You are on page 1of 32

WIRELESS AND CELLULAR COMMUNICATION

• Understand the concepts of propagation over wireless channels from a physics


standpoint
Mobile Radio Propagation –
Large Scale Path Loss - Free Space Propagation Model, Relating Power to Electric Field,
Three Basic Propagation Mechanisms – Reflection (Ground Reflection) , Diffraction,
Scattering, Practical Link Budget, ( Text 1 - 2.2 and Ref1 - Chapter 4).
Fading and Multipath – Broadband wireless channel, Delay Spread and Coherence
Bandwidth, Doppler Spread and Coherence Time, Angular spread and Coherence Distance
(Text 1 – 2.4) ,
Statistical Channel Model of a Broadband Fading Channel
(Text 1 – 2.5.1)
The Cellular Concept – Cellular Concept , Analysis of Cellular Systems, Sectoring (Text 1-
2.3)

Mobile Radio Propagation


– Large Scale Path Loss

1. Radio Wave Propagation Fundamentals

The radio frequency channel places fundamental limitations on the


performance of wireless communication systems. The transmission path
between the transmitter and receiver can vary from simple line-of-sight to
one that is severely obstructed by buildings, mountains, and foliage. Unlike
wired channels, that are stationary and predictable, radio channels are
extremely random and do not offer simple analysis. Even the speed of
motion impacts how rapidly the signal level fades as a mobile terminal
moves in space. Modeling the radio channel has historically been one of
the most difficult parts of wireless communication system design, and is
often done statistically, based on measurements made specifically for an
intended communication system or spectrum allocation.

1.1 Introduction to Radio Wave Propagation


The mechanisms of electromagnetic wave propagation are diverse, but can
generally be attributed to reflection, diffraction, and scattering. For
example, in urban areas where there is no direct line-of-sight (LOS) path
between transmitter and receiver, the presence of buildings causes severe
diffraction loss. Due to multiple reflections from various objects, the
electromagnetic waves travel along different paths of varying lengths. The
interaction between these waves causes multipath fading at specific
locations, and the signal strengths decrease as the distance between
transmitter and receiver increases. A discussion of wave propagation
models, more detailed than what follows, can be found in [Geng and
Wiesbeck, 1998].

Why it is important to understand propagation

The radio spectrum has a large range of applications in many operational


environments. The frequency bands cover many orders of magnitude in
frequency. Radio systems require transmitting sources and receivers to
provide wireless communications links. Understanding the radio channel
between the transmitter and the receiver is critical in designing any radio
system.

Service providers for example broadcasters, private radio users like taxi
firms, mobile phone network operators, the MOD and a host of others all
want to know how to get radio coverage for their particular application.
When planning a service they need to know where to put their masts, how
many masts will be needed, what antennas should be used, how much
transmitter power is going to be needed and how reliable their radio links
will be.

Regulators who manage the radio spectrum are interested in making the
best use of this limited and valuable resource. They need to regulate so
that as many as possible can share the radio spectrum and so are interested
in predicting interference between users. Where do signals go beyond
where they are intended to. Also what is this or that bit of spectrum worth
and how much benefit can the taxpayer gain from selling the rights?

It is frequently forgotten how important propagation is to the overall


performance of a communications network. Radiowave propagation
studies allow us to estimate and evaluate the radio channel and so design
systems that work as well as possible. Some people, especially the
regulators may prefer propagation to be simple, but in practice
propagation is fairly complicated, so we will start with the
fundamentals….
2.1 Large Scale Path Loss
Propagation models have traditionally focused on predicting the average
signal strength at a given distance from the transmitter, as well as the
variability of the signal strength in close spatial proximity to a particular
location. Propagation models can be divided as follows:

• large-scale propagation models predict the mean signal strength for


large T-R

(i.e. transmit-receive) separation distances (several hundreds or


thousands of meters). They are useful in estimating the radio coverage
area of a transmitter, since they characterize signal strength. small-scale
or fading models characterize the rapid fluctuations of the received signal
over very short travel distances (a few wavelengths). As a receiver (e.g.,
mobile station) moves over very small distances, the instantaneous
received signal may fluctuate rapidly giving rise to small-scale fading. The
reason is that the received signal is a sum of many contributions coming
from different directions (multipath).

Figure 1.1: Small-scale fading and large-scale fading.

Since the phases are random, the sum of the contributions varies widely.
In small-scale fading (e.g., Rayleigh fading), the received signal power may
vary by as much as three or four orders of magnitudes (e.g., up to 30–
40dB) when the receiver is moved by only a fraction of a wavelength. As
the receiver moves away from the transmitter over much larger distances,
the local average received signal will gradually decrease (and increase in
some parts), and it is this local average signal level that is predicted by
large-scale propagation models. Typically, the local average is computed
by averaging signal measurements over a measurement track of 5λ to 40λ.
Fig. 1.1 illustrates small-scale fading and the large-scale variations. This
chapter covers large-scale propagation and presents some common
modeling techniques used to predict the received power in radio
communication systems. Chapter 3 describes small-scale multipath fading.
Take care: Small scale fading is not a synonym for fast fading, and large
scale fading is not a synonym for slow fading! In chapter 3 the differences
will be made clear.

1.2 Free Space Propagation Model


The free-space propagation model is used to predict the received signal strength when except
for a line-of-sight path no other signal contributions have to be considered. This is the e.g. the
case for communication and microwave line-of-sight radio links. The freespace model predicts
that the received power decays as a function of T-R separation raised to the power of two.
The free-space power received by a receiving antenna which is separated from a transmitting
antenna by a distance d, is given by the Friis free-space equation [Geng and Wiesbeck, 1998]

(2.1a)

(2.1b)
where PT is the transmitted power, PR is the received power (assuming a polarization matched
receiving antenna, optimum orientation of transmitting and receiving antennas, and conjugate
complex impedance matching of the receiver), GT is the gain of the transmitting antenna, GR
is the receiving antenna gain, d is the T-R separation, and λ is the wavelength. The available
received power is given by the product of the effective aperture AeR = GR · λ2/4π of the
receiving antenna and the power density SR at the receiver location. The Friis free-space
equation (2.1) shows that the received power falls off as the square of the T-R separation
distance. This implies that the received power decays by 20dB/decade (concerning the
distance).
An isotropic radiator is an idealized antenna which radiates power with unit gain uniformly in
all directions and is often used to reference antenna gains. The effective isotropic radiated
power (EIRP) is defined as
EIRP = PT GT (2.2)
and represents the transmitted power necessary at the input of an idealized isotropic antenna,
so that the far-field power density is equal to that radiated by the real transmitting antenna in
the direction of maximum gain.
The path loss, which represents signal attenuation as a positive quantity measured in dB, is
defined as the difference (in dB) between the transmitted power and the received power. It
may, but does not have to, include the effect of the antenna gains. The path loss for the free-
space model when antenna gains are included is given by

. (2.3)
Drill Problem 14 Assume that a GSM900 (f = 900 MHz) link budget allows a maximum path
loss of 143 dB. Which is the corresponding maximum cell radius for undisturbed free space
propagation assuming isotropic antennas?
By

1.3 Relating Power to Electric Field and Voltage


In section 2.2 the available received power has been derived from the principle of power
conservation. In practice, often the field strength at the receiver position is needed instead. It
can be proven that any radiating structure produces electric and magnetic fields, where most Commented [sg1]: Relating Power to Electric Field,
often only the electric field strength is utilized. For a single plane wave in free space, the
complex electric and magnetic field strength phasors E and H, respectively, are related to the
power flux density S by [Balanis, 1989, Geng and Wiesbeck, 1998].

(2.4)
where η0 denotes the free-space intrinsic impedance (i.e. η0 = (µ0/ǫ0)1/2 = 120πΩ ≈ 377Ω)
and the subscript” eff” characterizes effective field strengths (i.e. RMS values) to distinguish
between the RMS values and magnitudes of the field strength phasors.
Assuming only a single signal at the receiver, i.e. considering only one quasi-plane wave
incident on the receiving antenna, the received power PR (again assuming optimum orientation,
polarization and a conjugate complex impedance match of the receiver to the receiving
antenna) is related to the electric field strength at the receiver location by [Balanis, 1982,
Geng and Wiesbeck, 1998]

(2.5a)
(2.5b)
. (2.5c)

Drill Problem 15 Show how to derive the (2.5b) from (2.5a).


Often, it is useful to relate the (maximum available) received power PR to the open circuit
voltage at the receiving antenna port. If the receiver input impedance RR +jXR is
conjugate complex to the receiving antenna impedance , i.e. , then the receiving
antenna will induce a voltage into the receiver which is half of the open-circuit voltage at the
antenna output (conjugate complex impedance matching). Thus, the received power is
maximum and given by

. (2.6)
Through the equations (2.5)– (2.6) it is possible to relate the (maximum available) received
power to the electric field strength at the receiver location (assuming a single incident quasi-
plane wave) or the RMS open-circuit voltage at the receiving antenna terminal.
Drill Problem 16 Show, using the Thévenin equivalent antenna circuit, that the received power
PR is related to the open circuit voltage VR at the receiving antenna port according to (2.6).

2. Three Basic Propagation Mechanisms


Reflection, diffraction and scattering are the basic propagation mechanisms which might have
an impact on propagation in terrestrial radio communication systems operating in the 30 MHz
to 10 GHz frequency range (most mobile communication systems). Propagation mechanisms
affecting systems at low frequencies (e.g., ground wave propagation, ionospheric reflection) Commented [sg2]: Three Basic Propagation Mechanisms
and high frequencies (e.g., gaseous absorption, rain attenuation, troposcatter) are not discussed – Reflection (Ground Reflection) , Diffraction, Scattering,
here. Details can be found in the literature [Giger, 1991]. Reflection, diffraction, and scattering
however, are briefly explained in the next subsections.

2.1 Reflection (Ground Reflection)


When a radio wave propagating through a medium encounters a boundary layer which is plane
and large compared to the wavelength the wave is partially reflected and partially transmitted.
If the second medium is perfectly electric conducting (PEC), then all incident energy is
reflected back into the first medium without any loss of energy. The electric field strength of
the reflected and transmitted waves may be related to the incident wave through the Fresnel
reflection and transmission coefficients Rk,⊥ and Tk,⊥ respectively. The Fresnel coefficients
depend on the material properties, wave polarization, angle of incidence, and material
parameters which are frequency dependent. A polarized electromagnetic wave may be
mathematically represented as a sum of two orthogonal components, such as vertical and
horizontal, or left-hand and right-hand circularly polarized components. For an arbitrary
polarization, linear superposition may be used to compute the reflected fields from a surface.
Fig. 2.2 shows an electromagnetic wave incident at an angle θi relative to the normal on the
interface between medium 1 and 2. Part of the energy is reflected back to medium 1 at an angle
θr=θi (law of reflection), and part of the energy is transmitted (refracted) into the second
medium at an angle satisfying Snell’s law of refraction [Geng and Wiesbeck, 1998]:

(2.7)
where is the relative complex permittivity, including also losses due to a
non-vanishing conductivity σ, and is the complex permeability of the medium for
most cases in radio communications).

(a) parallel polarization (b) perpendicular polarization


Figure 2.2: Geometry for reflection and transmission of a plane wave incident upon a plane
interface between two media 1 and 2 of different material properties.
The nature of reflection varies with the polarization of the electric field. In Fig. 2.2a, the electric
field vector is parallel to the plane of incidence (i.e. the plane containing incident, reflected,
and transmitted rays), and in Fig. 2.2b, the electric field vector is perpendicular to the plane of
incidence. Using superposition, only these two orthogonal and linear independent polarizations
need be considered to solve a general reflection problem. Considering an interface between air
(i.e. and a non-magnetic (but possibly lossy) half space 2 (i.e. ), the
Fresnel reflection/transmission coefficients for the electric field parallel (denoted as k) and
perpendicular (denoted as ⊥) to the plane of incidence are given by [Geng and Wiesbeck,
1998]:

and

and , (2.8b)

where is the wave impedance of the propagation medium. For the reflection
between air and the earth surface, typical values for the soil permittivity and the conductivity
are ǫr2 = 3...25 and σ2 = 10−4S/m...0.1S/m, respectively. Keep in mind that these values are
dependent on the frequency.
Drill Problem 17 Automotive rain sensors detect water on the windshield of a vehicle. The
infrared beam is radiated at a certain angle (θc) from inside of the car on the glass. If the
windshield is dry, all the energy comes back to the sensor; in case of water presence on the
glass surface, less energy comes back and the wipers are activated.
What is the name of the phenomenon, when no energy is transfered form one dielectric medium
to the other? Calculate the value of the angle θc if the permittivity (assume no losses) of the
windshield glass is ǫw = 7.6.
Fig. 2.3 shows plots of the reflection and transmission coefficients (here magnitude only) for
both parallel and perpendicular E-field polarization (e.g. vertical and horizontal, if the interface
is parallel to the xy-plane) as a function of the incident angle for the special case when a wave
propagates in free space ( =1) and the reflecting soil half space is characterized by
=4,36-j3 and =1. For parallel polarization, there exists usually an angle for which the
reflection coefficient shows a local minimum. In case of a lossless medium 2, this minimum
reduces to zero, and the corresponding incident angle is called Brewster angle [Balanis, 1989].
Therefore, the Brewster angle is the angle at which in medium 1 no reflection occurs. For non-
magnetic (generally if µr1 = µr2) and lossless half spaces 1 and 2, the Brewster angle only
exists for parallel polarization and is given by

. (2.9)
Drill Problem 18 A parallel polarized electromagnetic wave radiated from a submerged
submarine impinges on a planar water-air interface. The dielectric constant of water is ǫr = 81
at the frequency of the impinging wave. Assume that the impinging wave is a plane wave at
the interface. Determine the angle of incidence to allow complete transmission of energy.
In most practical cases of reflection and transmission, the horizontal and vertical axes of the
spatial coordinates do not coincide with the perpendicular and parallel axes of the propagating
waves. In addition, the interface between medium 1 and 2 is in general not perpendicular to the
z-axis of a global coordinate system (e.g., reflection from a sidewall of a building). Therefore,
the plane of incidence, field components and angles have to be transformed first (according to
the true geometry) before utilizing the Fresnel coefficients in (2.8) [Geng and Wiesbeck, 1998].

1
reflection parallel
0.9 reflection perpendicular

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 10 20 30 40 50 60 70 80 90
incident angle [deg]

Figure 2.3: Magnitude of reflection and transmission coefficients as a function of the incident
angle for an interface between air ( =1) and a non-magnetic
(but lossy) medium characterized by a complex relative permittivity of and =4,36-j3 and a
relative permeability of
Since there can’t be an electric field in a perfect conductor, all energy is reflected when a ray
encounters a PEC i.e. |R| = 1. As the electric field tangential to the surface must vanish in order
to satisfy Maxwell’s equations, the reflection coefficients for parallel and perpendicular
polarization must be Rk = +1 and R⊥ = −1 respectively, regardless of the incident angle. The
± signs here are valid for the reference directions defined in Fig. 2.2. Commented [sg3]: Practical Link Budget, ( Text 1 - 2.2

2.2 Diffraction
Diffraction occurs when the radio path between transmitter and receiver is obstructed (e.g.,
buildings, hills etc.). The secondary waves resulting from the irregularities are present
throughout the space and even behind the obstacle, giving rise to a bending of waves around
the obstacle, even when a line-of-sight path does not exist between transmitter and receiver.
Although the received field strength decreases rapidly as a receiver moves deeper into the
shadow (obstructed) region, the diffraction field still exists and often has sufficient strength to
produce a useful (or in other cases an unwanted) signal. The phenomenon of diffraction can be
explained by Huygen’s principle, which states that all points on a wavefront can be considered
as point sources producing secondary spherical waves, and that these secondary waves combine
to produce a new wavefront in the direction of propagation. Diffraction is caused by the
propagation of these secondary waves into the shadow region. The total field strength of the
diffracted wave in the shadow region is the vector sum of the electric field components of the
individual secondary waves.
Consider a transmitter and receiver separated in free space by a distance d = dT +dR as shown
in Fig. 2.8. Let a semi-infinite screen of effective height H (height above lineof-sight path) and
infinite width in the y-direction be placed between them at distances dT from the transmitter
and dR from the receiver. Assuming H ≪ dT and H ≪ dR , the field strength E at the receiver,
relative to the field strength in the absence of the knife edge, is given by [Geng and
Wiesbeck, 1998]

with
5
0
-
-
-
-
-
Figure 2.9: Normalized magnitude of knife-edge diffracted electric field (or similar for
magnetic field) as a function of the knife-edge diffraction parameter ν.
-- - - - - 0 1 2 3 4 5
where C(ν) and S(ν) are real Fresnel integrals. ν
Approximations
parameter and tabulated values for Fresnel
integrals can be found in [Abramowitz, 1972].

Fig. 2.9 shows the dependency of from the diffraction parameter ν. In the lit region, the
linear superposition of direct line-of-sight and diffracted signal leads to a spatial interference
pattern of the electric field, oscillating around the free-space reference field strength (Fig. 2.9
for ν < 0).

For a receiving antenna located at the shadow boundary (i.e ν = H = 0), the magnitude of
the field strength is half of the reference magnitude without knife edge, independent of the
frequency (cf. Fig. 2.10).
The field strength decreases monotonically when the receiver moves away from the lit-shadow
boundary further into the shadow region (Fig. 2.9 for ν > 0). The signal strength in the deep
shadow region decreases with 1/√f. Fig. 2.11 shows the relative magnitude of the electric field
for three different frequencies (1 GHz, 3 GHz, 10 GHz), and in Fig. 2.10 which shows the
height dependence of the field strength magnitude for the example in Fig. 2.11 at a distance of
100 m behind the semi-infinite knife edge.
To avoid the calculation of Fresnel integrals, the dash-dotted line in Fig. 2.9 is often used as an
approximation for the knife-edge diffraction in practice. The corresponding
Figure 2.10: Height dependence of the normalized field strength 100m behind
the semi-infinite screen (knife edge) in Fig. 2.11 for five different frequencies between
100MHz and 10GHz
equations are given by [Geng and Wiesbeck, 1998]:
for ν < −0.78
(2.14)
for ν ≥ −0.78
For ν < −√2 , the difference between the field strength level in the presence of the knife edge
and without knife edge is less than ± 1.1 dB. Under these circumstances, the knife edge can be
often neglected in practice. This limiting case is strongly related to the concept of Fresnel zones
[Geng and Wiesbeck, 1998] as discussed in the following. The concept of diffraction as a
function of the path difference around the obstacle is explained by Fresnel ellipsoids (see Fig.
2.12). Fresnel ellipsoids represent successive surfaces for which the total diffraction path length
from the transmitter to the receiver is Nλ0/2 (with N = 1,2,3,...) larger than the total path length
of the line-of-sight path (in the absence of the knife edge). The successive Fresnel ellipsoids
alternately provide constructive and destructive interference to the total received signal. The
radius of the Nth Fresnel ellipsoid is denoted by RFN and can be expressed in terms of N, λ0,
dT , and dR by [Geng and Wiesbeck, 1998]

with N = 1,2,3..., (2.15)


where this is only valid for dT ≫ RFN and dR ≫ RFN. Comparing the knife-edge diffraction
parameter ν in (2.13) and the Fresnel zone radius in (2.15), it follows that
Figure 2.11: Normalized field strength produced by an isotropic transmitting antenna in
the presence of a semi-infinite absorbing knife edge in a vertical plane cross section for the
three frequencies1 GHz, 3 GHz, and 10 GHz

corresponds to a knife edge of ”height” H≤ RF1, i.e. the top of the knife-edge touches
the 1st Fresnel ellipsoid. In summary, if an obstacle does not block any of the space contained
within the 1st Fresnel ellipsoid, the diffraction losses will be small, and diffraction effects may
be neglected for practical purposes.
Drill Problem 21 Determine the knife edge diffraction loss for a GSM900 system caused by an
obstacle which is located at 10 km distance from the transmitter and 2 km distance from the
receiver and surmounts the direct connection between the antennas by 50 m. How large is the
radius of the first Fresnel Zone at the point of the obstacle?
In practice, the propagation path may consist of more than one obstacle, in which case the total
diffraction loss must be computed. This multiple diffraction problem
RF1

Tx dT dR Rx

1st Fresnel ellipsoid


Nth Fresnel ellipsoid

Figure 2.12: Geometry for the definition of the Nth Fresnel ellipsoid
is very difficult to solve. The problem of multiple knife-edges for example, cannot be solved
by applying the single knife-edge equation (2.13) several times [Gen98]. Specialized multiple
knife-edge models have been developed for this purpose. However, the limiting case of
propagation over a single knife edge described above gives good insight into the order of
magnitude of the diffraction losses. For a more detailed discussion, the reader is referred to the
literature on multiple knife-edge diffraction [Deygout, 1991].

2.3 Scattering
Scattering occurs when the medium through which the electromagnetic wave travels or with
which the wave interacts consists of objects with dimensions in the order of or smaller
compared to the wavelength. Scattered waves are produced by rough surfaces, small objects,
or by other irregularities in the channel (Fig. 2.13).

Figure 2.13: Scattering from relatively small objects (point scatterer), from statistically
rough surfaces (rough surface scattering), and volumes containing many objects (volume
scattering)
In practice, foliage, street signs, individual trees, and lamp posts induce scattering in a mobile
radio communication system. The actual received signal in a radio system is therefore often
different from what is predicted by free-space propagation, reflection, and diffraction models
alone. This is because when an electromagnetic wave impinges on a single object small
compared to the wavelength, a rough surface, or a volume containing many individual objects
(e.g., trunks, branches, and leaves in a forest), the energy is spread out in all directions due to
scattering, thereby providing additional radio energy at the receiver. Analytical methods for
the scattering of electromagnetic waves are only known for very few canonical targets. The
computational complexity of numerical techniques, which are in principle capable of solving
scattering problems to an arbitrary accuracy, rises very quickly if the problem size increases
beyond several wavelengths. Therefore, in the following only some remarks on rough surface
scattering are given.
Flat smooth surfaces that have much larger dimensions than a wavelength may be modeled as
reflective surfaces, i.e. the Fresnel reflection coefficients (2.8) and the 2-ray propagation model
(2.10) can be utilized. However, the roughness of such surfaces often induces effects different
from the specular reflection described earlier. As the roughness increases, more and more
energy is spread out in directions different from the specular direction (Fig. 2.14). The
roughness of the surface can be only neglected as long as the standard deviation σ of the surface
height (i.e. RMS value of the deviation from the mean height) satisfies [Beckmann and
Spiyyichino, 1987]

(Rayleigh) or (Fraunhofer) (2.16)


where the Fraunhofer roughness criterion is generally the better choice.

specular reflection coherent scattering diffuse scattering


Figure 2.14: Specular reflection for ideal flat smooth surface, primarily coherent scattering for
slightly rough surface, and exclusively incoherent scattering for very rough surface
Drill Problem 22 A plane wave is impinging on a rough surface. Determine the maximum
allowed standard deviation of the surface roughness σ at which the surface is still considered
to be flat. Assume that the allowed averaged phase differences are π/2 according to the Rayleigh
criterion and π/8 according to the Fraunhofer criterion, respectively.
As a larger percentage of the total energy is spread out for increased surface roughness, the
amount of energy in the specular direction reduces. As a first approximation
Multipath and Spatial Interference Pattern
to account for this effect, the flat surface Fresnel reflection coefficient is multiplied by a factor
depending on the surface roughness. The resulting modified Fresnel reflection coefficient is
given by [Beckmann and Spiyyichino, 1987]

(2.17)
However, these modified reflection coefficients only account for the decrease of energy in the
specular direction but do not include the spread of energy into other directions. Details on more
sophisticated rough surface scattering models are given in the literature [Geng and Wiesbeck,
1998].
Drill Problem 23 Determine the difference between the standard and the modified Fresnel
reflection coefficients at the Rayleigh and Fraunhofer criteria. Which criterion would the better
choice if you want a reliable result?

2.4 Practical Link Budget


This is all about finding the signal level received from the signal level transmitted. A link
budget is a formal way of calculating the expected received signal to noise ratio. This is
something designers generally want to know to make design decisions like what antenna gain
and how much transmitter power is needed. This effects the hardware cost and is important in
satisfying the license conditions etc. Knowing how to properly make a link budget is a very
important skill for a communications system design engineer. Some people make a lot of
money out of being able to do it well. It is at the basis of antennas and propagation studies as
the path loss between the terminals depends only on the propagation loss and the antenna gain.

Link budgets usually start with the transmitter power and sum all the gains and losses in the
system accounting for the propagation losses to find the received power. Then the noise level
at the receiver is estimated so we can take the ratio of the signal power to the noise power and
work out the performance of the link. This procedure is shown for the generic system below:
The 3 steps are

 find the signal power at the receiver by subtracting the path loss from the transmitted
power, remembering to account for antenna gains and feeder losses.
 find the noise power from the antenna and add to this any noise generated within the
system
 Calculate the ratio of signal power to noise power

What is not included yet in the above in order to avoid confusion is the interference.
Interference can often be treated like additional noise, but the effect of interference depends
very much on the modulation scheme being used. With digital systems, interference can be
treated as noise, but beware of pulse type interference, which may have a low average power
but can completely disrupt services like DTT and DAB through causing bursts of
unrecoverable errors that prevent the highly compressed content from being decoded.

2.4.1 EIRP
The transmitter parameters are often further simplified using the concept of EIRP. This is
useful as it allows us to treat systems with very different antenna characteristics similarly.
In radio systems, the Equivalent Isotropically Radiated Power (EIRP) is the amount of power
that would have to be radiated by an isotropic antenna to produce the equivalent power density
observed from the actual antenna in a specified direction. The EIRP is still a function of
direction, we are not assuming power is radiated isotropically. Usually EIRP is quoted for bore
sight, defined as the axis of maximum radiation. Occasionally we need to refer to the off axis
EIRP which may be in the direction of another system that is suffering interference.

The EIRP is usually quoted in decibels compared to a reference power, e.g. 1watt, 0dBW or 1
milliwatt 0dBm. The EIRP is a useful quantity for comparing systems as it is system
independent, that is we do not need to know anything else in order to calculate the radiated
field strength.
2.4.2 Path Losses
We now need to consider the link parameters - the path loss, which we know this already, it is
the sum of all the losses between transmitter and receiver that are not to do with the antennas
or feeders.

Path loss = Free space loss + Gas loss + Additional path loss

Signal power at receiver

We now have enough information to calculate the signal power at the receiver:

Received power = EIRP - Path Loss + Receiver antenna gain

E.g. Handheld radio 448 MHz, EIRP ~ 0.5 Watt = -3 dBW, Antenna gain, 0 dB (Isotropic) so
for a 1km line of sight path, the loss = 85 dB and the received power = -88 dBW (That is a
strong signal). It is easy, but we have assumed the receiver is linear. With high received signal
powers from -40 dBW upwards, this becomes less likely. Many strong signals at a receiver
may cause undesirable intermodulation products to be generated which will degrade the
performance.

Noise power at the receiver

We are half way to finishing our link budget with the signal power at the receiver. We need to
know the noise power to find the signal to noise ratio. Noise comes from several sources, there
is natural noise from the environment, noise generated within the receiver itself and man made
noise. Everything with a temperature will generate noise - Boltzmann’s law says the noise
power per unit bandwidth = kT where k is Boltzmann’s constant and T is the absolute
temperature in Kelvin.
Important features of this type of noise is that noise is additive, if you have two noise sources
you get the sum of the noise power from each, and noise has a flat spectrum, so if you increase
the bandwidth you increase the noise power in proportion.

Boltzmann’s constant is often expressed in the units of dB Watts per Hz per Kelvin, that is how
many watts you get per Hz of bandwidth for each Kelvin of temperature. Its value is -228.6
dBw/HzK. For example you might an antenna looking at the ground has a noise temperature
of 290K. The noise power received in a 1MHz bandwidth at a noise temperature of 290K is:

Noise power per MHz = -228.6 + 60 + 24.6 = -144 dBW

 Other external noise sources not part of the system include


 The atmosphere including the ionosphere
 The Earth ~ 290K
 The Sun (it is very hot!)
 Galactic sources (Crab Nebula, etc)
 Cosmic background of ~ 2K
 Man made noise (ranging from negligible to very high)
There is some evidence that the man made noise levels are increasing in some environments.
This is a hot topic as Ultra-Wideband systems intend to operate below the noise floor of
conventional systems and there is some disagreement between the UWB and conventional
camps over what that level is.

Another issue is over power line transmission PLT technology, which used mains wiring to
send data in the vein hope that the wires will not radiate. They do radiate of course and the aim
is to keep the additional noise to below the current noise floor – this dispute is quite heated
because prototype PLT systems have been demonstrated by the BBC to be severely damaging
to broadcast reception.

Noise at the antenna

The antenna picks up noise from the sources in the previous slide, depending on its radiation
pattern, it also generates noise through its own temperature and losses and picks up noise from
the Earth at 290K in the sidelobes. To estimate the noise picked up by an antenna, a quick
method is to take the antenna efficiency as an indication of the sidelobe power. So, if the
efficiency is 60%, the external noise in the direction of the antenna accounts for 60% and the
sidelobes represent 40% of the noise pickup. It is further assumed that half of the sidelobes are
looking towards the sky and half towards the ground. For example, with a 60% efficient dish
as might be used for satellite TV reception at 12GHz, the sky noise temperature may be ~15K
and the ground noise temperature 200K. The total antenna noise is estimated as:

Antenna noise temperature = 0.6 x 15K + 0.4 x (15K + 200K)/2 = 52K

The ratio is nearly 4 times compared to what it would be without the sidelobes, so dish
efficiency can sometimes matter even more for noise than it does for received signal power.

Noise generated in the receiver

Noise is generated by all inline devices, for example passive devices including attenuators,
waveguides, cables, filters etc. or active devices, amplifiers, mixers etc. Each can be
represented reasonably accurately as an additional noise source (resistor) at the input to the
system:

Passive Devices

5/8
If a passive device has loss it will add noise to the system proportional to its temperature
(Assumed 290K unless known) and the loss.

E.g. For a gain of 0.9 (10% power loss), T Device = 29K.For a feeder loss of 1 dB the noise
temperature increase works out as 75K.Nothing adds no noise unless it has a temperature of
absolute zero. Feeders always have loss. The loss at microwave frequencies is higher and feeder
lengths need to be minimised to obtain a low overall system noise temperature. Radio
astronomy stations whose performance would be totally devastated by a 75K feeder
temperature dispense with the feeder altogether. They use beam waveguides where the only
loss is in the reflecting surfaces – which is low because of their size plus they can be easily
cooled.
Active Devices

All active devices generate noise internally, the reasons are complex, but it can be modeled
as an effective noise temperature, e.g. for an amplifier:

kTrx Brx = amplifier noise power


Typical noise temperatures for real amplifiers are in range 10K - 1000K. The Noise Factor is
a measure of how much noise is added by an active device. When the receiver is matched by
a load resistor at standard temperature T0 290K noise power input is:

Noise Factor F = (Noise Out / Noise In) – referenced to input!


Noise Factor:

Which if expressed in dB is FdB = 10log(F) we call this the noise figure. in terms of noise
power after substituting for k and T0:
N0 = FdB -204 + 10log(B) dBW

Alternatively, we can consider an Effective noise temperature:

N0 = kT0B + kTeB = k(T0 + Te)B and

6/8
Cascaded Sources

When a device is described as having a noise temperature or a noise figure, this is always
relative to the INPUT of the device. When summing noise contributions we need to be careful
about gain and loss, an amplifier will amplify input noise as well as the signal and a lossy
device will attenuate input noise as well as adding noise due to its own noise temperature.

Ptotal = G1G2G3...GnkT0B + G1G2G3...GnkT1B + G2G3...GnkT3B + G3...GnkTnB + …

Input x total gain 1st x total gain 2nd x (total gain – gain of 1st stage)...

This tells us something useful – if we have enough gain in our front end low noise
amplifier, the noise figure of the rest of the receiver is of secondary importance. There is a
trade off though as too much gain is bad for performance. The problem is that putting gain
at the front end of the receiver before the filtering reduces the systems immunity to strong
out of band signals. Too much overall gain will add to the inter-modulation distortion
generated from in-band signals and thereby reduce the dynamic range of the receiver.

Example

Find the overall noise figure of a receiver with a 10 dB noise figure preceded by an
amplifier with a noise figure of 0.5 dB and a gain of 20 dB ?
To solve this we need to sum the input noise plus the noise temperature of the amplifier
multiplied by the gain of the amplifier plus the noise temperature of the original receiver.
(Rearrange the equation Ptotal = G1kT0B + G1kT1B + kT2B)
T0 = 290K
T1 0.5 dB NF = noise temp of (100.5/10/ - 1) x 290 = 35K
T2 10 dB NF = noise temp of (1010/10 - 1) x 290 = 2610K G2 20 dB Gain = 100 x
Tadded = 35 + 2610/100 = 61K, Equivalent to a system noise figure of 0.8 dB
We have assumed a 60% antenna efficiency, that is quite a challenge for a mass market
product. When satellite TV first became popular the original LNBs had noise temperatures

7/8
of about 250K so the total noise temperature was 300K and the extra noise from the antenna
really didn’t make that much difference. Nowadays LNBs are apparently available with
noise temperatures of around 50K giving a total noise temperature of 100K. Now the
antenna is responsible for half of the noise power. The LNB improvement equates to an
improvement of 4.7 dB in SNR. That is more than enough gained to permit the dish size
to be reduced to 45cm. An even smaller dish could be used if it was not for the congestion
in the Geostationary orbit. Instead, we now use 45cm and the improvements in LNBs have
allowed an increased data rate.

Why worry about noise figure?

Say we have a receiver with a noise figure of 10 dB and 20MHz bandwidth. What is the
equivalent noise power?

The receiver noise floor is – 121 dBW. Say we are looking for a TV satellite, with a good
antenna so the antenna noise temperature is low, maybe 50K, if we work out the input noise
power we find it is:

So our receiver noise is much higher than the input noise, with at –121 dBW being 18 dB worse
than it would be an ideal noiseless receiver. An input signal of –120 dBW from a transponder
would have an SNR of 1 dB, if our system was noiseless this could have been 19 dB. To fix
this the manufacturer fits a 0.5 dB NF LNA , which as we know gave us 61K of additional
noise, which when added to the antenna noise gives a a total of 111K.

The signal would now have an SNR of 15 dB! That is the difference between a good watchable
picture and none at all. If we did not have the low noise amplifier the satellite operator would
need to compensate with 14 dB more power, which would be economic suicide.

In perspective
Remember to keep noise temperatures in perspective, there is usually no need for one part
of a system to have a vastly better noise performance than another. Low noise receivers are
important in space and satellite systems but much less important for terrestrial systems in

8/8
noisy environments.
In a typical business area, at UHF the noise temperature seen by the antenna will be
1000K or more. This is similar to Te for our example so improving our 10 dB NF receiver is
not going to make such a large difference to the signal to noise ratio. Adding a 20 dB gain
amplifier in front would actually make the receiver worse as it would affect the strong signal
handling – there lots of strong signals around in industrial areas.

Things are different for SETI trying to look for ET at π times the hydrogen line (~4.5 GHz)
from a quiet location with a good antenna with input noise ~ 10K. Even the receiver noise
temperature of 61K would be considered poor and Cryogenic cooling and very low noise
systems are needed for radio astronomy.

Summary

Do link budgets in dB as it is easier. The steps are:


Find the Signal (dBW)

1. Work out the transmitter EIRP


2. Work out the path loss
3. Add the receiver antenna gain
4. Subtract the feeder losses

Find the Noise (dBW)

1. Find the natural noise


2. Add the noise from the antenna
3. Add the noise from the feeders etc.
4. Add the noise from the receiver

Find the SNR (dB) = Signal (dBW) – Noise (dBW)

Some Examples
Voyager - again. Last time we calculated Goldstone received a signal level of -180 dBW from Voyager. Building on our previous example, what is the data rate one
might expect to be able to receive from a space probe.

9/8
Goldstone uses the best LNAs available and the system noise temperature is around 30k. The
antenna is designed for very low sidelobe noise
and it tends to not be used at low elevation angles. Goldstone uses cryogenic cooling on the
LNAs, much of the noise power comes from the waveguide loss of 0.2 dB, equivalent to a
noise temperature of 14K. The rest comes from the atmospheric loss and antenna noise
temperature, both of which depend on the elevation angle. This equates to a noise power of:

N = -228.6 + 10log(30) = -213.8 dBW/Hz

The S/N is the difference:

S/N = -180 – (-213.8) = 33.8 dB/Hz

Shannon's theory says in the limit we get no errors with a bit energy to noise energy ratio of -
1.6 dB. Some modern exotic schemes get to within a dB or so of this, but remember Voyager
was designed in the 1970s. The coding scheme deployed needs a bit energy to noise energy
ratio of 2.5 dB which gives us a maximum data rate 33.8 – 2.5 = 31.3 dB/Hz, which is
equivalent to 1.35 kb/s.

(note: we have ignored several degradations, for example Goldstone’s antenna efficiency, in
this link budget to avoid clouding the point)

1. A PMR system

Back to our PMR448 - What is the maximum line of sight range? Radio specification:
Handheld radio 448 MHz, TX ~ 0.5 Watt Antenna gain, 0 dB (Isotropic)
Noise figure = 6 dB, Bandwidth = 25kHz Minimum usable SNR 12 dB

10/
8
Find the noise floor:

Assume the antenna is seeing half ground (290K), half sky (30K) = 160K average Noise figure
6 dB gives (4 -1) x 290 = 870K noise temp
Total noise temperature = 870 + 160 = 1030 K
Noise power = kTB = 1.38x10-23 x 1030 x 25,000 = 3.55x10-16 W = -155 dBW
Transmitted Power (EIRP):
Our EIRP is the transmitter power + the antenna gain EIRP = -3 dBW power + 0 dBi antenna
gain = -3 dBW
Maximum path loss capability:
This is simply the difference between the EIRP and the required power, so we need to find
minimum required signal power: We need 12 dB SNR to communicate so the signal power
must be at least 12 dB above the noise floor
Required power = –155+12 = -143 dBW
Path loss capability = (-3) - (-143) = 140 dB
Path loss with distance:
At this frequency gases can be ignored, so the line of sight path loss is given by Path Loss =
32.4 + 20log(d) + 20log(448) = 85.4 + 20log(d) dB
This should equal 140 dB, so 20log(d) = 54.6
And d = 10(54.6/20) = 537km
The reason that PMR 448 handies don’t work over 537km because there are not usually any
537km line of sight paths. We have of course completely forgotten about interference here.
Interference is frequently the major limiting factor in mobile communications systems. That is
why it is important to understand the propagation characteristics of unwanted co-channel
signals as well as wanted ones.

Signals and Noise


An exact copy of the signal transmitted is not what is received at the receiver. Noise is added
to the signal, the amplitude and phase of the signal varies with time and location, there is a time
delay which may be variable and the signal shape is distorted. So far we have treated noise as
a purely random signal which is added at the receiver input, we call this Additive White
Gaussian Noise (AWGN).

11/
8
I and Q representation
Typically, we represent signals as their complex baseband equivalents, I and Q representation
where I and Q are orthogonal
and represented as a complex number:
n(t) = xn(t) + jyn(t)
Where x and y are random variables of zero mean. The mean noise power is related to the noise
variance.

We will almost always treat noise as additive and Gaussian.

Showing
We know we can represent a random variable by its Probability Density function p(x). The
probability the value lies between two values a and b is:

All PDFs by definition have a total integral of 1

So the probability a value is less than a is:

12/
8
We call P the cumulative distribution and so also the area between a and b is the probability
below b minus probability below a;

Differentiating:

The Expectation of some function f(x) of a random variable x is:

Expectation is a statistical term referring to the Expected (average) value of the function. Here
we are interested in the Expectation (mean) of x,
which we get if we set f(x) = x. This is often called the first moment of x in statistics.

The variance is given by:

To find the power of complex signal n(t) we multiply by the complex conjugate.

The mean power is the expectation of N(t):

13/
8
Expectation is a statistical term, according to Wikipedia:

In probability theory the expected value (or mathematical expectation) of a random variable is
the sum of the probability of each possible outcome of the experiment multiplied by its payoff
("value"). Thus, it represents the average amount one "expects" as the outcome of the random
trial when identical odds are repeated many times. Note that the value itself may not be
expected in the general sense; it may be unlikely or even impossible. For example, the expected
value from the roll of an ordinary six-sided die is 3.5, which is not one of the possible outcomes.
Remember that xn and yn oscillate about zero, which means their mean values are zero.

so

It is assumed σx and σy have the same variance σn

so

3. PROPAGATION MODELS FOR WIRELESS


COMMUNICATION SYSTEM
4. Introduction channel model
The wireless communication system possesses several challenges for the reliable and a high
speed communication. It is not receptive of noise channel and other channel hindrance, but
these obstacle changes with time in unforeseeable ways due to user movement. We will
characterize in detail the variation in the received signal power over the distance due to path
loss and shadowing. Path loss models describe the signal attenuation between a transmitter
and receiver antenna as a function of propagation distance and other parameters which is
caused by the dissipation of the power radiated by the transmitter as well as effects of the
propagation channel. Shadowing is caused by obstruction between the transmitter and the
receiver that attenuate the signal power through absorption, reflection, scattering, and

14/
8
diffraction. A very important practical issue is to test and validate the ability of the “smart”
antenna array to meet performance requirements. For this purpose, a channel model is
needed to take into account the temporal and spatial characteristics of radio propagation.

2.2
Three Basic Propagation Mechanisms
Reflection, diffraction and scattering are the basic propagation mechanisms which might have
an impact on propagation in terrestrial radio communication systems operating in the 30 MHz
to 10 GHz frequency range (most mobile communication systems). Propagation mechanisms
affecting systems at low frequencies (e.g., ground wave propagation, ionospheric reflection) Commented [sg4]: Three Basic Propagation Mechanisms
and high frequencies (e.g., gaseous absorption, rain attenuation, troposcatter) are not discussed – Reflection (Ground Reflection) , Diffraction, Scattering,
here. Details can be found in the literature [Giger, 1991]. Reflection, diffraction, and scattering
however, are briefly explained in the next subsections.
Reflection
When a radio wave propagating through a medium encounters a boundary layer which is plane
and large compared to the wavelength the wave is partially reflected and partially transmitted.
If the second medium is perfectly electric conducting (PEC), then all incident energy is
reflected back into the first medium without any loss of energy. The electric field strength of
the reflected and transmitted waves may be related to the incident wave through the Fresnel
reflection and transmission coefficients Rk,⊥ and Tk,⊥ respectively. The Fresnel coefficients
depend on the material properties, wave polarization, angle of incidence, and material
parameters which are frequency dependent. A polarized electromagnetic wave may be
mathematically represented as a sum of two orthogonal components, such as vertical and
horizontal, or left-hand and right-hand circularly polarized components. For an arbitrary
polarization, linear superposition may be used to compute the reflected fields from a surface.
Fig. 2.2 shows an electromagnetic wave incident at an angle θi relative to the normal on the
interface between medium 1 and 2. Part of the energy is reflected back to medium 1 at an angle
θr=θi (law of reflection), and part of the energy is transmitted (refracted) into the second
medium at an angle satisfying Snell’s law of refraction [Geng and Wiesbeck, 1998]:

(2.7)
where is the relative complex permittivity, including also losses due to a
non-vanishing conductivity σ, and is the complex permeability of the medium for
most cases in radio communications).

(a) parallel polarization (b) perpendicular polarization

15/
8
Figure 2.2: Geometry for reflection and transmission of a plane wave incident upon a plane
interface between two media 1 and 2 of different material properties.
The nature of reflection varies with the polarization of the electric field. In Fig. 2.2a, the electric
field vector is parallel to the plane of incidence (i.e. the plane containing incident, reflected,
and transmitted rays), and in Fig. 2.2b, the electric field vector is perpendicular to the plane of
incidence. Using superposition, only these two orthogonal and linear independent polarizations
need be considered to solve a general reflection problem. Considering an interface between air
(i.e. and a non-magnetic (but possibly lossy) half space 2 (i.e. ), the
Fresnel reflection/transmission coefficients for the electric field parallel (denoted as k) and
perpendicular (denoted as ⊥) to the plane of incidence are given by [Geng and Wiesbeck,
1998]:

and

and , (2.8b)

where is the wave impedance of the propagation medium. For the reflection
between air and the earth surface, typical values for the soil permittivity and the conductivity
are ǫr2 = 3...25 and σ2 = 10−4S/m...0.1S/m, respectively. Keep in mind that these values are
dependent on the frequency.
Drill Problem 17 Automotive rain sensors detect water on the windshield of a vehicle. The
infrared beam is radiated at a certain angle (θc) from inside of the car on the glass. If the
windshield is dry, all the energy comes back to the sensor; in case of water presence on the
glass surface, less energy comes back and the wipers are activated.
What is the name of the phenomenon, when no energy is transfered form one dielectric medium
to the other? Calculate the value of the angle θc if the permittivity (assume no losses) of the
windshield glass is ǫw = 7.6.
Fig. 2.3 shows plots of the reflection and transmission coefficients (here magnitude only) for
both parallel and perpendicular E-field polarization (e.g. vertical and horizontal, if the interface
is parallel to the xy-plane) as a function of the incident angle for the special case when a wave
propagates in free space ( =1) and the reflecting soil half space is characterized by
=4,36-j3 and =1. For parallel polarization, there exists usually an angle for which the
reflection coefficient shows a local minimum. In case of a lossless medium 2, this minimum
reduces to zero, and the corresponding incident angle is called Brewster angle [Balanis, 1989].
Therefore, the Brewster angle is the angle at which in medium 1 no reflection occurs. For non-
magnetic (generally if µr1 = µr2) and lossless half spaces 1 and 2, the Brewster angle only
exists for parallel polarization and is given by

. (2.9)
Drill Problem 18 A parallel polarized electromagnetic wave radiated from a submerged
submarine impinges on a planar water-air interface. The dielectric constant of water is ǫr = 81

16/
8
at the frequency of the impinging wave. Assume that the impinging wave is a plane wave at
the interface. Determine the angle of incidence to allow complete transmission of energy.
In most practical cases of reflection and transmission, the horizontal and vertical axes of the
spatial coordinates do not coincide with the perpendicular and parallel axes of the propagating
waves. In addition, the interface between medium 1 and 2 is in general not perpendicular to the
z-axis of a global coordinate system (e.g., reflection from a sidewall of a building). Therefore,
the plane of incidence, field components and angles have to be transformed first (according to
the true geometry) before utilizing the Fresnel coefficients in (2.8) [Geng and Wiesbeck, 1998].

1
reflection parallel
0.9 reflection perpendicular

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0 10 20 30 40 50 60 70 80 90
incident angle [deg]

Figure 2.3: Magnitude of reflection and transmission coefficients as a function of the incident
angle for an interface between air ( =1) and a non-magnetic
(but lossy) medium characterized by a complex relative permittivity of and =4,36-j3 and a
relative permeability of
Since there can’t be an electric field in a perfect conductor, all energy is reflected when a ray
encounters a PEC i.e. |R| = 1. As the electric field tangential to the surface must vanish in order
to satisfy Maxwell’s equations, the reflection coefficients for parallel and perpendicular
polarization must be Rk = +1 and R⊥ = −1 respectively, regardless of the incident angle. The
± signs here are valid for the reference directions defined in Fig. 2.2. Commented [sg5]: Practical Link Budget, ( Text 1 - 2.2

Practical Link Budget, ( Text 1 - 2.2 and Ref1 - Chapter 4).

17/
8
What is free space?

Free space in this context means space with nothing at all in it, it does not exist in the
known universe but interstellar space is a good approximation. We start at this level as
there is nothing there to make the maths even more complicated than it already is.

The important features of free space:

Uniform everywhere
Contains no electrical charge
Carries no current
Infinite extent in all dimensions
http://www.mike-willis.com/Tutorial/PF3.htm

18/
8

You might also like