlmnlmnH¯=H¯αl=H¯iαγlmna=al=almnE1=Eα1=Eiαγ1

3.1. Scalar basis DFHT

For the case of the scalar basis decomposition, the matrix elements for: the input samples, the forward DFHT, the inverse DFHT, and the output samples are given by,

InputSamples&ForwardDFHTInverseDFHT&OutputSamplesEijklmn0=δil{Ez(ρ1mk,ϕj,z0)l=1ηHz(ρ2mk,ϕj,z0)l=2H¯iαγlmn=δilψlmn(ραγ,ϕαγ)Hlmnijk=δil2π2R2JBs2Clmk2ψlmn*(ρlmk,ϕj)Eiαγ1={Ez(ραγ,ϕαγ,z1)i=1ηHz(ραγ,ϕαγ,z1)i=2

3.2. Vector basis DFHT

For the case of the vector basis decomposition, the matrix elements for: the input samples, the forward DFHT, the inverse DFHT, and the output samples are given by,

InputSamples&ForwardDFHTInverseDFHT&OutputSamplesEijklmn0=Et(ρlmk,ϕj,z0)e^iH¯iαγlmn=Ψlmn(ραγ,ϕαγ)e^iHlmnijk=2π2R2JBs2Clmk2Ψlmn*(ρlmk,ϕj)e^iEiαγ1=Et(ραγ,ϕαγ,z1)e^i
where ê1 = ρ̂, and ê2 = ϕ̂. The fields and basis functions are assumed to be zero when the radius is greater than the domain radius, R. If the functions used to compute the matrix elements are defined otherwise, the radius should be tested and the element set to zero appropriately. This can be done for either the E0 or H0 elements. It is not required for both, since when computing the diagonal of the matrix product, the element product pairs always have the same radius.

One might assume, that since the l = 2 scalar basis functions are non-zero (but with zero radial derivative) at the the domain boundary, ρ = R, that one can accurately find cylindrical harmonic coefficients for Hz fields that are finite at ρ = R. In fact, the accuracy will be poor in this case. Only when fields are near zero at the boundary, can the DFHT provide accurate coefficients. The usual derivation of the discrete Hankel transform involves approximating a finite radial integration by an infinite one. This approximation is only accurate when the functions being transformed is zero outside the specified domain radius. Due to the intrinsic periodicity, an analogous problem exists for functions with non-zero amplitude at the boundary of a Fast Fourier Transform (FFT).

4. Cylindrical harmonic mode propagation (2.5D) & construction of transfer matrix T

The vector wave propagation of electromagnetic fields through an inhomogeneous medium is a well known problem and is often solved using finite difference time domain (FDTD) or finite element methods (FEM). However, this problem becomes computationally challenging when the size of this inhomogeneous structure is fairly large compared to wavelength. This forces one to exploit symmetries in the structure in order to reduce the computational load [21]. In this section we describe a technique to model axially symmetric structures using quasi-two-dimensional modeling method (2.5D modeling). This approach exploits the rotational symmetry of the structure (about the z axis) without imposing any restrictions on the source excitation field.

In section 2 and 3, we described the decomposition of arbitrary fields in terms of cylindrical harmonics. The TM and TE coefficients determined by this method can be used to completely describe the source excitation field. Similarly, the scattered fields due to the test object can again be decomposed into TM and TE cylindrical harmonic modes and their respective coefficients can be found. These scattered field coefficients form the transfer matrix T shown in Eq. (1). By combining the excitation and the scattered field coefficients appropriately with their respective basis functions one can determine the total fields due to the scattering test object. When the inhomogeneous scattering structure is rotationally symmetric independent), a cylindrical harmonic mode of a given azimuthal mode number, m, couples only to modes of the same mode number, m. In such a structure, the electric field in cylindrical coordinate varies with the azimuthal mode m as:

E(ρ,ϕ,z)=E˜(ρ,z)exp(jmϕ)
Using Eq. (34), the wave equation can be written in the form [22],
[jmρϕ^]×[1μ(jmρϕ^)×E˜]β2E˜=0
where, and μ are the relative electric permittivity and magnetic permeability second-order tensors of the medium. For each m, Eq. (35) can be solved on a two dimensional = 0) plane for any arbitrary excitation field provided the scattering medium under test is axially symmetric. For a given domain with maximum radius R having a maximum radial spatial bandwidth B, there exists approximately S number of modes, S ≈ 2MN. So, one needs to perform S number of independent 2D simulations (easily parallelizable), and hence the name 2.5 D. Also, there exists a parity relation between the propagating modes with positive and negative azimuthal mode numbers and this fact can be exploited to further reduce the number of 2D simulations by a factor of 2 (see Appendix B). Such an axi-symmetric medium can be solved numerically using finite element methods which are available in commercial software like COMSOL Multiphysics [22]. The scattered electric field formulation in COMSOL axi-symmetric solver (RF module) is used to specify a single cylindrical harmonic (with unit mode harmonic coefficient), whose the background input electric fields Ebg are given based on equations in section 2,
Elmnbg(ρ,z)=[ϒlmnΨlmn(ρ,0)+δl1ψ1mn(ρ,0)z^]exp(jβzlmn(zz0))
where ψlmn, ϒlmn, and Ψlmn are given from Eqs. (4), (13 and (14) respectively. The output of each such cylindrical harmonic 2D axi-symmetric simulation is a 1D function fl′lmn at z = z1, where z1 is output plane at which net response from an axi-symmetric scatterer is reconstructed. This 1D function at z = z1 can again be decomposed with 1D Hankel transforms (as the fields are independent of ϕ) to compute the output coefficients and inturn the transfer matrix T. The output coefficients and transfer matrix (for unity input harmonic coefficient) are given by,
Tlmnlmn=blmnlmn=0Rππψlmn*(ρ,ϕ)exp(jmϕ)fllmndϕρdρ=2πδmmClmn0RJm(βlmnρ)fllmnρdρ
where the function fl′lmn is given by the sum of background field and the calculated scattered field Ezsc or Hzsc,
fllmn(ρ)=δllψlmn(ρ,0)exp(jβzlmn(z1z0))+{Ezsc(ρ,z1)l=1ηHzsc(ρ,z1)l=2
Since the azimuthal mode number is preserved in the axi-symmetric simulation, in the above equation, m is used in the basis function instead of m′. The transfer function can also be given in terms of a discrete Hankel transform,
Tlmnlmn=k=1KHlmnmkFkllmnHlmnmk=δmm4π2R2Bs2ClmnClmk2Jm(βlmnρlmk)Fkllmn=fllmn(ρlmk)
The transfer matrix T, constructed using the scattered fields from all S simulations (negative azimuthal modes need not be simulated) can be written in a lexicographical order to form a sparse block diagonal matrix as shown in Fig. 1. In Fig. 1, T is a sparse, block-diagonal transfer matrix of size [S×S] constructed from the COMSOL 1D scattered fields. Each input mode generates a column vector in T matrix. It is important to note that in a linear axially symmetric scattering medium excited with a particular cylindrical harmonic mode with a specific mode type l, always preserves the azimuthal mode number m, but need not necessarily preserve the mode type (TM mode can partially or fully get converted to TE mode or vice versa). Hence the scattered fields have to be decomposed again into TM and TE modes for each background field mode type to construct a complete transfer matrix. There exist 2M number of blocks. This matrix T can be pre-computed for a given spatial bandwidth B and later reused for different excitations by appropriately changing the right hand side column vector a, i.e the input cylindrical harmonic coefficients.

 figure: Fig. 1

Fig. 1 Transfer Matrix T in its sparse, block-diagonal form. The m = 0th block is highlighted in magenta.

Download Full Size | PDF

The computational complexity of this numerical technique is mainly dependent on the total number of band-limited modes; S in a given space-bandwidth product; R × B. The total number of modes increases quadratically with space-bandwidth product and this is shown in the Fig. 2. The number of 2D cylindrical harmonic simulations and in turn the size of the transfer matrix T are directly dependent on total number of modes, S as shown in Fig. 1. An important advantage of this approach is that the 2D cylindrical harmonic simulations are independent of one another and can be completely parallelized. The computational complexity of this method can now be written as O(S).

 figure: Fig. 2

Fig. 2 Shows the total number of band-limited modes; S as a function of space-bandwidth product R × B along with a quadratic fit.

Download Full Size | PDF

5. Numerical validation

The following examples are considered to numerically verify the techniques described in the previous sections.

5.1. Gradient index hemispherical Maxwell’s fish-eye lens

Maxwell’s fish-eye lens is a radially symmetric GRIN lens whose refractive index profile n, is of the form, n(r)=21+(r/a)2, where, a is the radius of the lens and r=x2+y2+z2 is the radial coordinate. We simulate the hemispherical version of this lens excited by an offset Gaussian vector beam. For this example, we perform a detailed computational performance analysis between complete 3D full-wave and 2.5 D cylindrical harmonic propagation technique. We also compare the numerical results between the two methods. For this purpose, we chose a nominal lens radius, a = 4λ and an incident offset Gaussian full beam-waist of 2λ propagating along +z direction. (An offset Gaussian beam ensures a non-symmetric excitation field with harmonic components that have azimuthal mode numbers m greater than 1). We first simulate the blue region of the lens (one half of the 2D slice of the lens named as 2.5 D COMSOL Simulation Domain) as shown in Fig. 3 and construct the transfer matrix T based on all possible propagating modes inside this simulation domain (R=8λ). The input cylindrical harmonic mode coefficients of the source field and the transfer matrix are used to generate output plane mode coefficients. These output modes are then expanded (and also independently further propagated) using their respective basis functions to calculate the full wave vector fields post the lens. This procedure is as explained in section 4. The electric field norm on the xz plane (y = 0) is shown in Fig. 4(a). The internal fields of the lens for the 2.5D case are not shown as the transfer matrix was constructed for a plane just outside of the lens (z = 4λ) and the resulting output modes were propagated. We also simulated the full-wave complete 3D structure in COMSOL. These results are shown in Fig. 4(b). We plot the transverse plane electric field norm (xy plane) at the focal plane of the lens and at the source plane in Fig. 4(c) and (d) respectively. At the focal plane, we subtract the results from cylindrical harmonic propagator technique (2.5D) and 3D full wave simulations and show that the results are indeed accurate within a maximum error < 1 %. The normalized absolute error map at the focal plane of the lens is shown in Fig. 4(e).

 figure: Fig. 3

Fig. 3 Hemispherical Maxwell’s fish-eye gradient index lens schematic for 2.5 D technique. Blue outline box indicates the simulated domain , z, ϕ = 0).

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Shows the comparison between 2.5D technique and complete 3D simulations. (a). Electric field norm on xz plane (y = 0) using 2.5 D technique (Only fields pre and post the scattering medium are computed using the transfer matrix. Hence the fields inside the lens in (a) are blank). (b). Electric field norm on xz plane (y = 0) computed using conventional 3D approach in COMSOL. (c). Electric field norm at the focal plane of the lens. The white dashed line indicates the focal line. (d). Offset Gaussian beam with a beam waist=2λ propagating along +z direction is the source excitation field in both 2.5D and 3D simulations. (e). Normalized absolute error between 2.5D technique and 3D simulations at the focal plane. Electric fields are in the units of V/m.

Download Full Size | PDF

5.1.1. 3D versus 2.5D based transfer matrix

To compare a complete 3D full-wave simulation with a cylindrical harmonic propagator, using the 2.5D method and a transfer matrix approach, an example with a specific space-bandwidth product was analyzed. Performance analysis comparison between the two methods is presented in Table 2 on page 15. The cylindrical harmonic technique has superior performance in terms of computational burden, simulation time, memory usage, computational scaling, result re-usability, and parallelizability.

Tables Icon

Table 2. Comparison of Computational Performance between complete 3D simulation and Cylindrical Harmonic Mode Propagation technique (2.5D technique) for a domain radius R = 8λ.

5.1.2. Plane wave versus cylindrical harmonic basis

As stressed in previous sections, construction of a reusable transfer matrix for a given axi-symmetric medium is one of the main goals of this article. The transfer matrix, T, for a given medium can be constructed using any complete basis, including plane waves, which comprise the most commonly used basis. Since plane waves cannot be simply represented in cylindrical coordinates, one is at first tempted to use a full-wave 3D solver (rather than an axi-symmetric solver). Such an approach is indeed more straightforward, but does not fully exploit the axial symmetry of the mediums of interest in this article. One can employ one mirror plane symmetry but the solution domains will remain three dimensional, with the associate unfavorable size scaling. Somewhat surprisingly, it is more advantageous to expand a plane wave in a cylindrical-harmonic basis and use the axi-symmetric solver as we now describe. The z-components of a plane wave incident at an arbitrary angle, θi (angle between the wave vector and ), can be expanded in cylindrical coordinates using cylindrical harmonics as [19,21],

Ez(ρ,ϕ,z0)orη0Hz(ρ,ϕ,z0)=E0sinθiexp(jβ0z0cosθi)m=jmJm(β0ρsinθi)exp(jmϕ)
where, β0=ω0μ0 is the free-space wave number. Even though the above equation is an infinite sum of azimuthal mode numbers, for a given finite space-bandwidth product (and a reasonable tolerance), the above equation converges quickly for a finite set of azimuthal mode numbers.

In order to construct the transfer matrix T for an axi-symmetric medium based on plane wave propagator, each cylindrical harmonic that contributes to a plane wave (every plane wave has a discrete value of θi in the range 0 to π/2) has to be propagated using an axi-symmetric solver. The 2D scattered fields due to a plane wave at z = z1 plane is calculated by revolving the 1D fields in = 0, z = z1) cut-line (i.e multiplying by the factor ejmϕ, with ϕ discretely varying from 0 to 2π) and summing all the 2D field contributions from each cylindrical harmonic. These summed 2D scattered fields due to a single plane wave input need to be regridded and interpolated to transform from , ϕ) cylindrical coordinates to (x,y) Cartesian coordinates. Then, 2D FFT can be applied on these fields to calculate the output harmonic coefficients for a particular input harmonic. (Parity conditions exist and can be exploited for −ve azimuthal mode numbers). Once the output coefficients are known, inverse FFT can be applied to calculate the vector field results post the medium. In this approach as one can very evidently note, there is a back and forth change of basis. Hence, this approach to construct T may not be the best choice in terms of computational load for a given space-bandwidth product inspite of superior, well known FFT algorithms. Table 3 illustrates the case for cylindrical harmonic propagator by comparing the computational load required to construct T using the two methods for a specific (Maxwell’s Fish-eye with R = 8λ) example.

Tables Icon

Table 3. Comparison of computational load between Plane wave propagator and Cylindrical harmonic propagator using the same 2D axi-symmetric solver (2.5D technique) for a domain radius R = 8λ.

5.2. Spherical invisibility cloak

A spherical invisibility cloak is a transformation optics designed media whose material properties are anisotropic but yet radially symmetric. This example was chosen to show the versatility of cylindrical harmonic propagator technique in exploiting axial symmetry and in dealing with such sophisticated medium. The second-order electric permittivity tensor, (or the magnetic permeability, μ) for a spherical invisibility cloak in spherical coordinates (r, θ, ϕ) is given by [23],

=μ=[rrrθrϕθrθθθϕϕrϕθϕϕ]=[bba(ra)2r2000bba000bba]
where a is the radius of the sphere to be hidden and the cloaking region is contained in the annulus a < r < b. In cylindrical coordinates, , ϕ, z), the tensor in Eq. (41) transforms to,
[ρρρϕρzϕρϕϕϕzzρzϕzz]=[12(rr+θθ+(θθrr)cos2θ)0(rrθθ)cosθsinθ0ϕϕ0(rrθθ)cosθsinθ0rrcos2θ+θθsin2θ]
The 2.5D simulation setup for a spherical cloak is shown in Fig. 5. The simulation domain just consists of one half of a 2D slice of the cloak. The inner cloaked region has a radius a = 2λ and is set to perfect electric boundary condition (PEC). The cloaking region annulus is between the radii a = 2λ and b = 4λ. The material property tensor components used in the cloaking annulus is shown in sub-figures. 6. The effect of cloaking is shown in Fig. 6. The incident beam is an offset y-polarized Gaussian vector beam propagating along +z direction. The electric fields shown in Fig. 6 correspond to the real part of y-component on xz-plane (y = 0).

 figure: Fig. 5

Fig. 5 Spherical invisibility cloak schematic for 2.5 D technique. Blue outline box indicates the simulated domain.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Shows the real part of the field (in V/m) due to incident offset Gaussian vector beam on xz plane (y = 0) inside and outside the spherical cloak simulated using 2.5D technique. The subfigures below show the individual component of the material property tensors required inside the cloaking region.

Download Full Size | PDF

5.3. Free space impedance matched thin aspheric plano-convex lens

In this case, we consider a simple aspheric, thin plano-convex lens with a focal length, f = 15λ, refractive index of n = 3.58 and relative impedance Z = 1 (∊r = μr = n) as the homogeneous linear scattering medium. The lens is axially symmetric with a maximum thickness of 0.8λ and a diameter of 18λ. The lens is illuminated normally by a ŷ-polarized Gaussian beam of waist diameter = 5λ at the source plane (40λ behind the lens). This example was chosen to demonstrate the versatility of this method when the transverse size and range are large compared to λ. Figure 7 shows the complete schematic diagram of the setup. The blue colored rectangular box in the figure shows the COMSOL 2.5 D simulation domain along with the boundary conditions that were applied. The simulation domain has a transverse radius Rdomain = 30λ. Figure 8(a) shows the numerically computed electric field norm pre- and post-lens. The input Gaussian beam is focused at z = 15λ in front of the lens as per the design. Since a reflection-less lens in general converts a normally incident input Gaussian beam to another output Gaussian beam, one can derive the complete analytical solution for this setup using Gaussian beam optics theory [24]. Using the parameters of the considered lens and the excitation Gaussian beam, the waist radius of the output beam as a function of z was analytically calculated. Figure 8(b) shows the comparison between the numerical technique and the exact analytical result. The numerical and the analytical output beam radius are in good agreement with each other within an RMS error of 3%.

 figure: Fig. 7

Fig. 7 Aspheric thin plano-convex lens schematic for 2.5 D technique. Blue outline box indicates the simulated domain.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 (a). Numerically computed fields norm of electric field pre- and post-lens (in V/m). (b). Comparison between the output Gaussian beam radius as a function of z calculated numerically (black solid) and analytically (dashed red).

Download Full Size | PDF

5.4. Negative refractive index perfect lens

In this example, we choose a perfect negative index lens excited by a completely arbitrary source excitation field. A perfect lens has a straightforward analytical relationship between the image and source fields. A perfect lens (∊r = μr = n = −1) acts as a focusing slab in which the image is the exact reproduction of the source (when there is no loss) [25]. The perfect lens disk has a thickness z2 = 3.6λ and a radius Rdomain = 30λ. The distances from the edge of the lens slab to the object and image plane are z1 = 0.45λ and z3 = 3.15λ respectively, such that z2 = z1 + z3. Figure 9 shows the schematic diagram of the setup along with the COMSOL simulation domain and the boundary conditions used.

 figure: Fig. 9

Fig. 9 Perfect Lens schematic for 2.5D technique. Blue outline box shows the simulated domain.

Download Full Size | PDF

Figure 10(a) shows the source plane electric field norm of the excitation field. Since the image is the object for a perfect lens with no loss, the fields at the image plane should be exactly equal to the source fields. Figure 10(b) shows the difference between the image and source fields and the absolute peak error about 1.5%.

 figure: Fig. 10

Fig. 10 (a) Shows the source plane electric field norm (in V/m) of the chosen arbitrary excitation source at z = 0 plane. (b) Shows the difference map between the image and the object plane electric field norms (in V/m).

Download Full Size | PDF

6. Conclusion

In this article we develop a complete formulation for both continuous and discrete cylindrical harmonic decomposition. We formulate this decomposition method for both scalar and vector basis fields. Such harmonic decomposition methods (like Fourier transforms and spherical harmonics) play a vital role in many computational physics problems. Some future developments regarding the decomposition will include reducing the computational burden further and parallelizing the decomposition technique on a Graphical Processing Unit (GPU). We use this decomposition method to propagate the individual harmonics through an axi-symmetric medium using a 2D axi-symmetric EM solver as a part of a pre-computational step (2.5D technique). The fields from this solver are inturn used to construct an efficient, modular, and re-useable transfer matrix, which can be later used to calculate the total scattered fields due to the inhomogeneous axi-symmetric medium for any given arbitrary incident excitation and a finite space-bandwidth product. We also present a composite index ordering strategy to perform the required forward and/or inverse DFHT. This ordering helps to reduce the operations to simple matrix multiplications. This overall approach was numerically verified on well-known and relevant axi-symmetric media. We also compare our method with other possible techniques in terms of computational load. The exploitation of axial symmetry along with transfer matrix approach using CHD help to significantly reduce the computational burden/resources without sacrificing accuracy.

A. Orthogonality Properties of Bessel and Exponential Functions

The following orthogonality properties of the Bessel and exponential functions are used in the article,

0RJm(χ1mnρ)Jm(χ1mnρ)ρdρ=R22Jm+12(χ1mn)δnn
0RJm(χ2mnρ)Jm(χ2mnρ)ρdρ=R22[Jm2(χ2mn)Jm1(χ2mn)Jm+1(χ2mn)]δnn
ππexp(jmϕ)exp(jmϕ)dϕ=2πδmm

B. Parity Condition between positive and negative azimuthal mode numbers

The following Tables 4 and 5 show the parity condition between positive and negative azimuthal mode numbered basis functions for TM and TE modes.

Tables Icon

Table 4. Parity Condition for TM modes.

Tables Icon

Table 5. Parity Condition for TE modes.

C. Composite Indices for DFHT

Based on the example composite indices shown in Table 1 on page 9, the transfer matrix T takes the representative form shown in Eq. 46, where, ⋈ and # in T matrix indicates the TM (l′ = 1) and TE (l′ = 2) mode coefficients respectively. The rest of the values in the 17 × 17 transfer matrix are zeros, making it a sparse, block-diagonal matrix.

The functional form of the composite indices [l, i, α] mapping are neither illuminating nor easy to specify in mathematical notation. It is best to generate a vector of triplets that store the mapping and use elements of this vector for building up the needed matrices. Some pseudo-code for generating this composite index vector l is shown in Algorithm 1.

[b]=[###############################]17×17[a2,4,1a2,3,1a1,2,1a2,2,1a1,1,1a2,1,1a2,1,2a1,0,1a1,0,2a2,0,1a2,0,2a1,1,1a2,1,1a2,1,2a1,2,1a2,2,1a2,3,1]17×1

Tables Icon

Algorithm 1. Finding Composite Index l

Funding

This work was supported by the Department of Homeland Security, Science and Technology Directorate (Contract No. HSHQDC-12-C-00049). The published material represents the position of the authors and not necessarily that of the DHS or S&T.

Acknowledgments

The support and resources from the Center for High Performance Computing at the University of Utah are gratefully acknowledged. Authors would like to acknowledge Richard Edwards for some useful comments on the manuscript.

References and links

1. O. Furxhi, D. L. Marks, and D. J. Brady, “Echelle crossed grating millimeter wave beam scanner,” Opt. Express 22, 16393–16407 (2014). [CrossRef]   [PubMed]  

2. O. S. Cossairt, D. Miau, and S. K. Nayar, “Scaling law for computational imaging using spherical optics,” J. Opt. Soc. Am. A 28, 2540–2553 (2011). [CrossRef]  

3. S. Venkatesh, N. Viswanathan, and D. Schurig, “W-band sparse synthetic aperture for computational imaging,” Opt. Express 24, 8317–8331 (2016). [CrossRef]   [PubMed]  

4. M. Meeks and J. Ruze, “Evaluation of the haystack antenna and radome,” Antennas and Propagation, IEEE Transactions on 19, 723–728 (1971). [CrossRef]  

5. G. Crone, A. Rudge, and G. Taylor, “Design and performance of airborne radomes: A review,” Communications, Radar and Signal Processing, IEE Proceedings F 128, 451–464 (1981). [CrossRef]  

6. M. Ibanescu, Y. Fink, S. Fan, E. L. Thomas, and J. D. Joannopoulos, “An All-Dielectric coaxial waveguide,” Science 289, 415–419 (2000). [CrossRef]   [PubMed]  

7. B. Thomas, G. L. James, and K. J. Greene, “Design of wide-band corrugated conical horns for cassegrain antennas,” Antennas and Propagation, IEEE Transactions on 34, 750–757 (1986). [CrossRef]  

8. G. Lipworth, J. Ensworth, K. Seetharam, D. Huang, J. S. Lee, P. Schmalenberg, T. Nomura, M. S. Reynolds, D. R. Smith, and Y. Urzhumov, “Magnetic Metamaterial Superlens for Increased Range Wireless Power Transfer,” Scientific Reports 4, 3642 (2014). [CrossRef]   [PubMed]  

9. R. Merlin, “Radiationless electromagnetic interference: Evanescent-Field lenses and perfect focusing,” Science 317, 927–929 (2007). [CrossRef]   [PubMed]  

10. Y. Ye, Z. J. Wong, X. Lu, X. Ni, H. Zhu, X. Chen, Y. Wang, and X. Zhang, “Monolayer excitonic laser,” Nat Photon 9, 733 (2015). [CrossRef]  

11. Y. Zhao, J. Wyrick, F. D. Natterer, J. F. Rodriguez-Nieva, C. Lewandowski, K. Watanabe, T. Taniguchi, L. S. Levitov, N. B. Zhitenev, and J. A. Stroscio, “Creating and probing electron whispering-gallery modes in graphene,” Science 348, 672–675 (2015). [CrossRef]   [PubMed]  

12. J. B. Pendry, A. Aubry, D. R. Smith, and S. A. Maier, “Transformation optics and subwavelength control of light,” Science 337, 549–552 (2012). [CrossRef]   [PubMed]  

13. C. Ciraci, R. T. Hill, J. J. Mock, Y. Urzhumov, A. I. Fernández-Dominguez, S. A. Maier, J. B. Pendry, A. Chilkoti, and D. R. Smith, “Probing the Ultimate Limits of Plasmonic Enhancement,” Science 337, 1072–1074 (2012). [CrossRef]   [PubMed]  

14. Y. Ou, D. Pardo, and Y. Chen, “Fourier finite element modeling of light emission in waveguides: 2.5-dimensional fem approach,” Opt. Express 23, 30259–30269 (2015). [CrossRef]   [PubMed]  

15. C. C. Johnson, Field and Wave Electrodynamics (McGraw-Hill, 1965).

16. A. F. Peterson, S. L. Ray, and R. Mittra, Computational Methods for Electromagnetics, Volume 4 of IEEE Press Series on Electromagnetic Wave Theory (Wiley, 1998).

17. R. Song, J. Zhu, and X. Zhang, “Full-vectorial modal analysis for circular optical waveguides based on the multidomain chebyshev pseudospectral method,” J. Opt. Soc. Am. B 27, 1722–1730 (2010). [CrossRef]  

18. S. Feng, K. Halterman, P. L. Overfelt, and D. Bowling, “Cyclic sommerfeld resonances in nanorods at grazing incidences,” Opt. Express 17, 19823–19841 (2009). [CrossRef]   [PubMed]  

19. C. A. Balanis, Advanced Engineering Electromagnetics, 2nd Edition (Wiley, 2012).

20. S. Ali, W. Chew, and J. Kong, “Vector hankel transform analysis of annular-ring microstrip antenna,” IEEE Transactions on Antennas and Propagation 30, 637–644 (1982). [CrossRef]  

21. C. Ciraci, Y. Urzhumov, and D. R. Smith, “Far-field analysis of axially symmetric three-dimensional directional cloaks,” Opt. Express 21, 9397–9406 (2013). [CrossRef]   [PubMed]  

22. COMSOL Multiphysics, RF Module User’s Guide (COMSOL, 2012).

23. J. B. Pendry, D. Schurig, and D. R. Smith, “Controlling electromagnetic fields,” Science 312, 1780–1782 (2006). [CrossRef]   [PubMed]  

24. B. E. A. Saleh and M. C. Teich, Fundamentals of Photonics (Wiley, 2007).

25. J. B. Pendry, “Negative Refraction Makes a Perfect Lens,” Physical Review Letters 85, 3966–3969 (2000). [CrossRef]   [PubMed]  

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 Transfer Matrix T in its sparse, block-diagonal form. The m = 0th block is highlighted in magenta.
Fig. 2
Fig. 2 Shows the total number of band-limited modes; S as a function of space-bandwidth product R × B along with a quadratic fit.
Fig. 3
Fig. 3 Hemispherical Maxwell’s fish-eye gradient index lens schematic for 2.5 D technique. Blue outline box indicates the simulated domain , z, ϕ = 0).
Fig. 4
Fig. 4 Shows the comparison between 2.5D technique and complete 3D simulations. (a). Electric field norm on xz plane (y = 0) using 2.5 D technique (Only fields pre and post the scattering medium are computed using the transfer matrix. Hence the fields inside the lens in (a) are blank). (b). Electric field norm on xz plane (y = 0) computed using conventional 3D approach in COMSOL. (c). Electric field norm at the focal plane of the lens. The white dashed line indicates the focal line. (d). Offset Gaussian beam with a beam waist=2λ propagating along +z direction is the source excitation field in both 2.5D and 3D simulations. (e). Normalized absolute error between 2.5D technique and 3D simulations at the focal plane. Electric fields are in the units of V/m.
Fig. 5
Fig. 5 Spherical invisibility cloak schematic for 2.5 D technique. Blue outline box indicates the simulated domain.
Fig. 6
Fig. 6 Shows the real part of the field (in V/m) due to incident offset Gaussian vector beam on xz plane (y = 0) inside and outside the spherical cloak simulated using 2.5D technique. The subfigures below show the individual component of the material property tensors required inside the cloaking region.
Fig. 7
Fig. 7 Aspheric thin plano-convex lens schematic for 2.5 D technique. Blue outline box indicates the simulated domain.
Fig. 8
Fig. 8 (a). Numerically computed fields norm of electric field pre- and post-lens (in V/m). (b). Comparison between the output Gaussian beam radius as a function of z calculated numerically (black solid) and analytically (dashed red).
Fig. 9
Fig. 9 Perfect Lens schematic for 2.5D technique. Blue outline box shows the simulated domain.
Fig. 10
Fig. 10 (a) Shows the source plane electric field norm (in V/m) of the chosen arbitrary excitation source at z = 0 plane. (b) Shows the difference map between the image and the object plane electric field norms (in V/m).

Tables (6)

Tables Icon

Table 1 Composite indices l, i, and α constructed for an example case: R = 1, B = 2π, Bs = 1.1B and δ = 0.5. Highlighted (grey colored rows) in l index indicate that the harmonics have space-bandwidth product less than or equal to RB. Highlighted (grey colored rows) in α index indicate that the spatial coordinate is within the domain radius R.

Tables Icon

Table 2 Comparison of Computational Performance between complete 3D simulation and Cylindrical Harmonic Mode Propagation technique (2.5D technique) for a domain radius R = 8λ.

Tables Icon

Table 3 Comparison of computational load between Plane wave propagator and Cylindrical harmonic propagator using the same 2D axi-symmetric solver (2.5D technique) for a domain radius R = 8λ.

Tables Icon

Table 4 Parity Condition for TM modes.

Tables Icon

Table 5 Parity Condition for TE modes.

Tables Icon

Algorithm 1 Finding Composite Index l

Equations (50)

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

E t ( ρ , ϕ , z 0 ) E t ( ρ , ϕ , z 1 ) or a l m n T b l m n or E z ( ρ , ϕ , z 0 ) , H z ( ρ , ϕ , z 0 ) E z ( ρ , ϕ , z 1 ) , H z ( ρ , ϕ , z 1 )
E z ( ρ , ϕ , z 0 ) = m , n a 1 m n ψ 1 m n ( ρ , ϕ )
η 0 H z ( ρ , ϕ , z 0 ) = m , n a 2 m n ψ 2 m n ( ρ , ϕ )
E z ( ρ , ϕ , z 1 ) = m , n b 1 m n ψ 1 m n ( ρ , ϕ )
η 0 H z ( ρ , ϕ , z 1 ) = m , n b 2 m n ψ 2 m n ( ρ , ϕ )
ψ l m n ( ρ , ϕ ) = C l m n J m ( β l m n ρ ) exp ( j m ϕ )
J m ( β 1 m n R ) = 0 and J m ( β 2 m n R ) = 0
J m ( χ 1 m n ) = 0 and J m ( χ 2 m n ) = 0
C 1 m n 1 = π R J m + 1 ( χ 1 m n )
C 2 m n 1 = π R [ J m 2 ( χ 2 m n ) J m 1 ( χ 2 m n ) J m + 1 ( χ 2 m n ) ] 1 / 2
ψ l m n | ψ l m n = 0 R π π ψ l m n * ( ρ , ϕ ) ψ l m n ( ρ , ϕ ) d ϕ ρ d ρ = δ m m δ n n
a l m n = ψ l m n | E z l = 1 η 0 H z l = 2 = 0 R π π ψ l m n * ( ρ , ϕ ) { E z ( ρ , ϕ , z 0 ) l = 1 η 0 H z ( ρ , ϕ , z 0 ) l = 2 } d ϕ ρ d ρ
E t = E E z z ^
E t ( ρ , ϕ , z 0 ) = l , m , n a l m n ϒ l m n Ψ l m n ( ρ , ϕ )
E t ( ρ , ϕ , z 1 ) = l , m , n b l m n ϒ l m n Ψ l m n ( ρ , ϕ )
ϒ l m n = j ω 2 0 μ 0 β l m n 2 δ l 1
Ψ 1 m n ( ρ , ϕ ) = C 1 m n [ J m ( β 1 m n ρ ) ρ ^ + j m J m ( β 1 m n ρ ) β 1 m n ρ ϕ ^ ] exp ( j m ϕ )
Ψ 2 m n ( ρ , ϕ ) = C 2 m n [ j m J m ( β 2 m n ρ ) β 2 m n ρ ρ ^ + J m ( β 2 m n ρ ) ϕ ^ ] exp ( j m ϕ )
Ψ l m n | Ψ l m n = 0 R π π Ψ l m n * ( ρ , ϕ ) Ψ l m n ( ρ , ϕ ) d ϕ ρ d ρ = δ l l δ m m δ n n
a l m n ϒ l m n = Ψ l m n | E t = 0 R π π Ψ l m n * ( ρ , ϕ ) E t ( ρ , ϕ , z 0 ) d ϕ ρ d ρ
b l m n = l , m , n T l m n l m n a l m n
T l m n l m n = δ l l δ m m δ n n exp ( j β z l m n ( z 1 z 0 ) ) where , β z l m n = ω 2 μ β l m n 2
M = max { m | χ 1 m 1 B R or χ 2 m 1 BR } N = max { n | χ 10 n B R or χ 21 n BR }
{ l } = { l , m , n | χ l m n BR } { l = 1 , 2 } { M m M 1 } { 1 n N }
J = max { j | χ 1 j 1 B s R or χ 2 j 1 B s R } B s B M K = max { k | χ 10 k B s R or χ 21 k B s R } B s B N
ρ l m k = χ l m k B s and ϕ j = j π j
{ i } = { i , j , k } = { i = 1 , 2 } { J j J 1 } { 1 k K }
x α = α δ and y γ = γ δ
ρ α γ = δ α 2 + γ 2 and ϕ α γ = atan 2 ( γ , α )
{ α } = { i , α , γ | ρ α γ R } { i = 1 , 2 } { A α A } { A γ A }
a = diag ( HE 0 )
b = Ta
E 1 = H ¯ b
E 1 = H ¯ T diag ( HE 0 )
E 0 = E il 0 = E i j k l m n 0 b = b l = b l m n H = H li = H l m n i j k T = T l l = T l m n l m n H ¯ = H ¯ α l = H ¯ i α γ l m n a = a l = a l m n E 1 = E α 1 = E i α γ 1
Input Samples & Forward DFHT Inverse DFHT & Output Samples E i j k l m n 0 = δ i l { E z ( ρ 1 m k , ϕ j , z 0 ) l = 1 η H z ( ρ 2 m k , ϕ j , z 0 ) l = 2 H ¯ i α γ l m n = δ i l ψ l m n ( ρ α γ , ϕ α γ ) H l m n i j k = δ i l 2 π 2 R 2 J B s 2 C l m k 2 ψ l m n * ( ρ l m k , ϕ j ) E i α γ 1 = { E z ( ρ α γ , ϕ α γ , z 1 ) i = 1 η H z ( ρ α γ , ϕ α γ , z 1 ) i = 2
Input Samples & Forward DFHT Inverse DFHT & Output Samples E i j k l m n 0 = E t ( ρ l m k , ϕ j , z 0 ) e ^ i H ¯ i α γ l m n = Ψ l m n ( ρ α γ , ϕ α γ ) e ^ i H l m n i j k = 2 π 2 R 2 J B s 2 C l m k 2 Ψ l m n * ( ρ l m k , ϕ j ) e ^ i E i α γ 1 = E t ( ρ α γ , ϕ α γ , z 1 ) e ^ i
E ( ρ , ϕ , z ) = E ˜ ( ρ , z ) exp ( j m ϕ )
[ j m ρ ϕ ^ ] × [ 1 μ ( j m ρ ϕ ^ ) × E ˜ ] β 2 E ˜ = 0
E l m n bg ( ρ , z ) = [ ϒ l m n Ψ l m n ( ρ , 0 ) + δ l 1 ψ 1 m n ( ρ , 0 ) z ^ ] exp ( j β z l m n ( z z 0 ) )
T l m n l m n = b l m n l m n = 0 R π π ψ l m n * ( ρ , ϕ ) exp ( j m ϕ ) f l l m n d ϕ ρ d ρ = 2 π δ m m C l m n 0 R J m ( β l m n ρ ) f l l m n ρ d ρ
f l l m n ( ρ ) = δ l l ψ l m n ( ρ , 0 ) exp ( j β z l m n ( z 1 z 0 ) ) + { E z sc ( ρ , z 1 ) l = 1 η H z sc ( ρ , z 1 ) l = 2
T l m n l m n = k = 1 K H l m n m k F k l l m n H l m n m k = δ m m 4 π 2 R 2 B s 2 C l m n C l m k 2 J m ( β l m n ρ l m k ) F k l l m n = f l l m n ( ρ l m k )
E z ( ρ , ϕ , z 0 ) or η 0 H z ( ρ , ϕ , z 0 ) = E 0 sin θ i exp ( j β 0 z 0 cos θ i ) m = j m J m ( β 0 ρ sin θ i ) exp ( j m ϕ )
= μ = [ r r r θ r ϕ θ r θ θ θ ϕ ϕ r ϕ θ ϕ ϕ ] = [ b b a ( r a ) 2 r 2 0 0 0 b b a 0 0 0 b b a ]
[ ρ ρ ρ ϕ ρ z ϕ ρ ϕ ϕ ϕ z z ρ z ϕ z z ] = [ 1 2 ( r r + θ θ + ( θ θ r r ) cos 2 θ ) 0 ( r r θ θ ) cos θ sin θ 0 ϕ ϕ 0 ( r r θ θ ) cos θ sin θ 0 r r cos 2 θ + θ θ sin 2 θ ]
0 R J m ( χ 1 m n ρ ) J m ( χ 1 m n ρ ) ρ d ρ = R 2 2 J m + 1 2 ( χ 1 m n ) δ n n
0 R J m ( χ 2 m n ρ ) J m ( χ 2 m n ρ ) ρ d ρ = R 2 2 [ J m 2 ( χ 2 m n ) J m 1 ( χ 2 m n ) J m + 1 ( χ 2 m n ) ] δ n n
π π exp ( j m ϕ ) exp ( j m ϕ ) d ϕ = 2 π δ m m
[ b ] = [ # # # # # # # # # # # # # # # # # # # # # # # # # # # # # # # ] 17 × 17 [ a 2 , 4 , 1 a 2 , 3 , 1 a 1 , 2 , 1 a 2 , 2 , 1 a 1 , 1 , 1 a 2 , 1 , 1 a 2 , 1 , 2 a 1 , 0 , 1 a 1 , 0 , 2 a 2 , 0 , 1 a 2 , 0 , 2 a 1 , 1 , 1 a 2 , 1 , 1 a 2 , 1 , 2 a 1 , 2 , 1 a 2 , 2 , 1 a 2 , 3 , 1 ] 17 × 1
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.