-
Notifications
You must be signed in to change notification settings - Fork 68
/
fdtd-near-to-far.tex
1368 lines (1285 loc) · 58.7 KB
/
fdtd-near-to-far.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
\chapter{Near-to-Far-Field Transformation \label{chap:nfffTrans}}
\renewcommand{\thefootnote}{\fnsymbol{footnote}}
\footnotetext{Lecture notes by John Schneider. {\tt
fdtd-near-to-far.tex}}
\section{Introduction}
As we have seen, the FDTD method provides the fields throughout some
finite region of space, i.e., the fields throughout the computational
domain. However, in practice, we are often interested in the fields
far away from the region we have modeled. For example, an FDTD
implementation may have modeled an antenna or some scatterer. But,
the fields in the immediate vicinity of that antenna or scatterer may
not be the primary concern. Rather, the distant or ``far'' fields may
be the primary concern. In this chapter we show how the near fields,
which are essentially the fields within the FDTD grid, can be used to
obtain the far fields. We start with a brief review of the underlying
theory that pertains in the continuous world and then discuss the
implementation details for the FDTD method.
\section{The Equivalence Principle}
Recall the boundary conditions that pertain to the electric and
magnetic fields tangential to an interface:
\begin{eqnarray}
\nunit\times(\Evec_1 - \Evec_2) &=& -\Mvec_s, \label{eq:eBoundary}\\
\nunit\times(\Hvec_1 - \Hvec_2) &=& \Jvec_s, \label{eq:hBoundary}
\end{eqnarray}
where $\nunit$ is normal to the interface, pointing toward region
$1$. The subscript $1$ indicates the fields immediately adjacent to
one side of the interface and the subscript $2$ indicates the fields
just on the other side of the interface. The ``interface'' can either
be a physical boundary between two media or a fictitious boundary with
the same medium to either side. The current $\Mvec_s$ is a magnetic
surface current, i.e., a current that only flows tangential to the
interface. In practice there is no magnetic charge and thus no
magnetic current. Therefore \refeq{eq:eBoundary} states that the
tangential components of $\Evec$ must be continuous across the
boundary. However, in theory, we can imagine a scenario where the
tangential fields are discontinuous. If this were the case, the
magnetic current $\Mvec_s$ must be non-zero to account for this
discontinuity. In a little while we will see why it is convenient to
envision such a scenario. The current $\Jvec_s$ in \refeq{eq:hBoundary}
is the usual electric surface current.
As depicted in Fig.\ \ref{fig:equivalence}(a), consider a space in
which there is a source or scatterer that radiates (or scatters) some
fields. We can define a fictitious boundary that surrounds this
source or scatterer. Let us then imagine that the fields exterior to
this boundary are unchanged but the fields interior to the boundary
are set to zero as depicted in Fig.\ \ref{fig:equivalence}(b). By
setting the fields interior to the boundary to zero, we will create
discontinuities in the tangential components on either side of the
fictitious boundary. These discontinuities are perfectly fine provided
we account for them by having the appropriate surface currents flow
over the boundary. These currents are given by
\refeq{eq:eBoundary} and \refeq{eq:hBoundary} where the fields in
region $2$ are now zero. Thus,
\begin{eqnarray}
\Mvec_s &=& -\nunit\times\Evec_1, \\
\Jvec_s &=& \nunit\times\Hvec_1.
\end{eqnarray}
\begin{figure}
\begin{center}
\epsfig{width=2.65in,file=Figures/Fdtd-near-to-far/radiation-cartoon.eps}\\
(a)\\
\epsfig{width=2.65in,file=Figures/Fdtd-near-to-far/radiation-cartoon-b.eps}\\
(b)\\
\epsfig{width=2.65in,file=Figures/Fdtd-near-to-far/radiation-cartoon-c.eps}\\
(c)
\end{center}
\caption{(a) A space containing a source or scatterer that is
surrounded by a fictitious boundary which is indicated by the dashed
line. The fields are continuous across this boundary. (b) The
fields are set to zero within the boundary. Surface currents must
be used to account for the discontinuity across the boundary. (c)
Since the fields are zero within the boundary, any inhomogeneities
within the boundary can be discarded.} \label{fig:equivalence}
\end{figure}
As you may recall and as will be discussed in further detail below, it
is fairly simple to find the fields radiated by a current (whether
electric or magnetic) when that current is radiating in a homogeneous
medium. Unfortunately, as shown in Fig.\ \ref{fig:equivalence}(b),
the surface currents are not radiating in a homogeneous medium. But,
in fact, since the fields within the fictitious boundary are zero, we
can place anything, or nothing, within the boundary and that will have
no effect on the fields exterior to the boundary. So, let us maintain
the same surface currents but discard any inhomogeneity that were
within the boundary. That leaves a homogeneous region as depicted in
Fig.\ \ref{fig:equivalence}(c) and it is fairly straightforward to find
these radiated fields.
\section{Vector Potentials}
Before proceeding further, let us briefly review vector potentials.
First, consider the case (which corresponds to the physical world) where
there is no magnetic charge---electric currents can flow but magnetic
currents cannot. Thus,
\begin{equation}
\nabla\cdot \Bvec_A = \nabla\cdot \mu \Hvec_A = 0 \label{eq:divBA}
\end{equation}
where the subscript $A$ indicates we are considering the case of no
magnetic charge. There is a vector identity that the divergence of
the curl of any vector field is identically zero. Therefore
\refeq{eq:divBA} will automatically be satisfied if we write
\begin{equation}
\Hvec_A = \frac{1}{\mu}\nabla\times\Avec
\end{equation}
where $\Avec$ is a yet-to-be-determined field known as the magnetic
vector potential. Now, using Faraday's law we obtain
\begin{equation}
\nabla\times\Evec_A = -j\omega\mu\Hvec_A = -j\omega\nabla\times\Avec.
\end{equation}
Using the terms on the left and the right and regrouping yields
\begin{equation}
\nabla\times(\Evec_A + j\omega\Avec) = 0.
\label{eq:derivingAI}
\end{equation}
The curl of the gradient of any function is identically zero. Thus we
can set the term in parentheses equal to the (negative of the)
gradient of some unknown scalar electric potential function $\Phi_e$
and in this way \refeq{eq:derivingAI} will automatically be satisfied.
Therefore we have
\begin{equation}
\Evec_A + j\omega\Avec = -\nabla\Phi_e
\end{equation}
or, after rearranging,
\begin{equation}
\Evec_A = -j\omega\Avec -\nabla\Phi_e.
\end{equation}
Using the remaining curl equation, Ampere's law, we can write
\begin{eqnarray}
\nabla\times\Hvec_A &=& \Jvec + j\omega\epsilon\Evec_A \\
\nabla\times\frac{1}{\mu}\nabla\times\Avec &=&
\Jvec + j\omega\epsilon(-j\omega\Avec -\nabla\Phi_e)
\end{eqnarray}
Multiplying through by $\mu$ and expanding the curl operations yields
\begin{equation}
\nabla(\nabla\cdot\Avec)-\nabla^2\Avec
= \mu\Jvec
+ \omega^2\mu\epsilon\Avec
- j\omega\mu\epsilon\nabla\Phi_e.
\end{equation}
Regrouping terms yields
\begin{equation}
\nabla^2\Avec + \omega^2\mu\epsilon\Avec
= -\mu\Jvec
+\nabla(\nabla\cdot\Avec + j\omega\mu\epsilon\Phi_e).
\label{eq:aVecDiffEqInitial}
\end{equation}
So far we have said what the curl of $\Avec$ must be, but that does
not fully describe the field. To fully describe a vector field one
must specify the curl, the divergence, and the value at a point (which
we will ultimately assume is zero at a infinite distance from the
origin). We are free to make the divergence of $\Avec$ any convenient
value. Let us use the ``Lorentz gauge'' of
\begin{equation}
\nabla\cdot\Avec = -j\omega\mu\epsilon\Phi_e.
\end{equation}
By doing this, \refeq{eq:aVecDiffEqInitial} reduces to
\begin{equation}
\nabla^2\Avec + k^2\Avec = -\mu\Jvec.
\label{eq:aVecDiffEq}
\end{equation}
where $k=\omega\sqrt{\mu\epsilon}$.
Distinct from the scenario described above, let us imagine a situation
where there is no free electric charge. Magnetic currents can flow,
but electric currents cannot. Thus the divergence of the electric
flux density is
\begin{equation}
\nabla\cdot \Dvec_F = \nabla\cdot \epsilon \Evec_F = 0 \label{eq:divDF}
\end{equation}
where here the subscript $F$ is used to indicate the case of no
electric charge. Again, this equation will be satisfied automatically
if we represent the electric field as the curl of some potential
function $\Fvec$. To this end we write
\begin{equation}
\Evec_F = -\frac{1}{\epsilon}\nabla\times\Fvec
\end{equation}
where $\Fvec$ is known as the electric vector potential.
Following steps similar to the ones we used to obtain
\refeq{eq:aVecDiffEq}, one can obtain the differential equation
that governs $\Fvec$, namely,
\begin{equation}
\nabla^2 \Fvec + k^2 \Fvec = -\epsilon \Mvec.\label{eq:fVecDiffEq}
\end{equation}
Thus both $\Avec$ and $\Fvec$ are governed by the wave equation. We
see that the source of $\Avec$, i.e., the forcing function that
creates $\Avec$ is the electric current $\Jvec$. Similarly, the
source of $\Fvec$ is the magnetic current $\Mvec$. (We have not yet
restricted these currents to be surface currents. At this point they
can be any current distribution, whether distributed throughout a
volume, over a surface, or along a line.)
Note that both the Laplacian ($\nabla^2$) and the constant $k^2$ that
appear in \refeq{eq:aVecDiffEq} and \refeq{eq:fVecDiffEq} are scalar
operators. They do not change the orientation of a vector. Thus, the
$x$ component of $\Jvec$ gives rise to the $x$ component of $\Avec$,
the $y$ component of $\Mvec$ gives rise to the $y$ component of
$\Fvec$, and so on. In this way, \refeq{eq:aVecDiffEq} and
\refeq{eq:fVecDiffEq} could each be broken into their three
Cartesian components and we would be left with six scalar equations.
These equations have relatively straightforward solutions. Let us
consider a slightly simplified problem, the solution of which can
easily be extended to the full general problem. Consider the case of
an incremental current of length $d\ell$ that is located at
the origin and oriented in the $z$ direction. In this case
\refeq{eq:aVecDiffEq} reduces to
\begin{equation}
\nabla^2 A_z(\rvec) + k^2 A_z(\rvec) = -\mu I d\ell \delta(\rvec)
\label{eq:azGoverningEq}
\end{equation}
where $I$ is the amount of current and $\delta(\rvec)$ is the 3D Dirac
delta function. The Dirac
delta function is zero expect when its argument is zero. For an
argument of zero, $\delta(\rvec)$ is singular, i.e., infinite.
However, this singularity is integrable. A volume integral of any
region of space that includes the Dirac delta function at the origin
(i.e., $\rvec=0$) will yield unit volume. For any observation point
other than the origin, \refeq{eq:azGoverningEq} can be written
\begin{equation}
\nabla^2 A_z(\rvec) + k^2 A_z(\rvec) = 0 \qquad \rvec\neq 0.
\end{equation}
It is rather easy to show that a general solution to this is
\begin{equation}
A_z(\rvec) = C_1\frac{e^{-jkr}}{r} + C_2\frac{e^{jkr}}{r}.
\end{equation}
We discard the second term on the right-hand side since that
represents a spherical wave propagating in toward the origin. Thus we
are left with
\begin{equation}
A_z(\rvec) = C_1\frac{e^{-jkr}}{r}
\end{equation}
where we must now determine the constant $C_1$ based on the ``driving
function'' on the right side of \refeq{eq:azGoverningEq}.
To obtain $C_1$, we integrate both side of \refeq{eq:azGoverningEq}
over a small spherical volume of radius $r_0$ and take the limit as
$r_0$ approaches zero:
\begin{equation}
\lim_{r_0\rightarrow 0} \int_V
\left[\nabla^2 A_z + k^2 A_z\right] dv
=
\lim_{r_0\rightarrow 0}
\int_V -\mu I d\ell \delta(\rvec) dv.
\label{eq:intOfAz}
\end{equation}
Using the sifting property of the delta function, the right-hand side
of \refeq{eq:intOfAz} is simply $-\mu I d\ell$. For the left-hand side,
we first expand the integral associated with the second term in the
square brackets
\begin{equation}
\lim_{r_0\rightarrow 0}
\int_{r=0}^{r_0}
\int_{\theta=0}^{\pi}
\int_{\phi=0}^{2\pi}
k^2 C_1 \frac{e^{-jk r}}{r} r^2\sin\theta\,d\phi\,d\theta\,dr.
\end{equation}
Including the term $r^2\sin\theta$, which is contributed by the
volume element $dv$, the entire integrand is proportional to $r$.
Therefore as $r_0$ (the upper limit of integration in $r$) goes to
zero, this integral goes to zero.
The remaining integral on the left side of \refeq{eq:intOfAz} is
\begin{equation}
\lim_{r_0\rightarrow 0} \int_V
\nabla\cdot\nabla A_z dv =
\lim_{r_0\rightarrow 0}
\oint_S
\nabla A_z \cdot \mathbf{ds}
\end{equation}
where we have used the divergence theorem to convert the volume
integral to a surface integral (and used the fact that $\nabla^2 =
\nabla\cdot\nabla$). The integrand of the surface integral is given
by
\begin{equation}
\left.\nabla A_z\right|_{r=r_0} =
\left.\hat{r}\frac{\partial A_z}{\partial r}\right|_{r=r_0} =
\hat{r}\left(-\frac{e^{-jk r_0}}{r_0^2}
-jk\frac{e^{-jk r_0}}{r_0}\right)C_1.
\end{equation}
The surface element $\mathbf{ds}$ is given by $\hat{r}\,r_0^2 \sin\theta
d\phi\,d\theta$ so that the entire surface integral is given by
\begin{equation}
\lim_{r_0\rightarrow 0}
\int_{\theta=0}^{\pi}
\int_{\phi=0}^{2\pi}
C_1 e^{-jk r_0}(-1-jk r_0)
\sin\theta d\phi\,d\theta =
\lim_{r_0\rightarrow 0}
-C_1 e^{-jk r_0}(1+jk r_0) 4 \pi = -C_1 4 \pi.
\end{equation}
Equating this with the right-hand side of \refeq{eq:azGoverningEq}, we
can solve for $C_1$. The final result is
\begin{equation}
C_1 = \frac{\mu I d\ell}{4 \pi}.
\end{equation}
It should be noted that there is actually nothing special about having
$r_0$ approach zero. The same coefficient is obtained for any value
of $r_0$. Letting $r_0$ approach zero merely simplifies the problem a
bit.
We now have that a filamentary current $I$ of length $d\ell$
located at the origin and oriented in the $z$ direction produces the
vector potential
\begin{equation}
A_z(\rvec) = \frac{\mu I d\ell}{4 \pi}\frac{e^{-jkr}}{r}.
\end{equation}
This is simply a spherical wave radiating symmetrically away from the
origin. If the source is located at the point $\rvec'$ instead of the
origin, one merely needs to account for this displacement. The vector
potential in that case is
\begin{equation}
A_z(\rvec) = \frac{\mu I d\ell}{4 \pi}
\frac{e^{-jk|\rvec-\rvec'|}}{|\rvec-\rvec'|}.
\end{equation}
If the current was oriented in the $x$ or $y$ direction, that would
produce a vector potential that only had a $x$ or $y$ component,
respectively.
For a current source, the ``strength'' of the source is, in one way of
thinking, determined by the amount of current that is flowing times
the length over which that current flows. For a filament we have that
the ``source strength'' is given by $I d\ell$. For a surface
current the equivalent concept is $J_s ds$ where $J_s$ is the
surface current density in Amperes/meter and $ds$ is an
incremental surface area. Similarly, for a volumetric current, the
equivalent term is $J dv$ where $J$ is the current density in
Amperes/meter$^2$ and $dv$ is an incremental volume.
Instead of having just a point source, currents can be distributed
throughout space. To get the corresponding vector potentials, we
merely have to sum the contributions of the current wherever it
exists, accounting for the location (displacement from the origin),
the orientation, and the amount of current.
For surface currents, the vector potentials are given by
\begin{eqnarray}
\Avec(\rvec) &=&\mu \oint_S \Jvec_s(\rvec')
\frac{e^{-jk|\rvec-\rvec'|}}{4\pi|\rvec-\rvec'|}
ds', \label{eq:aVecThreeD} \\
\Fvec(\rvec) &=&\epsilon \oint_S \Mvec_s(\rvec')
\frac{e^{-jk|\rvec-\rvec'|}}{4\pi|\rvec-\rvec'|}
ds', \label{eq:fVecThreeD}
\end{eqnarray}
where, as before, $\rvec$ is the observation point, $\rvec'$ is the
location of the source point (i.e., the location of the currents), and
$S$ is the surface over which the current flows.
Equations \refeq{eq:aVecThreeD} and \refeq{eq:fVecThreeD} give the
vector potentials at an arbitrary observation point $\rvec$ in three
dimensions. The surface $S$ is a surface that exists in the 3D space
(such as the surface of sphere or a cube).
Let us consider the two-dimensional case. We can consider the 2D case
as a special case in 3D in which there is no variation in the $z$
direction. The observation point is a point in the $xy$ plane
specified by the vector $\rhovec$. Thus, the magnetic vector
potential could be written as
\begin{eqnarray}
\Avec(\rhovec) &=& \mu \oint_L\int_{z'=-\infty}^\infty
\Jvec_s(\rhovec')
\frac{e^{-jk\sqrt{|\rhovec-\rhovec'|^2+z'^2}}}
{4\pi\sqrt{|\rhovec-\rhovec'|^2+z'^2}}
dz'd\ell',\\
&=& \mu \oint_L\Jvec_s(\rhovec')
\left(\,\,\int_{z'=-\infty}^\infty
\frac{e^{-jk\sqrt{|\rhovec-\rhovec'|^2+z'^2}}}
{4\pi\sqrt{|\rhovec-\rhovec'|^2+z'^2}}
dz'\right)d\ell'.
\end{eqnarray}
The term in parentheses can be integrated to obtain
\begin{equation}
\int_{z'=-\infty}^\infty
\frac{e^{-jk\sqrt{|\rhovec-\rhovec'|^2+z'^2}}}
{4\pi\sqrt{|\rhovec-\rhovec'|^2+z'^2}}
dz' = -\frac{j}{4}H_0^{(2)}(k|\rhovec-\rhovec'|)
\end{equation}
where $H_0^{(2)}$ is the zeroth-order Hankel function of the second
kind. This represents a cylindrical wave radiating from the point
$\rhovec'$. Similar steps can be done for $\Fvec$ and in this way $z$
is eliminated from the expressions for the vector potentials. We are
left with a 2D representation of the fields, namely,
\begin{eqnarray}
\Avec(\rhovec) &=& -j\frac{\mu}{4} \oint_L \Jvec(\rhovec')
H_0^{(2)}(k|\rhovec-\rhovec'|)
d\ell', \label{eq:aVecTwoD} \\
\Fvec(\rhovec) &=&-j\frac{\epsilon}{4} \oint_L \Mvec(\rhovec')
H_0^{(2)}(k|\rhovec-\rhovec'|)
d\ell'. \label{eq:fVecTwoD}
\end{eqnarray}
where the explicit $s$ subscript has been dropped from the currents.
The approximation of the zeroth-order Hankel function of the second
kind as the argument $\xi$ gets large is
\begin{equation}
H_0^{(2)}(k\xi) \approx \sqrt{\frac{j 2}{\pi k\xi}}e^{-jk\xi}.
\end{equation}
Now let $\xi=|\rhovec - \rhovec'|$ where $\rho$ is large
enough that the following approximations are valid:
\begin{equation}
\xi \approx \left\{
\begin{array}{ll}
\rho-\rho'\cos\psi & \mbox{for the phase}, \\
\rho & \mbox{for the magnitude}.
\end{array}\right.
\end{equation}
where, referring to Fig.\ \ref{fig:ntffGeomTwoD}, $\psi$ is the angle
between the observation angle and the angle to the source point.
(Because we are taking the cosine of $\psi$, and cosine is an even
function, it doesn't matter if we define $\psi$ as $\phi-\phi'$ or
$\phi'-\phi$ but we will take $\psi$ to be $\phi - \phi'$.) Thus the
Hankel function can be written
\begin{equation}
H_0^{(2)}(k|\rhovec - \rhovec'|) \approx
\sqrt{\frac{j 2}{\pi k\rho}}e^{-jk\rho}e^{jk\rho'\cos\psi}.
\end{equation}
The $\mathbf{A}$ vector potential for a 2D problem can thus be written
\begin{eqnarray}
\mathbf{A}(\rhovec) &=&
-j\frac{\mu}{4}\oint_L \mathbf{J}(\rhovec')
H_0^{(2)}(k|\rhovec - \rhovec'|)d\ell', \\
&\approx&
-j\frac{\mu}{4}\sqrt{\frac{j 2}{\pi k\rho}}e^{-jk\rho}
\oint_L \mathbf{J}(\rhovec')
e^{jk\rho'\cos\psi} d\ell', \\
&=&
-j\frac{\mu}{4}\sqrt{\frac{j 2}{\pi k\rho}}e^{-jk\rho}
\mathbf{N}_{2D},
\label{eq:afar}
\end{eqnarray}
where
\begin{equation}
\mathbf{N}_{2D} =
\oint_L \mathbf{J}(\rhovec') e^{jk\rho'\cos\psi} d\ell'.
\end{equation}
Correspondingly, the $\mathbf{F}$ vector potential can be written
\begin{eqnarray}
\mathbf{F}(\rhovec) &=&
-j\frac{\epsilon}{4}\oint_L \mathbf{M}(\rhovec')
H_0^{(2)}(k|\rhovec - \rhovec'|)d\ell', \\
&\approx&
-j\frac{\epsilon}{4}\sqrt{\frac{j 2}{\pi k\rho}}e^{-jk\rho}
\oint_L \mathbf{M}(\rhovec')
e^{jk\rho'\cos\psi} d\ell', \\
&=&
-j\frac{\epsilon}{4}\sqrt{\frac{j 2}{\pi k\rho}}e^{-jk\rho}
\mathbf{L}_{2D},
\label{eq:ffar}
\end{eqnarray}
where
\begin{equation}
\mathbf{L}_{2D} =
\oint_L \mathbf{M}(\rhovec') e^{jk\rho'\cos\psi} d\ell'.
\end{equation}
Nominally $\mathbf{N}_{2D}$ and $\mathbf{L}_{2D}$ are functions of
$\rhovec$. However, within these functions the only thing that
depends on $\rhovec$ is $\psi$. The angle $\psi$ only changes for
large changes in $\rhovec$---incremental changes of $\rhovec$ will not
affect $\psi$. Thus, derivatives of $\mathbf{N}_{2D}$ or
$\mathbf{L}_{2D}$ with respect to $\rho$, $\phi$, or $z$ (i.e.,
derivatives with respect to the unprimed coordinates) are zero.
The geometry is depicted in Fig.\ \ref{fig:ntffGeomTwoD}.
\begin{figure}
\begin{center}
\epsfig{width=3.0in,file=Figures/Fdtd-near-to-far/ntff-2d-geometry.eps}
\end{center}
\caption{Geometry associated with the near-to-far-field
transformation in 2D.} \label{fig:ntffGeomTwoD}
\end{figure}
It is convenient to think of the currents, and subsequently
$\mathbf{N}_{2D}$ and $\mathbf{L}_{2D}$ (and ultimately the
potentials), in terms of cylindrical coordinates, i.e.,
\begin{eqnarray}
\mathbf{J}(\rhovec) &=& J_\rho \unitvec{\rho} +
J_\phi \unitvec{\phi} +
J_z \zunit, \\
\mathbf{M}(\rhovec) &=& M_\rho \unitvec{\rho} +
M_\phi \unitvec{\phi} +
M_z \zunit.
\end{eqnarray}
Combining the contributions from both $\Avec$ and $\Fvec$, the
electric and magnetic fields are given by
\begin{eqnarray}
\mathbf{E}(\rhovec) &=&
-j\omega\left[\mathbf{A} +
\frac{1}{k^2}\nabla(\nabla\cdot\mathbf{A})\right]
-\frac{1}{\epsilon}\nabla\times\mathbf{F}, \label{eq:eFieldFar}
\\
\mathbf{H}(\rhovec) &=&
-j\omega\left[\mathbf{F} +
\frac{1}{k^2}\nabla(\nabla\cdot\mathbf{F})\right]
+\frac{1}{\mu}\nabla\times\mathbf{A}. \label{eq:hFieldFar}
\end{eqnarray}
By plugging \refeq{eq:afar} and \refeq{eq:ffar} into
\refeq{eq:eFieldFar} and \refeq{eq:hFieldFar} and performing the various
operations in cylindrical coordinates and discarding any terms that
fall off faster than $1/\sqrt{\rho}$, one can obtain expressions for
the electric and magnetic fields in the far field. (Note that the
$\nabla$ operator acts on the unprimed coordinates and, as mentioned
above, $\mathbf{N}_{2D}$ and $\mathbf{L}_{2D}$ are not considered
functions of the unprimed coordinates.)
\section{Electric Field in the Far-Field}
Following the steps outlined in the previous section, the scattered
electric field $\Evec^s$ at the far-field point $\rhovec$ can be
obtained from the scattered ``near'' fields using
\begin{equation}
\Evec^s(\rhovec) =
\sqrt{\frac{j}{8\pi k}} \frac{e^{-jk\rho}}{\sqrt{\rho}}
\left\{\unitvec{\phi}\left(\omega\mu_0\Ntwod\cdot\unitvec{\phi}
+k\Ltwod\cdot\zunit\right)
- \zunit\left(\omega\mu_0\Ntwod\cdot\zunit
-k\Ltwod\cdot\unitvec{\phi}\right)
\right\},
\label{eq:esfar}
\end{equation}
where, as stated previously,
\begin{eqnarray}
\Ntwod &=&
\oint_L \Jvec(\rhovec') e^{jk\rho'\cos\psi} d\ell', \\
\Ltwod &=&
\oint_L \Mvec(\rhovec') e^{jk\rho'\cos\psi} d\ell',
\end{eqnarray}
$L$ is the closed path of integration, $\psi$, given by $\phi-\phi'$,
is the angle between the source point and observation point, $\Mvec =
-\nunit\times\Evec$, $\Jvec = \nunit\times\Hvec$, and
$\nunit$ is a unit vector normal to the integration contour on
the same side of the contour as the observation point (i.e., the
outward normal). Unprimed coordinates correspond to the observation
point while primed coordinates indicate the ``source'' location (i.e.,
points along the contour).
Let us now restrict consideration to TM$^z$ polarization where the
non-zero field are $H_x$, $H_y$, and $E_z$. Since the outward normal
is restricted to exist in the $xy$ plane, $\nunit\times\Hvec$ only has
a non-zero component in the $z$ direction while $-\nunit\times\Evec$
only has non-zero components in the $xy$ plane. Thus, for this
polarization only the $z$ component of the electric field is
non-zero---the $\phi$ component of \refeq{eq:esfar} is zero. The
electric field can be written
\begin{equation}
E^s_z(\rhovec) =
-\sqrt{\frac{j}{8\pi k}} \frac{e^{-jk\rho}}{\sqrt{\rho}}
\oint_L
\left(\omega\mu_0\Jvec(\rhovec')\cdot\zunit
-k\Mvec(\rhovec')\cdot\unitvec{\phi}\right)
e^{jk\rho'\cos\psi} d\ell'.
\label{eq:ezfar}
\end{equation}
Usually the scattering width is of more interest than the field
itself. For TM$^z$ polarization the two-dimensional scattering width
is defined to be
\begin{equation}
\stwo(\phi) = \lim_{|\rhovec-\rhovec'|\rightarrow\infty} 2\pi \rho
\frac{|E^s_z(\rhovec)|^2}{|E^i_z|^2} \label{eq:swidth}
\end{equation}
Noting that $\omega\mu_0=k\eta_0$ and plugging \refeq{eq:ezfar} into
\refeq{eq:swidth} and normalizing by the wavelength yields
\begin{equation}
\frac{\stwo(\phi)}{\lambda} = \frac{1}{8\pi|E^i_z|^2}
\left|\oint_{L}\left\{\eta_0\Jvec(\rhovec')\cdot\zunit
- \Mvec(\rhovec')\cdot\unitvec{\phi}\right\}
e^{jkr'\cos(\phi-\phi')}kd\ell'\right|^2.
\label{eq:stwo}
\end{equation}
The term $r'\cos(\phi-\phi')$ which appears in the
exponent can be written as $\runit\cdot\rhovec = \runit\cdot(x'\xunit
+ y'\yunit) = x'\cos\phi + y'\sin\phi$. This last form is especially
useful since $\phi$ is fixed by the observation direction and
therefore the sine and cosine functions can be determined outside of
any loop (rather than over and over again as we move along the
integration contour).
For TM$^z$ polarization the unit vector normal to the integration
path is restricted to lie in the $xy$ plane, i.e., $\nunit = n'_x\xunit
+ n'_y\yunit$ where $({n'}_x^2+{n'}_y^2)^{1/2}=1$. Thus, the electric
current $\Jvec$ is given by
\begin{equation}
\Jvec = \nunit\times\Hvec =
\left|
\begin{array}{ccc}
\xunit & \yunit & \zunit \\
n'_x & n'_y & 0 \\
H_x & H_y & 0
\end{array}
\right|
= \zunit(n'_x H_y - n'_y H_x).
\end{equation}
The dot product of $\zunit$ and $\Jvec$ yields
\begin{equation}
\Jvec \cdot \zunit =
n'_x H_y - n'_y H_x.
\label{eq:jvecDotZ}
\end{equation}
The magnetic current is given by
\begin{equation}
\Mvec = -\nunit\times\Evec =
\left|
\begin{array}{ccc}
\xunit & \yunit & \zunit \\
n'_x & n'_y & 0 \\
0 & 0 & E_z
\end{array}
\right|
= -(\xunit n'_y E_z - \yunit n'_x E_z)
\end{equation}
The dot product of $\unitvec{\phi}$ and $\Mvec$ yields
\begin{equation}
\Mvec\cdot\unitvec{\phi} =
-\xunit\cdot\unitvec{\phi} n'_y E_z +
\yunit\cdot\unitvec{\phi} n'_x E_z =
(n'_y \sin\phi + n'_x \cos\phi) E_z
\label{eq:mvecDotPhi}
\end{equation}
Incorporating \refeq{eq:jvecDotZ} and \refeq{eq:mvecDotPhi} into
\refeq{eq:stwo} yields a general expression for the scattering width:
\begin{equation}
\frac{\stwo(\phi)}{\lambda} = \frac{1}{8\pi|E^i_z|^2}
\left|\oint_{L}\left\{\eta_0(n'_x H_y - n'_y H_x)
- (n'_y \sin\phi + n'_x \cos\phi) E_z\right\}
e^{jkr'\cos(\phi-\phi')}kd\ell'\right|^2.
\label{eq:stwoGeneral}
\end{equation}
We now want to specialize this equation to a rectangular box which is
typical of the integration boundary which would be employed in an FDTD simulation.
Assume the integration boundary corresponds to the dashed box shown
in Fig.\ \ref{fig:nfftBoundary}.
\begin{figure}
\begin{center}
\setlength{\unitlength}{2pt}
\begin{picture}(220,110)(-25,-10)
% axes
\put(-10,0){\vector(1,0){200}}
\put(182,-6){$x=m'\Delx$}
\put(0,-5){\vector(0,1){100}}
\put(-5,96){$y=n'\Dely$}
% integration boundary
\put(.5,.5){\dashbox(180,90)}
\put(2,4){$(0,0)$}
\put(155,4){$(L_x\Delx,0)$}
\put(144.5,84){$(L_x\Delx,L_y\Dely)$}
\put(2,84){$(0,L_y\Dely)$}
% side 1
\put(80,4){Side $L_4$}
\put(90,0){\vector(0,-1){9}}
\put(92,-6){$\nunit=-\yunit$}
% side 2
\put(160,43){Side $L_3$}
\put(180,45){\vector(1,0){9}}
\put(182,47){$\nunit=\xunit$}
% side 3
\put(80,84){Side $L_2$}
\put(90,90){\vector(0,1){9}}
\put(92,95){$\nunit=\yunit$}
% side 2
\put(2,43){Side $L_1$}
\put(0,45){\vector(-1,0){9}}
\put(-25,47){$\nunit=-\xunit$}
\end{picture}
\end{center}
\caption{Depiction of integration boundary for near-to-far-field
transformation in the FDTD grid. \label{fig:nfftBoundary}}
\end{figure}
The width of this rectangle is $L_x\Delx$ and the height is
$L_y\Dely$. In an FDTD grid there are $L_x+1$ samples of the fields
along the top and bottom (i.e., spanning the width) and $L_y+1$ total
samples along the left and right (i.e., spanning the height). The
integration over the closed path $L$ consists of the integration over
the four sides of this box, i.e., $L = L_1 + L_2 + L_3 + L_4$. Using
this geometry, the quantities needed to perform each integral are
presented in the following two tables.
\begin{center}
\begin{tabular}{c|l|l|r|r}
& \multicolumn{2}{c|}{$\nunit$}
& $\Jvec\cdot\zunit$
& \multicolumn{1}{c}{$\Mvec\cdot\unitvec{\phi}$}
\\ \hline
Side $L_1$ \rule[-7pt]{0pt}{20pt}
& $n'_x=-1$ & $n'_y=0$
& $-H_y$
& $-\cos\phi E_z$
\\\hline
Side $L_2$ \rule[-7pt]{0pt}{20pt}
& $n'_x=0$ & $n'_y=1$
& $-H_x$
& $\sin\phi E_z$
\\\hline
Side $L_3$ \rule[-7pt]{0pt}{20pt}
& $n'_x=1$ & $n'_y=0$
& $H_y$
& $\cos\phi E_z$
\\\hline
Side $L_4$ \rule[-7pt]{0pt}{20pt}
& $n'_x=0$ & $n'_y=-1$
& $H_x$
& $-\sin\phi E_z$
\end{tabular}
\end{center}
\begin{center}
\begin{tabular}{c|c|c|c}
& $\phi'$ & $\rho'$ & $\unitvec{\rho}\cdot\rhovec'$
\\ \hline
Side $L_1$ \rule[-7pt]{0pt}{20pt}
& $\pi/2$
& $n'\Dely$
& $n' \Dely \sin\phi$
\\\hline
Side $L_2$ \rule[-7pt]{0pt}{20pt}
& $\tan^{-1}(L_y/m')$
& $\sqrt{(m'\Delx)^2+(L_y\Dely)^2}$
& $m' \Delx \cos\phi + L_y \Dely \sin\phi$
\\\hline
Side $L_3$ \rule[-7pt]{0pt}{20pt}
& $\tan^{-1}(n'/L_x)$
& $\sqrt{(L_x\Delx)^2+(n'\Dely)^2}$
& $L_x \Delx \cos\phi + n' \Dely \sin\phi$
\\\hline
Side $L_4$ \rule[-7pt]{0pt}{20pt}
& $0$
& $m' \Delx \cos\phi$
& $m'\Dely$
\end{tabular}
\end{center}
Note that in the second table the value $n'$ represents the
vertical displacement along Side $L_1$ or $L_3$. It varies between
$0$ and $L_y$ and should not be confused with the outward normal
$\nunit$ and its components $n_x'$ and $n_y'$.
Assume that the spatial step sizes are equal so that
$\Delx=\Dely=\delta$. Further assume that the problem has be
discretized using $\ppw$ points per wavelength so that the wavenumber
can be written $k=2\pi/\lambda=2\pi/(\ppw\delta)$. Combining all
together, the scattering width is given by
\begin{eqnarray}
\frac{\stwo(\phi)}{\lambda} &=& \frac{1}{8\pi|E^i_z|^2}
\left|\,\,
\int^{L_x}_{m'=0}
% side 4
\left\{\left[\eta_0 H_x(m',0) + \sin\phi E_z(m',0)\right]
e^{j\frac{2\pi}{\ppw}m'\cos\phi}\right.\right. \nonumber\\
&&
\mbox{\hspace{.48in}} -
% side 2
\left.\left[\eta_0 H_x(m',L_y) + \sin\phi E_z(m',L_y)\right]
e^{j\frac{2\pi}{\ppw}(m'\cos\phi + L_y\sin\phi)}
\right\}\frac{2\pi}{\ppw}dm' \nonumber\\
&&\mbox{\hspace{.13in}} +
% side 1
\left.\int^{L_y}_{n'=0}
\left\{-\left[\eta_0 H_y(0,n') - \cos\phi E_z(0,n')\right]
e^{j\frac{2\pi}{\ppw}n'\sin\phi}
\right.\right. \nonumber\\
&&
\mbox{\hspace{.48in}} +
% side 3
\left.\left.\left[\eta_0 H_y(L_x,n') - \cos\phi E_z(L_x,n')\right]
e^{j\frac{2\pi}{\ppw}(L_x\cos\phi+n'\sin\phi)}\right\}
\frac{2\pi}{\ppw}dn'\right|^2\,\,\,\mbox{}
\label{eq:sall}
\end{eqnarray}
Again we note that the integration variable for the second integral is
$n'$ which corresponds to displacement along the vertical sides. The
fields in this expression are phasor (frequency-domain) quantities and
hence are complex. Although this equation may look rather messy, as
described in the next two sections, the phasors can be obtained rather
simply with a running DFT and the integration can be calculated with a
sum.
Because of the staggered nature of the FDTD grid, the electric and
magnetic fields will not be collocated on the integration
boundary---they will be offset in both space and time. A spatial
average of either the electric or magnetic field will have to be
performed to obtain the fields at the proper location. In the past, a
simple arithmetic average was typically used to account for the
spatial offsets. However, as will discussed, one can do better by
using a harmonic average in space. For the temporal offset, a simple
phase correction can be used to collocate the fields in time.
\section{Simpson's Composite Integration}
Assume we wish to integrate the function $f(x)$ over the interval
$0\leq x\leq L$ where $L$ is an even integer. The integral can be
obtained using Simpson's composite integration as follows
\begin{equation}
\int_0^L f(x) dx \approx
\frac{1}{3}\left[
f(0) + 2\sum_{m=1}^{L/2-1}f(2m) + 4\sum_{m=1}^{L/2}f(2m-1) + f(L)
\right]
\end{equation}
Note that this approximation requires a total of $L+1$ samples of the
function (so we need an odd number of samples). Using Simpson's
approximation yields quite a bit of additional accuracy over what
would be obtained using a straight Riemann sum and it costs
essentially nothing (it just requires slightly more bookkeeping).
%===================================================================
\section{Collocating the Electric and Magnetic Fields: The Geometric
Mean}
Figure \ref{fig:geometryNTFF} depicts an integration boundary in a
TM$^z$ grid.
\begin{figure}
\begin{center}
\epsfig{width=4in,file=Figures/Fdtd-near-to-far/geometryNTFF.eps}
\vspace{-.1in}
\end{center}
\caption{Depiction of a TM$^z$ grid showing the integration boundary.
The boundary is assumed to be aligned with electric-field nodes. The
expanded views show the offset of the magnetic-field nodes from the
boundary.}
\label{fig:geometryNTFF}
\end{figure}
The boundary is assumed to be aligned with the electric-field nodes,
i.e., the $E_z$ nodes. The expanded views show a portion of the
boundary along the right side and the bottom. The field notation
employs superscripts to indicate time steps while spatial indices are
given as arguments within parentheses. Half-step spatial offsets are
implicitly understood. Thus, the nodes in space-time and the
corresponding notation are
\begin{eqnarray}
H_x^{(q-1/2)\Delt}(m\Delx,[n+1/2]\Dely) &=&
H_x^{q-1/2}(m,n), \\
H_y^{(q-1/2)\Delt}([m+1/2]\Delx,n\Dely) &=&
H_y^{q-1/2}(m,n), \\
E_z^{q\Delt}(m\Delx,n\Dely) &=&
E_z^{q}(m,n),
\end{eqnarray}
where $\Delx$ and $\Dely$ are the spatial steps in the $x$ and
$y$ directions, respectively, and $\Delt$ is the temporal step.
The index $q$ indicates the temporal step and we assume it varies
between $1$ and $N_T$ which is the total number of time-steps.
Near-to-far-field (NTFF) transforms require that the fields be defined
over a single surface and use the same phase reference. For harmonic
fields, the temporal offset can be easily accounted for with a phase
factor. Assume the magnetic fields have been recorded at times of
$q=1/2, 3/2, 5/2,\ldots$ while the electric field has been recorded at
times of $q=1, 2, 3,\ldots$\@ For the harmonic transforms of interest
here, the time-domain near-fields are Fourier transformed to the
frequency domain. For example, the harmonic electric field on the
boundary is given by
\begin{eqnarray}
\hat{E}^k_z(m,n)
&=& {\cal F}\left[E_z^q(m,n)\right], \\
&=& \frac{1}{N_T}
\sum_{q=\langle N_T\rangle} E_z^q(m,n) e^{-jk\frac{2\pi}{N_T}q}.
\label{eq:dft}
\end{eqnarray}
where $\cal F$ is the discrete Fourier transform. For situations
where the entire spectrum is not of interest, typically a running
discrete Fourier transform will be used only at the particular
frequencies of interest. A frequency is specified by the index $k$
which varies between zero (dc) and $N_T-1$. Regardless of the
implementation used, the resulting spectral terms $\hat{E}^k_z$ will be
the same.
The time-domain series $E_z^q(m,n)$ can be obtained from
$\hat{E}^k_z(m,n)$ via
\begin{equation}
E_z^q(m,n) = \sum_{k=\langle N_T\rangle}
\hat{E}^k_z(m,n) e^{jk\frac{2\pi}{N_T}q}.
\label{eq:ezq}
\end{equation}
Because of the temporal offset between the electric and magnetic
fields, the desired field is actually $E_z^{q-1/2}(m,n)$, thus,
plugging $q-1/2$ into \refeq{eq:ezq} yields
\begin{equation}
E_z^{q-1/2}(m,n) =
\sum_{k=\langle N_T\rangle}
\left(\hat{E}^k_z(m,n)e^{-jk\frac{\pi}{N_T}}\right)
e^{jk\frac{2\pi}{N_T}q}.
\end{equation}
In practice one calculates $\hat{E}^k_z(m,n)$ and the spectral
representation of the magnetic fields in the same way, i.e., as in
\refeq{eq:dft}. Then one multiplies $\hat{E}^k_z(m,n)$ by
$\exp(-jk\pi/N_T)$ to account for the temporal offset.
The spatial offset is slightly more problematic than the temporal
offset. As shown in Fig.\ \ref{fig:geometryNTFF}, the integration
boundary can be aligned with only one of the fields. The magnetic
field tangential to the integration boundary is found from the nodes
that are a spatial half-step to either side of the boundary.
To obtain the magnetic field on the boundary, the traditional approach
has been to use a spatial average of the nodes to either side of the
boundary. For example, along the right side of the boundary, the
harmonic magnetic field would be given by
\begin{equation}
\hat{H}^k_y(m,n) =
\frac{1}{2}{\cal F}\left[H_y^{q-1/2}(m-1,n)+H_y^{q-1/2}(m,n)\right].
\label{eq:arithmetic}
\end{equation}
Because of this spatial average, $\hat{H}_y(m,n)$ and $\hat{E}_z(m,n)$
are assumed to be collocated and, with the temporal phase correction,
can be used to determine the equivalent currents over the integration
boundary (which are then used in the NTFF transform itself).
Unfortunately the arithmetic mean used in \refeq{eq:arithmetic}
introduces errors. To illustrate this, assume a harmonic plane wave
is propagating in the grid. The temporal frequency $\omega$ is $2\pi
k'/N_T$ where $k'$ is an integer constant and, as before, $N_T$ is the
total number of time-steps in a simulation. The $y$ component of the
magnetic field is given by
\begin{eqnarray}
\lefteqn{H_y^{q-1/2}(m,n) =
\cos\left(\omega\left[q-\half\right]\Delt-\xi\right)}, \\
&=&\!\!\!\!
\frac{e^{j\left(k'\frac{2\pi}{N_T}[q-\half]\Delt+\xi\right)}+
e^{-j\left(k'\frac{2\pi}{N_T}[q-\half]\Delt+\xi\right)}}{2},
\label{eq:hyPlaneWave}
\end{eqnarray}
where $\xi=\beta_x(m+1/2)\Delx + \beta_y n\Dely$, and $\beta_x$ and
$\beta_y$ are the $x$ and $y$ components of the wave vector,
respectively. Taking the discrete Fourier transform of
\refeq{eq:hyPlaneWave}, i.e.,
\begin{equation}
\hat{H}^k_y(m,n)
= \frac{1}{N_T}
\sum_{q=\langle N_T\rangle} H_y^{q-1/2}(m,n)
e^{-jk\frac{2\pi}{N_T}\left(q-\half\right)},
\end{equation}
one notes that the sum yields zero when $k$ is anything other than
$k'$ or $N_T-k'$. The values of $k$ that yield non-zero correspond
to the positive and negative frequency of the continuous world and,
like the continuous world, the corresponding spectral values are
complex conjugates. Without loss of generality, we will continue the
discussion in terms of the spectral component corresponding to the
positive frequency, i.e.,
\begin{equation}
\hat{H}^{k'}_y(m,n) =
\half \exp(-j[\beta_x(m+1/2)\Delx + \beta_y n\Dely]).
\end{equation}
(Note that since the time-domain functions are real-valued, in
practice one does not need to calculate the transform at any of the
negative frequencies. They are merely the complex conjugates of the
values at the positive frequencies.)
Because the Fourier transform is a linear operator, using
\refeq{eq:hyPlaneWave} in \refeq{eq:arithmetic} yields
\begin{eqnarray}
\hspace{-.15in}
\hat{H}^{k'}_y(m,n) \!\!\!\!\!&=&\!\!\!\!
e^{-j\beta_y n\Dely}
\frac{e^{-j\beta_x\!\left(m-\frac{1}{2}\right)\!\Delx} +
e^{-j\beta_x\!\left(m+\frac{1}{2}\right)\!\Delx}}{4}, \\
&=&\!\!\!\!
\half e^{-j(\beta_x m \Delx + \beta_y n\Dely)}
\cos\!\left(\frac{\beta_x\Delx}{2}\right).
\label{eq:arithmeticPlane}
\end{eqnarray}
The exact expression for the magnetic field on the integration
boundary is $\exp(-j[\beta_x m \Delx + \beta_y n\Dely])/2$. Thus, the
cosine term represents an error---one which vanishes only in the limit
as the spatial-step size goes to zero.
Instead of taking the Fourier transform of the average of the
time-domain fields, let us take the Fourier transform of the fields to
either side of the boundary. We define the transforms as
\begin{eqnarray}
\hat{H}^+_y(m,n) &=&
{\cal F}\!\left[H_y^{q-1/2}(m,n)\right], \\
\hat{H}^-_y(m,n) &=&
{\cal F}\!\left[H_y^{q-1/2}(m-1,n)\right].
\end{eqnarray}
Still assuming a single harmonic plane wave, for the ``positive
frequency'' corresponding to $k=k'$, $\hat{H}^+_y(m,n)$ and
$\hat{H}^-_y(m,n)$ are given by
\begin{eqnarray}
\hat{H}^+_y(m,n) &=&
\half e^{-j(\beta_x(m+1/2)\Delx + \beta_y n\Dely)}, \\
\hat{H}^-_y(m,n) &=&
\half e^{-j(\beta_x(m-1/2)\Delx + \beta_y n\Dely)}.
\end{eqnarray}
Were one to calculate the arithmetic mean of $\hat{H}^+_y(m,n)$ and
$\hat{H}^-_y(m,n)$, the result would be the same as given in
\refeq{eq:arithmeticPlane}. However, consider the geometric mean
(where the geometric mean of $a$ and $b$ is $\sqrt{ab}$) of