Difference between revisions of "A Review of Maxwell's Equations & Computational Electromagnetics (CEM)"

From Emagtech Wiki
Jump to: navigation, search
 
(21 intermediate revisions by the same user not shown)
Line 1: Line 1:
[[Image:Maxwell1.png|right|600px]]
+
[[Image:Maxwell1.png|right|720px]]
 
<table>
 
<table>
 
<tr>
 
<tr>
<td>[[image:Cube-icon.png | link=Getting_Started_with_EM.CUBE]] [[image:cad-ico.png | link=CubeCAD]] [[image:fdtd-ico.png | link=EM.Tempo]] [[image:prop-ico.png | link=EM.Terrano]] [[image:static-ico.png | link=EM.Ferma]] [[image:planar-ico.png | link=EM.Picasso]] [[image:metal-ico.png | link=EM.Libera]] [[image:po-ico.png | link=EM.Illumina]] </td>
+
<td>[[image:Cube-icon.png | link=Getting_Started_with_EM.Cube]] [[image:cad-ico.png | link=Building Geometrical Constructions in CubeCAD]] [[image:fdtd-ico.png | link=EM.Tempo]] [[image:prop-ico.png | link=EM.Terrano]] [[image:static-ico.png | link=EM.Ferma]] [[image:planar-ico.png | link=EM.Picasso]] [[image:metal-ico.png | link=EM.Libera]] [[image:po-ico.png | link=EM.Illumina]] </td>
 
<tr>
 
<tr>
 
</table>
 
</table>
[[Image:Back_icon.png|40px]] '''[[EM.Cube | Back to EM.Cube Main Page]]'''
 
== Maxwell's Equations in Differential Form ==
 
  
Maxwell's equations form the basis for the mathematical formulation of almost all electromagnetic modeling problems. The differential form of Maxwell's equations relates the electric and magnetic fields and sources locally at every point in the space. In an isotropic, time-invariant and homogeneous medium, they are given by:
 
  
:<math> \nabla . \mathbf{D} = \rho </math>
+
The following links provide a basic review of electromagnetic theory and the various numerical techniques used by [[EM.Cube]]'s simulation engines:
  
:<math> \nabla . \mathbf{B} = 0 </math>
+
[[Basic Electromagnetic Theory]]
  
:<math> \nabla \times \mathbf{E} = - \dfrac{\partial \mathbf{B}}{\partial t} </math>
+
[[The Far-Field Approximation for Radiation & Scattering Problems]]
  
:<math> \nabla \times \mathbf{H} = \dfrac{\partial \mathbf{D}}{\partial t} + \mathbf{J} </math>
+
[[Basic Principles of The Finite Difference Time Domain Method]]
  
where '''&nabla;''' is the gradient operator:
+
[[Basic Principles of SBR Ray Tracing]]
  
:<math> \nabla = \dfrac{\partial}{\partial x}\hat{\mathbf{x}}  + \dfrac{\partial}{\partial y}\hat{\mathbf{y}}  + \dfrac{\partial}{\partial z}\hat{\mathbf{z}}  </math>
+
[[Basic Principles of The Method of Moments]]
  
'''&nabla;.''' denotes the divergence operation, '''&nabla;'''&times; denotes the curl operation, '''E''' and '''H''' are the electric and magnetic fields in V/m and A/m, respectively, '''D''' and '''B''' are the electric and magnetic flux densities in C/m<sup>2</sup> and Wb, respectively, '''J''' is the electric volume current density in A/m<sup>2</sup>, &rho; is the electric volume charge density in C//m<sup>3</sup>, and the following constitutive relationships hold:
+
[[Basic Principles of Physical Optics]]
  
:<math> \mathbf{D} = \epsilon \mathbf{E}, \quad \quad \mathbf{B} = \mu \mathbf{H}, \quad \quad \mathbf{J} = \sigma \mathbf{E} </math>
+
[[Electrostatic_%26_Magnetostatic_Field_Analysis | Basic Principles of Electrostatic & Magnetostatic Field Analysis]]
  
where &epsilon; is the permittivity in F/m, &mu; is the permeability in H/m, and &sigma; is the electric conductivity of the medium in S/m.
+
[[Steady-State_Thermal_Analysis | Basic Principles of Steady-State Thermal Analysis]]
  
Although real magnetic charges and currents do not exist in nature, using the electromagnetic equivalence theorem, it is convenient to introduce a magnetic volume charge density in C/m<sup>3</sup>, and a magnetic volume current density '''M''' in V/m<sup>2</sup> to preserve the symmetry and duality of Maxwell's equations in the following form:
 
  
:<math> \nabla . \mathbf{D} = \rho </math>
+
{{Note|The above pages contain a large number of mathematical expressions and equations. Sometimes, the mathematical notations might not load properly the first time you open a page. In that case, you need to use the "reload" or "update" feature of your web browser to refresh the page, or return to the gateway page and try to open the link one more time.}}  
 
+
:<math> \nabla . \mathbf{B} = \rho_m  </math>
+
 
+
:<math> \nabla \times \mathbf{E} = - \dfrac{\partial \mathbf{B}}{\partial t} - \mathbf{M} </math>
+
 
+
:<math> \nabla \times \mathbf{H} = \dfrac{\partial \mathbf{D}}{\partial t} + \mathbf{J} </math>
+
 
+
The following constitutive relationships now hold:
+
 
+
:<math> \mathbf{D} = \epsilon \mathbf{E}, \quad \quad \mathbf{J} = \sigma \mathbf{E} </math>
+
 
+
:<math> \mathbf{B} = \mu \mathbf{H}, \quad \quad \mathbf{M} = \sigma_m \mathbf{H} </math>
+
 
+
where &sigma;<sub>m</sub> is the magnetic conductivity of the medium in &Omega;/m.
+
 
+
Additionally, one can write the following continuity equations:
+
 
+
:<math> \nabla . \mathbf{J} - \dfrac{\partial \rho}{\partial t} = 0 </math>
+
 
+
:<math> \nabla . \mathbf{M} - \dfrac{\partial \rho_m}{\partial t} = 0 </math>
+
 
+
== The Wave Equations ==
+
 
+
Combining Maxwell's equations, we can arrive at the electric and magnetic wave equations:
+
 
+
:<math> \nabla^2 \mathbf{E} - \epsilon \mu \dfrac{\partial \mathbf{E}}{\partial t} = 0 </math>
+
 
+
:<math> \nabla^2 \mathbf{H} - \epsilon \mu \dfrac{\partial \mathbf{H}}{\partial t} = 0 </math>
+
 
+
where '''&nabla;'''<sup>2</sup> is the Laplacian operator. The wave equations are hyperbolic partial differential equation in space coordinates and time, which must be solved subject to the proper initial and boundary conditions.
+
 
+
== Electric & Magnetic Boundary Conditions ==
+
 
+
The electric field boundary conditions at the interface between two material media are:
+
 
+
<math> \hat{\mathbf{n}} . [ \mathbf{D_2(r)} - \mathbf{D_1(r)} ] = \rho_s (\mathbf{r})  </math>
+
 
+
<math> \hat{\mathbf{n}} \times [ \mathbf{E_2(r)} - \mathbf{E_1(r)} ] =  - \mathbf{M_s(r)}  </math>
+
 
+
where <math>\hat{\mathbf{n}}</math> is the unit normal vector at the interface pointing from medium 1 towards medium 2, and &rho;<sub>s</sub> is the electric surface charge density, and <b>M<sub>s</sub></b> is the magnetic surface current density at the interface.
+
 
+
The magnetic field boundary conditions at the interface between two material media are:
+
 
+
<math> \hat{\mathbf{n}} . [ \mathbf{B_2(r)} - \mathbf{B_1(r)} ] = \rho_{ms} (\mathbf{r}) </math>
+
 
+
<math> \hat{\mathbf{n}} \times [ \mathbf{H_2(r)} - \mathbf{H_1(r)} ] = \mathbf{J_s(r)}  </math>
+
 
+
where <math>\hat{\mathbf{n}}</math> is the unit normal vector at the interface pointing from medium 1 towards medium 2, &rho;<sub>ms</sub> is the magnetic surface charge density, and <b>J<sub>s</sub></b> is the electric surface current density at the interface.
+
 
+
<table>
+
<tr>
+
<td> [[File:BC1.png|thumb|left|420px|The interface between two material media and definition of unit normal vector.]]
+
</td>
+
</tr>
+
</table>
+
 
+
== Maxwell's Equations in Integral Form ==
+
 
+
In certain applications, it is advantageous to cast Maxwell's equations in an integral form. These can be done using the theorems of vector calculus. In an isotropic, time-invariant and homogeneous medium, the integral forms of Maxwell's equations are given by:
+
 
+
:<math> \int\int_S \mathbf{D} . \mathbf{ds} = \int\int\int_V \rho dv </math>
+
 
+
:<math> \int\int_S \mathbf{B} . \mathbf{ds} = \int\int\int_V \rho_m dv </math>
+
 
+
:<math> \int_C \mathbf{E} . \mathbf{dl} = - \dfrac{\partial}{\partial t} \int\int_S \mathbf{B} . \mathbf{ds} - \int\int_S \mathbf{M} . \mathbf{ds} </math>
+
 
+
:<math> \int_C \mathbf{H} . \mathbf{dl} =  \dfrac{\partial}{\partial t} \int\int_S \mathbf{D} . \mathbf{ds} - \int\int_S \mathbf{J} . \mathbf{ds} </math>
+
 
+
where V is a closed region of the space, S the surface boundary and C is a path.
+
 
+
== Time-Harmonic Form of Maxwell's Equations ==
+
 
+
In a time-harmonic system operating at a given frequency f, the time dependence of the fields takes the form of <math> e^{j\omega t} </math>, where <math> j = \sqrt{-1} </math>, and &omega; = 2&pi;f is the angular frequency. In that case, the time derivative is <math> {\partial}/{\partial t} = j\omega </math>, and Maxwell's curl equations reduce to:
+
 
+
:<math> \nabla \times \mathbf{E} = - j\omega \mathbf{B} - \mathbf{M} </math>
+
 
+
:<math> \nabla \times \mathbf{H} = j\omega \mathbf{D} + \mathbf{J} </math>
+
 
+
and the continuity equations reduce to:
+
 
+
:<math> \nabla . \mathbf{J} = j\omega \rho </math>
+
 
+
:<math> \nabla . \mathbf{M} = j\omega \rho_m </math>
+
 
+
The wave equations then reduce to the Helmholtz equations given by:
+
+
:<math> \nabla^2 \mathbf{E} + k^2 \mathbf{E} = 0 </math>
+
 
+
:<math> \nabla^2 \mathbf{H} + k^2 \mathbf{H} = 0 </math>
+
 
+
where <math> k = \omega \sqrt{\epsilon \mu} </math> is the propagation constant in the medium. 
+
 
+
== Electric and Magnetic Potentials ==
+
 
+
Under the time-harmonic assumption, the electric and magnetic fields can further be expressed in terms of an electric scalar potential &Phi;, a magnetic scalar potential &Psi;, a vector electric potential '''F''' and a vector magnetic potential '''A''' in the following form:
+
 
+
:<math> \mathbf{E(r)} = - \nabla \times \mathbf{F(r)} - \nabla \Phi(\mathbf{r}) - j\omega\mu \mathbf{A(r)} </math>
+
 
+
:<math> \mathbf{H(r)} = \nabla \times \mathbf{A(r)} - \nabla \Psi(\mathbf{r}) - j\omega\epsilon \mathbf{F(r)} </math>
+
 
+
with the additional gauge relations:
+
 
+
:<math> \nabla \times \mathbf{A(r)} = - j\omega\epsilon \Phi(\mathbf{r}) </math>
+
 
+
:<math> \nabla \times \mathbf{F(r)} = - j\omega\mu \Psi(\mathbf{r}) </math>
+
 
+
All the potential functions satisfy the Helmholtz equation:
+
 
+
:<math> \nabla^2 \mathbf{A} + k^2 \mathbf{A} = - \mathbf{J} </math>
+
 
+
:<math> \nabla^2 \mathbf{F} + k^2 \mathbf{F} = - \mathbf{M} </math>
+
 
+
:<math> \nabla^2 \Phi + k^2 \Phi = - \frac{\rho}{\epsilon} </math>
+
 
+
:<math> \nabla^2 \Psi + k^2 \Psi = - \frac{\rho_m}{\mu} </math>
+
 
+
Sometimes it is useful to define the Hertz vector potentials as:
+
 
+
:<math> \mathbf{\Pi_e} = \frac{1}{j\omega\epsilon} \mathbf{A(r)} </math>
+
 
+
:<math> \mathbf{\Pi_m} = \frac{1}{j\omega\mu} \mathbf{F(r)} </math>
+
 
+
In that case, the electric and magnetic fields can be fully expressed in terms of these two vector potentials:
+
 
+
:<math> \mathbf{E(r)} = -j\omega\mu \nabla \times \mathbf{\Pi_m} + k^2 \mathbf{\Pi_e} + \nabla \nabla . \mathbf{\Pi_e} </math>
+
 
+
:<math> \mathbf{H(r)} = j\omega\epsilon\nabla \times \mathbf{\Pi_e} + k^2 \mathbf{\Pi_m} + \nabla \nabla . \mathbf{\Pi_m} </math>
+
 
+
{{Note|For historical reasons, it is customary in electrostatic problems to directly set the magnetic flux density '''B'''  equal to &nabla; &times; '''A'''. [[EM.Cube]]'s Static Module ([[EM.Ferma]]) uses that convention for definition of the vector magnetic potential '''A''', which is different by a factor of &mu; from the definition of the '''A''' vector in the electrodynamic discussion presented in this section. That would also change the source term of the Helmholtz equation by the same factor.}} 
+
 
+
== Green’s Function Representations ==
+
 
+
The Green’s functions are the analytical solutions of boundary value problems when they are excited by an elementary source. This is usually an infinitesimally small vectorial point source.  The total electric ('''E''') field and total magnetic ('''H''') field can be expressed in terms of the volume electric current source '''J''' and volume magnetic current source '''M''' in the following way:
+
 
+
:<math> \mathbf{E = E^{inc}} +  \mathbf{\iiint_{V_J} \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') } d \nu' +  \mathbf{\iiint_{V_M} \overline{\overline{G}}_{EM}(r|r') \cdot M(r') } d \nu' </math>
+
 
+
 
+
:<math> \mathbf{H = H^{inc}} +  \mathbf{\iiint_{V_J} \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') } d \nu' +  \mathbf{\iiint_{V_M} \overline{\overline{G}}_{HM}(r|r') \cdot M(r') } d \nu' </math>
+
<!--[[File:PMOM1(1).png]]-->
+
 
+
 
+
where '''E<sup>inc</sup>''' and '''H<sup>inc</sup>''' are the incident electric and magnetic fields, respectively, and V<sub>J</sub> and V<sub>M</sub> are the volumes containing the electric and magnetic current sources, respectively. The above equations involve four dyadic Green's functions types: dyadic electric-field Green’s functions due to electric current sources '''G<sub>EJ</sub>''', dyadic magnetic-field Green’s functions due to electric current sources '''G<sub>HJ</sub>''', dyadic electric-field Green’s functions due to magnetic current sources '''G<sub>EM</sub>''', and dyadic magnetic-field Green’s functions due to magnetic current sources '''G<sub>HM</sub>''' . The incident or impressed fields represent the source terms and provide the excitation of the structure.
+
 
+
In order for the Green’s functions to be computationally useful, they must have analytical closed forms. This can be a mathematical expression or a more complex recursive process. It is no surprise that only very few electromagnetic boundary value problems have closed-form Green’s functions. Among [[EM.Cube]]'s computational modules, [[EM.Libera]] is based on the free-space Green's functions, whereas [[EM.Picasso]] is based on the dyadic Green's functions of an arbitrary multilayer planar structure.
+
 
+
== Free-Space Field Solutions ==
+
 
+
The simplest background structure is the unbounded free space, which is represented by the following Green’s function:
+
 
+
:<math> \mathbf{ \overline{\overline{G}}_{EJ}(r|r') = (\overline{\overline{I}} + \nabla\nabla) } G_{\Lambda} (\mathbf{r|r'}), \quad G_{\Lambda} (\mathbf{r|r'}) = \frac{ e^{-jk_0 \mathbf{|r-r'|}} }{ 4\pi \mathbf{|r-r'|} } </math>
+
<!--[[File:03_freespace_tn.gif]]-->
+
 
+
where <math>\mathbf{\overline{\overline{I}}}</math> is the unit dyad, <math>\nabla</math> is the gradient operator, '''r''' and '''r'''' are the position vectors of the observation and source points, respectively, and k<sub>0</sub> is the free-space propagation constant. This implies that electromagnetic waves propagate in free space in a spherical form away from the source. Note that the Green’s function has a singularity at the source, <i>i.e.</i> when '''r''' = '''r''''.
+
 
+
Assuming electric and magnetic surface current sources '''J''' and '''M''' residing on surfaces S<sub>J</sub> and S<sub>M</sub>, respectively, the near-field equations reduce to:
+
 
+
 
+
:<math> \begin{align} \mathbf{ E(r) = E^{inc}(r) }  & - jk_0 Z_0 \iint_{S_J} \left\{ \left[ 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right] \mathbf{J(r')} -  \left[ 1 - \frac{3j}{k_0 R} - \frac{3}{(k_0 R)^2} \right] \mathbf{ (\hat{R} \cdot J(r')) \hat{R} } \right\} \frac{e^{-jk_0 R}}{4\pi R} ds' \\ & + jk_0 \iint_{S_M} \left[ 1-\frac{j}{k_0 R} \right] \mathbf{ (\hat{R} \times M(r')) } \frac{e^{-jk_0 R}}{4\pi R} ds' \end{align} </math>
+
 
+
 
+
:<math> \begin{align} \mathbf{ H(r) = H^{inc}(r) }  & - jk_0 Y_0 \iint_{S_M} \left\{ \left[ 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right] \mathbf{M(r')} -  \left[ 1 - \frac{3j}{k_0 R} - \frac{3}{(k_0 R)^2} \right] \mathbf{ (\hat{R} \cdot M(r')) \hat{R} } \right\} \frac{e^{-jk_0 R}}{4\pi R} ds' \\ & - jk_0 \iint_{S_J} \left[ 1-\frac{j}{k_0 R} \right] \mathbf{ (\hat{R} \times J(r')) } \frac{e^{-jk_0 R}}{4\pi R} ds' \end{align} </math>
+
<!--[[File:PO7.png]]-->
+
 
+
 
+
where <math> R=|r-r'| \text{, } k_0 = \frac{2\pi}{\lambda_0} \text{ and } Z_0 = 1/Y_0 = \eta_0 </math>.
+
 
+
== Free-Space Wave Propagation ==
+
 
+
In a free-space line-of-sight (LOS) communication system, the signal propagates directly from the transmitter to the receiver without encountering any obstacles (scatterers). Electromagnetic waves propagate in the form of spherical waves with a functional dependence of e<sup>j(&omega;</sup><sup>t-k<sub>0</sub>R)</sup>/R, where R is the distance between the transmitter and receiver, <math>\omega = 2\pi f</math>, f is the signal frequency, <math>k_0 = \frac{\omega}{c} = \frac{2\pi}{\lambda}</math>, c is the speed of light, and &lambda;<sub>0</sub> is the free-space wavelength at the operational frequency. By the time the signal arrives at the location of the receiver, it undergoes two changes. It is attenuated and its power drops by a factor of 1/R<sup>2</sup>, and additionally, it experiences a phase shift of <math>\frac{2\pi R}{\lambda_0}</math>, which is equivalent to a time delay of R/c. The signal attenuation from the transmitter to the receiver is usually quantified by '''Path Loss''' defined as the ratio of the received signal power (P<sub>R</sub>) to the transmitted signal power (P<sub>T</sub>). Assuming isotropic transmitting and receiving radiators (<i>i.e.</i> radiating uniformly in all directions), the Path Loss in a free-space line-of-sight communication system is given by Friis’ formula:
+
 
+
:<math> \frac{P_R}{P_T} = \left( \frac{\lambda_0}{4\pi R} \right)^2 </math>
+
The above formula assumes that the receiving antenna is polarization-matched. Normally, there is a polarization mismatch between the transmitting and receiving antennas. In the case of directional transmitting and receiving antennas, Friis’ formula takes the following form:
+
 
+
:<math>  P_R = P_T \, G_T G_R \left( \frac{\lambda_0}{4\pi R} \right)^2 \left| \mathbf{ \hat{u}_T \cdot \hat{u}_R } \right|^2 </math>
+
 
+
where '''u<sub>T</sub>''' and '''u<sub>R</sub>''' are the unit polarization vectors of the transmitting and receiving antennas, and G<sub>T</sub> and G<sub>R</sub> are their gains, respectively.
+
 
+
<table>
+
<tr>
+
<td>
+
[[Image:los.png|thumb|left|640px|A line-of-sight (LOS) propagation scenario.]]
+
</td>
+
</tr>
+
</table>
+
 
+
== Basic Wave Interaction Mechanisms ==
+
 
+
[[EM.Terrano]] discretizes all of the objects in your propagation scene into flat triangular facets. Obviously, rectangular and cubic objects preserve their geometric shapes through this discretization. Objects with curved surfaces such as cylinders, cones or spheres, are approximated by triangular surface mesh representations. The geometric fidelity of the resulting mesh depends on the specified mesh edge length. When a ray hits a triangular facet, the propagating spherical wave is approximated as a plane wave at the specular point. The reflection and transmission coefficients of the surface are calculated at the operational frequency and at the particular ray incident angle.
+
 
+
A new reflected ray is generated at the specular point, which starts traveling and bouncing around in the scene. If the obstructing surface is penetrable, a second transmitted ray is generated and added to the scene. If the ray hits the edge of an obstacle, it is diffracted from that edge. This leads to the creation of a cone of new rays, which greatly complicate the computational problem. The Uniform Theory of Diffraction (UTD) is used to calculate the wedge diffraction coefficients at the edges of scattering blocks. Note that reflection, transmission and diffraction coefficients are all dependent on the polarization of the incident plane wave.
+
 
+
A receiver may receive a large number of rays: direct line-of-sight rays from the transmitter, rays reflected or diffracted off the ground or terrain, rays reflected or diffracted from buildings or rays transmitted through buildings. Each received ray is characterized by its power, delay and angles of arrival, which are the spherical coordinate angles &theta; and &phi; of the incoming ray. The actual signal received and detected by the receiver is the superposition of all these rays with different power levels and different time delays. Most of the time, you will be interested in the coverage map of an area, which shows how much power is received by a grid of receivers spread over the area from a given fixed transmitter.
+
 
+
== Ray Reflection & Transmission ==
+
 
+
The incident, reflected and transmitted rays are each characterized by a triplet of unit vectors:
+
 
+
* <math>( \mathbf{ \hat{u}_{\|}, \hat{u}_{\perp}, \hat{k} } )</math> representing the incident parallel polarization vector, incident perpendicular polarization vector and incident propagation vector, respectively.
+
* <math>( \mathbf{ \hat{u}_{\|}^{\prime}, \hat{u}_{\perp}', \hat{k}' } )</math> representing the reflected parallel polarization vector, reflected perpendicular polarization vector and reflected propagation vector, respectively.
+
* <math>( \mathbf{ \hat{u}_{\|}^{\prime\prime}, \hat{u}_{\perp}^{\prime\prime}, \hat{k}^{\prime\prime} } )</math> representing the transmitted parallel polarization vector, transmitted perpendicular polarization vector and transmitted propagation vector, respectively.
+
 
+
<table>
+
<tr>
+
<td>
+
[[Image:reflect.png|thumb|left|480px|The incident, reflected and transmitted rays at the interface between two dielectric media.]]
+
</td>
+
</tr>
+
</table>
+
 
+
The reflected ray is assumed to originate from a virtual image source point. The three triplets constitute three orthonormal basis systems. Below, it is assumed that the two dielectric media have permittivities &epsilon;<sub>1</sub> and &epsilon;<sub>2</sub>, and permeabilities &mu;<sub>1</sub> and &mu;<sub>2</sub>, respectively. A lossy medium with a conductivity &sigma; can be modeled by a complex permittivity &epsilon;<sub>r</sub> = &epsilon;'<sub>r</sub> –j&sigma;/&epsilon;<sub>0</sub>. Assuming '''n''' to be the unit normal to the interface plane between the two media, and Z<sub>0</sub> = 120&Omega; , the incident polarization vectors as well as all the reflected and transmitted vectors are found as:
+
 
+
:<math> \mathbf{ \hat{u}_{\perp} = \frac{\hat{k} \times \hat{n}}{|\hat{k} \times \hat{n}|} } </math>
+
 
+
:<math> \mathbf{ \hat{u}_{\|} = \hat{u}_{\perp} \times \hat{k} } </math>
+
<!--[[File:frml1.png]]-->
+
 
+
The reflected unit vectors are found as:
+
 
+
:<math> \mathbf{ \hat{k}' = \hat{k} - 2(\hat{k} \cdot \hat{n}) \hat{n} } </math>
+
 
+
:<math> \mathbf{ \hat{u}_{\perp}' = \hat{u}_{\perp} } </math>
+
 
+
:<math> \mathbf{ \hat{u}_{\|}' = \hat{u}_{\perp}' \times \hat{k}' } </math>
+
<!--[[File:frml2.png]]-->
+
 
+
The transmitted unit vectors are found as:
+
 
+
:<math> \mathbf{ \hat{k}^{\prime\prime} = \hat{n} \times a - \sqrt{1-a \cdot a} \; \hat{n} } </math>
+
+
:<math> \mathbf{ \hat{u}_{\perp}^{\prime\prime}  = \hat{u}_{\perp} } </math>
+
 
+
:<math> \mathbf{ \hat{u}_{\|}^{\prime\prime}  = \hat{u}_{\perp}^{\prime\prime}  \times \hat{k}^{\prime\prime}  } </math>
+
<!--[[File:frml3.png]]-->
+
 
+
where
+
 
+
:<math> \mathbf{a} = (k_1/k_2) \mathbf{\hat{k} \times \hat{n}}</math>
+
 
+
:<math> k_1 = k_0 \sqrt{\varepsilon_1 \mu_1} </math>
+
 
+
:<math> k_2 = k_0 \sqrt{\varepsilon_2 \mu_2} </math>
+
 
+
:<math> \eta_1 = Z_0 \sqrt{\mu_1 / \varepsilon_1} </math>
+
 
+
:<math> \eta_2 = Z_0 \sqrt{\mu_2 / \varepsilon_2} </math>
+
 
+
 
+
:<math> \sin\theta^{\prime\prime} = \frac{k_1}{k_2}\sin\theta \text{ if } \sin\theta \le k_2/k_1</math>
+
<!--
+
[[File:frml4.png]]
+
 
+
[[File:frml5.png]]
+
-->
+
 
+
The reflection coefficients at the interface are calculated for the two parallel and perpendicular polarizations as:
+
 
+
:<math> R_{\|} = \frac { \eta_2(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) - \eta_1(\mathbf{ \hat{k} \cdot \hat{n} }) }  { \eta_2(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) + \eta_1(\mathbf{ \hat{k} \cdot \hat{n} }) } = \frac{\eta_2 \cos\theta^{\prime\prime}  - \eta_1 \cos\theta} {\eta_2 \cos\theta^{\prime\prime}  + \eta_1 \cos\theta} = \frac{Z_{2\|} - Z_{1\|}} {Z_{2\|} + Z_{1\|}} </math>
+
 
+
 
+
:<math> R_{\perp} = \frac { \eta_2(\mathbf{ \hat{k} \cdot \hat{n} }) - \eta_1(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) }  { \eta_2(\mathbf{ \hat{k} \cdot \hat{n} }) + \eta_1(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) } = \frac{\eta_2 / \cos\theta^{\prime\prime}  - \eta_1 / \cos\theta} {\eta_2 / \cos\theta^{\prime\prime}  + \eta_1 / \cos\theta} = \frac{Z_{2\perp} - Z_{1\perp}} {Z_{2\perp} + Z_{1\perp}} </math>
+
<!--[[File:frml6.png]]-->
+
 
+
 
+
== Penetration through Thin Walls or Surfaces ==
+
 
+
In &quot;Thin Wall Approximation&quot;, we assume that an incident ray gives rise to two rays, one is reflected at the specular point, and the other is transmitted almost in the same direction as the incident ray. The reflected ray is assumed to originate from a virtual image source point. Similar to the case of reflection and transmission at the interface between two dielectric media, here too we have three triplets of unit vectors, which all form orthonormal basis systems.
+
 
+
<table>
+
<tr>
+
<td>
+
[[Image:thinwalltrans.png|thumb|left|480px|The incident and transmitted rays through a thin wall.]]
+
</td>
+
</tr>
+
</table>
+
 
+
The transmission coefficients are calculated for the two parallel and perpendicular polarizations as:
+
 
+
:<math> T_{\|} = \frac{(1-{\Gamma_{\|}}^2) \exp(-jk_2 d (\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n}}))} { 1-{\Gamma_{\|}}^2 \exp( -2jk_2 d (\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) ) } </math>
+
 
+
 
+
:<math> T_{\perp} = \frac{(1-{\Gamma_{\perp}}^2) \exp(-jk_2 d (\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n}}))} { 1-{\Gamma_{\perp}}^2 \exp( -2jk_2 d (\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) ) } </math>
+
<!--[[File:frml20.png]]-->
+
 
+
where
+
 
+
:<math> \Gamma_{\|} = \frac{ \eta_2(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) - \eta_1(\mathbf{ \hat{k} \cdot \hat{n} }) } { \eta_2(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) + \eta_1(\mathbf{ \hat{k} \cdot \hat{n} }) } = \frac{\eta_2 \cos\theta^{\prime\prime}  - \eta_1 \cos\theta} {\eta_2 \cos\theta^{\prime\prime}  + \eta_1 \cos\theta} = \frac{Z_{2\|} - Z_{1\|}} {Z_{2\|} + Z_{1\|}} </math>
+
 
+
 
+
:<math> \Gamma_{\perp} = \frac{ \eta_2(\mathbf{ \hat{k} \cdot \hat{n} }) - \eta_1(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) } { \eta_2(\mathbf{ \hat{k} \cdot \hat{n} }) + \eta_1(\mathbf{ \hat{k}^{\prime\prime}  \cdot \hat{n} }) } = \frac{\eta_2 / \cos\theta^{\prime\prime}  - \eta_1 / \cos\theta} {\eta_2 / \cos\theta^{\prime\prime}  + \eta_1 / \cos\theta} = \frac{Z_{2\perp} - Z_{1\perp}} {Z_{2\perp} + Z_{1\perp}} </math>
+
<!--[[File:frml21.png]]-->
+
 
+
 
+
== Wedge Diffraction from Edges ==
+
 
+
For the purpose of calculation of diffraction from building edges, we define a &quot;Wedge&quot; as having two faces, the 0-face and the ''n''-face. The wedge angle is a = (2-''n'')p, where the parameter ''n'' is required for the calculation of diffraction coefficients. All the diffracted rays lie on a cone with its vertex at the diffraction point and a wedge angle equal to the angle of incidence in the opposite direction. A diffracted ray is assumed to originate from a virtual image source point. Three triplets of unit vectors are defined as follows:
+
 
+
* <math>\mathbf{(\hat{u}_0, \hat{u}_l, \hat{t})}</math> representing the unit vector normal to the edge and lying in the plane of the 0-face, the unit vector normal to the 0-face, and the unit vector along the edge, respectively.
+
* <math>\mathbf{(\hat{u}_f, \hat{u}_b, \hat{t})}</math> representing the incident forward polarization vector, incident backward polarization vector and incident propagation vector, respectively.
+
* <math>\mathbf{(\hat{u}_f', \hat{u}_b', \hat{t}')}</math> representing the diffracted forward polarization vector, diffracted backward polarization vector and diffracted propagation vector, respectively.
+
 
+
<table>
+
<tr>
+
<td>
+
[[Image:diffract.png|thumb|left|480px|The incident ray and diffracted ray cone at the edge of a building.]]
+
</td>
+
</tr>
+
</table>
+
 
+
The three triplets constitute three orthonormal basis systems. The propagation vector '''k'''' of the diffracted ray has to be constructed based on the diffraction cone as follows:
+
 
+
:<math> \mathbf{\hat{k}'} = \cos\phi_w \mathbf{\hat{u}_0} + \sin\phi_w \mathbf{\hat{u}_l} + \mathbf{(\hat{k} \cdot \hat{t}) \hat{t}}, \quad 0 \le \phi_w \le \alpha</math>
+
<!--[[File:frml8.png]]-->
+
 
+
where the resolution of the angle &theta;<sub>w</sub> is chosen to be the same as the resolution of the incident ray.
+
 
+
The other unit vectors for the incident and diffracted rays are found as:
+
 
+
:<math> \mathbf{ \hat{u}_f = \frac{\hat{k} \times \hat{t}}{|\hat{k} \times \hat{t}|} } </math>
+
 
+
:<math> \mathbf{ \hat{u}_b = \hat{k} \times \hat{u}_f } </math>
+
 
+
:<math> \mathbf{ \hat{u}_f' = \frac{\hat{k}' \times \hat{t}}{|\hat{k}' \times \hat{t}|} } </math>
+
 
+
:<math> \mathbf{ \hat{u}_b' = \hat{k}' \times \hat{u}_f' } </math>
+
 
+
The diffraction coefficients are calculated in the following way:
+
 
+
:<math> D_s = \frac{-e^{-j\pi/4}}{2n \sqrt{2\pi k} \sin\beta_0'} \left\lbrace \begin{align} & \cot \left(\frac{\pi + (\phi-\phi')}{2n}\right) F[kLa^+(\phi-\phi')] +  \cot \left(\frac{\pi - (\phi-\phi')}{2n}\right) F[kLa^-(\phi-\phi')] + \\ & R_{0 \perp} \cot \left(\frac{\pi - (\phi+\phi')}{2n}\right) F[kLa^-(\phi+\phi')] +  R_{n \perp} \cot \left(\frac{\pi + (\phi+\phi')}{2n}\right) F[kLa^+(\phi+\phi')] \end{align} \right\rbrace </math>
+
 
+
 
+
:<math> D_h = \frac{-e^{-j\pi/4}}{2n \sqrt{2\pi k} \sin\beta_0'} \left\lbrace \begin{align} & \cot \left(\frac{\pi + (\phi-\phi')}{2n}\right) F[kLa^+(\phi-\phi')] +  \cot \left(\frac{\pi - (\phi-\phi')}{2n}\right) F[kLa^-(\phi-\phi')] + \\ & R_{0 \|} \cot \left(\frac{\pi - (\phi+\phi')}{2n}\right) F[kLa^-(\phi+\phi')] +  R_{n \|} \cot \left(\frac{\pi + (\phi+\phi')}{2n}\right) F[kLa^+(\phi+\phi')] \end{align} \right\rbrace </math>
+
 
+
where ''F(x)'' is the Fresnel Transition function:
+
 
+
:<math> F(x) = 2j \sqrt{x} e^{jx} \int_{\sqrt{x}}^{\infty} e^{-j\tau^2} \, d\tau </math>
+
 
+
In the above equations, we have
+
 
+
:<math> \begin{align} s = |\rho_D - \rho_S| \\ s' = |\rho_D - \rho_r| \end{align} </math>
+
 
+
:<math>L = \frac{s s' \sin^2 \beta'}{s + s'} </math>
+
 
+
:<math>a^{\pm}(\nu) = 2\cos^2 \left( \frac{2n\pi N^{\pm} - \nu}{2} \right), \quad \nu = \phi \pm \phi' </math>
+
 
+
where <math>N^{\pm}</math> are the integers which most closely satisfy the equations <math> 2n\pi N^{\pm} - \nu = \pm \pi </math>.
+
 
+
== Multilayer Green’s Functions ==
+
 
+
[[Image:PMOM14.png|thumb|400px|A typical planar layered structure.]]
+
The Green’s functions are the solutions of boundary value problems when they are excited by an elementary source. This is usually assumed to be an infinitesimally small vectorial point source. In order for Green’s functions to be computationally useful, they must have analytical closed forms like a mathematical expression, or one should be able to compute them using a recursive process. It turns out that only very few boundary value problems have closed-form Green’s functions. Planar layered structures with laterally infinite extents are one of those few cases, which can be represented by recursive dyadic Green's functions.
+
 
+
In general, a structure may support both electric ('''J''') and magnetic ('''M''') currents. The total electric ('''E''') and magnetic ('''H''') fields can be expressed in terms of the electric and magnetic currents in the following way:
+
 
+
:<math>E = E^{inc} + \iiint\limits_V \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') \, dv' + \iiint\limits_V \overline{\overline{G}}_{EM}(r|r') \cdot M(r') \, dv'</math>
+
 
+
:<math>H = H^{inc} + \iiint\limits_V \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') \, dv' + \iiint\limits_V \overline{\overline{G}}_{HM}(r|r') \cdot M(r') \, dv'</math>
+
<!--[[File:PMOM1(1).png]]-->
+
 
+
where '''G<sub>EJ</sub>''', '''G<sub>EM</sub>''', '''G<sub>HJ</sub>''', '''GH<sub>M</sub>''' are the dyadic Green’s functions for the electric and magnetic currents due to electric and magnetic current source, respectively, and '''E<sup>i</sup>''' and '''H<sup>i</sup>''' are the incident or impressed electric and magnetic fields, respectively. In these equations, '''r''' is the position vector of the observation point and '''r'''' is the position vector of the source point. V is the volume that contains all the sources and the volume integration is performed with respect to the primed coordinates. The incident or impressed fields provide the excitation of the structure. They may come from an incident plane wave or a gap source on a microstrip line, a short dipole, etc. The complexity of the Green’s functions depends on what is considered as the background structure. If you remove all the unknown currents from the structure, you are left with the background structure.
+
 
+
== Planar Integral Equations ==
+
 
+
To derive a system of integral equations, we enforce the boundary conditions on the integral definitions of the '''E''' and '''H''' fields as follows:
+
 
+
:<math>L_E(E) = L_E \bigg\{ E^{inc} + \iiint\limits_V \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') \, dv' + \iiint\limits_V \overline{\overline{G}}_{EM}(r|r') \cdot M(r') \, dv' \bigg\} </math>
+
 
+
:<math>L_H(H) = L_H \bigg\{ H^{inc} + \iiint\limits_V \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') \, dv' + \iiint\limits_V \overline{\overline{G}}_{HM}(r|r') \cdot M(r') \, dv' \bigg\} </math>
+
<!--[[File:PMOM4(2).png]]-->
+
 
+
where '''L<sub>E</sub>''' is the boundary value operator for the electric field and '''L<sub>H</sub>''' is the boundary value operator for the magnetic field. For example, '''L<sub>E</sub>''' may require that the tangential components of the '''E'''field vanish on perfect conductors:
+
 
+
:<math> \hat{n} \times \hat{n} \times \mathbf{E} = 0, \quad \mathbf{r} \in PEC </math>
+
<!--[[File:PMOM65.png]]-->
+
 
+
Or '''L<sub>E</sub>''' and '''L<sub>H</sub>''' may require that the tangential components of the '''E''' and '''H''' fields be continuous across an aperture in a perfect ground plane:
+
 
+
:<math>\begin{cases} \hat{n} \times \hat{n} \times (\mathbf{E}^+ - \mathbf{E}^-) = 0 \\ \hat{n} \times \hat{n} \times (\mathbf{H}^+ - \mathbf{H}^-) = 0 \end{cases}  \quad \Rightarrow \quad \mathbf{M}^+(r) = \mathbf{M}^-(r), \quad r \in PMC </math>
+
<!--[[File:PMOM66(1).png]]-->
+
 
+
Given the fact that the dyadic Green’s functions and the incident or impressed fields are all known, one can solve the above system of integral equations to find the unknown currents '''J''' and '''M'''.
+
 
+
In [[EM.Picasso]], magnetic currents are always surface current with units of V/m. Electric currents, however, can be surface currents with units of A/m as in the case of metallic traces like microstrip lines, or they can be volume currents with units of A/m<sup>2</sup> as in the case of perfectly conducting vias. Dielectric inserts are modeled as volume polarization currents that are related to the electric field '''E''' in the following manner:
+
 
+
:<math>\mathbf{J}_p(r) = jk_0 Y_0(\varepsilon_r - \varepsilon_b)\mathbf{E}(r)</math>
+
<!--[[File:PMOM5.png]]-->
+
 
+
where k<sub>0</sub> is the free space propagation constant, <math>Y_0 = \tfrac{1}{Z_0} = \tfrac{1}{120\pi}</math> is the free space intrinsic admittance, &epsilon;<sub>r</sub> is the permittivity of the dielectric insert, and &epsilon;<sub>b</sub> is the permittivity of its background layer. In a 2.5-D formulation, it is assumed that the volume currents have only a vertical component along the Z direction, and their circumferential components are negligible.
+
 
+
== Numerical Solution of Integral Equations ==
+
 
+
The planar integral equations derived earlier can be solved numerically by discretizing the unknown currents using a proper meshing scheme. The original functional equations are reduced to discretized linear algebraic equations over elementary cells. The unknown quantities are found by solving this system of linear equations, and many other [[parameters]] can be computed thereafter. This method of numerical solution of integral equations is known as the Method of Moments (MoM). In this method, the unknown electric and magnetic currents are represented by expansions of basis functions as follows:
+
 
+
:<math>J(r) = \sum_{n=1}^N I_n^{(J)} f_n^{(J)} (r)</math>
+
 
+
:<math>M(r) = \sum_{k=1}^K V_k^{(M)} f_k^{(M)} (r)</math>
+
<!--[[File:PMOM2.png]]-->
+
 
+
where <math>f_n^{(J)}</math> and <math>f_k^{(M)}</math> are the generalized vector basis functions for the expansion of electric and magnetic currents, respectively, and <math>I_n^{(J)}</math> and <math>V_k^{(M)}</math> are the unknown amplitudes of these basis functions, which have to be determined. Substituting these expansions into the integral equations generates a set of discretized integral equations, which can further be converted to a system of linear algebraic equations. This is accomplished by testing the discretized integral equations using the a set of test functions. In the method of moments, the Galerkin technique is typically used, which chooses the expansion basis functions as test functions. This leads to the following linear system:
+
 
+
:<math> \begin{bmatrix} Z^{(EJ)} & T^{(EM)} \\ U^{(HJ)} & Y^{(HM)} \end{bmatrix} \cdot \begin{bmatrix} I^{(J)} \\ V^{(M)} \end{bmatrix} = \begin{bmatrix} V^{(E)} \\ I^{(M)} \end{bmatrix} </math>
+
<!--[[File:PMOM3.png]]-->
+
 
+
where
+
 
+
:<math> Z_{ij}^{(EJ)} = \iiint\limits_{V_i} dv f_i^{(J)}(r) \cdot  \iiint\limits_{V_j} dv' \overline{\overline{G}}_{EJ}(r|r') \cdot f_i^{(J)}(r')</math>
+
<!--[[File:PMOM6.png]]-->
+
 
+
and
+
 
+
:<math> V_i^{(E)} = \iiint\limits_{V_i} dv f_i^{(J)}(r) \cdot E^{inc}(r) </math>
+
 
+
 
+
:<math> I_i^{(H)} = \iiint\limits_{V_i} dv f_i^{(M)}(r) \cdot H^{inc}(r) </math>
+
<!--[[File:PMOM7.png]]-->
+
 
+
Similar expressions can be derived for the T<sup>(EM)</sup>, U<sup>(HJ)</sup> and Y<sup>(HM)</sup>elements of the MoM matrix.
+
 
+
== Discretization Of Electric &amp; Magnetic Currents ==
+
 
+
The right choice of the basis functions to represent the elementary currents is very important. It will determine the accuracy and computational efficiency of the resulting numerical solution. Rooftop basis functions are one of the most popular types of basis functions used in a variety of MoM formulations. The surface currents (whether electric or magnetic) are discretized using 2D rooftop basis functions shown in the figure below:
+
 
+
[[File:image055_tn.png]]
+
 
+
Rooftop or RWG basis functions built over two rectangular, triangular or mixed cells.
+
 
+
The rooftop basis functions are defined over two adjacent cells with a common edge of length. If the two cells are triangular, then the so-called RWG functions are obtained. It is also possible to define rooftop functions over two adjacent rectangular cells or two adjacent rectangular and triangular cells with a common edge. On a rectangular cell, the function is defined as having a (descending or ascending) linear profile in one direction and a constant profile in the other perpendicular direction.
+
 
+
The volume polarization currents in 2.5-D MoM have a vertical direction along the Z-axis. These are discretized using prismatic basis functions that have either a rectangular or triangular base with a constant profile along the Z-axis.
+
 
+
[[File:image065_tn.png]][[File:image066_tn.png]]
+
 
+
Prismatic basis functions built over single triangular and rectangular cells.
+
 
+
== The Rectangular Mesh Advantage ==
+
 
+
Rectangular cells offer a major advantage over triangular cells for numerical MoM simulation of planar structures. This is due to the fact that the dyadic Green's functions of planar layered background structures are space-invariant on the transverse plane. Recall that the elements of the moment matrix are given by the following equation:
+
 
+
:<math> Z_{ij}^{(\mu \nu)} = \iiint_{V_i} d\nu f_i^{(\mu)}(r) \cdot \iiint_{V_j}d\nu ' \overline{\overline{G}}_{\mu \nu}(r|r') \cdot f_j^{(v)}(r') </math>
+
<!--[[File:PMOM24(1).png]]-->
+
 
+
where the spatial-domain dyadic Green's functions are a function of the observation and source coordinates, '''r'''and '''r' '''. The MoM matrix elements can indeed be interpreted as interactions between two elementary basis functions '''f<sub>i</sub>(r)''' and '''f<sub>j</sub>(r')''' on that particular background structure. The spatial-domain dyadic Green's functions can themselves be expressed in terms of the spectral-domain dyadic Green's functions as follows:
+
 
+
:<math> \overline{\overline{G}}_{\mu \nu}(r|r') = \frac{1}{(2\pi)^2} \int\limits_{-\infty}^{\infty} \int\limits_{-\infty}^{\infty}  \tilde{\overline{\overline{G}}}_{\mu \nu} (k_p, z|z') e^{-j[k_x(x-x')+k_y(y-y')]} \, dk_x \, dk_y , \quad {k_p}^2 = {k_x}^2 + {k_y}^2 </math>
+
<!--[[File:PMOM26.png]]-->
+
 
+
where the doubly infinite integration is performed with respect to the spectral [[variables]] k<sub>x</sub> and k<sub>y</sub>. As can be seen from the above expression, the spatial-domain dyadic Green's functions are functions of z, z', as well as (x-x') and (y-y'). The MoM matrix elements can now be transformed into the spectral domain as
+
 
+
:<math> Z_{ij}^{(\mu \nu)} = \dfrac{1}{(2\pi)^2} \int\limits_{-\infty}^{\infty} \int\limits_{-\infty}^{\infty} \tilde{f}_i^{(\mu)} (k_x, k_y) \cdot \tilde{\overline{\overline{G}}}_{\mu \nu} (k_{\rho}, z|z') \cdot \tilde{f}_j^{(\nu)} (k_x, k_y) \, dk_x \, dk_y </math>
+
<!--[[File:PMOM27.png]]-->
+
 
+
where the tilde symbol signifies the Fourier transform of a function defined as
+
 
+
:<math> \tilde{f}(k_x, k_y) = \dfrac{1}{(2\pi)^2} \int\limits_{-\infty}^{\infty} \int\limits_{-\infty}^{\infty} f(x,y) e^{j(k_x x + k_y y)} \, dx \, dy </math>
+
<!--[[File:PMOM28(1).png]]-->
+
 
+
Rectangular cells have simple Fourier transforms. The rooftop basis functions are triangular functions in the direction of current flow and constant in the perpendicular direction. This means that their Fourier transform is a product of a sinc-squared function along one spectral direction and a sinc function along the other. You can see from the figure below that if one deals with a rectangular mesh of identical cells (all equal and parallel), then the interactions among the rooftop basis functions become a functions of the index differences and not the absolute indices:
+
 
+
:<math> Z_{(i,k)|(j,l)} = Z \Big\langle f_{i,k}(x,y)| f_{j,l}(x', y') \Big\rangle = Z_{(i-j)|(k-l)} </math>
+
<!--[[File:PMOM29.png]]-->
+
 
+
In the above equation, the vectorial rooftop basis functions have explicit, double indices: i and k along the local X and Y directions, respectively, for the test (observation) basis function, and j and l along the local X and Y directions, respectively, for the expansion (source) basis function. Thus, uniform rectangular cells, i.e. structured rectangular cells of identical size aligned in the same direction, can speed up the planar MoM simulation significantly due to these symmetry and the invariance properties. For example, all the self-interactions are identical regardless of the location of a rooftop basis function. This reduces the matrix fill process for a total of N rooftop basis functions from an N2 process to one of order N.
+
 
+
[[File:PMOM25.png]]
+
 
+
Pairs of rooftop basis functions that have identical MoM interactions.
+
 
+
== Computing The Near Fields in Planar MoM ==
+
 
+
Once all the current distributions are known in a planar structure, the electric and magnetic fields can be calculated everywhere in that structure using the dyadic Greens's functions of the background structure:
+
 
+
:<math> \begin{align} \mathbf{E(r) = E_{inc}(r)} +  & \sum_{n=1}^N I_n^{(J)} \iiint_V \overline{\overline{G}}_{EJ}(r|r') \cdot f_n^{(J)}(r') \, d\nu' + \\ & \sum_{k=1}^K V_n^{(M)} \iiint_V \overline{\overline{G}}_{EM}(r|r') \cdot f_k^{(M)}(r') \, d\nu' \end{align} </math>
+
 
+
:<math> \begin{align} \mathbf{H(r) = H_{inc}(r)} +  & \sum_{n=1}^N I_n^{(J)} \iiint_V \overline{\overline{G}}_{HJ}(r|r') \cdot f_n^{(J)}(r') \, d\nu' + \\ & \sum_{k=1}^K V_n^{(M)} \iiint_V \overline{\overline{G}} {HM}(r|r') \cdot f_k^{(M)}(r') \, d\nu' \end{align} </math>
+
<!--[[File:PMOM92(2).png]]-->
+
 
+
The above equations can be cast into the spectral domain as follows:
+
 
+
:<math> \begin{align} \mathbf{E(r) = E_{inc}(r)} + \frac{1}{(2\pi)^2} \int\limits_{-\infty}^{\infty} \int\limits_{-\infty}^{\infty} \bigg[ & \sum_{n=1}^N I_n^{(J)} \tilde{\overline{\overline{G}}}_{EJ}(k_{\rho}, z|z') \cdot \tilde{f}_n^{(J)}(k_x, k_y) + \\ & \sum_{k=1}^K V_n^{(M)} \tilde{\overline{\overline{G}}}_{EM}(k_{\rho}, z|z') \cdot \tilde{f}_k^{(M)}(k_x, k_y) \bigg] \, dk_x \, dk_y \end{align} </math>
+
 
+
:<math> \begin{align} \mathbf{H(r) = H_{inc}(r)} + \frac{1}{(2\pi)^2} \int\limits_{-\infty}^{\infty} \int\limits_{-\infty}^{\infty} \bigg[ & \sum_{n=1}^N I_n^{(J)} \tilde{\overline{\overline{G}}}_{HJ}(k_{\rho}, z|z') \cdot \tilde{f}_n^{(J)}(k_x, k_y) + \\ & \sum_{k=1}^K V_n^{(M)} \tilde{\overline{\overline{G}}}_{HM}(k_{\rho}, z|z') \cdot \tilde{f}_k^{(M)}(k_x, k_y) \bigg] \, dk_x \, dk_y \end{align} </math>
+
<!--[[File:PMOM93(1).png]]-->
+
 
+
Calculation of the near-zone fields (fields at the vicinity of the unknown currents) is done at the post-processing stage and in a Cartesian coordinate systems. These calculations involve doubly infinite spectral-domain integrals, which are computed numerically. As was mentioned earlier, [[EM.Cube]]'s planar MoM engine rather uses a polar integration scheme, where the radial spectral variable k<sub>&rho;</sub> is integrated over the interval [0, Mk<sub>0</sub>], M being a large enough number to represent infinity, and the angular spectral variable t is integrated over the interval [0, 2&pi;]. You also saw some of the numerical [[parameters]] related to this spectral-domain integration scheme.
+
 
+
{{Note|When the observation plane is placed very close to the radiating J and M currents, the Green's functions exhibit singularities, which translate to very slow convergence or divergence of the integrals. You need to be careful to place field sensors at adequate distances from these radiating sources.}}
+
 
+
== Computing The Far Fields in Planar MoM ==
+
 
+
Unlike differential-based methods, MoM simulators do not need a radiation box to calculate the far field data. The far-zone fields are calculated directly by integrating the currents on the traces and across the embedded objects using the asymptotic form of the background structure’s dyadic Green's functions:
+
 
+
:<math> \mathbf{E^{ff}(r)} = \iiint_V \mathbf{ \overline{\overline{G}}_{EJ,ff}(r|r') \cdot J(r') } \, d\nu ' + \iiint_V \mathbf{ \overline{\overline{G}}_{EM,ff}(r|r') \cdot M(r') } \, d\nu '</math>
+
 
+
:<math> \mathbf{H^{ff}(r)} = \dfrac{1}{\eta_0} \mathbf{ \hat{r} \times E^{ff}(r) }</math>
+
<!--[[File:PMOM112.png]]-->
+
 
+
where &eta;<sub>0</sub> = 120&pi; is the characteristic impedance of the free space. As can be seen from the above equations, the far fields have the form of a TEM wave propagating in the radial direction away from the origin of coordinates. This means that the far-field magnetic field is always perpendicular to the electric field and the propagation vector, which in this case happens to be the radial unit vector in the spherical coordinate system. In other words, one only needs to know the far-zone electric field and can easily calculate the far-zone magnetic field from it. In [[EM.Cube]]'s mixed potential integral equation formulation, the far-zone electric field can be expressed in terms of the asymptotic form of the vector electric and magnetic potentials '''A''' and '''F''':
+
 
+
:<math>\mathbf{E^{ff}}(x,y,z) = j k_0 \eta_0 \hat{r} \times [\hat{r} \times \mathbf{A}(r \to \infty)] + j k_0 \hat{r} \times \mathbf{F}(r \to \infty)</math>
+
<!--[[File:PMOM113.png]]-->
+
 
+
The asymptotic form of these vector potentials are calculated using the &quot;'''Method of Stationary Phase'''&quot; when k<sub>0</sub>r &rarr; &infin;. In that case, one can use the approximation:
+
 
+
:<math> k_0 |\mathbf{r-r'}| \approx k_0 (r - \mathbf{\hat{r} \cdot r'}) </math>
+
<!--[[File:PMOM115.png]]-->
+
 
+
After applying the stationary phase method, one can extract the spherical wave factor exp(-jk<sub>0</sub>r)/r from the far-zone electric field, leaving the rest as functions of the spherical angles &theta; and &phi;. In other words, the far field is normalized to r, the distance from the field observation point to the origin. It is customary to express the far fields in spherical components E<sub>&theta;</sub> and E<sub>&phi;</sub>. Note that the outward propagating, TEM-type, far fields do not have radial components, i.e. E<sub>r</sub> = 0.
+
 
+
:<math> \mathbf{E_{\theta}}(\theta, \phi) = \cos\theta \cos\phi E_x + \cos\theta \sin\phi E_y - \sin\theta E_z </math>
+
:<math> \mathbf{E_{\phi}}(\theta, \phi) = -\sin\phi E_x + \cos\phi E_y </math>
+
<!--[[File:PMOM114.png]]-->
+
 
+
== Periodic Planar  MoM Simulation ==
+
 
+
In the case of an infinite periodic planar structure, the field equations can be written in the following form:
+
 
+
:<math> \mathbf{E(r) = E^{inc}(r)} +  \sum_{m=-\infty}^{\infty} \sum_{n=-\infty}^{\infty} \bigg[ \iiint_V \mathbf{ \overline{\overline{G}}_{EJ}(r|r') \cdot J_{mn}(r') } \, d\nu' +  \iiint_V \mathbf{ \overline{\overline{G}}_{EM}(r|r') \cdot M_{mn}(r') } \, d\nu' \bigg] </math>
+
 
+
 
+
:<math> \mathbf{H(r) = H^{inc}(r)} +  \sum_{m=-\infty}^{\infty} \sum_{n=-\infty}^{\infty} \bigg[ \iiint_V \mathbf{ \overline{\overline{G}}_{HJ}(r|r') \cdot J_{mn}(r') } \, d\nu' +  \iiint_V \mathbf{ \overline{\overline{G}}_{HM}(r|r') \cdot M_{mn}(r') } \, d\nu' \bigg] </math>
+
<!--[[File:PMOM94.png]]-->
+
 
+
where
+
 
+
:<math>\mathbf{J_{mn}(r) = J_{mn}}(x,y,z) = \mathbf{J_{00}}(x+m S_x, y+n S_y, z) e^{j(m k_{x00} S_x + n k_{y00} S_y)}</math>
+
 
+
:<math>\mathbf{M_{mn}(r) = M_{mn}}(x,y,z) = \mathbf{M_{00}}(x+m S_x, y+n S_y, z) e^{j(m k_{x00} S_x + n k_{y00} S_y)}</math>
+
 
+
and
+
 
+
:<math> -\infty < m, n < \infty </math>
+
<!--[[File:PMOM95(1).png]]-->
+
 
+
In the above equations, <math>\mathbf{J_{00}(r)}</math> and <math>\mathbf{M_{00}(r)}</math> are the periodic unit cell's electric and magnetic currents that are repeated everywhere in space on a rectangular lattice with periods S<sub>x</sub> and S<sub>y</sub> along the X and Y directions, respectively. <math>k_{x00}</math> and <math>k_{y00}</math> are the periodic propagation constants along the X and Y directions, respectively, and they are given by:
+
 
+
:<math> k_{x00} = k_0 \sin\theta \cos\phi </math>
+
 
+
:<math> k_{y00} = k_0 \sin\theta \sin\phi </math>
+
<!--[[File:PMOM96(1).png]]-->
+
 
+
where &theta; and &phi; are the beam scan angles in the case of periodic excitation of lumped sources, or they are the spherical angles of incidence in the case of a plane wave source illuminating the periodic structure. Using the infinite summations, one can define periodic dyadic Green's functions in the spectral domain in the following manner:
+
 
+
:<math> \mathbf{ \overline{\overline{G}}_{\mu \nu}^{PER} (r|r') } = \frac{1}{S_x S_y} \sum_{m=-\infty}^{\infty} \sum_{n=-\infty}^{\infty} \mathbf{ \tilde{\overline{\overline{G}}}_{\mu \nu} } (k_x, k_y, z|z') e^{-j[k_{xm}(x-x') + k_{yn}(y-y')]} </math>
+
 
+
where
+
:<math> k_{xm} = k_{x00} + \frac{2\pi m}{S_x} \quad \text{and} \quad k_{ym} = k_{y00} + \frac{2\pi m}{S_y} </math>
+
<!--[[File:PMOM97.png]]-->
+
 
+
The above doubly infinite periodic Green's functions are said to be expressed in terms of &quot;Floquet Modes&quot;. The exact formulation involves an infinite set of these periodic Floquet modes. During the MoM matrix fill process for a periodic structure, a finite number of Floquet modes are calculated. By default, [[EM.Cube]]'s planar MoM engine considers M<sub>x</sub> = M<sub>y</sub> = 25. This implies a total of 51 modes along the X direction and a total of 51 modes along the Y direction, or a grand total of 51<sup>2</sup> = 2,601 Floquet modes. You can increase the number of Floquet modes for your project from the Planar MoM Engine Settings Dialog. In the section titled &quot;Periodic Simulation&quot;, you can change the values of '''Number of Floquet Modes''' in the two boxes designated X and Y.
+
 
+
== EM.Picasso's Linear System Solvers ==
+
 
+
After the MoM impedance matrix '''[Z]''' (not to be confused with the impedance parameters) and excitation vector '''[V]''' have been computed through the matrix fill process, the planar MoM simulation engine is ready to solve the system of linear equations:
+
 
+
:<math> \mathbf{[Z]}_{N\times N} \cdot \mathbf{[I]}_{N\times 1} = \mathbf{[V]}_{N\times 1} </math>
+
<!--[[File:PMOM81.png]]-->
+
 
+
where '''[I]''' is the solution vector, which contains the unknown amplitudes of all the basis functions that represent the unknown electric and magnetic currents of finite extents in your planar structure. In the above equation, N is the dimension of the linear system and equal to the total number of basis functions in the planar mesh. [[EM.Cube]]'s linear solvers compute the solution vector'''[I]''' of the above system. You can instruct [[EM.Cube]] to write the MoM matrix and excitation and solution vectors into output data files for your examination. To do so, check the box labeled &quot;'''Output MoM Matrix and Vectors'''&quot; in the Matrix Fill section of the Planar MoM Engine Settings dialog. These are written into three files called mom.dat1, exc.dat1 and soln.dat1, respectively.
+
 
+
There are a large number of numerical methods for solving systems of linear equations. These methods are generally divided into two groups: direct solvers and iterative solvers. Iterative solvers are usually based on matrix-vector multiplications. Direct solvers typically work faster for matrices of smal to medium size (N&lt;3,000). [[EM.Cube]]'s [[Planar Module]] offers five linear solvers:
+
 
+
# LU Decomposition Method
+
# Biconjugate Gradient Method (BiCG)
+
# Preconditioned Stabilized Biconjugate Gradient Method (BCG-STAB)
+
# Generalized Minimal Residual Method (GMRES)
+
# Transpose-Free Quasi-Minimum Residual Method (TFQMR)
+
 
+
Of the above list, LU is a direct solver, while the rest are iterative solvers. BiCG is a relatively fast iterative solver, but it works only for symmetric matrices. You cannot use BiCG for periodic structures or planar structures that contain both metal and slot traces at different planes, as their MoM matrices are not symmetric. The three solvers BCG-STAB, GMRES and TtFQMR work well for both symmetric and asymmetric matrices and they also belong to a class of solvers called '''Krylov Sub-space Methods'''. In particular, the GMRES method always provides guaranteed unconditional convergence.
+
 
+
[[EM.Picasso]], by default, provides a &quot;'''Automatic'''&quot; solver option that picks the best method based on the settings and size of the numerical problem. For linear systems with a size less than N = 3,000, the LU solver is used. For larger systems, BiCG is used when dealing with symmetric matrices, and GMRES is used for asymmetric matrices. If the size of the linear system exceeds N = 15,000, the sparse version of the iterative solvers is used, utilizing a row-indexed sparse storage scheme. You can override the automatic solver option and manually set you own solver type. This is done using the '''Solver Type''' drop-down list in the &quot;'''Linear System Solver'''&quot; section of the Planar MoM Engine Settings dialog. There are also a number of other parameters related to the solvers. The default value of '''Tolerance of Iterative Solver''' is 1E-3, which can be increased for more ill-conditioned systems. The maximum number of iterations is usually expressed as a multiple of the systems size. The default value of '''Max No. of Solver Iterations / System Size''' is 3. For extremely large systems, sparse versions of iterative solvers are used. In this case, the elements of the matrix are thresholded with respect to the larges element. The default value of '''Threshold for Sparse Solver''' is 1E-6, meaning that all the matrix elements whose magnitude is less than 1E-6 times the large matrix elements are set equal to zero. There are two more parameters that are related to the Automatic Solver option. These are &quot;''' User Iterative Solver When System Size &gt;'''&quot; with a default value of 3,000 and &quot;''' Use Sparse Storage When System Size &gt;''' &quot; with a default value of 15,000. In other words, you control the automatic solver when to switch between direct and iterative solvers and when to switch to the sparse version of iterative solvers.
+
 
+
==  Free-Space Green’s Function ==
+
 
+
The Green’s functions are the analytical solutions of boundary value problems when they are excited by an elementary source. This is usually an infinitesimally small vectorial point source. In order for the Green’s functions to be computationally useful, they must have analytical closed forms. This can be a mathematical expression or a more complex recursive process. It is no surprise that only very few electromagnetic boundary value problems have closed-form Green’s functions. The total electric ('''E''') field can be expressed in terms of the electric current in the following way:
+
 
+
:<math> \mathbf{E = E^{inc}} +  \mathbf{\iiint_V \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') } d \nu' +  \mathbf{\iiint_V \overline{\overline{G}}_{EM}(r|r') \cdot M(r') } d \nu' </math>
+
 
+
 
+
:<math> \mathbf{H = H^{inc}} +  \mathbf{\iiint_V \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') } d \nu' +  \mathbf{\iiint_V \overline{\overline{G}}_{HM}(r|r') \cdot M(r') } d \nu' </math>
+
<!--[[File:PMOM1(1).png]]-->
+
 
+
where is the dyadic Green’s functions for electric fields due to electric current sources and '''E<sup>i</sup>''' is the incident or impressed electric field. The incident or impressed field provides the excitation of the structure. It may come from an incident plane wave or a gap source on a line, etc. The simplest background structure is the unbounded free space, which is represented by the following Green’s function:
+
 
+
:<math> \mathbf{ \overline{\overline{G}}_{EJ}(r|r') = (\overline{\overline{I}} + \nabla\nabla) } G_{\Lambda} (\mathbf{r|r'}), \quad G_{\Lambda} (\mathbf{r|r'}) = \frac{ e^{-jk_0 \mathbf{|r-r'|}} }{ 4\pi \mathbf{|r-r'|} } </math>
+
<!--[[File:03_freespace_tn.gif]]-->
+
 
+
where <math>\mathbf{\overline{\overline{I}}}</math> is the unit dyad, <math>\nabla</math> is the gradient operator, '''r''' and '''r'''' are the position vectors of the observation and source points, respectively, and k<sub>0</sub> is the free-space propagation constant. This implies that electromagnetic waves propagate in free space in a spherical form away from the source. Note that the Green’s function has a singularity at the source, i.e. when '''r''' = '''r''''. This singularity must be removed when solving the integral equations.
+
 
+
==  3D Integral Equations ==
+
 
+
In the more general formulation of the field integration equations, both electric and magnetic currents are included. In that case, the total electric and magnetic fields are given by the following equations:
+
 
+
:<math> \mathbf{E = E^{i}} +  \mathbf{\iiint_V \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') } +  \mathbf{\iiint_V \overline{\overline{G}}_{EM}(r|r') \cdot M(r') } </math>
+
 
+
 
+
:<math> \mathbf{H = H^{i}} +  \mathbf{\iiint_V \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') } +  \mathbf{\iiint_V \overline{\overline{G}}_{HM}(r|r') \cdot M(r') } </math>
+
<!--[[File:image001_tn.gif]]-->
+
 
+
The above coupled equations involve four types of dyadic Green's functions that represent the electric and magnetic field radiated by an electric or a magnetic current. The incident or impressed electric and magnetic fields Ei and Hi exist independently of the given structures and are related to each other depending on the type of excitation source.
+
 
+
Enforcing the boundary conditions on the integral definitions of the '''E''' and '''H''' fields results in a system of integral equations as follows:
+
 
+
:<math> \mathcal{L}_E(\mathbf{E}) = \mathcal{L}_E \left( \mathbf{E^{i}} +  \mathbf{\iiint_V \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') } +  \mathbf{\iiint_V \overline{\overline{G}}_{EM}(r|r') \cdot M(r') } \right) = 0 </math>
+
 
+
 
+
:<math> \mathcal{L}_H(\mathbf{H}) = \mathcal{L}_H \left( \mathbf{H^{i}} +  \mathbf{\iiint_V \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') } +  \mathbf{\iiint_V \overline{\overline{G}}_{HM}(r|r') \cdot M(r') } \right) = 0 </math>
+
<!--[[File:image016_tn.gif]]-->
+
 
+
where <math>\mathcal{L}_E(E)</math> is the boundary value operator for the electric field and <math>\mathcal{L}_H(H)</math> is the boundary value operator for the magnetic field. For example, they may require that the tangential components the '''E''' field vanish on perfect electric conductors. Or they may require that the tangential components the '''E''' and '''H''' fields be continuous across an aperture in a perfect ground plane. Given the fact that the dyadic Green’s functions and the incident or impressed fields are all known, one can solve the above system of integral equations to find the unknown currents '''J''' and '''M'''. Therefore, through these relationships you can easily cast the above integral equations in terms of unknown '''E''' and '''H''' fields.
+
 
+
== Galerkin Testing ==
+
 
+
The integral equation derived in the previous section can be solved numerically by discretizing the computational domain using a proper meshing scheme. The original functional equation is reduced to a set of discretized linear algebraic equations over elementary cells. The unknown quantities are found by solving this system of linear equations, and many other parameters can be computed thereafter. This method of numerical solution of integral equations is known as the Method of Moments (MoM). In this method, the unknown electric current is represented by an expansion of basis functions as follows:
+
 
+
:<math> \mathbf{J(r)} = \sum_{n=1}^N {I_n}^{(J)} \mathbf{ {f_n}^{(J)}(r) }</math>
+
<!--[[File:07_numerical-solutions_tn.gif]]-->
+
 
+
where <math>\mathbf{ {f_n}^{(J)} }</math> are the generalized vector basis functions for the expansion of electric currents, and <math>{I_n}^{(J)}</math> are the unknown complex amplitudes of these basis functions, which have to be determined. Substituting these expansions yields the following discretized integral equation:
+
 
+
:<math> \mathcal{L}_E \left( \mathbf{E^i} +\iiint_V \mathbf{ \overline{\overline{G}}_{EJ}(r|r') }  \cdot \sum_{n=1}^N {I_n}^{(J)} \mathbf{ {f_n}^{(J)}(r') } \, d\nu' \right) = 0 </math>
+
<!--[[File:10_numerical-solution_tn.gif]]-->
+
 
+
In order to solve the above equation, the method of moments uses Galerkin's technique to turn it into a set of linear algebraic equations. This is accomplished by testing the above equations using the basis functions, leading to the following linear system:
+
 
+
:<math>\mathbf{[Z] \cdot [I] = [V]}</math>
+
<!--[[File:11_numerical-solution_tn.gif]]-->
+
 
+
where
+
 
+
:<math> Z_{ij} = \iiint_{V_i} \mathbf{ {f_i}^{(J)}(r) } \, d\nu \cdot  \iiint_{V_j} \mathbf{ \overline{\overline{G}}_{EJ}(r|r') \cdot {f_j}^{(J)}(r') } \, d\nu' </math>
+
<!--[[File:12_numerical-solution_tn.gif]]-->
+
 
+
and
+
 
+
:<math> V_i = \iiint_{V_i} \mathbf{ {f_i}^{(J)}(r) \cdot E^i(r) } \, d\nu </math>
+
<!--[[File:13_numerical-solution_tn.gif]]-->
+
 
+
Using a rooftop expansion of the currents on the wires, we can discretize the Pocklington integral equation. In order to convert the discretized integral equation into a system of linear system of algebraic equations, we use Galerkin’s testing process, in which the testing functions are chosen to be identical to the expansion basis functions. However, to avoid the source singularity at r=r’, the expansion functions are placed at the center of the wires, while the test functions are evaluated on the surface of the wires, assuming a finite non-zero radius for all wires. The solution vector [I] is then found as:
+
 
+
:<math>\mathbf{[I] = [Z]^{-1} \cdot [V] } </math>
+
<!--[[File:24_galerkin_tn.gif]]-->
+
 
+
where [Z]<sup>-1</sup> is the inverse of the impedance matrix and [V] is the excitation vector.
+
 
+
==  Pocklington’s Integral Equations for Wire Structures ==
+
 
+
Wire structures are made of linear PEC elements. These may consist of actual physical wires such as a dipole or loop antenna or a wireframe representation of a surface or solid object. In a wire structure, the unknown electric currents are one-dimensional. The integral equation is derived by forcing the tangential component of the electric field to vanish on the surface of the wire. This leads to the following simpler integral equation:
+
 
+
:<math> \mathbf{ \hat{I} \cdot E^i } - jk_0 Z_0 \int_C \left( G_A \mathbf{(r|r')} I(l') \mathbf{ \hat{l} \cdot \hat{l}' } + \frac{1}{{k_0}^2} \frac{\partial G_A}{\partial l} \frac{\partial I}{\partial l'} \right) \, dl' = 0 </math>
+
<!--[[File:14_pocklingtons_tn.gif]]-->
+
 
+
where G<sub>A</sub> is the free space Green’s function, I(l) is the unknown linear current in the wire and C is the contour of the wire.  and <math>\hat{l}'</math> are the unit vectors along the wire contour. Note that G<sub>A</sub> has a singularity when r = r’, which must be either removed or avoided as will be explained later.
+
 
+
==  Discretization Of Wire Structures ==
+
 
+
The right choice of the basis functions that are used to represent the elementary currents is very important. It will determine the accuracy and computational efficiency of the resulting numerical solution. Rooftop basis functions are one of the more popular types of basis functions used in a variety of MoM formulations. The simplest rooftop function is the one-dimensional triangular functions defined as in the figure below:
+
 
+
[[File:18_meshing_tn.gif]]
+
 
+
This function provides a linear interpolation of the unknown currents or fields in one dimension. Note that the function vanishes at it two ends. This is a desirable feature for basis functions that represent electric currents on metallic wires as the current must vanish at the two ends of a wire. The total current on the wire can be approximated in a linear fashion by a set of one-dimensional rooftop functions as shown in the figure below:
+
 
+
[[File:19_meshing_tn.gif]]
+
 
+
This can be written as
+
 
+
:<math> I(l) = \sum_{n=1}^N a_n f_n(l) \mathbf{\hat{s}_n} </math>
+
<!--[[File:20_meshing_tn.gif]]-->
+
 
+
where l is the length coordinate along the wire with l=0 at its start point. <math>f_n(l)</math> is the scaled and translated version of the linear basis function <math>f(l)</math> shown in the previous figure. <math>\mathbf{\hat{s}_n}</math> is the unit vector along wire.
+
 
+
== Conventional Physical Optics (GO-PO) ==
+
 
+
[[Image:po_manual_1.png|thumb|500px|A diagram showing a scatterer lit by a source.]]
+
The following analysis assumes a general impedance surface. To treat an object with an arbitrary geometry using PO, the object is first decomposed into many small elementary patches or cells, which have a simple geometry such as a rectangle or triangle. Then, using the tangent plane approximation, the electric and magnetic surface currents, '''J(r)''' and '''M(r)''', on the lit region of the scatterer are approximated by:
+
 
+
:<math> \mathbf{J(r)} = (1+\alpha) \mathbf{\hat{n} \times H(r)} </math>
+
 
+
:<math> \mathbf{M(r)} = -(1-\alpha) \mathbf{\hat{n} \times E(r)} </math>
+
<!--[[File:PO1(1).png]]-->
+
 
+
where '''E(r)''' and '''H(r)''' are the incident electric and magnetic fields on the object and '''n''' is the local outward normal unit vector as shown in the figure below. a is a parameter related to the impedance Z of the surface (expressed in Ohms), which is defined in the following way:
+
 
+
:<math> \alpha = \frac{1-Z/\eta_0}{1+Z/\eta_0} </math>
+
<!--[[File:PO2.png]]-->
+
 
+
where <math>\eta_0 = 120\pi \; \Omega</math> is the intrinsic impedance of the free space. Then, the electric and magnetic currents reduce to:
+
 
+
:<math> \mathbf{J(r)} = \frac{2\eta_0}{\eta_0 + Z} \mathbf{\hat{n} \times H(r)} </math>
+
 
+
:<math> \mathbf{M(r)} = - \frac{2Z}{\eta_0 + Z} \mathbf{\hat{n} \times E(r)} </math>
+
<!--[[File:PO3.png]]-->
+
 
+
 
+
Two limiting cases of an impedance surface are perfect electric conductor (PEC) and perfect magnetic conductor (PMC) surface. For a PEC surface, Z = 0,  &alpha; = 1, and one can write:
+
 
+
:<math> \mathbf{J(r)} = 2 \mathbf{\hat{n} \times H(r)} </math>
+
 
+
:<math> \mathbf{M(r)} = 0 </math>
+
<!--[[File:PO4.png]]-->
+
 
+
while for a PMC surface, Z = &infin;,  &alpha; = -1, and one can write:
+
 
+
:<math> \mathbf{J(r)} = 0 </math>
+
 
+
:<math> \mathbf{M(r)} = -2 \mathbf{\hat{n} \times E(r)} </math>
+
<!--[[File:PO5.png]]-->
+
 
+
 
+
Another special case is a Huygens surface with equivalent electric and magnetic surface currents. In that case, Z =  &eta;<sub>0</sub>, &alpha; = 0, and one can write:
+
 
+
:<math> \mathbf{J(r) = \hat{n} \times H(r)} </math>
+
 
+
:<math> \mathbf{M(r) = -\hat{n} \times E(r)} </math>
+
<!--[[File:PO10.png]]-->
+
 
+
 
+
A major difficulty encountered in determining the PO currents of the scatterer is identification of lit and shadowed facets. Determination of lit and shadowed regions for simple, stand-alone, convex objects is rather simple. Denoting the incidence direction from a source to a point on the scatterer by the unit vector '''k''', the point is considered lit if '''n.k'''&lt; 0, and shadowed if '''n.k'''&gt; 0. These conditions, however, are only valid if there is a direct line of sight (LOS) between the source and the centroid of the cell under consideration. They cannot predict if there are any obstructing objects in the path of the incident beam or ray. For simple convex objects, a Geometrical Optics (GO) approach can be used to finds the optical LOS lines and determine the lit and shadowed areas on the object. The conventional PO can then be used to find the electric and magnetic surface currents.
+
 
+
== Calculating Near &amp; Far Fields In PO ==
+
 
+
Once the electric and magnetic surface currents are determined in the lit regions of the scatterer(s), they act as secondary sources and radiate into the free space. These secondary fields are the scattered fields that are superposed with the primary incident fields. The near fields at every point '''r''' in space are calculated from:
+
 
+
 
+
:<math> \mathbf{ E(r) = E^{inc}(r) } +  \iint_{S_J} \mathbf{ \overline{\overline{G}}_{EJ}(r|r') \cdot J(r') } ds' +  \iint_{S_M} \mathbf{ \overline{\overline{G}}_{EM}(r|r') \cdot M(r') } ds' </math>
+
 
+
:<math> \mathbf{ H(r) = H^{inc}(r) } +  \iint_{S_J} \mathbf{ \overline{\overline{G}}_{HJ}(r|r') \cdot J(r') } ds' +  \iint_{S_M} \mathbf{ \overline{\overline{G}}_{HM}(r|r') \cdot M(r') } ds' </math>
+
<!--[[File:PO6.png]]-->
+
 
+
 
+
where '''G<sub>EJ</sub>''', '''G<sub>EM</sub>''', '''G<sub>HJ</sub>''', '''G<sub>HM</sub>''' are the dyadic Green's functions of electric and magnetic fields due to electric and magnetic currents, respectively. In [[EM.Cube]]'s [[PO Module]], the background structure is the free space. Therefore, all these dyadic Green's functions reduce to the simple free-space Green's function of the form <math>\exp(-jk_0r)/(4\pi r)</math> and the near fields reduce to:
+
 
+
 
+
:<math> \begin{align} \mathbf{ E(r) = E^{inc}(r) }  & - jk_0 Z_0 \iint_{S_J} \left\{ \left[ 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right] \mathbf{J(r')} -  \left[ 1 - \frac{3j}{k_0 R} - \frac{3}{(k_0 R)^2} \right] \mathbf{ (\hat{R} \cdot J(r')) \hat{R} } \right\} \frac{e^{-jk_0 R}}{4\pi R} ds' \\ & + jk_0 \iint_{S_M} \left[ 1-\frac{j}{k_0 R} \right] \mathbf{ (\hat{R} \times M(r')) } \frac{e^{-jk_0 R}}{4\pi R} ds' \end{align} </math>
+
 
+
 
+
:<math> \begin{align} \mathbf{ H(r) = H^{inc}(r) }  & - jk_0 Y_0 \iint_{S_M} \left\{ \left[ 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right] \mathbf{M(r')} -  \left[ 1 - \frac{3j}{k_0 R} - \frac{3}{(k_0 R)^2} \right] \mathbf{ (\hat{R} \cdot M(r')) \hat{R} } \right\} \frac{e^{-jk_0 R}}{4\pi R} ds' \\ & - jk_0 \iint_{S_J} \left[ 1-\frac{j}{k_0 R} \right] \mathbf{ (\hat{R} \times J(r')) } \frac{e^{-jk_0 R}}{4\pi R} ds' \end{align} </math>
+
<!--[[File:PO7.png]]-->
+
 
+
 
+
where <math> R=|r-r'| \text{, } k_0 = \frac{2\pi}{\lambda_0} \text{ and } Z_0 = 1/Y_0 = \eta_0 </math>.
+
 
+
 
+
When k<sub>0</sub>r &gt;&gt; 1, i.e. in the far-zone field of the scatterer, one can use the asymptotic form of the Green's functions and evaluate the radiation integrals using the stationary phase method to obtain far-field expressions for the electric and magnetic fields as follows:
+
 
+
 
+
:<math> \mathbf{E^{ff}(r)} = \frac{jk_0 e^{-jk_0 r}}{4\pi r}  \left\{ Z_0 \mathbf{ \hat{r} \times \hat{r} } \times \iint_{S_J} \mathbf{J(r')} e^{-jk_0 \mathbf{\hat{r}\cdot r'}} ds' + \mathbf{\hat{r}} \times \iint_{S_M} \mathbf{M(r')} e^{-jk_0 \mathbf{ \hat{r} \cdot r' } } ds' \right\} </math>
+
 
+
:<math> \mathbf{H^{ff}(r)} = \frac{1}{Z_0} \mathbf{\hat{r} \times E^{ff}(r)} </math>
+
<!--[[File:PO8.png]]-->
+
 
+
== Iterative Physical Optics (IPO) ==
+
 
+
The induced electric and magnetic surface currents on each point of the scatterer object can be calculated from the Magnetic and Electric Field Integral Equations (MFIE &amp; EFIE):
+
 
+
 
+
:<math> \mathbf{J(r)} = (1+\alpha)\mathbf{\hat{n}} \times \left\lbrace  \begin{align} & \mathbf{ H^{inc}(r) } - jk_0 \iint_{S_J} \left( 1 - \frac{j}{k_0 R} \right)  (\mathbf{ \hat{R} \times J(r') }) \frac{e^{-jk_0 R}}{4\pi R} \,ds' \\ & -j k_0 Y_0 \iint_{S_M} \left[ \left( 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right) \mathbf{M(r')} - \left( 1-\frac{3j}{k_0 R}-\frac{3}{(k_0 R)^2} \right) \mathbf{ (\hat{R} \cdot M(r')) \hat{R} } \right] \frac{e^{-jk_0 R}}{4\pi R} \,ds' \end{align} \right\rbrace </math>
+
 
+
 
+
:<math> \mathbf{M(r)} = -(1-\alpha)\mathbf{\hat{n}} \times \left\lbrace  \begin{align} & \mathbf{ E^{inc}(r) } + jk_0 \iint_{S_M} \left( 1 - \frac{j}{k_0 R} \right) (\mathbf{ \hat{R} \times M(r') }) \frac{e^{-jk_0 R}}{4\pi R} \,ds' \\ & -j k_0 Z_0 \iint_{S_J} \left[ \left( 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right) \mathbf{J(r')} - \left( 1-\frac{3j}{k_0 R}-\frac{3}{(k_0 R)^2} \right) \mathbf{ (\hat{R} \cdot J(r')) \hat{R} } \right] \frac{e^{-jk_0 R}}{4\pi R} \,ds' \end{align} \right\rbrace </math>
+
<!--[[File:PO9(1).png]]-->
+
 
+
where '''R''' ='''r''' - '''r'''', R = |'''R'''|, and
+
 
+
:<math>\mathbf{ \hat{R} = \frac{R}{|R|} = \frac{r-r'}{|r-r'|} }</math>
+
<!--[[File:PO11.png]]-->
+
 
+
The shadowing phenomenon can indeed be attributed to near-field interaction of surface currents. The current on the lit region produces a scattered field in the forward direction that is almost equal and out of phase with the incident wave. Hence, the sum of the scattered field and incident field over the shadowed region almost cancel each other, giving rise to a very small field there. This suggests that keeping track of multiple scattering can take care of shadowing problems automatically. In addition, the effects of multiple scattering can be readily accounted for by an iterative PO approach to be formulated next.
+
 
+
The starting point for the iterative PO solution is the above MFIE and EFIE integral equations. To the first (zero-order) approximation, we can write
+
 
+
:<math> \begin{align} & \mathbf{J^{(0)}(r)} = (1+\alpha) \mathbf{ \hat{n} \times H^{inc}(r) } \\ & \mathbf{M^{(0)}(r)} = -(1-\alpha) \mathbf{ \hat{n} \times E^{inc}(r) } \end{align} </math>
+
<!--[[File:PO13.png]]-->
+
 
+
which are the conventional PO currents. However, this approximation does not formally recognize the lit and shadowed areas. Instead of identifying the exact boundaries of the lit and shadowed areas over a complex target, a simple condition is used first to find the primary shadowed areas. Then, through PO iterations all shadowed areas are determined automatically. When calculating the field on the scatterer for every source point, a primary shadowing condition given by '''n.k'''&lt; 0 is examined. In complex scatterer geometries, there are shadowed points in concave regions where '''n.k'''&gt; 0, but the correct shadowing is eventually achieved through the iteration of the PO currents. Therefore, in computation of the above equations, only the contribution of the points that satisfy the following condition are considered:
+
 
+
:<math>\mathbf{ \hat{n} \cdot \hat{R}} < 0 \quad \text{or} \quad \mathbf{\hat{n} \cdot (r-r')} < 0</math>
+
<!--[[File:PO12.png]]-->
+
 
+
At the subsequent iterations, the higher order PO currents are given by;
+
 
+
[[File:PO14(1).png]]
+
 
+
For most practical applications, iterations up to the second order is sufficient. The iterative solution will not only account for double-bounce scattering over the lit regions but it also removes the lower order currents erroneously placed over concave shadowed areas.
+
 
+
== General Huygens Sources ==
+
 
+
According to the electromagnetic equivalence theorem, if we know the tangential components of E and H fields on a closed surface, we can determine all the E and H fields inside and outside that surface in a unique way. Such a surface is called a Huygens surface. At the end of a full-wave FDTD or MoM solution, all the electric and magnetic fields are known everywhere in the computational domain. We can therefore define a box around the radiating (source) structure, over which we can record the tangential E and H field components. The tangential field components are then used to define equivalent electric and magnetic surface currents over the Huygens surface as:
+
 
+
:<math> \begin{align} & \mathbf{ J(r) = \hat{n} \times H(r) } \\ & \mathbf{ M(r) = -\hat{n} \times E(r) } \end{align} </math>
+
 
+
In the physical optics domain, the known equivalent electric and magnetic surface currents (or indeed the known tangential E and H field components) over a given closed surface S can be used to find reradiated electric and magnetic fields everywhere in the space as follows:
+
 
+
:<math> \mathbf{E^{inc}(r)} = -jk_0 \sum_j \iint_{\Delta_j} \, ds' \frac{e^{-jk_0 R}}{4\pi R} \left\lbrace \begin{align} & Z_0 \left[ 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right] \mathbf{J_j(r')} \\ & -Z_0 \left[ 1 - \frac{3j}{k_0 R} - \frac{3}{(k_0 R)^2} \right] \mathbf{ (\hat{R} \cdot J_j(r')) \hat{R} } \\ & - \left[ 1 - \frac{j}{k_0 R} \right] \mathbf{ (\hat{R} \times M_j(r')) } \end{align} \right\rbrace </math>
+
 
+
 
+
:<math> \mathbf{H^{inc}(r)} = -jk_0 \sum_j \iint_{\Delta_j} \, ds' \frac{e^{-jk_0 R}}{4\pi R} \left\lbrace \begin{align} & Y_0 \left[ 1 - \frac{j}{k_0 R} - \frac{1}{(k_0 R)^2} \right] \mathbf{M_j(r')} \\ & -Y_0 \left[ 1 - \frac{3j}{k_0 R} - \frac{3}{(k_0 R)^2} \right] \mathbf{ (\hat{R} \cdot M_j(r')) \hat{R} } \\ & + \left[ 1 - \frac{j}{k_0 R} \right] \mathbf{ (\hat{R} \times J_j(r')) } \end{align} \right\rbrace </math>
+
 
+
 
+
where the summation over index ''j'' is carried out for all the elementary cells &Delta;<sub>j</sub> that make up the Huygens box. In [[EM.Cube]] Huygens surfaces are cubic and are discretized using a rectangular mesh. Therefore, &Delta;<sub>j</sub> represents any rectangular cell located on one of the six faces of Huygens box. Note that the calculated near-zone electric and magnetic fields act as incident fields for the scatterers in your EM.Ilumina project. The Huygens source data are normally generated in one of [[EM.Cube]]'s full-wave computational modules like FDTD, Planar or MoM3D. Keep in mind that the fields scattered (or reradiated) by your physical structure do not affect the fields inside the Huygens source.   
+
 
+
The far fields of the Huygens surface currents are calculated from the following relations:
+
 
+
 
+
:<math> \mathbf{E^{ff}(r)} = \frac{jk_0}{4\pi} \frac{e^{-jk_0 r}}{r} \sum_j \iint_{\Delta_j} \left[ Z_0 \, \mathbf{ \hat{r} \times \hat{r} \times J_j(r') } +  \mathbf{ \hat{r} \times M_j(r') } \right] e^{ -jk_0 \mathbf{\hat{r} \cdot r'} } \, ds' </math>
+
 
+
:<math>\mathbf{H^{ff}(r)} = \frac{1}{Z_0} \mathbf{\hat{r} \times E^{ff}(r)} </math>
+
  
 
<br />
 
<br />
Line 841: Line 32:
 
<hr>
 
<hr>
  
[[Image:Top_icon.png|48px]] '''[[#Maxwell's Equations in Differential Form | Back to the Top of the Page]]'''
+
[[Image:Back_icon.png|30px]] '''[[EM.Cube | Back to EM.Cube Main Page]]'''
 
+
[[Image:Back_icon.png|40px]] '''[[EM.Cube | Back to EM.Cube Main Page]]'''
+

Latest revision as of 02:38, 17 June 2018

Maxwell1.png
Cube-icon.png Cad-ico.png Fdtd-ico.png Prop-ico.png Static-ico.png Planar-ico.png Metal-ico.png Po-ico.png


The following links provide a basic review of electromagnetic theory and the various numerical techniques used by EM.Cube's simulation engines:

Basic Electromagnetic Theory

The Far-Field Approximation for Radiation & Scattering Problems

Basic Principles of The Finite Difference Time Domain Method

Basic Principles of SBR Ray Tracing

Basic Principles of The Method of Moments

Basic Principles of Physical Optics

Basic Principles of Electrostatic & Magnetostatic Field Analysis

Basic Principles of Steady-State Thermal Analysis


Attention icon.png The above pages contain a large number of mathematical expressions and equations. Sometimes, the mathematical notations might not load properly the first time you open a page. In that case, you need to use the "reload" or "update" feature of your web browser to refresh the page, or return to the gateway page and try to open the link one more time.



Back icon.png Back to EM.Cube Main Page