-
Notifications
You must be signed in to change notification settings - Fork 3
/
Copy path6002x-notes-exscape.tex
3087 lines (2057 loc) · 162 KB
/
6002x-notes-exscape.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
\author{Thomas Backman, exscape@gmail.com}
\date{\today}
\title{6.002x notes}
% Use a nice font
\documentclass[12pt,a4paper]{report}
% Custom package!
% For strikethrough.
\usepackage{cancel}
% Create clickable URLs
%\usepackage{url}
% Multiline comments
\usepackage{verbatim}
\usepackage{graphicx}
%\usepackage{pgfplots}
% Aligned equations and more
\usepackage{amsmath}
\usepackage{mathtools}
% Links in the table of contents
\usepackage[colorlinks=true, urlcolor=blue,linkcolor=red]{hyperref}
% Fix margins and indentation
\usepackage[cm]{fullpage}
%\addtolength{\oddsidemargin}{-0.75cm}
%\addtolength{\evensidemargin}{-0.75cm}
%\addtolength{\topmargin}{-0.5cm}
\usepackage[
top = 1cm,
bottom = 2cm,
left = 1cm,
right = 1cm]{geometry}
\setlength{\parindent}{0in}
% For circuit diagrams
\usepackage{siunitx}
\usepackage[american,siunitx]{circuitikz}
% \bar{} is a bit too small (and grey?), \overline{} is a bit too wide!
% Here's a compromise:
\newcommand{\overbar}[1]{\mkern 1.5mu\overline{\mkern-1.5mu#1\mkern-1.5mu}\mkern 1.5mu}
\begin{document}
\maketitle
\tableofcontents
%%%%%%%%%%%%%%%%%%%%%%%%
%% Formulas and such %%%
%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Formulas and such}
\section{Intro}
This chapter contains several common formulas etc., without any further explanation of how they are derived.
\section{Components in series and parallel}
Series resistances add:
\[ R_s = R_1 + R_2 + R_3 + \cdots \]
Parallel resistances diminish:
\[ R_p = \frac{1}{ \frac{1}{R_1} + \frac{1}{R_2} + \frac{1}{R_3} + \cdots} \]
Impedances behave like resistances.\\
Series capacitances diminish:
\[ C_s = \frac{1}{ \frac{1}{C_1} + \frac{1}{C_2} + \frac{1}{C_3} + \cdots} \]
Parallel capacitances add:
\[ C_p = C_1 + C_2 + C_3 + \cdots \]
Series inductances add:
\[ L_s = L_1 + L_2 + L_3 + \cdots \]
Parallel inductances diminish:
\[ L_p = \frac{1}{ \frac{1}{L_1} + \frac{1}{L_2} + \frac{1}{L_3} + \cdots} \]
For the special case of two components of the diminshing type (x being a dummy variable):
\[ \frac{1}{ \frac{1}{x_1} + \frac{1}{x_2} } = \frac{x_1 \cdot x_2}{x_1 + x_2} \]
\section{Digital}
\subsection{Thresholds}
\begin{itemize}
\item $V_{OL}$: the \emph{highest} allowed output voltage for a logical 0 (lower is OK)
\item $V_{IL}$: the \emph{highest} allowed input voltage for a logical 0 (higher would be in the forbidden region)
\item $V_{IH}$: the \emph{lowest} allowed input voltage for a logical 1 (lower would be in the forbidden region)
\item $V_{OH}$: the \emph{lowest} allowed output voltage for a logical 1 (higher is OK)
\end{itemize}
\[ 0 \le V_{OL} < V_{IL} < V_{IH} < V_{OH} \le V_S \]
\subsection{Noise margins}
\begin{align*}
NM_H &= V_{OH} - V_{IH}\\
NM_L &= V_{IL} - V_{OL}\\
\text{Forbidden region} &= V_{IH} - V_{IL}
\end{align*}
\section{MOSFETs}
\subsection{SCS model}
\[
i_{DS} = \begin{cases}
\frac{K}{2}(v_{GS} - V_T)^2 & \text{for $v_{GS} \ge V_T, v_{DS} \ge v_{GS} - V_T$} \\
0 & \text{for $v_{GS} < V_T$}
\end{cases}
\]
The above covers saturation and cutoff regions only (the SCS model).
\subsection{SR model}
\[
i_{DS} = \begin{cases}
\frac{V_{DS}}{R_{ON}} & \text{for $V_{GS} \ge V_T$} \\
0 & \text{otherwise}
\end{cases}
\]
Alternatively:
\[
R_{DS} = \begin{cases}
R_{ON} & \text{for $V_{GS} \ge V_T$} \\
\infty & \text{otherwise}
\end{cases}
\]
Used for digital circuits only (in this course).
\section{State devices / energy storage devices}
This section assumes time-invariant devices, i.e. capacitance/inductance is a fixed value, and not a function of time.
\subsection{Capacitors}
The current through a capacitor is a function of the \emph{rate of change} of voltage:
\[ i(t) = C \frac{dv(t)}{dt} \]
To find the voltage over a capacitor, we need to know its full history:
\[ v(t) = \frac{1}{C} \int_{-\infty}^{t} i(t) dt \]
... or we can simply do it by knowing the current through it between $t_1$ and $t_2$, plus the initial voltage:
\[ v(t_2) = \frac{1}{C} \int_{t_1}^{t_2} i(t) dt + v(t_1) \]
The energy stored in a capacitor is:
\[ E = \frac{1}{2} C v^2 \]
\subsection{Inductors}
The voltage over an inductor is a function of the \emph{rate of change} of current:
\[ v(t) = L \frac{di(t)}{dt} \]
To find the current through an inductor, we need to know its full history:
\[ i(t) = \frac{1}{L} \int_{-\infty}^t v(t) dt \]
... or we can simply do it by knowing the voltage over it between $t_1$ and $t_2$, plus the initial current through it:
\[ i(t_2) = \frac{1}{L} \int_{t_1}^{t_2} v(t) dt + i(t_1) \]
The energy stored in an inductor is:
\[ E = \frac{1}{2} L i^2 \]
\section{First order circuits}
General equation for an increasing/decaying exponential in a RC/RL circuit:
\[ v = V_S + (V_0 - V_S) e^{-t/RC} \]
\[ i = \frac{V_S}{R} + (i_0 - \frac{V_S}{R}) e^{-Rt/L} \]
where v = the voltage across the capacitor. For RL circuits, replace $V_S$ by the maximum current and $V_0$ by the initial current through the inductor. Above is for a thevenin equivalent circuit.\\
ZIR / Zero Input Response, same as above with $V_S = 0$:
\[ V_0 e^{-t/RC} \]
ZSR / Zero State Response, same as above with $V_0 = 0$:
\[ V_S(1 - e^{-t/RC}) \]
\newpage
\section{Second order circuits, impedance, filters}
Canonical form of the characteristic equation for second-order circuits; use this to match up the values of $\alpha$ and $\omega_0$ for a circuit:
\[ s^2 + 2\alpha s + {\omega_0}^2 = 0 \]
\subsection{For all LC and RLC circuits:}
\[ \text{Natural/undamped resonant radian frequency: } \omega_0 \text{ (rad/s)} \]
\[ \text{Damping factor: } \alpha \text{ (rad/s)} \]
Note that zeta ($\zeta$) is used as a damping factor in many texts; it is defined as
\[ \zeta = \frac{\alpha}{\omega_0} \text{ (dimensionless)} \]
The bandwidth $\Delta \omega$, i.e. the \emph{width} of the frequency \emph{band} that is above $\displaystyle \frac{1}{\sqrt{2}}$ times the input amplitude, is given by
\[ \Delta \omega = 2 \alpha \text{ (rad/s) (measured at } \frac{1}{\sqrt{2}} \text{ points)} \]
\[ \text{Quality factor: } Q = \frac{\omega_0}{2\alpha} \text{ (dimensionless)} \]
RLC circuits can be underdamped, overdamped, or critically damped.
\[ \text{Underdamped: } \omega_0 > \alpha \text{ or, equivalently, } Q > \frac{1}{2} \text { or, equivalently, } \zeta < 1 \]
\[ \text{Overdamped: } \omega_0 < \alpha \text{ or, equivalently, } Q < \frac{1}{2} \text { or, equivalently, } \zeta > 1 \]
\[ \text{Critically damped: } \omega_0 = \alpha \text{ or, equivalently, } Q = \frac{1}{2} \text { or, equivalently, } \zeta = 1 \]
When they are \emph{underdamped}, the \emph{damped resonant frequency} $\omega_d$ is given by
\[ \omega_d = \sqrt{{\omega_0}^2 - \alpha^2} \]
The naming here might be confusing; the \emph{damped} frequency is used for \emph{underdamped} systems. The reason is that the \emph{undamped} frequency is used for systems with no damping whatsoever, i.e. LC circuits with no resistor.\\
Underdamped RLC circuits are the only kind that oscillate, so the natural frequency is less interesting for overdamped circuits.
\subsection{Series RLC circuits}
\[ \omega_0 = \frac{1}{\sqrt{LC}} \text{ rad/s} \]
\[ f_0 = \frac{1}{2\pi \sqrt{LC}} \text { Hz} \]
\[ \alpha = \frac{R}{2L} \text{ rad/s} \]
\[ \Delta \omega = 2\alpha = \frac{R}{L} \text { rad/s} \]
\[ \text{Period: } \frac{2\pi}{\omega_0} = 2\pi \sqrt{LC} \text { seconds} \]
\[ Q = \frac{\omega_0}{2\alpha} = \frac{L}{R\sqrt{LC}} \text { (dimensionless)} \]
\subsection{Parallel RLC circuits}
\[ \omega_0 = \frac{1}{\sqrt{LC}} \text{ rad/s} \]
\[ f_0 = \frac{1}{2\pi \sqrt{LC}} \text { Hz} \]
\[ \alpha = \frac{1}{2RC} \text{ rad/s} \]
\[ \Delta \omega = 2\alpha = \frac{1}{RC} \text { rad/s} \]
\[ \text{Period: } \frac{2\pi}{\omega_0} = 2\pi \sqrt{LC} \text { seconds} \]
\[ Q = \frac{\omega_0}{2\alpha} = \frac{RC}{\sqrt{LC}} \text { (dimensionless)} \]
\subsection{Frequency- to time-domain conversion}
You can find the time-domain behavior of a circuit to sinusoidal input from nothing but a complex amplitude of the form $V_x$:
\[ v_X(t) = |V_x| \cos{(\omega t + \angle V_x)} \]
See below for information about how to calculate the magnitude $|z|$ and the angle $\angle z$ of a complex number.
\subsection{Complex algebra}
A few properties of complex numbers that are necessary to know:
\[ |a + jb| = \sqrt{a^2 + b^2} \]
\[ \angle (a + jb) = \arctan{(\frac{b}{a})} \text{ or, preferably, } \text{atan2}(a, b) \]
\[ |a + j0| = a \text { if $a > 0$; otherwise, the magnitude is just the absolute value } |a| \]
\[ |0 + jb| = b \]
\[ |0 - jb| = b \]
\[ \angle (a + j0) = 0 \]
\[ \angle (0 + jb) = \frac{\pi}{2} \]
\[ \angle (0 - jb) = -\frac{\pi}{2} \]
\[ |z_1 \cdot z_2| = |z_1| \cdot |z_2| \]
\[ \left| \frac{z_1}{z_2} \right| = \frac{|z_1|}{|z_2|} \]
\[ \angle (z_1 \cdot z_2) = \angle z_1 + \angle z_2 \]
\[ \angle \left( \frac{z_1}{z_2} \right) = \angle z_1 - \angle z_2 \]
When finding angles, it's best to use a function which is capable of giving correct answers in all quadrants of the unit circle. The ``atan2'' function was created in many computer languages for this purpose. It is equal to $\displaystyle \arctan{(\frac{b}{a})}$ for some inputs, but not all; arctan() cannot differentiate between all quadrants, because $\displaystyle \arctan{\left( \frac{-a}{-b}\right )} = \arctan{\left( \frac{a}{b}\right) }$ and likewise, $\displaystyle \arctan{\left( \frac{-a}{b}\right )} = \arctan{\left( \frac{a}{-b}\right) }$. Thus, the atan2 function has two arguments, and when using it, the angle of a complex number $a + jb$ is simply $\text{atan2}(a, b)$.\\
The angle of a fully real number is always $0$, and the angle of a fully imaginary number always either $\displaystyle \frac{\pi}{2}$ (for positive imaginary numbers) or $\displaystyle -\frac{\pi}{2}$ (for negative imaginary numbers).
One definition of atan2, if you are not using math software that has it, is:
\[ \text{atan2}(b, a) = 2\arctan{(\frac{b}{\sqrt{a^2 + b^2} + a})} \]
However, do note that these notes use an atan2 function that is defined with the variables in the opposite order, i.e. atan2(a, b); specifically, Mathematica's ArcTan[a, b] function.
\subsection{Impedances}
\[ \text{Resistor: } Z_R = R \]
\[ \text{Capacitor: } Z_C = \frac{1}{j\omega C} \]
\[ \text{Inductor: } Z_L = j\omega L \]
%%%%%%%%%%%%%%%%%%%%%%%%
%%% Circuit analysis %%%
%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Circuit analysis}
\section{Thevenin equivalent circuits}
Say we have an capacitor circuit to analyze:
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [V=$V_S$] (0,3)
to [R=$R_1$] (3,3)
to [R=$R_2$] (6,3)
to [C=$C$, v=$v_C$] (6,0);
\draw (3,3) to [R=$R_3$] (3,0);
\draw (0,0) to (6,0);
\end{circuitikz}
\\
Since this is a linear network, we can simplify it by calculating its \emph{Thevenin equivalent}. Consider the network as seen from the port where the capacitor is attached:
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [V=$V_S$] (0,3)
to [R=$R_1$] (3,3)
to [R=$R_2$, -o] (6,3);
\draw (3,3) to [R=$R_3$] (3,0);
\draw (0,0) to [short, -o] (6,0);
\draw (6,0) to [open, v>=$V_{TH}$] (6,3);
\end{circuitikz}
\\
$V_{TH}$, the open circuit voltage, will be given by the voltage divider formed by $R_3$ and $R_1$:
\[ V_{TH} = \frac{R_3}{R_1 + R_3} \cdot V_S \]
Since no current flows at the port (for the \emph{open circuit} voltage!), $R_2$ doesn't contribute at all.
We also want to measure the resistance ``looking in'' to this port; this will be the Thevenin resistance $R_{TH}$. To do this, we turn off all \emph{independant} voltage and current sources, by replacing all current sources with \emph{opens} and all voltage sources with \emph{short circuits}.\\
Leave dependant sources in the circuit!
\begin{circuitikz}[scale=1.2]
\draw (0,0) to [short] (0,3)
to [R=$R_1$] (3,3)
to [R=$R_2$, -o] (6,3);
\draw (3,3) to [R=$R_3$] (3,0);
\draw (0,0) to [short, -o] (6,0);
\draw (6,0) to [open] (6,3);
\end{circuitikz}
\[ R_{TH} = R_2 + (R_1 || R_3) = R_2 + \frac{R_1 \cdot R_3}{R_1 + R_3} \]
Now that we know the Thevenin voltage $V_{TH}$ and the Thevenin resistance $R_{TH}$, we can replace the circuit with a voltage source of voltage $V_{TH}$ volts in series with a resistor of value $R_{TH}$ ohm, and place the capacitor back into the circuit:
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [V=$V_{TH}$] (0,3)
to [R=$R_{TH}$] (3,3)
to [C=$C$, v=$v_C$] (3,0);
\draw (0,0) to (3,0);
\end{circuitikz}
Our previous circuit has now turned into a simple RC circuit, which is easier to analyze. See the chapter on RC circuits.\\
As a side note, another way of measuring the Thevenin resistance is to short circuit the output node, calculate/measure the short-circuit current (with all sources left intact, of course), and calculate $R_{TH}$ as $\frac{V_{TH}}{I_{SC}}$.\\
In summary:
$\bullet$ Calculate/measure the open circuit voltage $V_{TH}$ at the port\\
$\bullet$ Turn off all independent sources (make short circuits of voltage sources, and open circuits of current sources), but leave dependant sources intact\\
$\bullet$ Calculate/measure the resistance $R_{TH}$ at the port terminal pair\\
$\bullet$ Replace the original circuit with a series circuit of a voltage source (voltage $V_{TH}$ volts), a resistor (resistance $R_{TH}$ ohm) and the element you want to analyze.
\newpage
\section{Norton equivalent circuits}
Nortan equivalent circuits are very similar to Thevenin equivalents, but use a \emph{current source} in \emph{parallel} with a resistor rather than a \emph{voltage source} in \emph{series} with a resistor.
To convert a circuit to its Norton equivalent:
$\bullet$ Calcurate/measure the \emph{short circuit current}, i.e. the current that would flow through the output port if we were to short-circuit it. The result is the Norton current $I_N$.\\
$\bullet$ Turn off all independent sources (make short circuits of voltage sources, and open circuits of current sources), but leave dependant sources intact.\\
$\bullet$ Calculate/measure the resistance at the port terminal pair; the result is the Norton resistance $R_N$.\\
$\bullet$ Replace the original circuit with a parallel circuit of a current source (current $I_N$ amperes), a resistor (resistance $R_N$ ohm) and the element you want to analyze.\\
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [I=$I_N$] (0,3);
\draw (0,3) to (3,3);
\draw (3,3) to [R=$R_N$] (3,0);
\draw (6,3) to [C=$C$, v=$v_C$] (6,0);
\draw (6,0) to (0,0);
\draw (6,3) to (3,3);
\end{circuitikz}
\\
Note that since the method for calculating the equivalent resistance is identical for the Thevenin and Norton methods, $R_N = R_{TH}$.
It is easy to convert between one and the other:
\[ R_N = R_{TH} \]
\[ I_N = \frac{V_{TH}}{R_{TH}} \]
\[ V_{TH} = I_N \cdot R_N \]
%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% Small signal method %%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Small signal method}
\section{Deriving small signal models}
For a device with $i_D = f(v_D)$, the small signal current $i_d$ is given by
\[ \underbrace { \left. \frac{\partial f(v_D)}{\partial v_D} \right|_{v_D=V_D} }_{\displaystyle g_m} \cdot v_d \]
where $v_d$ is the small signal input voltage.\\
In other words, take the partial derivative of the V-I relation, with respect to the voltage. That gives $g_m$, the transconductance. The transconductance multiplied by the small signal input voltage $v_d$ gives the small signal current.\\
As an example, a MOSFET in the saturation region has $i_{DS} = f(v_{GS}) = \frac{K}{2}(v_{GS} - V_T)^2$:
\[ i_{ds} = \underbrace { \left. \frac{ \partial \frac{K}{2}(v_{GS} - V_T)^2 }{\partial v_{GS}} \right|_{v_{GS}=V_{GS} }}_{\displaystyle g_m} \cdot v_{gs} = \underbrace{ K(V_{GS} - V_T)}_{\displaystyle g_m} \cdot v_{gs} \]
Note the difference between $v_{GS}$ (the total gate-to-source voltage), $V_{GS}$ (the DC bias voltage) and $v_{gs}$ (the small signal / incremental voltage). \\
Also note that the value of $g_m$ depends not only on the MOSFET parameters $K$ and $V_T$, but also on the user-chosen DC bias voltage $V_{GS}$.\\
Here's a table of common circuit elements and their small signal equivalents:\\
\begin{tabular}{| l | l |}
\hline
Device & Small signal replacement \\ \hline
Resistor, R $\Omega$ & Resistor, R $\Omega$ \\ \hline
Voltage source, V volts & Short circuit \\ \hline
Current source, I amps & Open circuit \\ \hline
MOSFET in saturation region & VCCS, $i_{ds} = g_m v_{gs}$, $g_m = K(V_{GS} - V_T)$ \\ \hline
MOSFET with gate/drain tied together & Resistor, $\displaystyle \frac{1}{K(V_{DS} - V_T)} \Omega$ (for $v_{DS} > V_T$) \\ \hline
\end{tabular}
\newpage
\section{Multivariable small signal models}
In some cases, it might be necessary to make a small signal model of a device where the current depends on more than one variable. An example (that will be used here) in 6.002x is the (probably hypothetical) ``NewFET'' in the week 6 homework.\\
In this case, we take the partial derivative of the gate-to-source voltage at the bias point times the small signal voltage $vgs$, \emph{plus} the partial derivative of the drain-to-source voltage at the bias point times the small signal voltage $vds$.
First, the properties of the NewFET:
\[
iDS = \begin{cases}
0 & \text{for $v_{GS} < V_T$} \\
K(v_{GS} - V_T) {v_{DS}}^2 & \text{for $v_{GS} \ge V_T$}
\end{cases}
\]
\\
The small signal model will look like this:
\begin{circuitikz}[scale=1.2]
\draw (3,2.5) to [cI=$g_m v_{gs}$, *-*] (3,0);
\draw (5,2.5) to [R=$r_o$, *-*] (5,0);
\draw (5,2.5) to [short, i<=$i_{ds}$] (6,2.5) node {\quad D};
\draw (5,0) to (6,0) [short] node {\quad S};
\draw (7,2.5) to [open, v=$v_{ds}$] (7,0);
\draw (3,2.5) to (5,2.5);
\draw (0,0) to (5,0);
\draw (-0.25,0) node {S} [open] to (0,0);
\draw (0,2.5) [open, v=$v_{gs}$, *-*] to (0,0);
\draw (-0.25,2.5) node {G} [open] to (0,2.5);
\end{circuitikz}
\\
The small signal current $i_{ds}$ will depend on both $v_{gs}$ and $v_{ds}$, in this manner:
\[
i_{ds} = v_{gs} \underbrace{ \left. \frac{\partial i_D}{\partial v_{GS}} \right|_{v_{GS}=V_{GS}}}_{\displaystyle g_m} + v_{ds} \underbrace{ \left. \frac{\partial i_D}{\partial v_{DS}} \right|_{v_{DS}=V_{DS}}}_{\displaystyle 1/r_o}
\]
$g_m$ will be the regular transconductance, calculated the same as with MOSFETs (see the above section):
\[
g_m = \left. \frac{\partial i_D}{\partial v_{GS}} \right|_{v_{GS}=V_{GS}} = \left. \frac{\partial K(v_{GS} - V_T) {v_{DS}}^2}{ \partial v_{GS}} \right|_{v_{GS}=V_{GS}} = K {V_{DS}}^2
\]
$r_o$ will be the \emph{reciprocal} of the partial of $i_D$ with respect to $v_{DS}$:\\
(In other words, the partial will give a conductance, and we want a resistance.)
\[
\frac{1}{r_o} = \left. \frac{\partial i_D}{\partial v_{DS}} \right|_{v_{DS}=V_{DS}} = \left. \frac{\partial K(v_{GS} - V_T) {v_{DS}}^2}{ \partial v_{DS}} \right|_{v_{DS}=V_{DS}} = 2 K V_{DS} (V_{GS} - V_T)
\]
So
\[ r_o = \frac{1}{2 K V_{DS} (V_{GS} - V_T)} \]
From these equations and the circuit diagram, we see that
\[
i_{ds} = \frac{v_{ds}}{r_o} + g_m v_{gs} = v_{ds} \cdot 2 K V_{DS} (V_{GS} - V_T) + K {V_{DS}}^2 \cdot v_{gs}
\]
Note that, although the expression contains a square term (${V_{DS}}^2$), it is still linear, as the square term is a constant - the bias voltage $V_{DS}$ should not change, or the entire small signal model will be incorrect either way.
So, the result is not quite the simplest of expressions, but when the bias constants are replaced with their actual bias values, the result is of the form
\[ i_{ds} = C_{1} v_{ds} + C_{2} v_{gs} \]
where $C_1$ and $C_2$ are constants.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% First-order circuits %%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{First-order circuits}
\section{Series RC circuits}
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [V=$V_S$] (0,3)
to [R=$R$] (3,3)
to [C=$C$, v<=$v_C$] (3,0);
\draw (3,0) to (0,0);
\end{circuitikz}
\\
We start off by writing down a KCL equation for the unknown node voltage $v_C$:\\
\[ \frac{v_C - V_S}{R} + C \frac{dv_C}{dt} = 0 \]
Rewrite the equation to get it in the form we prefer:
\[ RC \frac{dv_C}{dt} + v_C = V_S \]
To solve this first-order linear differential equation, we'll use the method of particular and homogeneous solutions, where we need to find two solutions to the differential equation: the first (the \emph{particular} solution) is \emph{any} solution that makes the equation true:
\[ RC \frac{dv_{Cp}}{dt} + v_{Cp} = V_S \]
We see here that if we pick $v_{Cp} = V_S$, where $V_S$ is a constant, thus making $\frac{dV_S}{dt} = 0$, this equation is indeed true; since the first term becomes 0, all that remains is $V_S = V_S$ - which is clearly true!\\
Thus, we've found the particular solution:
\[ v_{Cp} = V_S \]
The next step in this method is to find a solution to the homogeneous equation, where $V_S$ (the ``input drive'') is zero:
\[ RC \frac{dv_{Ch}}{dt} + v_{Ch} = 0 \]
We need a function such that its derivative is the function itself times a constant. $e^x$ comes to mind - more specifically, the solution will have some (still unknown) coefficients $A$ and $s$, such that:
\[ v_{Ch} = Ae^{st} \]
We substitute $Ae^{st}$ into the homogeneous equation and end up with:
\[ RC \frac{dAe^{st}}{dt} + Ae^{st} = 0 \]
We calculate the derivative and replace the $\frac{d}{dt}$ term with it:
\[ RCAs e^{st} +Ae^{st} = 0 \]
Divide both sides by $Ae^{st}$:
\[ RCs + 1 = 0 \]
\[ RCs = -1 \]
\[ s = -\frac{1}{RC} \]
We've thus found one of our two constants.\\
The total solution to the differential equation will be the \emph{sum} of the particular and homogeneous solutions, so the next step is to add them together:
\[ v_C(t) = v_{Cp}(t) + v_{Ch}(t) \]
\[ v_C(t) = V_S + Ae^{-\frac{1}{RC} \cdot t} \]
All that remains to do is to find the value of the constant $A$. To do so, we substitute $v_C$ for the given initual condition $v_C(0) = V_0$, while setting $t = 0$:
\[ V_0 = V_S + Ae^0 \]
\[ V_0 = V_S + A \]
\[ A = V_0 - V_S \]
We've thus found the full solution:
\[ v_C(t) = V_S + (V_0 - V_S)e^{-\frac{t}{RC}} \]
\newpage
\section{Parallel RC circuit with a current source}
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [I=$I$] (0,3);
\draw (0,3) to (2,3);
\draw (2,3) to [R=$R$] (2,0);
\draw (2,3) to (4,3);
\draw (4,3) to [C=$C$, v=$v_C$] (4,0);
\draw (4,0) to (0,0);
\end{circuitikz}
\\
To save time (and space): the differential equation we end up with is exactly the same as for the series circuit above, with the sole difference that $IR$ replaces $V_S$, where $I$ is the current source drive current, and $R$ is the parallel resistor's resistance.\\
Since the resulting equation
\[ RC \frac{dv_C}{dt} + v_C = IR \]
is of the same form as the one for the series circuit, the solution is also the same:
\[ v_C(t) = IR + (V_0 - IR)e^{-\frac{t}{RC} } \]
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% Second-order circuits %%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Second-order circuits}
\section{Series RLC circuits}
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [V=$V_S$] (0,3)
to [L=$L$, v<=$v_L$] (2,3)
to [R=$R$, v<=$v_R$, i>=$i$] (4,3)
to [C=$C$, v<=$v$] (4,0);
\draw (4,0) to (0,0);
\end{circuitikz}
We know that the current $i$ relates to the capacitor voltage:
\[ i = C \frac{dv}{dt} \]
Since this is a series circuit, that current goes through all elements.
By KVL, we can add up the voltage drops around the loop, with the proper sign. I'll go clockwise, and start in the bottom left:
\[ -V_S + v_L + v_R + v = 0 \]
If we solve for $V_S$:
\[ V_S = v_L + v_R + v \]
Now, let's use the element laws for the inductor and resistor, and substitute them into the above:
\[ V_S = L \frac{di(t)}{dt} + R i(t) + v(t) \]
\newpage
We know that $i(t) = C \frac{dv}{dt}$, so let's substitute that back in. Let's also differentiate, in case of the inductor, to reduce the mess of nested differentiation operators:
\[ V_S = LC \frac{d^2v(t)}{dt^2} + RC \frac{dv}{dt} + v(t) \]
There we go; we now have a second-order, linear, constant coefficient ordinary differential equation.\footnote{I just love these overly verbose classifications.}\\
We'll use the same method to solve it as we did for the first-order ones, namely the method of particular and homogeneous solutions. First we first the particular solution (\emph{any} solution that makes the equation true), then the homogeneous solution (a solution to the equation with the input drive $V_S$ set to 0), and finally we add the two together to find the total solution.\\
As for the particular solution, just as in the first-order case, we try $v(t) = V_S$ and see that the two derivatives both go to zero (as $V_S$ is a constant), and so we end up with
\[ V_S = V_S \]
which tells us that $V_S$ indeed is a particular solution. On to the homogeneous one, i.e. the solution to
\[ LC \frac{d^2v(t)}{dt^2} + RC \frac{dv}{dt} + v(t) = 0\]
Again, like in the first-order case, we try the form
\[ v(t) = Ae^{st} \]
where $A$ and $s$ are constants we'll have to find later. So, we substitute $Ae^{st}$ for $v(t)$, and differentiate where necessary, which gives us
\[ LC As^2 e^{st} + RC As e^{st} + A e^{st} = 0 \]
We can cancel out a ton of the above, by dividing both sides by $A e^{st}$:
\[ LC \cancel{A} s^2 \cancel{e^{st}} + RC \cancel{A}s \cancel{e^{st}} + \cancel{A e^{st}} = 0 \]
What remains is
\[ LC s^2 + RC s + 1 = 0 \]
Divide by LC throughout:
\[ s^2 + \frac{R}{L} s + \frac{1}{LC} = 0 \]
We've thus arrived at the \emph{characteristic equation}, which describes most details of the circuit's behavior. There's a canonical form to write such an equation:
\[ s^2 + 2\alpha s + {\omega_0}^2 = 0 \]
So, in the case of the series RLC circuit, the values of $\alpha$ and $\omega_0$ would have to be
\[ \alpha = \frac{R}{2L} \]
and
\[ \omega_0 = \frac{1}{\sqrt{LC}} \]
$\alpha$ relates to the damping factor of the circuit (sometimes the damping factor, $\displaystyle \zeta = \frac{\alpha}{\omega_0}$, is used) - that is, the ringing second-order circuits can produce will decay faster for a larger value of $\alpha$. Meanwhile, $\omega_0$ is the \emph{undamped resonance frequency} of the circuit (in radians/second), the same as in undriven LC circuits. However, this might not be the frequency of interest in a RLC circuit - more on that later.\\
Let's get back to solving the differential equation. Remember, we only found the homogeneous solution - we still haven't figured out the values of $s$ or $A$.
The next step would be to find the roots of the characteristic equation:
\[ s^2 + 2\alpha s + {\omega_0}^2 + 0 \]
The quadratic formula will work nicely on this. The two roots are:
\[ s_1 = -\alpha + \sqrt{\alpha^2 - {\omega_0}^2} \]
\[ s_2 = -\alpha - \sqrt{\alpha^2 - {\omega_0}^2} \]
The full solution to the homogeneous solution will then be
\[ v_H = A_1 e^{s_1 t} + A_2 e^{s_2 t} \]
However, we've now found $s_1$ and $s_2$, so we can fill them in:
\large
\[ v_H = A_1 e^{(-\alpha + \sqrt{\alpha^2 - {\omega_0}^2}) t} + A_2 e^{(-\alpha - \sqrt{\alpha^2 - {\omega_0}^2}) t} \]
\normalsize
Then, the total solution (prior to finding $A_1$ and $A_2$, and also prior to simplifying) is:
\[ v(t) = V_S + A_1 e^{(-\alpha + \sqrt{\alpha^2 - {\omega_0}^2}) t} + A_2 e^{(-\alpha - \sqrt{\alpha^2 - {\omega_0}^2}) t} \]
After some magic\footnote{Sorry, but I don't quite follow the entire process myself yet. Since this part was cut out from the lectures in order to save time (after ~14 videos going through this process so far), I will do the same.} and making the definition $\omega_d = \sqrt{{\omega_0}^2 - \alpha^2}$, the total solution ends up looking like this, for the case $\omega_0 > \alpha$, i.e. the underdamped case:
\[ v(t) = V_I + K_1 e^{-\alpha t} \cos{\omega_d t} + K_2 e^{-\alpha t} \sin{\omega_d t} \]
Evaluating the above at t = 0 with the initial condition $v(0) = 0$ gives us:
\[ 0 = V_I + K_1 \]
\[ K_1 = -VI \]
Evaluating at t = 0 with the other initial condition, $\displaystyle i(0) = C \frac{dv}{dt} = 0$, gives, after the differentiation and substitution for $t = 0$:
\[ 0 = -K_1 \alpha + K_2 \omega_d \]
We know that $K_1 = -V_I$:
\[ 0 = V_I \alpha + K_2 \omega_d \]
\[ K_2 \omega_d = -V_I \alpha \]
\[ K_2 = -\frac{V_I \alpha}{\omega_d} \]
Thus, finally, the full equation that governs the capacitor voltage of the series RLC circuit, in the underdamped case ($\omega_0 > \alpha$) is:
\[ v(t) = V_I - V_I e^{-\alpha t} \cos{\omega_d t} - \frac{V_I \alpha}{\omega_d} e^{-\alpha t} \sin{\omega_d t} \]
Using the scaled sum of sines trig identity, we can turn this cosine minus sine business into a single cosine, and end up with:
\[ v(t) = V_I - V_I \frac{\omega_0}{\omega_d} e^{-\alpha t} \cos{(\omega_d t - \arctan{(\frac{\alpha}{\omega_d})})} \]
While the equation is rather involved, it's still very clear that the two main features are a cosine multiplied by a decaying exponential, which will give us exactly the kind of waveform we see in a damped second-order system.\\
However, another extraordinary detail hides in the equation. Note how the capacitor voltage is $V_I$ minus $V_I$ times a bunch of things that is always positive... except for the cosine. In fact, if $V_I$ is 5 volts, this circuit will oscillate between 0 and $10$ volts - yes, the capacitor voltage will be twice as high as the input step! Rather incredible.
\subsection{Summary}
Since this section was by \emph{far} the longest of this document so far, I thought a summary of the relevant formulas could be useful. Remember that most of these only apply to series RLC circuits, though some definitions (such as $\omega_d$) are universal.\\
Where v(t) denotes the capacitor node voltage in the circuit shown (far) above - also, remember that this is for the \emph{underdamped} case, i.e. $\displaystyle \omega_0 > \alpha$:
\[ v(t) = V_I - V_I \frac{\omega_0}{\omega_d} e^{-\alpha t} \cos{(\omega_d t - \arctan{(\frac{\alpha}{\omega_d})})} \]
where
\[ \alpha = \frac{R}{2L} \text{ rad/s}\]
\[ \omega_0 = \frac{1}{\sqrt{LC}} \text{ rad/s} \]
\[ \omega_d = \sqrt{{\omega_0}^2 - \alpha^2} \text{ rad/s} \]
Other formulas that may be useful are:
\[ Q = \frac{\omega_0}{2\alpha} \text{ (dimensionless)} \]
\[ \text{Natural period} = \frac{2\pi}{\omega_0} \text{ seconds} \]
\[ \text{Natural frequency} = \frac{\omega_0}{2\pi} = \frac{1}{2\pi \sqrt{LC}} \text{ Hz} \]
The quality factor, $Q$, can (for now) be thought of as the approximate number of oscillations that will occur before the ringing dies out due to the damping, though the actual meaning is more precisely defined.\\
One thing that is important to note, and that in truth should have been brought up in more detail, is that for the special case where $R = 0$, the circuit becomes a pure LC circuit, which will (in theory) oscillate forever. In practice, of course, parasitic resistances will make sure that doesn't happen - unless the circuit is superconducting.\\
Why does it oscillate forever? Well, it's rather intuitive, at least if you accept that it will oscillate at all: it oscillates until the energy stored in the circuit is small enough that we see the voltages in the circuit to be constants. The only way energy stored is reduced at all is by resistances; without them, it rings forever.\\
Mathematically, this is easy to see from the above equations: $\displaystyle Q = \frac{\omega_0}{2\alpha} \to \infty$ as $2\displaystyle \alpha = \frac{R}{L} \to 0$.
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%% Sinusoidal Steady State, Impedance and Filters %%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\chapter{Sinusoidal Steady State, Impedance and Filters}
\section{Sinusoidal Steady State}
Instead of showing the entire, rather long path from a drawn circuit to the impedance model, I will take considerable shortcuts by simply skipping some parts. The reason is that I don't find it that important to remember fully; any good book on the subject should show it, though. I'll try to cover what \emph{is} important to remember, however!\\
Thus, the beginning of this chapter might be a bit hard to follow. However, this isn't an electronics book (which I would be very underqualified to write) - I assume that all readers already know most of this, and just wanted a little refresher.
We'll try to analyze this circuit, for sinusoidal input:\\
\begin{circuitikz}
\draw (0,0) node [ground] {} to [sV=$v_I$] (0,3)
to [R=$R$] (3,3)
to [C=$C$, v<=$v_C$] (3,0);
\draw (3,0) to (0,0);
\end{circuitikz}
where $v_I$ is a sinusoidal input, e.g. $V_I \cos{(\omega t)}$.
So, let's get started. We write down the differential equations governing the dynamics of this circuit:
\[ RC \frac{dv_C(t)}{dt} + v_C(t) = V_I \cos{(\omega t)} \]
If this were anything like the time-domain RC circuit analysis for step inputs, we'd now go and find particular and homogeneous solutions for the circuit, add them, and solve for constants using initial conditions.\\
However, this would not be a very nice analysis due to the massive amount of trigonometry that would turn out to be involved due to the sinusoidal drive. Instead, we'll go down a different path, and see that doing so brings us some rather massive advantages in simplicity - especially later on, when we've derived and understood the impedance model, that removes the need to use differential equations for many kinds of analyses!\\
So, instead of analysing the circuit for the input signal $V_I \cos{(\omega t)}$, we'll analyse it for the input $V_i e^{s t}$.\\
Our first analysis step is to attempt to find a particular solution that works with the given differential equation, which we've now modified (by choosing a different input) to be:
\[ RC \frac{dv_C(t)}{dt} + v_C(t) = V_i e^{s t} \]
Let's try a solution of the form $V_p e^{s t}$, and substitute that in:
\[ RC \frac{d V_p e^{s t}}{dt} + V_p e^{s t} = V_i e^{s t} \]
Let's carry out the differentiation:
\[ RC V_p s e^{s t} + V_p e^{s t} = V_i e^{s t} \]
Let's cancel out the $\displaystyle e^{s t}$ terms that are common to all terms:
\[ RC V_p s \cancel{e^{s t}} + V_p \cancel{e^{s t}} = V_i \cancel{e^{s t}} \]
\[ RC V_p s + V_p = V_i \]
Rearrange, then solve for $V_p$:
\[ V_p (RC s + 1) = V_i \]
\[ V_p = \frac{V_i}{1 + sRC} \]
However, remember that we crossed out all the $\displaystyle e^{s t}$ terms. The full particular solution is
\[ V_p e^{s t} = \frac{V_i}{1 + sRC} e^{s t} \]
Now, let's assume that $s = j \omega$, where j is the imaginary unit $j = \sqrt{-1}$. We do the substitution, and get the particular solution
\[ \text{Particular solution } = \frac{V_i}{1 + j \omega RC} e^{j \omega t} \]
Note that because this chapter deals with sinusoidal \emph{steady state} analysis, we will ignore the homogeneous solution as it relates only to transients, i.e. what happens in the circuit \emph{before} it reaches steady state.
Now we have a complex number. $\displaystyle V_p = \frac{V_i}{1 + j \omega RC}$ is the \emph{complex amplitude}, which is an extremely useful variable. We'll see later that it can describe the entire frequency-domain behavior of the circuit to sinusoidal input, \emph{and} most time-domain behavior as well.\\
However, remember that our goal was to find the response to sinusoidal input, not exponential input as we've actually done! What is the connection?\\
The answer lies in Euler's formula, which states
\[ e^{j \omega t} = \cos{(\omega t)} + j \sin{(\omega t)} \]
which can be proved e.g. via Taylor series.
Thus, we can take the real part of the exponential input, to yield the input we sought after:
\[ V_i \cos{(\omega t)} = Re\left[ V_i e^{j \omega t} \right] \]
where Re[z] gives the real part of the complex number $z$, i.e. $Re[x + jy] = x$.
Thus, we can use a ``inverse superposition'' argument: for linear circuits, the response to the input $Re[z]$ should be of the form $Re[f(z)]$. That is, the real part of the output should be a function of the real part of the input.\\
\subsection{Time-domain analysis}
I promised earlier that $V_p$ could describe the time domain behaviour of the circuit. Let's see how! Also note that we will soon see a much faster and easier method of writing $V_p$ down, than to work with the differential equations for a circuit.\\
The relation between $V_p$ and the time-domain behavior of the circuit is:
\[ v_C(t) = \left| V_p \right| \cos{(\omega t + \angle V_p)} \]
Note that we derived the time-domain behavior for the circuit simply by taking the magnitude and angle of the complex amplitude $V_p$. However, the above equation only specifies the relation, without using circuit values (i.e. $R$, $C$ and $V_i$). Let's calculate the magnitude and phase angles.
\[ | V_p | = \left| \frac{Vi}{1 + j \omega RC} \right| = \frac{|V_i|}{|1 + j \omega RC|} \]
The magnitude of $V_i$ is simply the absolute value, and so we'll leave it as-is and assume it is positive.\\
The denominator is a complex number, and so we calculate the magnitude as always with complex numbers:
\[ |a + jb| = \sqrt{a^2 + b^2} \]
\[ |1 + j \omega R C| = \sqrt{1 + (\omega R C)^2} \]
Combining the two answers:
\[ |V_p| = \frac{Vi}{\sqrt{1 + (\omega R C)^2}} \]
Thus step one is complete. Next, let's find the phase angle.\\
The angle of divided complex numbers subtracts (while magnitude divides). The angle of $V_i$, being fully real, is 0. (The angle of a fully imaginary number is $\displaystyle \frac{\pi}{2}$, since the imaginary axis is the up/down axis, and $\displaystyle \frac{\pi}{2}$ is the number of radians to rotate from the real (x) to the imaginary (y) axis.)\\
\begin{align*}
\angle V_p &= \angle V_i - \angle (1 + j \omega R C) \\
&= 0 - \arctan{(1, \omega R C)} \\
&= -\arctan{(1, \omega R C)}
\end{align*}
... where arctan is the two-argument arctan function. $\arctan{(x, y)}$ is equal to $\displaystyle \arctan{(\frac{y}{x})}$ for some inputs, but not all, which is why the two-argument one (also known as ``atan2'' in many programming languages) is used; the regular form can't differentiate between quadrants uniquely.\\
Thus, combining the time-domain equation with the now-known values of $|V_p|$ and $\angle V_p$:
\[ v_C(t) = \left| V_p \right| \cos{(\omega t + \angle V_p)} \]
\[ v_C(t) = \frac{Vi}{\sqrt{1 + (\omega R C)^2}} \cos{(\omega t - \arctan{(1, \omega R C)} )} \]
As for the frequency response, we can also calculate and plot the magnitude of the \emph{transfer function}, i.e. the ratio between the capacitor voltage and the input voltage.
\[ \left| \frac{V_p}{V_i} \right| = \left| \frac{1}{1 + j \omega R C} \right| = \frac{1}{\sqrt{1 + (\omega R C)^2}} \]
... and the same goes for the phase of the transfer function:
\[ \angle \frac{V_p}{V_i} = \angle \left( \frac{1}{1 + j \omega R C} \right) = \angle 1 - \angle (1 + j \omega R C) = 0 - arctan(1, \omega R C) = -arctan(1, \omega R C) \]
Now that we know how to find the magnitude and phase of the complex amplitude rather easily (compared to solving the differential equations!), let's see if we can find an easier way to writing the expression down in the first place, so that we can solve linear circuits with resistors, capacitors and inductors in no-time.
\section{The Impedance Model}
For the resistor, Ohm's law applies equally well to the complex amplitudes of voltages and currents as it does to the usual inputs:
\[ V_r = I_r \cdot R \]
However, the interesting things turn up when we look at the capacitor and inductor. Let's consider a lone capacitor and its relation between current and voltage.\\
\begin{circuitikz}
\draw (0,0) to [C=$C$, v>=$v_C$, i<=$i_C$] (0,3);
\end{circuitikz}
We know from earlier that
\[ i_C = C \frac{dv_C}{dt} \]
But what happens when we use complex amplitudes instead? Say the current is of the form $\displaystyle I_c e^{s t}$ and the voltage across it of the form $V_c e^{st}$ - what then? Well, let's see. We start by substituting the values:
\[ I_c e^{st} = C \frac{d V_c e^{s t}}{dt} \]
Perform the differentiation:
\[ I_c e^{s t} = C V_c s e^{st} \]
Cancel out the $\displaystyle e^{s t}$ terms from both sides, and solve for $V_c$:
\[ I_c \cancel{e^{s t}} = C V_c s \cancel{e^{st}} \]
\[ I_c = C V_c s \]
\[ V_c = \frac{1}{sC} I_c \]
Hmm, it looks like they have a very simple relation indeed! To simplify even further, let's set $\displaystyle Z_c = \frac{1}{sC}$:
\[ V_c = Z_c I_c \]
Incredible! The voltage is proportional to the current through this $Z_c$ variable, just like a resistor's voltage drop is $V = IR$!\\
$Z_c$ is known as the capacitor's \emph{impedance}, and is frequency dependant: remember that $\displaystyle Z_c = \frac{1}{sC}$, and $s = j \omega$, where $\omega$ denotes the angular frequency of the input sinusoid. Therefore, we see that the higher the frequency, the lower the impedance.\\
What this means is that the capacitor will have high impedance - which is similar (but not equal!) to high resistance - to low-frequency signals, to the point where it completely blocks DC ($\omega = 0$), but has a low impedance for high-frequency signals. As the frequency increases, the capacitor impedance decreases, and as $\omega \to \infty$, the impedance goes to zero and the capacitor acts as a short circuit.\\
Let's see how an inductor deals with this.\\
\begin{circuitikz}
\draw (0,0) to [L=$L$, v>=$v_L$, i<=$i_L$] (0,3);
\end{circuitikz}
Again, we know that
\[ v_L = L \frac{di_L}{dt} \]
Let's see what happens for complex input, where $v_L = V_l e^{st}$ and $i_L = I_l e^{s t}$:
\[ V_l e^{s t} = L \frac{d I_l e^{s t}}{dt} \]
Same as before; we perform the differentiation, cancel out the $e^{st}$ terms, and solve for $V_l$:
\[ V_l e^{s t} = L \frac{d I_l e^{s t}}{dt} \]
\[ V_l e^{s t} = L I_l s e^{s t} \]
\[ V_l \cancel{e^{s t}} = L I_l s \cancel{e^{s t}} \]
\[ V_l = L I_l s \]
\[ V_l = I_l \cdot s L \]
Again, the voltage is related to the current through a simple algebraic expression! The impedance of the inductor is $Z_L = sL$, where (again) $s = j \omega$.
\[ V_l = Z_L I_l \]
This is an incredible powerful technique, as we'll soon see. We can now write down the frequency (and time) behaviour of circuits by inspection, without writing a single differential equation!\\
Impedances not only look like resistances (impedance times current equals voltage), but can be used in the same way, to create things like voltage dividers using capacitors and inductors! This is the basis of \emph{filters}, which we'll look at soon.
So, in summary, the relevant models are:\\
\begin{circuitikz}
\draw (0,0) to [C=$C$, v>=$V_c$, i<=$I_c$] (0,3);
\draw (2,0) to [L=$L$, v>=$V_l$, i<=$I_l$] (2,3);
\draw (4,0) to [R=$R$, v>=$V_r$, i<=$I_r$] (4,3);
\end{circuitikz}
Where the relations between voltage amplitudes and current amplitudes are given as
\[ V_r = I_r Z_R \]
\[ V_c = I_c Z_L \]
\[ V_l = I_l Z_L \]
... and the impedences are given as
\[ Z_R = R \]
\[ Z_C = \frac{1}{j\omega C} \]
\[ Z_L = j \omega L \]
These impedances can be used exactly as resistances. For example, in a series RC circuit, the current through the circuit is given by
\[ \frac{V_i}{R + Z_C} \]
just as you would expect.
\subsection{RC circuit example}
\begin{circuitikz}[scale=1.2]
\draw (0,0) node [ground] {} to [sV=$V_i$] (0,3)
to [R=$R$] (3,3)
to [C=$Z_C$, v<=$V_c$] (3,0);
\draw (3,0) to (0,0);
\end{circuitikz}
Let's analyze the above circuit, to find the voltage across the capacitor. The impedance is $Z_C = \frac{1}{sC}$, and we can use that to create a voltage divider relationship:
\[ V_c = V_i \cdot \frac{Z_C}{Z_C + R} = V_i \cdot \frac{ \frac{1}{sC} }{ \frac{1}{sC} + R} \]
Let's multiply out by $\displaystyle \frac{sC}{sC}$ to clean things up:
\[ V_c = V_i \cdot \frac{1}{1 + sRC} \]
This is another HUGE result; look at the denominator: what he have there is the characteristic equation for the series RC circuit! We can use that to find everything we want to know about the time-domain behavior of the circuit, such as the natural frequency, bandwidth, damping factor and more - without a single differential equation; not to mention that the same technique works in more complex circuits, such as a series-parallel combination of resistors, capacitors and inductors!
\subsection{Impedance Model Summary}
Before trying a few more complex circuits, let's summarize the method we'll use, and clarify the meaning of a few variables.
\begin{enumerate}
\item Replace sinusoidal sources by their complex amplitudes. For example, $v_I \cos{(\omega t)}$ would simply become $V_i$, which represents a complex amplitude (capital letter, small subscript).
\item Replace the elements with their impedance models; resistors don't change, capacitors turn in to $\displaystyle Z_C = \frac{1}{sC} $ ``resistors'' and inductors into $Z_L = sL$ ``resistors'', where $s = j \omega$.
\item Calculate the complex amplitudes for the nodes that matter, using the normal circuit analysis methods e.g. the node method, Thevenin equivalents, etc, treating impedances exactly as resistances.
\item If necessary, find the time-domain behavior from the complex amplitudes: if you have the complex amplitude $V_a$, the time-domain behavior of that amplitude is described by
\[ v_A(t) = |V_a| \cos{(\omega t + \angle V_a)} \]
\end{enumerate}
\subsection{Finding the characteristic equation}
What we generally want to do here is to write the expression as a proper polynomial: no $\displaystyle \frac{1}{s}$ terms are allowed, and the denominator must be of a higher degree than the numerator. Note that there may be $s$ terms in the numerator, which are then to be ignored, if the polynomial is in proper form.\\
As an example, let's look at the \emph{parallel} RLC circuit, to avoid overfocusing on the series one.\\