Download as pdf or txt
Download as pdf or txt
You are on page 1of 92

Frankfurt University of Applied Sciences

Faculty of Computer Science and Engineering

RF-Engineering

Prof. Dr.-Ing. G. Zimmer


Contents

1 Introduction 1
1.1 History of wireless communication . . . . . . . . . . . . . . . . . 1
1.2 Wireless Communication Systems . . . . . . . . . . . . . . . . . 2
1.2.1 Terrestrial Mobile Telecommunication . . . . . . . . . . . 2
1.2.2 Wireless Local Area Networks (WLANs) . . . . . . . . . 2
1.2.3 Satellite Communication . . . . . . . . . . . . . . . . . . 2
1.3 Circuit Technoloy Options for Wireless RF/Microwave Circuit
Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3.1 Semiconductor Devices . . . . . . . . . . . . . . . . . . 3

2 Fundamentals of Transmission Line Theory 5


2.1 Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Discussion on the Propagation Constant and Charateristic Wave
Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.1 Lossless Line . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2.2 Lossy Transmission Line . . . . . . . . . . . . . . . . . . 10
2.2.3 The Reflection Coefficient in the Frequency Domain . . . 13
2.2.4 The Smith Chart . . . . . . . . . . . . . . . . . . . . . . 16
2.2.5 Mathematical Impedance Transformation along a Trans-
mission Line . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3 Line Transformer . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.3.1 Quarter-Wave Transformer . . . . . . . . . . . . . . . . . 26

3 Scattering Parameters 29
3.1 Definition of Wave Amplitudes . . . . . . . . . . . . . . . . . . . 30
3.2 Definition of S-Parameters . . . . . . . . . . . . . . . . . . . . . 31
3.2.1 2-Port . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2.2 N-Port . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2.3 Useful Definitions . . . . . . . . . . . . . . . . . . . . . 32
3.3 Lossless Networks . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4 Signal Flow Graph . . . . . . . . . . . . . . . . . . . . . . . . . 35

I
3.4.1 Definition of Graphical Elements . . . . . . . . . . . . . 35
3.4.2 Special Devices and their graphical Representation . . . . 37
3.5 Network Analyzer . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4 Microwave Amplifier 46
4.1 Y-Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.1.1 2-Port described by y-parameters . . . . . . . . . . . . . 48
4.2 Stability of a 2-Port . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.1 Stability Factor . . . . . . . . . . . . . . . . . . . . . . . 53
4.2.2 Stabilization of an Active Device . . . . . . . . . . . . . 54
4.3 Power Gain Definitions . . . . . . . . . . . . . . . . . . . . . . . 55
4.3.1 Unilateral Transistor . . . . . . . . . . . . . . . . . . . . 57
4.3.2 Bilateral Transistor . . . . . . . . . . . . . . . . . . . . . 58
4.3.3 Optimum Input and Output Matching . . . . . . . . . . . 59
4.3.4 Maximum Available Power Gain . . . . . . . . . . . . . . 60
4.4 Differnt Amplifier Concepts . . . . . . . . . . . . . . . . . . . . 61

5 Plane Wave 65
5.1 Wave Equation of the Plane Wave . . . . . . . . . . . . . . . . . 65
5.2 Weakly damped Plane Wave . . . . . . . . . . . . . . . . . . . . 67
5.2.1 Plane Wave in free Space . . . . . . . . . . . . . . . . . . 68
5.2.2 Plane Wave in a Lossless Material . . . . . . . . . . . . . 69
5.2.3 Polarization . . . . . . . . . . . . . . . . . . . . . . . . . 71
5.2.4 The Wave Vector . . . . . . . . . . . . . . . . . . . . . . 73
5.3 Plane Wave in a good Conductor . . . . . . . . . . . . . . . . . . 75
5.3.1 Ideal Conductor . . . . . . . . . . . . . . . . . . . . . . . 77

6 Antennas 78
6.1 Transmitter Receiver System . . . . . . . . . . . . . . . . . . . . 78
6.1.1 Transmitter-Receiver-System as Quadrupole . . . . . . . 81
6.2 The Hertzian Dipole . . . . . . . . . . . . . . . . . . . . . . . . 83
6.2.1 The Far Field . . . . . . . . . . . . . . . . . . . . . . . . 86

II
Chapter 1

Introduction

The rapid development of wireless communication products over the last ten years
has led to an explosion of interest in improved circuit design approaches in the ra-
dio frequency (RF) and the microwave area. The development of the field is char-
acterized by a shift from purely analog techniques to those employing digital or
hybrid digital/analog approaches. This shift is further accelerated by the advance
in digital integrated circuit technology and monolithic microwave integrated cir-
cuits.

1.1 History of wireless communication


In the following section a short history of wireless communictaion is given.

1864 Theoretical prediction of electromagnetic waves by James Clerk Maxwell.

1887 Experimental verification of electromagnetic waves by Heinrich Hertz.

1897 First practical transmission experiment by Guglielmo Marconi.

1901 First transmission experiments from a truck.

1926 First telephone on a train in Germany.

1958 Installation of the A-Netz by the German Bundespost

1972 Installation of the B-Netz by the German Bundespost

1986 Installation of the C-Netz by the German Bundespost

1991 Introduction of the GSM-Standard (Global System for mobile Communica-


tion). Deregulation of the telecommunication industry

1
1.2 Wireless Communication Systems
1.2.1 Terrestrial Mobile Telecommunication
GSM has become a worldwide used standard for terrestrial mobile telephony. In
the following table some technical data of the GSM-Standard is given in compar-
ison to IS-54 an U.S. Interim Standard.

IS-54 GSM
Accces methods TDMA/FDMA TDMA/FDMA
Frequency bands 824 - 848 890 - 915
in MHz 869 - 894 935 - 960
Channel spacing 30kHz 200kHz
π
Modulation 4 QPSK GMSK
Transmitted Power 600mW 1W
Bit Rate 8kbps 13kbps
Time Slots per
Channel 3 8
Table 1.1: Comparison of different mobile systems

1.2.2 Wireless Local Area Networks (WLANs)


The most active frequency bands used today for the development of WLANs are
the so-called ISM-Bands ( Industrial Scientific Medical-Bands). The following
table shows some technical features of the 915 MHz, 2.4GHz and the 5.8GHz
bands.

1.2.3 Satellite Communication


One must distinguish between two different forms of satellite communication.
The first is the well known traditional satellite communication which is used for:

• Point to point voice video and data transmission using geosynchronous


satellites
• Point to multipoint video transmission with fixed earth transmitting and re-
ceiving stations

2
915MHz 2,4GHz 5,8GHz
Frequencies 902 - 928 2,4 - 2,4835 5,725 - 5,825
Bandwidth 26MHz 83,5MHz 150MHz
Availability US/Canada Worldwide US/Canada
Transmission Range Greatest 95% of 915MHz 80% of 915MHz
Technologies Si Si/GaAs GaAS
Interference Celluar Microwave Some Radar
Sources Phones Oven Application
Table 1.2: WLAN in different ISM-Bands

the second is the not yet established mobile satellite communication described as:

• Geosynchronous mobile satellite service (e.g. Inmarsat) for voice, data and
video services operating with frequencies from one to six GHz (Geosyn-
chronous Earth Orbit, GEO)

• Low Earth Orbit (LEO) mobile satellite service (Iridium system of Mo-
torola) with about 66 satelites in orbit

• Medium Earth Orbit (MEO) satellite systems (Odyssey system proposed by


TRW) with about 12 satellites in orbit

1.3 Circuit Technoloy Options for Wireless RF/Microwave


Circuit Implementation
1.3.1 Semiconductor Devices
The choice of the optimum semiconductor technology for wirelss applications
is a hotly contested issue. In the last few years wireless components have been
announced or proposed using the following technologies:

• Silicon bipolar

• Complementary metal oxide semiconductor (CMOS)

• Si/SiGe heterojunction bipolar transistors (HBTs)

• GaAs MESFETs and HBTs

The key transistor technoloy figures of merit for RF/Microwave systems are:

3
• Minimum necessary supply voltage

• Short circuit current bandwidth ( fT ), important for both analog and digital
circuits

• Maximum power gain bandwidth ( fmax ), important for both analog and dig-
ital circuits

• Minimum noise figure, LNAs, wide dynamic range front ends and mixers

• Maximum power-added efficiency, power amplifier

• Linearity, AGC amplifier, front ends

• Yield

The following table shows a comparison between the different technologies.

Si-BJT Si/SiGe HBT GaAs-MesFet GaAs HBT BiCMOS


Supply Voltage/V 4 4 8 15 6
fT / GHz 32 55 50 30 13
fmax /GHz 35 55 60 70 11
Gmax /dB (2GHz) 24 28 20 19 17
NF /dB /2GHz) 1 0,5 0,3 1,5 1,0
IP3/P(-1dB)/dB 9 9 12 16 9
Power added ef.(%) - 70 60 70 40
Table 1.3: Figures of merit for different technologies

4
Chapter 2

Fundamentals of Transmission Line


Theory

Figure 2.1: Electric and magnetic field lines of a two wire line

Electrical transmission lines exist within all forms of electro-technical equip-


ment. A majority of the everyday amount of energy used is provided by electrical
transmission lines. The connection of all telephone subscribers to the end office
is made by transmission lines. With the help of transmission lines TV and radio
programs are distributed, antennas are fed and components are interconnected.
In this chapter, we want to use a circuit description to deduce the voltage/current
waves that may exist on a transmission line. To do this, Fig. 2.1 shows the electric
and magnetic field lines of a two wire line. If we consider a line section with the
length ∆z, the effect of the electric field can be represented by the capacitance per
unit length C′ and the effect of the magnetic field can be represented by the induc-
tance per unit length L′ . As a result we end up in the equivalent circuit, shown in
Fig. 2.2, of a loss less line section of length ∆z.

5
i(z,t) L′ ∆z i(z + ∆z,t)
- s 

 s s -

u(z,t) C′ ∆z u(z + ∆z,t)


? ?
s s s

z z + ∆z

Figure 2.2: Equivalent circuit for a loss less line section of length ∆z

2.1 Analysis
Fig. 2.2 refers to the time domain, this means that the function u(z,t) can show
different values for any point z or time t. Considering that all occurring time func-
tions are related to sinusoidal time functions by the Fourier analysis or Fourier
transform, it is sufficient to restrict ourself to sinusoidal time functions to de-
velop the transmission line theory. If the function u(z,t) has a sinusoidal time-
dependency, we can use the following representation:
[ ]
u(z,t) = Û(z) sin(ωt + φ(z)) = Im Û(z) exp( jφ(z)) exp( jωt) =

Im [U(z) exp( jωt)] (2.1)


In the last equation, Û(z) represents the amplitude of the function u(z,t) at certain
point z, in addition, the phasor function U(z) = Û(z) exp( jφ) is introduced. Its
absolute value is equal to the amplitude function Û(z), while its phase is equal to
the phase of the time function sin(ωt + φ). Similar considerations also apply to
the space and time function i(z,t), which is also represented by a complex phasor
I(z), being a function of the z-coordinate.

The Wave Equation in the Frequency Domain


In order to deduce the wave equation, we will use the equivalent circuit shown
in figure 2.3. In contrast to Fig. 2.2, two additional elements are introduced, de-
scribing the losses of a real transmission line. On the one hand, we have the series
resistance per unit length R′ , which accounts for all losses caused by the current
flowing along the line and, on the other hand, we have the parallel conductance
per unit length G′ , which accounts for the losses due to the voltage between the
two conductors. The derivation of the transmission line equations for sinusoidal
time functions is carried out almost like in the time domain, with the only differ-
ence that the derivation with respect to time is replaced by a multiplication with

6
I(z) R′ ∆z L′ ∆z I(z + ∆z)
- s 

 s s s -

U(z) C′ ∆z G′ ∆z U(z + ∆z)


? ?
s s s s

z z + ∆z

Figure 2.3: Equivalent circuit of a lossy line section of length ∆z

jω. Considering the resistance R′ ∆z in series with the inductance L′ ∆z, we will
get in the limit ∆z → 0:

dU(z)
− = (R′ + jωL′ )I(z) (2.2)
dz
An equivalent equation can be setup for the local decrease of current:

dI(z)
− = (G′ + jωC′ )U(z) (2.3)
dz
By derivation of the equation 2.2 with respect to z and taking into account equa-
tion 2.3 we obtain a homogeneous linear differential equation of second order for
phasor of the complex voltage U(z).

d 2U(z)
2
= (R′ + jωL′ )(G′ + jωC′ )U(z) (2.4)
dz
It is well known that the exponential function is a possible solution for the above
differential equation. So we will use the following function as trial function:

U(z) = U o exp(−γ z) (2.5)

In this equation, U o is a constant complex amplitude which must be adapted to the


physical reality, while γ describes the local dependence of the wave. It is denoted
as the complex propagation constant of the wave. If the trail function is included
in the differential equation 2.4, we end up with following equation:
[ ]
γ2 − (R′ + jωL′ )(G′ + jωC′ ) U(z) = 0

In principle two solution exist. The equation can be fulfilled if there exists no
voltage across the transmission line (U(z) ≡ 0) - a trivial solution of little interest

7
for an electrical engineer. Instead the non-trivial solutions defines the value of the
complex constant of propagation γ.

γ = ± (R′ + jωL′ )(G′ + jωC′ ) (2.6)

Equation 2.6 is called the characteristic equation of the differential equation 2.4.
As in the case of time domain, the wave equation presents two different solutions
in the frequency domain. The physical significance of the solution due to the two
different signs will be treated in the next chapter. To calculate the complex current
distribution, I(z) along the transmission line, we start with equation 2.2:

1 dU(z) γ
I(z) = − = U(z)
R′ + jωL′ dz R′ + jωL′
Considering the result of equation 2.6, we obtain:

G′ + jωC′ 1
I(z) = ′ ′ U(z) = U(z)
R + jωL ZC

The complex constant ZC , introduced in the previous equation, shows the unit of
an impedance and is called complex characteristic wave impedance of the trans-
mission line. √
R′ + jωL′
ZC = (2.7)
G′ + jωC′

2.2 Discussion on the Propagation Constant and Chara-


teristic Wave Impedance
The propagation constant γ possesses a complex value, which can be separated in
its real and imaginary part.
γ = α + jβ (2.8)
It will be shown that α describes the damping of the wave while β describes its
phase shift. α is called damping constant and β is called phase constant.

2.2.1 Lossless Line


To discuss the physical significance of α and β, we will start by examining the
case of a lossless transmission line R ≡ 0, G ≡ 0. According to the equation 2.6
we obtain for β and α: √
β = ω L′C′ α≡0 (2.9)

8
By using equation 2.5 and equation 2.1 to convert the complex voltage distribution
U(z) to the voltage function depending on z and t, we obtain:

u(z,t) = Im[Û0 exp(− jβz) exp( jωt)] = Û0 sin(ωt − βz) (2.10)

The complex constant U 0 was considered to be real. This may be achieved by


choosing an appropriate time scale. If the result of equation 2.10 is compared
to Fig. 2.4, you will see that the function exp(− jβz) describes a wave in the
frequency domain which propagates in the positive z-direction. In accordance to
this result, the function exp(+ jβz) represents a wave propagating in the negative
z-direction. In Fig. 2.4, the voltage u(z,t) according to equation 2.10 is shown

Figure 2.4: Voltage distribution along a loss less transmission line for t = 0 and
t = ∆t > 0

for t = 0 and a t = ∆t > 0. A sinusoidal voltage distribution can be observed


along the entire transmission line which propagates in the positive z-direction. To
determine the velocity of a point with constant phase,we consider the point with
the total phase φ = 0. According to the equation 2.10 this results in:
ω
φ = ωt − βz → für φ = 0 → z = t
β

Apparently, the factor ω/β shows the unit of a velocity which is denoted as phase
velocity v ph of the transmission line wave.

ω 1
vPh = = √ (2.11)
β L′C′

9
The phase velocity v ph of the loss less transmission line as defined in equation
2.11 is equal to the propagation velocity v of pulses along a loss less transmission
line. Moreover its value is identical to the propagation velocity of TEM waves
according to Maxwells theory. We therefore need:
1 c0
vPh = √ = √ (2.12)
L′C′ εr
The last equation shows that there exists a relation between the inductance and
capacitance per unit length of a transmission line according to Maxwells theory.
Another important characteristic value of transmission line theory is the wave-
length λ of the wave on a transmission line. The wavelength corresponds to a
length ∆z where the wave will show a phase shift of 2π. From this definition, we
obtain:
2π vPh
λ = = (2.13)
β f
which results in the following equation using the results of Maxwell:
vPh c0
λ = = √ (2.14)
f f εr
If we focus on the equation 2.7 in the lossless case, we obtain a real value for the
characteristic wave impedance.

L′
ZC = ZC = (2.15)
C′
In the lossless case, the characteristic wave impedance is real and independent of
the frequency.

2.2.2 Lossy Transmission Line


In the following we will not consider the unrealistic hypothesis of a lossless trans-
mission (R′ ̸= 0, G′ ̸= 0). However, we want to anticipate that the losses are at a
low level and that the following prerequisites are fulfilled:
R′ ≪ ωL′
G′ ≪ ωC′
These conditions certify that the fields still show essentially the properties of TEM
waves. If we consider the propagation constant according to equation 2.6, this
equation can be rewritten as follows:

√ R′ G′
′ ′
γ = jω L C (1 − j ′ )(1 − j ′ )
ωL ωC

10
Considering the above conditions, both brackets within the last root can be multi-
plied and as a first-order approximation the product of the small-sized terms can
be neglected. √
√ R′ G′
γ ≈ jω L′C′ 1 − j ′ − j ′
ωL ωC

The following approximation 1 + ε ≈ 1 + ε/2 is used for further simplification:
[ ]
√ R′ G′
′ ′
γ ≈ jω L C 1 − j( + )
2ωL′ 2 jωC′

The real-valued constants β and α can be determined by multiplying and by sep-


arating the last equation into real and imaginary part.
√ 2π
β = ω L′C′ = (2.16)
λ
R′ 1
α = + G′ ZC (2.17)
2ZC 2
A comparison of equation 2.9 with the equations 2.16 and 2.17 shows that the
phase propagation constant β shows the same value in the case of low losses, while
the constant α is greater than zero in the lossy case. To be able to recognize the
physical meaning of these constants, we again consider the voltage at an arbitrary
point in space and time u(z,t):

u(z,t) = Im[Ûo exp(−(α + jβ)z) exp( jωt)] = Ûo exp(−αz) sin(ωt − βz) (2.18)

By comparing the last equation with the result of the loss less circuit (equation
2.10), it is obvious that the losses are reflected in the term exp(−αz). The value of
the exponential function becomes lower with increasing z, so that the amplitude
of the wave and as a result the transported power will decrease. Therefore, the
constant α is denoted as damping constant of the transmission line. To illustrate
again the physical process of a damped wave according to equation 2.18 Fig. 2.5
shows the space and time function u(z,t) at time t = 0 and at time t = ∆t > 0.
It becomes evident that the amplitude continuously decreases while the wave is
propagating in the z-direction. Hence a part of the power transported by the wave
is transformed into heat.

Example 2.1: Calculation of the damping constants α


A CATV cable manufacturer (Community Antenna Television) defines the damp-
ing of its cable to be 3 dB/100 m at 200 MHz. Find the value of the damping

11
Figure 2.5: Voltage distribution along a lossy transmission line t = 0 and t = ∆t >
0

constant α at 200 MHz.

|U(0)| |U(0)|
|in dB = 3 → = 103/20 = exp(α · 100 m)
|U(100m)| |U(100 m)|
1 1
α = ln(103/20 ) ≈ 3, 5 · 10−3
100 m m

To focus on the characteristic wave impedance in the lossy case, we refer to the
equation 2.7. Provided that the losses are low, we use the following approxima-
tion: v
u
u R′ R′
u1− j ′ 1 − j
u ωL ≈ Z 2ωL′
ZC = ZC u
t G′ G′
C
1− j ′ 1− j
ωC 2ωC′
If we consider the last equation, we see that the numerator of the fraction will
result in a negative phase change, while the denominator is rotating in the positive
direction. Both rotations will esentially compensate each other while each rotation
itself will decrease with increasing frequency. A more detailed analysis shows [1]
that for frequencies above 10 kHz, the complex characteristic impedance ZC will
become real value ZC .
ZC ≈ ZC (2.19)

12
Therefore, we will assume that for RF- and microwave frequencies the character-
istic wave impedance of a transmission line will always show a real value.

2.2.3 The Reflection Coefficient in the Frequency Domain


Remembering the results of the lossless transmissin line in the time domain, re-
flections will occur when the load impedance terminating the transmission line
is not equal to its characteristic impedance. To calculate the voltage distribution
U(z) along the line, both waves, the incident and reflected wave, must be consid-
ered:
U(z) = U i exp(−γz) + U r exp(γz) (2.20)
The constants U i and U r introduced into equation 2.20 define the complex am-
plitude of forward- and backward-traveling waves at z = 0. Therefore, the input
reflection coefficient rin of a transmission line is defined by:
Ur
rin = (2.21)
Ui

In case of a reflection taking place the same approach must be used for the current
distribution I(z) along the transmission line:

I(z) = I i exp(−γz) + I r exp(γz) (2.22)


Remembering that the relation between voltage and current amplitudes U i , I i , U r ,
and I r is given by the characteristic impedance ZC = U i /I i = −U r /I r , the equation
for the current distribution I(z) along the transmission line can be transcribed as
follows:
1 [ ]
I(z) = U exp(−γz) − U r exp(γz) (2.23)
ZL i
Fig. 2.6 shows the end of the transmission line, which is terminated into a com-

I(z) IL
- -
s s

U(z) ZC , γ UL ZL
? ?
s s
-
0 l z

Figure 2.6: Transmission line with arbitrary load impedance Z L

13
plex load impedance Z L determining the ratio between voltage U L and current I L .
Equation 2.20 can be used to calculate the voltage U L :

U L = U i exp(−γl) + U r exp(γl) (2.24)

The current I L is obtained by applying 2.23:


1 [ ]
IL = U i exp(−γl) − U r exp(γl) (2.25)
ZC
If we form the ratio of the preceding equations, we obtain:
U i exp(−γl) + U r exp(γl)
1 [ ]
ZL =
U i exp(−γl) − U r exp(γl)
ZC
In the last equation, only the complex amplitudes of the forward- and backward-
traveling waves U i and U r are unknown. For this reason, this term can be used to
determine the input reflection factor rin :
Ur Z − ZC
rin = = exp(−2γl) L (2.26)
Ui Z L + ZC
Considering the last term of equation 2.26, it appears obvious that this term defines
the load reflection coefficient rL due to the arbitrary load impedance Z L :
Z L − ZC
rL = (2.27)
Z L + ZC
On the other hand, the first term of equation 2.26 describes the transformation of
the load reflection coefficient rL into the input reflection coefficient rin :

rin = exp(−2γl) rL = |rL | exp(−2αl) exp[ j(φrL − 2βl)] (2.28)

Equation 2.28 demonstrates again the interrelation between the output reflection
factor rA and the input reflection factor rin of a transmission line; two effects
are obvious. One of these effects is that the load reflection coefficient will be
rotated in a mathematical negative sense (clockwise direction) by 2βl through the
transmission line; the second effect, damping of the transmission line will reduce
the absolute value of the reflection coefficient. In Fig. 2.7 the impact caused by
the transformation line on the reflection coefficient along the line is depicted. To
explain the impact of the output reflection factor rL on the voltage distribution
U(z) along the transmission line U(z), equation 2.20 is rearranged as follows:
[ ]
Ur
U(z) = U i exp(−γz) 1 + exp(2γz)
Ui

14
r_L
r_in

Figure 2.7: Transformation of the reflection coefficient rL along a transmission


line

The ratio U r /U i can be replaced by the load reflection coefficient:


[ ]
U(z) = U i exp(−γz) 1 + rL exp(−2γl) exp(2γz)

This leads to an equation describing the voltage distribution along the transmission
line: [ ]
U(z) = U i exp(−γz) 1 + rL exp(−2γ(l − z)) (2.29)

The last term becomes particularly concise if we consider a lossless line (γ = jβ)
and focus on the absolute value of the voltage |U(z)|.

|U(z)| = |U i | | exp(− jβz)| | [1 + rL exp(− j2β(l − z))] | (2.30)

Since the term | exp(− jβz)| is equal to one, the bracket term is particularly rele-
vant. Due to the exponential function and the space coordinate z it will change its
value between 1 − |rL | and 1 + |rL |. This is also illustrated by Fig. 2.8. Here the
voltage distribution U(z) is shown normalized to the incident wave amplitude U i
for different real values of rL . The diagram shows that the voltage U(z) shows a

15
r_L

Figure 2.8: Voltage distribution along a transmission line for different reflection
coefficients rL

constant amplitude equal to U i for rL = 0 along the transmission line. If reflec-


tions occur, the amplitudes will oscillate between the values Ûmin = |U i |(1 − |rL |)
and Ûmax = |U i |(1 + |rL |). The ratio Ûmax /Ûmin is called Voltage Standing Wave
Ratio (vswr). In the case |rL | = 1 the maximum voltage along the transmission
line can become two times the value of the incident wave amplitude. In this case
a standing wave exists along the transmission line. If it is possible to measure
the values of the position-dependent voltage with the help of a slotline [6], the
absolute value of reflection coefficient can be determined from the acquired vswr.
vswr − 1
|rL | = (2.31)
vswr + 1

2.2.4 The Smith Chart


With the help of equation 2.27, the reflection coefficient can be calculated on the
basis of the given load resistance Z L and characteristic impedance ZC of the trans-
mission line. If we define z = Z/ZC to be an arbitrary impedance Z normalized to
the characteristic impedance ZC of the transmission line, the reflection coefficient
is calculated by:
z−1
r = (2.32)
z+1
In a mathematical sense, the equation above is a conformal mapping of the z-plane
onto the r-plane. It preserves angles locally and circles will be mapped onto cir-
cles again. In the frame of this theory a straight line is nothing else but a circle

16
with an infinite radius. If two curves intersect in the z-plane under a certain angle,
the local preservation of angles guarantees that the maped curves will intersect in
the r-plane under the same angle.

To visualize the properties of the term 2.32, we consider at first some important
points of the z-plane:

matching point • z = 1 r = 0
short circuit ⊗ z = 0 r = −1
open circuit ◦ z→∞ r=1

Figure 2.9: Conformal mapping of the z-plane onto the r-plane

These points are depicted in 2.9 to show their mapping from the z- to the r-plane.
In addition to the analysis of these discrete points, it is helpful to have a look on
the transformation of entire coordinate lines.

The positive real axis Im(z) = 0

0 ≤ Re(z) ≤ ∞ → −1 ≤ Re(r) ≤ +1 und Im(r) = 0

The imaginary axis Re(z) = 0

−∞ ≤ Im(z) ≤ +∞ → unit circle

In 2.9, two further coordinate lines are represented. We can see that the entire
half-plane Re(z) ≥ 0 on the right side of the imaginary axis is displayed within
the unit circle of the r plane.

17
Impedance Transformation with the Help of the Smith Chart
Terminating a transmission line with an impedance Z L ,which is not equal to its
characteristic impedance, the voltage U(z) will vary along the transmission line
(see figure 2.8). Similarly, the function I(z) will also change along the transmis-
sion line and hence the input impedance Z in will become a space function depen-
ing on the length of the transmission line. In the following, we will examine how
to apply the Smith Chart in order to determine the input impedance of a lossless
transmission line. In this context, we rewrite equation 2.29 for the lossless case.
l
rin = exp(− j2βl)rL = exp(− j4π ) rL (2.33)
λ
As shown by equation 2.33, a transformation of the load reflection coefficient rL
along the transmission line into the input reflection coefficient rin may be obtained
by just conducting a simple rotation (clockwise) round the angle ∆φ = −4πl/λ.
The only complication may be the calculation of the reflection coefficient rL and
the reconversion of the input reflection coefficient rin , into the associated input
impedance Z in . As pointed out in the previous chapter, the Smith Chart allows a
graphic calculation of the reflection coefficient. Therefore, an impedance trans-
formation can easily be performed with the help of a Smith Chart. To do this, the
following steps should be carried out:
1. We normalize the load impedance Z L to the characteristic impedance ZC
of the transmission line and then draw the normalized impedance into the
Smith Chart. After entering this information, we are able to read out the
load reflection coefficient rL .
2. Transformation of the reflection coefficient along a circle with a constant
reflection radius in clockwise direction using the angle ∆φ = 4πl/λ.
3. Inverse transformation by converting the input reflection coefficient into the
normalized impedance zin and de-scaling to the impedance Z in by multipli-
cation with the value of the characteristic impedance of the transmission
line.
Example: Transformation of an impedance with the help of the Smith Chart
As example, we want to consider an air filled transmission line with the charac-
teristic impedance ZC = 50 Ω terminated in an impedance Z L = 10 Ω. The length
is assumed to be l =0,75 m and the transmission line is operated at a frequency of
f =75 MHz.
1. For the normalized load impedance zL , we obtain:
10 Ω
zL = = 0, 2
50 Ω

18
If we insert this value into a Smith Chart, as shown in Fig. 2.10, it will be
possible to determine the output reflection coefficient to be rA ≈ −0, 66.

2. To calculate the angle ∆φ of transformation, we must determine the wave-


length:
c0 l
λ = λ0 = =4m → = 0, 1875
f λ
Hence we get for the angle of transformation: ∆φ ≈ 2, 356=135 ˆ 0 . The

input reflection coefficient rin can be determined by rotating (clockwise) of


the load reflection coefficient by 1350 .

r_in

r_L

Figure 2.10: Transformation of an impedance with the help of the Smith Chart

3. After transformation, the input reflection coefficient rin ≈ 0, 66 exp(+ j450 )


can be read out from the Smith Chart and also the normalized associated
input impedance zin = 1, 1 + j1, 85. Hence we get for the input impedance
Z in :
Z in = 50 Ω(1, 1 + j1, 85) = (55 + j92, 5) Ω

Dealing with Admittances


The Smith Chart in the present form is chart for impedances. For operations based
on admittances, another kind of diagram would be necessary. To avoid this, we

19
will give an advice how to the use the Smith Chart for impedances to work also
with admittances. For this purpose, we now look again at equation 2.32 normally
used for the calculation of coordinate lines of the Smith-Chart. This equation can
be expressed as follows:
z−1 1−y y−1
r = = = (−1)
z+1 1+y y+1
If we examine the last term of the above stated equation, we have the same form
as equation 2.32 except for the factor −1. It is evident that the Smith Chart for

Figure 2.11: Reflection coefficient of an admittance

impedances can also be used for admittances. However, if we want to determine


the reflection coefficient ry corresponding to the admittance y, the phasor r′ in-
dicating the position of the normalized admittance must be rotated by 1800 or a
mirrored at the point ”1” of the Smith Chart. If one reads out the numerical value
one may notice that it is equal to the normalized impedance z = 1/y. Mirroring at
point ”1” in the Smith Chart correspond to the transition from the admittance plane
to the impedance plane. To demonstrate our approach, we consider equation 2.11.
Included was the admittance y = 1 + j2 displayed by r′ . The reflection coefficient
ry , associated to this admittance, will be obtained by mirroring of r′ at point ”1”.
The reflection coefficient ry ≈ 0.7 exp(− j1350 ) is shown in the diagram. Further-
more, we can read the normalised impedance value z = 1/y ≈ 0, 2 − j0, 4 which
corresponds to the admittance y.

20
2.2.5 Mathematical Impedance Transformation along a Trans-
mission Line
In order to obtain a mathematical form for the input impedance of a transmission
line, it is recommended to deduce the ratio of voltage U in and current I in as a
function of the ratio of voltage U L and I L at the load of the transmission line. For
the definition of the positive defined directions see Fig. 2.12. For this purpose we

I in IL
-s s -

U in ZC , γ UL
? ?
s s
- z
0 l

Figure 2.12: Definition of the direction of voltage and current at the transmission
line

start from equations 2.24 and 2.25 and calculate the sum and difference in order
to determine the complex amplitudes of forward- and backward-traveling waves.

U i = 12 (U L + I L ZC ) exp(+γl)
(2.34)
U r = 12 (U L − I L ZC ) exp(−γl)

The last two equations can be used in equations 2.20 and 2.23. In this way the
voltage and current distribution along the transmission line can be determined
according to the value of voltage and current at the ouput of the line. For the
voltage distribution, we obtain:
1 1
U(z) = (U L + I L ZC ) exp(+γl) exp(−γz) + (U L − I L ZC ) exp(−γl) exp(γz)
2 2
If we order the terms of the voltage U L and current I L and take into account the
definitions cosh(x) = [exp(x) + exp(−x)]/2 and sinh(x) = [exp(x) − exp(−x)]/2,
we obtain for the voltage distribution U(z) along the transmission line:

U(z) = U L cosh[γ(l − z)] + I L ZC sinh[γ(l − z)] (2.35)

Similar considerations lead to the equation for the current distribution I(z):
UL
I(z) = sinh[γ(l − z)] + I L cosh[γ(l − z)] (2.36)
ZC

21
If we include the parameter into equations 2.35 and 2.36, we obtain, as wished,
the input voltage U in and current I in as a function of the output voltage U L and
current I L of the transmission line:
U in = U L cosh(γl) + I L ZC sinh(γl)
U (2.37)
I in = Z L sinh(γl) + I L cosh(γl)
C

In case the line shows no losses, (γ → jβ) the above equations change to:

U in = U L cos(βl) + jI L ZC sin(βl)
U (2.38)
I in = j Z L sin(βl) + I L cos(βl)
C
With the help of equations 2.37 we are able to determine the input resistance Z in
of a transmission line in case it is terminated with the impedance Z A . Fig. 2.13
illustrates the circuit setup. Shown is a transmission line of the length l, which

Z in
s s

- ZC , γ ZL
s s
- z
0 l

Figure 2.13: Transmission line terminated with the resistance Z L

is described by its characteristic impedance ZC and its propagation constant γ,


which is terminated in the impedance Z A . To determine the input impedance, we
calculate the ratio between input voltage U in and input current I in .
U in U L cosh(γl) + I L ZC sinh(γl)
Z in = =
I in UL
sinh(γl) + I L cosh(γl)
ZC
If we divide the above term by I L cosh(γl), taking into account that the output
impedance Z L is given by the ratio of the output voltage U L and the output current
I L , we obtain for the input resistance Z in of the circuit:
ZL
+ tanh(γl)
Z
Z in = ZC C (2.39)
Z
1 + L tanh(γl)
ZC

22
Figure 2.14: Input impedance of a transmission line in a complex Z plane

In the loss less case the equation changes to:


ZL
+ j tan(βl)
ZC
Z in = ZC (2.40)
Z
1 + j L tan(βl
ZC
Example: Input impedance of an air filled (εr = 1) transmission line.
Transmission line: ZC = 60 Ω, l =1 m and Z L = 20 Ω

Figure 2.14 shows the input impedance of the above transmission line in the
frequency range from 1 MHz to 130 MHz in a complex impedance plane and how
it may be determined by means of a simulation program. In this chart, the mark
”>” specifies, in each case, the position of the reference values in x direction (real
part) and in y direction (imaginary part). Both show a value of 0 Ω in this chart.
The shown part of the plane Z in is subdivided in squares, the boundary lines show
a distance of 20 Ω.
As we can see from the diagram, the calculated value for low frequencies f ≈1 MHz
is almost equal the load resistance. With increasing frequency it becomes induc-

23
tive and and shows a real maximum of 180 Ω at a frequency of f = 75 MHz.
The marker does not indicate exactly this value because due to the setsize of the
frequency the value for 75 MHz was not calculated. When calculating the wave-
length λ for 75 MHz, the result is λ =4 m. The length l of the line is at this
frequency equal to λ/4. If the frequency exceeds 75 MHz, the input resistance
becomes capacitive. Obviously, the plot of the input impedance shows the form
of a circle in the impedance plane. The circle would close for a frequency of
150 MHz. (l = λ/2).

Short and Open Circuit Stub


Particularly simple equations for the input impedance of a transmission line can
be obtained in case of a short-circuit Z L = 0 or in case of an open circuit Z L → ∞.
We first consider the case of a short circuit. In this case equation 2.39 will reduce
to:
Z in = ZC tanh(γl) (2.41)
To illustrate this equation, we consider the following example.

Example: Input resistance of a short-circuit stub


Data: ZC = 50 Ω, l =1 m, α = ˆ 1 dB/m
Figure 2.15 shows that the input impedance is inductive for low frequencies. If the

Figure 2.15: Input impedance of a short circuit stub

frequency continues to rise, the impedance shows at 75 MHz a resonance. Around


this frequency, a short-circuit stub behaves like a parallel-resonant circuit. Since

24
the line length is λ/4 at this frequency, it is also called λ/4-resonator. With further
increasing frequency, the stub shows capacitive behaviour. At a frequency of ≈
150 MHz the length of l approaches a value of λ/2 and the short-circuit stub
shows the behaviour as a series-resonant circuit.
From this example we can conclude that a short-circuit stub allows to realize a
inductive reactance, parallel resonance circuit, capacitance and series resonance
circuit, even if the length of the transmission line has to become longer. For
this purpose, the circuit elements are set above ca. 300 MHz in order to obtain
concentrated reactive components. Now we want to examine in more detail which
inductance values can be realized with a short-circuit lossless transmission line.
For the input impedance of this circuit we get:
l
Z inS = jZC tan(βl) = jZC tan(2π ) (2.42)
λ
Equation 2.42 shows that the input impedance of the circuit for l < λ/4 behaves
like an inductive reactance. Defining fλ/4 as frequency the length of the line
becomes a fourth of the wave length, we obtain as value of a corresponding in-
ductance:
ZC π f
LS = tan( ) (2.43)
ω 2 fλ/4
In contrast to ideal lumped inductance, the inductance value LS ( f ) of a short-
circuit stub becomes a function of the frequency. Its functional dependence is
illustrated in figure 2.16. We can see that the inductance Lk shows a value which is
for low frequencies approximately frequency-independent but, however, increases
considerably starting at f / fλ/4 ≈ 0.5. If an inductance is realised by a short-
circuit stub, we should keep the additional frequency response as low as possible.
Accordingly, a lower limit can be determined for the characteristic impedance of
the stub:
ZC ≥ π fλ/4 LS (2.44)
The description of the behaviour of an open-circuit stub can be deduced on the
same basis of the description applied to the short-circuit stub if the admittance
is considered instead of the impedance. According to equation 2.39, following
applies to the input admittance of the open-circuit stub:
l
Y inO = jYC tan(2π ) (2.45)
λ
In principle the admittance of the open-circuit stub is governed by the same equa-
tion already known from the short-circuit stub. By changing the interpretation
of figure 2.15, we can see that an open-circuited stub for low frequencies oper-
ates as a capacitance. Does the frequency further increase it shows a resonance

25
Figure 2.16: Normalized inductance of a short-circuit stub

at l = λ/4. This resonance corresponds to a series resonance circuit. Above this


frequency, the behaviour becomes inductive. A stub open-circuit stub below the
λ/4-resonance may thus be used to realise a capcitance. For its value we obtain:
YC π f
CO = tan( ) (2.46)
ω 2 fλ/4
With the same arguments, a lower limit for the characteristic admittance of the
stub can be deduced:
YC ≥ π fλ/4 CO (2.47)

2.3 Line Transformer


A problem appearing frequently in the RF circuit technology is illustrated in Fig.
2.17. An ohmic load with resistance RL must be matched to a generator with
internal resistance RG . The problem is to find a suitable network, that is able
to transform the load resistance RL into an input resistance Rin = RG at a prede-
termined frequency f0 . The so-called λ/4-transformer provides a narrow-band
solution to this problem.

2.3.1 Quarter-Wave Transformer


An approach to understand the operation principle of a λ/4-transformator is given
by equation 2.40. If we take the line length l in a wise that it is equal to λ/4 at the

26
Rin
RG
s s
 matching
-
 network RL
s s

Figure 2.17: Block diagram of a matching circuit

desired frequency, we obtain for the input resistance Rin :

ZC2
Rin = for f0 = fλ/4 (2.48)
RL
As we can see, the input resistance is real at the operating frequency. With the
converse equation 2.48 we determine the necessary characteristic impedance ZC
of the transmission able to transform the given load resistance RL into the desired
input resistance Rin . √
ZC = RL Rin (2.49)
Example: λ/4-Transformator for matching RL (Fig. 2.17)
RL = 200 Ω, RG = 50 Ω and f0 =3 GHz

ZC = 200 50 Ω = 100 Ω

To determine the line length l, simplifying we suppose a air-filled transmission


line (εr = 1).
c0 3 · 108
l = = m = 2, 5 cm
4 f0 12 109
Fig. 2.18 shows the input reflection coefficient in dB calculated with the help of a
simulation tool. One can see that the reflection coefficient in the frequency range
between 2.55 GHz and 3.45 GHz is below -15 dB. This corresponds to a band-
width of approx. 900 MHz. This bandwidth will be sufficient for many applica-
tions. However, the behaviour of the curve in 2.18 is very frequency-dependant.
Due to its resonant behaviour, a single-level λ/4-transformator enables only a
narrow-band matching.

27
Figure 2.18: Frequency response of a single step λ/4-transformator

28
Chapter 3

Scattering Parameters

The linear and linearised circuits of RF- and microwave engineering can be sub-
divided into two groups, one group using voltages and currents to describe their
behaviour leading to the well known classical z-, h- or y-parameters [2] and a
second group, e.g. waveguide circuits, where the description is essentially based
on waves, because it is not possiple to define voltages and/or currents in a unique
way. Here usually s-parameters are used for description [3]. During the last years
the importance of the s-parameter has signifanctly grown due to the availability of
modern vector network analyser (VNA).
While below approximately 10 MHz passive or linearised active two ports are still
described using the classical parameters, with increasing frequency this becomes
more and more difficult. We will discuss briefly the situation for y-parameters,
which where traditionally used in the RF-domain.

• With increasing frequency the measurement of the y-parameter becomes


impratical, because short-circuits are hard to realise while measuring volt-
ages and currents.

• Active elements e.g. transistors tend to become unstable if they are termi-
nated in short-circuits.

• In elements like a rectangular waveguide it is no more possible to define


voltages and currents in a unique way.

To avoid all these problems the characterisation of circuits must be based on pa-
rameters which are easy to measure. At the dawn of microwave engineering it was
noticed that it is easier to measure powers associated with certain waves. With the
help of directional couplers it is possible to separate forward and backward travel-
ling waves. This was the starting point to develop a measurement technique based
on wave amplitudes and scattering parameters.

29
3.1 Definition of Wave Amplitudes
As already dicussed in the previous section the concept of waves is of essential
importance for the definition of the s-parameter and the assiocated measurement
technique. Instead of using the voltages U 1 , U 2 or the currents I 1 , I 2 defined at the
input or output one introduces wave amplitudes a1 , a2 und b1 , b2 at the input and
output port. With the help of the transmission line shown in Fig. 3.1 their defini-
tion will be explained. In the following we will describe the transmission line in
e 
a1 - -b
2
Uf -  Ur
? b1   a2 ?
e 
-
z1 z2 z
a1- b2-
 (S) 
b1 a2

Figure 3.1: Wave amplitudes on a transmission line

the region from z = z1 to z = z2 by an equivalent two port and its s-parameters. As


already stated in the previous chapter on a transmission line there generally exit
two waves: the forward and backward travelling wave (reflected wave). With the
help of the wave amplitude a1 the forward travelling wave will be described at the
location z = z1 while the wave amplitude b1 shall represent the backward travel-
ling wave at the same location. Since the measurement technique was originally
based on power measurements one defines:
1 √
Pf = |a1 |2 → |a1 | = 2Pf (3.1)
2
1 √
Pr = |b1 |2 → |b1 | = 2Pr (3.2)
2
By the equations 3.1 and 3.2 only the absolute values of the complex wave am-
plitudes a1 and b1 are defined. By assigning to the wave amplitude a1 the phase
of the voltage wave U f at z = z1 and to the wave amplitude b1 the phase of the
voltage wave U r at the same location, a unique phase can be assigned to both. The
planes transverse to the waves at z1 and z2 are called reference planes of the two
port. Remembering the following equations for the power transported by waves,
|U f |2 |U r |2
Pf = Pr =
2ZC 2ZC

30
we will get for the wave amplitudes:
U f (z1 ) U r (z1 )
a1 = √ b1 = √ (3.3)
ZC ZC
Using the same argumentation we get for the wave amplitudes a2 and b2 :
U r (z2 ) U f (z2 )
a2 = √ b2 = √ (3.4)
ZC ZC

3.2 Definition of S-Parameters


3.2.1 2-Port
According to Fig. 3.1 a transmission line of length l = z2 − z1 forms a two port.
In general the emerging wave amplitudes b1 and b2 are given as functions of the
incident wave amplitudes a1 and a2 with the help of the s-parameters.
b1 = S11 a1 + S12 a2
(3.5)
b2 = S21 a1 + S22 a2
Using the defining equation one can already deduce the meaning of the single
parameter.
b Input reflection coefficient of port 1, while port
S11 = 1 |a2 =0
a1 2 is matched
b Input reflection coefficient of port 2, while port
S22 = 2 |a1 =0
a2 1 is matched
b Transmission coefficient from port 1 to port 2,
S21 = 2 |a2 =0
a1 while port 2 is matched
b Transmission coefficient from port 2 to port 1,
S12 = 1 |a1 =0
a2 while port 1 is matched
If we now represent the emerging wave amplitudes b1 and b2 of the lossless trans-
mission line section (Fig. 3.1) as a function of the incident wave amplitudes a1
and a2 we get:
U f (z2 ) U f (z1 )
b2 = √ = exp[− jβ(z2 − z1 )] √ = exp(− jβl)a1
ZC ZC
U (z1 ) U (z2 )
b1 = √r = exp[+ jβ(z1 − z2 )] √r = exp(− jβl)a2
ZC ZC
As result a lossless transmission line section of length l is described by the fol-
lowing s-matrix. ( )
0 exp(− jβl)
(S) =
exp(− jβl) 0

31
3.2.2 N-Port
It is easy to extend the equations 3.5 of a 2-port to an arrangement with an arbitrary
number N of ports:
    
b1 S11 S12 · · ·S1N a1
 b   S  
 2   21 S22 · · ·S2N   a2 
 ..  =  .. .. .  .  (3.6)
 .   . . · · · ..   .. 
bN SN1 SN2 · · ·SNN aN

Like in the case of a 2-port the meaning of the single elements can be deduced
from the defining equations. The element Snn represents the input reflection co-
efficient of port n, while all other ports must be matched. Analogous the element
Snm descibes the transmission coefficient from port m to port n, while again all
other ports must be matched. To reduce the writing effort the last equation can be
put into matrix notation. If we define a complex vector of the emerging wave am-
plitudes ⃗b by (b1 , b2 · · · bN ) and a complex vector of the incident wave amplitudes
⃗a by (a1 , a2 · · · aN ) and the scattering matrix is denoted by (S), we can write:

⃗b = (S)⃗a (3.7)

3.2.3 Useful Definitions


To characterise the properties of N-ports described by s-parameters some useful
definitions are used, which will be introduced briefly:

• One denotes an N-port as free of selfreflections, if the following conditions


hold true Sn,n ≡ 0 for all n ∈ 1 · · · N.

• An N-port is said to be reciprocal if for all n, m ∈ 1 · · · N and n ̸= m we have


Sn,m ≡ Sm,n .

• An N-port is said to be lossless, if the power sum of all incident waves an is


equal to the power sum of all emerging waves bn .

• A 2-port is said to be unilateral or free of internal feedback if S21 ̸= 0 and


S12 ≡ 0

32
3.3 Lossless Networks
Knowing the S-matrix of a network it is easy to decide for example wether a
network is reciprocal or not because of the obvious properties of the s-parameters.
In case of a lossless network the situation is more complicated. Nevertheless
the property of losslessness will impose certain constrains on the values of the s-
parameters. To study this in more detail we will first have a look on the properties
of a lossless 1-port and 2-port and than we will extend our knowledge to the case
of a lossless N-port.

1-port
An arbitrary 1-port is described by the following equation:

b1 = S11 a1

According to equation 3.1 and 3.2 the absolute values of the wave amplitudes a1
and b1 are directly related to the incident power Pin and the emitted power Pout of
a 1-port.
1 1
Pin = |a1 |2 Pout = |b1 |2
2 2
If the 1-port is lossless the following equation must hold true:

|b1 |2 = |S11 |2 |a1 |2 = |a1 |2 → |S11 | = 1

To be a lossless 1-port the reflection coefficient S11 must show an absolute value
of 1, while the phase may have an arbitray value.

2-port
To go on in our discussion, we now want to examine which conditions are imposed
onto the s-parameters of a 2-port if we know it to be lossless. For the sake of
simplicity we will first discuss the case where only the wave described by a1 will
enter port 1, while a2 is assumed to be zero. According to equation 3.5 there will
still exist the wave amplitudes b1 and b2 which will be emitted from port 1 and
port 2. As a consequence we get for the power Pin entering the 2-port and the
power Pout emitted by the 2-port:
1 1[ 2 ]
Pin = |a |2 Pout = |b1 | + |b2 |2
2 1 2
If we replace the wave amplitude b1 by S11 a1 and b2 by S21 a1 we find in the case
of a lossless 2-port:

Pin = Pout → |S11 |2 + |S21 |2 = 1 (3.8)

33
Equation 3.8 is termed the first absolute value condition of a lossless 2-port. If we
now assume only a wave is entering port 2 (a2 ̸= 0, a1 ≡ 0) we can conclude:

|S12 |2 + |S22 |2 = 1 (3.9)

Equation 3.9 is termed the second absolute value condition of a lossless 2-port. Of
course in general waves will enter port 1 and port 2 simultaneously. But to reduce
the writing effort we will now use matrix notation according to equation 3.7. If
one defines ⃗a† to be the row vector (a∗1 , a∗2 ) in mathematical terms it will be the
adjoint vector to the column vector ⃗a. With the introduced vectors one can write
the power entering the 2-port:
1[ 2 ] 1
Pin = |a1 | + |a2 |2 = ⃗a†⃗a
2 2
Analogous one gets for the power emitted from the 2-port Pout :
1[ 2 ] 1 †
Pout = |b1 | + |b2 |2 = ⃗b ⃗b
2 2
The condition for the 2-port to be lossless can now be reformulated :

⃗b†⃗b = [(S)⃗a]† (S)a = ⃗a†⃗a

Bearing in mind the mathematical identity [(S)⃗a]† = ⃗a† (S)† and ⃗a†⃗a = ⃗a† (E)⃗a
with (E) being the unity-matrix, the above equation can be rewritten as:
[ ]
⃗a† (S)† (S) − (E) ⃗a = 0

As a consequence we now get the general condition for a 2-port to be lossless.


Since we used matrix notation, it directly applies also to an N-port (equation 3.10).

(S)† (S) = (E) (3.10)

A matrix fulfilling the condition stated in equation 3.10 is called a unitary matrix.
In detail equation 3.10 reads in case of a 2-port:
( ∗ )( ) ( )
S11 S∗21 S11 S12 1 0
=
S∗12 S∗22 S21 S22 0 1

In addition to the already known absolute value conditions 3.8 and 3.9 there exist
two further equations:

S11 S∗12 + S21 S∗22 = 0 und S∗11 S12 + S∗21 S22 = 0 (3.11)

34
Since the second equation is just the conjugate complex of the first equation, it
is sufficient only to take the fisrt equation into account, which imposes further
conditions on the s-paramter. This condition is called phase condition. In the
case of a N-Port to be lossless the N × N s-parameterss must fulfill N(N + 1)/2
conditions.

Example 3.2: S-parameter of a series inductance inside a transmission line


Data: L =20 nH, ZC = 50 Ω, f =100 MHz.
The parameter S11 is equal to the input reflection coefficient of port 1 if port 2 is

 l≈0 -
s  s
L
ZL = 50Ω ZL = 50Ω
s s

Figure 3.2: Series inductance in a transmission line

matched.
( jωL + ZC ) − ZC ωL
S11 = ≈ j = j0, 126
( jωL + ZC ) + ZC 2ZC
Due to the setup there is essentially no difference between port 1 and port 2, thus
we have S22 = S11 = j0, 126. To calculate the parameter S21 one can use the fact
that an ideal inductance is lossless and hence the first absolute value condition can
be applied. √
|S21 | = 1 − |S11 |2 = 0, 992
If one chooses the parameter S21 to be real also the phase condition according to
equation 3.11 is fulfilled. A series inductance inside a 50 Ω transmission line at a
frequency of a 100 MHz thus can be described by the following s-matrix:
( )
j0, 126 0, 992
(S) =
0, 992 j0, 126

3.4 Signal Flow Graph


3.4.1 Definition of Graphical Elements
It is always of advantage if the abstract mathematical operations defined for exam-
ple by equation 3.5 can be represented by graphical elements in a more descriptive

35
way. This is done with the help of the signal flow graph. In a signal flow graph
wave amplitudes like an or bm are represented using dots, while a scattering pa-
rameter like Smn will be represented by an arrow starting at the dot with index n
and ending at a dot with the index m. Fig. 3.3 gives some simple graphical rep-
resentations and the corresponding equations. Fig. 3.3a shows for example the
s
al H
HHSil
Smn SikHHH
an s -s b
m
ak s

js b
-

* i
Si j
a) bm = Smn an b) s 
aj 
bi = Si j a j + Sik ak + Sil al

Figure 3.3: Simple equations and their graphical representation

graphical representation of the equation bm = Smn an , while Fig.3.3b is equivalent


to the equation bi = Si j a j + Sik ak + Sil al . Using the simple rules of linear equa-

S′mn @@
Skk
an bk bm @ an s @ b
s -s -s an s ′′ @
Rs bm @
Rs
- -s m
Skn Smk @ @
S mn  Skn bk Smk
H
H H
H
H
H
ans b
-sm
an s b an s
-s m
b
-s m
′ ′′ Smk Skn
Smk Skn Smn + Smn
a) b) c) 1 − Skk

Figure 3.4: Examples to simplify signal flow graphs

tions one can deduce some simple approaches to reduce the complexity of a given
signal flow graph. Fig. 3.4a for example shows how the wave amplitude bk can
be omitted by a resulting scattering parameter Smn = Smk Skn . In Fig. 3.4b a signal
flow graph is shown representing a wave amplitude bm which is composed via
two different branches from the wave amplitude an . It is obvious that the effec-
tive transmission parameter Smn is given by the sum of S′mn und S′′mn . A somehow
more complicated flow graph is shown in Fig. 3.4c. Here for the first time a loop
occurs. To simplify the structure one can split node bk in a further node b′k with a
transmission coefficient of 1 between the nodes. For the node bk we get:

Skn
bk = Skn an + Skk b′k = Skn an + Skk bk → bk = a
1 − Skk n

36
Taking into account the transmission factor Smk between the nodes bk and bm , we
end up in the result depicted in Fig.3.4c. With the help of the shown examples all
signal flow graphs discussed in this lecture can be understood. In flow graphs with
a lot of nodes the operations will become laborious and it will be of advantage to
use the rules given by Mason [6].

3.4.2 Special Devices and their graphical Representation


s ps -s
S11 @ 
(S) 4 3 4 @ 3
s
? s @ s
a) 1   2 @I @
@
I


@
R @ @
S
s 21 -s - s @ @
Rs
-
S21 @
6 f)
(S) S11 S22 1 @ 2
?
s s - S31 s
 @s
b) S12
 s -s s
j exp(− jβl) 
 S11 = 0
 s
 s s?
c) g)
s -s s

- S21 = 1 ? S11 = 0

s s s
?
d) h)
s -s s 1 a=s

φ exp(− jφ) 6
G 0
e) 
s s i  s

Figure 3.5: Block diagram and signal flow graphs of different devices

To analyse circuits of RF- and Microwave engineering with the help of the s-
parameters and their signal flow graph Fig. 3.5 shows the block diagram of some
special devices and their signal flow graph. We want to discuss the properties of
these devices using their representation in a signal flow graph without going into
the details of their physical realisation.

a) 1-port
Arbitrary 1-port, described by its frequency dependent input reflection co-
efficient S11 .

b) 2-port

37
Arbitrary 2-port, described by its frequency dependent s-parameters S11 ,
S12 , S21 and S22 .

c) Transmission line
Lossless transmission line of length l. It is free of self-reflections S11 =
S22 = 0. The incident waves suffer a phase change of φ = −βl. The kind of
line, e.g. circular wave guide, rectangular wave guide, coaxial transmission
line, is denoted by a small additional symbol.

d) One-way attenuator
In a one-way attenuator a wave only can propagate in the declared direction,
while the reflected wave is attenuated and finally absorbed in the device.

e) Phase-shifter
With the help of a phase-shifter the phase of a wave is changed. In most
cases it is realised by a transmission line of variable length.

f) Directional coupler
Directional couplers are realised using two coupled transmission lines. It
is free of self-reflections and the coupling is chosen to decouple two ports.
With the help of a directional coupler it is possible to determine the wave
travelling in different directions. The ideal directional coupler is described
by its transmission-factor and coupling coefficient.

g) Absorber
In the case of a transmission line a resistor with a value equal to the char-
acteristic impedance of the transmission line is an ideal absorber. The wave
entering the absorber will be totally absorbed, so its input reflection coeffi-
cient is zero.

h) Power meter
The ideal power meter shows no self-reflections and is able to determine the
power of the incident wave.

i) Generator
It provides the necessary power to drive a microwave circuit. The ideal
generator will be free of self-reflections.

As an example to apply the signal flow graph we will have a look on the setup
shown in Fig. 3.6. It shows a generator which feeds its wave using a one-way
attenuator and a directional coupler to a 1-port, which is described by its reflection
coefficient rA . The ports 3 and 4 of the directional coupler are terminated by ideal

38
power meters, which are able to measure the absolute values of the emerging
wave amplitudes b3 and b4 . Also shown in Fig. 3.6 is the associated signal flow
graph. To enhance the clearness of the flow graph only nodes and branches are
included which are of importance for the measurement setup. To understand the
operation principle of the setup we will calculate the wave amplitudes b4 and b3
in dependence of the wave a0 generated by the rf generator.

 

 

? ?


-   rA
G @
I


@
R


:s b3
s b4 
S H
YH 
a0s S
-10
s
a1 31 HH S21
-s -s 2
b
HH
S31HH rA
H
s a2
H?

Figure 3.6: Principle measurement setup to measure a reflection coefficient and


the associated signal flow graph

b3 = S31 S10 a0

b4 = S31 rA S21 S10 a0


According to the signal flow graph the wave amplitude b4 will be proportional to
the reflection coefficient rA and the wave amplitude a0 , while the wave amplitude
b3 is only proportional to the wave amplitude a0 . If one determines the ratio of the
wave amplitudes b4 and b3 and accounts for the fact that due to the power meters
only their absolute value can be measured one gets for the absolute value of the
measured reflection coefficient rM :
b4
|rM | = | | = |S21 ||rA | (3.12)
b3
The result is essentially proportional to the reflection coefficient rA . Unfortunately
the measured result is distorted by the unknown transmission factor S21 of the used
directional coupler, which introduces a systematic error. To eliminate this system-
atic error in a first step a known reflection coefficient e.g. a short circuit can be
measured which allows the determination of the absolute value of S21 (calibra-
tion of the measurement setup ) and its value can be used to correct the measured

39
reflection coefficient rM . In the next section we will discuss this in more detail,
but for now we will discuss a further systematic error which will occur if a real
directional coupler is used in the measurement setup. Fig. 3.7 shows essentially
b4
s :s b3


Y
HHH S31
 
S41 H

a0 S10   HH S21 -s b
s -s -s H 2
HH
S31 H rA
HHs?
a2
Figure 3.7: Real directional coupler with finite directivity

the same signal flow graph already given in Fig. 3.6. The only difference is the in-
clusion of a further branch termed S41 . It accounts for the finite isolation between
port 4 and port 1 of a real directional coupler. Using the flow graph of Fig. 3.7
one again can calculate the ratio of the wave amplitudes b4 /b3 .
b4 S
= rM = S21 rA + 41
b3 S31
The finite isolation between port 4 and port 1 results in an additional term. Its in-
fluence becomes dominant for small reflection coefficients to be measured |rA | ≈ 0
and suggests a measured reflection coefficient of the object even if its value is zero.
S41
|rM | ≈ | |
S31

The abolute value of the ratio S41 /S31 of a directional coupler will thus set a lower
limit to the reflection coefficient which can be measured without calibration. This
ratio is thus a kind of quality factor of a real directional coupler. One denotes the
value aD = −20 log(|S41 /S31 |) as the directivity of a directional coupler.

3.5 Network Analyzer


To measure the linear properties of passive circuits or the linearized properties
of active devices in the rf- and microwave domain modern network analyzers are
used today. One can distinguish between scalar (SNA) or vector network ana-
lyzers (VNA). While an SNA is only able to measure the absolute values of the
s-parameters, with a VNA the complex values of the s-parameters (absolute value,
phase) can be measured. Due to the build-in processors the elimination of system-
atic errors is also possible and post processing of the measured data is possible.
So today a VNA is a indispensable equipment in an rf- or microwave lab. Fig. 3.8

40
Internal Computer
PP 
P
Signal Source

Forward ? 
s Backward 
s s B
B
B
a0 ? B

j
b0 B

j A/D B
a3 -×
j 

b3 -×
j 
Test-Set
B B
B B Receiver

Port 1 Port 2

Figure 3.8: Principle setup of a modern VNA

shows the principle setup of a modern VNA. It consists of the following building
blocks:

• Signal Source
The rf-signal for operation is normally provided by a synthesizer. Due to the
build in phase look loops rf-signal with low phase noise and high frequency
accuracy can be generated. In order not to lose the phase information during
the mixing process the synthesizer will generate a second signal with a fixed
phase relation to the rf-signal for measurement.

• Test-Set
The rf-heart of the VNA is the Test-Set. The rf-signal provided from the
synthesizer can be switched between a forward (S11 , S21 ) and backward
(S22 , S12 ) branch. To provide the reference signal for the forward branch
a0 or the backward branch a3 often a lossy power divider is used. Direc-
tional couplers are used to do the seperation of the forward and backward
travelling waves.

• Receiver
The conditioning of the measured signals takes place in the receiver. The
rf-signals are converted down to an intermediate frequency, while the phase
relation of the signal is kept constant. After converting to digital form they
are fed to the internal computer.

41
• Internal Computer
The controlling of the whole measurement setup is done by the internal
computer. It coordinates the collaboration of the different building blocks,
carries out all error correction and processes the data, so it can be displayed
in various formats.

Today VNAs with coaxial connectors are build covering the frequency range from
10 MHz to approximately 60 GHz. Beyond this frequency range Test-Sets are
usually realised in wave guide technique.

1-Port Calibration
As already discussed, a real directional coupler shows properties which lead to
systematic errors when measuring e.g. the reflection coefficient of a device under
test (DUT). So for example the finite directivity of the coupler limits the lowest
measureable reflection coefficient. Due to the available data processing capability
of modern computers it is now possible to setup complicated error models, which
describe the systematic errors and which are used to eliminate them. The concept
a0 b0 a0 b0

 HH 
  HH 
 error
G rA G rA
 real  2-port
ideal
a) b)

a0 s e10 -s
@
@ @
@
e00 e11 rA
c) b0  e
s 01 @
@s

Figure 3.9: 1-port error model

of a 1-port error model is depicted in Fig. 3.9. In Fig. 3.9 a) the principle setup
of a real reflectometer is shown, which will result in systematic errors of the mea-
surement. As shown in Fig. 3.9 b) one can think of the systematic errors to be due
to an error 2-port, which is placed between the DUT and the now assumed ideal
reflectometer. Fig. 3.9 c) shows the associated signal flow graph. With the help of
a calibration procedure the error parameters of the error 2-port can be determined
and a correction of the measured data can be carried out. The error 2-port will
result in a difference between the real reflection coefficient rA of the DUT and the
measured reflection coefficient rM . Carrying out the calculation shows that three

42
error terms will have an influence: e00 , e10 e01 and e11 as shown in equation 3.13.
e10 e01
rM = e00 + r (3.13)
1 − e11 rA A

In the following section we will briefly discuss the influence of the three parame-
ters:

• e00
According to equation 3.13 the value of the parameter e00 becomes domi-
nant when the reflection coefficient of the DUT is small. In the limit rA → 0
we have rM = e00 . Thus it accounts for the finite directivity of the direc-
tional couplers used in the measurement setup.

• e10 e01
The influence of the above product becomes most significant in the limit e00
= e11 = 0. In this case we get rM = e10 e01 rA . Comparing this result with
equation 3.12 shows that the product e10 e01 accounts for the transmission
factors of the used directional couplers.

• e11
This parameter accounts for the finite input reflection coefficient of the mea-
surement setup.

In the ideal case the parameters of the error 2-port (equation 3.13) should show
the following values: e00 ≡ 0, e10 e01 ≡ 1 and e11 ≡ 0. In this limit the measured
reflection coefficient rM would be identical to the reflection coefficient rA of the
DUT.

OSL-calibration According to equation 3.13 the error 2-port is described by


three independent parameters which are frequency dependent. To carry out a
correction all three terms have to be identified during a calibration process with
three independent calibration standards. An often used calibration method is the
so-called OSL-calibration. The three letters correspond to the used calibration
standards, which are Open Short and Load. In the following section we want to
examine how one can determine the error parameters e00 , e10 e01 and e11 using the
measurement results of the calibration standards.

• ”Load” - Measurement rML


To carry out this measurement the port of the network analyzer has to be
terminated with its characteristic impedance, which in the ideal case results
in a reflection coefficient of rA = 0. This is one of the drawbacks of the
OSL-calibration, because with increasing frequency it becomes more and

43
more complicated to really match the port. As result of an ideal load we
will get as measured reflection coefficient:

rML = e00 (3.14)

Measuring the ”Load”-standard thus directly results in the knowledge of


the error parameter e00 . A more accurate result in determining e00 can be
obtained if instead of a load a so-called sliding load is used [2].

• ”Open” - Measurement rMO


The realization of an open standard is almost impossible under a practical
point of view. For example an open coaxial transmission line would show
fringing fields at its end and the aperture would radiate resulting in a re-
flection coefficient smaller than one. Instead one can realize the standard as
schematically sketched in Fig. 3.10. Here the outer conductor of the coaxial
transmission line is closed to avoid radiation and the inner conductor is not
connected to the outer conductor resulting in a capacitance between them.
To use such a kind of ”Open” as calibration standard the electrical behaviour
of the capacitance must be known very precisely. In the case of no losses its

-
-
- CE

Figure 3.10: Example for an ”Open” calibration standard in coaxial transmission


line technique

reflection coefficient will have the following form rA = exp(− jδC ). The fre-
quency dependent phase change δC due to the capacitance must be known
very accurately to use it as calibration standard. Using equation 3.14 we
will get:
e10 e01
rMO = rML + exp(− jδC ) (3.15)
1 − e11 exp(− jδC )
• ”Short” - Measurement rMS
The realisation of a ”Short” calibration standard is less challenging, at least
at low frequencies. Again using equation 3.13 we will get for the measured
reflection coefficient:
e e
rMS = rML − 10 01 (3.16)
1 + e11

44
The equations 3.15 and 3.16 can be used to determine the still unkown error pa-
rameters e11 and e10 e01 :

rM0 exp( jδC ) + rMS + rML [1 + exp( jδC )]


e11 = (3.17)
rMO − rMS

(rML − rMS )[1 + exp( jδC )]


e10 e01 = (rMO − rML ) (3.18)
rMO − rMS
With the help of equations 3.14, 3.17 and 3.18 and the measured reflection coeff-
cients rML , rMO and rMS all error parameters can be calculated for every frequency
point of interest. The denominator of equation 3.18 and 3.17 is of special impor-
tance. It is given by the difference of rMO − rMS . Under normal conditions the
absolute value of the difference will be close to 2. A problem will occur if the
phase due to the capacitance will become δC = π. In this case the measured
reflection coefficients rMO and rMS would be almost identical and the error pa-
rameters can no longer be determined. This puts an upper frequency limit to the
OSL calibration process. If for all frequency points the error parameters according
to Fig. 3.9 are known a correction can be carried out using the following equation:
rM − e00
rA = (3.19)
e11 (rM − e00 ) + e10 e01

45
Chapter 4

Microwave Amplifier

The small signal behaviour of active elements like bipolar junction transistor
(BJT) or field effect transistors (FET) can be described using different physically
based equivalent circuits. But to allow a unified treatment it is of advantage to use
2-port parameters like y- or s-parameters. Based on the used parameters equations
for some characteristic properties of amplifier e.g. input- and output impedance,
voltage gain or power gain can be deduced without reference to the physics of
the device. Beside the s-parameters also the y-parameters can be used to deduce
some properties of the amplifiers. Here again we want to point out that the small
signal parameters not only depend on the considered frequency but also depend
on the chosen point of operation. Unfortunately this is not directly obvious at the
deduced equations.
Even so today amplifiers working up to the GHz-range are available as monolithic
integrated circuits we will focus on simple amplifiers with only a single transistor
to explain the basics of amplifier design. Furthermore we will not discuss how to
establish the necassary point of operation, because this has already been done in
the lecture of electronics. As an example Fig. 4.1 shows the s-parameters of the
BJT BFR 280 in a point of operation with UCE = 5V and IC = 3mA according to a
data set provided by the manufacturer.

4.1 Y-Parameters
Besides the s-parameters especially in the rf-domain, y-parameters are used to
decribe the linearised behaviour of active elements. Equation 4.1 gives the cur-
rents of a 2-port independence of the voltage, when y-parameters are used for its
description.
I 1 = y11U 1 + y12U 2
(4.1)
I 2 = y21U 1 + y22U 2

46
Figure 4.1: S-parameters of the BJT BFR 280

The meaning of the single parameter can be explained by using equation 4.1 under
certain conditions:
I
y11 = U1 |U 2 =0 short-circuit input admittance
1
I
y12 = U1 |U 1 =0 short-circuit reverse transfer admittance
2 (4.2)
I2
y21 = U |U 2 =0 short-circuit forward tranfer admittance
1
I2
y22 = U |U 1 =0 short-circuit output admittance
2

By looking at equation 4.2 it becomes obvious that all currents have to be mea-
sured using an rf-short-circuit. In the lower rf-domain this can be realised us-
ing large capacitors. But with increasing frequency this becomes more and more
complicated. A further drawback is the tendency of active devices to become
unstable when operating under short-ciruit conditions. These are some reasons
why today active devices are in most cases described by s-parameters. Never-
theless y-parameters are sometimes still of interest because certain situations are
easierer to describe. Fig. 4.2 shows an equivalent circuit exclusively based on the
defining equations 4.1. In principle it is simply their graphical representation. A
further equivalent circuit is shown in Fig. 4.3 which is called π-equivalent circuit
and which is used to describe the linearised behaviour of BJTs. Essentially it is
equivalent to the physical motivated equivalent circuit given by Giacoletto [4].

47
y12U 2 y21U 1 I2
I1
s- s s  s
? ?
 
U1 y11 y22 U2
 
? ?
s s s s s s

Figure 4.2: Formal equivalent circuit according to the y-paramters

−y12 (y − y )U I2
I
s-
1 s 21
s s12 1
 s
?
 y22
U1 y11 + y12 U2
 +y12
? ?
s s s s s

Figure 4.3: π-equivalent circuit

4.1.1 2-Port described by y-parameters


In the following section we want to deduce some properties of 2-ports using y-
parameters. Starting point will be the circuit shown in Fig. 4.4. Here a 2-port is
driven by an rf-generator with the inner admittance Y G while it is terminated in
the load admittance Y L . It is assumed that the 2-port is formed by an active device

I0
- s s

Y G U1 2-port U2 YL

? ?
s s

Figure 4.4: Active 2-port driven by an rf-generator

or even by a complete amplifier, which is described using y-parameters.

48
Voltage Gain
The voltage gain vu ( jω) is defined as the ratio of the output voltage U 2 to the input
voltage U 1 . Using the second equation of 4.1 and the fact that the load admittance
is given by Y L = −I 2 /U 2 we get for the voltage gain of the circuit in Fig. 4.4:
U2 y21
vu ( jω) = =− (4.3)
U1 y22 +Y L

Input Admittance
The influence of the 2-port onto the rf-generator is always of interest. It is de-
scribed by the input admittance yin . Starting from the first equation of 4.1 and
using equation 4.3 we find:
I1 y y
yin = = y11 + y12 vu ( jω) = y11 − 12 21 (4.4)
U1 y22 +Y L

Output Admittance
The output admittance yout is given by the ratio of current I 2 to voltage U 2 . Essen-
tially it can be deduced in the same manner as the input admittance. One just has
to replace index 1 by 2 and vice versa and the load admittance has to be replaced
by the generator admittance Y G .
I2 y y
yout ( jω) = = y22 − 21 12 (4.5)
U2 y11 +Y G

4.2 Stability of a 2-Port


A 2-port can only work as an amplifier if it contains some active devices like BJTs
or FETs. Unfortunately for certain load admittances Y L or for certain generator
admittances Y G the 2-port may become unstable - it will show oscillations. In
a more mathematical formulation this means that the voltages U 1 and U 2 exist
even when there is no excitation (I 0 = 0 in Fig. 4.4). To explain the possibilty of
oscillations we have a look at Fig. 4.5. Here the 2-port is sketched divided into a
forward branch and a backward branch. The forward branch will show the voltage
gain:
U2 y21
= −
U1 y22 +Y L

49
1’ s s
YG ?Y 12U 2

1s U ′1 Y 11  s s
?Y U
?  21 1
U1 s s s Y 11 U2

? ? YL
s s s s s

Figure 4.5: Forward and backward branch of a 2-port

Due to the reverse transfer admittance y12 a voltage U ′1 will be excited. For the
ratio U ′1 /U 2 we get:
U ′1 y12
= −
U2 y11 +Y G
The voltage U 1 at point 1 will result in the voltage U ′1 at point 1’. If both voltages
are equal in value and phase the points 1 and 1’ can be connected and as a result
we will have a self-sustainable oscillation.
y12 y12 y21
U ′1 = − U2 = U1
y11 +Y G (y11 +Y G )(y22 +Y L )

As a result we can formulate a condition for a selfsustainable oscillation expressed


using the y-parameters:
y12 y21
1 = (4.6)
(y11 +Y G )(y22 +Y L )

The equation above can be rearranged into two different equivalent forms:
y y21
Y G + y11 − y 12+Y =0
22 L (4.7)
YG + yin =0
y y12
Y L + y22 − y 21+Y =0
11 G (4.8)
YL + yout =0
The new forms can also be used to formulate a condition for stability:

50
One says a 2-port is unconditionally stable, if for an arbi-
trary load admittance Y L ( generator admittance Y G ) the
condition gin = Re(yin ) > 0 (gout = Re(yout ) > 0) is fulfilled.
A necessary condition for unconditional stability is g11 > 0
and g22 > 0.
Only a few devices will be unconditionally stable. So it makes sense to discuss
for a given set of y-parameters which load admittance Y L will result in a stable
and which one will result in an unstable circuit. To deduce an expression we
want to assume the necessary conditions g11 > 0 and g22 > 0 to be fulfilled. As
a starting point for the derivation we will use the equation for the real part of the
input admittance yin :
y y
gin = g11 − Re( 12 21 ) > 0
y22 +Y L
After several conversions one gets:

Re{y12 y21 }(g22 + GL ) + Im{y12 y21 }(b22 + BL )


gin = g11 −
(g22 + GL )2 + (b22 + BL )2

The denominator of the equation above shows a positive real value. Hence the
product with gin remains positive and can be divided by the positive parameter g11
without changing its sign.

gin [(g22 + GL )2 + (b22 + BL )2 ]


=
g11

Re(y12 y21 ) Im(y12 y21 )


(g22 + GL )2 + (b22 + BL )2 − (g22 + GL ) − (b22 + BL ) > 0
g11 g11
(4.9)
Looking at equation 4.9 makes clear that the middle term must show a value
greater than zero if the load admittance shall result in a stable circuit. Completing
the square yields the following expression:
[ ]2 [ ]2
Re(y12 y21 ) Im(y12 y21 ) |y12 y21 |2
GL − (−g22 + ) + BL − (−b22 + ) ≥
2g11 2g11 4g211
(4.10)
In case of an equal sign the equation above defines a circle in the plane of the load
admittance. According to equation 4.10 all admittances outside the circle will
result in a stable circuit, while a load admittance from inside the circle will lead
to a potential unstable circuit. Comparing equation 4.10 with the normal form of

51
a circle equation enables us to define an admittance to the center of the circle in
the load plane Y CL and the radius of the circle ρYL .
Re(y12 y21 )
Re(Y CL ) = −g22 + 2g11
Im(y y )
Im(Y CL ) = −b22 + 12 21
2g11
(4.11)
|y y |
ρYL = 2g12 21
11

Example 4.3: Stability circle of the transistor BFR 280


At 300 MHz the transistor is described by the following set of y-parameters:

y11 = (1, 96 + j3, 197) mS y12 = 0, 54 mS exp(− j90.90 )


y21 = 86, 8 mS exp(− j170 ) y22 = (0, 07 + j0, 995) mS
With the help of these values one can calculate the real part GCL and the imaginary
part BCL of the center of the stability circle:
46, 96 exp(− j107, 90 )
GCL = −0, 0695 mS + Re{ } mS = −3, 76 mS
3, 92

46, 96 exp(− j107, 90 )


BCL = −0, 995 + Im{ } mS = −12, 36 mS
3, 92
For the radius of the stability circle in the load plane we get:
46, 96
ρYL = mS = 11, 98 mS
3, 92
Fig. 4.6 shows the stability circle in the load plane Y L . As already explained all
loads lying inside the circle will result in an unstable circuit.

Up to now the discussion concentrated on the location and size of the stability
circles in the load plane. Using the same argumentation one can also deduce
stability circles in the generator plane. On a more formal base on can get the
equations by interchanging port 1 and port 2.
Re(y12 y21 )
Re(Y CG ) = −g11 + 2g22
Im(y y )
Im(Y CG ) = −b11 + 12 21
2g22
(4.12)
|y y |
ρYG = 2g12 21
22

52
Figure 4.6: Stability circle of BFR 280 at f = 300 MHz

Since admittances are not in favour of impedances the stability circles of the load
or generator admittance can also be mapped onto the corresponding impedance
planes. In a mathematical sense this mapping is an inversion and it is well known
[5] that by an inversion circles are again mapped onto circles. So in the plane of
the generator and load impedance stability circles will also exist.

4.2.1 Stability Factor


To reach stability at a certain frequency the whole stability circle must be located
in the open left half of the load plane. In this case the negative real part of the cen-
ter admittance must be larger than the radius of the stability circle. This relation
is formulated in the following condition:
Re(y12 y21 ) |y12 y21 |
g22 − > (4.13)
2g11 2g11
With the help of the relation above we are able to define the so-called stability
factor K. For stability its value must exceed 1.
2g11 g22 − Re(y12 y21 )
K= > 1 (4.14)
|y12 y21 |

53
Fig. 4.7 shows the stability factor K of the BJT BFR 280 in the frequency range

Figure 4.7: Stability factor of the BJT BFR 280

from 100 MHz to 3 GHz. According to the chart shown in Fig. 4.7 the BJT
BFR 280 becomes stable for frequencies higher than 1.9 GHz. At lower frequen-
cies a circuit using the BJT may become unstable depending on the load admit-
tance Y L used.

4.2.2 Stabilization of an Active Device


In principle the internal feedback of an active device can be neutralised at a certain
frequency [6] by a additional circuit section. Of course one must take care that the
realized circuit does not become unstable at other frequencies. In the microwave
domain this technique becomes almost impossible.
A more convenient possibility to stabilize a given transistor arises from the
knowledge of the stability circle. In the last example we determined the loca-
tion and the radius of the stability circle at 300 MHz for the BJT BFR 280. With
the help of todays simulation tools it is easy to determine the stability circle in the
whole frequency range. In case of the BJT BFR 280 we notice that the stability
circle at 300 MHz reaches furthest into the right half plane of the load admittance.
An additional load admittance of 8.25 mS, which corresponds to a parallel resis-
tor of 121 Ω at the output will be sufficient to stabilize the BJT over the whole
frequency range. Fig. 4.8 shows the absolute value of the parameter s21 before
and after stabilisation with the additional load at the output. Due to the additional
load the parameter s21 will be reduced by 2.8 dB at the frequency of 300 MHz.

54
Figure 4.8: Forward transmission s21 and stability factor K of the BJT BFR 280

In Fig. 4.8 the stability Factor K is also shown. At 300 MHz it shows its lowest
value while over the whole frequency range it is greater than one and one can use
the transistor to build up a stable amplifier.
In the previous example an admittance in the load plane was used to stabilize the
transistor. If we start with stability circles in the generator plane we can determine
an additional admittance to be put parallel to the input to reach stability. But the
stability circles in the impedance can also be used as the starting point of the con-
siderations, resulting in series resistors to be put at the output or the input of the
BJT. Thus in general there exist at least four different methods to reach stability,
which can be chosen taking into account the application the amplifier is to be built
for.

4.3 Power Gain Definitions


One of the basic features of an amplifier in rf- and microwave-systems is for ex-
ample to amplify the power received by an antenna. Depending on the application
different forms of power gain have been defined.

We will now introduce a simple method to define the power gain of an rf-
amplifier. In a first step a power meter is used to measure the available power of a
signal source. In general this will be the maximum available Pav power since sig-
nal source and power meter are matched. In a second step the amplifier is placed
between signal source and power meter and the output power PL is measured. The

55
ratio between the two powers is called transducer power gain GT of the amplifier.
PL
GT = (4.15)
Pav
In the following we want to assume that the internal resistance of the generator
Z G as well as the load resistance Z L are not equal to the characteristic impedance
ZC of the transmission line system used to define the s-parameters. In this case in
a signal flow graph describing the wave between the generator and transistor as
well as between the transistor and load the reflection coefficients rG and rL have
to be considered. To calculate the available power of the source Pav we can use the
signal flow graph shown in Fig. 4.9. The generator will excite the wave amplitude

1 a
a0 s -s1 -s b2
 @@
1
rG r∗G
@
@s 1 s a2
Figure 4.9: Signal flow graph to calculate the available power Pav

a0 . Using a short transmission line (l ≈ 0) with characteristic impedance ZC the


wave will enter the power meter. To reach matching in general the power meter
must show an internal resistance of Z ∗G . The available power Pav can be calculated
by the power difference of the waves b2 and a2 .
1[ 2 ] 1
Pav = |b2 | − |a2 |2 = (1 − |rG |2 )|b2 |2 (4.16)
2 2
The wave amplitude b2 can be expressed by the wave amplitude a0 using the
following relation:
a0
b2 = a0 + |rG |2 b2 → b2 =
1 − |rG |2
At the end we get under the condition |rG | < 1 for the available source power:

1 |a0 |2
Pav = (4.17)
2 1 − |rG |2
The power PL entering the load can also be expressed by the power difference of
the waves bL and aL , as shown in Fig. 4.10.
1[ ] 1
PL = |bL |2 − |aL |2 = (1 − |rL |2 )|bL |2
2 2
56
This results in the following general equation to calculate the transducer gain GT
of an amplifier:
b
GT = (1 − |rG |2 ) · | L |2 · (1 − |rL |2 ) (4.18)
a0

4.3.1 Unilateral Transistor


If we neglect the internal feedback s12 we speak of a unilateral transistor. To
calculate the wave amplitude bL we will use the signal flow graph shown in Fig.
4.10. For the wave amplitude bL in dependence of the wave amplitude a1 we get:

a0 s 1 -s 1 -as1 s21 b-
2 s 1 -s b
L
 @
@  @
@
rG s11 s22 rL
@
@s
s @
@s 1
s
b1 a2 aL

Figure 4.10: Signal flow graph to calculate bL for a unilateral transistor

s21
bL = s21 a1 + s22 rL bL → bL = a
1 − s22 rL 1
Also the wave amplitude a1 can be expressed by the wave amplitude a0 , which
gives us the final result for bL :
1 1
bL = · s21 · a (4.19)
1 − s11 rG 1 − s22 rL 0
Remembering equation 4.18 we finally get in the case of a unilateral transistor:

1 − |rG |2 1 − |rL |2
GT = · |s 21 |2
· (4.20)
|1 − s11 rG |2 |1 − s22 rL |2
In the unilateral case equation 4.20 can be split up into the product of three inde-
pendent terms:
the generator gain GG ,
1 − |rG |2
GG = (4.21)
|1 − s11 rG |2
the gain due to the active device,

G0 = |s21 |2 (4.22)

57
and the load gain
1 − |rL |2
GL = (4.23)
|1 − s22 rL |2
We will get the maximum transducer gain if the input as well as the output of
the active device is matched. For a unilateral transistor in principle it is easy to
realise since the input reflection coefficient s11 does not depend on the load and
the output reflection coefficient s22 does not depend on the internal resistance of
the generator.
rGm = s∗11 rLm = s∗22
In Fig. 4.11 the principle setup of a single stage rf-amplifier is shown. To each
term of equation 4.20 a 2-port can be assigned. With the help of the input matching

Z G = ZC Z = ZC
s s s s L
 Input Ouput
 active device -
 matching matching
s s s s

s∗11 s∗22

Figure 4.11: Principle setup of a single stage rf-amplifier

network the generator impedance is transformed into s∗11 to match the transistor
input, while the output matching network will transform the load impedance into
s∗22 to match the transistor output. This results in an equation for the so-called
unilateral transducer gain GTu , which only depends on the properties of the active
device.
1 1
GTu = · |s21 |2 · (4.24)
1 − |s11 |2
1 − |s22 |2

4.3.2 Bilateral Transistor


In general the used transistors will always show an internal feedback (s12 ̸= 0)
which results in the signal flow graph shown in Fig. 4.12. In contrast to the
previous section one must account for the dependence of the wave amplitude a1
on the wave amplitude bL
a0 + s12 rL rG rL bL
a1 = a0 + rG (s11 a1 + s12 rG bL ) → a1 =
1 − rG s11
As a result we get:
s21
bL = a (4.25)
(1 − s22 rL )(1 − s11 rG ) − s21 s12 rG rL 0

58
a0 s 1 -s 1 -as1 s21 b-
2 s 1 -s b
L
 @
@  @
@
rG s11 s22 rL
@
@s
s s12 @
@s 1 s
b1 a2 aL

Figure 4.12: Signal flow graph to calculate bL for a bilateral transistor

Using again equation 4.18 we get for the transducer gain in case of internal feed-
back:

|s21 |2
GT = (1 − |rG |2 ) (1 − |rL |2 ) (4.26)
|(1 − s22 rL )(1 − s11 rG ) − s21 s12 rG rL |2

4.3.3 Optimum Input and Output Matching


Due to the internal feedback of the active device the load reflection coefficient rL
will change the input reflection coefficient rin and the reflection coefficient of the
generator rG will have an influence on the output reflection coefficient rout . To
get the maximum transducer gain the input as well as the output must be matched
simultaneously. If we define rGm and rLm to be the reflection coefficient at the gen-
erator and load side to reach matching, the following equations must be fulfilled
simultaneously:

s s s −∆ r
r∗Gm = s11 + 1 −12s 21r rLm = 111− s sr Lm
22 Lm 22 Lm
s12 s21 s22 − ∆s rGm
r∗Lm = s22 + 1 − s r rGm = 1 − s r
11 Gm 22 Gm

Into the last equations the following short cut was introduced ∆s = s11 s22 −s12 s21 .
Using the above equations one in principle is able to find the reflection coefficients
rGm and rLm . This leads for example to a quadratic equation for rLm , which only
gives values |rLm | < 1 if the active device is stable. For illustration Fig. 4.13
shows rGm for the transistor BFR 280 which was stabilised using a parallel resistor
of 121Ω at its output. Also shown is the complex conjugated value of s11 . One
notices a strong deviation between the two curves for frequncy values with a low
stability factor K (compare Fig. 4.8)

59
Figure 4.13: rGm and s∗11 of transistor BFR 280

4.3.4 Maximum Available Power Gain


If the generator reflection coefficient shows the value rGm and the load reflection
coefficient shows the value rLm the power gain will be at its maximum, which is
denoted as maximum available power gain GMAG . For its value we get:

(1 − |rGm |2 ) |s21 |2 (1 − |rLm |2 )


GMAG = (4.27)
|(1 − s11 rGm )(1 − s22 rLm ) − s21 s12 rGm rLm |2

The above equation can be simplified by introducing the stability factor K [2]:

|s21 | √
GMAG = (K − K 2 − 1 ) (4.28)
|s12 |

It is obvious that equation 4.28 will only result in meaningful results for GMAG as
long as the active device is stable (K > 1). On the other hand we learned in the
previous sections how to stabilize an active device using parallel or series resistors
at the input or output. Using equation 4.28 for the case K = 1 thus can gives us a
crude approximation of the achievable power gain after stabilisation. It is termed
maximum stable gain GMSG :

|s21 |
GMSG = (4.29)
|s12 |

60
Figure 4.14: GMSG , GMAG of transistor JS8818A

Fig. 4.14 shows the values of GMSG and GMAG for the MesFet JS8818A in the
frequency range from 1 to 50 GHz. For lower frequencies the transistor is unstable
(K < 1) and only the maximum stable gain GMSG can be calculated. It drops from
approximately 23 dB at 1 GHz to approximately 12 dB at 13 GHz. Above this
frequency the transistor becomes stable and the maximum available gain GMAG
can be calculated. Above roughly 40 GHz the transistor no longer can be used to
realise an amplification.

4.4 Differnt Amplifier Concepts


To realize rf- or microwave amplifiers in hybrid or integrated technique different
concepts are used. The most common are:
• single-ended amplifier
• feedback amplifier
• balanced amplifier
• travelling-wave amplifier
A more detailed description of the different amplifier types can be found in [2].
As a simple example in the following section we want to design the rf-section of
a single-ended amplifier, without discussing the necessary DC-circuit to establish
the point of operation.

61
Single-Ended Amplifier
Its principal setup is shown in Fig. 4.11. It consists of three different parts, the
input matching network, the active device and the output matching network. As
already discussed in the previous section input and output matching network are
used to maximize the generator gain GG and the load gain GL as defined in equa-
tion 4.20. To design the matching networks we can use the technique discussed
in the section of transmission line theory and its application with the help of the
smith-chart. To demonstrate the procedure we will start with the stabilized tran-
sistor BFR 280 discussed in section 4.2.2 and design the matching networks to
operate at a frequency of 500 MHz. At this frequency the transistor is described
by the following set of s-parameters:
{ }
0, 67 exp(− j410 ) 0, 05 exp( j690 )
(S) =
4, 6 exp( j1330 ) 0, 34 exp(− j310 )

With the help of a simulation tool it is posible to calculate the optimum laod re-

Figure 4.15: Constant gain contours in the generator plane. Stabilized transistor
BFR 280, f = 500 MHz, GA = 18 dB, ∆G = 1 dB

flection coefficient as well as the optimum generator reflection coefficient. Fig.

62
4.15 shows contours of constant gain in the generator plane for a frequency of
500 MHz. Also the values for the optimum generator and load reflection coeffi-
cient are displayed.

rGopt ≈ 0, 8 exp(480 ) rBopt ≈ 0, 57 exp(700 )

With the help of the Smith-Chart it is possible to design the input and output
matching networks. One possible realisation example is shown in Fig. 4.16.
Fig. 4.17 shows the computed absolute values of the parameters s21 , s11 and

19nH
s s 
 s
BFR280
50Ω 40nH
s 
 s 121Ω 


 @
R
@ 106nH
 50Ω
 
 
 177nH 


Figure 4.16: Single-ended amplifier using the transistor BFR 280

Figure 4.17: s21 -, s11 - and s22 of single-ended amplifier with BFR 280

s22 using the circuit according to Fig. 4.16. At the frequency of operation the
amplifier will show a gain of approximately 17 dB. The frequency dependence of

63
the input and output matching network is clearly displayed in the behaviour of the
parameters s11 and s22 , which show a narrow band characteristic.

64
Chapter 5

Plane Wave

To find a simple solution for Maxwell’s equation we use the differential form of
Faradays and Amperes law []. The electromagnetic field will be represented using
⃗ All other fields will
the electric field strength ⃗E and the magnetic field strength H.
be deduced from this basic fields. In general the phasor of the electric field will be
described by three independent functions E x , E y , E z which are themselves func-
tions of the three space coordinates x, y, z. To find a simple solution for Maxwell’s
equation we assume the existence of only the x-component of the electric field E x
which only depends on the z-coordinate.
⃗E(z) = E x (z)⃗ex (5.1)
The electrical properties of space will be described by its conductivity κ and its
permittivity ε = εr ε0 , which are assumed to be constant in the considered space.
In this case we will get the following equations for the conduction current density
J⃗C and the electric displacement density field ⃗D:
J⃗C = κE x (z)⃗ex ⃗D = εE x (z)⃗ex (5.2)
The magnetic properties of space are described by the scalar permeability µ = µr µ0
which is also assumed to be constant in the considered space. Hence the following
equation gives the relation between the magnetic flux density ⃗B and the magnetic

field H:
⃗B = µH ⃗

5.1 Wave Equation of the Plane Wave


Since we already defined the direction and space dependence of the electric field
we can use Farady’s law to calculate the associated magnetic flux density field:
∂ ∂E (z)
jω⃗B = −⃗∇ × ⃗E(z) = −(⃗ez ) × (E x (z)⃗ex ) = − x ⃗ey
∂z ∂z

65
According to the equation above the electric field with only one x-component with
a z-dependence in space will result in a magnetic field having only a y-component:
1 ∂E x (z)

H(z) = − ⃗ey (5.3)
jωµ ∂z
After realising that the considered electric field results in a magnetic field with
y-component and z-dependence, we now use the magnetic field and Ampere’s law
to calculate the associated electric field.

J⃗C + jω⃗D = (κ + jωε)⃗E = (⃗ez ) × (H y (z)⃗ey )
∂z

It makes sense to combine the conduction current density J⃗C = κ⃗E and the dis-
placement current density J⃗D = jωε⃗E by the introduction of a complexed valued
permittivity εr :
κ
κ + jωε = jωε0 (εr − j ) = jωε0 εr (1 − j tan(δε ) = jωε
ωε0
Comparing the coefficients we find for the complex valued permittivity ε:
κ
ε = ε− j (5.4)
ω
This results in the follwoing equation for the electric field:
∂H y (z)
⃗E(z) = − 1 ⃗ex (5.5)
jωε ∂z
According to equation 5.5 the y-orientated magnetic field with z-dependence will
result in an electric field with x-component and also a z-dependence. One observes
that both fields excite each other. As a result equation 5.3 can be combined with
equation 5.5 and one ends up in an ordinary differential equation for the electric
field:
1 d 2 E x (z)
E x (z) =
−ω2 εµ dz2
As final result we find the wave equation of the plane wave:

d 2 E x (z)
2
+ ω2 µε E x (z) = 0 (5.6)
dz
Comparing equation 5.6 with the differential equation for a wave on a transmis-
sion line shows a correspondence between the voltage distribution U(z) and the
electric field E x (z), while the inductance per unit length L′ and the capacitance

66
per unit length C′ correspond to the space constants µ and ε. Due to this analogy
one can use the same function to solve the differential equation:

E x (z) = E x0 exp(−γz) = E x0 exp(−αz) exp(− jβz) (5.7)

One can introduce the same complex valued propagation constant γ, the attenua-
tion constant α and the phase propagation constant β. In case of the plane wave
their numerical values depend of the space constants µ and ε.

√ √ κ √ √
γ = jω µε = jω µε 1 − j = jω µε 1 − j tan(δε ) (5.8)
ωε
Due to the square root in equation 5.8 the sign of the complex valued propagation
constant γ can be positive or negative, corresponding to two different waves, one
travelling in the positive z-direction and one travelling in the negative z-direction.
According to equation 5.3 we get for the magnetic field H y (z) in dependence of
the electric field E x (z):

1 d(E x0 exp(−γz)) γ ε
H y (z) = − = E x (z) = E (z)
jωµ dz jωµ µ x

It makes sense in analogy to the characteristic impedance of a transmission line to


introduce the complex valued field impedance Z F .

µ
ZF = (5.9)
ε

Using the field impedance we get the following equation for the magnetic field
H y (z):
1
H y (z) = E (z) = H y0 exp(−γz) (5.10)
ZF x
According to equation 5.8 the complex propagation constant becomes imaginary
in the limit κ → 0 which corresponds to a lossless wave propagation. For a certain
value of conductivity κ the wave propagation will no longer be lossless.

5.2 Weakly damped Plane Wave


A plane wave shows low losses if the displacement current density is dominant
compared to the conduction current density and the following relation holds true:
κ
|J⃗C | ≪ |J⃗D | → ≪ 1
ωε

67
Using this limit we get for the complex valued propagation constant 5.8 the fol-
lowing approximation:
√ κ
γ ≈ jω µε(1 − j )
2ωε
For the phase constant β and the damping constant α we get in this case:

β = ω µε
√ (5.11)
κ
α = 2 ε =
µ tan(δε )
β
2
If one wants to answer the question with which velocity a point of constant phase
ϕ0 is propagating, we can in analogy to transmission line theory formulate the
equation for the phase velocity vPh :
ω 1
vPh = = √ (5.12)
β µε

For the field impedance Z F we get for the case of dominant displacement current
density: √
v µ
u µ ε ZF
ZF = ut κ ≈ 1− j κ = 1− j κ (5.13)
ε(1 − j ) 2ωε 2ωε
ωε

5.2.1 Plane Wave in free Space


In the case of free space the equation describing the plane wave reduces even to a
simpler relations. The phase velocity of the plane wave will be equal to the speed
of light c0 in free space.
1 m
vPh = √ = c0 = 2, 99792458 · 108 (5.14)
µ0 ε0 s

The phase propagation constant β and the damping constant α will reduce to:

β = ω µ0 ε0
(5.15)
α ≡ 0

According to the last equation the plane wave in free space will not be attenuated.
The length ∆z in which the phase is changed by 2π is known as the wavelength λ0
of the plane wave in free space. Using equation 5.15 we find:
2π 1 c0
λ0 = = √ = (5.16)
β f µ0 ε0 f

68
The equation for the field impedance reduces to:

µ0
ZF = = Z0 (5.17)
ε0
In free space the field impedance shows a real value of Z0 ≈ 377 Ω. Fig. 5.1
shows in its lower part the space function of the electric field at t = 0, which can
be deduced from the complex valued equation 5.7 (α = 0). In the upper part the
corresponding field lines of the electric field are shown. Since the electric field
consists only of a x-component the electric field lines are parallel to the x-axis. In
principle they have an extension from x = −∞ to x = ∞.

Figure 5.1: Field lines and space function of the electric field of a plane wave

5.2.2 Plane Wave in a Lossless Material


If one assumes the plane wave to propagate in a lossless material which is de-
scribed by the following constants µr ̸≡ 1, εr ̸≡ 1 and κ ≡ 0, it will not be atten-
uated, but the phase velocity will be reduced compared to free space. Its value is
given by:
1 c0
vPh = √ = √ (5.18)
µε µr εr
Also the wavelength λ in a lossless material is reduced:
2π λ0
λ = = √ (5.19)
β µr εr
For the majority of materials used in the rf-domain we have µr ≡ 1 and εr > 1 as
a result also the field impedance reduces compared to its value in free space:
√ √
µ µr
ZF = = Z0 (5.20)
ε εr

69
From a physical point of view the real value of the field impedance reflects the
fact that the real space and time function of the electric field Ex (z,t) as well as the
magnetic field Hy (z,t) are in phase, as illustrated in Fig. 5.2.

Figure 5.2: Space and time function of the electric and magnetic field of a plane
wave

To develop an idea of how a wave transports power, we will calculated the time
averaged Poynting vector < ⃗S >. We start with the solution given in equation 5.7
for the case without losses (α ≡ 0).
1 ∗ 1
< ⃗S > = Re(⃗E × ⃗H ) = Re[E x0 exp(− jβz)E ∗x0 exp(+ jβz)]⃗ez
2 2ZF
For the time averaged Poynting vector we get:
1 1
< ⃗S > = |E x0 |2 = ZF |H y0 |2 (5.21)
2ZF 2
According to 5.21 the discussed plane wave in a lossless medium will transport
a constant mean power density in the direction of propagation (z-direction). For
illustration Fig. 5.3 shows the electric and magnetic field lines at a certain time,
showing their negative and positive maximum. Easy to recognise are the planes in
three dimensional space, which are reponsible for the name of the wave and which
are also planes of constant phase. Also shown is the instantaneous Poynting vector
which is perpendicular to the planes of constant phase. It is worth to mention that
the Poynting vector always points in the same direction even when the directions
of the field lines change.

Exmaple Power density of the satellite TV SAT 2


The satellite TV SAT 2 generates a power density of approx. −100 dBW/m2 in

70
6
- 
-
6x 6 
⃗E - - - -
6- - - 
- - -
6
- - - - ⃗S 
- - - -
6
- - - - - 
- - - - -
6
- - - - - - 
- - - - - -
- - - - - - - - - -?
- - - - - - - -? -
- - - - - -? z

H - - - -?
- -?

y  λ/2 -?

Figure 5.3: Electric and magnetic field lines in field strength maximum with their
respective present Poynting vector

Germany. Calculate the maximum receiveable power with the help of a parabolic
antenna with a diameter of D = 80 cm.
| < ⃗S > | = 10−10 W/m2 = 0, 1nW/m2 →
Pmax = | < ⃗S > |πD2 /4 = 16 pW =
ˆ −78 dBm

5.2.3 Polarization
To discuss the terms polarization and wave vector of a plane wave we again use
the solution for a plane wave in a lossless medium (γ = jβ, Z F = ZF ). To find a
simple solution of Maxwell’s equation we arbitrarily started with a electric field
in x-direction, which only shows a z-dependence. As a result we found a plane
wave, which propagates in the z-direction and which has a y-orientated magnetic
field also with a z-dependence. As starting point we also could have chosen a y-
orientated electric field with z-dependence. As a possible solution for Maxwell’s
equation we would have found a plane wave, but now with a magnetic field in
the negative x-direction. The values for the other parameters describing the wave
e.g. phase constant β or field impedance ZF do not change. Extending our knowl-
edge we could also use the following more general field distribution which solves
Maxwell’s equation and which describes a plane wave in a more general form:
( )
⃗E(z) = E x0
exp(− jβz) = ⃗E 0 exp(− jβz) (5.22)
E y0
The values E x0 and E y0 defined in equation 5.22 are arbitrarily selectable and
define the complex electric field vector ⃗E 0 at the space coordinate z = 0. Using

71
the law of induction we find for the magnetic field:

⃗H(z) = − 1 (⃗ez ∂ ) × ⃗E(z) = 1 (⃗ez × ⃗E 0 ) exp(− jβz) (5.23)


jωµ ∂z ZF

To discuss the property of polarisation of a plane wave we will follow [7] and
have a closer look at the vector ⃗E 0 . As a complex vector its square will alos be
complex, which can be brought into the following form:

⃗E 20 = |⃗E 20 | exp(−2 jδ) = (⃗E exp(− jδ))2

2 2
The term |⃗E 0 | represents the absolute value of the field strength |⃗E 0 | and is real.
The complex valued field strength ⃗E shows a phase difference of δ with respect
to the original field strength ⃗E 0 but the square of it is also real, because of:

⃗E 2 = |⃗E 2 |
0

Being a complex vector ⃗E it can be decomposed in its real and imaginary part:
⃗E = ⃗Er + j⃗E j

Using its real- and imaginary for forming the square of the vector ⃗E , shows that
the vectors ⃗Er and ⃗E j are perpendicular to each other.

⃗E 2 = (⃗Er + j⃗E j )2 = ⃗E 2 − ⃗E 2 + j2⃗Er ⃗E j = reell → ⃗Er ⃗E j = 0


r j

The equation of an arbitrary plane wave (5.22) may thus be written in the follow-
ing form:

⃗E(z) = (⃗Er + j⃗E j ) exp(− jβz − jδ) with ⃗Er ⃗E j = 0

In the last equation the electric field consists of two field components ⃗Er and ⃗E j ,
which are perpendicular to each other and which show a phase shift of 900 . If a
unit vector ⃗eξ is introduced, which has the same direction as the vector ⃗eξ and a
further unit vector ⃗eζ which is parallel to ⃗E j , the equation of the electric field can
be rewritten as:
⃗E(z) = (Er⃗eξ + jE j⃗eζ ) exp(− jβz − jδ) (5.24)
For the real time functions of the electric field we will now get:
( )
⃗E(z,t) = Re[⃗E(z) exp( jωt)] = Eξ cos(ωt − βz − δ)
(5.25)
Eζ sin(ωt − βz − δ)

72
Equation 5.25 gives the space and time function of the electric field strength in a
general form. To discuss the different forms of polarisation of a plane wave Fig.
5.4 shows some spezial cases. In Fig. 5.4a the direction of the unit vector ⃗eξ coin-
cides with the x-direction of the coordinate system, while the ζ-component does
not exist. In this case it is said the wave shows a linear polarisation in x-direction.
If the directions of the vectors are given as shown in Fig. 5.4b one speaks of a
linear polarised wave in y-direction. If the shown x-direction coincides with the

Figure 5.4: Different polarisations of a plane wave

vertical axis and the y-direction coincides with the horizontal axis, one also speaks
of vertical or horizontal polarisation. The existence of two polarisations which do
not influence each other can be used to double the channel bandwidth in radio link
systems [6]. The most general case of polarisation is shown in Fig. 5.4c. Here
the end of the vector of the electric field strength will move on an ellipse. In that
case one speaks of a wave with elliptical polarisation. In the special case when
the absolute values of ξ and ζ are equal the tip of the vector of the electric field
strength will move on a circle, as shown in Fig. 5.4d. In this case one speaks of a
circular polarised wave.

5.2.4 The Wave Vector


By choosing arbitrarily the electric field to be only a function of the z-direction, as
the solution of Maxwell’s equation we found waves propagating in ±z-direction.
In the free isotropic space any direction can be chosen to be identical with the

73
z-direction. This leads to the following definition of the wave vector ⃗k:

⃗k = β⃗ez = 2π⃗ez (5.26)


λ
To describe the same phase change of the wave, the product βz in the exponential
function must be replaced by the dot product between the wave vector ⃗k and the
space vector⃗r. In this case the electric field of a linear polarised wave propagating
in an arbitrary direction in space can be written as:

⃗E(z) = E 0⃗eE exp(− j⃗k⃗r) with ⃗eE ⊥ ⃗k (5.27)

In the last equation a real unit vector ⃗eE was introduced to denote the direction of
the electric field ⃗E 0 = E 0⃗eE . Since the direction of the magnetic field is perpen-
dicular to the electric field and the direction of propagation one gets the following

 @
 @
 ⃗E@
⃗r  
I
@@ @ ⃗
 @
@ @ @ k

z′ 6  @ @
⃗ @
@ ?H planes

 @ of constant
 @
@ phase

 @
 - y′
@
@
@
R x′
@

Figure 5.5: Plane wave propagating in an arbitrary direction in space

equation for the magnetic field of a plane wave in general:


⃗H(z) = k ×⃗eE E 0 exp(− j⃗k⃗r) (5.28)
|⃗k| ZF

Fig. 5.5 shows a plane wave, which propagates in the coordinate system x′ , y′ , z′
in the direction given by the wave vector ⃗k. For illustration a plane of constant
phase is shown, which is located at a distance of ⃗r · ⃗k/|⃗k| from the origin of the
coordinate system. Within this plane the vectors of the magnetic field H ⃗ and
the electric field ⃗E are located, which are perpendicular to each other and also
perpendicular to the direction of propagation.

74
5.3 Plane Wave in a good Conductor
The properties of a plane wave change drastically if the conduction current density
exceeds the displacement current density and the following relation is fulfilled:
κ
≫ 1
ωε
In this case the equation 5.8 for the complex propagation constant changes to:

γ ≈ jωκµ (5.29)

This results in the following value for the phase and damping constants:

ωκµ
β = α = (5.30)
2
Since the displacement current density is neglected ε is not present in equation

Figure 5.6: Penetration of a plane wave into a conductor

5.30. To further explain the different behaviour we have a look at Fig. 5.6. It
shows a space described by a cartesian coordinate system. In the region z > 0 a
good conductor is present. We assume the existence of the magnetic field H y0 at
the boundary. Since both the magnetic as well as the electric field are described
by the same propagation constant, we find according to equation 5.7:

H y (z) = H y0 exp(−γz)

Fig. 5.6 shows the normalized absolute value of the magnetic field H y (z) as a
function of the normalized space coordinate. The length ∆z resulting in a drop of

75
the magnetic field at the surface |H y0 | by a factor of 1/e is called the skin depth δ
of the plane wave for a given conductor.

1 2
δ = = (5.31)
α ωµκ

Example: Skin depth of a plane wave for copper


Data: κCu = 6 · 107 S/m, µ = µ0 = 4π · 10−7 Vs/Am, f = 100 MHz

2
δ = m ≈ 6, 5 µm
2π 10 6 107 4π10−7
8

With the help of the calculated value we can setup an equation to calculate the
skin depth for copper δCu with a specific set of units:
65 µm
δCu = √
f
MHz
The last equation essentially shows that an rf or microwave field cannot penetrate
into a good conductor. This effect is called skin effect, because all fields will only
exist next to the surface of the conductor.
The complex valued field impedance Z F of a conductor posses real and imaginary
parts with the same value, as does the complex valued propagation constant γ.
Starting from equation 5.9 we get:
v
u µ √
u
t ε 1 + j µω
ZF = κ = √2
− j ωε κ
(5.32)

Example: Field impedance of copper for f = 100 MHz


1+ j
ZF = = (1 + j)0, 26 mΩ
6 · 107 S/m 65µm
With the help of the field impedance Z F we get for the electric field inside the
conductor: √
µω
E x (z) = (1 + j) H exp(−γz) (5.33)
2κ y0
The electric field inside the conductor will according to 5.2 result in a conduction
current density J x (z). Its absolute value will also exponentially decay:
J x (z) = κE x0 exp(−αz) exp(− jβz) (5.34)
Thus Fig. 5.6 also shows its principle dependence.

76
5.3.1 Ideal Conductor
An ideal conductor is described in the limit κ → ∞. According to equation 5.33
the tangential electric will no longer exist on the surface as well as inside the
conductor. This result can be extended to the ideal conductor with an arbitrarily
shaped surface. If ⃗n is equal to the surface normal in a considered point on the
surface, the electric field on the surface of an ideal conductor must fulfill the
following boundary condition:

⃗n × ⃗E = 0 (5.35)

The behaviour of the current density will be more complicated. Using equation
5.34 it can be deduced using the magnetic field at the surface:

µωκ 1+ j
J x (z) = (1 + j) H y0 exp(− z)
2 δ

At the one end its absolute value will tend to infinite as κ and on the other
√ hand
the thickness of the current carrying layer will decay according to 1/ κ due to
the skin depth δ. It is of advantage to introduce a so-called surface current density
J Sx with the unit A/m, which can be defined by a one dimensional integration of
the current density J x :
∫ ∞
J Sx = J x (z) dz = H y0
0

The surface current density thus shows the same value like the magnetic field,
but its direction is perpendicular to the magnetic field. Again using ⃗n to define a
normal to the surface, we get the following relation between the surface current
density ⃗J S and the magnetic field:

⃗n × ⃗H = ⃗J S (5.36)

77
Chapter 6

Antennas

Antennas are the obvious components of our modern telecommunication infras-


tructure. Even today the skyline of our cities are dominated by Yagi-Antennas for
receiving the different terrestrial distributed TV programs. Today they are going
to be replaced by parabolic antennas, capable for receiving programs distributed
by satellites. Beside these pure receiving antennas transmitting and receiving an-
tennas are installed to support the mobile telecommunication networks. Despite
the differences, antennas always have to fulfill similar general tasks in a commu-
nication system. The transmitting antenna has to transfer the guided TEM wave
to a plane wave in free space, while the receiving antenna will transfer power
transported by a plane wave into a guided TEM wave. Thus antennas are also
sometimes denoted as wave type converter [8]. Since the spatial extension of an-
tennas are always in the order of the wave length, their mathematical treatment is
complicated. To calculate their field distribution Maxwell’s equations have to be
solved numerical. To find a simple description of a transceiver system we will use
a phenomenological model.

6.1 Transmitter Receiver System


In this section we will setup a phenomenological model of a transmitter receiver
system. Starting point of the considerations will be the phenomenon that a trans-
mitting antenna will transmit an electromagnetic wave carrying the power P1 . The
simplest possibility to describe such a phenomenon is to assume that the power
will be transmitted homogeneously into a sphere surrounding the antenna e. g.
the antenna is assumed to be an isotropic point radiator. This concept is illustrated
in Fig. 6.1a. Assuming a lossless medium the radiated power of the point source
will be homogeneously distributed over the surface of a sphere with raduis r. The
assumption leads to the following equation describing the radiated power density

78
Figure 6.1: Radiated power density of an isotropic point radiator and an antenna
described by the gain function G(φ, ϑ)

⃗S(r) at a distance r from the point source.

⃗S(r) = P1
⃗er (6.1)
4πr2
We setup the radiation model described by equation 6.1 without any reference to
Maxwell’s equations. So it is not suprising that a real antenna for electromagnetic
radiation does not behave as decribed by equation 6.1. Nevertheless it is a usefull
model and it is denoted as isotropic point radiator. As illustrated in Fig. 6.1b a
real antenna will always focus the radiation to a certain space region. It can be
described by introducing the so-called gain function G(φ, ϑ) of the transmitting
antenna with reference to the isotropic point radiator. Using the gain function one
can now modify the radiated power density to be transmitted in the space direction
described by (φ, ϑ):
⃗S(r) = G1 (φ, ϑ) P1 ⃗er (6.2)
4πr2
The product of transmitted power P1 and gain in main beam direction G1 is called
Effective Isotropic Radiated Power (EIRP). Its value is equal to the power to be
transmitted by an isotropic point radiator to result in the same power density in
main beam direction.
EIRP = G1 P1 (6.3)
To describe a receiving antenna we will have a look at the parabolic antenna shown
in Fig. 6.2, which is assumed to be far away from the transmitting antenna. In
this case the elctromagnetic radiation will exist as a plane wave having a power
density S2 . The parabolic antenna will focus the power density and result in the
power P2 at the anteanna ouput. This behaviour can be described by assigning an

79
Figure 6.2: Operation principle of a parabolic antenna

effective receiving area Aw2 to the receiving antenna.


P2 = Aw2 S2 (6.4)
According to Fig. 6.2 the effective receiving area of the parabolic antenna will
be in the order of its cross-sectional area πD2 /4. Using equation 6.2 and 6.4 it is
easy to describe the properties of a Transmitter-Receiver-System.
Aw2 G1
P2 = P1 (6.5)
4πr2
Of course the transmitting antenna can be used as receiving antenna and vice
versa. This will lead to the following equation:
Aw1 G2 ′
P1′ = P2
4πr2
If we assume in both cases to be the same transmitting power P1 ≡ P2′ due to the
reciprocity law [3], we must have the same value of the received power. This leads
to:
Aw1 G2 Aw2 G1 Aw1 Aw2
2
= 2
→ = = const.
4πr 4πr G1 G2
According to the last equation the ratio between effective area and gain of an
antenna in direction of the main beam shows a constant value. According to
Maswell’s theory this ratio is given by:
Aw λ20
= (6.6)
G 4π

80
This results in the following relation for the received power of a transmitter-
receiver-system:
Aw2 Aw1 G2 G1
P2 = P1 = r P1 (6.7)
λ0 r
2 2
(4π )2
λ0
Example: Received power of a satellite receiver station
The satellite DFS-Kopernikus transmits at a frequency of 12 GHz with an EIRP
of 56 dBW with its main beam to Europe. (Diameter of the receiving antenna
D = 80 cm)
EIRP
56 = 10 log( ) → EIRP = 1 W · 105,6 = 398 kW
1W
For the power density S2 at the receiver side we get:
398kW pW
S2 = 2
= 24, 4 2
4π(36000km) m
which yields for the received power:

24, 4pW/m2
P2 = = 12, 3pW
π(0, 4m)2

6.1.1 Transmitter-Receiver-System as Quadrupole


In the previous section we concentrated on the descrition of the power transfer
between a transmitter and the receiver. But for designing the circuits matching
between transmitter and transmitting antenna and receiver and receiving antenna
must be discussed and will have a significant influence on the overall system. A
simple model to discuss this aspects is shown in Fig. 6.3. It shows the a tramitting
and receiving antenna which have a mutual distance of r with optimal alignment
considering their main beams. The terminla pairs of the transmitter and receiver
antenna can be considered to form a quarupole. If the quadrapole is described
using Z-parameter we have:

U 1 = Z 11 I 1 − Z 12 I 2
(6.8)
U 2 = Z 21 I 1 − Z 22 I 2

The following denotation can be used for the signle parameters:


Z 11 , Z 22 - Open circuit input impedances of the antennas
Z 21 , Z 12 - Coupling impedances between the antennas
The values of the impedances Z 11 and Z 22 depend on the size and geometrie of
the transmitting and receiving antenna, while for the impedances Z 21 and Z 12 the

81
A A

I1 - s  r - s I-
2

U1 U
?
s ?
s 2

Figure 6.3: Qudrupole of a Transmitter-Reseiver-System

mutual distance r and the mutual orientation play a siginificant role. Due to the
reciprocity principle and the introduced posistive directions of current we have:

Z 21 = −Z 12 (6.9)

Equation 6.9 guarantees that the influence the current I 1 on the antenna 2 is equal
to the influence of the current I 2 on the antenna 1.
Since the mutual distance between the transmitting and receiving antenna will
always be large in a real system the quations will reduce to:
U 1 = Z 11 I 1 (a)
(6.10)
U 2 = Z 21 I 1 −Z 22 I 2 (b)
If we think of equation 6.10b to be the result of a mesh equation we can find the
following equivalent circuit for the receiving antenna. The receiving antenna will

Z 22
I
s 2-


U 20 U2

?
?
s

Figure 6.4: Equivalent circuit of the receiving antenna

show the open circuit voltage U 20 = Z 21 I 1 and the internal impedance Z 22 . To


define a realtion to field at the location of the receiver one introduces the effective
hight he f f 2 of the receiving antenna by the quotient of the open circuit voltage to
the electrical field strength at the receiving antenna E 2 .
|U 20 |
he f f 2 = (6.11)
|E 2 |

82
As we will see in the next section the effective height of a Hertzian Dipol is equal
to its physical length. In the case the receiving antenna is terminated with the
impedance Z ∗22 , we will get for the received power P2 :
( )
|U 02 | 2
1 2 |E 2 |2 h2e f f 2
P2 = = = Aw2 S2 (6.12)
2 Re(Z 22 ) 8Re(Z 22 )
With the help of equation 6.12 a relation between the effective height of the re-
ceiving antenna and its effective area can be deduced:

Re(Z 22 )
he f f 2 = 2 Aw2 (6.13)
Z0

While the term effective height of antenna in the case of a rod antenna is motivi-
ated by its geometrie, it is not quite clear how to understand its effective area. For

Figure 6.5: Effective area of a rod antenna above ground

illustration Fig. 6.5 according to [9] the time averaged flow lines of the poynting
vector in the vicinty of a rod antenna. The current distribution of the receiving
antenna results in a distortion of the flow lines of the poynting vector. All the
power density entering hatched area will contribute to the received power of the
antenna, hence it is equal to the effective area of the rod antenna.

6.2 The Hertzian Dipole


In a mathematical sense a Hertzian Dipole is defined by a current element with
the amplitude I and a length ∆l in the limit ∆l → 0 while the product I∆l is keeped
constant. Fig. 6.6a shows a current element of finite length with orientation in
the z-direction of a coordinate system. Fig. 6.6b shows its current distribution

83
6z 6z z
6
δ ``` I u
6 y |I(z)| |σ(z)| 6
- ∆l - - j

∆l

u


x
a) b) c) d)

6z
@
I 6 @ I(z)
j @ -

e) f)

Figure 6.6: Current and charge distribution of an ideal Hertzian dipole and a short
linear antenna

in z-direction, which is constant. According to the continuity equation [7] the


assioated charge distribution σ(z) can by differentiation of the current distribu-
tion with respect to the z-coordinate. In the case of the ideal Hertzian dipole this
results in infinte high charge densities σ(z) at z = ±∆l/2 (point charges). This
becomes more obvious if we consider the current distribution shown in Fig.6.6b
and the associated charge distribution (6.6c) in the limit δ → ∞. A pratical reali-
sation of a Hertzian dipole is shown in Fig. 6.6d under the assumption ∆l ≪ λ0 .
It consists of a conductor of length ∆l with little spheres loacted at both ends to
carry the charges and which is driven by a current source I in its center. If only a
small wire is driven by a current source (Fig. 6.6e) we will get approximately the
current distribution shown in Fig. 6.6f. In this case the conduction current will be
smoothly converted into a displacement current.

To find a mathematical representation of a Hertzian dipole we consider the


current element I of length ∆l shown in Fig. 6.7. According to [7] this current el-
ement will result in z-orientated vector potentail function ⃗A decreasing reciprocal
to the distance in the limit ω → 0. In the case of a time varrying current the vector
poetntail function must be multiplied by the term exp(− jβr) to account for the
phase delay βr a change of current is noticed at a certain distance r. As a result
we get for the vector potentail function of a Hertzian dipole at a distance r:

⃗A(r) = I∆l exp(− jβr)⃗ez (6.14)


4πr

84
6z

*

⃗r 


ϑ 


X -y
X
6 XXX
φ XXX
XXX
X

x
Figure 6.7: Hertzian dipole in the origin of a coordinate system

According to [7] the magnetic field is given as the cross product of the operator ⃗∇
and the vector potential. To do this it is of advantage to use the representaion of
the operator in a spherical coordinate system [10].

⃗H = ⃗∇ × Az (r)⃗ez = −⃗ez × (⃗∇Az ) = −⃗ez ×⃗er [ ∂ Az (r)]


∂r
Applying the identity ⃗ez ×⃗er = sin ϑ⃗eφ and carrying out the differentiation yields
for the magnetic field:
( )
⃗H = I∆l sin ϑ 1 + j β exp(− jβr) sin ϑ⃗eφ = H φ (r, ϑ)⃗eφ (6.15)
4π r2 r
Analog to a current carrying wire in z-direction also the z-directed current element
shows only a φ-component of the magnetic field. Its magnitude depence on the
distance r to the current element and the angle ϑ of the spherical coordinate sys-
tem. Hence the magnetic fieldlines form circle in planes of constant z-coordinate,
while their centre lie on the z-axis. Using Maxwells equation the electric field can
be deduced form the magnetic field:
1 ⃗ ⃗ 1
⃗E = ∇×H = {−⃗eφ × [⃗∇H φ (r, ϑ)] + H φ (r, ϑ)[⃗∇ ×⃗eφ ]}
jωε jωε
Carrying out the differentiation, one notices that the eletric field consists of a r-
component as well as of a ϑ-component
( )
I∆l 1 β jβ2
E ϑ (r, ϑ) = + + exp(− jβr) sin ϑ (6.16)
4ωε0 π jr3 r2 r
( )
I∆l 2 2β
E r (r, ϑ) = + exp(− jβr) cos ϑ (6.17)
4ωε0 π jr3 r2
Hence the electric field forms lines in planes of constant φ in the spherical coor-
dinate system. The planes are also called E-planes.

85
6.2.1 The Far Field
According to equation 6.15, 6.16 and 6.17 all field components of the Hertzian
Dipole show a strong dependence on r, the distance from the dipole. If one is in-
teressted in the time averaged power a Hertzian dipole radiates, one must examine
the fields for big distances from the dipole. In this limit the equations will reduce
to:
1 β λ0
2
≪ → r≫ (6.18)
r r 2π
One denotes the fields in the region r ≤ λ0 /(2π) as near field. Especially the
elctric field can show here very high values. But for the far field only components
decaying as 1/r are of interest.
I∆l
H φ (r, ϑ) = j sin ϑ exp(− jβr)
2λ0 r
I∆l
E ϑ (r, ϑ) = jZ0 sin ϑ exp(− jβr) (6.19)
2λ0 r
E r (r, ϑ) = 0
According to equation 6.19 the Hertzial dipole possess only a H φ and a E ϑ com-
ponent. Both fields are perpendicular to each other while the absolute value of
their ratio is given by the field impedance Z0 of the plane wave in free space.
Eϑ β
= = Z0 (6.20)
Hφ ωεo
For the time averaged power density of the Hertzian dipol we get:
( )
⃗S = 1 Re ⃗E × ⃗H ∗ = Z0 |H φ (r, ϑ)|2⃗er
2 2
It becomes obvious that the poynting vector in the far field of the dipole possess
only one radial component Sr .
Z0 |I|2 ∆l 2 2
Sr (r, ϑ) = sin ϑ (6.21)
2 4λ20 r2
Using an integration over a sphere in the limit r → ∞ one can calculate the total
power Ps radiated by the dipole.
∫ π ( )
2π ∆l 2 |I|2
Ps = 2π Sr (r, ϑ)r sin ϑdϑ =
2
Z0 (6.22)
0 3 λ 2
The last equation describes the power radiated by the dipol current. So the term in
front of the current must have the unit of a resistance. It is denoted as the radiation
resistor Rs of the hertzian dipole.
( )2
2π ∆l
Rs = Z0 (6.23)
3 λ

86
Gain and Radiation Pattern
With the help of the equations 6.21 and 6.22 we are able to calculate the radiated
power density Sr as a function of the radiadted total power of the Hertzian dipole.
3 2 Ps
Sr = sin ϑ (6.24)
2 4πr2
If we compare equation 6.24 with equation 6.2 we are able to define the gain
GHz (φ, ϑ) of the Hertzian dipole with respected to the isotropic point radiator.

3 2
GHz = sin ϑ (6.25)
2
Additional to the gain also the radiation pattern C(φ, ϑ) is defined. For a transmit-
ting antenna the radiation pattern gives essentially the field amplitudes in the far
field of the antenna. In the case of a receiving antenna it gives essentially the open
circuit voltage of the antenna as a function of the receiing direction. Due to theo-
rem of reciprocity the radiation pattern of the transmitting and receiving antenna
are the same. In general the radiation pattern is given in a normalised form. Here

Figure 6.8: Gain and radiation pattern of a Hertzian dipole

the far field strength is given normalised to the highest occuring value. In the case
of the Hertzian dipole we the maximum is found in the plane ϑ = π/2.

|E ϑ (r → ∞, ϑ)| |H φ (r → ∞, ϑ)|
C(φ, ϑ) = = = | sin ϑ| (6.26)
|E ϑ (r → ∞, π/2)| |H φ (r → ∞, π/2)|

Fig. 6.8 shows the gain and the radiation pattern of the Herzian dipole in a E-
plane. According to this Fig. shows that the simplest radiating element of the
electromagnetic field does not behave like a isotropic point radiator, which does
not exits in the case of electromagnetic radiation.

87
Effective Height and Area
To discuss the effective height of a Hertzian dipole consider Fig. 6.9. It shows the
undistored electric field lines of the field E e which can be ecited by a transmit-
ting antenna at the location of a receiver. If one places the two metallic halfs of
a short dipole into the field we will endup in the distored field distribution shown
approximately in Fig. 6.9b. The original field will result in a open circuit voltage

Figure 6.9: Short dipole as receiving antenna

between the halfs of the dipole. As a result the effective hight of the Hertzian
dipol is equal to its length ∆l. Using equation 6.13 and equation 6.23 allows us to
caluclate the effective area of the Hertzian dipole.

Z0 ∆l 2 3 2
Aw = = λ (6.27)
4Rs 8π 0

88
Bibliography

[1] Zinke, O.; Brunswig, H.:Hochfrequenztechnik 1 u. 2, Springer-Verlag, 1995

[2] Pozar, M. D.: Microwave Engineering, John Wiley & Sons, 2011.

[3] Collin, E. R.: Foundations for Microwave Engineering, McGraw Hill, 1992.

[4] Giacoletto, L. J.: Study of p-n-p allowy junction transistors from DC trough
medium frequencies, RCA Rev, 1954.

[5] Zeidler, E.: Teubner-Taschenbuch der Mathematik, B. G. Teubner Verlag,


1996.

[6] Meinke, H.; Gundlach, F. W.: Taschenbuch der Hochfrequenztechnik,


Springer-Verlag, 1986.

[7] Landau, L.D; Lifschitz, E.M.: Klassische Feldtheorie, Akademie-Verlag,


1981.

[8] Ein neuer Weg zur Loesung des Problems der Breitbandantenne. Nachricht-
entechnische Zeitschrift, Heft 12, 1957, 594-601.

[9] Landstorfer, F. et al.: Energiestroemung in elektromagnetischen Wellen-


feldern. Nachrichtentechnische Zeitschrift, Heft 5, 1972, 225-231.

[10] Bourne, D. E.; Kendall, P. C.: Vektoranalysis, B. G. Teubner, 1973.

89

You might also like