txline (3)
txline (3)
txline (3)
Shouri Chatterjee
October 20, 2024
∆x
1
One can construct other distributed circuits. The analyses of other dis-
tributed circuits are typically similar to the analysis we present here for
transmission lines.
2 Analysis technique
Let us first evaluate Kirchoff’s equations for the transmission-line section in
Fig. 1. Kirchoff’s voltage and current laws give the following equations, (1),
(2), respectively.
di(x, t)
v(x, t) − v(x + ∆x, t) = L′ ∆x + i(x, t)R′ ∆x (1)
dt
dv(x + ∆x, t)
i(x, t) − i(x + ∆x, t) = C ′ ∆x + v(x + ∆x, t)G′ ∆x (2)
dt
For this treatment of transmission lines, let us assume that they are loss-less,
i.e., R′ and G′ are zero. We will extend our analysis of loss-less transmission
lines to lossy transmission lines later. Further, let us analyze transmission-
line sections of infinitesimally small length, at the limit of ∆x tending to
zero. (1) then reduces to:
2
The Telegrapher’s equations:
∂v ∂i
= −L′ (6)
∂x ∂t
∂i ′ ∂v
= −C (7)
∂x ∂t
The solution to the Telegrapher’s equations is the solution to the
transmission-line problem.
∂ 2v ∂ 2i
= −L′ 2
∂x∂t ∂t
∂ 2i 2
′ ∂ v
and = −C
∂x2 ∂t∂x
2 2
∂ i ′ ′∂ i
This gives: = LC 2 (8)
∂x2 ∂t
Likewise, we can take the time derivative of (7) and compare it with the
spatial derivative of (6). This results in a similar equation:
∂ 2v 2
′ ′∂ v
= L C (9)
∂x2 ∂t2
(8) and (9) are not coupled to each other. (8) and (9) happen to be the
same equation, one with voltage as the variable and the other with time.
In mathematics literature, the form of this equation is known as the wave
equation.
The solution to the wave equation is the sum of a forward-moving wave
and a backward-moving wave. If f (x) is any function of the variable x, then
f (x − ct) is a forward-moving wave, f (x + ct) is a backward-moving wave. c
is the velocity of the wavefront.
Plugging in f (x − ct) as a solution to i(x, t) in (8), we get:
3
f (x − ct0 )
f (x − ct1 )
f (x − ct2 )
Figure 2: One can jerk one end of a rope to release a wavefront on the rope.
The wave is an example of f (x − ct).
√
In other words, the velocity of the wavefront, c, is given by 1/ L′ C ′ . Plug-
ging f (x+ct) as a solution gives an identical result. Working with (9) instead
of (8) gives the same result, but for v(x, t).
For a second-order differential equation, there should be two degrees of
freedom. The two degrees of freedom are accounted for when we indepen-
dently choose functions for the forward and backward moving waves.
The general solution for both, v(x, t) and i(x, t), is therefore:
f1 (x − ct) + f2 (x + ct)
√
f1 and f2 are arbitrary functions. c = 1/ L′ C ′ .
4
The equations (10), (11), now can be plugged into the Telegrapher’s equa-
tions, (6) and (7).
V+′ (x − ct) + V−′ (x + ct) = −L′ (−cI+′ (x − ct) + cI−′ (x + ct)) (12)
I+′ (x − ct) + I−′ (x + ct) = −C ′ (−cV+′ (x − ct) + cV−′ (x + ct)) (13)
In (12) and (13), we can solve for I√+′ and I−′ in terms of V+′ and V−′ .
Further, plugging in the value of c as 1/ L′ C ′ , we obtain the following.
length, for simple geometries of transmission lines, will reveal that c is the speed of light
in the given medium. Sometimes, this can be a quick way to estimate C ′ if one knows L′ ,
or vice-versa.
5
ZS
Vin Z0 ZL
6 Reflection coefficient
Consider a transmission line of characteristic impedance Z0 , terminated by
a load, ZL , as shown in Fig. 3.
At the load end, if the forward and backward moving waves are V+ and
V− respectively, then the voltage is V+ + V− , and the current into the load
is (V+ + V− )/ZL . The current is also the sum of the forward and backward
moving current waves, (V+ − V− )/Z0 .
V+ + V− V+ − V−
=
ZL Z0
V+ + V− ZL
or =
V+ − V− Z0
V− ZL − Z0
or ΓL = = (17)
V+ ZL + Z0
6
When ZL = Z0 , the transmission line is said to be matched at the load.
The transmission line is matched at the source if ZS = Z0 .
V0 − Vx Vx Z0
= or, Vx = V0 (18)
ZS Z0 ZS + Z0
(18) shows that the voltage appearing at the input of the transmission line is
the result of a resistive division of the applied voltage as if the transmission
line is a resistor of value Z0 . The transmission line is not a resistor Z0 , and
the resistive division appears to hold only initially.
The forward-moving wave, Vx , moves forward at a velocity c. After a
duration of l/c (where l is the length of the transmission line), it reaches
the load. At the load end, the wave Vx impinges on the load and creates a
reflection, ΓL Vx , at time t = (c/l)+ . Now the voltage at the load is Vx (1+ΓL ).
The backward-moving wave is now ΓL Vx .
The backward-moving wave, ΓL Vx , moves backward and impinges on the
source resistor at t = 2l/c. The reflection coefficient at the source (now look
at the transmission line backward, the backward-moving wave is your new
−Z0
forward direction, the source resistor is your new load resistor) is ΓS = ZZSS +Z0
.
The impinging backward-moving wave will cause a reflection, a new forward-
moving wave, ΓS ΓL Vx , at time t = (2l/c)+ . The voltage at the source is now
going to be Vx (1 + ΓL + ΓL ΓS ).
The waves slosh back and forth forever - till the voltages settle to their
final values. Graphically, this process is described in the space-time chart
of Fig. 4. At any point in space-time, the voltage is the sum of the voltage
waves that have passed till then.
The voltage at the load as the system settles to steady-state is going to
7
t=0 V+
ΓL V + t = l/c
t = 2l/c ΓL ΓS V +
Γ2L ΓS V+ t = 3l/c
t = 4l/c Γ2L Γ2S V+
Γ3L Γ2S V+ t = 5l/c
t = 6l/c Γ3L Γ3S V+
Γ4L Γ3S V+ t = 7l/c
t = 8l/c Γ4L Γ4S V+
x=0 x=l
be given by:
∞
X ∞
X
lim V (l) = (Vx ΓiS ΓiL + Vx ΓiS Γi+1
L ) = Vx ΓiS ΓiL (1 + ΓL )
t→∞
i=0 i=0
∞
X 1 + ΓL
= Vx (1 + ΓL ) (ΓL ΓS )i = Vx (19)
i=0
1 − ΓL ΓS
The result is indeed intuitive. After all, at DC, the inductors and capacitors
in the transmission line are all short-circuits and open-circuits, respectively.
The short and open substitutions imply that at DC, the transmission line is
just a wire. Voltage division between ZL and ZS is therefore natural.
8
v(l)
0
v(l/2) T 3T 5T 7T 9T 11T 13T
2
1
0
Figure 5: Voltage waveforms at the far end of the transmission line and at
the middle of the transmission line when the load is an open circuit, and the
source resistance is 0. T is the time the wave takes to travel from one end of
the line to the other.
9
ZS L0.5 L1.5 L2.5
vin (t) C0 C1 C2 C3 ZL
Exercise 1. Show that the voltage at the source end of the transmission
line also settles to Vin ZL /(ZL + ZS ) as time progresses. (i.e., the same
result as the load end of the transmission line.)
Exercise 2. Plot the voltage at the load end of the transmission line as
a function of time, when Z0 is 50 Ω, ZL is 100 Ω, and ZS is 200 Ω.
8 Numerical simulation
Numerically simulating transmission lines is not simple. However, we can
break the transmission line into discrete sections and simulate it like an ordi-
nary RLC circuit with techniques studied in Chapter num-. Unfortunately,
simulation is always heavy on resources and slow.
The first step for numerical simulation is to approximate the transmission
line as a cascade of K sections (Fig. 6). K, the number of sections, should
be chosen such that the length of each section is much smaller than the
maximum frequency of interest. The simulation will not be able to capture
frequencies higher than this.
In Fig. 6, all the inductors labeled Lx.5 are of value L′ d/K where d is
the length of the line, L′ is the inductance per unit length, K is the number
of sections. The capacitors C0 and CK are 12 C ′ d/K, all other capacitors are
C ′ d/K, where C ′ is the capacitance per unit length. We have broken up
the first and last capacitors into halves to add symmetry to the model. One
may visualize each section as a pi-network of 21 C ′ d/K, L′ d/K, and 12 C ′ d/K.
Other components, such as R′ and G′ , could be added into the model.
10
The second step is to write out the A matrix to solve the circuit using
the state-space equations of (ttt-6). For this, we identify the state variables
as all the inductor currents (left to right) and all the capacitor voltages (top
to bottom). In our circuit, we will have 2K + 1 state variables. In the case
of K = 3, ′
X = v0 i0.5 v1 i1.5 v2 i2.5 v3
Then we can write out the KVL/KCL equations to obtain the derivatives of
the state variables. This operation is readily done as follows.
11
8 Zs = 25; Zl = 100;
9
10 % I n d u c t a n c e and c a p a c i t a n c e p e r s e c t i o n
11 Lps = Lp * d / K ; Cps = Cp * d / K ;
12 Rps = Rp * d / K ;
13
14 % Number o f s t a t e v a r i a b l e s = 2K+1
15 Xinit = zeros (2* K +1 , 1);
16
17 % Time r e s o l u t i o n
18 t_res = d / c / K /100;
19 t_stop = 3* d / c ;
20
21 % Input
22 Vinput = zeros (1 ,round( t_stop / t_res )+1);
23 Vinput (1 ,round(10 e -9/ t_res ):end)=1; % S t e p f u n c t i o n
24 % Any o t h e r i n p u t waveform can be c r e a t e d . For example :
25 % Gaussian p u l s e a p p l i e d from 10 ns t o 30 ns
26 % t t = −10e −9: t r e s : 1 0 e −9;
27 % Vinput ( 1 , round (10 e−9/ t r e s ) : round (30 e−9/ t r e s ) ) = exp (−( t t /2 e − 9 ) . ˆ 2 ) ;
28
29 % Initialize
30 Xs = zeros (2* K +1 ,round( t_stop / t_res +1));
31 A = zeros (2* K +1 , 2* K +1); BU = zeros (2* K +1 ,1);
32
33 % Define A
34 f o r n =1: K
35 A (2* n ,2* n -1)=1/ Lps ;
36 A (2* n ,2* n )= - Rps / Lps ;
37 A (2* n ,2* n +1)= -1/ Lps ;
38 end;
39 f o r n =1: K -1
40 A (2* n +1 ,2* n )=1/ Cps ;
41 A (2* n +1 ,2* n +2)= -1/ Cps ;
42 end;
43 A (1 ,1)= -2/ Cps / Zs ; A (1 ,2)= -2/ Cps ;
44 A (2* K +1 ,2* K )=2/ Cps ; A (2* K +1 ,2* K +1) = -2/ Zl / Cps ;
45
46 % Prepare f o r s i m u l a t i o n
47 h = t_res ;
48 inv_be = inv (eye (2* K +1) - h * A ); % For BE
49 %i n v t r = i n v ( e y e (2∗K+1)−h /2∗A) ; % For t r a p e z o i d a l
50 %i n v g 2 = i n v (3∗ e y e (2∗K+1)−2∗h∗A) ; % For Gear2
51
52 % Simulate
53 f o r k =2:(round( t_stop / t_res ))
54 BU (1) = 2* Vinput (1 , k )/ Zs / Cps ;
55 %Xs ( : , k ) = i n v t r ∗ ( e y e (2∗K+1)+h /2∗A)∗ Xs ( : , k−1)+h∗BU; %Trap
56 %Xs ( : , k ) = i n v g 2 ∗ (4∗ Xs ( : , k−1)−Xs ( : , k −2)+2∗h∗BU) ; % Gear2
57 Xs (: , k ) = inv_be * ( Xs (: ,k -1)+ h * BU ); % BE
58 end;
59
60 % P l o t v o l t a g e s a t i n p u t , b e g i n n i n g o f t x l i n e , end o f t x l i n e
61 vbegin = Xs (1 ,:); % V o l t a g e w i t h time a t i n p u t
62 vend = Xs (2* K +1 ,:); % V o l t a g e w i t h time a t o u t p u t
63 f i g u r e (1); c l f ;
64 plot ([0: t_res : t_stop ] , Vinput , ’r ’ ); hold on ; grid on ;
65 plot ([0: t_res : t_stop ] , vend , ’b ’ ); hold on ; grid on ;
66 plot ([0: t_res : t_stop ] , vbegin , ’k ’ ); hold on ; grid on ;
67
68 % Animate , v o l t a g e w i t h d i s t a n c e
69 x = 0: d / K : d ;
12
1 vin (t)
v0 (t)
0.8 vL (t)
Voltage (V)
0.6
0.4
0.2
70 f i g u r e (2); c l f ;
71 f o r k =1:100:round( t_stop / t_res +1)
72 plot (x , Xs (1:2:end, k ));
73 ylim ([ -0.2 , 1.2]);
74 pause (0.001);
75 end;
On simulation, one will notice that the trapezoidal method gives signif-
icant ringing; ringing with Gear-2 is somewhat lesser, and the least ringing
is in BE. However, this ringing is unavoidable. As soon as we have sectioned
the transmission line, we have limited the maximum observable simulation
frequency. As such, the frequencies higher than this limit are not observed
in the output of the simulation. Since a step input has infinite bandwidth,
the manifestation of this band-limiting is a ringing in the simulation. The
ringing behavior would be absent If the input waveform had been a more
gentle waveform with less bandwidth, such as a Gaussian pulse (Fig. 7) or a
sinusoid.
Numerical simulations are resource-intensive. However, we require nu-
merical simulations in most practical situations, such as where the solution
is not a simple wave equation, the transmission line is lossy, and the source
and load impedances are not resistive.
13
9 Steady-state sinusoid stimulation
Without any loss of generality, the forward and backward moving waves can
be phasors. In other words, V+ (x − ct) and V− (x + ct) may be V+ ej(ωt−kx)
and V− ej(ωt+kx) , respectively. Here k is the wavenumber and is given as
k = ω/c = 2πf /c = 2π/λ, where λ is the wavelength corresponding to the
frequency of excitation. The phasor corresponding to the forward-moving
wave is, then, V+ e−j2πx/λ . The backward-moving wave is V− ej2πx/λ . V+
and V− are complex numbers, in general. The current at any point is the
difference of the forward and backward moving voltage waves, divided by Z0 .
We are working entirely in the phasor domain.
The vital thing to note about analysis with phasors is the reflection co-
efficient. The reflection coefficient at the load is still given by:
ZL − Z0
ΓL =
ZL + Z0
ZL can now be any complex impedance. The same analysis as (17) holds.
The forward-moving wave at any point ahead of the load will be the
forward-moving wave at the load, only advanced in time. At any point
ahead of the load, the backward-moving wave will be delayed in time. For
example, at the source end, the forward-moving wave will be ej2πl/λ times the
forward-moving wave at the load. The backward-moving wave will be e−j2πl/λ
times the backward-moving wave at the load. The reflection coefficient at
the source is the ratio of the backward-moving to the forward-moving waves
at the source. This tells us that Γ(0) = ΓL e−j4πl/λ . Now we can compute
the reflection coefficient at any point in space over the entire length of the
transmission line.
14
0.6
normalized |V (x)|
0.58
0.56
0.54
0.52
0.5
0 0.5 1 1.5 2 2.5
distance in λ
We have given a small Octave routine showing all the steps to solve the
problem in Prog. 2.
We have shown a plot of the amplitude of the voltage along the trans-
mission line in Fig. 8. In a laboratory, one can experimentally estimate
the VSWR from the plot of the voltage amplitude along the length of
the transmission line. Notice that the amplitude is periodic, with a peri-
odicity of λ/2. There are two peaks and two valleys in every λ length of
the line. The amplitude along the transmission line is a standing wave
because of the (constructive and destructive) interference of the forward
and backward moving waves. The VSWR is 1.2.
15
d
Z0 ZL
Zin
16
30 mm Z0 = 500 Ω
Z0 = 5 Ω
input 50 Ω output
30 mm
Figure 10: A circuit with open circuit and short circuit stubs.
10 Quarter-wavelength lines
From the preceding discussion, it follows that transmission lines of length
λ/4 are unique. For such transmission lines, from (20), Γ(0) is given by −ΓL .
The input impedance, looking into the source end of the transmission line,
is given by:
V+ + V− 1 + Γ(0)
Zin = V /I = Z0 = Z0 (22)
V+ − V− 1 − Γ(0)
Given that Γ(0) is nothing but −ΓL , we obtain the input impedance as
1−ΓL
Z0 1+Γ L
. With the value of ΓL from (17), the input impedance is Z02 /ZL .
For an open-circuit load, the input impedance looks like a short-circuit.
For a short-circuit load, the input impedance looks like an open-circuit. For
a capacitive load, the input impedance looks like an inductor and vice-versa.
A λ/4 long transmission line inverts the load impedance, normalized to Z0 .
Note, this property is valid only at frequencies at which the length of the
transmission line is exactly an odd multiple of λ/4.
17
50 nH
output
input 20 pF 50 Ω
18
0
Transfer function, |H(jω)|
LC-Model
−20
−40
−60
Actual
−80
107 108 109 1010
Angular frequency, ω, rad/s
Figure 12: The system transfer function may be computed using the exact
impedances of the open and short-circuit stubs. We can also compute the
system transfer function by approximating tan(ωd/c) to be ωd/c for ω ≪ c/d.
|H(jω)| for the circuit in Example 5 has been shown, with and without the
approximate LC-model.
∆x
19
The Telegrapher’s equations: In the limit, as ∆x tends to 0, (25) and
(26) can be simplified.
dV(x)
= −(jωL′ + R′ )I(x) (27)
dx
dI(x)
= −(jωC ′ + G′ )V(x) (28)
dx
The equations (27) and (28) are the Telegrapher’s equations. They rep-
resent the phasor forms of (6) and (7).
The two equations (27) and (28) can be solved by taking a second deriva-
tive and decoupling them.
d2 V(x)
= (jωL′ + R′ )(jωC ′ + G′ )V(x) (29)
dx2
d2 I(x)
2
= (jωL′ + R′ )(jωC ′ + G′ )I(x) (30)
dx
The solution to both equations is as follows:
20
V(x) = V+ e−γx + V− eγx
V+ −γx V− γx
and, I(x) = e − e (33)
Z0 Z0
s
jωL′ + R′
where Z0 = (34)
jωC ′ + G′
V− eγL ZL − Z0
Γ(L) = −γL
= = ΓL (35)
V+ e ZL + Z0
V− eγx
Γ(x) = = ΓL e2γ(x−L) (36)
V+ e−γx
Transmission line analysis using phasors is used all the time in power
transmission and microwave circuits. This analysis is necessary when the
wavelength is comparable or smaller than the overall length of the cable.
The electrical length of a cable is the number of wavelengths in the length
of the cable. If the electrical length of the cable is anything more significant
than 0.1, one should model the distributed effects of the transmission line.
Transmission line effects come up in long power lines2 , in data cables such
as USB and SATA3 , on the back-plane of a computer motherboard, and in
microwave systems.
2
The free-space λ is roughly 6000 km at 50 Hz. In the transmission-line medium, λ
might reduce to 2000 km. 200 km long cables may have an electrical length ≥ 0.1λ.
3
In USB-3.1, the data rate is 10 Gbit/s. At the speed of light, each bit will use a length
of 3 cm (bit-length, as opposed to wavelength). 10 bits will be traversing a 0.3 m cable at
a time.
21
Example 6. Consider a 50 Hz power transmission system where the
source is at 1 MV and has a source resistance of 100 Ω. The transmission
line is 200 km long, with a copper resistance of 0.05 Ω/km, an insulation
loss of 100 nS/km, a series inductance of 4 mH/km, a capacitance of
10 nF/km. Let us assume a load resistance of 10 kΩ. Let us find the
voltage at the source end of the line, the voltage at the load-end of the
line, the power delivered to the load, the power input to the transmission
line. Let us also plot the magnitude of the voltage along the length of
the transmission line. We have solved the numerical in Prog. 3. We
have plotted the magnitude of the voltage along the transmission line in
Fig. 14.
22
1.06
|V (x)| in MV
1.04
1.02
M M
L L
12 Smith charts
The Smith chart is named after Philip Smith. The Smith chart is a graph-
ical calculator for solving problems that involve transmission lines. Philip
Smith was an engineer at the Bell Laboratories when he came up with the
Smith Chart in 1936 [3, 4, 5]. Later he started his own company, Analog
Instruments, which sold aircraft navigational instruments and radar-related
23
j1
j0.5 j2
j0.2 j5
j0.1 j10
0.2
0.1 0.5 1 2 5 10
−j0.1 −j10
−j0.2 −j5
−j0.5 −j2
−j1
(a) (b)
Figure 16: The graphs are maps from the impedance plane to the reflection-
coefficient plane. (a) The impedance is a constant resistance in series with a
variable reactance. (b) The impedance is a constant reactance in series with
a variable resistor. We have normalized all marked resistances and reactances
with respect to Z0 , the characteristic impedance of the transmission line. The
point marked as 1, in (a), corresponds to a resistance Z0 and is a reflection
coefficient of 0.
items. At its heart, the Smith chart is nothing but a plot of the complex
reflection coefficient, Γ, at different locations on the transmission line. Intri-
cate complex-number calculations reduce to operations that require a scale
and a compass on the Smith chart.
At the outset, we need to appreciate that the function Γ = (ZL (s) −
Z0 )/(ZL (s) + Z0 ) is a conformal mapping4 in the complex plane that takes
us from the Z-plane to the Γ-plane. We can move back and forth between
impedance and reflection coefficient.
The graphs in Fig. 16 are plots of the reflection coefficient, when (a) the
impedance is R0 ± jX, and (b) the impedance is R + jX0 . All the plots
are arcs of circles. These contours are the circles of constant resistance,
in the lower Γ-plane the circles of constant capacitance, and, in the
upper Γ-plane the circles of constant inductance respectively. The big
circle is for |Γ| = 1. The point at the center of the big circle marks Γ = 0
and ZL = Z0 . Regions outside |Γ| = 1 are irrelevant because the reflection
coefficient cannot be greater than 1 for passive loads.
4
A conformal mapping is a 1-to-1 mapping from the complex plane to the complex
plane. f : C → C.
24
0.12 0.13 0.14
0.11
0.1 0.15
0.38 0.37 0.36 0.1
0.0
9 0.39 6
0.4 90 80 0.35
100
8 0.3 0.1
0.0 1 70 7
0.4 110 4
1
0.9
1.2
0.3
0.8
2 0.1
7 0.4 120 + j X /Z0
60 3
0.0
1.4
8
0.7
E NT
PON
COM 0.3
1.6
3 CE
0.4 0
0.6
TA
N 2
50
1.8
AC
0.
06
13 RE
19
0.
2
IV A
0.5
0.
N
44
CT
31
DU
0.
G
L
IN
5
0.2
0.0
O
14
F
4
5
0.3
0.
0.4
R
E
F
3
LE
0.2
CT
→
1
0.2
6
150
0.4
0.3
IO
OR
9
N
AT
CO
0.22
ER
EF
0.28
0.47
EN
FI
5
S G
160
0.2
CI
ARD
EN
0.23
0.27
0.48
T OW
T
IN
TH S
10
170
D EGR
0.1
0.24
→ WA V E L E N G
0.26
0.49
EE S
20
50
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.2
1.4
1.6
1.8
10
20
50
0.25
0.25
180
5
0.5
0
50
RESISTIVE COMPONENT, R/Z0
20
←
0.24
0.26
L OAD
0.49
−170
−10
10
0.1
S
ARD
0.23
0.27
0.48
T OW
0
−20 2
−16
5
0.2
S
0.2
H
GT
0.28
0.47
4
EN L
VE
50
0
−3
.
−1
2
0.3
A
1
0
W
0.2
6
0.4
3
←
9
5
0.2
0.0
0
4
−4 0
0.
06 −14
5
0.3
0 .19
0.4
0
Z
X/
−j
0.
2
0.5
T
EN −5
1.8
0.
30
44
ON
31
−1 7 MP 0.1 0
0.
CO
1.6
0.6
0.0 CTA
NC
E 8
R EA
1.4
0.3
0.7
3 20 I VE
C A P AC I T 0.1 −60
0.4 −1 0.08 2
1.2
0.8
7
0.9
Fig. 17 is an example of a Smith chart that may be printed and used for
transmission-line problem-solving.
25
L
50 Ω, d 500 − j20 Ω
Let us also try to solve this with Octave. We will iterate over many
possible lengths of added transmission line, and select the length at which
the real part of the input impedance is just right.
26
0.12 0.13 0.14
0.11
0.1 0.15
0.38 0.37 0.36 0.1
0.0
9 0.39 6
0.4 90 80 0.35
100
8 0.3 0.1
0.0 1 70 7
0.4 110 4
1
0.9
1.2
0.3
7 2
0.4 120 0.8 + j X /Z0
60 3 0.1
0.0
1.4
8
0.7
E NT
PON
COM 0.3
1.6
3 E
0.4 0
0.6
A NC 2
CT 50
1.8
0.
06
A
13 RE
19
0.
2
IV A
0.5
0.
N
44
CT
31
U
0.
G
ND
L
I
5
0.2
0.0
O
14
F
4
5
0.3
0.
0.4
R
E
F
3
LE
0.2
CT
→
1
0.2
6
150
0.4
0.3
IO
R
9
TO
N
4
RA
CO
0.22
1 + jX
NE
EF
0.28
0.47
E
FI
5
S G
160
0.2
CI
ARD
EN
0.23
0.27
0.48
T OW
T
IN
TH S
10
170
D EGR
0.1
0.24
→ WA V E L E N G
0.26
0.49
EE
20
S
50
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.2
1.4
1.6
1.8
10
20
50
0.25
0.25
180
5
0.5
0
10 − j0.2
50
RESISTIVE COMPONENT, R/Z0
20
←
0.24
0.26
L OAD
0.49
−170
−10
10
0.1
S
ARD
0.23
0.27
0.48
T OW
0
−20 2
−16
5
0.2
S
0.04842λ
0.2
H
GT
0.28
0.47
4
1 − jX
N
LE E
50
0.2
−3
AV
−1
0.3
1
0
W
0.2
6
0.4
3
←
−j2.83
5
0.2
0.0
0
4
−4 0
0.
06 −14
5
0.3
0 .19
0.4
0
Z
X/
−j
0.
2
0.5
T
EN −5
1.8
0.
30
44
ON
31
−1 7 MP 0.1 0
0.
CO
1.6
0.6
0.0 CTA
NC
E 8
R EA
1.4
0.3
0.7
3 20 I VE
C A P AC I T 0.1 −60
0.4 −1 0.08 2
1.2
0.8
7
0.9
Figure 19: Example 7 worked out with the help of a Smith chart. One can
readily obtain approximate values graphically. If required, one may further
refine these approximate values.
27
Listing 4. Octave work in lieu of the Smith chart
octave:1> Z0 = 50; ZL = 500-j*20;
octave:2> GL = (ZL-Z0)/(ZL+Z0);
octave:3> l = 0:0.00001:0.5;
octave:4> Gin = GL*exp(-j*4*pi*l);
octave:5> Zin = (1+Gin)./(1-Gin);
octave:6> index = min(find((real(Zin)<1.0002) & (real(Zin)>0.9998)));
octave:7> length = l(index)
length = 0.048070
octave:8> real(Zin(index))
ans = 0.99988
octave:9> imag(Zin(index))
ans = -2.8487
octave:10> X = -imag(Zin(index)) * Z0
X = 142.43
13 Important concepts
• The transmission line is a distributed circuit with an infinite cascade
of R-L-C-G sections. Analysis of a single section leads us to a pair of
coupled partial differential equations in current and voltage over space
and time. These are called the Telegrapher’s equations.
28
• The Telegrapher’s equations can be differentiated once more to arrive
at the wave equation. The wave equation has a simple mathematical
solution as the superposition of a forward-moving and a backward-
moving wave. The wave function can be any arbitrary function, as
long as it satisfies
√ the boundary conditions. The wave moves at the
velocity of 1/ L C ′ , which also happens to be the velocity of light in
′
the medium.
• The forward and backward moving waves are related to each other
through
p the reflection coefficient, Γ. The characteristic impedance is
Z0 = L′ /C ′ .
• The two boundaries are at the source and the load. The boundary
conditions constrain Γ to be (ZL − Z0 )/(ZL + Z0 ) at the load. At
the source, only if we are interested in a time-domain analysis, the
reflection coefficient is (ZS − Z0 )/(ZS + Z0 ). If we are considering
sinusoidal steady-state analysis instead, the reflection coefficient is not
so.
29
12.5 nH
i(t) Z0 = 50 Ω
1 cos(109 t) 40 pF 200 Ω
λ/4
14 Exercises
1. A transmission line of characteristic impedance 200 Ω is of length such that it
takes a signal 1 ms to traverse from one end to the other. The transmission
line is terminated at the load-end in an open circuit. A voltage source,
v(t) = u(t), with a source impedance of 50 Ω is applied at the input end.
Draw a graph of the voltage at the load-end as a function of time, from
t = −1 ms to t = 10 ms.
4. In the circuit of Fig. 20, the transmission line is of length λ/4 at the given
operating frequency. Find the current i(t).
30
P1 Z0 = 50 Ω P2
λ/4
C ′ ∆x C ′ ∆x/2 C ′ ∆x/2
(a) (b)
7. Show that even if we start from any one of the transmission line sections
shown in Fig. 22(a), (b), we will reach the same Telegrapher’s equations, (6)
and (7).
10. Use a Smith chart or any other technique to match a (40 + j80) Ω load
to a 50 Ω source impedance for maximum power transfer. With the Smith
chart, first, plot the load impedance. Second, using a compass, draw a circle
31
50 Ω Z0 = 20 Ω Z0 = 120 Ω Z0 = 20 Ω
VS (jω) 50 Ω VL
2 mm 6 mm 2 mm
through the point with the center at the origin to cut the circle going through
the origin. The earlier step will give the length of the transmission line to
be added in series. Next, traverse to the center along the circle of constant
resistance. The last step will give the reactance to be added in series.
32
References
[1] M. Pupin, “Art of reducing attenuation of electrical waves and apparatus
therefore,” Jun. 1900, united States Patent No. 0 652 230.
[3] P. H. Smith, “Transmission line calculator,” Electronics, vol. 12, pp. 29–
31, Jan. 1939.
[4] ——, “An improved transmission line calculator,” Electronics, vol. 17,
pp. 130–133, Jan. 1944.
[5] ——, Electronic applications of the Smith Chart in waveguide, circuit and
component analysis. Raleigh, NC, USA: SciTech Publishing, 1969.
33