-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy pathRF-discharge-theory.tex
1496 lines (1308 loc) · 75.5 KB
/
RF-discharge-theory.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
%% LyX 2.3.2 created this file. For more info, see http://www.lyx.org/.
%% Do not edit unless you really know what you are doing.
\documentclass[12pt,english,titlepage, captions=tableheading, bibliography=totoc, usenames, dvipsnames]{scrartcl}
\usepackage{lmodern}
\renewcommand{\sfdefault}{lmss}
\renewcommand{\ttdefault}{lmtt}
\usepackage[T1]{fontenc}
\usepackage[latin9]{inputenc}
\setlength{\parskip}{\medskipamount}
\setlength{\parindent}{0pt}
\usepackage{color}
\definecolor{note_fontcolor}{rgb}{0, 0, 1}
\usepackage{babel}
\usepackage{booktabs}
\usepackage{textcomp}
\usepackage{enumitem}
\usepackage{amsmath}
\usepackage{graphicx}
\PassOptionsToPackage{version=3}{mhchem}
\usepackage{mhchem}
\usepackage[unicode=true,
bookmarks=true,bookmarksnumbered=true,bookmarksopen=true,bookmarksopenlevel=2,
breaklinks=false,pdfborder={0 0 1},backref=false,colorlinks=true]
{hyperref}
\hypersetup{
pdfauthor={Uwe Stöhr},
linkcolor=black, citecolor=black, urlcolor=blue, filecolor=blue, pdfpagelayout=OneColumn, pdfnewwindow=true, pdfstartview=XYZ, plainpages=false}
\makeatletter
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
\newcommand{\noun}[1]{\textsc{#1}}
%% Because html converters don't know tabularnewline
\providecommand{\tabularnewline}{\\}
%% The greyedout annotation environment
\newenvironment{lyxgreyedout}
{\textcolor{note_fontcolor}\bgroup\ignorespaces}
{\ignorespacesafterend\egroup}
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Textclass specific LaTeX commands.
\newlength{\lyxlabelwidth} % auxiliary length
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
% Formatierung von Legenden
\usepackage[labelfont={bf,sf}]{caption}[2004/07/16]
% ermöglicht das Berechnen von Werten
\usepackage{calc}
%Vergrößert den Teil der Seite, in dem Gleitobjekte
% unten angeordnet werden dürfen
\renewcommand{\bottomfraction}{0.5}
% Vermeidet, dass Gleitobjekte vor ihrem Abschnitt gedruckt werden
\let\mySection\section\renewcommand{\section}{\suppressfloats[t]\mySection}
% Setzt den Link fuer Spruenge zu Gleitabbildungen
% auf den Anfang des Gelitobjekts und nicht aufs Ende
\usepackage[figure]{hypcap}
% Linkfläche für Querverweise vergrößern und automatisch benennen,
\AtBeginDocument{\renewcommand{\ref}[1]{\mbox{\autoref{#1}}}}
\AtBeginDocument{%
\renewcommand*{\equationautorefname}[1]{}
\renewcommand{\sectionautorefname}{sec.\negthinspace}
\renewcommand{\subsectionautorefname}{sec.\negthinspace}
\renewcommand{\subsubsectionautorefname}{sec.\negthinspace}
\renewcommand{\figureautorefname}{Fig.\negthinspace}
}
% the pages of the TOC is numbered roman
% and a PDF-bookmark for the TOC is added
\let\myTOC\tableofcontents
\renewcommand\tableofcontents{%
\pdfbookmark[2]{\contentsname}{}
\myTOC
\newpage}
% used to have extra space in table cells
\@ifundefined{extrarowheight}
{\usepackage{array}}{}
\setlength{\extrarowheight}{2pt}
% color table cells
\usepackage{colortbl}
% enable comma as decimal separator
\usepackage{icomma}
% increase formula line separation
\setlength{\jot}{2mm+3pt}
\@ifundefined{showcaptionsetup}{}{%
\PassOptionsToPackage{caption=false}{subfig}}
\usepackage{subfig}
\makeatother
\begin{document}
\title{Theory of RF-plasma processes}
\author{Uwe Stöhr, Ph.D.\\
\\
Version 1.0}
\maketitle
\tableofcontents{}
\section{Plasma Frequency\label{sec:Plasma-Frequency}}
The plasma frequency $\omega_{e}$ is the eigenfrequency of the electrons.
Without external forces they oscillate with this frequency around
the ions. The derivation is simple: Assuming the electrons are shifted
by a length $x$ away from their position. The equation of motion
is then
\begin{equation}
m_{e}\ddot{x}=-eE\label{eq:motion}
\end{equation}
where $m_{e}=9.1\cdot10^{-31}\,$kg is the electron mass, $e=1.602\cdot10^{-19}\,$As
the elementary charge and $E$ the electric field strength.
Using the electron density $n_{e}$ in the \noun{Gauss} law we get
\begin{equation}
E=\frac{en_{e}}{\epsilon_{0}}\,x\label{eq:E-frequency}
\end{equation}
where $\epsilon_{0}=8.854\cdot10^{-12}\,$As/Vm is the vacuum permittivity.
Putting (\ref{eq:E-frequency}) into (\ref{eq:motion}) gives
\begin{equation}
0=\underbrace{\frac{e^{2}n_{e}}{m_{e}\epsilon_{0}}}_{\mathrm{eigenfrequency}}\,x+\ddot{x}
\end{equation}
Out of this oscillation equation we can directly read the expression
for the eigenfrequency:
\begin{equation}
\omega_{e}=\sqrt{\frac{e^{2}n_{e}}{m_{e}\epsilon_{0}}}
\end{equation}
Similarly to $\omega_{e}$ the eigenfrequencies for the ions can be
determined:
\begin{equation}
\omega_{i}=\sqrt{\frac{e^{2}n_{i}}{M\epsilon_{0}}}\label{eq:ion-plasma-frequency}
\end{equation}
where $M$ is the mass of an ion. Due to the global quasi-neutrality
$n_{e}\approx n_{i}$ for normal plasmas.
For a \href{https://en.wikipedia.org/wiki/Capacitively_coupled_plasma}{capacitively coupled plasma}
(CCP) RF-discharge at a pressure of a few pascals the typical electron
density is in the range of $10^{16}\,$1/m\textthreesuperior . This
leads to $\omega_{e}=5.6\,$GHz. For $\text{\ensuremath{O_{2}^{-}}}$-ions
with $M=5.312\cdot10^{-26}\,$kg we then have $\omega_{i\,\text{\ensuremath{O_{2}^{-}}}}=23.36\,$MHz.
\section{Plasma Sheath}
The sheath is the zone above the electrode and the chamber wall surface
where the electron density $n_{e}$ is much lower than the ion density
$n_{i}$. This zone is also visible as the ``dark zone'' when looking
into the chamber when a plasma is burning. The reason for the sheath
is that the RF of $f=\omega/2\pi=13.56\,$MHz is greater than the
ion plasma frequency $\omega_{i}$ but below the electron plasma frequency
$\omega_{e}$. Therefore the electrons can follow the electrical field
while the ions cannot.\footnote{\textbf{Note: }Hydrogen atoms are light enough to follow the electrical
field: $\omega_{H^{+}}\approx132\,$MHz.} So at the time where the electrode is positively charged, electrons
can stream out of the plasma to the electrode so that there is a region
where $n_{e}\approx0$ while $n_{i}\gg n_{e}$. The density distribution
at the electrode/walls is schematically shown in \ref{fig:Scheme-sheath}.
\begin{figure}[h]
\begin{centering}
\includegraphics[width=0.5\columnwidth]{clipart/Sheath-scheme}
\par\end{centering}
\caption{\label{fig:Scheme-sheath}Scheme of the particle densities in a sheath.
The wall is at $x=s$, the sheath in the range $0\le x\le s$. Image
from \cite{Keudell12}.}
\end{figure}
The thickness of the sheath is important to know because it determines
the distance from the wall in which no substrates can be placed in
for coating. It also helps to approximate the dimensions of holes
and recessed areas in which strong plasma discharges will burn (hollow
cathode effect). This topic is discussed in \ref{sec:Hollow-cathode-effect}.
\subsection{Requirements for a Sheath}
What is necessary to get a sheath? It is known that there is no plasma
burning in small holes, so obviously there is not enough space to
form a sheath. The reason could be that the ions cannot be accelerated
to reach a level to build up the positive potential at the wall. That
this is indeed the reason is shown in the following.
The energy conservation leads to
\begin{equation}
E(x)=\frac{Mv_{i}^{2}(x)}{2}+e\,\varPhi(x)=\frac{Mv_{0}^{2}}{2}=E_{0}\label{eq:energy-conservation}
\end{equation}
where $M$ is the mass of an ion, $E(x)$ is the electric field strength
in the sheath, $\varPhi(x)$ the potential in the sheath, $v_{i}(x)$
the velocity of the ions in the sheath and $v_{0}$ the velocity of
the ions in the plasma bulk streaming into the sheath.
The charge also needs to be conserved so that the current $I$ through
the area $A$ of the sheath needs to be constant:
\begin{eqnarray}
\frac{I_{i}}{A} & = & \frac{I_{0}}{A}=\frac{N_{0}\cdot e}{tA}=n_{o}\cdot e\cdot v_{0}\\
n_{i}(x)v_{i}(x) & = & n_{0}v_{0}\label{eq:current-conservation}
\end{eqnarray}
($N$ is the number of particles and $t$ the time.) Putting (\ref{eq:current-conservation})
into (\ref{eq:energy-conservation}) we get
\begin{equation}
n_{i}(x)=n_{0}\,\left(1-\frac{e\,\varPhi(x)}{E_{0}}\right)^{-1/2}
\end{equation}
The plasma frequency of electrons is above the applied RF so that
they can follow this frequency. Therefore their density follows the
\noun{Boltzmann} relation:\footnote{See \cite[sections 2.2.4 and 2.2.5]{Keudell12-2} for a detailed derivation
of the \noun{Boltzmann} relation.}
\begin{eqnarray}
en_{e}\nabla\varPhi(x) & = & \nabla p(x)=\nabla n_{e}(x)k_{B}T\\
\frac{\mathrm{d}n_{e}(x)}{n_{e}(x)} & = & \frac{e\,\mathrm{d}\,\varPhi(x)}{k_{B}T_{e}}\\
n_{e}(x) & = & n_{0}\exp\left(\frac{e\,\varPhi(x)}{k_{B}T_{e}}\right)
\end{eqnarray}
with the \noun{Boltzmann} constant $k_{B}=1.38\cdot10^{-23}\,$J/K.
The \noun{Gauss} law is then
\begin{eqnarray}
\frac{\mathrm{d}^{2}\varPhi(x)}{\mathrm{d}x^{2}} & = & \frac{e}{\epsilon_{0}}\left(n_{e}(x)+n_{i}(x)\right)\label{eq:Gauss-law1}\\
& = & \frac{en_{0}}{\epsilon_{0}}\,\left(\exp\left(\frac{e\,\varPhi(x)}{k_{B}T_{e}}\right)+\left(1-\frac{e\,\varPhi(x)}{E_{0}}\right)^{-1/2}\right)
\end{eqnarray}
($\epsilon_{0}$ is the vacuum permittivity.) It cannot be solved
analytically. For a derivation of an approximation at $\varPhi(x\approx0)$
see \cite[sec. "Raumladungszone einer Ionenrandschicht"]{Keudell12}.
This approximation delivers the relation
\begin{eqnarray}
\frac{\left(e\,\varPhi(x)\right)^{2}}{k_{B}T_{e}}-\frac{\left(e\,\varPhi(x)\right)^{2}}{2E_{0}} & > & 0\nonumber \\
\underbrace{\sqrt{\frac{k_{B}T_{e}}{M}}}_{v_{B}} & < & v_{0}\label{eq:Bohm-velocity}
\end{eqnarray}
$v_{B}$ is named \noun{Bohm} velocity and is the velocity that ions
must at least have so that a sheath can be formed. This velocity is
also the speed of sound of the ions.
\subsection{Mean Thickness of Sheaths}
We are working in a low pressure regime so that we can assume that
the mean free path $\lambda$ of the particles is independent of the
location in the plasma chamber. For the ion collision frequency we
can now write $\nu=\cfrac{v_{i}}{\lambda}$. With the ion velocity
$v_{i}(x)=\mu E(x)$, the ion drift mobility $\mu=\cfrac{e}{M\nu}$
and the field strength $E$ we can write
\begin{eqnarray}
v_{i}(x) & = & \frac{e\lambda}{Mv_{i}(x)}\,E(x)\nonumber \\
& = & \sqrt{\frac{e\lambda}{M}\,E(x)}
\end{eqnarray}
where $M$ is the mass of an ion and $e$ the electron charge. The
ion density can now be calculated according to (\ref{eq:current-conservation}):
\begin{equation}
n_{i}(x)=\frac{v_{B}}{v_{i}(x)}\,n_{0}\label{eq:n_i}
\end{equation}
where $v_{B}$ is the \noun{Bohm} velocity that is necessary to get
a sheath, $T_{e}$ the electron temperature, $k_{B}$ the \noun{Boltzmann}
constant and $n_{0}\approx n_{e}\approx n_{i}$ the density in the
plasma bulk.
The simplest model of a sheath is that the electron density in the
sheath is zero. Using this we can neglect the electron density $n_{e}$
in the \noun{Gauss} law:
\begin{eqnarray}
\epsilon_{0}\,\frac{\mathrm{d}E}{\mathrm{d}x} & = & e\,\left(n_{i}(x)+\underbrace{n_{e}(x)}_{\approx0}\right)=\cfrac{e\,n_{0}\,\sqrt{\cfrac{k_{B}T_{e}}{M}}}{\sqrt{\frac{e\lambda}{M}\,E}}\label{eq:Gauss-law2}\\
\epsilon_{0}\sqrt{E}\,\mathrm{d}E & = & en_{0}\,\sqrt{\cfrac{k_{B}T_{e}}{e\lambda}}\,\mathrm{d}x=j_{0}\,\sqrt{\cfrac{M}{e\lambda}}\,\mathrm{d}x\label{eq:current-density}
\end{eqnarray}
($\epsilon_{0}$ is the vacuum permittivity and $j_{0}$ the current
density.) We integrate from the border of the sheath at $x=0$ to
the chamber border at $x=s$:
\begin{equation}
E^{3}=\frac{9}{4}\,s^{2}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\label{eq:E-sheath}
\end{equation}
At the electrode we apply the voltage $U$ so that we have $E(s)=\cfrac{U}{s}$.
We can now calculate the sheath thickness $s$ to
\begin{equation}
s=\left(\frac{4U^{3}\epsilon_{0}^{2}\lambda}{9n_{0}^{2}ek_{B}T_{e}}\right)^{1/5}\label{eq:sheath-thickness}
\end{equation}
Using the relation for the mean free path assuming a \noun{Maxwell}
distribution for the molecule energies
\begin{equation}
\lambda=\cfrac{k_{B}T}{\sqrt{2}\pi d^{2}p}\label{eq:lambda}
\end{equation}
($d$ is the diameter of the gas molecules.) (\ref{eq:sheath-thickness})
can be transformed to
\begin{equation}
s=\left(\frac{4U^{3}\epsilon_{0}^{2}T}{9\sqrt{2}n_{0}^{2}\pi epd^{2}T_{e}}\right)^{1/5}\label{eq:s-mean}
\end{equation}
The values for $\lambda$ for different precursors are listed in \ref{sec:Mean-Free-Paths}.
The mean voltage $\bar{U}$ at the electrode is the bias voltage so
that this can be used to calculate the mean sheath thickness $s$.
Taking typical values: $\lambda_{\ce{O2}}(p=2\,\mathrm{Pa})\approx8\,$mm,
$T\approx300\,$K, $T_{e}\approx3\cdot10^{4}\,$K, $n_{0}\approx10^{16}\,$1/m\textthreesuperior ,
$p=2\,$Pa, $\bar{U}\approx400\,$V leads with (\ref{eq:sheath-thickness})
to $s\approx4.7\,$mm.
(\ref{eq:s-mean}) is the time-independent mean sheath thickness where
also the dependency on the electrode area is not taken into account.
An approximation of a time- and area-dependent sheath thickness is
derived in the following section.
\subsection{Time-dependent Thickness of Sheaths}
A CCP is kept burning by shifting the charge within the chamber in
the applied frequency $\omega_{\mathrm{RF}}$. This displacement current
$I_{a}$ is
\begin{equation}
I_{a}(t)=I_{\mathrm{RF}}\cos(\omega_{\mathrm{RF}}t)\label{eq:I-displacement}
\end{equation}
$I_{\mathrm{RF}}$ is the current in the circuit that delivers the
RF.
For the electrical field across the sheath we found the relation (\ref{eq:E-sheath}).
If we this time don't integrate (\ref{eq:current-density}) from $x=0..s$
but from $x=s..x$ we get
\begin{equation}
E(x,\,t)^{3}=\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\,\left(x^{2}-s(t)^{2}\right)
\end{equation}
The displacement current of the sheath above a wall is
\begin{eqnarray}
I_{a}(t) & = & \epsilon_{0}A_{w}\frac{\partial E(x,\,t)}{\partial t}\label{eq:I_a}\\
& = & -\epsilon_{0}A_{w}\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{1/3}\frac{\partial s(t)^{2/3}}{\partial t}\label{eq:I_a-final}
\end{eqnarray}
where $A_{w}$ is the area of the wall.
With (\ref{eq:I-displacement}) we can now calculate the time-dependent
$s_{w}(t)$:
\begin{eqnarray}
s_{w}(t)^{2/3} & = & -\underbrace{\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{-1/3}\frac{I_{\mathrm{RF}}}{\epsilon_{0}A_{w}\omega_{\mathrm{RF}}}}_{{\displaystyle s_{0}}}\,\sin(\omega_{\mathrm{RF}}t)+C\label{eq:s(t)}
\end{eqnarray}
At one time the sheath must collapse to enable electrons to leave
the plasma compensating the ion current. So at one time we have $s(t)=0$
leading to $C=s_{0}$. The mean sheath thickness $\bar{s}$ is then
\begin{eqnarray}
\bar{s}_{w} & = & \left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{-1/2}\left(\frac{I_{\mathrm{RF}}}{\epsilon_{0}A_{w}\omega_{\mathrm{RF}}}\right)^{3/2}\underbrace{\left\langle \left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{3/2}\right\rangle _{t}}_{{\displaystyle \approx1.2}}\nonumber \\
& = & \frac{2.4}{3n_{0}}\,\sqrt{\frac{\lambda}{\epsilon_{0}e\,k_{B}T_{e}}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{3/2}\label{eq:s_precise}
\end{eqnarray}
The dime-dependent voltage across the sheath above a wall is
\begin{eqnarray}
U_{w}(t) & = & \int_{0}^{s(t)}E(x,\,t)\,\mathrm{d}x\\
& = & \int_{0}^{s(t)}\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{1/3}\left(x^{2}-s_{w}(t)^{2}\right)^{1/3}\,\mathrm{d}x\nonumber \\
& = & \frac{3}{5}\,s_{w}(t)^{5/3}\,\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{1/3}+C_{U}\label{eq:U_w-intermediate}\\
& = & \frac{3}{5}\,\left(-s_{0}\,\sin(\omega_{\mathrm{RF}}t)+s_{0}\right)^{5/2}\,\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{1/3}+C_{U}\nonumber \\
& = & \frac{3}{5}\,\left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}\left(\frac{I_{\mathrm{RF}}}{\epsilon_{0}A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\,\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{-1/2}+C_{U}
\end{eqnarray}
At one time per cycle the sheath must collapse so that $U(t)=0$.
This is already the case because of the sine term so that $C_{U}=0$
and we finally get
\begin{eqnarray}
U_{w}(t) & = & \frac{2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\label{eq:U_w-(t)}\\
\bar{U}_{w} & \approx & \frac{1.92\cdot2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\label{eq:U_w-precise}
\end{eqnarray}
where it was used that $\left\langle \left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}\right\rangle _{t}\approx1.92$.
But what is about the sheath at the electrode?\\
We know that the charge in the plasma must be conserved to fulfill
the quasi-neutrality. So the current towards the wall $I_{a}$ must
be the opposite of the one towards the electrode $I_{e}$. With (\ref{eq:I_a-final})
we find
\begin{eqnarray}
I_{a} & = & -I_{e}\nonumber \\
-A_{w}\frac{\partial s_{a}(t)^{2/3}}{\partial t} & = & A_{e}\frac{\partial s_{e}(t)^{2/3}}{\partial t}\nonumber \\
s_{e}(t)^{2/3} & = & -\frac{A_{w}}{A_{e}}s_{a}(t)^{2/3}\nonumber \\
s_{e}(t)^{2/3} & = & \frac{A_{w}}{A_{e}}\left(s_{0}\,\sin(\omega_{\mathrm{RF}}t)-C\right)
\end{eqnarray}
We hereby paid attention that due to the differentiation we cannot
know the sign of the integration constant in (\ref{eq:s(t)}) that
we identified as $C=s_{0}$. The sheath thickness cannot be negative
and at one time the sheath must collapse. So we get
\begin{eqnarray}
s_{e}(t)^{2/3} & = & \frac{A_{w}}{A_{e}}\,s_{0}\left(\sin(\omega_{\mathrm{RF}}t)+1\right)\nonumber \\
s_{e}(t) & = & \left(\frac{A_{w}}{A_{e}}\right)^{3/2}\,\frac{2}{3n_{0}}\,\sqrt{\frac{\lambda}{\epsilon_{0}e\,k_{B}T_{e}}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{3/2}\left(\sin(\omega_{\mathrm{RF}}t)+1\right)^{3/2}\\
\bar{s}_{e} & = & \left(\frac{A_{w}}{A_{e}}\right)^{3/2}\,\frac{2.4}{3n_{0}}\,\sqrt{\frac{\lambda}{\epsilon_{0}e\,k_{B}T_{e}}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{3/2}=\left(\frac{A_{w}}{A_{e}}\right)^{3/2}\bar{s}_{w}\label{eq:s_e-mean}
\end{eqnarray}
where it was used that $\left\langle \left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{3/2}\right\rangle _{t}\approx1.2$.
The voltage across the sheath at the electrode $U_{e}$ can now be
calculated using (\ref{eq:U_w-intermediate}):
\begin{eqnarray}
U_{e}(t) & = & \frac{3}{5}\,s_{e}(t)^{5/3}\,\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{1/3}+C_{U}\nonumber \\
& = & \frac{3}{5}\,\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\left(s_{0}\,\sin(\omega_{\mathrm{RF}}t)+s_{0}\right)^{5/2}\,\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{1/3}+C_{U}\nonumber \\
& = & \frac{3}{5}\,\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}\left(\frac{I_{\mathrm{RF}}}{\epsilon_{0}A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\left(\frac{9}{4}\,n_{0}^{2}\,\frac{e\,k_{B}T_{e}}{\epsilon_{0}^{2}\lambda}\right)^{-1/2}+C_{U}
\end{eqnarray}
Also in this case at one time the voltage must vanish so that $C_{U}=0$
and we get
\begin{eqnarray}
U_{e}(t) & = & \left(\frac{A_{w}}{A_{e}}\right)^{5/2}\frac{2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\nonumber \\
\bar{U}_{e} & = & \left(\frac{A_{w}}{A_{e}}\right)^{5/2}\bar{U}_{w}\label{eq:5/2-law}
\end{eqnarray}
\subsection*{Conclusion}
The wall area, the applied frequency and the input current (and thus
the input power) have a strong influence on the sheath voltage, while
the influence of the mean free length of path is low.
The voltage across the sheath at the electrode strongly depends on
the ratio of the electrode area to the wall area. For example for
the 80\,l chamber device ``Domino''\footnote{Manufacturer: Plasma Electronic; device details: \href{https://www.plasma-electronics.com/domino.html}{https://www.plasma-electronics.com/domino.html}}
$A_{w}/A_{e}\approx6$ so that $U_{e}$ is about 88~times greater
than $U_{w}$.
The expression for $\bar{U}_{w}$ contains with $n_{0}$ and $T_{e}$
2~variables which we usually don't measure. The electron temperature
only appears because we respected that the ions in the sheath can
collide with each other so that they are decelerated. As an approximation
we could now neglect collisions. We then have in the sheath $n_{e}=0$
and $n_{i}=n_{0}=\mathrm{constant}$. In this so-called ``matrix
sheath'' model the electron temperature and the mean free path do
not influence the sheath voltage and thickness.
\subsection*{Approximation}
Using the matrix model we can write
\begin{eqnarray}
\frac{\mathrm{d}\,E(x,\,t)}{\mathrm{d}x} & = & \frac{en_{0}}{\epsilon_{0}}\nonumber \\
E(x,\,t) & = & \frac{en_{0}}{\epsilon_{0}}\,(x-s(t))
\end{eqnarray}
using again (\ref{eq:I_a}) we get
\begin{equation}
s_{\mathrm{matrix}}(t)=-\underbrace{\frac{I_{\mathrm{RF}}}{e\,n_{0}A_{w}\omega_{\mathrm{RF}}}}_{{\displaystyle s_{0\,\mathrm{matrix}}}}\,\sin(\omega_{\mathrm{RF}}t)+\frac{I_{\mathrm{RF}}}{e\,n_{0}A_{w}\,\omega_{\mathrm{RF}}}\label{eq:s-approx}
\end{equation}
and
\begin{eqnarray}
U_{w\,\mathrm{matrix}}(t) & = & \frac{en_{0}}{2\epsilon_{0}}\,s_{\mathrm{matrix}}^{2}(t)\label{eq:U-sheath-approx}\\
& = & \frac{en_{0}}{2\epsilon_{0}}\left(\frac{I_{\mathrm{RF}}}{e\,n_{0}A_{w}\,\omega_{\mathrm{RF}}}\,\left(1-\sin(\omega_{\mathrm{RF}}t)\right)\right)^{2}\label{eq:U_sheath-approx-detail}\\
\bar{U}_{w\,\mathrm{matrix}} & = & \frac{1}{2e\epsilon_{0}n_{0}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\,\omega_{\mathrm{RF}}}\right)^{2}\label{eq:U_w-matrix-mean}
\end{eqnarray}
We further on get
\begin{equation}
\bar{U}_{e\,\mathrm{matrix}}=\left(\frac{A_{w}}{A_{e}}\right)^{2}\bar{U}_{w\,\mathrm{matrix}}\label{eq:U-matrix-relation}
\end{equation}
\section{Bias Voltage}
If the electrode has the same area $A_{e}$ than the area $A_{w}$
of the grounded chamber walls, the voltage across both sheaths is
the same. But if the electrode area is smaller, there the is less
space to transport the same charge as to the chamber walls. But across
the sheaths the same charge must be transported otherwise the charge
conservation would be violated. This leads us to (\ref{eq:5/2-law}):
\begin{equation}
\bar{U}_{e}=\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\bar{U}_{w}
\end{equation}
What we measure as bias voltage is the mean voltage at the electrode
\begin{equation}
U_{\mathrm{bias}}=\bar{U}_{e}=\bar{U}_{w}\,\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\label{eq:U_bias-general}
\end{equation}
Using (\ref{eq:U_w-precise}) we have
\begin{equation}
U_{\mathrm{bias}}=\frac{1.92\cdot2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\,\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\label{eq:U_bias}
\end{equation}
We know that the applied voltage (the one in the circuit) is the difference
of the sheath voltages. With (\ref{eq:U_w-(t)}) we can therefore
write
\begin{eqnarray}
U_{\mathrm{RF}}(t) & = & U_{e}(t)-U_{w}(t)\nonumber \\
& = & \frac{2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\nonumber \\
& & \cdot\,\left(\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}-\left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}\right)\label{eq:U_RF(t)}
\end{eqnarray}
$U_{\mathrm{RF}}(t)$, $U_{w}(t)$ and $U_{e}(t)$ are shown in \ref{fig:Temporal-progression}.
It can be seen that $U_{e}(t)-U_{w}(t)$ is a distorted sine that
has an offset of $\left(\cfrac{A_{w}}{A_{e}}\right)^{5/2}-1$ and
the maximum at $\pi/2$. In our calculation we applied a cosine current.
That the voltages follow a sine visualizes that the plasma sheath
model treats the plasma as a pure capacity. A possible inductance
due to inhomogeneities in the plasma bulk (for example if metal substrates
are used) are not taken into account.
\begin{figure}[h]
\begin{centering}
\includegraphics[width=0.75\columnwidth]{clipart/Bias-RF-voltage}
\par\end{centering}
\caption{\label{fig:Temporal-progression}Temporal progression of $U_{\mathrm{RF}}=U_{e}-U_{w}$,
$U_{e}$, $U_{e}$ and $P$ for $A_{w}/A_{e}=2$.}
\end{figure}
The power we have to input is
\begin{eqnarray}
P(t) & = & I_{\mathrm{RF}}\cos(\omega_{\mathrm{RF}}t)\cdot U_{\mathrm{RF}}(t)\label{eq:P-general}
\end{eqnarray}
In (\ref{eq:U_RF(t)}) we see that the term $\left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}$
can be neglected for $\left(\cfrac{A_{w}}{A_{e}}\right)^{5/2}>50$
(error is then < 2\,\%). Using this in (\ref{eq:P-general}) gives
\begin{equation}
P(t)=I_{\mathrm{RF}}\cos(\omega_{\mathrm{RF}}t)\,\frac{2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}
\end{equation}
The temporal progression of $P(t)$ is shown in \ref{fig:Temporal-progression}.
The generators measure the root mean square (RMS) of $P(t)$ as effective
power. The RMS of the term $\cos(\omega_{\mathrm{RF}}t)\cdot\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{5/2}$
is $\approx1.436$. We can therefore write
\begin{eqnarray}
P_{\mathrm{measured}} & = & I_{\mathrm{RF}}\,\frac{1.436\cdot2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\nonumber \\
& = & I_{\mathrm{RF}}^{7/2}\,\frac{1.436\cdot2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{1}{A_{e}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\\
I_{\mathrm{RF}} & = & \left(\frac{1.436\cdot2}{5P_{\mathrm{measured}}n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{1}{A_{e}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\right)^{-2/7}\nonumber \\
& = & \left(\frac{5P_{\mathrm{measured}}n_{0}\epsilon_{0}^{3/2}}{1.436\cdot2}\right)^{2/7}\left(\frac{e\,k_{B}T_{e}}{\lambda}\right)^{1/7}\left(A_{e}\omega_{\mathrm{RF}}\right)^{5/7}
\end{eqnarray}
This could now be put into (\ref{eq:U_bias}):
\begin{eqnarray}
U_{\mathrm{bias}} & = & \frac{1.92\cdot2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{\left(\frac{5P_{\mathrm{measured}}n_{0}\epsilon_{0}^{3/2}}{1.436\cdot2}\right)^{2/7}\left(\frac{e\,k_{B}T_{e}}{\lambda}\right)^{1/7}\left(A_{e}\omega_{\mathrm{RF}}\right)^{5/7}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\,\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\nonumber \\
& = & \frac{0.831}{n_{0}\epsilon_{0}^{3/2}\omega_{\mathrm{RF}}^{5/2}}\,\left(\left(P_{\mathrm{measured}}n_{0}\epsilon_{0}^{3/2}\right)^{2/7}\left(\frac{e\,k_{B}T_{e}}{\lambda}\right)^{1/7}\left(A_{e}\omega_{\mathrm{RF}}\right)^{5/7}\right)^{5/2}\sqrt{\frac{\lambda}{e\,k_{B}T_{e}}}\,\left(\frac{1}{A_{e}}\right)^{5/2}\nonumber \\
& = & \frac{0.831}{n_{0}^{2/7}\epsilon_{0}^{3/7}}\,\left(\frac{P_{\mathrm{measured}}}{\omega_{\mathrm{RF}}A_{e}}\right)^{5/7}\left(\frac{\lambda}{e\,k_{B}T_{e}}\right)^{1/7}\label{eq:U-Bias-power}
\end{eqnarray}
and (\ref{eq:s_e-mean}) but this doesn't help us as long as we don't
know the electron temperature. We therefore go back to the matrix
model approximation.
\subsection{Approximation}
We already found (\ref{eq:U_sheath-approx-detail}):
\begin{equation}
U_{w\,\mathrm{matrix}}(t)=\frac{en_{0}}{2\epsilon_{0}}\left(\frac{I_{\mathrm{RF}}}{e\,n_{0}A_{w}\omega_{\mathrm{RF}}}\,\left(1-\sin(\omega_{\mathrm{RF}}t)\right)\right)^{2}
\end{equation}
and with (\ref{eq:U-matrix-relation}) we get
\begin{equation}
U_{\mathrm{RF}}(t)=\frac{en_{0}}{2\epsilon_{0}}\left(\frac{I_{\mathrm{RF}}}{e\,n_{0}A_{w}\omega_{\mathrm{RF}}}\right)^{2}\left(\left(\frac{A_{w}}{A_{e}}\right)^{2}\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{2}-\left(1-\sin(\omega_{\mathrm{RF}}t)\right)^{2}\right)
\end{equation}
To calculate the power we assume that $\left(\cfrac{A_{w}}{A_{e}}\right)^{2}\gg1$
so that we have to find only the RMS of the term $\cos(\omega_{\mathrm{RF}}t)\cdot\left(1+\sin(\omega_{\mathrm{RF}}t)\right)^{2}\raisebox{5mm}{}$.
This is $\approx1.146$ so that we can now write
\begin{eqnarray}
P_{\mathrm{measured}} & = & I_{\mathrm{RF}}\,\frac{1.146\cdot en_{0}}{2\epsilon_{0}}\left(\frac{I_{\mathrm{RF}}}{e\,n_{0}A_{w}\omega_{\mathrm{RF}}}\right)^{2}\left(\frac{A_{w}}{A_{e}}\right)^{2}\nonumber \\
& = & I_{\mathrm{RF}}^{3}\,\frac{1.146}{2\epsilon_{0}en_{0}}\left(\frac{1}{A_{e}\omega_{\mathrm{RF}}}\right)^{2}\\
I_{\mathrm{RF}} & = & \left(A_{e}\,\omega_{\mathrm{RF}}\right)^{2/3}\,\left(\frac{2\epsilon_{0}en_{0}P_{\mathrm{measured}}}{1.146}\right)^{1/3}\label{eq:I_RF-approx}
\end{eqnarray}
Putting this relation for $I_{\mathrm{RF}}$ into (\ref{eq:U_w-matrix-mean})
gives
\begin{eqnarray}
U_{\mathrm{bias}} & = & \frac{1}{2e\epsilon_{0}n_{0}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\,\omega_{\mathrm{RF}}}\right)^{2}\left(\frac{A_{w}}{A_{e}}\right)^{2}\\
& = & \frac{1}{2e\epsilon_{0}n_{0}}\,\left(\frac{\left(A_{e}\,\omega_{\mathrm{RF}}\right)^{2/3}\,\left(\frac{2\epsilon_{0}en_{0}P_{\mathrm{measured}}}{1.146}\right)^{1/3}}{A_{e}\,\omega_{\mathrm{RF}}}\right)^{2}\left(\frac{A_{w}}{A_{e}}\right)^{2}\nonumber \\
& = & \left(\frac{1}{2e\epsilon_{0}n_{0}}\right)^{1/3}\,\left(\frac{P_{\mathrm{measured}}}{1.146\cdot A_{e}\,\omega_{\mathrm{RF}}}\right)^{2/3}\left(\frac{A_{w}}{A_{e}}\right)^{2}
\end{eqnarray}
and this can be transformed to deliver the plasma density:
\begin{eqnarray}
n_{0}^{1/3} & = & \frac{1}{U_{\mathrm{bias}}}\,\left(\frac{1}{2e\epsilon_{0}}\right)^{1/3}\,\left(\frac{P_{\mathrm{measured}}}{1.146\cdot A_{e}\,\omega_{\mathrm{RF}}}\right)^{2/3}\left(\frac{A_{w}}{A_{e}}\right)^{2}\nonumber \\
n_{0} & = & \frac{1}{2e\epsilon_{0}U_{\mathrm{bias}}^{3}}\,\left(\frac{P_{\mathrm{measured}}}{1.146\cdot A_{e}\,\omega_{\mathrm{RF}}}\right)^{2}\left(\frac{A_{w}}{A_{e}}\right)^{6}\label{eq:n_0-approx}
\end{eqnarray}
With this approximated plasma density we can also approximate the
ionization ratio $\Gamma_{i}$:
\begin{equation}
\Gamma_{i}=\frac{N}{Vn_{0}}=\frac{p}{k_{B}T\,n_{0}}\label{eq:ionization-ratio}
\end{equation}
where $N$ is the number of particles in the plasma chamber, $p$
is the pressure, $V$ the volume of the chamber and $T$ the global
temperature in the chamber.
Although (\ref{eq:n_0-approx}) is an approximation, we can put it
into the precise relation (\ref{eq:U_w-precise}) to get a feeling
how the electron temperature behaves:
\begin{eqnarray}
U_{\mathrm{bias}} & = & \frac{1.92\cdot2}{5n_{0}\epsilon_{0}^{3/2}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\,\left(\frac{ek_{B}T_{e}}{\lambda}\right)^{-1/2}\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\nonumber \\
\frac{ek_{B}T_{e}}{\lambda} & = & \left(\frac{1.92\cdot2}{5n_{0}\epsilon_{0}^{3/2}U_{\mathrm{bias}}}\,\left(\frac{I_{\mathrm{RF}}}{A_{w}\omega_{\mathrm{RF}}}\right)^{5/2}\,\left(\frac{A_{w}}{A_{e}}\right)^{5/2}\right)^{2}\nonumber \\
T_{e} & = & \left(\frac{1.92\cdot2}{5}\right)^{2}\left(\frac{I_{\mathrm{RF}}}{A_{e}\,\omega_{\mathrm{RF}}}\right)^{5}\frac{\lambda}{n_{0}^{2}\,\epsilon_{0}^{3}\,U_{\mathrm{bias}}^{2}\,ek_{B}}\label{eq:T_e-wrong}\\
& = & \left(\frac{1.92\cdot2}{5}\right)^{2}\left(\frac{\left(\frac{5P_{\mathrm{measured}}n_{0}\epsilon_{0}^{3/2}}{1.436\cdot2}\right)^{2/7}\left(\frac{e\,k_{B}T_{e}}{\lambda}\right)^{1/7}\left(A_{e}\omega_{\mathrm{RF}}\right)^{5/7}}{A_{e}\,\omega_{\mathrm{RF}}}\right)^{5}\frac{\lambda}{n_{0}^{2}\,\epsilon_{0}^{3}\,U_{\mathrm{bias}}^{2}\,ek_{B}}\nonumber \\
& = & \left(\left(\frac{1.92\cdot2}{5}\right)^{2}\left(\frac{5P_{\mathrm{measured}}n_{0}\epsilon_{0}^{3/2}}{1.436\cdot2}\right)^{10/7}\left(\frac{e\,k_{B}}{\lambda}\right)^{5/7}\frac{1}{\left(A_{e}\omega_{\mathrm{RF}}\right)^{10/7}}\,\frac{\lambda}{n_{0}^{2}\,\epsilon_{0}^{3}\,U_{\mathrm{bias}}^{2}\,ek_{B}}\right)^{7/2}\nonumber \\
& = & \left(\frac{1.92\cdot2}{5}\right)^{7}\left(\frac{5P_{\mathrm{measured}}}{1.436\cdot2}\right)^{5}\frac{\lambda}{n_{0}^{2}ek_{B}\,\left(A_{e}\omega_{\mathrm{RF}}\right)^{5}\epsilon_{0}^{3}\,U_{\mathrm{bias}}^{7}}
\end{eqnarray}
\subsection{Conclusions}
\begin{figure}[h]
\begin{centering}
\includegraphics[width=0.75\columnwidth]{clipart/Bias-Geometry}
\par\end{centering}
\caption{\label{fig:Spatial-progression}Spatial progression of the voltage
in a plasma between symmetric and asymmetric electrode areas. Image
from \cite{Keudell12}.}
\end{figure}
\ref{fig:Spatial-progression} shows the spatial progression of the
voltage inside a plasma schematically. Due to the smaller electrode
area the voltage across the sheath above the electrode is much higher.
This is the effect we use to accelerate ions towards substrates to
coat them. (\ref{eq:U_bias}) shows that a smaller $A_{e}$ leads
to less necessary input current (and thus input power) to get the
same bias voltage.
Interesting is the inverse cubic dependence of the bias voltage in
(\ref{eq:n_0-approx}). So a low plasma density results in a high
bias. This is an important result because this is the reason why a
dense plasma is burning into holes of substrates at the electrode
for low ionized plasmas. These are typically plasmas of mainly larger
precursor molecules. This topic is discussed in detail in \ref{sec:Hollow-cathode-effect}.
\begin{figure}[h]
\begin{centering}
\includegraphics[width=1\columnwidth]{clipart/plasma-density-Domino}
\par\end{centering}
\caption{\label{fig:Plasma-density-Domino}Plasma density for the plasma device
``Domino'' with a chamber volume of 80\,l ($A_{e}\approx0.138\,$m\texttwosuperior ,
$A_{w}\approx0.832$\,m\texttwosuperior ) in dependence of the applied
plasma power and the measured $U_{\mathrm{bias}}$.}
\end{figure}
\begin{figure}[!h]
\begin{centering}
\includegraphics[width=1\columnwidth]{clipart/plasma-density-Porta900}
\par\end{centering}
\caption{\label{fig:Plasma-density-Porta}Plasma density for the plasma device
``Porta~900'' with one large electrode ($A_{e}=0.64\,$m\texttwosuperior ,
$A_{w}=0.9^{2}\cdot6-0.64=4.22$\,m\texttwosuperior ) in dependence
of the applied plasma power and the measured $U_{\mathrm{bias}}$.
For the setup with 3~chambers ($A_{e}=0.07\,$m\texttwosuperior ,
$A_{w}=2.594$\,m\texttwosuperior ) the values in this plot must
be multiplied with the factor 198.16.}
\end{figure}
The plasma density $n_{0}$ for the 80\,l chamber device ``Domino''\footnote{Manufacturer: Plasma Electronic; device details: \href{https://www.plasma-electronics.com/domino.html}{https://www.plasma-electronics.com/domino.html}}
is shown in \ref{fig:Plasma-density-Domino}, the plasma density for
a 720\,l chamber device ``Porta~900''\footnote{Manufacturer: Plasma Electronic; cubic chamber with an edge length
of 0.9\,m;\\
device details: \href{https://www.plasma-electronics.com/porta.html}{https://www.plasma-electronics.com/porta.html}} in \ref{fig:Plasma-density-Porta}. One can see that if one measures
for a low power a relatively high bias, the plasma density is low
meaning the ionization rate is also low. By modifying the ``Porta~900''
so that the device chamber is divided by 3~equally sized inner chambers
with each a smaller electrode, the plasma density is increased by
about a factor~200. So the then different ratio $A_{e}/A_{w}$ change
the plasma parameters drastically.
Taking the case that in the ``Domino'' we measured a bias of 500\,V
for pure Ar at $p=2\,$Pa with a power of 190\,W. With \ref{fig:Plasma-density-Domino}
we get $n_{0}\approx1.5\cdot10^{14}\,$1/m\textthreesuperior{} and
therefore with (\ref{eq:T_e-wrong}) $T_{e}\approx7.6\cdot10^{5}\,$K.
But this about 100~times higher than values found in literature and
with (\ref{eq:Phi_float}) we would get a plasma potential of about
327\,V. This shows that the electron temperature cannot be estimated
using the matrix model in its derivation. This result could have been
expected because the matrix model does not take care of $T_{e}$.
In fact one cannot estimate $T_{e}$ by measuring only the plasma
power and the bias voltage.
\subsection{Applications}
\subsubsection{ICP + CCP}
Looking at (\ref{eq:U-Bias-power}):
\begin{equation}
U_{\mathrm{bias}}=\frac{0.831}{n_{0}^{2/7}\epsilon_{0}^{3/7}}\,\left(\frac{P_{\mathrm{measured}}}{\omega_{\mathrm{RF}}A_{e}}\right)^{5/7}\left(\frac{\lambda}{e\,k_{B}T_{e}}\right)^{1/7}
\end{equation}
we see that $U_{\mathrm{bias}}\sim n_{0}^{-2/7}$, $U_{\mathrm{bias}}\sim P_{\mathrm{measured}}^{5/7}$
and $U_{\mathrm{bias}}\sim T_{e}^{-1/7}$. For a setup where the ions
are generated in an \href{https://en.wikipedia.org/wiki/Inductively_coupled_plasma}{inductively coupled plasma}
(ICP), the ion energy is a constant while the ion density increases
with increasing input power for the ICP. To get a defined and adjustable
ion impact energy on the substrate one can apply an RF-voltage to
the substrate. We then have a \href{https://en.wikipedia.org/wiki/Capacitively_coupled_plasma}{capacitively coupled plasma}
(CCP) above the substrate but $n_{0}$ is almost independent of the
CCP parameters.
For a glass etching application this setup was chosen. For the ICP
and the CCP the same frequency was used. As both plasmas did not interfere
in this setup although the same frequency was used, they can be treated
as being independent from each other. Then for a constant ICP power
and thus a constant $n_{0}$ the bias voltage should have the dependency
$U_{\mathrm{bias}}\sim P_{\mathrm{measured}}^{5/7}$ for the case
that the CCP does not change the electron temperature $T_{e}$. The
measurements to prove this theory was performed using an ICP source
model ``Copra DN 160''\footnote{Manufacturer: CCR technology;\\
device details: \href{https://www.ccrtechnology.de/copra-products/copra-round-sources/}{https://www.ccrtechnology.de/copra-products/copra-round-sources/}}. The result is plotted in \ref{fig:Dependency-Bias-ICP}. Fitting
the curves in this plot with the function $U_{\mathrm{bias}}=K\cdot P_{\mathrm{measured}}^{X}$
gives the results listed in \ref{tab:Fit-parameters}. The expected
exponent $5/7\approx0.71$ is the one for 300\,W. For larger ICP
powers the exponent increases which implies that $T_{e}$ is then
smaller.
\begin{figure}[h]
\begin{centering}
\includegraphics[width=0.7\columnwidth]{clipart/ICP-CCP-Bias}
\par\end{centering}
\caption{\label{fig:Dependency-Bias-ICP}Dependency of the bias voltage on
the CCP power for different ICP powers.}
\end{figure}
\begin{table}[h]
\caption{\label{tab:Fit-parameters}Fit parameters for the measurement curves
from \ref{fig:Dependency-Bias-ICP}.}
\centering{}%
\begin{tabular}{|c|c|c|c|}
\hline
ICP power in W & Parameter $X$ & Parameter $K$ & $\chi_{\mathrm{red}}^{2}$ of the fit\tabularnewline
\hline
\hline
300 & $0.71\pm0.01$ & $73.71\pm9.09$ & 1.12\tabularnewline
\hline
400 & $0.79\pm0.01$ & $22.82\pm2.06$ & 1.58\tabularnewline
\hline
500 & $0.82\pm0.01$ & $14.82\pm1.16$ & 1.31\tabularnewline
\hline
\end{tabular}
\end{table}
The ion current density in the used ICP source for a fixed pressure
of 0.1\,Pa is for an input power of 300\,W about 0.1\,mA/cm\texttwosuperior{}
and for 500\,W about 0.21\,mA/cm\texttwosuperior . Although we measured
at a pressure of about 0.4\,Pa we assume that also $n_{0}$ is increased
by a factor 1.75 if going from 300 to 500\,W ICP power. For low CCP
powers we can neglect the influence of the CCP on $n_{0}$ and expect
that the bias voltage for 500\,W is $1.75^{2/7}\approx0.44$ lower
than for 300\,W. Looking at \ref{fig:Dependency-Bias-ICP} we get
a factor 0.53 -- 0.67 lower bias voltage, depending on the CCP power,
see \ref{tab:Factors-of-the}. Even for the lowest measurable CCP
power of 10\,W we see again that an increasing ICP power decreases
$T_{e}$ as the factor is much greater than expected. For higher CCP
powers we have also a lower $T_{e}$. This is an interesting result
because a higher CCP power always leads to a higher $n_{0}$ and this
decreases the bias.
\begin{table}[h]
\caption{\label{tab:Factors-of-the}Factors of the decreasing bias voltages
when going from 300 to 500\,W ICP power; in dependence of the CCP
power.}
\centering{}%
\begin{tabular}{|c|c|c|c|c|c|c|c|}
\hline
CCP power in W & 10 & 15 & 20 & 30 & 40 & 50 & 60\tabularnewline
\hline
\hline
Factor & 0.53 & 0.55 & 0.56 & 0.62 & 0.64 & 0.65 & 0.67\tabularnewline
\hline
\end{tabular}
\end{table}
\subsubsection{CCP}
In contrary to the combination of ICP~+~CCP, for a pure CCP the
Bias voltage behave differently. According to (\ref{eq:U-Bias-power}):
\begin{equation}
U_{\mathrm{bias}}=\frac{0.831}{n_{0}^{2/7}\epsilon_{0}^{3/7}}\,\left(\frac{P_{\mathrm{measured}}}{\omega_{\mathrm{RF}}A_{e}}\right)^{5/7}\left(\frac{\lambda}{e\,k_{B}T_{e}}\right)^{1/7}
\end{equation}
one gets the relations $U_{\mathrm{bias}}\sim n_{0}^{-2/7}$ and
$U_{\mathrm{bias}}\sim P_{\mathrm{measured}}^{5/7}$. In a CCP an
increased $P$ will also always result in an increased $n_{0}$ and
a changed $T_{e}$. For practical applications it is helpful to know
how the Bias is changed by changing the input power and the pressure
$p$.
For the measurement the device ``Labor~1'' (80~liter vacuum chamber)
was used. Pure argon was used as process gas because it is a single-atom
gas (to keep the pressure constant because no molecule fragmentation
can occur). The Bias voltage was measured at constant pressures while
the input power was changed. The resulting datasets were fit using
these equations
\begin{equation}
U_{\mathrm{bias}}=K\cdot\left(P_{\mathrm{measured}}-x_{0}\right)^{B_{\mathrm{powerr}}}+y_{0}
\end{equation}
\begin{equation}
U_{\mathrm{bias}}=K\cdot\left(p-x_{0}\right)^{B_{\mathrm{pressure}}}+y_{0}
\end{equation}
\ref{fig:Dependency-Bias-CCP} shows as example how the results looked.
It can be seen that the fit at constant pressure well describes the
change of the Bias whereas the fit at constant pressure shows that
the fit model is not optimal ($\chi_{\mathrm{red}}>1$). In \ref{fig:Dependency-Bias-CCP-final}
the fit parameters $B_{\mathrm{power}}$ and $B_{\mathrm{pressure}}$
are plotted for the different pressures and powers, receptively. One
can see that the fit parameters are relatively constant for the different
pressures and powers. As final conclusion one gets for a CCP with
argon:
\begin{equation}
U_{\mathrm{bias}}\sim P_{\mathrm{measured}}^{{\textstyle 0.5360\pm0.0152}}\label{eq:U-P-exponent}
\end{equation}
\begin{equation}
U_{\mathrm{bias}}\sim p^{{\textstyle -0.1831\pm0.0016}}\label{eq:U-p-exponent}
\end{equation}
This shows that the plasma density $n_{0}$ is indeed increased if
the pressure is increased. The electron temperature $T_{e}$ is increased
too so that the exponent in (\ref{eq:U-p-exponent}) is greater than
$-2/7$. If the CCP-power is increased $n_{0}$ and $T_{e}$ are increased
too so that the exponent in (\ref{eq:U-P-exponent}) is smaller than
$5/7$.
\begin{figure}[p]
\begin{centering}
\subfloat[Dependency of the bias voltage on the input power for a fixed pressure
of 4.14\,Pa.]{\begin{centering}
\includegraphics[width=0.8\columnwidth]{clipart/Power-Bias-at-4-14Pa}
\par\end{centering}
}
\par\end{centering}
\begin{centering}
\subfloat[Dependency of the bias voltage on the pressure for a fixed input power
of 400\,W.]{\begin{centering}
\includegraphics[width=0.8\columnwidth]{clipart/Pressure-Bias-at-400W}
\par\end{centering}
}
\par\end{centering}
\caption{\label{fig:Dependency-Bias-CCP}Dependency of the bias voltage on
the pressure and the input power for 2~different cases.}
\end{figure}
\begin{figure}[p]
\begin{centering}
\subfloat[Fit of the fit-parameters $B_{\mathrm{power}}$ at different pressures.]{\begin{centering}
\includegraphics[width=0.85\columnwidth]{clipart/Bias-Exponent-Pressure}
\par\end{centering}
}
\par\end{centering}
\begin{centering}
\subfloat[Fit of the fit-parameters $B_{\mathrm{pressure}}$ at different input
powers.]{\begin{centering}
\includegraphics[width=0.85\columnwidth]{clipart/Bias-Exponent-Power}
\par\end{centering}
}
\par\end{centering}
\caption{\label{fig:Dependency-Bias-CCP-final}Linear fit of the different
fit-parameters $B_{\mathrm{power}}$ and $B_{\mathrm{pressure}}$.}
\end{figure}
\newpage{}
\section{Floating Potential}
When treating polymer substrates they are often placed into the chamber
without an electrical connection to the chamber walls or the electrode.
At the substrate there is therefore no sheath. The question is now
how large the potential between the plasma and the substrate will
be.
As there is no electric connection, the substrate is on a so-called
floating potential and the ion current $I_{i}$ must be equal the
electron current $I_{e}$ onto the substrate surface. So we can write
\begin{eqnarray}
I_{e} & = & I_{i}\nonumber \\
A\,n_{0}v_{e,\,\mathrm{therm}} & = & A\,n_{0}v_{B}\label{eq:charge-conservation}
\end{eqnarray}
where we used that the ions within the plasma must have the \noun{Bohm}
velocity.
Within the plasma we can assume that the electrical field is independent
of the position. Therefore the velocity of the electrons is their
thermal one $v_{e,\,\mathrm{therm}}$.
The mean $v_{e,\,\mathrm{therm}}$ can be calculated using the Maxwell-Boltzmann
distribution $f(v)$:
\begin{eqnarray}
\left\langle v_{e,\,\mathrm{therm}}\right\rangle & = & \int vf(v)\,\mathrm{d}v\nonumber \\
& = & \int v_{e}\,\sqrt{\frac{m_{e}}{2\pi k_{B}T_{e}}}\exp\left(\frac{-m_{e}v_{_{e}}^{2}}{2k_{B}T_{e}}\right)\,\mathrm{d}v_{e}\nonumber \\
& = & -\sqrt{\frac{k_{B}T_{e}}{2\pi m_{e}}}\exp\left(\frac{-e\,\varPhi}{k_{B}T_{e}}\right)\label{eq:<e-therm>}
\end{eqnarray}
Where the relation $\cfrac{m_{e}v_{e}^{2}}{2}=e\,\varPhi$ was used.
(\ref{eq:charge-conservation}) becomes now
\begin{eqnarray}
\sqrt{\frac{k_{B}T_{e}}{M}} & = & -\sqrt{\frac{k_{B}T_{e}}{2\pi m_{e}}}\exp\left(\frac{-e\,\varPhi_{\mathrm{float}}}{k_{B}T_{e}}\right)\nonumber \\
\varPhi_{\mathrm{float}} & = & \frac{k_{B}T_{e}}{e}\ln\left(\sqrt{\frac{2\pi m_{e}}{M}}\right)
\end{eqnarray}
In our devices the electrode is set to a certain voltage to apply
power to the plasma. Therefore we get a bias which is negative. Electrons
are therefore reflected and cannot reach the electrode surface (except
of one point in time). We therefore have instead of (\ref{eq:charge-conservation})
\begin{eqnarray}
A_{w}\,n_{0}v_{e,\,\mathrm{therm}} & = & n_{0}v_{B}\,(A_{w}+A_{e})
\end{eqnarray}
which leads to
\begin{eqnarray}
\sqrt{\frac{k_{B}T_{e}}{M}}\,(A_{w}+A_{e}) & = & -\sqrt{\frac{k_{B}T_{e}}{2\pi m_{e}}}\exp\left(\frac{-e\,\varPhi_{\mathrm{float}}}{k_{B}T_{e}}\right)A_{w}\nonumber \\
\varPhi_{\mathrm{float}} & = & \frac{k_{B}T_{e}}{e}\ln\left(\frac{A_{w}+A_{e}}{A_{w}}\,\sqrt{\frac{2\pi m_{e}}{M}}\right)\label{eq:Phi_float}
\end{eqnarray}
The influence of the ion mass $M=N\,$u ($\mathrm{u}=1.66\cdot10^{-27}\,$kg
is the atomic mass unit) on $\varPhi_{\mathrm{float}}$ is shown in
\ref{fig:floating-potential}. $\ce{O2}$ has a mass of 32\,u while
a large molecule like \href{https://en.wikipedia.org/wiki/Hexamethyldisiloxane}{HMDSO}
has a mass of 162~u. It can be seen that the potential does not increase
a lot with higher ion masses and for large molecules one has to have
in mind that they are cracked into smaller ones in the plasma so that
the real floating potential will not be far above the one of oxygen.
\begin{figure}[h]
\begin{centering}
\includegraphics[width=1\columnwidth]{clipart/Floating-potential}
\par\end{centering}
\caption{\label{fig:floating-potential}$\Phi_{\mathrm{float}}$ in V in dependence
of the ion mass $M=N\,$u for different fractions $\cfrac{A_{w}}{A_{e}}$.
The electron temperature was set in the calculation to $T_{e}=3.48\cdot10^{4}\,$K.}
\end{figure}
As the ions must have the Bohm velocity, the potential to accelerate
the ions to the velocity is the plasma potential:
\begin{eqnarray}
e\,\varPhi_{\mathrm{plasma}} & = & \frac{Mv_{B}^{^{2}}}{2}\nonumber \\
\varPhi_{\mathrm{plasma}} & = & \frac{k_{B}T_{e}}{2e}
\end{eqnarray}
For a typical electron temperature of $T_{e}=3.48\cdot10^{4}\,$K
we have $\varPhi_{\mathrm{plasma}}\approx1.5\,$V.
\subsection*{Conclusion}
The floating potential is often greater than the voltage across the
sheath of a grounded chamber wall. Assuming we have a setup with $A_{w}=6A_{e}$
and measure $U_{\mathrm{bias}}=400\,$V we can calculate with (\ref{eq:U_bias-general})
that $U_{w}\approx4.5\,$V. If the electrode area is much larger than
the chamber wall area the plasma potential can be positive.
By measuring the floating potential with e.\,g.\ a \noun{Langmuir}
probe, we could measure the electron temperature and could then calculate
plasma potential, the real sheath thickness $s$ according to (\ref{eq:s_precise})
and also the real plasma density $n_{0}$ according to (\ref{eq:U_w-precise}).
\section{Dissipated Power}
The power in a CCP is dissipated because energy is consumed to ionize
molecules, the ions collide and therefore loose energy and the electrons
are accelerated. So in summary we can write
\begin{equation}
P=\int E\,\mathrm{d}t=\int\left(E_{\mathrm{ionization}}+E_{e,\,\mathrm{therm}}+E_{\mathrm{sheath}}\right)\,\mathrm{d}t
\end{equation}
The power can in principle be split into an ohmic and a stochastic
part.
\subsection{Ohmic}
The ohmic part describes the heating of the plasma due to the RF-electrical
field accelerating charged particles. The equation of motion is
\begin{equation}
m\ddot{x}+m\nu\dot{x}=eE_{0}\sin(\omega_{\mathrm{RF}}t)\label{eq:Force_ohm}
\end{equation}
where $m$ is the mass of a particle, $\nu$ the collision rate of
the particles and $E_{0}$ the applied field strength.
The power is the energy per time $t$ and the energy is the force
along a path $x$:
\begin{equation}
P_{\mathrm{ohm,\,particle}}(t)=\frac{F(t)\,\mathrm{d}x}{\mathrm{d}t}=eE_{0}\sin(\omega_{\mathrm{RF}}t)\dot{x}
\end{equation}
$\dot{x}$ can be derived by solving (\ref{eq:Force_ohm}). It is
\begin{equation}
\dot{x}=-\,\frac{eE_{0}\omega_{\mathrm{RF}}}{m\left(\omega_{\mathrm{RF}}^{2}+\nu^{2}\right)}\,\left(\cos(\omega_{\mathrm{RF}}t)-\frac{\nu}{\omega_{\mathrm{RF}}}\sin(\omega_{\mathrm{RF}}t)\right)
\end{equation}
so that we get for a single particle
\begin{eqnarray}
\bar{P}_{\mathrm{ohm,\,particle}} & = & \frac{\omega_{\mathrm{RF}}}{2\pi}\int_{0}^{2\pi/\omega_{\mathrm{RF}}}P_{\mathrm{ohm,\,particle}}\,\mathrm{d}t\nonumber \\
& = & \frac{e^{2}E_{0}^{2}}{2m}\,\frac{\nu}{\omega_{\mathrm{RF}}^{2}+\nu^{2}}
\end{eqnarray}
The power per volume is simply
\begin{eqnarray}
\bar{P}_{\mathrm{ohm}} & = & \bar{P}_{\mathrm{ohm,\,particle}}\cdot n_{0}\nonumber \\
& = & \frac{n_{0}e^{2}E_{0}^{2}}{2m}\,\frac{\nu}{\omega_{\mathrm{RF}}^{2}+\nu^{2}}\label{eq:P_ohm}
\end{eqnarray}
This result can be interpreted by splitting it into a part of the
electric field and one of the energy transfer rate due to the collisions:
\begin{eqnarray}
\bar{P}_{\mathrm{ohm}} & = & \frac{\epsilon_{0}E_{0}^{2}}{2}\cdot\,\underbrace{\frac{n_{0}e^{2}}{m\epsilon_{0}}}_{\omega_{\mathrm{plasma}}^{2}}\frac{\nu}{\omega_{\mathrm{RF}}^{2}+\nu^{2}}\nonumber \\
& = & \underbrace{\frac{\epsilon_{0}E_{0}^{2}}{2}}_{\text{electric field}}\cdot\underbrace{\frac{\omega_{\mathrm{plasma}}^{2}\,\nu}{\omega_{\mathrm{RF}}^{2}+\nu^{2}}}_{\text{collisions}}\label{eq:P_ohm-rate}
\end{eqnarray}
((\ref{eq:ion-plasma-frequency}) was used.)
The collision term shows that the ohmic heating is maximal if the
collision rate $\nu=\cfrac{v}{\lambda}$ is equal to the applied frequency
$\omega_{\mathrm{RF}}$.
It is interesting to see how electrons and ions are heated. We take
as example an argon plasma ($M=40\,$u) and $\omega_{\mathrm{RF}}=85.2\,$MHz.
With (\ref{eq:ion-plasma-frequency}) we get
\begin{equation}
\frac{\omega_{i}}{\omega_{e}}=\sqrt{\frac{m_{e}}{M}}=3.7\cdot10^{-3}
\end{equation}
Within the plasma bulk the ions must have the \noun{Bohm} velocity
so that we get with (\ref{eq:Bohm-velocity}) and (\ref{eq:lambda})
for the ion collision rate
\begin{eqnarray}
\nu_{i} & = & \frac{v_{B}}{\lambda_{i}}=\sqrt{\frac{k_{B}T_{e}}{\lambda_{i}^{2}M}}=\sqrt{\frac{T_{e}2\pi^{2}D^{4}p^{2}}{k_{B}T^{2}M}}\label{eq:nu-expanded}
\end{eqnarray}
Putting (\ref{eq:Phi_float}) into (\ref{eq:<e-therm>}) leads to
\begin{equation}
\nu_{e}=\frac{\left\langle v_{e,\,\mathrm{therm}}\right\rangle }{\lambda_{e}}=\frac{A_{w}}{A_{w}+A_{e}}\,\sqrt{k_{B}T_{e}M}\,\frac{1}{2\pi m_{e}\lambda_{e}}
\end{equation}
With (\ref{eq:nu-expanded}) we get
\begin{eqnarray}
\frac{\nu_{i}}{\nu_{e}} & = & \frac{\lambda_{e}}{\lambda_{i}}\,\frac{2\pi m_{e}}{M}\,\frac{A_{w}+A_{e}}{A_{w}}\nonumber \\
& = & \frac{R_{i}^{2}}{R_{e}^{2}}\,\frac{2\pi m_{e}}{M}\,\frac{A_{w}+A_{e}}{A_{w}}\nonumber \\
& = & \frac{8\pi m_{e}}{M}\,\frac{A_{w}+A_{e}}{A_{w}}\approx3.4\cdot10^{-4}\,\frac{A_{w}+A_{e}}{A_{w}}
\end{eqnarray}
where $R$ is the radius of the of the cross-sectional area. For ions
this is twice the molecule radius so that $R_{i}=d$ ($d$ is the