You are on page 1of 203

Automatic

Cell
Planning

Dr Rupert Rawnsley

Originally submitted in June 2001 for
the degree of Doctor of Philosophy to
the Department of Computer Science,
University of Wales, Cardiff, UK.

II
Abstract
An effective and efficient mobile telephone service relies, in part, on the deployment
and configuration of network infrastructure. This thesis describes some of the problems
faced by the network planners in providing such a service. Novel algorithmic methods
are presented for designing networks that meet both the subscribers demand for a high
quality of service and the operators requirement for low infrastructure overheads.
Automatic Cell Planning (ACP) is the generic name for these network optimising
techniques. Existing research into ACP is extended through the introduction of
interference surrogates, which give the network designer greater control over
interference in a wireless network. Interference management is the central theme of this
thesis. The accurate simulation and control of interference leads to a more efficient use
of the available radio spectrum (a limited and expensive commodity). Two approaches
to ACP are developed: in the first, network performance (in terms of subscriber
coverage) is given priority over the cost of network infrastructure; in the second, these
priorities are reversed. These two distinct approaches typify a trade off that lies at the
heart of the problem of network design a trade off between spectral efficiency and the
cost of network infrastructure.
Keywords: automatic cell planning, ACP, cellular mobile telephone network
optimisation, cell plan optimisation, cell optimisation, cell dimensioning, base station
placement, wireless networks, spectral efficiency, spectrum efficiency, channel
assignment, CAP, frequency assignment, FAP, interference surrogates, signal freedom.

III
Contents
ABSTRACT II
CONTENTS III
LIST OF FIGURES V
LIST OF TABLES VIII
GLOSSARY OF TERMS IX
GLOSSARY OF SYMBOLS X
CHAPTER 1 INTRODUCTION 1
1.1 OUTLINE OF PROBLEM 1
1.2 PREVIOUS WORK 7
1.3 THESIS AIMS AND OBJECTIVES 19
1.4 TEST PLATFORM 20
1.5 THESIS OUTLINE 21
CHAPTER 2 NETWORK MODEL 23
2.1 SIMULATION ISSUES 23
2.2 DESCRIPTION OF MODEL 31
2.3 PERFORMANCE EVALUATION 51
2.4 DATA SETS 56
CHAPTER 3 NETWORK DESIGN PART 1 65
3.1 CHOICE OF OPTIMISATION ALGORITHMS 65
3.2 INITIALISATION ALGORITHMS 66
3.3 SEARCH ALGORITHMS 84
3.4 FACTORS LIMITING SEARCH PERFORMANCE 100
CHAPTER 4 CHANNEL ASSIGNMENT 103
4.1 PROBLEM DEFINITION 103
4.2 PERFORMANCE EVALUATION 109
4.3 DATA SETS 112
4.4 COMMON ALGORITHMIC CONCEPTS 114

IV
4.5 INITIALISATION ALGORITHMS 115
4.6 SEARCH ALGORITHMS 118
4.7 COMPARISON OF RESULTS 123
CHAPTER 5 INTERFERENCE SURROGATES 125
5.1 ROLE OF THE INTERFERENCE SURROGATE 126
5.2 INTERFERENCE SURROGATE DESCRIPTIONS 128
5.3 METHODOLOGY OF COMPARISON 133
5.4 RESULTS 136
5.5 CONCLUSION 146
CHAPTER 6 NETWORK DESIGN PART 2 148
6.1 EVALUATION OF TWO-STAGE NETWORK DESIGN PROCESS 148
6.2 NETWORK DESIGNS WITH HIGH COMPUTATIONAL INVESTMENT 152
6.3 NETWORK DESIGNS WITH MINIMAL INFRASTRUCTURE 156
CHAPTER 7 CONCLUSIONS AND FUTURE WORK 164
7.1 THESIS CONCLUSIONS 164
7.2 NETWORK MODEL ENHANCEMENTS 178
7.3 IMPROVEMENTS TO OPTIMISATION METHODOLOGY 182
ACKNOWLEDGMENTS 184
BIBLIOGRAPHY 185


V
List of Figures
Figure 1.1: The simplified path of a call between a wireless telephone and a wireline telephone. ........................................2
Figure 1.2: An example of different switching technologies connecting both wireless and wireline telephones..................3
Figure 1.3: A pictorial illustration of radio signal propagation in wireless and wireline networks.......................................6
Figure 1.4: The relationship between re-use distance D and cell radius R in co-channel cells. ..........................................10
Figure 1.5: A channel re-use pattern for 7 different sets of channels.......................................................................................11
Figure 1.6: The translation of a political map into a colouring constraint graph..................................................................12
Figure 1.7: Some received power levels at a mobile station; this data may be measured or predicted. .............................13
Figure 2.1: An abstracted traffic cover problem using multiple servers..................................................................................28
Figure 2.2: Rasterisation of the network working area. ............................................................................................................32
Figure 2.3: The test point layout in NetworkA3. .........................................................................................................................33
Figure 2.4: The test point traffic volumes in NetworkA3. ..........................................................................................................34
Figure 2.5: The test point service thresholds in NetworkB1......................................................................................................34
Figure 2.6: The site positions in NetworkB1. ..............................................................................................................................35
Figure 2.7: The geometry of the horizontal loss calculation for directional antennae...........................................................36
Figure 2.8: The geometry of the vertical loss calculation for antennae...................................................................................37
Figure 2.9: The angles-of-incidence for a site in NetworkA4....................................................................................................38
Figure 2.10: Two active base stations in NetworkA4.................................................................................................................42
Figure 2.11: The propagation gain from a site in NetworkA2..................................................................................................43
Figure 2.12: The signal strength due to an active base station in NetworkA4........................................................................45
Figure 2.13: An example of serviced test points in NetworkA4. ...............................................................................................46
Figure 2.14: An example CIR calculation for two active base stations in NetworkA4. .........................................................49
Figure 2.15: An example of test point cover for two base stations in NetworkA4. .................................................................50
Figure 2.16: Comparison operator for network performance metrics....................................................................................55
Figure 2.17: NetworkA1 (a) test points, (b) traffic densities, and (c) candidate site positions. .............................................57
Figure 2.18: NetworkA2 (a) test points, (b) traffic densities, and (c) candidate site positions. .............................................59
Figure 2.19: NetworkA3 (a) test points, (b) traffic densities, and (c) candidate site positions. .............................................60
Figure 2.20: NetworkA4 (a) test points, (b) traffic densities, and (c) candidate site positions. .............................................61
Figure 2.21: NetworkB1 (a) test points, (b) traffic densities, (c) candidate site positions, and (d) service thresholds. ......62
Figure 3.1: Algorithm InitBlank....................................................................................................................................................66
Figure 3.2: Algorithm InitRandomA. ...........................................................................................................................................68
Figure 3.3: An example network design produced by InitRandomA. ......................................................................................69
Figure 3.4: The distribution of traffic cover for InitRandomA on NetworkA4........................................................................70
Figure 3.5: The distribution of test point cover for InitRandomA on NetworkA4...................................................................70
Figure 3.6: The distribution of site occupancy for InitRandomA on NetworkA4. ..................................................................71

VI
Figure 3.7: The distribution of number of base stations active for InitRandomA on NetworkA4.........................................71
Figure 3.8: Graph of traffic cover vs. base stations active for InitRandomA on NetworkA4................................................72
Figure 3.9: Graph of traffic cover vs. sites occupied for InitRandomA on NetworkA4. ........................................................72
Figure 3.10: Graph of test point cover vs. base stations active for InitRandomA on NetworkA4. .......................................73
Figure 3.11: Graph of test point cover vs. sites occupied for InitRandomA on NetworkA4..................................................73
Figure 3.12: A subsection of the example network generated by InitRandomA.....................................................................74
Figure 3.13: Algorithm InitRandomB..........................................................................................................................................76
Figure 3.14: An example network design produced by InitRandomB.....................................................................................77
Figure 3.15: The distribution of traffic cover for InitRandomB on NetworkA4......................................................................78
Figure 3.16: The distribution of test point cover for InitRandomB on NetworkA4. ...............................................................78
Figure 3.17: Graph of traffic cover vs. base stations active for InitRandomB on NetworkA4..............................................79
Figure 3.18: Graph of traffic cover vs. sites occupied for InitRandomB on NetworkA4.......................................................79
Figure 3.19: Graph of test point cover vs. base stations active for InitRandomB on NetworkA4. .......................................80
Figure 3.20: Graph of test point cover vs. sites occupied for InitRandomB on NetworkA4..................................................80
Figure 3.21: Algorithm InitRandomC..........................................................................................................................................81
Figure 3.22: An example network design produced by InitRandomC.....................................................................................82
Figure 3.23: The distribution of traffic cover for InitRandomC on NetworkA4......................................................................83
Figure 3.24: The distribution of test point cover for InitRandomC on NetworkA4................................................................83
Figure 3.25: Algorithm SearchRandomB....................................................................................................................................85
Figure 3.26: The performance of the best design produced vs. time for SearchRandomB on NetworkA4.........................86
Figure 3.27: The performance of the best design produced vs. time for SearchRandomC on NetworkA4.........................87
Figure 3.28: Algorithm SearchHillClimbA. ................................................................................................................................89
Figure 3.29: The performance of the best design produced vs. time for SearchHillClimbA on NetworkA4 (run 1). ........91
Figure 3.30: The performance of the best design produced vs. time for SearchHillClimbA on NetworkA4 (run 2). ........91
Figure 3.31: The performance of the best design produced vs. time for SearchHillClimbA on NetworkA4 (run 3). ........92
Figure 3.32: Algorithm SearchConstructA. ................................................................................................................................94
Figure 3.33: The performance of the best design produced vs. time for SearchConstructA on NetworkA1.......................95
Figure 3.34: The performance of the best design produced vs. time for SearchConstructB on NetworkA1.......................98
Figure 4.1: The comparison operator for assignment performance metrics. .......................................................................110
Figure 4.2: CAP problem CapA210 generated from NetworkA2. ......................................................................................113
Figure 4.3: Algorithm InitRandomD..........................................................................................................................................115
Figure 4.4: Algorithm InitConstructA........................................................................................................................................117
Figure 4.5: Algorithm SearchRandomD. ..................................................................................................................................119
Figure 4.6: The performance of the best assignment produced vs. time for SearchRandomD on CapA290.................120
Figure 4.7: Algorithm SearchHillClimbB. ................................................................................................................................122
Figure 4.8: The performance of the best assignment produced vs. time for SearchHillClimbB on CapA390...............123
Figure 5.1: The performance of an ideal (hypothetical) interference-surrogate.................................................................127
Figure 5.2: An illustration of the attenuation of interference by channel separation...........................................................130
Figure 5.3: The performance of an imperfect interference surrogate, in which one pair of designs violate the IS
Condition............................................................................................................................................................................134
Figure 5.4: The performance of an imperfect interference surrogate, in which three pairs of designs violate the IS
Condition............................................................................................................................................................................135
Figure 5.5: The normalised inaccuracy of three ISs vs. time for NetworkA1........................................................................137
Figure 5.6: Graph of overlap vs. test point cover for 10 designs in NetworkA1...................................................................138
Figure 5.7: Graph of non-dominance vs. test point cover for 10 designs in NetworkA1.....................................................138
Figure 5.8: Graph of signal freedom vs. test point cover for 10 designs in NetworkA1. .....................................................139
Figure 5.9: The normalised inaccuracy of three ISs vs. time for NetworkA2........................................................................140
Figure 5.10: Graph of overlap vs. test point cover for 10 designs in NetworkA2.................................................................140

VII
Figure 5.11: Graph of non-dominance vs. test point cover for 10 designs in NetworkA2...................................................141
Figure 5.12: Graph of signal freedom vs. test point cover for 10 designs in NetworkA2. ...................................................141
Figure 5.13: The normalised inaccuracy of three ISs vs. time for NetworkA3. ....................................................................142
Figure 5.14: Graph of overlap vs. test point cover for 10 designs in NetworkA3.................................................................142
Figure 5.15: Graph of non-dominance vs. test point cover for 10 designs in NetworkA3...................................................143
Figure 5.16: Graph of signal freedom vs. test point cover for 10 designs in NetworkA3. ...................................................143
Figure 5.17: The normalised inaccuracy of three ISs vs. time for NetworkA4. ....................................................................144
Figure 5.18: Graph of overlap vs. test point cover for 10 designs in NetworkA4.................................................................144
Figure 5.19: Graph of non-dominance vs. test point cover for 10 designs in NetworkA4...................................................145
Figure 5.20: Graph of signal freedom vs. test point cover for 10 designs in NetworkA4. ...................................................145
Figure 6.1: The performance of the best design produced vs. time for SearchRandomC on NetworkA4.........................150
Figure 6.2: The performance of the best design produced vs. time for SearchHillClimbA on NetworkA4.......................150
Figure 6.3: Algorithm SearchPruneA........................................................................................................................................153
Figure 6.4: The performance of the best BSCP solution vs. time for composite search on NetworkA2............................155
Figure 6.5: Algorithm SearchPruneB........................................................................................................................................158
Figure 6.6: Algorithm SearchHillClimbC.................................................................................................................................159
Figure 6.7: Graph of network infrastructure vs. time for algorithm SearchPruneB on NetworkA1..................................160
Figure 6.8: Algorithm SearchConstructC. ................................................................................................................................162
Figure 7.1: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA1. ..........................167
Figure 7.2: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA2. ..........................167
Figure 7.3: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA3. ..........................168
Figure 7.4: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA4. ..........................168
Figure 7.5: A pathological example of receiver cover for two transmitters..........................................................................170
Figure 7.6: A pathological example of receiver cover for three transmitters........................................................................170
Figure 7.7: Graph of subscriber cover vs. network infrastructure. ........................................................................................171
Figure 7.8: Graph of the effect of computational investment on the pareto-optimal trade off between subscriber cover
and network infrastructure...............................................................................................................................................172
Figure 7.9: Target performance positions on the pareto-optimal graph...............................................................................173
Figure 7.10: Number of base stations vs. computational investment for a fixed subscriber cover.....................................176
Figure 7.11: Infrastructure cost vs. number of base stations...................................................................................................176
Figure 7.12: The determination of optimal computational investment by comparison of computational cost and network
infrastructure cost. ............................................................................................................................................................177


VIII
List of Tables
Table 1.1: Test platform specification. .........................................................................................................................................21
Table 2.1: Traffic capacity vs. number allocated base station channels..................................................................................39
Table 2.2: Summary of statistics for network data sets. .............................................................................................................56
Table 3.1: The statistical performance of algorithm InitRandomA..........................................................................................69
Table 3.2: The statistical performance of algorithm InitRandomB..........................................................................................77
Table 3.3: The statistical performance of algorithm InitRandomC..........................................................................................82
Table 3.4: A summary of results for initialisation algorithms. ..................................................................................................84
Table 3.5: The performance of the best network designs for SearchRandomB. .....................................................................86
Table 3.6: The performance of the best network designs for SearchRandomC......................................................................87
Table 3.7: Example moves for SearchHillClimbA......................................................................................................................88
Table 3.8: The performance of network designs for SearchHillClimbA..................................................................................90
Table 3.9: The initial and final traffic covers for multiple executions of SearchHillClimbA.................................................92
Table 3.10: The performance of the best network designs for SearchConstructA..................................................................95
Table 3.11: The performance of the best network designs for SearchConstructB..................................................................97
Table 3.12: A summary of results for all search algorithms on all networks. .........................................................................99
Table 3.13: Computational complexity of network performance evaluation. .......................................................................100
Table 4.1: Summary statistics for CAP problem instances......................................................................................................113
Table 4.2: The performance of the best assignments produced by InitConstructA. .............................................................118
Table 4.3: The performance of the best assignments produced by SearchRandomD..........................................................120
Table 4.4: The performance of the best assignment produced by SearchHillClimbB. ........................................................122
Table 5.1: An example of IS performance vs. time....................................................................................................................136
Table 5.2: The final IS inaccuracy values for NetworkA1. ......................................................................................................137
Table 5.3: The final IS inaccuracy values for NetworkA2. ......................................................................................................139
Table 5.4: The final IS inaccuracy values for NetworkA3. ......................................................................................................141
Table 5.5: The final IS inaccuracy values for NetworkA4. ......................................................................................................143
Table 5.6: A summary of IS inaccuracy results.........................................................................................................................146
Table 6.1: The performance of the best BSCP solutions for SearchRandomC and SearchHillClimbA............................149
Table 6.2: The performance of the best network design (after channel assignment) for SearchRandomB.......................151
Table 6.3: The performance of the best BSCP solution for composite search. .....................................................................154
Table 6.4: The performance of the best network design for seven day optimisation............................................................155
Table 6.5: A comparison of actual network infrastructure with theoretic minimum............................................................156
Table 6.6: The performance of the final network designs produced by SearchPruneB.......................................................160
Table 6.7: The effect on network infrastructure of algorithm SearchPruneB. ......................................................................161
Table 6.8: The performance of the best network designs produced by the composite constructive search algorithm.....163

IX
Glossary of Terms
ACP Automatic Cell Planning
AI Angle of Incidence
BNDP Broadcast Network Design Problem
BSPP Base Station Placement Problem
CAP Channel Assignment Problem
CDMA Code Division Multiple Access
CIR Carrier to Interference Ratio
CNDP Capacitated Network Design Problem
FAP Frequency Assignment Problem
FDMA Frequency Division Multiple Access
FSP Fixed Span Problem
GCP Graph Colouring Problem
GOS Grade Of Service
GSM Global System for Mobile communications
ICR Interference to Carrier Ratio
QOS Quality Of Service
TDMA Time Division Multiple Access
VSP Variable Span Problem
WCNDP Wireless Capacitated Network Design Problem

X
Glossary of Symbols
CV

CIR cover threshold
NC

Average CIR of non-covered test points
AV

Average ICR of all test points
AI
il
a Angle-of-Incidence from site
l
s to test point
i
r
B Set of all base stations available in N
AC
B Set of active base stations
j
b
j
th
base station in B
AC
b Active status of base station b
AD
b Antenna used by base station b
AZ
b Azimuth of the antenna used by base station b
CL
b Cell of base station b
CH
jk
b The k
th
channel used by base station
j
b
DM
b Demand (number of channels) required by base station b
ST
b Site occupied by base station b
TL
b Tilt of the antenna used by base station b
TX
b Transmission power used by base station b
C Set of all channels available in N

XI
k
c
k
th
channel in C
D Set of all antenna types available in N
OM
D Set of omni-directional antennae
m
d
m
th
antenna in D
( )
HG
d
Horizontal loss function for antenna type d
( )
VG
d
Vertical loss function for antenna type d
( )
CS
G
Function describing signal attenuation due to channel separation
PG
il
g Propagation loss from site
l
s to test point
i
r
FZ
g
Fuzzy server threshold gain constant
ND
I Total non-dominance
OV
I Total overlap
SF
I Total signal freedom
GP
l Length of side of each grid point in the network working area
N Mobile telephone network
( ) t N
TR

Number of channels required to accommodate call traffic t
AC
n Number of active base stations
BSPERSITE
n Maximum number of base stations per site
CHANPERBS
n Maximum number of channels per base station
CV
n Number of covered test points in N
OC
n Number of occupied sites
PX
n Horizontal (east-west) resolution of network working area
PY
n Vertical (north-south) resolution of network working area
SV
n Number of serviced test points in N
TXSTEP
n Number of discrete power steps
FS
ij
p Signal strength at test point
i
r due to base station
j
b

XII
TXMAX
p
Maximum transmission power
TXMIN
p
Minimum transmission power
R Set of all test points in N
CV
R Set of covered test points in N
SV
R Set of serviced test points in N
i
r
i
th
test point in R

r CIR at test point r


BS
r Best server of test point r
ND
r Service non-dominance at test point r
OV
r Service overlap at test point r
SF
r Signal freedom at test point r
SQ
r Field strength service threshold of test point r
SS
r Fuzzy server set for test point r
TR
r Expected volume of call traffic in test point r
XY
r
Position vector of r within the working area of N
S Set of all sites available in N
OC
S Set of occupied sites in S
l
s
l
th
site in S
XY
s
Position vector of s within the working area of N
( ) n T
CH

Traffic carrying capacity of n channels
CV
t Total volume of call traffic covered in N
SV
t Total volume of call traffic serviced in N
CV
w Weighted test point cover

1
Chapter 1

Introduction
An effective and efficient mobile telephone
1
service relies, in part, on the deployment
and configuration of network infrastructure. This thesis tackles some of the problems
faced by the network planners in providing that service. Novel algorithmic methods are
presented for automatically designing networks that meet both the subscribers demand
for a high quality service and the operators requirement for low infrastructure
overheads. Automatic Cell Planning (ACP) is the generic name for these network
optimising techniques.
The structure of this chapter is as follows: Section 1.1 presents an outline of the problem
domain considered, Section 1.2 contains a review of resources and research related to
the problem domain, Section 1.3 sets out the aims and objectives of the thesis, and
Section 1.4 summarises the contents of each subsequent thesis chapter.
1.1 Outline of Problem
The growth of the mobile telephone industry has been enormous. In the early 1980s
cellular
2
mobile telephones were introduced; by the Year 1990 the number of
subscribers worldwide was 11 million, and by the Year 2000 that figure had risen to 500

1
A telephone with a wireless connection to its network.
2
A technology where services are provided in discrete geographical-regions, known as cells.

2
million. This growth shows no sign of slowing, and it is estimated that by 2010 the
number of mobile-telephone subscribers will be 2 billion when it will exceed the
number of subscribers to wireline telephones
1
[1]. In order to keep pace with the
enormous increase in demand, faster and more efficient methods are required for the
design of mobile telephone networks [2,3].
Wireless communication requires sophisticated technology. Consider GSM
2
networks,
they employ advanced devices such as Very High Frequency (VHF) radio transceivers,
high-speed digital signal processors, optical fibre, microwave dishes, and data
compression. Until recently the cost of using mobile telephone networks has been too
high for the majority of consumers, however, advances in integrated-circuit design
(amongst other things) have reduced these costs sufficiently to make mass-market
wireless communications viable.

Base
Transceiver
Station
Base
Station
Controller
Mobile
Switching
Centre
Public
Land
Network
VHF
Radio
Waves
Microwave
Radio Link
Optical
Fibre
Optical
Fibre
Copper
Wire
Wireline
Telephone
Wireless
Telephone

Figure 1.1: The simplified path of a call between a wireless telephone and a wireline telephone.

1
Telephones that are connected to their network by wire.
2
Global System for Mobile communications (formerly Global System Mobil) - the European mobile
standard.

3
Consider the simplified example, shown in Figure 1.1, of a call between a wireless
telephone and a wireline telephone (this example is adapted from [4]). The call passes
through a variety of network switches. Speech data leaving the wireline telephone
travels as electrical impulses through copper wire; these impulses are digitally encoded
by the Public Land Network (PLN)
1
and transmitted, via optical fibre, to the wireless
subscribers Mobile Switching Centre (MSC). The MSC determines the destination of
the data
2
and re-transmits it to the correct Base Station Controller (BSC), and then the
optically encoded data is decoded and transmitted via a microwave link to the Base
Transceiver Station (BTS) currently hosting the wireless telephone. The BTS identifies
which of the wireless telephones in its service area (known as a cell) is the intended
recipient, and forwards the data using a VHF radio link.

BTS
MSC
PLN
PLN
MSC
BSC
BSC
BSC
BTS
BTS
BTS
BTS

Figure 1.2: An example of different switching technologies connecting both wireless and wireline
telephones.
The worldwide wireless and wireline telephone networks contain millions of
communication switches, and care is required in their deployment. Figure 1.2 shows an

1
The conversion from analogue to digital may happen before the data enters the telephone exchange.
2
The telephone number of the recipient will determine the destination of the data; the number will be
used as an index to a dynamic database of subscriber locations maintained by the network.

4
example of the connectivity in a simplified communications network. Both wireless and
wireline equipment is connected together using switches, each of which has a maximum
call carrying capacity. In addition to creating an effective network (one which can
handle all the call traffic) the network designer must strive to minimise infrastructure
costs. Striking the correct balance between these competing goals takes skill and effort
both of which are expensive commodities in themselves.
1.1.1 The Capacitated Network Design Problem
All communications switches have a limited call carrying capacity at any one time
each switch may act as a conduit for only a finite number of telephone calls. Although
this maximum capacity varies widely for different technologies, it is a fundamental
limitation in the design of all communication systems (see [5] for a more detailed
explanation).
In abstraction, the Capacitated Network Design Problem
1
(CNDP) involves the
provision of communication conduits between pairs of users
2
(see [6] for a generic
problem formulation). A network with insufficient, or poorly deployed, switching
elements will not be able to transport all of the required call traffic, which will lead to
lost revenue and customer dissatisfaction. Conversely, a network that deployed too
many switches would have a high running cost and, hence, reduced profitability. A
network designer must also take account of many other issues including fault-tolerance
(by including redundant switching elements), call delay from long wires, and switch
site availability. These problems tend to be technology specific (and hence, network
specific) so are often ignored in the study of pure CNDPs.
Note that the CNDP is similar to the Broadcast Network Design Problem
3
(BNDP), in
which a large number of users receive identical data. However, solutions for the BNDP
often exploit economies of scale that are not present in the CNDP, and separate

1
Also known as the point to point communication problem.
2
Communication may involve more than two users, but these party calls are not common.
3
Also known as the point to multi-point communication problem.

5
solutions must be sought for this problem (this is discussed in more detail in Section
1.2.3.1).
1.1.2 The Wireless Capacitated Network Design Problem
Wireless communication has two main advantages over wireline communication:
Mobility: a wireless user is freer to move around than their wireline counterpart.
This is even more relevant in the case of multiple mobile-users, where wires would
soon get tangled.
Economy: in a wireline network a significant portion of the wires are those that
terminate with the user. For example, from the switch on a telegraph pole to a
subscribers house (commonly known as the last mile problem). Using a wireless
connection for this portion of the network gives a significant saving in cable
installation
1
.
Conversely, wireless communication has two main disadvantages compared with
wireline communication:
Poor signal propagation: a signals strength and quality will deteriorate more for
wireless transmission than for wireline transmission over the same distance. The
wire acts as a wave guide for the signal, allowing precise propagation control.
Poor noise immunity: wireline signals are shielded from noise sources, whereas
wireless signals may be easily contaminated by stray radio waves.

1
It is worth noting that the cost of laying cable is usually significantly greater than the cost of the cable
itself.

6
Transmitters Transmitters
a) Wireless transmission b) Wireline transmission

Figure 1.3: A pictorial illustration of radio signal propagation in wireless and wireline networks.
Figure 1.3 shows the contrast in noise immunity between the two types of network. In
both cases there are three telephones (red, green and blue) receiving information from
three respective transmitters (shown as filled circles). In the wireline network,
telephones only receive a signal from the desired transmitter, whereas in the wireless
network each telephone receives additional signals from interfering transmitters. These
unwanted signals will cause the quality of the desired signal to degrade.
To take advantage of the benefits of wireless communication, the disadvantages must be
addressed in both the design of the technology (e.g. high quality receivers) and in its
deployment. In the Wireless Capacitated Network Design Problem (WCNDP) the
particular deficiencies inherent in deploying wireless technology are addressed,
principally:
1. The reliable propagation of data between wireless users and network base stations.
2. The management of the radio spectrum to avoid interference between wireless users.

7
The solution of the WCNDP requires the location and configuration of network base
stations such that mobile subscriber demand is satisfied. This must be done in such a
way as to minimise the cost of deploying and running the network.
A key factor in creating high quality solutions to the WCNDP is the management of the
available radio spectrum. The restrictions imposed on base station deployment will vary
between different technologies, but limited spectrum availability remains a problem in
all wireless networks. The division of the spectrum (known as the multiple access
scheme) might be frequency based, as in Frequency Division Multiple Access (FDMA),
or it might be code based, as in Code Division Multiple Access (CDMA) (see [7] for
further information). Regardless of the multiple access scheme used, the same piece
1

of radio spectrum cannot be re-used in the same place at the same time, as this would
result in interference.
1.2 Previous Work
This section contains an overview of previous approaches to solving the WCNDP, and
solutions to related problems are also included.
The exact formulation of the WCNDP is technology dependent and, as such, a variety of
models have arisen in which different aspects of the problem are emphasised. This has
led to the development of many different design techniques. The heterogeneity of
network models makes a quantitative comparison of design techniques impossible, and
therefore all comparisons must be made on a qualitative basis.
Some of the previous research (such as that on the problem of channel assignment)
relates only to a small part of the WCNDP. In this section these subproblems are
discussed and placed in the wider context of the WCNDP.

1
A channel in a FDMA scheme or a code in a CDMA scheme.

8
1.2.1 General Background Reading
A good introduction to radio communication can be found in [7]. More specific
information on mobile wireless communications can be found in [8-10], and a highly
accessible overview of the issues involved in designing wireless networks is given in
[11].
Information on the practical application of combinatorial optimisation can be found in
[12,13], and valuable insights into the theoretical aspects of combinatorial optimisation
can be found in [14,15]. An overview of the application of combinatorial optimisation
to problems in the field of telecommunications is given in [16].
1.2.2 Channel Assignment
The Channel Assignment Problem
1
(CAP) is a well studied area of wireless network
design, and it is a sub-problem of the WCNDP (as described in Section 1.1.2). In the
CAP, only the channels assigned to base stations may vary, and all other network design
parameters (such as antenna selection or transmitter placement) are fixed. This section
contains a brief overview of the CAP, but a more complete taxonomy of problems and
solutions can be found in [17-19].
In the CAP a network is assumed to consist of a set of transmitters (TXs) and receivers
(RXs), where each RX requires an interference free signal from one or more TX. The
strength of the radio signal from each TX to each RX is known. The interference each
RX experiences is dependent on the separation between the channel used by the serving
TX and the channels used by interfering TXs the greater the separation (in channel
space), the smaller the contribution of that signal to the total interference.
Given an unlimited radio spectrum, channel assignments could all be unique and
generously spaced, but unfortunately only a limited range of frequencies are suitable for
mobile communication, and it is smaller than the total number of channels required

1
Also known as the Frequency Assignment Problem (FAP).

9
worldwide. To maximise spectrum availability, portions of the spectrum must be reused.
This is only possible when RXs that share a common channel (or use channels that are
close enough to interfere) have sufficient electromagnetic isolation from one another.
As the range of wireless communication signals is small, typically line-of-site
1
, this is
not an insurmountable problem. Furthermore, in cities (where the demand for channels
is at its highest) the buildings provide RXs with extra shielding from interfering TXs,
and this allows the network designer more precise control over interference (an
observation made in [11]).
1.2.2.1 Fixed and Variable Spectrum
The CAP is usually stated in one of two forms: as a Variable Span Problem (VSP) or as
a Fixed Span Problem (FSP). The solution of the VSP requires all the RXs to have a
usable signal (i.e. to have acceptably low interference) whilst the difference between the
lowest and highest channels used (known as the span of the assignment) is minimised.
The solution of the FSP requires the maximum number of RXs to have a usable signal
given a fixed span of channels from which assignments may be made.
Network operators bidding for new spectrum licenses might adopt the VSP, as they
would need to know how much spectrum to buy. Another application might be in the
day-to-day allocation of spectrum in a theatre-of-war, as the reliance on radio
communications in modern warfare means that efficient spectrum management is
crucial.
Instances of the FSP are much more common than those of the VSP. The biggest users
of radio spectrum are commercial network operators (such as broadcasters, mobile
telephone companies, and taxi firms) and they usually have exclusive rights to a
particular portion of the radio spectrum
2
. Commercial networks typically have a

1
Some low-frequency radio signals have a much further range. However, high frequency signals (which
are most useful for high bandwidth communications) have a short range.
2
This may vary with location, for instance two taxi firms might use the same part of the spectrum but in
different cities.

10
consumer base of subscribers (or potential subscribers) who require interference free
radio-communication either one-way (e.g. TV broadcasts) or two-ways (e.g. mobile
telephone calls); they wish to maximise the number of subscribers covered (and hence
revenue) using the spectrum they have available. It will be assumed that all CAPs are
FSPs for the remainder of this thesis.
1.2.2.2 Geometric Solutions
One approach to the CAP has been the division of TX cells into discrete geometric
regions. Fundamental to this approach is the concept of re-use distance; this is the
minimum separation (in space) between cells that use the same channel or set of
channels. The ratio of re-use distance (D) to cell size (R) gives an indication of the
amount of interference expected; this is illustrated in Figure 1.4.

D
R
Co-channel Cells

Figure 1.4: The relationship between re-use distance D and cell radius R in co-channel cells.
The available spectrum is divided into N distinct sets, and this division will dictate the
acceptable D/R ratio. From this relationship the network designer can estimate the trade
off between capacity and interference. Cell plans for different D/R ratios may be created
using hexagonal tessellations (Figure 1.5 shows the re-use pattern for N = 7).

11

D

Figure 1.5: A channel re-use pattern for 7 different sets of channels.
For large areas with a near-uniform subscriber distribution, this approach is fast and
effective. If subscriber density varies then cell size and capacity must vary, and this
leads to a disjointed tessellation, which is difficult to analyse using geometry. For a
more in depth treatment of the geometric approach see [9]. Regular geometric shapes do
not bound real cells, and this simplification leads to inaccuracies in the models
prediction of interference (see [20] for further analysis of the shortcomings of this
approach).
1.2.2.3 Graph Colouring
As Figure 1.5 illustrates, the assignment of channels to cells has a similarity with the
problem of colouring political maps no adjacent regions may have the same colour. In
political maps this is to ensure that the contours of adjacent countries can be clearly
distinguished, but in cell plans it is to ensure that adjacent cells do not interfere with one
another.

12

Scotland
The North
North West
England
Wales
The South
West
The West
Midlands
Yorkshire &
Humberside
The East
Midlands
The
South
East
East
Anglia

Figure 1.6: The translation of a political map into a colouring constraint graph.
The general form of the political-map colouring problem is known as the Graph
Colouring Problem (GCP). Figure 1.6 shows the relationship between a political map
and a mathematical graph. Each region of the map becomes a vertex on the graph, and if
two regions are adjacent then an edge will exist between the corresponding vertices. A
colouring of the graph in which no two adjacent (i.e. edge sharing) vertices have the
same colour can be translated into a map colouring where no two adjacent regions share
a colour. Graphs that represent channel assignment constraints may be generated in the
same way, and this allows the migration of many results and techniques from graph
theory to the problem of channel assignment.
Translating a CAP into a GCP can be lengthy and complicated. Furthermore, the
method is usually specific to the type of wireless network in use; this leads to the
translation process being difficult to port between different types of network model.
The political-map colouring problem is a subset of the GCP, whereas the CAP is not a
subset of the GCP (there are many ways to create a GCP from a CAP). This has led to

13
variants of the GCP being adopted that better approximate the CAP. Examples include
the addition of weighting to the graph edges, as described in [21], and the generalisation
from binary graphs (graphs with only two vertices per edge) to hyper graphs (graphs
with many vertices per edge), as described in [18].
The advantage of adopting a graph theoretic approach lies in the wealth of techniques
already derived for the GCP. A good example of this is lower-bounding [22], which can
provide an estimate of the minimum number of distinct colours required to satisfy a
given graph.
1.2.2.4 Radio Signal Analysis
The ambiguity in translation between CAPs and GCPs means that even if the GCP is
solved to optimality, the corresponding CAP solution may not be optimal. A more exact
approach relies on the direct estimation of radio signal interference. Given a particular
channel assignment, the total interference each RX will experience can be estimated
based on measured or predicted signal strengths. Figure 1.7 shows an example of
received power levels for a single RX; in order to achieve a tolerable noise level, the
interfering signals must be small or sufficiently attenuated by channel separation.

Serving
Antenna
Interfering
Antennae
10W 2W
5W
1W

Figure 1.7: Some received power levels at a mobile station; this data may be measured or predicted.

14
This approach gives the most accurate representation of the CAP, as it takes account of
everything that can be either measured or predicted. Some studies of the effectiveness of
this approach have been done (examples include [23]), and it has been shown to deliver
the highest quality solutions of any of the channel assignment methods. The
computational requirements of this approach are high, and this has meant that less exact
methods (such as graph colouring) have dominated the creation of channel assignments.
On-going improvements in areas such as microprocessor speed, compiler technology,
and software engineering have increased the size of problem for which this technique
can deliver superior solutions; some studies have been made of the scalability of this
approach versus less exact methods (see [24]) but the results remain inconclusive.
1.2.3 Cell Planning
At the time of writing, there are many commercial Computer Aided Design (CAD)
systems for cell planning available; examples include ATOLL (by FORSK), Wizard (by
Agilent Technologies), deciBell Planner (by Northwood), and Asset (by AIRCOM). An
introduction to the issues involved in CAD for mobile telephone networks can be found
in [20,25]. All current CAD systems require a user to decide base station locations and
configurations; feedback about network performance (information such as area
coverage and infrastructure costs) comes from a computer simulation.
The elimination of the user from the design loop is a natural progression of the CAD
paradigm (as concluded in [20]). The user should be required to specify the desired
network performance, and not be required to specify the details of the network design
needed to achieve it. The following subsections describe research into the automation of
the network design process.
1.2.3.1 Automatic Base Station Placement
The Base Station Placement Problem (BSPP) is a subproblem of the WCNDP, and there
are two common formulations of this problem. In the first, known as uncapacitated, the
subscribers only require a usable (i.e. sufficiently strong) radio signal. In the second,
known as capacitated, base stations are limited by the maximum number of subscribers

15
they can serve. The uncapacitated problem (referred to as the BNDP in Section 1.1.1) is
typical of the requirements a television distribution network might have. The simplest
solution to this problem is typified by satellite television technology, in which one
transmitter (the satellite) can transmit information to an entire country. This type of
solution is not well suited to capacitated networks, where each subscriber requires
different data streams; in fact, a single-satellite solution would only allow the available
spectrum to be used once for the entire area covered by the satellite
1
.
Models for the BSPP are further divided in their approach to site availability. In some
previous work (such as [26]) a pure BSPP is formulated, in which base stations may be
placed anywhere within the working area of the network. Whereas in other work (such
as [27]) a subset BSPP is formulated, in which the set of candidate sites is decided by
the user and the computer selects a subset for occupation. The subset BSPP represents
the majority of practical problems, as network operators find it more efficient to select
sites based on their cost, availability, and practicality rather than on their suitability for
radio communications (see [3] for similar reasoning).
It is interesting to note that base station (or transmitter) placement in network design has
much in common with the Facility Location Problem (FLP) (see [28] for a
comprehensive description of the FLP).
Research on the BSPP for uncapacitated networks can be found in [29-32]; a common
goal in previous work has been the maximisation of cover (the number of receivers with
a sufficiently strong signal) coupled with the minimisation of infrastructure cost. The
application of Artificial Intelligence (AI) to the uncapacitated BSPP is explored in [33],
where an expert system is developed as a drop in replacement for the network
designer. In [34] graph theory is used to simplify the solution of the uncapacitated
BSPP; this work was part of the STORMS project [35]. Complexity analysis for a

1
It should be noted that this fundamental problem has not stopped commercial vendors utilising satellites
for mobile communications; although it is inefficient for densely populated areas it is well suited to sparse
ones. In this thesis, only densely populated areas are considered.

16
generic model of the uncapacitated BSPP is described in [36]. A comparison of heuristic
and meta-heuristic approaches can be found in [27], and the performance of a large
number of local-search algorithms is evaluated in [37]. A genetic algorithm is applied to
the uncapacitated BSPP in [38]. Interference is explicitly considered in [39], where the
co-channel interference experienced by a mobile receiver is calculated and used to give
an improved estimate of area coverage.
In [40] a small CDMA network is considered as a capacitated BSPP, where base
stations are located such that the predicted bit error rate of data from mobile subscribers
does not exceed a given threshold; a set covering approach is augmented with various
heuristics tailored to the problem. A broad approach to the problem is taken in [4],
where base station placement is included as part of the wider network design problem
(including the layout of the wireline backbone). In [41] the process of cell splitting
(dividing large macro-cells into smaller micro-cells) is automated; the goal being the
minimisation of call blocking probabilities (for new or handed-over calls) using the
minimum number of base stations.
1.2.3.2 Automatic Base Station Configuration
The Base Station Configuration Problem
1
(BSCP) is the selection of all operational
parameters of the base station except the assigned channels; these may include position,
transmission power, antenna type, or antenna down-tilt. The BSCP is a superset of the
BSPP (because it includes site selection) and a subset of the WCNDP (because it does
not include channel assignment).
Some work has been done on base station configuration using an uncapacitated network
model. However, the networks considered tend to be of a specific type and the solutions
derived may not necessarily be generic. A good example can be found in [42], which
assesses the advantage of deploying adaptive antenna arrays in an indoor environment.

1
Also known as the dimensioning problem or cell dimensioning.

17
The majority of work relating to the BSCP is based on a capacitated network model, and
some solutions to the general problem of base station configuration can be found in
[43,44]. In [45] the evolution of a network over time is considered, which is an
important concern in commercial networks where subscriber demand is constantly
changing and generally increasing. The connectivity of cell coverage is discussed in
[46]; cells that are irregular or fragmented have a detrimental effect on network
performance, specifically they make channel assignment more difficult and they
increase the probability of handover for mobile subscribers. In most of the network
models studied, only the transmission of a radio signal from the base station to the
mobile station (the downlink) is considered, and it is assumed that the uplink is either
easy to maintain or difficult to model. In [47] simulation of both the uplink and
downlink is considered, and its effect on the optimisation process discussed. The Pareto
optimality of the cell planning problem is discussed in [48]; this approach can provide
network designers with valuable insights into the intrinsic compromises that must be
made in solving a multi-objective problem such as cell planning.
Some work on the BSCP explicitly considers the channel assignment that follows the
completion of a cell plan. In [49] a simple network model is augmented with channel
assignment constraints. A novel approach to including channel assignment in the cell
planning process is used in [50], where lower bounding techniques (as described in
Section 1.2.2.3 and [22]) are used as an indicator of the difficulty of channel
assignment without actually producing a complete assignment. The most
comprehensive treatment of the full WCNP is detailed in [72], in which channels are
assigned to optimised cell plans, but the relationship between the cell plan and the
channel plan is not explored in detail.
1.2.4 Shortcomings of Previous Work
This subsection describes some of the shortcomings of previous approaches to the
WCNDP. It is important to separate the underlying network model, by which all
candidate network designs must be judged, from the optimisation algorithms employed
to create candidate designs. Much effort has been expended on the application of

18
combinatorial algorithms to classical problems, which, by definition, have a standard
formulation (see [51] for example). Optimisation algorithms can be tailored
1
to the
classical problems because the underlying evaluation function does not change, and this
allows assumptions to be made and shortcuts to be taken in the optimisation procedure.
However, the WCNDP has not reached a universally agreed formulation, and
assumptions may not be valid over the set of all formulations. This implies that the
tailoring of algorithms to a particular problem formulation may prove wasted effort
even if the formulation of the problem is altered only slightly. Worse still, the bias
introduced by a tailored algorithm could exclude sets of good solutions. In this thesis,
more attention is paid to the formulation of the problem than to the design of clever
optimisation techniques, and care has been taken to ensure that the optimisation
algorithms are generic; that is to say, they do not introduce bias by excluding possible
solutions and, hence, remain applicable to future formulations of the WCNDP. An
alternative approach would be to develop a simpler formulation of the WCNDP and
search for the ideal optimisation algorithm with which to solve it. This ideal algorithm
would be highly specific to the problem formulation and, hence, potentially inapplicable
to a wider set of problems, furthermore the particular WCNDP used may not be
sufficiently complex to be relevant to industry.
This thesis builds upon the model of a GSM network described in [2]. This model was
developed by industry, and is known to be used by at least two major European mobile
telephone network operators. The high availability of data (problem instances) for this
model also favoured its adoption. Other authors have extended this basic model to take
account of a variety of different problems (for example [45] describes a network roll-
out scenario), and in this thesis an interference model was added, which considered
channel assignments. Channel assignment has been considered in the wider context of
cell planning (in work such as [49] and [50]) but only simple models were used; no

1
Usually manifest as embedded heuristics that can reject possible solutions as inferior without resorting
to a (computationally intensive) full evaluation of the solutions performance.

19
work to date had explicitly considered the relationship between cell planning and
channel assignment. This significant omission was addressed in this thesis.
1.3 Thesis Aims and Objectives
The aim of this thesis was the development of techniques for solving the WCNDP, and
this general aim was divided into two objectives that were further divided into specific
tasks. The completion of tasks and objectives is explicitly mentioned in the thesis text,
and they are enumerated here to facilitate those references.
1.3.1 Objective O1: Statement of Problem
Previous formulations of the WCNDP do not completely represent the requirements of
the mobile telephone network designer, and an objective of this thesis was the
formulation of a more comprehensive WCNDP. The following tasks were completed in
order to meet this objective:
Task T1: Acquisition of industrial data for the simulation of mobile telephone
networks. The quality of the data obtained dictates the accuracy of the
simulations and, hence, the relevance of the network designs produced.
Task T2: Mathematical specification of the simulator. To communicate the details of
the network simulator concisely and without ambiguity a mathematical
description was derived. Without a clear description the model could not be
effectively criticised or accurately reproduced.
Task T3: Specification of network performance requirements. It was necessary to
identify a set of objectives that each network design must fulfil, as this is the
mechanism by which one candidate design may be judged against another.
Although such a specification of requirements is typical in the engineering
process, automatic design techniques require that it be expressed
algorithmically.

20
1.3.2 Objective O2: Development of Solutions
The formulation of the WCNDP derived to meet O1 provided a test-bed for the
development of optimisation algorithms. The effective application of automatic
optimisation to the WCNDP is broken down into the following stages:
Task T1: Simple optimisation algorithms were implemented and applied to the
WCNDP. The validity of this approach was demonstrated by the production
of high quality network designs in a reasonable time.
Task T2: Assessment of the advantages of separating the CAP from the rest of the
WCNDP. In Section 1.2.4 it was observed that the CAP is rarely considered
as part of the WCNDP, and channel assignment (if required) is only done
when all other parameters have been fixed.
Task T3: The optimisation procedures that produced the best network designs were
identified. Finding the best optimisation algorithm was not the principal
goal of this thesis (as observed in Section 1.2.4), but it was still of interest.
Task T4: The specification of improvements that may be desirable to both the
network model and the optimisation methodology.
1.4 Test Platform
When assessing the performance of an optimisation algorithm, the best solution
produced must be placed in context by the amount of computer power that was required
to produce it. It is not easy to accurately describe computer power by any single figure
(such as the number of floating point operations performed) as it depends on many
factors (such as memory size and hard disk access speed). All of the results generated in
this thesis were done using the same computer operating under the same conditions, as
this approach eliminates (as far as possible) the distortion of results due to platform
heterogeneity. Table 1.1 contains all of the defining characteristics of the test platform,
which corresponds to best practice as identified in [52].

21
Manufacturer Intel
Model SE440BX-2 Motherboard
Bus Speed 100MHz
Manufacturer Intel
Model P III Processor
Speed 700 MHz
Size 640 MB
Memory
Speed 100 MHz
Manufacturer Fujitsu
Model MPD3084AT
Size 84 GB
Bus EISA ATA-3
Hard Disk
Max Transfer Speed 666 MB/s
Vendor Microsoft
Name Windows NT
Version 4.00.1381 + Service Pack 6
Operating System
File System NTFS
Vendor Microsoft
Name Visual C++ Professional Edition
Version 6.0 with SP 4
Language Used C++
Compile Type Release
Compiler
Additional Flags /G6 /MD /GX /Ox /Ot /Og /Oi /Op /Ob2
Table 1.1: Test platform specification.
It is important to note that all of the data sets used in this thesis fitted within the
(relatively large) RAM of the test platform. Therefore, the experiment run times were
not hampered by overspills into virtual memory, which cause the hard disk to thrash.
1.5 Thesis Outline
This thesis is divided into eight chapters; this section contains a summary of each
chapters contents.
Chapter 2 describes the development of a simulation environment for modelling
mobile telephone networks, and it includes a description of data obtained from leading
industrial sources. The specific goals of the WCNDP are stated in this chapter.
Chapter 3 presents the results of applying combinatorial optimisation techniques to the
WCNDP. A number of simple techniques are described, tested, and compared.
Chapter 4 describes new approaches to the CAP (part of the WCNDP). The traditional
form in which this problem has been expressed is challenged, and a more realistic

22
approach is described that aims to better represent the problem of spectrum
management.
Chapter 5 describes interference surrogates and evaluates them with respect to their
ability to predict subscriber cover.
Chapter 6 contains a comparison of the approach described in Chapter 3 to the two-
stage process discussed in Chapter 5.
Chapter 7 contains the conclusions that may be drawn from the thesis results, and
suggestions for future research on the WCNDP (including improvements to the design
process, and extensions of the underlying network model).

23
Chapter 2

Network Model
This chapter describes a mathematical model of a cellular mobile-telephone network. It
is based around, but not exclusive to, GSM. The model was developed for the purpose
of this thesis, and it is intended to represent industrial models currently in use. It takes
account of network coverage and capacity requirements, and it also considers the effect
that channel assignment has on interference. The Wireless Capacitated Network Design
Problem (WCNDP), as outlined in Section 1.1.2, is formulated in this chapter.
The structure of this chapter is as follows: Section 2.1 explains some of the issues raised
by the specification of the model, Section 2.2 contains a description of the actual model
used, and Section 2.3 contains details of the network data sets used in this thesis.
2.1 Simulation Issues
This section contains an explanation of the key design decisions made during the
specification of the network simulator. All the data sets used have been obtained from
companies that operate mobile telephone networks, and this has meant that some design
decisions are inherited and, therefore, inflexible.
2.1.1 Limitations on Prediction Accuracy
Some aspects of mobile telephone networks are impossible to model exactly, for
example it is not possible to say at what times a particular subscriber will make or

24
receive calls. Quantities such as this may only be modelled on a probabilistic or
statistical basis, e.g. 80% of telephone calls will be made between the hours of 9am and
5pm. The prediction of network performance (the models output) is limited by the
accuracy of network data (the models inputs), but it is not easy to quantify the effect
that errors in the models input have on the models output.
2.1.2 Discretisation
For networks to be simulated using a computer, the data which defines them must be
discretised (or digitised). Some network data is inherently discrete, e.g. the number of
base stations, whereas other data is continuous (or analogue), e.g. the azimuth of a
directional antenna. Discretisation of an analogue quantity introduces inaccuracies (or
noise). All of the data used in this thesis had been prepared by network operators, and
most decisions regarding discretisation (such as the resolution of the terrain model)
cannot, therefore, be changed.
2.1.3 Rasterisation
A raster is a discrete, regular, and rectangular grid of data. Typically it is two-
dimensional (e.g. the picture elements, or pixels, in a television picture). A co-ordinate
vector can uniquely identify the data elements in the grid, and the datum at a particular
grid location may be digital or analogue. The process of converting data into a raster
format is known as rasterisation or, sometimes, rendering.
The model used for a mobile telephone network brings together data from a variety of
sources. This data does not necessarily conform to a common standard, which makes
interoperability difficult. Rasterisation facilitates interoperability between data sources
such as:
The terrain databases derived from topographical relief maps.
The vector based building models used for radio propagation prediction in urban
areas.

25
The call traffic density information. This is sometimes derived from postcode
information, which describes discrete regions but is not constrained by a raster grid.
Not all data needs to be rasterised, for instance the location of base station sites can be
described as a free-vector within the network region without conflicting with other data
formats.
2.1.4 Spatial Dimensionality
The model used is essentially two-dimensional. Although some data, such as
propagation information, may have been derived from three-dimensional models, the
result is ultimately amalgamated into cells of a two-dimensional raster. The choice of a
two-dimensional model limits the accuracy of prediction, and this is especially true in
dense, urban areas where calls may occur in tall buildings. The advantage of two-
dimensional models is in the speed of evaluation.
2.1.5 Call Traffic Prediction
A statistical approach is used to assess the number of subscribers covered. The call
traffic (number of calls per hour) expected during the busy hour is amalgamated into a
single figure (measured in Erlangs) for each geographic area. This is a typical,
conservative, approach to network capacity planning, i.e. it is implicitly assumed that a
network that can accommodate the peak traffic will be able to accommodate traffic for
all other times.
2.1.6 Uplink and Downlink
The network model only simulates the link from the base station to the mobile station
(known as the downlink or forward link), and the uplink is ignored. In general, the
quality of the circuitry in the base stations is higher than in the mobiles, and hence the
uplink signal can be more easily distinguished than that of the downlink (making the
downlink the limiting factor). The computational complexity of uplink simulation is
prohibitive, and the decision to ignore uplinks is inherited from the network operators
(i.e. they do not consider the uplink in the majority of their network simulations). In this

26
thesis, the simulation techniques used by operators have been adopted in order to
maximise the industrial relevance of the optimisation algorithms developed, but it is
worth noting that uplink is considered to be a limiting factor in CDMA networks and
may be explicitly considered during the optimisation process (see [40] for an example).
2.1.7 Fixed Base Station Sites
Base station site rental represents a significant portion of the cost of running a network,
typically the sites desired by network designers (based on their suitability for wireless
communications) are not always easy to obtain and may be subject to planning
regulations. A strategy popular with network operators has been to make site-acquisition
largely independent of the network design requirements, as this allows the identification
of a large number of potential sites that are easy to obtain (short lead-time) and
relatively cheap to occupy in the network model these are known as candidate sites.
This approach increases the burden on the network design process, as undesirable sites
(such as electricity pylons) must be more intelligently utilised than those that are highly
desirable (such as the Eiffel Tower).
2.1.8 Radio-Propagation Prediction
The radio signal propagation in the network area was precalculated and provided as part
of the network data set. The prediction algorithms used in the network model are based
on the COST231 Walfish-Ikegami models (see [53] for further information), but these
propagation algorithms are computationally intensive. The fast simulation of a network
relies on the fact that (potentially slow) propagation predictions can be precomputed,
and this is possible for the following reasons:
Only the downlink is simulated. This means that (potentially complex) uplink
propagation predictions can be avoided.
Base station sites are fixed. Downlink propagation predictions are made from each
of the candidate sites to each of the possible mobile station locations.

27
The propagation calculation can be performed independently of most base station
parameters. The relative (normalised) propagation gain can be included in the
calculation of signal strength as an independent factor.
Note, base station height does affect the relative propagation, however, the data sets
obtained for this thesis do not permit this parameter to vary. If height were desired to be
a variable then each different height could be treated as a different site, as this would
allow the pre-computation of propagation.
The frequency of a radio signal also affects its propagation, but across the range of
permitted frequencies the variation in propagation will be small, therefore the mean
frequency is used for the propagation calculation.
2.1.9 Best Server
When modelling the behaviour of base stations in a network it is necessary to know how
much call traffic they will be serving. This allows an estimate to be made of the number
of channels each base station will require (see Section 2.1.5). The best server of a
mobile station is the base station from which it receives the strongest radio signal, and it
is assumed that a mobile station will always make a call through its best server. This
implies that the set of all the mobile stations for which a base station is best server
gives a prediction of that base stations traffic-carrying requirement. The best server
model gives a conservative estimate of a networks traffic capacity because each base
station can only accommodate a fixed amount of traffic, and any traffic above that limit
must be considered lost. In the best server model this lost traffic is not allowed to be
covered by other base stations even if they offer interference-free communication
channels. A more accurate estimate of traffic cover can be achieved with a more
sophisticated model, but its disadvantages make best server the most sensible model to
adopt. These disadvantages are summarised in the following subsection, but they are not
central to the thesis and may be skipped.

28
2.1.9.1 The Multiple Server Problem
For the mobile telephone network model used in the thesis, only one base station is a
valid server for each test point
1
, whereas in the GSM standard up to six base stations
may be considered as valid servers. In practice, if a server were to reach its traffic
carrying capacity, then a mobile subscriber could connect to the network through an
alternate server. The conservative assumption in the thesis model means that the
predicted traffic service will always be less than or equal to the traffic service that could
occur in practice.
A subtle problem was encountered when attempting to extend the thesis model to
include the possibility of multiple servers, and no evaluation algorithm could be found
that did not have a worst case time-complexity (see [54] for background) that grew
exponentially with the number of potential servers.

Cell
A
Cell
B
Cell
C
r
6
r
5
r
7
r
1
r
3
r
4
r
2
r
8
r
9

Figure 2.1: An abstracted traffic cover problem using multiple servers.
Figure 2.1 presents an abstraction of the cell plan for 3 base stations A, B, and C, which
form cells Cell
A
, Cell
B
, and Cell
C
respectively. Cell
A
contains test points r
1
to r
6
and

1
Namely the base station that produces the highest field strength at that point.

29
Cell
B
contains test points r
4
to r
8
, in a multiple server model the points in the
intersection of Cell
A
and Cell
B
may be served by either base station A or base station B.
Assume that each test point represents 10 Er of traffic and that each base station may
serve up to 30 Er of traffic, therefore each base station may serve a maximum of 3 test
points. By inspection it can be seen that if base station A serves r
1
, r
2
, and r
3
, base
station B serves r
4
, r
5
, and r
7
, and base station C serves r
6
, r
8
, and r
9
then full coverage
can be achieved (i.e. 3 test points per base station). The problem lies in identifying
which points to allocate to which base stations; consider the evaluation procedure below
that allocates points to base stations in the (arbitrary) order they have been defined:
1. Points r
1
, r
2
, and r
3
are allocated to base station A.
2. Points r
4
, r
5
, and r
6
are allocated to base station B.
3. Point r
7
could only have gone to base station B, but B has already reached capacity
and hence it must be discarded
4. Points r
8
and r
9
are allocated to base station C.
Due to the discarding of test point r
7
, the total traffic coverage will be 10 Er less than
the potential maximum (found previously by inspection).
A variety of approaches were tried to overcome this problem but no suitable algorithm
was discovered, all heuristic based approaches tested (for example, ordering test points
by number of potential servers) were found to be defeated by pathological cases. An
algorithm based on hypergraphs was devised that guaranteed delivery of the correct
traffic coverage, however, analysis showed it to have a worst case time complexity that
increased exponentially with the number of potential servers, and testing proved this led
to long and unpredictable evaluation times. The performance of this algorithm was
compared to the simple algorithm employed in the thesis, and in no test case did the
traffic prediction from the best server model fall more than 5% below the actual
(multiple server model) value. This was felt to be an acceptable inaccuracy, especially

30
as the estimate was always conservative. Note that this problem can only occur in
situations where maximum base station capacity is reached by a base station, if all base
stations are operating below capacity then the best server prediction will be no less
accurate.
2.1.10 Fuzzy Servers
Although the best server assumption is essential in determining a base stations channel
requirements, it is not safe to assume that mobile stations will communicate exclusively
through their best server. The best server of a mobile station may not be the one
predicted by the simulator, due, for example, to inaccuracies in the prediction of radio-
signal strength. To decrease the chance of these server ambiguities creating unforeseen
interference, a set of potential servers is identified. This comprises base stations that
provide the mobile station with a sufficiently strong signal to capture it from its
(predicted) best server. The set of potential servers is known as the fuzzy-server set, and
it includes all the base stations that provide the mobile station with a signal within a
fixed tolerance of the best-server signal. This tolerance corresponds, in part, to the
cumulative error expected in the predicted radio-signal strengths.
2.1.11 Fixed Spectrum
A network operator will typically lease the use of a fixed portion of the radio spectrum,
and once the spectrum is leased there is no benefit in not using all of it. The wider the
portion of spectrum available, the easier it is to minimise signal interference. In the
network model it is assumed that a fixed spectrum is available for communication.
Section 1.2.2.1 contains a summary of the fixed and variable approaches to spectrum
management.
2.1.12 Handover
Handover (or handoff) is the transition of a mobile from one server to another. This is
most usually between adjacent base stations in a network, and occurs as the user passes
out of one cell and into another thus ensuring an uninterrupted service for mobile

31
users. Base stations have the ability to change communication channels mid-call, and
these changes may also be refered to as handover.
The network model takes no account of handover performance. Handover models
have been proposed (examples include [2]) but their accuracy is not well defined.
Suggestions for the addition of handover performance models are included in Section
7.2.1.
2.2 Description of Model
The model used for the simulation of mobile-telephone networks is an extension of the
one used in [2]. Network data sets that complement this model have been obtained from
two network operators: four data sets from Operator A and one data set from Operator B
(this corresponds to Task T1 of Objective O1 in Section 1.3). In the original model,
interference was not explicitly calculated, but in the new model it was decided to
include the assignment of base station channels, thereby allowing the simulation of
signal interference. The following thesis subsections describe all the elements of the
network model (this corresponds to Task T2 of Objective O1 in Section 1.3).
2.2.1 Summary of Data Set
The data sets for the network model consist of the following elements:
Network working area boundary.
Subscriber service and demand requirement maps.
Candidate sites for base station placement, including radio propagation information.
Base station antenna types.
Radio spectrum availabilty and characteristics.
The following subsections describe these elements in more detail.

32
2.2.2 Working Area
A network, N, contains a working area, and points in the working area may be described
by their x and y co-ordinates. Two types of point are located on the working area:
Points at which a wireless service is required (known as test points)
Candidate (or potential) sites where base stations may be positioned.
The test point data forms a two-dimensional raster, where the resolution of the raster
grid along the x-axis (east-west) and y-axis (north-south) is given by
PX
n and
PY
n
respectively. The size of each test point is the same for both the x and y axes, and is
denoted by
GP
l .

1
1
2
2
n
PY
l
GP
l
GP
n
PX

Figure 2.2: Rasterisation of the network working area.

33
2.2.3 Test Points
Network N contains a finite set of test points
1
R, where
i
r denotes the i
th
test point in R.
Test points discretely approximate a continuous subsection of the working area. A test
point, r , is uniquely defined by its position vector
XY
r .

Figure 2.3: The test point layout in NetworkA3.
Figure 2.3 shows an example set of test points; the black border defines the extent of the
working area, and test points are coloured light blue.
2.2.3.1 Traffic Volume
Telephone calls may arise
2
in the area represented by test point r , and the expected
volume of call traffic is given by
TR
r (measured in Erlangs). Some test points are not
expected to generate a significant volume of call traffic; hence the corresponding
TR
r
will be zero.

1
The term test point is slightly misleading as each point actually represents an area. The name is a
legacy of previous work.
2
Either instantiated or received by a subscriber.

34

Figure 2.4: The test point traffic volumes in NetworkA3.
Figure 2.4 shows an example of test point traffic volumes in a network. Test points with
no traffic are coloured light blue, and test points with traffic are coloured red in
proportion to the traffic volume (light red indicates low traffic volume and dark red
indicates high traffic volume).
2.2.3.2 Service Threshold
A test point, r , has a minimum field-strength requirement
SQ
r known as its service
threshold. If a test point receives a radio signal (from at least one base station) that is
greater than or equal to this threshold, it is said to be serviced.

Figure 2.5: The test point service thresholds in NetworkB1.

35
Figure 2.5 shows an example of the distribution of service thresholds in a network; dark
areas contain test points with high service thresholds and light areas contain test points
with low service thresholds. The pixel-intensity scale used in the diagram is
logarithmically proportional to the threshold value.
2.2.4 Candidate Sites
Candidate sites are the positions at which base stations may be placed. Network N
contains a finite set of candidate sites S, where
l
s denotes the l
th
site in S. Site s is
defined uniquely by its position vector
XY
s .

Figure 2.6: The site positions in NetworkB1.
Figure 2.6 shows an example of site positions in a network. Sites are shown as dark blue
squares with white centres.
2.2.5 Antenna Types
Network N contains definitions of two distinct classes of antenna: directional and omni-
directional. Antennae of both types are members of the set D, where
m
d denotes the m
th

antenna in D. An antenna is characterised by two diagrams: one tabulating the
horizontal loss and the other tabulating the vertical loss, and both diagrams are defined
between 0 and 359 with a resolution of 1.

36
2.2.5.1 Antenna Constraint
The following constraint exists between the two classes of antenna: if a site
accommodates a base station with an omni-directional antenna, it may
accommodate no more base stations. This represents a hard-constraint on the network
design, but it is enforced as a soft-constraint, i.e. network designs that break this rule are
valid but are always considered to be inferior to network designs that do not. This
approach makes the design of optimisation algorithms easier, as all possible network
designs are valid. However, no final design is accepted if it breaks the antenna
constraint.
2.2.5.2 Horizontal Loss
The horizontal loss of antenna d is characterised by the function ( )
RA HG
h d , where
RA
h
is the angle, projected onto the horizontal plane, between a vector from the antenna to
the receiver and a vector along which the antenna is pointing.

North (0)
Antenna
Direction
RA
h
s
RN
h
XY XY
s r
AZ
a
r

Figure 2.7: The geometry of the horizontal loss calculation for directional antennae.
Consider a directional antenna d at site s (as shown in Figure 2.7), the antenna is
pointing at an angle
AZ
a from true north (known as its azimuth). To calculate the
horizontal component of the antenna loss experienced by test point r it is necessary to
compute the angle
RA
h :
AZ RN RA
a h h =

37
where
RN
h is the angle, relative to north, of a vector to r :
( )
XY XY RN
s r h = arctan
where ( ) arctan is a function that returns the vector-safe inverse tangent
1
.
Note that function ( ) a d
HG
is defined as a look-up table with the angle given in degrees
between 0 and 359, therefore the angle calculation is ultimately rounded to the nearest
degree.
2.2.5.3 Vertical Loss
The vertical loss of antenna d is characterised by the function ( )
RA VG
v d , where
RA
v is
the angle between a vector from the antenna to the receiver (declination) and a vector
along which the antenna is pointing, projected onto a vertical plane along which the
antenna direction vector lies. Note that if no clear path exists between the antenna and
the receiver, the angle of declination is taken from the antenna to the top of the highest
incident peak on-route to the receiver.

Horizontal Vector
Antenna
Direction
RA
v
l
s
AI
il
a
i
r
TL
a

Figure 2.8: The geometry of the vertical loss calculation for antennae.

1
C-library equivalent is: double atan2(double x,double y).

38
Consider an antenna d at site
l
s (as shown in Figure 2.8), which is tilted at an angle
TL
a from the horizontal. To calculate the vertical component of the loss experienced by
test point
i
r it is necessary to compute the angle
RA
v :
TL AI
il
RA
a a v =
where
AI
il
a is the angle-of-incidence (AI) from site
l
s to test point
i
r . All of the AI
values are precomputed from the topographical data that describes the network area.
These precomputed values are stored in a look-up table, and de-referenced by the site
and test point indices.


Figure 2.9: The angles-of-incidence for a site in NetworkA4.
Figure 2.9 shows an example of the angle-of-incidence set for one site in a network; the
site position is marked by a dark-blue square with a white centre. Around the base of the
site, the angles-of-incidence are positive (shown in green); further away from the site,
the angles-of-incidence turn negative (shown in red); white areas are those in which the
angles-of-incidence are between -05 and +05.
Note that function ( ) a d
VG
is defined as a look-up table with the angle given in degrees
between -180 and +180; hence the angle calculation is ultimately rounded to the
nearest degree.

39
2.2.6 Channels
Network N contains the definition of a finite and ordered set of channels C, where
k
c
denotes the k
th
channel in C.
2.2.6.1 Channel Separation Gain
Receivers are tuned to receive signals on specific channels, and any received signal not
employing this channel will be attenuated the greater the difference (between the
channel to which the receiver is tuned and the channel on which the received signal was
transmitted), the greater the attenuation.
If a receiver, tuned to channel C c
x
, receives a signal, transmitted on channel
C c
y
, the effective power of the received signal is attenuated. The effective gain
applied to the received signal is given by the function ( )
y x
CS
c c G , . The range of the
function is given by:
( ) 1 0
y x
CS
c c G ,
2.2.6.2 Channel Traffic Capacity
The traffic carrying capacity of a given number of channels n is denoted ( ) n T
CH
, and is
described in Table 2.1. Note that these figures are derived by the operators, and the
method used for their generation is based on a heuristically adjusted version of Erlangs
formula (as described in [9]).
Number of channels n 1 2 3 4 5 6 7 8
Traffic Capacity ( ) n T
CH
in Erlangs 2.9 8.2 15 22 28 35.5 43 56
Table 2.1: Traffic capacity vs. number allocated base station channels.
The number of channels required by a base station is encapsulated in the function
( ) t N
TR
, which returns the number of channels whose capacity is greater than or equal to
call volume t.

40
2.2.7 Base Stations
Network N contains the definition of a finite set of base stations B, where
j
b denotes the
j
th
base station in B. Base stations occupy the candidate sites defined by S. The
maximum number of base stations that can occupy one site is constant and denoted by
BSPERSITE
n . Therefore the maximum number of base stations in a network is given by:
BSPERSITE
n S B =
2.2.7.1 Activation
Boolean parameter
AC
b is true if base station b is active; if a base station is not active it
has no influence on the network. The set of active base stations B B
AC
is given by:
{ } true = =
AC AC
b B b B :
and the number of active base stations is given by:
AC AB
B n =
2.2.7.2 Site Occupied
Base station b occupies site S b
ST
, note that
BSPERSITE
n base stations may share each
site. The set of occupied sites S S
OC
is given by:
{ } s b B b S s S
ST AC OC
= = :
and the number of occupied sites is given by:
OC OS
S n =
2.2.7.3 Antenna Type
The type of antenna used by base station b is given by D b
AD
.
The set of sites at which multiple base stations employ omni-directional antennae
(thereby breaking the antenna constraint described in Section 2.2.5.1) is given by:

41
( ) ( )

)

=
=
AD
y
AD
x
ST
y
ST
x
AC
y
AC
x
OC
OV
b b
b b y x B b B b S s
S
isOmni isOmni
:

where function ( ) d isOmni returns true if antenna D d is omni-directional,
otherwise it returns false.
The number of sites on which multiple base stations employ omni-directional antennae
is given by:
OV OV
S n =
2.2.7.4 Transmission Power
The transmission power employed by base station b is given by
TX
b . The lower limit of
TX
b is given by
TXMIN
p and transmission power is divided into
TXSTEP
n discrete intervals,
each one being 2dB (a factor of ~158) higher than its predecessor. From this we can
calculate the maximum transmission power:
( )
1
10
2
10

=
TXSTEP
n
TXMIN TXMAX
p p
2.2.7.5 Azimuth
The horizontal angle, or azimuth, of the antenna used by base station b is given by
AZ
b ,
and it is stored in one-degree intervals in the range 359 0
AZ
b .
2.2.7.6 Tilt
The vertical angle, or tilt, of the antenna used by base station b is given by
TL
b , and it is
stored in one-degree intervals in the range 15 0
TL
b .
2.2.7.7 Channel Assignment
A set of channels is allocated to each base station, where the k
th
channel allocated to
base station
j
b is denoted by
CH
jk
b . The maximum number of channels any base station
may use is given by
CHANPERBS
n .

42
The link is defined as being the part of the base station to which channels are assigned
each link requiring the allocation of exactly one channel. This convention is adopted to
simplify the description of the network design algorithms in Chapter 3.
2.2.7.8 Example
Figure 2.10 shows an example network with two active base stations:
One base station has a directional antenna, which is indicated by a magenta coloured
line originating from the base stations site. The angle at which it leaves the site is
the antenna azimuth (60) and its length is proportional to (the logarithm of) the base
station transmission power (04W)
The other base station has an omni-directional antenna, which is indicated by a
magenta coloured circle. The centre of the circle indicates the base stations site and
the circles radius is proportional to (the logarithm of) the base stations transmission
power (04W).

Figure 2.10: Two active base stations in NetworkA4.
Due to presentation limitations, tilt, channel assignment, and antenna type are not
shown in the diagram.

43
2.2.8 Signal Path Loss
Signal propagation information is defined in a look-up table. The normalised loss
experienced by a signal going from site
l
s to test point
i
r is given by
PG
il
g .

Figure 2.11: The propagation gain from a site in NetworkA2.
Figure 2.11 shows an example of the path loss from one site in a network (the site
position is marked with a dark blue square with a white centre). Around the site is the
area with the lowest propagation loss (shown in dark grey); the test points further from
the site have higher losses (shown light grey). A river (or, at least, its effect on signal
propagation) can be seen passing through the working area from the north-west to the
south-east. The pixel-intensity scale used in the diagram is logarithmically proportional
to the path loss.

44
2.2.9 Radio Field-Strength Calculation
Test points receive signals from every active base station. The field strength due to each
base station is a product of the base stations transmission power, the normalised path
loss, and the horizontal and vertical antenna losses
1
.
The field strength of the radio signal at test point R r
i
due to base station
AC
j
B b is denoted by
FS
ij
p :
( ) ( )
RA
ij
VL
m
RA
ij
HL
m
PG
il
TX
j
FS
ij
v d h d g b p =
where
l
ST
j
s b = and
m
AD
j
d b =
RA
ij
h and
RA
ij
v are the horizontal and vertical angles from the antenna to the test point, and
they are given by:
( )
AZ
j
XY
l
XY
i
RA
ij
b s r h = arctan and
TL
j
AI
il
RA
ij
b a v =
2.2.9.1 Example
Figure 2.12 shows a single active base station in an example network. The base station
has a directional antenna with an azimuth of 60, a transmission power of 04W, and a
tilt of 0. Each pixel represents a test point, and the pixel intensity is logarithmically
proportional to the signal strength (black being the highest).

1
The terms gain and loss are used interchangeably to refer to a multiplicative factor. Whether the
multiplied signal is amplified or attenuated will depend on whether the gain/loss factor is greater or less
than unity.

45

Figure 2.12: The signal strength due to an active base station in NetworkA4.
2.2.10 Serviced Test Points
Test point r is serviced if at least one received base station signal is greater than or
equal to its signal threshold
SQ
r . The set of serviced test points is therefore given by:
{ }
SQ
i
FS
ij
AC
j i
SV
r p B b R r R = :
and the number of serviced test points is given by:
SV SV
R n =
2.2.10.1 Example
Figure 2.13 shows an example network in which the serviced test points are coloured in
light green; the base station configuration is the same as was used for Figure 2.12.

46

Figure 2.13: An example of serviced test points in NetworkA4.
2.2.11 Best Server
The best server of a test point is the base station from which the signal with the highest
field strength is received. If the received signal from two or more base stations is equal,
then the one with the lowest index is chosen. For a serviced test point
SV
R r , the
best server of r is denoted by
BS
r :
( ) { } x j p p p p b B b B b r
FS
ix
FS
ij
FS
ix
FS
ij j
AC
x
AC
j
BS
i
< = > = \ :
2.2.11.1 Base Station Cell
The set of test points for which base station b is the best server is known as the cell of
b , and it is denoted by
CL
b :
{ } b r R r b
BS SV CL
= = :
2.2.12 Base Station Demand
Demand is the number of channels required by a base station. The demand,
DM
b , of
base station
AC
B b is dependent on the volume of traffic, denoted
BS
t , in its cell:


=
CL
b r
TR BS
r t
Now, the number of channels available is limited to
CHANPERBS
n , hence the traffic
successfully serviced by base station b is given by:

47
( ) { }
CHANPERBS CH BS TS
n T t b , min =
The number of channels, n , required to accommodate traffic volume
BS
t is given by:
( )
TS TR
b N n =
Base stations that cover no traffic still require at least one channel if they are to
contribute to general area coverage. The demand of base station b is therefore given by:
{ } n b
DM
, max 1 =
2.2.13 Traffic Service
The total volume of traffic serviced by all the active base stations in the network is
given by:


=
AC
B b
TS SV
b t
This corresponds to the maximum potential traffic cover, or, in other words, the traffic
cover that would be achieved in the absence of interference.
2.2.14 Fuzzy-Server Set
For test point
i
r a set of servers
SS
i
r is defined to contain all base stations that provide a
signal within the fuzzy-server threshold (the fuzzy server model is discussed in Section
2.1.10). This threshold is the product of
FZ
g (a gain constant) and the signal from the
best server of
i
r :
{ }
FS
ix
FZ FS
ij
AC
j
SS
i
p g p B b r = :
where
x
BS
i
b r =
2.2.15 Carrier to Interference Ratio
The carrier-to-interference-ratio (CIR)

i
r for test point
SV
i
R r is the worst-case CIR
from each of the servers in the fuzzy server set:

48

ij
r b
i
r r
SS
i j

= min
where

ij
r denotes the worst-case CIR found using each of the channels of base station
j
b :

ijk
b
k
ij
r r
DM
j
1 =
= min
where

ijk
r denotes the CIR at test point
i
r when it is tuned to the k
th
channel of base
station
j
b . It is given by the ratio of the server signal strength to the sum of all the other,
interfering, signals
1
:
N
S
ijk
p
p
r =


where
S
p is the strength of the serving signal:
FS
ij
S
p p =
and
N
p is the sum of all the interfering signals:
{
2
1
t
FS
ij
t
B b
N
x
N
p p p
AC
x
=


43 42 1

where
N
x
p is the interference due to all the channels of base station
x
b :
( )

=
=
DM
x
b
y
CH
xy
CH
jk
CS FS
ix
N
x
b b G p p
1
,
Note that, for simplicity of expression,
N
p is composed of two terms:
1
t and
2
t . Term
1
t
includes the server signal as part of the total noise, and then it is explicitly removed by
term
2
t .

1
Note that the contribution of interfering signals may be attenuated by channel separation.

49
The adoption of the worst-case CIR is pessimistic, and may lead to an underestimate of
test point cover. It might be argued that the best-case CIR is more representative
because only one interferference-free channel is required for communication. However,
if this channel was the only interference-free channel available across an entire cell it
might not be able to accommodate all of the call traffic in that cell.
2.2.15.1 Example
Figure 2.14 shows an example of the CIR calculation for two active base stations, where
the base stations have the same configuration as those described in Section 2.2.7.8. Each
pixel represents a test point, and the pixel intensity is logarithmically proportional to the
CIR at that test point (black being the highest). For the purpose of illustration, all the
networks test points are included in this figure, and not just those that are serviced.

Figure 2.14: An example CIR calculation for two active base stations in NetworkA4.
2.2.16 Test point Cover
A test point r is covered if its CIR,

r , is greater than or equal to the global CIR


threshold
CV
. The set of covered test points R R R
SV CV
is given by:
{ }
CV SV CV
r R r R

= :
therefore the number of covered test points,
CV
n , is given by:
CV CV
R n =

50
2.2.16.1 Example
Figure 2.15 shows an example network with two active base stations, where the base
stations have the same configuration as those described in Section 2.2.7.8. The non-
covered test points are coloured blue, the serviced test points are coloured light green,
and the covered test points are coloured dark green. It can be seen that cover is achieved
close to the base station sites, but cover becomes more difficult as the distance from the
base station increases.

Figure 2.15: An example of test point cover for two base stations in NetworkA4.
2.2.17 Traffic Cover
The traffic covered by base station
AC
j
B b is denoted
TC
j
b , it is calculated in the
same way as the traffic service (see Section 2.2.13) except that only covered test points
from the cell of
j
b may contribute:
( )

=


CHANPERBS CH
R b r
TR TC
j
n T r b
CV CL
j
, min
The total volume of traffic successfully covered by the whole network of base stations is
given by:


=
AC
B b
TC CV
b t

51
2.2.18 Network Designs
Most network parameters (such as path loss and test point positions) are fixed, except
for the following base station configuration parameters:
Activation,
AC
b
Antenna Type,
AD
b
Transmission Power,
TX
b
Azimuth,
AZ
b
Tilt,
TL
b
Channel Assignment,
CH
b
A complete and valid allocation of base station parameters is known as a network
design.
2.3 Performance Evaluation
To evaluate and compare design performances it is necessary to define what the design
goals are (this corresponds to Task T3 of Objective O1 in Section 1.3).
2.3.1 Design Goals
Like any business, the network operators wish to maximise profits. They must
maximise revenue and minimise overheads. For a network operator to succeed, the
network design goals must represent the economic goals as closely as possible.
2.3.1.1 Revenues and Overheads
Revenue is generated principally from telephone calls. The network model provides the
operator with a prediction of the volume of call traffic and, hence, the expected call
revenue. Examples of the overheads that contribute to the total cost of running the
network include:

52
Base station equipment leasing.
Base station site rental.
Staff office rental.
Personnel salaries.
Radio spectrum rental.
The overheads can be divided into two categories:
Dynamic costs, which grow with the size of the network (e.g. base station site rental,
equipment leasing).
Static costs, which must be met regardless of the network size (e.g. office rental,
salaries, radio-spectrum rental).
No overhead component can be exclusively categorised in this way, but it is a
distinction that lends itself neatly to analysis.
2.3.1.2 Profit and Loss
Consider a network of size n (arbitrary units), and let network revenue be a function of
n and some constant T that represents the normalised value of providing telephony:
n T = Revenue
Let network overheads be a function of n and some constant
D
C that represents the
normalised cost of providing telephony. The overheads also include the static cost
component
S
C :
S D
C n C + = Overhead
Profit is given by Revenue Overhead, therefore

53
( )
S D S D
C C T n C n C n T = = Profit
In order to turn a profit, the network operator must ensure that the revenue exceeds the
overhead. The equation above identifies two key-factors controlling a networks
profitability: the network size, n, and the call-profit-per-unit-network,
D
C T . These
expressions must both be maximised in order to maximise profits.
2.3.1.3 Translating Business Analysis into Engineering Requirements
The simple analysis in the previous section showed that high profitability is achieved in
a large network that provides efficient telephony. This must be translated into a network
design requirement. An ideal network design will:
Be available for use by the maximum number of subscribers every mobile station
in a network should be able to instantiate or receive telephone calls at any time.
Minimise the cost of the network infrastructure.
2.3.1.4 Data Set Limitations
The available data for the network design problem does not include details of network
infrastructure costs, furthermore the subscriber information is only given in terms of call
volume which is, at best, a relative indicator of call revenue. Because of the omission of
absolute monetary-values in the network data sets, the design requirements were
simplified. The design requirements are listed below in order of significance:
1. Maximisation of subscriber coverage.
2. Maximisation of area coverage.
3. Minimisation of the number of sites used.
4. Minimisation of the number of base stations deployed.

54
If more information were available then this design specification could be improved, for
example it would be simple to include an efficiency requirement such as the
maximisation of the ratio of call revenue to site rental. Note that site rental is assumed
to be considerably more expensive than the telephony equipment, hence the relative
significance of metrics 3 and 4; this assumption is based on the experience of network
operators.
2.3.2 Performance Comparison
Comparing the performance of different network designs is done by comparing network
performance metrics. Each performance metric has a relative significance, and the
networks are compared metric-by-metric. These metrics are listed below in order of
significance:
1. Number of co-sited omni-directional antennae,
OV
n
2. Volume of traffic covered,
CV
t
3. Volume of traffic serviced,
SV
t
4. Number of test points covered,
CV
n
5. Number of test points serviced,
SV
n
6. Number of occupied sites,
OS
n
7. Number of active base stations,
AB
n
Metrics 1,6 and 7 should be minimised and metrics 2,3,4 and 5 should be maximised.
The performance of a network design is expressed as a function ( ) ND p . If network
designs
a
ND and
b
ND contain no co-sited omni-directional antennae and produce
traffic cover values
CV
a
t and
CV
b
t respectively then
a
ND outperforms
b
ND if
CV
b
CV
a
t t > .
However, if
CV
b
CV
a
t t = then a comparison of the next significant metrics,
SV
a
t and
SV
b
t ,

55
must be performed; if these two metrics are also equal then comparison continues to the
next metric and so on. This relationship may be expressed as follows:
( ) ( )
( )
( )
|
|
|
.
|

\
|
<
= = =
= = =

|
|
.
|

\
|
< = =
= = =

|
|
.
|

\
|
> =
= = =

|
|
.
|

\
|
>
= = =
> = =
> =
<
= >
AB
b
AB
a
OS
b
OS
a
SV
b
SV
a
CV
b
CV
a
SV
b
SV
a
CV
b
CV
a
OV
b
OV
a
OS
b
OS
a
SV
b
SV
a
CV
b
CV
a
SV
b
SV
a
CV
b
CV
a
OV
b
OV
a
SV
b
SV
a
CV
b
CV
a
SV
b
SV
a
CV
b
CV
a
OV
b
OV
a
CV
b
CV
a
SV
b
SV
a
CV
b
CV
a
OV
b
OV
a
SV
b
SV
a
CV
b
CV
a
OV
b
OV
a
CV
b
CV
a
OV
b
OV
a
OV
b
OV
a
b a
n n
n n n n n n
t t t t n n
n n n n n n
t t t t n n
n n n n
t t t t n n
n n
t t t t n n
t t t t n n
t t n n
n n
ND ND p p

This expression is not particularly clear, but it is expressed better algorithmically, i.e. as
a function that compares the performance of two network configurations and returns the
result as a boolean (see Figure 2.16).
FunctionDef Boolean greaterThan(
a
ND ,
b
ND )
If
OV
b
OV
a
n n < Return true
If
OV
b
OV
a
n n > Return false
If
CV
b
CV
a
t t > Return true
If
CV
b
CV
a
t t < Return false
If
SV
b
SV
a
t t > Return true
If
SV
b
SV
a
t t < Return false
If
CV
b
CV
a
n n > Return true
If
CV
b
CV
a
n n < Return false
If
SV
b
SV
a
n n > Return true
If
SV
b
SV
a
n n < Return false
If
OS
b
OS
a
n n < Return true
If
OS
b
OS
a
n n > Return false
If
AB
b
AB
a
n n < Return true
If
AB
b
AB
a
n n > Return false
Return false
EndFunctionDef

Figure 2.16: Comparison operator for network performance metrics.

56
2.4 Data Sets
Data sets have been obtained from two leading network operators; four networks from
Operator A (NetworkA1, NetworkA2, NetworkA3, and NetworkA4) and one network
from Operator B (NetworkB1). There are some qualitative differences between the data
sets provided by the different operators, and the network model used in this thesis
represents a synthesis of ideas from both operators.
Network
Quantity
A1 A2 A3 A4 B1
PX
n 195 323 248 205 300
PY
n 844 287 229 245 300
GP
l 200 m 200 m 200 m 50 m 100 m
R 29954 74845 17393 48512 90000
TR
R r
r

min
0 Er 0 Er 0 Er 0 Er 0 Er
TR
R r
r

max
111 Er 064 Er 100 Er 143 Er 2739 Er

R r
TR
r
321069 Er 278733 Er 298782 Er 265151 Er 343760 Er
SQ
R r
r

min
90 dB 90 dB 90 dB 90 dB 134 dB
SQ
R r
r

max
90 dB 90 dB 90 dB 90 dB 96 dB
S 250 320 568 244 91
BSPERSITE
n 3 3 3 3 3
PG
il
S s R r
g
l i

min
212 dB 200 dB 184 dB 180 dB 209 dB
PG
il
S s R r
g
l i

max
67 dB 54 dB 49 dB 50 dB 77 dB
AI
il
S s R r
a
l i

min
6846 8491 8634 8618 4348
AI
il
S s R r
a
l i

max
486 1134 2447 8454 1523
D 3 3 3 3 3
OM
D 1 1 1 1 1
C 54 54 54 54 50
TXMIN
p 26 dBm 26 dBm 26 dBm 26 dBm 26 dBm
TXMAX
p 55 dBm 55 dBm 55 dBm 55 dBm 55 dBm
CHANPERBS
n 7 7 7 7 8
CV
14 dB 14 dB 14 dB 14 dB 12 dB
FZ
g 2 dB 2 dB 2 dB 2 dB 4 dB
k = 0 0 dB 0 dB 0 dB 0 dB 0dB
k = 1 18 dB 18 dB 18 dB 18 dB 9 dB
k = 2 50 dB 50 dB 50 dB 50 dB 41 dB
k = 3 58 dB 58 dB 58 dB 58 dB 48 dB
( )
k n n
CS
C C G
+
,
k > 3 0 0 0 0 0

Table 2.2: Summary of statistics for network data sets.

57
2.4.1 NetworkA1
Roads dominate NetworkA1, and a service is required on and around the roads
(indicated by the presence of test points). Significant call-traffic volume is seen on the
roads, probably in proportion to the volume of car traffic.
(a) (b) (c)
Figure 2.17: NetworkA1 (a) test points, (b) traffic densities, and (c) candidate site positions.

58
Candidate site density is high where there are test points, but it does not vary
significantly with traffic density.
The minimum number of base stations required to accommodate the total traffic is given
by:
( )
(
(
(
(
(

=


CHANPERBS CH
R r
TR
BSMIN
n T
r
n
hence, for NetworkA1:

(
75 67 74
43
91 3210
= =
(
(
(

= .
.
BSMIN
n base stations
The minimum number of sites is given by:
(
(
(

=
BSPERSITE
BSMIN
SITEMIN
n
n
n
hence, for NetworkA1:

(
25 25
3
75
= =
(
(
(

=
SITEMIN
n sites
which represents 10% of the candidate sites available.
2.4.2 NetworkA2
NetworkA2 represents a rural region. The majority of call-traffic is found on a strip
(possibly a road) that runs from west to south-east through the network, and traffic is
also found in a central cluster (possibly the town centre). Examination of the
propagation loss files suggests that a river runs through the centre of the network
alongside the road (see Figure 2.11).

59

(a)

(b)

(c)
Figure 2.18: NetworkA2 (a) test points, (b) traffic densities, and (c) candidate site positions.
Candidate sites are distributed evenly across the working area, but some sites lie in the
suspected river. It is possible that the river is in fact underground (and is still able to

60
effect propagation) or that candidate sites have been placed randomly for the purpose of
creating test data. Candidate site density does not vary significantly with test point
density or call traffic density.
The minimum number of base stations required to accommodate the total traffic in
NetworkA2 is 65 and, hence, the minimum number of sites required is 22 - this
represents 69% of the candidate sites.
2.4.3 NetworkA3
NetworkA3 probably represents a small town. The centre of the network contains two
large clusters of call-traffic; the highest density of call-traffic is on a strip that circles the
centre (possibly a ring-road). Significant call-traffic is found on several spokes
(probably roads) radiating from the networks centre.

(a)


(b)

(c)
Figure 2.19: NetworkA3 (a) test points, (b) traffic densities, and (c) candidate site positions.

61
Candidate-site density varies significantly with call-traffic density; in particular two
dense clusters of sites correspond very closely with dense regions of call traffic. The
remaining sites are distributed evenly throughout the working area.
The minimum number of base stations required to accommodate the total traffic in
NetworkA3 is 70 and, hence, the minimum number of sites required is 24 this
represents 42% of the candidate sites.
2.4.4 NetworkA4
NetworkA4 possibly represents an urban or suburban region; the call-traffic density is
fairly uniform across the working area, as is the candidate site density. The resolution of
this network (50m) is the highest of any of the data sets, and this suggests that it is a
close-up of an area with high traffic density.

(a)

(b)

(c)
Figure 2.20: NetworkA4 (a) test points, (b) traffic densities, and (c) candidate site positions.

62
The minimum number of base stations required to accommodate the total traffic in
NetworkA4 is 62 and, hence, the minimum number of sites required is 21 this
represents 86% of the candidate sites.
2.4.5 NetworkB1
In NetworkB1 high call-traffic density occurs in clusters, especially in the north-west
corner of the working area.
(a) (b)
(c)

(d)
Figure 2.21: NetworkB1 (a) test points, (b) traffic densities, (c) candidate site positions, and (d) service
thresholds.
Candidate-site density does not vary significantly with call-traffic density; in fact there
are very few sites in the region of highest call-density (the north-west). This almost
certainly makes covering the traffic in that region more difficult.

63
There are some significant differences between this network and the networks obtained
from Operator A:
The total traffic in this network is much higher more than 10 times the total of
NetworkA1.
The maximum traffic represented by one test point is 2739 Er, whereas the highest
value for the other networks was 143 Er almost 20 times smaller. This means that
some well-dimensioned cells (i.e. cells which contain nearly the maximum traffic
they can hold without covering an un-necessarily large area) could be 2 or 3 test
points in size. This makes good network designs harder to find, and it also indicates
that the resolution of the test point grid is too low to enable the simulation of reliable
cell-plans.
The number of candidate sites is much lower. This is possibly because the data is
derived from an existing network (i.e. candidate sites represent those currently
occupied) and not augmented with extra sites. In the other networks the addition of
extra sites increases the cell-planning options.
In the previous networks, the service threshold was a constant 90dBm, whereas in
this network less than 2% of test points are that easy to cover. In fact the average
service threshold is approximately 70dBm (a factor of 100 larger than the average
for the other networks). This makes the test points harder to cover, or, at least,
necessitates the use of higher transmission powers, which reduces the scope of
relative power trade-offs for interference management.
The channel separation gain is nearly 10 times as strict for this network as it is for
the other networks. This makes interference management by channel separation
more difficult.
The minimum number of base stations required to accommodate the total traffic in
NetworkB1 is 593 and, hence, the minimum number of sites required is 198. This is

64
more than twice as many sites as are defined for this network; therefore the total
network traffic is not coverable. The largest amount of traffic that may be covered
(assuming that each of the 91 sites accommodates 3 base stations, each of which can
handle 58 Er) is 91358 = 15834 Er. This is 4606% of the total network traffic.

65
Chapter 3

Network Design Part 1
This chapter describes the application of simple combinatorial optimisation techniques
to the Wireless Capacitated Network Design Problem (WCNDP), the details of which
are described in the previous chapter. The goal of the work described in this chapter was
the identification of optimisation algorithms that produce high quality network designs
in a reasonable time (this corresponds to Task T1 of Objective O2 in Section 1.3).
The structure of this chapter is as follows: Section 3.2 describes algorithms for
initialising network designs, which are improved by the search algorithms described in
Section 3.3, finally Section 3.4 discusses the factors that limit search efficiency.
3.1 Choice of Optimisation Algorithms
For all of the different problems described in this thesis, a wide variety of optimisation
algorithms were trialed. An empirical approach was taken to development, in which
common heuristics such as hillclimbing, simulated annealing, and genetic algorithms
were compared using short run-times. Some poorly performing algorithms could be
quickly rejected, allowing more time for the remaining algorithms to evolve. The
algorithms detailed in this chapter (and in the thesis as a whole) are those that produced
the best network designs.

66
3.2 Initialisation Algorithms
Initialisation is characterised by the creation of a single network design; no valid
intermediate designs are created during an initialisation procedure. Initialisation
algorithms create network designs, which may be modified by search algorithms.
3.2.1 Algorithm: InitBlank
A special case initialisation is the greenfield or empty network, this is a network design
in which no base stations are active and hence no cover can exist. These network
designs, although useless in themselves, are used as the starting point for other design
algorithms.
3.2.1.1 Description
Although no base stations are active, it is necessary to select valid parameters for in-
active base stations. If a subsequently applied search algorithm activates a base station,
then it is not necessarily going to adjust the base stations parameters. Choosing random
initial parameters reduces bias, e.g. if all base stations were allocated an azimuth of 0
this would give them a northern bias.

Algorithm InitBlank
Input: none
Output: network design ND
INIT


Declare network design ND
INIT
Deactivate all base stations in ND
INIT

For every base station b in ND
INIT
Choose random azimuth for b
Choose random tilt for b
Choose random transmit power for b
Choose random directional antenna for b
For every link k on b
Assign random channel to k from set of all channels
EndFor
EndFor
Return ND
INIT


Figure 3.1: Algorithm InitBlank.
No results are included for this algorithm, as the network designs it produces have no
effect.

67
3.2.2 Algorithm InitRandomA
Network design parameters (such as azimuth or antenna type) uniquely define a
candidate network design, and Algorithm InitRandomA (outlined in Figure 3.2) does not
favour the production of any particular network design. Algorithm InitRandomA aims to
sample with equal probability from the set of all possible network designs.
The number of active base stations in each candidate design is between 1 and B (with
a uniform probability distribution). Note that this requirement forces the probability
distribution of the number of occupied sites to be non-uniform
1
.
3.2.2.1 Description
A set of n base stations is randomly activated. For each active base station, all
parameters (azimuth, tilt, transmit power, and channel assignment) are chosen
randomly, and the antenna used by each active base station is chosen at random from
the set of directional antennae. If only one base station is active at a site then there is a
probability, fixed at the beginning of the algorithms execution, that the base station will
be re-allocated an omni-directional antenna (selected at random from the set of
directional antennae).

1
It is possible for this to be made uniform; however, the probability distribution of the number of active
base stations would then be non-uniform.

68

Algorithm InitRandomA
Input: none
Output: network design ND
INIT

Declare network design ND
INIT
Choose random number n from 1 to the number of possible base stations
Activate n base stations at random in ND
INIT
For every base station b in ND
INIT
Choose random azimuth for b
Choose random tilt for b
Choose random transmit power for b
Choose random directional antenna for b
For every link l on b
Assign random channel to l from set of all channels
EndFor
EndFor
Choose random probability p for omni-directional antennae
For every occupied site s in ND
INIT
Count number of active base stations t on s
If (t = 1) and (p occurs)
Choose antenna a at random from omni-directional set
Assign a to active base station on s
EndIf
EndFor
Return ND
INIT

Figure 3.2: Algorithm InitRandomA.
3.2.2.2 Example
The following is an example of the output from algorithm InitRandomA. The target
network was NetworkA4, the number of active base stations chosen was 21, and the
probability of mounting an omni-directional antenna on sites with one base station was
015. The network performance for the example configuration was: 163% test point
cover, and 106% traffic cover.

69

Figure 3.3: An example network design produced by InitRandomA.
It can be seen in Figure 3.3 that there is a mix of occupied and un-occupied sites. The
occupied sites have one or two active base stations, and some of those with just one
base station have an omni-directional antenna.
3.2.2.3 Results
In order to obtain a statistical view of the algorithms performance it was run on each
network for 20 hours. The results are summarised in Table 3.1.
Network
Performance Metric
A1 A2 A3 A4 B1
Min 000 154 144 051 000
Max 4451 4422 4282 5740 184
Average 2595 2802 2581 3273 061
Traffic Cover
(% of total)
s.d. 847 917 971 1242 032
Min 000 996 508 042 000
Max 7452 5609 5246 6411 1624
Average 3420 3843 3238 3914 644
Test Point Cover
(% of total)
s.d. 830 1045 1054 1291 296
Min 040 031 018 041 110
Max 10000 10000 10000 10000 10000
Average 7617 7430 7534 7265 7673
Occupied Sites
(% of total)
s.d. 2850 2892 2782 2875 2754
Min 013 010 006 014 037
Max 9920 9990 9988 10000 10000
Average 5157 5018 5010 4681 5207
Active Base Stations
(% of total)
s.d. 2895 2993 2868 2820 2918
Number of Design Evaluations 660 248 703 480 1292
Table 3.1: The statistical performance of algorithm InitRandomA.

70
The following graphs show, in more detail, the result of applying InitRandomA to
NetworkA4.
0.0%
0.5%
1.0%
1.5%
2.0%
2.5%
3.0%
3.5%
4.0%
4.5%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Traffic Cover (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 0.51% max = 57.40% average = 32.73% s.d. = 12.42% #iterations = 480

Figure 3.4: The distribution of traffic cover for InitRandomA on NetworkA4.
0%
1%
2%
3%
4%
5%
6%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Test Point Cover (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 0.42% max = 64.11% average = 39.14% s.d. = 12.91% #iterations = 480

Figure 3.5: The distribution of test point cover for InitRandomA on NetworkA4.

71
0%
2%
4%
6%
8%
10%
12%
14%
16%
18%
20%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Sites Occupied (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 0.41% max = 100.00% average = 72.65% s.d. = 28.75% #iterations = 480

Figure 3.6: The distribution of site occupancy for InitRandomA on NetworkA4.
0.0%
0.5%
1.0%
1.5%
2.0%
2.5%
3.0%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 0.14% max = 100.00% average = 46.81% s.d. = 28.20% #iterations = 480

Figure 3.7: The distribution of number of base stations active for InitRandomA on NetworkA4.
Figure 3.7 shows that the distribution of the number of active base stations is
approximately uniform (as intended), and consequently the distribution of the number of
occupied sites, as shown in Figure 3.6, is skewed, i.e. the instances of high site
occupancy occur most often.
The next set of graphs aims to show the correlation between the different network
performance metrics, and NetworkA4 is again given as an example.

72
0%
10%
20%
30%
40%
50%
60%
70%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
r
a
f
f
i
c

C
o
v
e
r

(
%
o
f

t
o
t
a
l
)
Summary: #iterations = 480

Figure 3.8: Graph of traffic cover vs. base stations active for InitRandomA on NetworkA4.
0%
10%
20%
30%
40%
50%
60%
70%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Sites Occupied (% of total)
T
r
a
f
f
i
c

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)
Summary: #iterations = 480

Figure 3.9: Graph of traffic cover vs. sites occupied for InitRandomA on NetworkA4.
It can be seen from Figure 3.8 that there is a strong positive correlation between the
number of active base stations and the traffic cover. A similar correlation between
traffic cover and site occupancy is shown in Figure 3.9.

73
0%
10%
20%
30%
40%
50%
60%
70%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)
Summary: #iterations = 480

Figure 3.10: Graph of test point cover vs. base stations active for InitRandomA on NetworkA4.
0%
10%
20%
30%
40%
50%
60%
70%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Sites Occupied (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)
Summary: #iterations = 480

Figure 3.11: Graph of test point cover vs. sites occupied for InitRandomA on NetworkA4.
It can be seen from Figure 3.10 that there is a strong positive correlation between the
number of active base stations and the test point cover. A similar correlation between
test point cover and site occupancy is shown in Figure 3.11.
The positive correlation between base station activity and traffic coverage suggests that
to maximise the average traffic coverage of network configurations produced by

74
initialisation method InitRandomA, all base stations should be active, and similarly all
sites should be occupied.
3.2.3 Algorithm: InitRandomB
Consider the example network configuration shown in Figure 3.3 a detail from which
is shown in Figure 3.12 indicating the active base station of interest. None of the test
points served by this base station (points where the field-strength due to the base station
is in excess of the service threshold) are covered, i.e. the carrier-to-interference-ratio
(CIR) at all of these points is insufficient. To understand why, it is necessary to examine
the channel assignment. The base station has a demand of 7 channels, and
8
c ,
9
c ,
18
c ,
21
c ,
27
c ,
2
c , and
9
c were allocated. Note that channel
9
c was allocated twice, resulting
in the co-channel rejection ratio (0dB) being the upper limit to the CIR for all points
served by this base station
1
. As the target CIR is 14dB it is impossible for this base
station to cover any points. Note also that channels
8
c and
9
c are adjacent; this leads to
a similar (but weaker) upper limit to the CIR, i.e. the adjacent-channel rejection ratio,
which for Operator A is 18dB. If the co-channel assignments were removed, this weaker
upper limit would not necessarily prevent test point coverage; however, it is within 4dB
of the CIR and would be best avoided. For Operator B the target CIR is 12dB and the
adjacent-channel rejection ratio is 9dB, therefore adjacent-channel allocations would
prevent coverage as surely as co-channel allocations.

Figure 3.12: A subsection of the example network generated by InitRandomA.

1
See Section 2.2.15 for details of CIR calculation.

75
Algorithm InitRandomB is a modified version of InitRandomA that employs an
improved channel selection process. It was shown above that a base station is
guaranteed to fail if any two allocated channels are equal, and likely to fail if any two
allocated channels are adjacent
1
. Algorithm InitRandomB selects channels at random
such that a two-channel separation is maintained between channels allocated to the same
base station this is an example of an embedded heuristic (see [12]). Note that the
channel-selection bias is only designed to avoid channel clashes in the same base
station, not between base stations (i.e. co-sector interference not co-site or far-site
interference).
3.2.3.1 Description
InitRandomB is identical to InitRandomA except for the channel selection process. For
each base station a list of all valid channels is constructed, and selections are made at
random from the channels in this list. Selected channels, and those adjacent to them, are
deleted from the list. This selection and deletion process is repeated until the base
station has received its full complement of channels.

1
Adjacent channel selection guarantees failure for Operator B networks.

76

Algorithm InitRandomB
Input: none
Output: network design ND
INIT

Declare network design ND
INIT
Choose random number n from 1 to the number of possible base stations
Activate n base stations at random in ND
INIT
For every base station b in ND
INIT
Choose random azimuth for b
Choose random tilt for b
Choose random transmit power for b
Choose random directional antenna for b
Construct list of all possible channels CL
For every link k on b
Choose channel c at random from CL
Assign c to k
Delete c from CL
If c-1 is in CL
Delete c-1 from CL
EndIf
If c+1 is in CL
Delete c+1 from CL
EndIf
EndFor
EndFor
Choose random probability p for omni-directional antennae
For every occupied site s in ND
INIT
Count number of active base stations t on s
If (t = 1) and (p occurs)
Choose antenna a at random from set of omni-directionals
Assign a to active base station on s
EndIf
EndFor
Return ND
INIT

Figure 3.13: Algorithm InitRandomB.
3.2.3.2 Example
The following is an example of the output from algorithm InitRandomB. The target
network was NetworkA4, the number of active base stations was chosen as 21, and the
probability of mounting an omni-directional antenna (on sites with one base station) was
015. The performance for the example design was: 296% test point cover, and 155%
traffic cover.

77

Figure 3.14: An example network design produced by InitRandomB.
3.2.3.3 Results
In order to obtain a statistical view of the algorithms performance it was run on each
network for 20 hours. The results are summarised in Table 3.2.
Network
Performance Metric
A1 A2 A3 A4 B1
Min 134 474 238 461 000
Max 5395 5058 4887 6055 469
Average 3358 3346 2940 4081 236
Traffic Cover
(% of total)
s.d. 905 920 998 1178 101
Min 1013 2394 1056 1367 002
Max 5726 5833 8205 6785 2342
Average 4148 4356 3593 4782 1305
Test Point Cover
(% of total)
s.d. 688 870 995 1052 523
Min 040 156 035 205 110
Max 10000 10000 10000 10000 10000
Average 7645 7283 7344 7854 7522
Occupied Sites
(% of total)
s.d. 2841 2918 2870 2667 2850
Min 013 052 012 068 037
Max 9987 9885 10000 9959 10000
Average 5221 4786 4845 5308 5043
Active Base Stations
(% of total)
s.d. 2930 2923 2912 2760 2919
Number of Design Evaluations 656 257 723 444 1338
Table 3.2: The statistical performance of algorithm InitRandomB.
The following graphs show, in more detail, the result of applying InitRandomB to
NetworkA4. Note that the number of active base stations and the number of sites
occupied occurred with the same frequencies as those observed for algorithm
InitRandomA, and therefore they are not included.

78
0.0%
0.5%
1.0%
1.5%
2.0%
2.5%
3.0%
3.5%
4.0%
4.5%
5.0%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Traffic Cover (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 4.61% max = 60.55% average = 40.81% s.d. = 11.78% #iterations = 444

Figure 3.15: The distribution of traffic cover for InitRandomB on NetworkA4.
0.0%
0.5%
1.0%
1.5%
2.0%
2.5%
3.0%
3.5%
4.0%
4.5%
5.0%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Test Point Cover (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 13.67% max = 67.85% average = 47.82% s.d. = 10.52% #iterations = 444

Figure 3.16: The distribution of test point cover for InitRandomB on NetworkA4.

79
0%
10%
20%
30%
40%
50%
60%
70%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
r
a
f
f
i
c

C
o
v
e
r

(
%
o
f

t
o
t
a
l
)
Summary: #iterations = 444

Figure 3.17: Graph of traffic cover vs. base stations active for InitRandomB on NetworkA4.
0%
10%
20%
30%
40%
50%
60%
70%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Sites Occupied (% of total)
T
r
a
f
f
i
c

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)
Summary: #iterations = 444

Figure 3.18: Graph of traffic cover vs. sites occupied for InitRandomB on NetworkA4.

80
0%
10%
20%
30%
40%
50%
60%
70%
80%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)
Summary: #iterations = 444

Figure 3.19: Graph of test point cover vs. base stations active for InitRandomB on NetworkA4.
0%
10%
20%
30%
40%
50%
60%
70%
80%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Sites Occupied (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)
Summary: #iterations = 444

Figure 3.20: Graph of test point cover vs. sites occupied for InitRandomB on NetworkA4.
It can be seen that the same correlations observed for InitRandomA network designs
(see Section 3.2.2.3) exist for InitRandomB, and therefore (for random networks at
least) the test point and traffic cover increases with the number of active base stations
and occupied sites.

81
3.2.4 Algorithm: InitRandomC
Algorithm InitRandomC is a slightly modified version of InitRandomB. It was
previously concluded that the maximum average traffic cover was achieved when all the
base stations were active, and therefore InitRandomC is identical to InitRandomB except
that all base stations are active (not just a randomly selected subset). Note that only
directional antennae are used in order to maximise the number of base stations that may
be active.
3.2.4.1 Description
The algorithm for InitRandomC, given in Figure 3.21, is the same as InitRandomB
except that all base stations are active.

Algorithm InitRandomC
Input: none
Output: network design ND
INIT

Declare network design ND
INIT
For every base station, b
Activate b
Choose random azimuth for b
Choose random tilt for b
Choose random transmit power for b
Choose random directional antenna for b
Construct list of all possible channels CL
For every link k on b
Choose channel c at random from CL
Assign c to k
Delete c from CL
If c-1 is in CL
Delete c-1 from CL
EndIf
If c+1 is in CL
Delete c+1 from CL
EndIf
EndFor
EndFor
Return ND
INIT

Figure 3.21: Algorithm InitRandomC.

82
3.2.4.2 Example
The following is an example of the output from algorithm InitRandomC. The target
network was NetworkA4, and the network performance for the example design was:
574% test point cover, 519% traffic cover.

Figure 3.22: An example network design produced by InitRandomC.
3.2.4.3 Results
In order to obtain a statistical view of the algorithms performance, it was run on each
network for 20 hours. The results are summarised in Table 3.3.
Network
Performance Metric
A1 A2 A3 A4 B1
Min 3469 4001 3458 4653 222
Max 5088 5144 4953 6215 483
Average 4328 4472 4262 5465 351
Traffic Cover
(% of total)
s.d. 295 217 245 260 044
Min 4087 5035 4155 5099 1437
Max 5677 6070 5633 6860 2402
Average 4888 5530 4911 6078 1906
Test Point Cover
(% of total)
s.d. 297 212 253 278 174
Number of Design Evaluations 429 160 439 313 551
Table 3.3: The statistical performance of algorithm InitRandomC.
The following graphs show, in more detail, the result of applying InitRandomC to
NetworkA4.

83
0%
2%
4%
6%
8%
10%
12%
14%
16%
18%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Traffic Cover (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 46.53% max = 62.15% average = 54.65% s.d. = 2.60% #iterations = 313

Figure 3.23: The distribution of traffic cover for InitRandomC on NetworkA4.
0%
2%
4%
6%
8%
10%
12%
14%
16%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Test Point Cover (% of total)
F
r
e
q
u
e
n
c
y

(
%

o
f

t
o
t
a
l

n
u
m
b
e
r

o
f

o
c
c
u
r
e
n
c
e
s
)
Summary: min = 50.99% max = 68.60% average = 60.78% s.d. = 2.78% #iterations = 313

Figure 3.24: The distribution of test point cover for InitRandomC on NetworkA4.
3.2.5 Comparison of Initialisation Algorithms
Table 3.4 gives a summary of the results for the three initialisation methods
(InitRandomA, B, and C).

84
Network
Initialisation
Algorithm
Average
Traffic Cover
(% of total)
Average Test
Point Cover
(% of total)
Average No.
Sites
Occupied
(% of total)
Average No.
Base Stations
Active
(% of total)
InitRandomA 2595 3420 7617 5157
InitRandomB 3358 4148 7645 5221 A1
InitRandomC 4328 4888 10000 10000
InitRandomA 2902 3843 7430 5018
InitRandomB 3346 4356 7283 4786 A2
InitRandomC 4472 5530 10000 10000
InitRandomA 2581 3238 7534 5010
InitRandomB 2940 3593 7344 4845 A3
InitRandomC 4262 4911 10000 10000
InitRandomA 3273 3914 7265 4681
InitRandomB 4081 4782 7854 5308 A4
InitRandomC 5465 6078 10000 10000
InitRandomA 061 644 7673 5207
InitRandomB 236 1305 7522 5043 B1
InitRandomC 351 1906 10000 10000
Table 3.4: A summary of results for initialisation algorithms.
Algorithm InitRandomC delivered network designs with the highest average traffic
cover and test point cover for all networks. This algorithm requires all base stations to
be active and, hence, introduces an element of bias in to the configuration parameters
selected. If such biasing is undesirable, InitRandomA or InitRandomB should be used
instead, however, InitRandomA produces network configurations with poor base station
channel assignments, so InitRandomB should always be used in preference to
InitRandomA.
3.3 Search Algorithms
Searching is characterised by the creation of a set of valid network designs. In assessing
the performance of a search algorithm two factors are considered:
The performance of the best design in the set (as defined by the evaluation function
in Figure 2.16).
The time taken to produce the whole set.
For the purposes of algorithm comparison each search algorithm is run for a fixed
length of time, therefore the performance of the best network designs produced by each

85
algorithm may be compared on an equal basis. Some search algorithms may have
alternative termination criteria, e.g. local-search algorithms may terminate once a local-
optima is reached, but for the tests described in this chapter no alternative termination-
criteria was invoked, i.e. in all cases termination was enforced, and indeed required,
after a fixed time.
3.3.1 Algorithm: SearchRandomB
Algorithm SearchRandomB is a pure random search, where network designs are
generated by calling InitRandomB. Successive configurations are independent of those
that have been created previously.
3.3.1.1 Description
The algorithm, detailed in Figure 3.25, begins with a call to InitRandomB to create an
initial network design that is recorded as the best design so far. Control then passes to
a loop, where it remains until a fixed time (specified by the user) has expired. In this
loop, InitRandomB is called and the resulting network design is evaluated. The best
design so far is replaced if its performance is exceeded by the performance of this new
network design.

Algorithm SearchRandomB
Input: execution time limit
Output: network design ND
BEST
Note: may be pre-emptively terminated

Declare network design ND
BEST = InitRandomB()
Repeat until time limit exceeded
Declare network design ND
TRIAL = InitRandomB()
If p(ND
TRIAL) > p(NC
BEST)
ND
BEST = ND
TRIAL
EndIf
EndRepeat

Figure 3.25: Algorithm SearchRandomB.

86
3.3.1.2 Results
Algorithm SearchRandomB was run for 20 hours on each network. The performance of
the best network design produced for each network is summarised in Table 3.5.
Network
Performance Metric
A1 A2 A3 A4 B1
Traffic Cover (% of total) 5395 5058 4887 6055 469
Test Point Cover (% of total) 5550 5819 5391 6785 1989
Occupied Sites (% of total) 10000 9969 10000 10000 10000
Active Base Stations (% of total) 9960 8615 9472 9549 9487
Number of Design Evaluations 656 257 723 444 1338
Table 3.5: The performance of the best network designs for SearchRandomB.
Figure 3.26 shows the performance of the best network design against the time taken to
produce it for NetworkA4.
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=60.55% Test-point Cover=67.85% Active Base Stations=95.49% Sites Occupied=100.00%
Traffic Cover
Test-point Cover
Active Base Stations
Sites Occupied

Figure 3.26: The performance of the best design produced vs. time for SearchRandomB on NetworkA4.
The fastest rate-of-improvement of traffic cover was seen in the early phase of the
algorithms execution. For all the networks, no improvement in design performance was
seen after approximately 2 hours.
3.3.2 Algorithm: SearchRandomC
Algorithm SearchRandomC is identical to SearchRandomB except that all calls to
InitRandomB are replaced with calls to InitRandomC.

87
3.3.2.1 Results
Algorithm SearchRandomC was run for 20 hours on each network. The performance of
the best network design produced for each network is summarised in Table 3.6.
Network
Performance Metric
A1 A2 A3 A4 B1
Traffic Cover (% of total) 5088 5145 4953 6215 483
Test Point Cover (% of total) 5677 5709 5518 6860 1989
Occupied Sites (% of total) 10000 9969 10000 10000 10000
Active Base Stations (% of total) 10000 10000 10000 10000 10000
Number of Design Evaluations 429 160 439 313 551
Table 3.6: The performance of the best network designs for SearchRandomC.
Figure 3.27 shows the performance of the best network design plotted against the time
taken to produce it for NetworkA4.
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=62.15% Test-point Cover=68.60%
Traffic Cover
Test-point Cover

Figure 3.27: The performance of the best design produced vs. time for SearchRandomC on NetworkA4.
SearchRandomC out-performed SearchRandomB on all but NetworkA2. However, the
difference in performance between the two algorithms was small. It is interesting to note
that SearchRandomC did not conclusively out-perform SearchRandomB even though
the initialisation technique that it is based on, InitRandomC, has a significantly higher
average performance than InitRandomB, on which SearchRandomB is based.

88
3.3.3 Algorithm: SearchHillClimbA
SearchHillClimbA is a local-search algorithm. An initial network design is modified by
small adjustments to base station parameters, which are repeated until no one-change to
a single parameter improves network performance. This final design is said to be a local
optima.
Local-search algorithms require the definition of a search neighbourhood, which is the
scope of changes made to input parameters during searching. For algorithm
SearchHillClimbA, the neighbourhood of a network design is defined as follows: a
network design, ND
NEW
, is in the neighbourhood of ND
OLD
if ND
OLD
can be transformed
into ND
NEW
by the adjustment of one base station parameter. For example, two network
designs would be neighbours if they were identical except for the type of antenna used
by a particular base station. The transformation of one network design into another,
neighbouring, design is known as a move.
3.3.3.1 Description
Algorithm SearchHillClimbA has two separate control sequences. During the first, the
set of all possible moves is generated. During the second, these moves are applied and
the resulting network designs evaluated. This cycle is repeated until a complete set of
moves has been applied with no improvement in network performance. The list of
moves is executed in random order, which reduces bias in the hill climbing procedure.
Moves are of the form: <base station> <parameter> <value> except for channel
assignment moves which are of the form: <base station> channel <link index>
<channel>. Table 3.7 lists some examples.
Encoding in Move Format
Action
Base Station Parameter Link Index Value
= 43
79
AZ
b
79
b azimuth 43
= 0
131
TL
b
131
b tilt 0
W 10
43
=
TX
b
43
b txPower 10W
2 69
d b
AD
=
69
b antDef
2
d
17 2 18
c b
CH
=
18
b channel 2
17
c
true
218
=
AC
b
218
b active TRUE
Table 3.7: Example moves for SearchHillClimbA.

89
Algorithm SearchHillClimbA is described by the pseudo-code in Figure 3.28.

Algorithm SearchHillClimbA
Input: network design NDINIT
Output: network design NDBEST
Note: may be pre-emptively terminated

Declare network design NDBEST = NDINIT
Declare boolean termination flag fCONTINUE = true
Create empty move list moves
For every base station b in NDBEST
Add move b active TRUE to moves
Add move b active FALSE to moves
For every possible azimuth az
Add move b azimuth az to moves
EndFor
For every possible tilt tl
Add move b tilt tl to moves
EndFor
For every possible transmit power tx
Add move b txPower tx to moves
EndFor
For every possible antenna type ad
Add move b antDef ad to moves
EndFor
For every link k on b
For every possible channel c
Add move b channel k c to moves
EndFor
EndFor
EndFor
Repeat while fCONTINUE
fCONTINUE = false
Shuffle moves into random order
For every move m in moves
Declare network design NDTRIAL = NDBEST + m
If p(NDTRIAL) > p(NDBEST)
NDBEST = NDTRIAL
fCONTINUE = true
EndIf
EndFor
EndRepeat
Return NDBEST

Figure 3.28: Algorithm SearchHillClimbA.
Note that all moves applied to inactive base stations (unless activating the base station)
will have no effect and may be ignored. Adjustments to base station channels that are
not in use (i.e. not in the base stations working set) will also have no effect and may be
ignored.

90
3.3.3.2 Results
Algorithm SearchHillClimbA was run three times for each network, each run used
different initial network designs and lasted for 20 hours. The designs were generated by
a modified version of InitRandomB, which allowed the fraction of base stations active in
the network to be specified at run-time. The values used were: 0% (Run 1, no base
stations active), 50% (Run 2, half the base stations active), and 100% (Run 3, all base
stations active). The performance of the best network designs produced is summarised
in Table 3.8, and the performance of the initial designs is given for comparison.
Network
Run Performance Metric
A1 A2 A3 A4 B1
Initial 000 000 000 000 000 Traffic Cover
(% of total) Final 5213 4797 5177 5254 536
Initial 000 000 000 000 000 Test Point Cover
(% of total) Final 6996 6153 7642 6948 1832
Initial 000 000 000 000 000 Occupied Sites
(% of total) Final 2080 1563 1215 2213 5385
Initial 000 000 000 000 000 Active Base Stations
(% of total) Final 747 552 411 779 2088
1
Number of Design Evaluations 3584 2060 4219 2797 4893
Initial 3736 3432 3104 4207 259 Traffic Cover
(% of total) Final 5412 3866 4318 5416 533
Initial 4293 4355 3855 4831 1150 Test Point Cover
(% of total) Final 5832 4742 4739 6156 2075
Initial 8480 8781 8856 8811 8462 Occupied Sites
(% of total) Final 8480 8781 8856 8811 8462
Initial 5000 5000 5000 5000 4982 Active Base Stations
(% of total) Final 5000 5000 5006 5027 5055
2
Number of Design Evaluations 749 266 827 508 1555
Initial 3866 4612 4558 53.66 367 Traffic Cover
(% of total) Final 5098 4978 5145 6114 589
Initial 4393 5641 4882 6188 1886 Test Point Cover
(% of total) Final 5356 6031 5453 6866 2433
Initial 10000 10000 10000 10000 10000 Occupied Sites
(% of total) Final 10000 10000 10000 10000 10000
Initial 10000 10000 10000 10000 10000 Active Base Stations
(% of total) Final 9987 10000 9994 9886 10000
3
Number of Design Evaluations 504 212 585 408 736
Table 3.8: The performance of network designs for SearchHillClimbA.
The following graphs show the change in network performance over time for algorithm
SearchHillClimbA on NetworkA4.

91
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=52.54% Test-point Cover=69.48% Active Base Stations=7.79% Sites Occupied=22.13%
Traffic Cover
Test-point Cover
Active Base Stations
Sites Occupied

Figure 3.29: The performance of the best design produced vs. time for
SearchHillClimbA on NetworkA4 (run 1).
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=54.16% Test-point Cover=61.56% Active Base Stations=50.27% Sites Occupied=88.11%
Traffic Cover
Test-point Cover
Active Base Stations
Sites Occupied

Figure 3.30: The performance of the best design produced vs. time for
SearchHillClimbA on NetworkA4 (run 2).

92
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=61.14% Test-point Cover=68.66% Active Base Stations=99.86% Sites Occupied=100.00%
Traffic Cover
Test-point Cover
Active Base Stations Sites Occupied

Figure 3.31: The performance of the best design produced vs. time for
SearchHillClimbA on NetworkA4 (run 3).
Network Run
Initial Traffic Cover
(% of total)
Final Traffic Cover
(% of total)
Improvement in
Traffic Cover
(Final Initial)
1 000 5213 5213
2 3736 5412 1676 A1
3 3866 5098 1232
1 000 4797 4797
2 3432 3866 434 A2
3 4612 4978 366
1 000 5177 5177
2 3104 4318 1214 A3
3 4558 5145 587
1 0.00 5254 5254
2 4207 5416 1209 A4
3 53.66 6114 7.48
1 000 536 536
2 259 533 274 B1
3 367 589 222
Table 3.9: The initial and final traffic covers for multiple executions of SearchHillClimbA.
The traffic cover results given in Table 3.8 are repeated in Table 3.9, where a
calculation of the improvement in traffic cover is included.
Run 1 produced the best design for NetworkA3, Run 2 produced the best design for
NetworkA1, and Run 3 produced the best design for NetworkA2, NetworkA4, and
NetworkB1. The average traffic cover (across all five networks) from Run 1 is 42%, for

93
Run 2 it is 39%, and for Run 3 it is 44%. Run 3 produces the best average performance
and the highest number of best network performances.
It can be seen from Table 3.9 that the traffic cover of the final network design is, to
some extent, dependent on that of the initial design, as the improvement in traffic cover
is inversely related to the amount of traffic cover in the initial design. This relationship
might be expected when one considers the following extreme case: if the traffic cover of
the initial design were maximal then no improvement could be made to traffic cover,
and, conversely, if it were minimal then a high proportion of moves would improve
traffic cover.
For Run 3 of the algorithm test, the initial base station activity was 100%, and this did
not decrease significantly throughout the algorithms execution. It is interesting to
compare the results of Run 3 with the results obtained for SearchRandomC (see Section
3.3.2.1) where base station activity is also 100% (by design). Now, as both algorithms
ran for the same length of time and used (nearly) the same number of base stations, one
would expect the total number of design evaluations to be comparable, however,
SearchHillClimbA made a total of 2445 (504+212+585+408+736) evaluations whereas
SearchRandomC made only 1892 (429+160+439+313+551). The difference is due to
the ability of SearchHillClimbA (or any local-search algorithm) to exploit coherence
between consecutive (neighbouring) network designs, i.e. field strengths are re-
calculated only when a base stations configuration changes. A greater number of
network evaluations does not necessarily improve the effectiveness of a search-
algorithm, but it is likely to help.
3.3.4 Algorithm: SearchConstructA
SearchConstructA is a local-search algorithm and is detailed in Figure 3.32. A green-
field network (see Section 3.2.1) is modified by the sequential activation of base
stations.

94
3.3.4.1 Description
During execution every base station is activated once per iteration, and the resulting
network designs evaluated. Once all base stations have been tried, the one that produced
the best improvement in performance is selected. This new design is only accepted if it
produces an improvement over the performance of the network design from the
previous iteration, i.e. network performance is better with it than without it. Iterations
continue until a (user specified) fixed time has elapsed.

Algorithm SearchConstructA
Input: execution time limit
Output: network design ND
BEST
Note: may be pre-emptively terminated

Declare network design ND
BEST = InitBlank()
Repeat until fixed time has elapsed
Declare network design ND
NEW = ND
BEST
Declare network design ND
TRIAL = ND
BEST
For every inactive base station b in N
TRIAL
Activate b
Choose random azimuth for b
Choose random tilt for b
Choose random transmit power for b
Choose random directional antenna for b
Construct list of all possible channels CL
For every link k on b
Choose channel c at random from CL
Assign c to k
Delete c from CL
If c-1 is in CL

Delete c-1 from CL
EndIf
If c+1 is in CL

Delete c+1 from CL
EndIf
EndFor
If p(ND
TRIAL) > p(ND
NEW)

ND
NEW = ND
TRIAL
EndIf
Deactivate b
EndFor
If p(ND
NEW) > p(ND
BEST)
ND
BEST = ND
NEW
EndIf
EndRepeat
Return ND
BEST

Figure 3.32: Algorithm SearchConstructA.

95
3.3.4.2 Results
Algorithm SearchConstructA was run for 20 hours on each network. The performance
of the best network design produced for each network is summarised in Table 3.10.
Network
Performance Metric
A1 A2 A3 A4 B1
Traffic Cover (% of total) 2411 1542 2159 2270 489
Test Point Cover (% of total) 9633 9392 9419 9733 2316
Occupied Sites (% of total) 720 313 264 533 2857
Active Base Stations (% of total) 240 104 088 191 1062
Number of Design Evaluations 12776 9441 24837 9826 7471
Table 3.10: The performance of the best network designs for SearchConstructA.
Figure 3.33 shows the performance of the best network design plotted against the time
taken to produce it for NetworkA1.
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=24.11% Test-point Cover=96.33% Active Base Stations=2.40% Sites Occupied=7.20%
Traffic Cover
Test-point Cover
Active Base Stations
Sites Occupied

Figure 3.33: The performance of the best design produced vs. time for SearchConstructA on NetworkA1.
Figure 3.33 shows that the rate of improvement in traffic cover slows during algorithm
execution and that high test point cover is achieved early and maintained throughout the
run.
An interesting effect occurs early in the algorithms execution cycle. It may be described
by examining the first iteration of the algorithm on NetworkB1, in which 9 base stations
(out of the 273 that were tried) delivered the maximum possible traffic cover (limited by

96
the traffic capacity of a single base station, i.e. 58 Erlangs). The number of test points
covered by these base stations ranged between 74 and 1031, and, as the traffic cover for
each base station was equal, the base station with the highest test point cover was
accepted. If this was the final network design produced then this would be the best
choice, however, as the algorithm continues, the extra coverage achieved early on
becomes an obstacle to further coverage, i.e. it increases the likelihood of interference in
the network. Essentially, the dimensions (i.e. the ratio of traffic cover to test point
cover) of the early cells are distorted by the requirements of the evaluation function.
3.3.5 Algorithm: SearchConstructB
In Section 3.3.4.2 it was postulated that during the initial stages of execution of
SearchConstructA, high test point cover was detrimental to the performance of the final
network design. Algorithm SearchConstructB attempts to avoid this problem by a
temporary adjustment of the evaluation function.
3.3.5.1 Description
The following list shows the default evaluation criteria in order of significance:
1. Higher traffic cover
2. Higher traffic service
1

3. Higher test point cover
4. Higher test point service
2

5. Lower site occupancy
6. Lower number of active base stations

1
The traffic that would be covered in the absence of interference.
2
The test points that would be covered in the absence of interference.

97
At the beginning of SearchConstructBs execution, changes are made (highlighted in
italics in the following list) to the evaluation criteria:
1. Higher traffic cover
2. Higher traffic service
3. Lower test point service
4. Higher test point cover
5. Lower site occupancy
6. Lower number of active base stations
The reversal in requirement from highertest point service to lower test point service
is intended to encourage well-dimensioned cells to be chosen. Notice that the test point
cover metric is not reversed as this could also be achieved by (and hence encourage)
poor channel assignments it is, however, demoted to 4
th
in the ranking of metric
significance.
The algorithm pseudo-code is identical to that of SearchConstructA (see Figure 3.32), as
only the evaluation function has changed.
3.3.5.2 Results
Algorithm SearchConstructB was run for 20 hours on each network. The performance
of the best network design produced for each network is summarised in Table 3.11.
Network
Performance Metric
A1 A2 A3 A4 B1
Traffic Cover (% of total) 4847 3846 4312 4348 743
Test Point Cover (% of total) 3889 2859 4199 4389 1295
Occupied Sites (% of total) 1480 781 528 1107 4615
Active Base Stations (% of total) 493 260 176 369 1722
Number of Design Evaluations 26441 22939 49958 19032 11596
Table 3.11: The performance of the best network designs for SearchConstructB.

98
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 10000 20000 30000 40000 50000 60000 70000
Time Elapsed (seconds)
Final: Traffic Cover=48.47% Test-point Cover=38.89% Active Base Stations=4.93% Sites Occupied=14.80%
Traffic Cover
Test-point Cover
Active Base Stations
Sites Occupied

Figure 3.34: The performance of the best design produced vs. time for SearchConstructB on NetworkA1.
SearchConstructB significantly outperforms SearchConstructA in terms of traffic cover.
This is may be partly due to the avoidance of poorly dimensioned cells (as intended),
but an additional factor is the increase in the number of network designs evaluated. The
designs produced by SearchConstructB have less serviced test points (on average) than
the designs of SearchConstructA, and this means that less CIR calculations are required
and less time is taken during design evaluations. This speeding up of the evaluation time
led to more iterations and better final designs. This consequence of the enhancements
made to SearchConstructB highlights the (relative) high computational cost of the CIR
calculation.
3.3.6 Comparison of Search Algorithms
This section provides a comparison of the performance of the search algorithms. Table
3.12 gives a summary of the results for each algorithm.

99
Network Performance
Metric
Search Algorithm
A1 A2 A3 A4 B1
Average
SearchRandomB 5395 5058 4887 6055 469 4373
SearchRandomC 5088 5145 4953 6215 483 4377
SearchHillClimbA (Run 1) 5213 4797 5177 5254 536 4195
SearchHillClimbA (Run 2) 5412 3866 4318 5416 533 3909
SearchHillClimbA (Run 3) 5098 4978 5145 6114 589 4385
SearchConstructA 2411 1542 2159 2270 489 1774
Traffic
Cover
(% of total)
SearchConstructB 4847 3846 4312 4348 743 3625
SearchRandomB 5550 5819 5391 6785 1989 5107
SearchRandomC 5677 5709 5518 6860 1989 5151
SearchHillClimbA (Run 1) 6996 6153 7642 6948 1832 5914
SearchHillClimbA (Run 2) 5832 4742 4739 6156 2075 4709
SearchHillClimbA (Run 3) 5356 6031 5453 6866 2433 5228
SearchConstructA 9633 9392 9419 9733 2316 8099
Test Point
Cover
(% of total)
SearchConstructB 3889 2859 4199 4389 1295 3326
SearchRandomB 10000 9969 10000 10000 10000 9994
SearchRandomC 10000 10000 10000 10000 10000 10000
SearchHillClimbA (Run 1) 2080 1563 1215 2213 5385 2491
SearchHillClimbA (Run 2) 8480 8781 8856 8811 8462 8678
SearchHillClimbA (Run 3) 10000 10000 10000 10000 10000 10000
SearchConstructA 720 313 264 533 2857 937
Occupied
Sites
(% of total)
SearchConstructB 1480 781 528 1107 4615 1702
SearchRandomB 9960 8615 9472 9549 9487 9417
SearchRandomC 10000 10000 10000 10000 10000 10000
SearchHillClimbA (Run 1) 747 552 411 779 2088 915
SearchHillClimbA (Run 2) 5000 5000 5006 5027 5055 5013
SearchHillClimbA (Run 3) 9987 10000 9994 9886 10000 9973
SearchConstructA 240 104 088 191 1062 337
Active
Base Stations
(% of total)
SearchConstructB 493 260 176 369 1722 604
Table 3.12: A summary of results for all search algorithms on all networks.
The best average traffic cover and, hence, the best average network design (based on the
structure of the evaluation function) was produced by Run 3 of SearchHillClimbA,
however, it did not significantly outperform SearchRandomB or SearchRandomC.
Although SearchConstructA did not produce designs with high traffic cover, it did
quickly produce designs with high test point cover. In addition, it produced the designs
with the lowest number of occupied sites and active base stations. This algorithm may
be better suited to network design problems in which traffic capacity is not an issue, e.g.
the placement of television broadcast masts (see Section 1.2.3 for a description of
similar problems).
It is likely that a 20 hour run-time is not long enough to definitively separate the
performances of SearchRandomB, SearchRandomC, and SearchHillClimbA.

100
3.4 Factors Limiting Search Performance
In order to improve the performance of the search algorithms, it is useful to identify
common factors that limit their effectiveness and efficiency.
3.4.1 Computational Complexity
Computational complexity analysis (see [54] for a comprehensive background) provides
information about the scalability of a calculation. In Table 3.13 the evaluation of
network performance is broken down into distinct computational steps; the complexity
of each step is given in terms of the number of base stations N (equivalent to B ) and
the number of test points M (equivalent to R ).
Calculation Computational Complexity
Field strengths from active base stations to test points O(NM)
Serviced test point set O(NM)
Traffic volume in each base station cell O(M)
Fuzzy server list for each test point O(NM)
CIR for each test point O(N
2
M)
Covered test point set O(NM)
Covered traffic volume in each cell O(M)
Table 3.13: Computational complexity of network performance evaluation.
It can be seen that the computational complexity of most of the steps does not exceed
O(NM), and in fact the only step to exceed this complexity is the CIR calculation, which
has complexity O(N
2
M) due to the inclusion of fuzzy servers. This implies that, for large
networks at least, the CIR calculation will dominate the evaluation time.
3.4.2 The Effect of Poor Channel Assignments
It was shown in Section 3.2.3 that poor channel assignments can cripple the operational
performance of base stations. This effect can obscure the potential improvements in
adjusting other base station parameters.

101
3.4.3 Conclusion
To avoid the difficulties associated with the channel model (expensive CIR calculation
and bad channel assignments) it is desirable to split the problem into two stages:
1. Base station configuration without channel assignment (and its associated CIR
calculations), known as the Base Station Configuration Problem or BSCP (see
Section 1.2.3.2 for description).
2. Channel assignment for the network design produced by stage 1. This is known as
the Channel Assignment Problem or CAP (see Section 1.2.2 for description).
The division of the WCNDP into these two sub-problems is not novel, and in fact most
network design problems are explicitly formulated with this two-part approach (see
Section 1.2.4). The novelty lies in the opportunity to quantify the advantages this
approach may offer (Chapter 6 contains details of this comparison).
In the two-stage approach to the WCNDP, performance in the CAP phase depends
heavily on the design produced to solve the BSCP, because with high traffic and test
point service is useless unless it is relatively easy to create good channel assignments.
To do this, an estimate must be made (during the cell planning process) of how difficult
it will be to produce a good channel assignment. These difficulty metrics shall be
referred to as interference surrogates. The interference surrogate must be included in
the definition of the BSCP, and the network design evaluation criteria must be as
follows (see Section 2.3.2 to compare this with the full WCNDP evaluation function):
1. High traffic service.
2. High test point service.
3. Low estimated interference.
4. Low site occupancy.

102
5. Low number of active base stations.
The interference surrogate is shown in italics, and the structure of the evaluation is such
that it would only become a deciding factor in networks that had identical traffic and
test point service. The interference surrogate function must have a computational
complexity that does not exceed O(NM), as this will ensure that it scales at least as well
as the other calculations.

103
Chapter 4

Channel Assignment
This chapter describes a novel formulation of the Channel Assignment Problem (CAP).
Channels represent a continuous portion of the radio spectrum, and channel assignment
is the allocation of channels to transmitters in a radio network such that the interference
between channels is minimised. The central frequency in a channel is sometimes used to
uniquely identify it, and for this reason the CAP is also know as the Frequency
Assignment Problem (FAP). An introduction to the CAP is given in Section 1.2.2.
The structure of this chapter is as follows: Section 4.1 gives a definition of the CAP as
formulated in this thesis, Section 4.2 describes how CAP solutions (i.e. assignments) are
compared, Section 4.3 summarises the problem instances studied in this chapter, Section
4.4 introduces some algorithmic concepts that prove useful in the definition of the
optimisation algorithms described in Sections 4.5 and 4.6, and the performance of these
algorithms is compared in Section 4.7.
4.1 Problem Definition
The definition of the CAP used in this thesis is not necessarily the same as the
definitions traditionally used in the literature. It is tailored to the cell planning problem
as described previously in this thesis and may not generalise to all CAP requirements.
The definition of the problem is such that: the solution of the CAP will maximise the

104
performance of the cell plan from which it was derived. It is hoped that this
definition has sufficient commonality with other CAPs to be applicable beyond the field
of mobile telephone networks.
This thesis introduces some novel additions to the basic CAP, including:
A detailed calculation of the Carrier-to-Interference-Ratio (CIR) for each radio link.
Consideration of the relative importance of receivers, i.e. it may be more important
to cover some receivers than others.
An assessment of channel assignments based on two additional performance
metrics: average CIR of non-covered receivers, and average Interference-to-Carrier-
Ratio (ICR) of all receivers.
In the CAP it is assumed that transmitter locations and parameters (such as transmission
power) are fixed, and this accounts for all factors that affect the strength of the received
radio signal. Each transmitter requires the allocation of a fixed number of channels -
know as its demand. In cellular mobile telephone networks the demand of a transmitter
is related to the size (in terms of peak call-volume) of the transmitters cell (see Section
2.2.12).
The CAP is a subset of the Wireless Capacitated Network Design Problem or WCNDP
(see Section 1.1.2). Chapter 3 described a solution to the WCNDP in which channel
selection was regarded as one of many network configuration parameters, but in Section
3.4 it was postulated that there are advantages in separating the CAP from the rest of the
network design problem. Historically network design has been approached in this way
transmitter configurations have been established and then channels have been allocated.
Instances of the CAP are generated from network designs that require a channel
assignment. All of the CAP data sets used in this chapter are derived from the cell-
planning data, as described in Section 2.4, obtained from Operators A and B.

105
The mathematical formulation of the CAP has much in common with the WCNDP (as
derived in Section 1.1.2), but the scope for ambiguity is sufficient to warrant the
repetition of common definitions. This also allows the chapter to stand on its own for
readers who are exclusively interested in channel assignment.
4.1.1 Service Region
A service region comprises:
A set of receivers.
A set of transmitters.
A finite set of channels to allocate to transmitters.
4.1.2 Receivers
A service region contains a finite set of receivers R, where
i
r denotes the i
th
receiver in
R. Each receiver, r, has weight
WT
r , and this represents the importance of its received
signals freedom from interference. In the case of the mobile telephone data, the
receivers correspond to test points and the weights correspond to the amount of traffic at
those test points.
4.1.3 Channels
A service region includes the definition of a finite and ordered set of channels C, where
k
c denotes the k
th
channel in C. For convenience, the first and last channels,
1
c and
C
c ,
are also denoted by
MIN
C and
MAX
C respectively.
4.1.3.1 Channel Separation Gain
Receivers are tuned to receive signals on specific channels, so that any received signal
not employing this channel will be attenuated - the greater the difference (between the
channel to which the receiver is tuned and the channel on which the received signal was
transmitted), the greater the attenuation.

106
If a receiver tuned to channel C c
x
receives a signal transmitted on channel
C c
y
, the effective power of the received signal is attenuated. The effective gain
applied to the received signal is given by the function ( )
y x
CS
c c G , . The range of the
function is given by:
( ) 1 0
y x
CS
c c G ,
4.1.4 Transmitters
The service region contains a finite set of transmitters B, where
j
b denotes the j
th

transmitter in B. A set of channels is allocated to each transmitter. The k
th
channel
allocated to transmitter
j
b is denoted by
CH
jk
b . The number of channels allocated to
transmitter b is the demand, and it is denoted by
DM
b .
4.1.5 Field-Strength
The field-strength of the radio signal at receiver
i
r due to transmitter
j
b is denoted by
FS
ij
p . Note that this is a pre-calculated input of the CAP, and the method used to derive
these values is not important to the solution of the CAP.
4.1.6 Best Server
The best server of a receiver is the transmitter from which the signal with the highest
field-strength is received. If the received signal from two or more transmitters is equal,
then the one with the lowest index is chosen. For test point R r
i
, the best server is
denoted by
BS
i
r :
( ) { } x j p p p p b B b B b r
FS
ix
FS
ij
FS
ix
FS
ij j x j
BS
i
< = > = \ :
4.1.7 Fuzzy-Server Set
For test point
i
r a set of servers,
SS
i
r , is defined to contain all transmitters that provide a
signal within the fuzzy-server threshold (the fuzzy server model is discussed in Section
2.1.10). This threshold is the product of
FZ
g (a gain constant) and the signal from the
best server of
i
r :
{ }
FS
ix
FZ FS
ij j
SS
i
p g p B b r = :

107
where
x
BS
i
b r =
4.1.8 Carrier-to-Interference Ratio
The CIR

i
r for receiver R r
i
is the worst case CIR from each of the servers in the
fuzzy server set
SS
i
r :

ij
r b
i
r r
SS
i j

= min
where

ij
r denotes the worst case CIR found using each of the channels allocated to
transmitter
j
b :

ijk
b c
ij
r r
CH
j k

= min
where

ijk
r denotes the CIR at receiver
i
r when it is tuned to the k
th
channel of transmitter
j
b . It is given by the ratio of the server signal strength to the sum of all the other,
interfering, signals
1
:
N
S
ijk
p
p
r =


where
S
p is the strength of the serving signal:
FS
ij
S
p p =
and
N
p is the sum of all the interfering signals:
{
2
1
t
FS
ij
t
B b
N
x
N
p p p
x
=


3 2 1

where
N
x
p is the interference due to all the channels of base station
x
b :
( )

=
=
DM
x
b
y
CH
xy
CH
jk
CS FS
ix
N
x
b b G p p
1
,

1
Note that the contribution of interfering signals may be attenuated by channel separation.

108
Note that, for simplicity of expression,
N
p is composed of two terms:
1
t and
2
t . Term
1
t includes the server signal as part of the total noise; it is then explicitly removed by
2
t .
4.1.9 Receiver Cover
Receiver r is covered if its CIR is greater than or equal to the global CIR threshold
CV
. The set of covered receivers R R
CV
is given by:
{ }
CV CV
r R r R

= :
therefore the number of covered receivers,
CV
n , is given by:
CV CV
R n =
and the weighted receiver cover,
CV
w , is given by:


=
CV
R r
WT CV
r w
4.1.10 Average Non-Covered Carrier-to-Interference Ratio
The set of receivers not covered,
NC
R , is given by:
CV NC
R R R =
The average CIR of receivers that are not covered is denoted
NC
and given by:


=
NC
R r
NC
NC
r
R

1

4.1.11 Average Interference-to-Carrier Ratio (ICR)
The ICR of receiver R r is denoted

r and given by:

r
r
1
=
The global average ICR,
AV
, is given by:


=
R r
AV
r
R

1


109
4.1.12 Assignment
All the service region parameters (such as field strengths and receiver weights) are
fixed, except for the allocation of transmitter channels. A complete and valid allocation
of channels is known as an assignment.
4.2 Performance Evaluation
Comparing the performance of different assignments is done by comparing their
respective network performance metrics. Each performance metric has a relative
significance, and networks are compared metric-by-metric. They are listed below in
order of significance:
1. weighted receiver cover,
CV
w
2. number of receivers covered,
CV
n
3. average non-covered CIR,
NC

4. average ICR,
AV

Metrics 1, 2 and 3 should be maximised and metric 4 should be minimised.
The performance of an assignment, A, is expressed as a function ( ) A p . If network
configurations
a
A and
b
A produce weighted-cover-metrics
CV
a
w and
CV
b
w respectively
then
a
A outperforms
b
A if
CV
b
CV
a
w w > . However, if
CV
b
CV
a
w w = then a comparison
of the next significant metrics,
CV
a
n and
CV
b
n , must be performed, and if these two metrics
are also equal then comparison progresses to the next metric and so on. This
relationship may be expressed as follows:
( ) ( )
( )
( )
( )
( )
AV
b
AV
a
NC
b
NC
a
CV
b
CV
a
CV
b
CV
a
NC
b
NC
a
CV
b
CV
a
CV
b
CV
a
CV
b
CV
a
CV
b
CV
a
CV
b
CV
a
b a
n n w w
n n w w
n n w w
w w
A A


< = = =
> = =
> =
>
= > p p

110
This expression is not particularly clear, and it is better expressed algorithmically, i.e. as
a function that compares the performance of two assignments and returns the result as a
boolean (see Figure 4.1).
FunctionDef boolean greaterThan(
a
A ,
b
A )
If
CV
b
CV
a
w w > Return true
If
CV
b
CV
a
w w < Return false
If
CV
b
CV
a
n n > Return true
If
CV
b
CV
a
n n < Return false
If
NC
b
NC
a
> Return true
If
NC
b
NC
a
< Return false
If
AV
b
AV
a
< Return true
If
AV
b
AV
a
> Return false
Return false
EndFunctionDef

Figure 4.1: The comparison operator for assignment performance metrics.
4.2.1 The Role of the Average Non-Covered CIR Metric
Consider two channel assignments that perform identically with respect to receiver
cover and weighted receiver cover. Using the first assignment, the CIR for the non-
covered receivers is high, e.g. all receivers have a CIR that is 5% below the cover
threshold, whereas using the second assignment, the CIR for the non-covered receivers
is low, e.g. all receivers have a CIR below 50% of the cover threshold. Although the
predicted performance (in terms of cover) of both assignments is identical, the first
assignment would be preferable for the following reasons:
It must be remembered that this is only a simulation of the real network and, as
such, it contains a degree of uncertainty. In the real network, a signal within 5% of
the target CIR may be sufficient for communication.
If an existing transmitter were removed (say, for repair) then this may remove
enough noise for previously uncovered receivers to be covered, and this is more
likely if they are already as close as possible to the CIR threshold.

111
Optimisation algorithms prefer smooth evaluation functions. The inclusion of this
CIR metric informs the algorithm when it is getting warm (i.e. when the CIR of
non-covered receivers is getting closer to the cover threshold) even if no direct
improvement in cover occurs.
To these ends, the average CIR of non-covered receivers is compared, and not the
average ICR. This means that receivers with a CIR near the threshold will dominate the
calculation, and hence will be encouraged. This approach favours the majority of
receivers having borderline cover, with a minority of receivers having a very low CIR.
If the ICR were used instead, then the minority of points with very low CIR would
dominate the calculation (and hence be discouraged), leaving fewer receivers with
borderline cover. For the purpose of this thesis the former choice is preferred, but the
latter could be employed if a minority of test points with a very-low CIR were deemed
the greater evil.
4.2.2 The Role Of The Average-ICR Metric
Consider two assignments that both give 100% receiver cover. The receiver CIR values
generated by the first assignment are all low, e.g. 5% above the CIR threshold, whereas
the receiver CIR values generated by the second assignment are all high, e.g. twice the
CIR threshold. The second assignment is preferable because it offers a more robust
solution for the following reasons:
Errors in simulation data (such as radio signal strengths) can be greater, on average,
for the second assignment than for the first whilst maintaining 100% cover.
Additional, unforeseen, noise sources (such as solar activity) can be better tolerated,
on average, by the networks that employ the second assignment rather than the first.
Note that the average ICR of receivers is compared, and not the average CIR. This
encourages receivers with a CIR just above the target threshold to improve rather than
to encourage a few receivers to have CIRs that massively exceed the target threshold.

112
It is worth emphasising that the addition of this metric to the CAP means that
optimisation does not stop with full coverage. This is a novel extension to the fixed-
spectrum CAP.
4.3 Data Sets
All problem instances are derived from the network data sets described previously in
this thesis (see Section 2.4). Two CAPs were generated from each of the five networks.
4.3.1 Methodology of Creation
Two random network designs were created for each network using a modified version
of algorithm InitRandomA. The first of each pair of designs had 10% of the potential
base stations active, and the second had 90% active. The problem instances are known
as CapXX10 and CapXX90 respectively, where XX is the identifier of the network
(i.e. A1, A2, A3, A4, or B1).
The test point field strengths due to the active base stations were calculated for every
network design. This allowed the creation of the set of serviced test points that became
the set of CAP receivers. The base station demands were also computed, which enabled
the translation of the set of active base stations into the set of CAP transmitters.
4.3.1.1 Example
The network design that produced problem CapA210 is shown in Figure 4.2.

113

Figure 4.2: CAP problem CapA210 generated from NetworkA2.
4.3.2 Service Region Statistics
The ten problem instances used to test algorithms in this chapter are summarised in
Table 4.1.
Transmitter Statistics Receiver Statistics
Data Set
B

B b
DM
b
R

R r
WT
r
CapA110 75 259 29726 319623
CapA190 675 1019 29954 321091
CapA210 96 299 70777 269841
CapA290 864 1129 74845 278797
CapA310 170 372 17393 298811
CapA390 1533 1818 17393 298811
CapA410 73 225 48388 264922
CapA490 658 884 48512 265244
CapB110 27 74 4806 65633
CapB190 245 523 32555 934662
Table 4.1: Summary statistics for CAP problem instances.
Problem instance data not listed in Table 4.1, such as the number of available channels,
is inherited from the networks (as summarised in Table 2.2) from which the CAP
problems were generated.

114
4.4 Common Algorithmic Concepts
The section contains a few definitions that help simplify the algorithm description of the
optimisation algorithms.
4.4.1 Cover Function
Some algorithm descriptions require the definition of a function, ( ) A cover , which
returns the number of receivers covered by assignment A:
CV
n A : cover
4.4.2 Links
Links represent the connection from transmitter to receiver. Each link requires the
assignment of a single channel, so a transmitter has one link for every channel it
employs. The number of links in the network is given by:


=
B b
DM LK
b n
This convention means that instead of referring to the k
th
channel of the j
th
base station,
one may simply refer to the l
th
link in the network. The relationship between these
indices is given by:

=
+ =
1
1
j
x
DM
x
b k l
4.4.3 Magic Channels
A magic channel can cause no interference nor be interfered with. Magic channels
represent an intermediate step in the assignment process, and can be used to exclude the
influence of some transmitters from the evaluation of network performance. The final
assignment produced by an algorithm may not contain any magic channels if it is to be
realised in practice.
If a magic channel is denoted by m then:
( ) ( ) C c m c G c m G
CS CS
= = 0 , ,

115
4.5 Initialisation Algorithms
Initialisation is characterised by the creation of a single assignment, and no valid
intermediate assignments are created during an initialisation procedure.
4.5.1 Algorithm: InitRandomD
InitRandomD creates a valid channel assignment chosen from the set of all possible
assignments with a uniform probability.
4.5.1.1 Description
This algorithm, detailed in Figure 4.3, simply assigns a channel to every link, where
channels are selected at random from the set of all available channels, C.

Algorithm InitRandomD
Input: none
Output: channel assignment A


Declare channel assignment A
For every link l in A
Choose random channel c between C
MIN
and C
MAX

Allocate c to l
EndFor
Return A

Figure 4.3: Algorithm InitRandomD.
InitRandomD is used by SearchRandomD (see Section 4.6.1), and it was not thought
necessary to include a separate table of results for InitRandomD alone.
4.5.2 Algorithm: InitConstructA
InitConstructA is a non-deterministic channel assignment algorithm that produces a
single valid assignment at the end of execution. During execution it produces
assignments that contain magic channels.
4.5.2.1 Description
At the beginning of the algorithms execution, every transmitter is allocated magic
channels. The algorithm progresses until all magic channels have been replaced with

116
real channels. Channel allocation occurs in two phases: the allocation of channels such
that no loss of cover occurs; and the allocation of channels such that the minimum loss
of cover occurs. In the first phase, channels are allocated in order from
MIN
C to
MAX
C .
A channel is allocated to a link and the network performance is evaluated, if cover is
lost then the link is re-assigned a magic channel. This continues until the channel has
been tried for all the links, at which point testing moves to the next channel. The order
in which links are allocated channels is decided by the demand of the transmitter from
which they came, as this gives high demand transmitters a good opportunity to spread
across the available spectrum. In the second phase, all links that have not been allocated
a channel (i.e. still have magic channels) are given the channel that causes the least loss
in cover (the order in which links are allocated is the same as in the first phase). The
algorithm is detailed in Figure 4.4.

117

Algorithm InitConstructA
Input: none
Output: channel assignment A
BEST

Declare channel assignment A
BEST
For every link l in A
BEST
Assign magic channel to l
EndFor
Create list of all links, links
Sort links by host transmitter demand
For every channel c from C
MIN
to C
MAX

For every link l in links
If channel assigned to l is magic
Declare assignment A
NEW = A
BEST
Assign c to l in A
NEW
If cover(A
NEW) = cover(A
BEST)
A
BEST = A
NEW
EndIf
Endif
EndFor
EndFor
For every link l in links
If channel assigned to l is magic
Assign C
MIN
to l in A
BEST
For every channel c from C
MIN
+1 to C
MAX

Declare assignment A
NEW = A
BEST
Assign c to l in A
NEW
If p(ANEW) > p(ABEST)
A
BEST = A
NEW
EndIf
EndFor
Endif
EndFor
Return A
BEST

Figure 4.4: Algorithm InitConstructA.

118
4.5.2.2 Results
InitConstructA was run once for each problem and the results are shown in Table 4.2.
Cap
Weighted
RX Cover
(% of total)
RX Cover
(% of total)
Average Non-
Covered CIR
Average ICR
Time
Elapsed
(seconds)
A110 9446 9350 1354 (1132dB) 003 (1591dB) 17617
A190 9730 9623 1410 (1149dB) 002 (1808dB) 148360
A210 7655 7809 1147 (1060dB) 005 (1270dB) 140115
A290 8856 8940 1211 (1083dB) 003 (1602dB) 795548
A310 6324 6714 1053 (1023dB) 009 (1061dB) 49890
A390 8939 8824 1200 (1079dB) 003 (1555dB) 976560
A410 8272 8257 1160 (1064dB) 005 (1320dB) 39250
A490 8525 8973 1144 (1059dB) 003 (1579dB) 267721
B110 9968 9694 497 (697dB) 001 (1833dB) 141
B190 7253 7359 531 (725dB) 015 (827dB) 168991
Table 4.2: The performance of the best assignments produced by InitConstructA.
For problem CapB110, the speed with which a solution was found and the high quality
of the assignment produced indicates that this problem was not significantly
challenging. It would be imprudent to use the results obtained for this problem instance
for comparing the performance of algorithms.
The weighted cover due to all the assignments produced by InitConstructA did not fall
below 60%, the average being 8333% (excluding CapB110). The time taken to
produce an assignment is large and varies significantly with the size of the network. It is
worth noting that the total time taken to generate all these results was more than 30
days.
4.6 Search Algorithms
Searching is characterised by the creation of a set of valid channel assignments. In
assessing the performance of a search algorithm, two factors are considered: the
performance of the best assignment in the set; and the time taken to produce the whole
set. For the purposes of algorithm comparison, each search algorithm is run for a fixed
length of time so that the performance of the best assignment produced by each
algorithm can then be compared on an equal basis.

119
4.6.1 Algorithm: SearchRandomD
SearchRandomD is a pure random search. Assignments are generated by calling
InitRandomD, and successive assignments are independent of those that have been
created previously.
4.6.1.1 Description
The algorithm, detailed in Figure 4.5, begins with a call to InitRandomD that creates an
initial assignment, this is recorded as the best assignment so far. Control then passes
to a loop, where it remains until a fixed time (specified by the user) has expired. In the
loop, InitRandomD is called and the best assignment so far is replaced if its
performance is exceeded by this new assignment.

Algorithm SearchRandomD
Input: none
Output: channel assignment A
BEST
Note: may be pre-emptively terminated

Declare channel assignment A
BEST = InitRandomD()
While time limit not exceeded
Call InitRandomD
Record assignment as A
NEW
If p(A
NEW) > p(A
BEST)
A
BEST = A
NEW
EndIf
EndWhile
Return A
BEST

Figure 4.5: Algorithm SearchRandomD.
4.6.1.2 Results
SearchRandomD was applied to each problem for 10 hours, the results are summarised
in Table 4.3.

120
Cap
Weighted
RX Cover
(% of total)
RX Cover
(% of total)
Average Non-
Covered CIR
Average ICR Iterations
A110 4546 4506 475 (677dB) 052 (287dB) 3916
A190 4452 4888 641 (807dB) 025 (607dB) 1144
A210 3117 3430 501 (700dB) 047 (331dB) 1483
A290 4774 5707 642 (807dB) 023 (637dB) 485
A310 2371 2453 477 (678dB) 055 (262dB) 4160
A390 4691 4969 722 (859dB) 022 (648dB) 1212
A410 2999 3265 540 (732dB) 038 (421dB) 2747
A490 5561 6175 706 (849dB) 018 (746dB) 846
B110 8193 8477 499 (698dB) 012 (938dB) 94024
B190 1525 2800 314 (497dB) 071 (150dB) 1359
Table 4.3: The performance of the best assignments produced by SearchRandomD.
Figure 4.6 shows the performance of the best assignment produced by SearchRandomD
plotted against the time taken to produce it for problem CapA290.
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 5000 10000 15000 20000 25000 30000 35000
Time Elapsed (seconds)
-8dB
-6dB
-4dB
-2dB
0dB
2dB
4dB
6dB
8dB
10dB
Final: Weighted Cover=47.74% Receiver Cover=57.07% Average N/C CIR=8.07dB Average ICR=-6.37dB
Weighted Cover
Receiver Cover
Average N/C CIR
Average ICR

Figure 4.6: The performance of the best assignment produced vs. time for SearchRandomD on
CapA290.
The average weighted cover due to SearchRandomD was 3782% (excluding CapB110
as recommended in Section 4.5.2.2).
4.6.2 Algorithm: SearchHillClimbB
SearchHillClimbB is a local-search algorithm, i.e. an initial assignment is modified by
single changes to assigned channels. These adjustments may continue until no one-

121
change to an assigned channel produces an improvement in assignment performance.
This final assignment is said to be a local optima.
Local-search algorithms require the definition of a search neighbourhood, this is the
scope of changes made to input parameters during searching. For algorithm
SearchHillClimbB, the neighbourhood of an assignment is defined as follows:
Assignment A
NEW
is in the neighbourhood of A
OLD
if A
OLD
can be transformed into A
NEW

by the re-assignment of only one transmitter channel. The transformation of one
assignment into another, neighbouring, assignment is known as a move.
4.6.2.1 Description
Algorithm SearchHillClimbB, as detailed in Figure 4.7, has two separate control
sequences:
1. The set of all possible moves is generated.
2. Each move is applied and the resulting assignment evaluated.
This cycle is repeated until a complete set of moves has been applied with no
improvement in assignment performance. The list of moves is executed in random
order, which reduces bias in the hill climbing procedure.
Moves are of the form: <link> <channel>

122

Algorithm SearchHillClimbB
Input: channel assignment A
INIT
Output: channel assignment A
BEST
Note: may be pre-emptively terminated

Declare channel assignment A
BEST = A
INIT
Repeat until no change is made to ABEST
Create empty move list, moves
For every link l in A
BEST
For every channel c from C
MIN
to C
MAX
Add move l c to moves
EndFor
EndFor
Shuffle moves
For every move m in moves
A
NEW = A
BEST + m
If p(A
NEW) > p(A
BEST)
A
BEST = A
NEW
EndIf
EndFor
EndRepeat
Return A
BEST

Figure 4.7: Algorithm SearchHillClimbB.
4.6.2.2 Results
SearchHillClimbB was applied to each problem for 10 hours, and the results are
summarised in Table 4.4. The initial assignments used were generated by InitRandomD.
Cap
Weighted
RX Cover
(% of total)
RX Cover
(% of total)
Average Non-
Covered CIR
Average ICR Iterations
A110 9395 9450 1168 (1067dB) 002 (1652dB) 3935
A190 6631 7108 818 (913dB) 011 (972dB) 1143
A210 5750 6461 763 (883dB) 017 (778dB) 1483
A290 5335 6258 758 (879dB) 016 (801dB) 484
A310 6005 6382 981 (992dB) 009 (1031dB) 4165
A390 5567 5805 777 (89dB) 017 (772dB) 1210
A410 7530 7496 784 (894dB) 021 (676dB) 2753
A490 6520 6833 737 (867dB) 014 (865dB) 847
B110 9998 9759 486 (687dB) 001 (2057dB) 14801
B190 3068 4123 379 (579dB) 069 (162dB) 1360
The algorithm terminated after 5647 seconds when the solution was found to be a local optima
Table 4.4: The performance of the best assignment produced by SearchHillClimbB.
Figure 4.8 shows the performance of the best assignment produced by
SearchHillClimbB plotted against the time taken to produce it for problem CapA390.

123
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 5000 10000 15000 20000 25000 30000 35000
Time Elapsed (seconds)
-10dB
-8dB
-6dB
-4dB
-2dB
0dB
2dB
4dB
6dB
8dB
10dB
Final: Weighted Cover=53.35% Receiver Cover=62.58% Average N/C CIR=8.79dB Average ICR=-8.01dB
Weighted Cover
Receiver Cover
Average N/C CIR
Average ICR

Figure 4.8: The performance of the best assignment produced vs. time for SearchHillClimbB on
CapA390.
The average weighted cover due to SearchHillClimbB was 6200% (excluding
CapB110 see Section 4.5.2.2).
4.7 Comparison of Results
InitConstructA produced the best assignments for all the CAP problems except for
CapB110, but it is likely that this problem was too easy to allow a fair comparison of
algorithms, and it will be excluded from further results comparison.
Of the two search algorithms
1
described in this chapter SearchHillClimbB performed
the best, as it produced the best assignment for all CAP problems over the 10 hour run-
time It gave an average weighted cover of 6200% compared with 3782%, which was
produced by SearchRandomD.
Of all the assignments produced by InitConstructA, only the one for CapA110 took
less than ten hours. This limits the scope for comparison between InitConstructA and
the search algorithms. It took just under five hours for InitConstructA to produce a
superior assignment to the one produced in ten hours by SearchHillClimbB. In addition,

1
Note that this comparison does not include the initialisation algorithms.

124
it is likely that the assignment produced for CapA410 by InitConstructA would still
outperform the one produced by SearchHillClimbB, even if the search algorithm were
given the extra time taken by InitConstructA. It is likely that InitConstructA is the most
computationally efficient way of assigning channels for this type of CAP (a conclusion
supported by the results obtained for problem CapA110), but the disadvantage of
InitConstructA is that no valid assignment is produced until the algorithm finally
terminates. Furthermore, the time it will take to terminate is not easily predicted.
Given an unlimited time in which to produce assignments, the most efficient method (of
those tested) is likely to be the application of InitConstructA followed by
SearchHillClimbB. If assignments were required in a fixed time, then the best approach
would be to call InitRandomD followed by SearchHillClimbB.

125
Chapter 5

Interference Surrogates
This chapter compares three different Interference Surrogates (ISs). In Section 3.4 it
was concluded that optimisation performance was limited by two factors:
The Carrier-to-Interference Ratio (CIR) calculation was dominating the time taken
to evaluate each network design.
Poor channel assignments were masking potential network design improvements
made by optimisation algorithms.
It was concluded that the loss of cover (due to interference) should be estimated using a
stand-in, or surrogate, function. This function should have a computational complexity
that is less than or equal to the complexity of the other functions required for network
simulation. In addition, the value of the surrogate function should be independent of
assigned channels. In this chapter, three ISs are investigated that meet these criteria:
Overlap a measure of the density of cell overlap in a network.
Non-dominance a measure of the server ambiguity seen by mobile stations.

126
Signal Freedom an estimate of the communication channels isolation from
interference.
The first two ISs are currently used in industry and they are employed by the operators
from whom the data for this thesis was acquired. The third surrogate is novel, and it was
derived from detailed analysis of the relationship between cell planning and channel
assignment.
5.1 Role of the Interference Surrogate
When comparing two candidate network designs, performance metrics are compared
one-at-a-time (see Section 2.3.2). When employing an IS during cell planning the order
of comparison is as follows:
1. Traffic service.
2. Test point service.
3. Interference surrogate performance.
4. Number of sites occupied.
5. Number of base stations active.
From this list it can be seen that the IS metric is only used when comparing network
designs with the same traffic and test point service. Suppose two such network designs
A and B have interference surrogate performances I
A
and I
B
respectively. A channel
assignment is produced for A that results in a test point cover CV
A
and a channel
assignment is produced for B that results in cover CV
B
. Now, if CV
A
> CV
B
then this

127
should be reflected by the IS, i.e. I
A
> I
B
. This relationship will be known as the IS
Condition.
1


Interference
Surrogate
Metric
Test point Cover

Figure 5.1: The performance of an ideal (hypothetical) interference-surrogate.
Figure 5.1 is a graph that indicates the performance of an IS. The value of the IS metric
is plotted against the test point cover (post channel-assignment) for 10 different network
designs. Designs with a high IS value have a correspondingly high test point cover. No
pair of network designs violate the IS Condition, which indicates that the IS is behaving
ideally.
5.1.1 Optimum Cover vs. Fixed Time Cover
The test point cover produced for a network will depend on the channel assignment
algorithm used and for how long it is run, this implies that there are two possible test
point covers to consider:
1. The cover due to the optimal channel assignment.
2. The best cover achievable in a fixed time.

1
Note that the > symbol used to compare ISs should be read as better than. It may or may not be
literally greater than as this will depend on whether the IS is designed to be maximised (as with signal
freedom) or minimised (as with overlap).

128
There is no known method that can deliver the first of these figures in a reasonable time.
To generate the second of these figures the best-known fixed time algorithm
(SearchHillClimbB) can be used. The longer the algorithm is run, the closer the result
will be to the optimum cover (although not guaranteed to ever equal it). The fixed time
performance is more realistic than the optimum performance because all real channel
assignments are produced in a fixed time.
5.2 Interference Surrogate Descriptions
The following section contains definitions of the three ISs investigated in this chapter.
The notation used for the mathematical descriptions is based on the definitions in
Chapter 2.
5.2.1 Overlap
Consider a subscriber served by a single base station, if that subscriber receives radio
signals from other base stations then the serving signal may be interfered with, i.e. the
CIR may fall below an acceptable threshold. The only way to attenuate the effective
power of the interfering signals is by channel separation, i.e. separation between those
channels used by interfering base stations and the channel on which the serving signal is
transmitted. The greater the number of interfering-signals that overlap serving signals,
the greater the number of channel separations required. The implicit assumption of the
overlap measure is that these increases in channel separation requirements will increase
the difficulty of channel assignment.
A non-serving signal is considered to overlap at a test point if it is above the test points
service threshold. The number of overlaps at each test point is summed to give the total
network overlap. If the set of base stations that overlap at test point
SV
i
R r is
denoted
OV
i
r and is given by:
{ }
SQ
i
FS
ij
AC
j
OV
i
r p B b r = :
then the total network overlap is given by:

129


=
SV
R r
OV OV
r I 1
Note that the serving base station does not contribute to the overlap, hence the
subtraction of 1 from the overlap figure for each test point.
The overlap metric is designed to be minimised.
5.2.2 Non-Dominance
The non-dominance metric takes account of the relative contribution of interfering
signals. The total noise a subscriber receives can be high as long as the server signal is
significantly higher. A subscribers server is said to be non-dominant if the number of
interferers within a certain threshold ( y dB) of the server signal is greater than x . Non-
dominance is sometimes known by its defining parameters, in the form x of y .
The set of base stations, denoted by
ND
i
r , that provide test point
SV
i
R r with a signal
within the non-dominance threshold is given by:
{ }
FS
iz
FS
ij
AC
j
ND
i
p y p B b r = :
where
z
BS
i
b r =
the set of test points with non-dominant servers, denoted by
ND
R , is given by:
{ } x r R r R
ND SV ND
> = :
and the total non-dominance is given by:
ND ND
R I =
For the purpose of this thesis, x was chosen to be 7 and y was chosen to be -9dB.
The non-dominance metric is designed to be minimised.

130
5.2.3 Signal Freedom
Signal freedom is a novel metric. In the signal freedom paradigm, networks are viewed
as a collection of links to which channels must be assigned
1
. The fact that links might
share base stations and that base stations might share sites is regarded as un-important.
Links simply represent conduits into the network.
5.2.3.1 A Single Test point with a Single Link
Consider a single test point r, a signal will be received by r for every active link
2
in the
network. Figure 5.2 shows the strengths of four such signals received at r, they are the
serving signal S and the three interfering signals
0
I ,
1
I , and
2
I .

P
O
W
E
R

Server Signal
Power
S
Target Noise
(with channel
separation)
S /
Total Noise (no
channel separation)
I
0
+ I
1
+ I
2
I
0
I
1
I
2
S

Figure 5.2: An illustration of the attenuation of interference by channel separation.
If the target CIR is then the total interference cannot exceed S (known as the
target noise). The interfering signals must be attenuated such that the sum of all three is
less than the target noise. The greater the required attenuation, the greater the channel
separation will need to be. Therefore the ratio of the target noise to the total

1
The concept of links is the same as the one used to simplify the description of channel assignment
algorithms in Section 4.4.2.
2
An active link represents a utilised channel of an active base station.

131
interference gives an estimate of the effort required in assigning channels. For the
example in Figure 5.2, is given by:
( )
2 1 0
I I I
S
+ +
=


Note that is the normalised CIR at r, therefore if 1 then no channel separation
would be required to cover r.
5.2.3.2 Normalised CIR Threshold
When calculating signal freedom, test point links for which 1 > are taken to have
1 = . This is to stop easy links (i.e. those requiring no channel separation) from
swamping the values of more demanding links in the final total.
5.2.3.3 Number of Links vs. Channel Separation
Signal freedom takes account of networks with different numbers of links by including
contributions from all possible links. Un-used links are considered perfect as they can
be assigned any channel. The two types of redundant links are:
Links hosted by inactive base stations.
Links that are not required to carry traffic but are hosted by an active base station.
A redundant link contributes a perfect value of 1 to the signal freedom metric.
Two factors control the difficulty of a channel assignment: the number of links that
require channels; and the separation requirements between the links. Consider two
example network designs that are both equally difficult to assign channels for:
Design 1. A design requiring many links, in which the channel assigned to each
link has a small impact on the channels assigned to other links.

132
Design 2. A design requiring few links, in which the channel assigned to each link
has a large impact on the channels assigned to other links.
If the signal freedom metric was simply the average of the contributions from active
links (and not the average of all possible links) the first design would always be
favoured, even though the assignment of channels is equally difficult for both (as
postulated).
5.2.3.4 Mathematical Derivation
The total field-strength of radio signals received by test point
SV
i
R r is denoted
TN
i
p
and is given by:


=
AC
x
B b
DM
x
FS
ix
TN
i
b p p
therefore the normalised CIR,
ij
, due to a channel of base station
j
b serving test point
i
r is given by:
( )
FS
ij
TN
i
CV
FS
ij
ij
p p
p

=


To calculate the signal freedom metric, a threshold must be applied to (see Section
5.2.3.2). In addition, an value of 1 should be assumed for all in-active links (see
Section 5.2.3.3). Both these requirements are encapsulated in the calculation of
ijk
,
which represents the signal freedom of the k
th
channel of base station
j
b serving test
point
i
r :
{ }


=
otherwise
if
1
1
DM
j
SS
i j ij
ijk
b k r b

, min

Hence, the signal freedom,
SF
i
r , for test point
i
r is given by the average of the signal
freedom due to every channel of every base station:

=

=
B b
M
k
ijk
SF
i
j
M B
r
1
1


133
where
CHANPERBS
n M =
The final signal freedom metric,
SF
I , is the average of the signal freedom for each of
the test points, i.e.


=
SV
i
R r
SF
i
SV
SF
r
R
I
1

The signal freedom metric is designed to be maximised.
Note, the computational complexity of the signal freedom metric is O(NM) (as
required). It manages to be less than that of the CIR calculation because the total noise
can be reused for every server in the fuzzy server set (as it is not effected by channel
assignment).
5.3 Methodology of Comparison
To test the performance of an IS it is necessary to determine how well it meets the IS
Condition, i.e. how reliably does the IS predict the post channel-assignment test point
cover?
For each network a set of ten designs with maximal test point and traffic service was
produced, and then a channel assignment was generated for each design (and the
resulting test point cover recorded). Designs were compared pair-wise to determine the
performance of the IS.

134

Interference
Surrogate
Metric
Test point Cover
e
d
1
d
2

Figure 5.3: The performance of an imperfect interference surrogate, in which one pair of designs violate
the IS Condition.
The IS illustrated in Figure 5.1 behaves ideally (i.e. the IS Condition is never violated).
Figure 5.3 shows a similar scenario except with a different IS. The IS value for d
1
is
below that of design d
2
despite d
1
s superior test point cover (i.e. a violation of the IS
Condition). This difference in cover (denoted by e in the diagram) shall be referred to as
an inaccuracy.
The difference in cover between all pairs of designs that violate the IS Condition shall
be referred to as the inaccuracy of the IS. Note that, for the purposes of calculating an
average, pairs of designs that do not violate the IS Condition are considered to have an
inaccuracy of zero.

135

Interference
Surrogate
Metric
Test point Cover
e
3
e
2
e
1

Figure 5.4: The performance of an imperfect interference surrogate, in which three pairs of designs
violate the IS Condition.
Figure 5.4 shows the test point cover for ten network designs; three pairs of designs
violate the IS Condition giving inaccuracies of e
1
, e
2
, and e
3
. The total number of
possible (unique) pairs is given by:
45
2
10 10
2
=

pairs
therefore the average inaccuracy of the IS is given by:
45
3 2 1
e e e + +

and the maximum inaccuracy is given by:
{ }
3 2 1
e e e , , max
5.3.1 Generation of Network Designs
For each network, a set of ten random network designs was created, in which a fixed
fraction of the potential base stations were active: 20 1 for the first design, 20 3 for the
next design and so-on until the last design, which used the fraction 20 19 . A high
variation in base station activity was chosen to give the widest range of final test point
covers, as this broadens the range of inaccuracies with which to compare ISs.

136
Each design was hill-climbed until both the traffic and test point service was within 1%
of the maximum possible; this ensured that the network designs were close enough in
terms of traffic and test point service performance to allow a fair comparison of the IS
values. Initially the requirement for service was set at 100%, however, it was found that
a large portion of the hill-climbing algorithms runtime was used to achieve the last 1%
(i.e. from 99% to 100%). Note that NetworkB1 was not included in the testing process
due to the impossibility of achieving 99% test point or traffic service (see Section
2.4.5).
5.3.2 Channel Assignment
A channel assignment was generated for each of the network designs using algorithm
InitRandomD. Control then passed to SearchHillClimbB, which ran for 10 hours per
network design. This combination was chosen because it was shown to deliver the best
fixed-time performance (see Section 4.7).
5.4 Results
Results are presented for each of the four networks tested, and a summary of the results
is also included to assist comparison. For the duration of each ten hour channel
assignment, the best test point cover was recorded every second; this gave one second
snapshots of test point cover, each of which could be used to calculate the inaccuracy
of the IS. Table 5.1 shows six such snapshots of the performance of an IS.
Time
(s)
Test point
cover for
design 1
(% of total)
Test point
cover for
design 2
(% of total)


Test point
cover for
design 10
(% of total)
Average IS
inaccuracy
(% of total)
Max IS
inaccuracy
(% of total)
Worst Case
IS
inaccuracy
(% of total)
393 2708 2860 6010 00410 205 1342
394 2708 2860 6013 00386 193 1347
395 2708 2860 6013 00386 193 1347
396 2708 2860 6013 00386 193 1347
397 2710 2860 6013 00383 193 1220
398 2710 2868 6013 00324 181 1220
Table 5.1: An example of IS performance vs. time.
The worst case IS inaccuracy figure quoted in Table 5.1 is the IS value that would be
obtained if the IS condition were violated for every pair of designs, and it is used to

137
normalise the IS values. These normalised values are plotted instead of the actual values
in the IS-versus-time graphs to allow comparison between different networks.
Although the plotting of IS performances over time is an interesting result, the
important comparison of ISs comes from the final IS inaccuracy values, i.e. those
obtained after the 10 hour run-time.
5.4.1 NetworkA1
The IS inaccuracy values for the ten network designs of NetworkA1 are given in Table
5.2. Overlap, non-dominance, and signal freedom are denoted OV, ND, and SF
respectively.
Max Inaccuracy in Test
Point Cover
Average Inaccuracy in
Test Point Cover
Normalised Average
Inaccuracy
OV ND SF OV ND SF OV ND SF
1659% 1353% 516% 645% 185% 037% 095 027 005
Table 5.2: The final IS inaccuracy values for NetworkA1.
The change in IS inaccuracy over time for the three ISs is shown in Figure 5.5. The
initial and final conditions for each IS are shown in Figure 5.6, Figure 5.7, and Figure
5.8.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 5000 10000 15000 20000 25000 30000 35000
Time (s)
N
o
r
m
a
l
i
s
e
d

I
n
-
a
c
c
u
r
a
c
y
Overlap
Signal Freedom
Non-Dominance

Figure 5.5: The normalised inaccuracy of three ISs vs. time for NetworkA1.

138
0
200000
400000
600000
800000
1000000
1200000
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
O
v
e
r
l
a
p
Initial Cover
Final Cover

Figure 5.6: Graph of overlap vs. test point cover for 10 designs in NetworkA1.
0
500
1000
1500
2000
2500
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
N
o
n
-
D
o
m
i
n
a
n
c
e
Initial Cover
Final Cover

Figure 5.7: Graph of non-dominance vs. test point cover for 10 designs in NetworkA1.

139
0.9984
0.9985
0.9986
0.9987
0.9988
0.9989
0.9990
0.9991
0.9992
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
S
i
g
n
a
l

F
r
e
e
d
o
m
Initial Cover
Final Cover

Figure 5.8: Graph of signal freedom vs. test point cover for 10 designs in NetworkA1.
5.4.2 NetworkA2
The IS inaccuracy values for the ten network designs of NetworkA2 are given in Table
5.3.
Max Inaccuracy in Test
Point Cover
Average Inaccuracy in
Test Point Cover
Normalised Average
Inaccuracy
OV ND SF OV ND SF OV ND SF
1145% 1127% 260% 376% 299% 025% 091 072 006
Table 5.3: The final IS inaccuracy values for NetworkA2.
The change in IS inaccuracy over time for the three ISs is shown in Figure 5.9. The
initial and final conditions for each IS are shown in Figure 5.10, Figure 5.11, and Figure
5.12.

140
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 5000 10000 15000 20000 25000 30000 35000
Time (s)
N
o
r
m
a
l
i
s
e
d

I
n
-
a
c
c
u
r
a
c
y
Overlap
Signal Freedom
Non-Dominance

Figure 5.9: The normalised inaccuracy of three ISs vs. time for NetworkA2.
0
500000
1000000
1500000
2000000
2500000
3000000
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
O
v
e
r
l
a
p
Initial Cover
Final Cover

Figure 5.10: Graph of overlap vs. test point cover for 10 designs in NetworkA2.

141
0
1000
2000
3000
4000
5000
6000
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
N
o
n
-
D
o
m
i
n
a
n
c
e
Initial Cover
Final Cover

Figure 5.11: Graph of non-dominance vs. test point cover for 10 designs in NetworkA2.
0.9991
0.9991
0.9992
0.9992
0.9993
0.9993
0.9994
0.9994
0.9995
0.9995
0.9996
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
S
i
g
n
a
l

F
r
e
e
d
o
m
Initial Cover
Final Cover

Figure 5.12: Graph of signal freedom vs. test point cover for 10 designs in NetworkA2.
5.4.3 NetworkA3
The IS inaccuracy values for the ten network designs of NetworkA3 are given in Table
5.4.
Max Inaccuracy in Test
Point Cover
Average Inaccuracy in
Test Point Cover
Normalised Average
Inaccuracy
OV ND SF OV ND SF OV ND SF
1954% 482% 433% 589% 049% 031% 083 007 004
Table 5.4: The final IS inaccuracy values for NetworkA3.

142
The change in IS inaccuracy over time is shown in Figure 5.13 for the three ISs. The
initial and final conditions for each IS are shown in Figure 5.14, Figure 5.15, and Figure
5.16.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 5000 10000 15000 20000 25000 30000 35000
Time (s)
N
o
r
m
a
l
i
s
e
d

I
n
-
a
c
c
u
r
a
c
y
Overlap
Signal Freedom
Non-Dominance

Figure 5.13: The normalised inaccuracy of three ISs vs. time for NetworkA3.
0
500000
1000000
1500000
2000000
2500000
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
O
v
e
r
l
a
p
Initial Cover
Final Cover

Figure 5.14: Graph of overlap vs. test point cover for 10 designs in NetworkA3.

143
0
500
1000
1500
2000
2500
3000
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
N
o
n
-
D
o
m
i
n
a
n
c
e
Initial Cover
Final Cover

Figure 5.15: Graph of non-dominance vs. test point cover for 10 designs in NetworkA3.
0.9993
0.9993
0.9994
0.9994
0.9995
0.9995
0.9996
0.9996
0.9997
0.9997
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
S
i
g
n
a
l

F
r
e
e
d
o
m
Initial Cover
Final Cover

Figure 5.16: Graph of signal freedom vs. test point cover for 10 designs in NetworkA3.
5.4.4 NetworkA4
The IS inaccuracy values for the ten network designs of NetworkA4 are given in Table
5.5.
Max Inaccuracy in Test
Point Cover
Average Inaccuracy in
Test Point Cover
Normalised Average
Inaccuracy
OV ND SF OV ND SF OV ND SF
2632% 992% 248% 925% 090% 009% 090 009 001
Table 5.5: The final IS inaccuracy values for NetworkA4.

144
The change in IS inaccuracy over time for the three ISs is shown in Figure 5.17. The
initial and final conditions for each IS are shown in Figure 5.18, Figure 5.19, and Figure
5.20.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1
0 5000 10000 15000 20000 25000 30000 35000
Time (s)
N
o
r
m
a
l
i
s
e
d

I
n
-
a
c
c
u
r
a
c
y
Overlap
Signal Freedom
Non-Dominance

Figure 5.17: The normalised inaccuracy of three ISs vs. time for NetworkA4.
0
500000
1000000
1500000
2000000
2500000
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
O
v
e
r
l
a
p
Initial Cover
Final Cover

Figure 5.18: Graph of overlap vs. test point cover for 10 designs in NetworkA4.

145
0
500
1000
1500
2000
2500
3000
3500
4000
4500
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
N
o
n
-
D
o
m
i
n
a
n
c
e
Initial Cover
Final Cover

Figure 5.19: Graph of non-dominance vs. test point cover for 10 designs in NetworkA4.
0.9985
0.9986
0.9987
0.9988
0.9989
0.9990
0.9991
0.9992
0.9993
0.9994
0 10 20 30 40 50 60 70 80 90 100
Test Point Cover (% of total)
S
i
g
n
a
l

F
r
e
e
d
o
m
Initial Cover
Final Cover

Figure 5.20: Graph of signal freedom vs. test point cover for 10 designs in NetworkA4.

146
5.4.5 Summary
The results are summarised in Table 5.6. For each result type, a maximum and average
(taken across all four networks) is included.
Maximum Inaccuracy in
Test Point Cover
Average Inaccuracy in
Test Point Cover
Normalised Average
Inaccuracy Scope
OV ND SF OV ND SF OV ND SF
NetA1 1659% 1353% 516% 645% 185% 037% 095 027 005
NetA2 1145% 1127% 260% 376% 299% 025% 091 072 006
NetA3 1954% 482% 433% 589% 049% 031% 083 007 004
NetA4 2632% 992% 248% 925% 090% 009% 090 009 001

Max 2632% 1353% 516% 925% 299% 037% 095 072 006
Average 1848% 989% 364% 634% 156% 026% 090 029 004
Table 5.6: A summary of IS inaccuracy results.
In comparing any two pairs of network designs, the signal freedom metric was never
incorrect (i.e. in violation of the IS condition) by more than 516% (in terms of actual
test point cover), non-dominance was never incorrect by more than 1353%, and overlap
by 2632%. The average of the maximum discrepancies for all four networks was 364%
for signal freedom, 989% for non-dominance, and 1848% for overlap.
The average in-accuracy of all network-design-pair comparisons never exceeded 037%
for signal freedom (in terms of actual test point cover), 299% for non-dominance, and
925% for overlap.
5.5 Conclusion
For all four networks, signal freedom outperformed non-dominance, which
outperformed overlap. Although non-dominance performed well on some networks
(NetworkA3 and NetworkA4) it performed badly on others, signal freedom performed
consistently well on all four networks. If high IS performance is required, then signal
freedom is the best choice. Signal freedom is difficult to implement (algorithmically),
and it may also be difficult to convert for use with other network models. If the
implementation costs were prohibitive, then non-dominance would be a suitable
replacement, but its erratic performance could lead to network designs with high
interference.

147
Overlap performed badly on all networks, and it must be concluded that overlap is a
poor IS (under the conditions applied in this test). However, the IS requirement used in
this thesis is based on the particular formulation of the evaluation function, whereas a
different evaluation function (even if only slightly different) may lead to a very different
IS condition, to which overlap may be better suited. A fundamental problem with
overlap is it inability to take into account strong server signals, and the problem of
interfering signals is only relative to the strength of the serving signals.
It might be argued that overlap also functions as an indicator of a network's ability to
handover subscribers between cells. However, in practice it is better to have separate
metrics for each network performance characteristic, as this allows the network designer
to control the relative significance of each characteristic independently.

148
Chapter 6

Network Design Part 2
The application of combinatorial optimisation to the Wireless Capacitated Network
Design Problem (WCNDP) is described in this chapter. In Chapter 3 the Base Station
Configuration Problem (BSCP) and the subsequent Channel Assignment Problem
(CAP) were solved as one problem (the WCNDP). In this chapter (following the
conclusion of Chapter 3) these two sub-problems are solved sequentially, and
comparisons drawn (this corresponds to Task T1 of Objective O2 in Section 1.3). The
evaluation function used throughout this chapter is the same as the one described in
Chapter 5.
The structure of this chapter is as follows: Section 6.1 evaluates the advantages of the
two-stage network design approach, Section 6.2 describes network designs that can be
produced with a high computational investment, and Section 6.3 assesses the impact
that minimising network infrastructure has on network performance.
6.1 Evaluation of Two-Stage Network Design Process
In Chapter 3, optimisation algorithms were described for the network design problem,
and it was concluded (see Section 3.4) that the problem should be split into two stages
(BSCP and CAP). Signal freedom is the interference surrogate (see Chapter 5) used
during the BSCP phase of optimisation described in this chapter.

149
As in Chapter 3, the total runtime allocated for each network design was limited to 20
hours. This ensured a fair comparison of results. It was not obvious how best to divide
the time between the BSCP and the CAP; different divisions were tried on each
problem, but no discernible pattern emerged, therefore both stages were allocated ten
hours.
6.1.1 Stage One: The Base Station Configuration Problem
The algorithms used for network design are taken from Chapter 3, and the (temporarily
redundant) channel assignment operations are ignored. Algorithms SearchRandomC and
SearchHillClimbA (seeded by InitRandomC) were chosen for testing as they both
performed well and there was no significant difference between their average
performances (see Section 3.3.6).
Both algorithms were tested on each network for ten hours, and the performance of the
best network designs produced is summarised in Table 3.5.
Network Performance
Metric
Search
Algorithm A1 A2 A3 A4 B1
SearchRandomC 9730 9753 9715 9852 908
Traffic
Service
(% of total)
SearchHillClimbA 9972 9491 9637 9878 1657
SearchRandomC 10000 10000 10000 10000 4412
Test Point
Service
(% of total)
SearchHillClimbA 10000 10000 10000 10000 5039
SearchRandomC 0999161 0999461 0999652 0999320 0995749
Signal
Freedom
SearchHillClimbA 0999322 0999476 0999716 0999384 0995500
SearchRandomC 10000 10000 10000 10000 10000
Occupied
Sites
(% of total)
SearchHillClimbA 10000 10000 10000 10000 10000
SearchRandomC 10000 10000 10000 10000 10000
Active
Base Stations
(% of total)
SearchHillClimbA 9960 10000 9965 10000 9853
SearchRandomC 821 246 613 492 909
Number of
Design
Evaluations
SearchHillClimbA 2294 663 1769 1325 3823
Table 6.1: The performance of the best BSCP solutions for SearchRandomC and SearchHillClimbA.
Figure 3.27 shows the performance of the best network design plotted against the time
taken to produce it for SearchRandomC on NetworkA4.

150
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 5000 10000 15000 20000 25000 30000 35000
Time Elapsed (seconds)
0.9990
0.9991
0.9992
0.9993
0.9994
0.9995
0.9996
0.9997
0.9998
0.9999
1.0000
Final: Traffic Service=98.52% Signal Freedom=0.999320
Traffic Service
Signal Freedom

Figure 6.1: The performance of the best design produced vs. time for SearchRandomC on NetworkA4.
Figure 6.2 shows the performance of the best network design plotted against the time
taken to produce it for SearchHillClimbA NetworkA4.
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 5000 10000 15000 20000 25000 30000 35000
Time Elapsed (seconds)
0.9990
0.9991
0.9992
0.9993
0.9994
0.9995
0.9996
0.9997
0.9998
0.9999
1.0000
Final: Traffic Service=98.78% Signal Freedom=0.999384
Traffic Service
Signal Freedom

Figure 6.2: The performance of the best design produced vs. time for SearchHillClimbA on NetworkA4.
6.1.2 Stage Two: The Channel Assignment Problem
Algorithm SearchHillClimbB was used, as it produced the best performance in a fixed
time (see Section 4.7). The performance of the final network designs produced is
summarised in Table 6.2.

151
Network Performance
Metric
Search Algorithm
Used for BSCP A1 A2 A3 A4 B1
Average
SearchRandomC 6576 4996 4896 7019 279 4753 Traffic Cover
(% of total) SearchHillClimbA 7651 5050 5411 7318 251 5136
SearchRandomC 6811 6010 5434 7686 1774 5553 Test Point Cover
(% of total) SearchHillClimbA 7788 6158 6131 7772 1363 5842
Table 6.2: The performance of the best network design (after channel assignment) for SearchRandomB.
6.1.3 Comparison With One-Stage Approach
The average traffic cover due to the two-stage design process was higher than that due
to the one-stage design process (see Section 3.3.6). The best average traffic cover
produced by the two-stage process was 5136%, and the best average traffic cover
achieved using the one-stage approach was 4385%, but to quantify the difference it is
best to compare relative traffic loss
1
:
Best traffic loss due to one-stage process = 100 4385 = 5615%
Best traffic loss due to two-stage process= 100 5136 = 4864%
Improvement in average traffic loss = 5615 4864 = 751%
Relative improvement in average traffic loss = 751 / 5615 = 1337%
This is a significant improvement in searching efficiency, and it was achieved with a
relatively small programming effort (i.e. the implementation of the interference
surrogate, the modification of the search procedures, and the modification of the
evaluation function).
Of the two algorithms tested in this chapter, SearchHillClimbA produced the highest
average performance for all network performance metrics, and this is due, in part, to the
ability of SearchHillClimbA (or any local-search algorithm) to exploit coherence
between consecutively evaluated network designs. The efficiency of caching is

1
This is because traffic loss becomes increasingly difficult to minimise the closer it gets to 0%, i.e.
optimisation effort gives a diminishing return.

152
demonstrated by the fact that approximately three times as many network evaluations
were made by SearchHillClimbA than were made by SearchRandomC. If the caching
of evaluation calculations were disabled it is certain that the performance of
SearchHillClimbA would decrease, and it is possible that it would fall below that of
SearchRandomC. Therefore, it should not be assumed that SearchHillClimbA would
always outperform SearchRandomC.
6.2 Network Designs with High Computational Investment
All the network designs produced previously in this thesis have been limited to 20 hours
of runtime, but in this section the runtime is extended to seven days: five days for the
BSCP and two days for the CAP. This relative division of time between network design
and channel assignment was found to yield good network design performance, however,
no optimal division of time between the two stages has been determined.
An unlimited time would ideally be devoted to finding the best possible network, but
obviously this is not possible in practice. Therefore it was decided that 35 days (7 days
5 networks) was the maximum time available for the generation of results for this
section. The two-stage network design process was adopted because of its superior
searching efficiency (see Section 6.1.3).
6.2.1 Algorithms for the BSCP
A combination of search algorithms was used to produce a solution to the BSCP for
each of the five networks. Although SearchHillClimbA was found to produce the best
final network design performance, it was observed that SearchRandomC produced the
best improvement in network design in the early stages of optimisation (for example,
compare Figure 6.1 with Figure 6.2), but then stalled after about an hour. Therefore the
first hour of searching was done by SearchRandomC, after which the control was passed
to SearchHillClimbA.
Experience from previous runs of SearchHillClimbA showed that approximately 50% of
moves in which base stations were deactivated produced an improvement in design

153
performance, and that the adjustment of parameters for these redundant base stations
(and their inclusion in the evaluation procedure) wasted time. Once the initial
optimisation algorithms (SearchRandomC and SearchHillClimbA) had achieved 99%
traffic and test point service, a new searching phase was started in which redundant base
stations were pruned from the network design. This process is embodied in the
algorithm SearchPruneA, which is a modified version of SearchHillClimbA where only
base station deactivation moves are permitted (see Figure 6.3 for details). SearchPruneA
was run to termination (local optimality), and then control was returned to
SearchHillClimbA for the remaining time.

Algorithm SearchPruneA
Input: network design ND
INIT
Output: network design ND
BEST
Note: may be pre-emptively terminated

Network design N
BEST = N
INIT
Repeat until no change is made to N
BEST
Create empty move list, moves
For every base station, b
Add move b active FALSE to moves
EndFor
Suffle moves
For every move, m, in moves
N
NEW = N
BEST + m
If p(N
NEW) > p(N
BEST)
N
BEST = N
NEW
EndIf
EndFor
EndRepeat
Return NBEST

Figure 6.3: Algorithm SearchPruneA.
6.2.2 Algorithms for the CAP
SearchHillClimbB is the only channel assignment algorithm (of those tested in Chapter
4) guaranteed to terminate in a fixed time, however, it was found that InitConstructA
terminates in a shorter (albeit still unpredictable) time on better cell plans than the
(random) ones used for testing in Chapter 4. Therefore due to InitConstructAs
superiority over SearchHillClimbB in terms of searching efficiency (see Section 4.7) it

154
was chosen for the initial assignment, after which the control passed to
SearchHillClimbB for the remainder of the two-day channel assignment period.
6.2.3 Results
The results are presented separately for the two optimisation stages: solution of the
BSCP and solution of the subsequent CAP.
6.2.3.1 BSCP
The performance of the best network designs produced during the five day period is
summarised in Table 6.3.
Network
Performance Metric
A1 A2 A3 A4 B1
Traffic Service
(% of total)
10000 10000 10000 10000 6880
Test Point Service
(% of total)
10000 10000 10000 10000 3853
Signal Freedom 0999628 0999707 0999863 0999694 0994096
Occupied Sites
(% of total)
8040 8969 5757 9713 10000
Active Base Stations
(% of total)
5187 5427 3175 6530 9341
Number of Design
Evaluations
40246 11095 44947 20286 135220
Table 6.3: The performance of the best BSCP solution for composite search.
Figure 6.4 shows the performance of the best network design plotted against the time
taken to produce it for NetworkA2, and the three optimisation phases are labelled as
follows: A denotes SearchRandomC followed by SearchHillClimbA, B denotes
SearchPruneA, and C denotes SearchHillClimbA.

155
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 50000 100000 150000 200000 250000 300000 350000 400000
Time Elapsed (seconds)
0.9994
0.9995
0.9996
0.9997
0.9998
0.9999
1.0000
Final: Traffic Service=100.00% Active Base Stations=54.27% Sites Occupied=89.69% Signal Freedom=0.999707
Traffic Service
Active Base
Stations
Sites Occupied
Signal Freedom
A B C

Figure 6.4: The performance of the best BSCP solution vs. time for composite search on NetworkA2.
6.2.3.2 CAP
The best channel assignment found for each BSCP solution was used to generate the
network performance metrics in Table 6.4.
Network
Performance Metric
A1 A2 A3 A4 B1
Traffic Cover (% of total) 10000 9567 9779 9977 1090
Test Point Cover (% of total) 10000 9676 9913 9982 2578
Table 6.4: The performance of the best network design for seven day optimisation.
6.2.4 Conclusion
Table 6.4 shows that, excluding NetworkB1, the performance of the network designs
produced during the seven day optimisation was high no more than 5% of call traffic
was lost from any network and no more than 4% of the test point area went uncovered.
The predicted performance of NetworkB1 suffered because of the characteristics of the
data (see Section 2.4.5).

156
A large amount of network infrastructure was required to achieve the high network
performance. In Table 6.5 the amount of infrastructure used is compared with the
theoretical minimum infrastructure dictated by capacity limitations (as calculated in
Section 2.4).
Network
Performance Metric
A1 A2 A3 A4
Actual 8040 8969 5757 9713 Occupied Sites
(% of total) Minimum 1000 688 423 861
Actual 5187 5427 3175 6530 Active Base Stations
(% of total) Minimum 1000 677 411 847
Table 6.5: A comparison of actual network infrastructure with theoretic minimum.
For all four networks taken together, approximately seven (20319 2935 = 692) times
as many base stations were used than the minimum required to serve all the network
traffic.
6.3 Network Designs with Minimal Infrastructure
In the previous section it was demonstrated that network designs with high performance
could be produced using automatic optimisation, however, the designs required
substantial network infrastructure. In this section, two approaches are presented that
explicitly address this infrastructure problem. The first approach attempts to minimise
the infrastructure in an existing design, and the second attempts to create a minimal
design from scratch. The number of base stations is explicitly targeted as the element
of infrastructure to be reduced, but the number of sites will implicitly be reduced by this
process.
NetworkB1 was excluded from this infrastructure minimisation study as it contains
more traffic than would be coverable using even a full complement of base stations (see
Section 2.4.5).

157
6.3.1 Algorithm: SearchPruneB
The goal of algorithm SearchPruneB, detailed in Figure 6.5, is the production of
network designs with minimal infrastructure. It is an extended version of
SearchPruneA.
SearchPruneB delivers a network design with a performance equal to (or greater than)
that of the initial network design used, and number of base stations in the new design
will be less than or equal to that of the initial design.
6.3.1.1 Description
A list of active base stations is made from the initial network design. The base stations
are then arranged in ascending order of the amount of traffic cover they provide, as the
removal of those covering the least traffic is likely to have the least impact on network
performance. The first base station in the list is deactivated and the resulting network
design evaluated. If network performance does not deteriorate (i.e. no loss of traffic or
test point cover), then this base station remains inactive, otherwise it is reactivated and
the next base station in the list is tested instead. If no superfluous base stations can be
found, the base station whose removal had the least impact (i.e. led to the least loss in
traffic cover) is deactivated, and the resulting network is then hill climbed until the
target network performance (that of the initial design) is restored. Whenever this target
performance is achieved (after the removal of a base station or after restoration by hill
climbing) the whole sequence is repeated on the new network design.

158

Algorithm SearchPruneB
Input: network design ND
INIT
Output: network design ND
PRUNE
Note: may be pre-emptively terminated

Declare network design NDPRUNE = NDINIT
Calculate total traffic in network t
TOT
Declare t
BSMAX = maximum traffic per base station
Calculate min number of base stations n
BSMIN = t
TOT / t
BSMAX(
Declare network design ND
TRIAL = ND
INIT
Repeat until number of active base stations in NDTRIAL = nBSMIN
Create list B of active base stations in ND
TRIAL
Sort B by increasing traffic cover
Declare base station b = first member of B
Remove b from B
Declare network design ND
FALLBACK = ND
TRIAL
Deactivate b in ND
FALLBACK
Repeat while p(ND
INIT) > p(ND
FALLBACK) and B not empty
Declare base station b = first member of B
Remove b from B
Deactivate b in ND
TRIAL
If p(ND
TRIAL) > p(ND
FALLBACK)
ND
FALLBACK = ND
TRIAL
EndIf
Activate b in ND
TRIAL
EndRepeat
ND
TRIAL = ND
FALLBACK
If p(ND
INIT) > p(ND
TRIAL)
ND
TRIAL = SearchHillClimbC(ND
TRIAL, p(ND
INIT))
EndIf
ND
PRUNE = ND
TRIAL
EndRepeat
Return ND
PRUNE

Figure 6.5: Algorithm SearchPruneB.
The hill climb procedure used by SearchPruneB to restore coverage is embodied in
algorithm SearchHillClimbC (detailed in Figure 6.6). This modified version of
SearchHillClimbA has no activation/deactivation moves (as these are explicitly handled
by SearchPruneB) and terminates once the performance of its current design is greater
than or equal to a given target performance.

159

Algorithm SearchHillClimbC
Input: network design NDINIT and target performance PTARGET
Output: network design NDBEST
Note: may be pre-emptively terminated

Declare network design NDBEST = NDINIT
Create empty move list moves
For every base station b in NDBEST
For every possible azimuth az
Add move b azimuth az to moves
EndFor
For every possible tilt tl
Add move b tilt tl to moves
EndFor
For every possible transmit power tx
Add move b txPower tx to moves
EndFor
For every possible antenna type ad
Add move b antDef ad to moves
EndFor
For every link k on b
For every possible channel c
Add move b channel k c to moves
EndFor
EndFor
EndFor
Repeat while PTARGET > p(NDBEST)
Shuffle moves
For every move m in moves
Declare network design NDTRIAL = NDBEST + m
If p(NDTRIAL) > p(NDBEST)
NDBEST = NDTRIAL
EndIf
EndFor
EndRepeat
Return NDBEST

Figure 6.6: Algorithm SearchHillClimbC.
Note that the full evaluation of CIR is necessary to compute the test point and traffic
covers for each iteration of these algorithms. Without this calculation the design
performance could not be guaranteed.
6.3.1.2 Results
The network designs produced in Section 6.2 were used as the input to SearchPruneB,
because these network designs had high performance but required large amounts of
infrastructure. SearchPruneB was run for seven days on each of the following networks:

160
NetworkA1, NetworkA2, NetworkA3, and NetworkA4. The results are summarised in
Table 6.6.
Network
Performance Metric
A1 A2 A3 A4
Traffic Cover (% of total) 10000 9568 9779 9977
Test Point Cover (% of total) 10000 9543 9577 9970
Occupied Sites (% of total) 6520 6594 3415 8484
Active Base Stations (% of total) 3733 3146 1631 4809
Number of Network Evaluations 17680 7744 33157 12834
Table 6.6: The performance of the final network designs produced by SearchPruneB.
Figure 6.7 shows the network infrastructure employed in the design for NetworkA1
plotted against the time taken to achieve it.
0%
10%
20%
30%
40%
50%
60%
70%
80%
90%
100%
0 100000 200000 300000 400000 500000 600000
Time Elapsed (seconds)
Final: Active Base Stations=37.33% Sites Occupied=65.20%
Active Base Stations
Sites Occupied

Figure 6.7: Graph of network infrastructure vs. time for algorithm SearchPruneB on NetworkA1.
Algorithm SearchPruneB decreased the amount of infrastructure required for each of
the four network designs whilst maintaining the original levels of performance. The new
design for NetworkA1 required 2803% less base stations than the original design,
4203% in NetworkA2, 4863% in NetworkA3, and 2636% in NetworkA4 (an average
decrease of 3626%). These results are presented in Table 6.7.

161
Network
Performance Metric
A1 A2 A3 A4
Before 8040 8969 5757 9713
After 6520 6594 3415 8484
Occupied Sites
(% of total)
Before - After 1520 2375 2342 1229
Occupied Sites %Difference 1891 2648 4068 1265
Before 5187 5427 3175 6530
After 3733 3146 1631 4809
Active Base
Stations
(% of total) Before After 1454 2281 1544 1721
Active Base Stations %Difference 2803 4203 4863 2636
Table 6.7: The effect on network infrastructure of algorithm SearchPruneB.
As well as requiring fewer base stations, fewer sites were required for each network.
The algorithm specifically targets base stations for removal, but a modified algorithm
that targeted site reduction rather than base station reduction might improve upon the
amount of site reduction (albeit at the expense of base station reductions).
Although the amount of infrastructure in the new network designs has fallen, it is still
significantly higher than the theoretical minimum dictated by capacity limitations (see
Section 2.4). For all four networks together, approximately four and a half (13319
2935 = 454) times as many base stations were used than is required to satisfy base
station capacity limitations.
6.3.2 Algorithm: SearchConstructC
In the previous section, the approach to infrastructure minimisation was the
modification of an existing, high performance, network design. An alternative approach
is the production of networks that employ minimal (or near minimal) infrastructure at
the (possible) cost of network coverage performance. Comparison of the previous
performance of the search algorithms (as summarised in Section 3.3.6) showed that
SearchConstructB produced network designs with a good ratio of coverage performance
to deployed base station infrastructure, therefore it was chosen as the core algorithm
for investigation.
For a fair comparison with the approach in the previous section, the optimisation of
each network ran for 14 days (the total time invested in creating each network design).

162
Algorithm SearchConstructB was found to be unsuitable for long optimisation runs and
was replaced with SearchConstructC, as detailed in Figure 6.8.
6.3.2.1 Description
Algorithm SearchConstructC is a modified version SearchConstructB, which reversed
the goal of maximising test point service in order to limit interference. In
SearchConstructC the construction of network designs occurs in two phases. In the first
phase, new base stations are selected based on high traffic service and low test point
service, and this phase ends once 99% of the traffic in the network is serviced. In the
second phase the evaluation function is restored and new base stations are selected
based on high test point service this hole filling stage ends once 99% test point
service is achieved.

Algorithm SearchConstructC
Input: none
Output: network design NDBEST
Note: may be pre-emptively terminated

Declare network design ND
BEST = InitBlank()
Reverse signifiance of test point service in evaluation function
Repeat until ND
BEST gives 99% traffic and test point service
If NDBEST gives 99% traffic service
Restore evaluation function to normal
Endif
Declare network design ND
NEW = ND
BEST
Declare network design ND
TRIAL = ND
BEST
For every inactive base station b in N
TRIAL
Activate b
Choose random azimuth for b
Choose random tilt for b
Choose random transmit power for b
Choose random directional antenna for b
If p(ND
TRIAL) > p(ND
NEW)

ND
NEW = ND
TRIAL
EndIf
Deactivate b
EndFor
If Peformance(ND
NEW) > Peformance(ND
BEST)
ND
BEST = ND
NEW
EndIf
EndRepeat
Return ND
BEST

Figure 6.8: Algorithm SearchConstructC.

163
Algorithm SearchConstructC was run to termination, and the resulting network design
given to algorithm SearchPruneB. If SearchPruneB terminated within the allotted time,
control was passed to SearchHillClimbC. This optimisation phase was terminated after
10 days regardless of which algorithm was currently active. A channel assignment was
created for the resulting network design using InitConstructA. If this did not require the
allotted four days, then the assignment was optimised by SearchHillClimbB for the
remaining time.
6.3.2.2 Results
Table 6.8 summarises the performance of the best network designs produced during the
full 14 day period. The results of the two optimisation phases (the BSCP solution and
the CAP solution) are combined in Table 6.8.
Network
Performance Metric
A1 A2 A3 A4
Traffic Cover (% of total) 8774 6021 6508 6235
Traffic Service (% of total) 9902 9907 9946 9947
Test Point Cover (% of total) 8101 6020 6566 6412
Test Point Service (% of total) 10000 9994 10000 10000
Signal Freedom 0998364 0998645 0999264 0998216
Occupied Sites (% of total) 2720 2031 1144 2295
Active Base Stations (% of total) 1000 677 411 847
No. Network Designs (BSCP) 309575 127796 366198 212731
Table 6.8: The performance of the best network designs produced by the composite constructive search
algorithm.
For all four networks, the minimum number of base stations (as dictated by capacity
limitations) was achieved whilst still servicing >99% of the network traffic. However,
the interference in these designs was much higher than in those found in Section 6.3.1.2
using the same amount of computational investment. This leads to a significant
reduction in test point and traffic cover in all four networks.

164
Chapter 7

Conclusions and Future Work
The structure of this chapter is as follows: Section 7.1 contains the thesis conclusions,
Section 7.2 recommends enhancements to the network model, and Section 7.3 presents
possible improvements to the optimisation methodology.
7.1 Thesis Conclusions
In this section, the thesis results are examined and conclusions are drawn about the
applicability of optimisation to the problem of mobile telephone network design.
7.1.1 Spectral Efficiency and Network Infrastructure
Although this thesis was primarily a study of ACP techniques, an important principle to
emerge from the results was a trade off between spectral efficiency and the quantity of
network infrastructure. It is useful to discuss this relationship first so that the remaining
conclusions, more directly relevant to ACP, may be better placed in context. The
underlying principal will be referred to as the efficiency conjecture, and it may be stated
as follows: in general, spectral efficiency in a capacitated network may be
improved by the addition of more network infrastructure.

165
In a wireless network, spectral re-use is the amount of information transferred
(measured in bits) divided by the amount of spectrum used (measured in Hertz) to
transfer it
1
. There is a fundamental limit to the amount of information a radio channel
(or concatenation of channels) can convey to the same place at the same time [5], but
the same channels may be used multiple times in different places if the information
receivers are sufficiently electromagnetically isolated from one another. Hence, there is
no theoretical upper-limit to spectral re-use. However, in a real wireless network the
receivers may not be sufficiently spread-out to ensure EM isolation, and the spectral
efficiency per unit area is a more accurate indicator of network performance
2
. This is
the measure referred to by the efficiency conjecture.
Now, the information bandwidth (or data rate) available to a receiver is dictated by the
CIR of its serving signal [55], and a fixed data rate is required for telephony (in a GSM
system at least). Therefore, a corresponding (fixed) target CIR may be derived (as was
the case for the data sets used in this thesis) that ensures a successful telephone
connection. For the thesis results, only receivers for which a sufficient CIR was
available were considered covered (see Section 2.2.16), and it was found that: the
addition of more base stations often improved the CIR in the network, which in turn
improved the number of receivers covered. This is the basis for the efficiency
conjecture, and it is this relationship that the following subsection attempts to establish.
It is important to note that the improvement in spectral efficiency predicted by the
efficiency conjecture is not due simply to the addition of traffic carrying transceivers;
instead it refers to the reduction of interference in a network that already has sufficient
transceivers to serve all the subscribers.
There are limits on the range of situations for which the efficiency conjecture is valid,
and there exists a point at which the addition of infrastructure will give no appreciable

1
The units being bits/Hertz.
2
The units being bits/Hertz/m
2
.

166
improvement in spectral efficiency. One limiting factor is the background noise even
if the interference due to other network elements where to disappear completely, the
CIR would be limited by unavoidable background noise.
7.1.1.1 Evidence
Figure 3.10 and Figure 3.19 show a positive correlation between test point cover and
base station activity for random network designs, but in the regions of the graph where
base station activity is low, a wide range of test point cover is seen. However,
comparison of designs in this region are (for the purposes of establishing a correlation)
invalid, because the subscriber capacity each design offers will vary widely. It is only
fair to compare network designs that have the same, or very similar, capacity
performance, i.e. the same traffic service. This is because interference, transceiver
allocations, and traffic are intimately linked.
To eliminate the effect (described in the previous example) that transceiver-based
capacity limitations have on the correlation graphs, it is necessary to compare network
designs in which the base station transceiver capacities have been matched to subscriber
demand (i.e. designs with the same traffic service). By chance, the network designs
produced for testing interference surrogates in Chapter 5 are ideally suited. Ten network
designs were produced (for each network data set), in which the number of active base
stations was widely varied (see Section 5.3) but the traffic service was the same. These
initial network designs were optimised until their traffic and test point service
performances both exceeded 99%. The correlations between base station activity and
test point cover are shown in Figure 7.1, Figure 7.2, Figure 7.3, and Figure 7.4. Clearly
a correlation does exist, but the stochastic nature of the process used to derive the
network designs means that some variance is present in the results, therefore the
correlation is not conclusive.

167
40%
45%
50%
55%
60%
65%
70%
75%
80%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)

Figure 7.1: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA1.
40%
45%
50%
55%
60%
65%
70%
75%
80%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)

Figure 7.2: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA2.

168
40%
45%
50%
55%
60%
65%
70%
75%
80%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)

Figure 7.3: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA3.
40%
45%
50%
55%
60%
65%
70%
75%
80%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Active Base Stations (% of total)
T
e
s
t

P
o
i
n
t

C
o
v
e
r

(
%

o
f

t
o
t
a
l
)

Figure 7.4: Graph of test point cover vs. base station activity for ten candidate designs in NetworkA4.
More sensitive and intelligent experimentation is required in order to prove or disprove
the efficiency conjecture, and is not considered in this thesis. The evidence presented
here, at best, implies the validity of the efficiency conjecture (or a closely related
mechanism). It is important to note that the thesis results are conveniently placed in
context by the efficiency conjecture, but they are not contingent upon it being true.

169
7.1.1.2 Explanation
It is useful to describe a scenario where the efficiency conjecture might be true. The
demonstration of a pathological case for which the conjecture is valid is by no means a
proof, but it does help elucidate the potential underlying mechanism. Consider a one-
dimensional array of evenly distributed test points indicated by the horizontal axes of
the two graphs in Figure 7.5 and Figure 7.6. Each test point requires an arbitrary CIR of
10 in order to be covered (shown as a bold horizontal line). The first graph shows the
effect of two transmitters TX A (shown in red) and TX B (shown in green). The
following values are plotted for all the receivers: the field strength of the signal from
each transmitter, the total noise (which corresponds to a constant noise floor plus the co-
channel interference from all other transmitters) that a receiver would experience if
served by the given transmitter, and (similarly) the CIR. It can be seen that 32% of test
points will be covered by one or other transmitter, i.e. 32% of test points have access to
a signal with a CIR greater than or equal to 10. The second graph shows the same
scenario with the inclusion of a third transmitter, TX C (shown in blue). The extra
interference generated by TX C has caused the CIRs of receivers served by TX A or TX
B to fall marginally, however this is more than compensated for by the additional
receiver cover in the immediate vicinity of TX C the total cover rising to 38%.
A simple propagation loss model was used to compute the field strengths in both cases,
the assumption being that signal strength was attenuated in proportion to

d where d is
the distance (in metres) and is 25. A background noise of 01W was assumed and
the transmitters were assumed to all be transmitting on the same channel at a power of
1W. The line of receivers was assumed to be 3 km from end to end. The figures were
chosen to be consistent with the data sets used in this thesis. Different figures could lead
to the opposite conclusion (i.e. that the introduction of TX C actually reduced receiver
cover), but the aim of this example was the demonstration of a possible scenario where
the efficiency conjecture might be true.

170
1.E-09
1.E-08
1.E-07
1.E-06
1.E-05
1.E-04
1.E-03
1.E-02
1.E-01
1.E+00
1.E+01
1.E+02
1.E+03
1.E+04
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Field strength due to A (W) Field strength due to B (W)
Total noise if served by A (W) Total noise if served by B (W)
CIR if served by A CIR if served by B

Figure 7.5: A pathological example of receiver cover for two transmitters.
1.E-09
1.E-08
1.E-07
1.E-06
1.E-05
1.E-04
1.E-03
1.E-02
1.E-01
1.E+00
1.E+01
1.E+02
1.E+03
1.E+04
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Field strength due to A (W) Field strength due to B (W) Field strength due to C (W)
Total noise if served by A (W) Total noise if served by B (W) Total noise if served by C (W)
CIR if served by A CIR if served by B CIR if served by C

Figure 7.6: A pathological example of receiver cover for three transmitters.

171
7.1.1.3 Implication
A base station has an upper traffic capacity limit (for a given grade of service), therefore
a minimum number of base stations for a network may be calculated from the total
network traffic (as shown in Section 2.4). The efficiency conjecture implies that
additional base stations may be required to meet CIR coverage targets. If the number of
base stations falls below this, CIR related, minimum then the number of subscribers
successfully covered must also fall. This trade off between infrastructure and subscriber
cover is illustrated by the pareto-optimal
1
curve shown in Figure 7.7. The shaded region
beyond this pareto-optimal front represents unrealisable network designs.

0%
100%
C
o
v
e
r

No. Base Stations
min no. BS for 100% cover
Theoretical Pareto Optima (PO)

Figure 7.7: Graph of subscriber cover vs. network infrastructure.
The curve shown in Figure 7.7 is theoretical. In practice there may be no method by
which these pareto-optimal designs could be created, or such a method may require an
unavailable quantity of computer power to achieve. Figure 7.8 shows the pareto-optimal
curves that may be obtained with finite computing resources, where one Computing
Unit (CU) is used to represent the minimum amount of effort required to create and

1
Limits may be quoted in terms of a single parameter (such as the speed of light), or in terms of a pareto
optimal trade off between competing parameters (such as speed and efficiency in a petrol engine). You
may optimise one parameter, but only at the expense of the others.

172
evaluate a single network design. As the computational investment increases, the
pareto-optimal curves approach (but do not necessarily meet) the theoretical maxima.

0%
100%
C
o
v
e
r

No. Base Stations
Theoretical Pareto Optima (PO)
PO for
1 CU
PO for
10 CU
PO for
100 CU
PO for 1000 CU

Figure 7.8: Graph of the effect of computational investment on the pareto-optimal trade off between
subscriber cover and network infrastructure.
In light of the efficiency conjecture, a network designer must decide what is more
important: a high level of cover or a minimal deployment of infrastructure? Two
designs that epitomise these aspirations are represented in Figure 7.9, where the target
performances are respectively labelled A and B. However, it is not possible to derive
these pareto-optimal curves for a given network, or to be sure that the design
performances shown are realisable. The graphs are only useful as an illustration of the
efficiency conjecture.

173

0%
100%
C
o
v
e
r

No. Base Stations
min no. BS for
TRX capacity
min no. BS for 100% cover
max
cover for
min no.
BS
A
B

Figure 7.9: Target performance positions on the pareto-optimal graph.
7.1.2 Effectiveness of Automatic Cell Planning
In the following subsections, methods of generating network designs with the
performances approaching those shown in Figure 7.9 are evaluated (this corresponds to
Task T3 of Objective O2 in Section 1.3).
7.1.2.1 Synthesising Network Designs with High Cover
In Section 6.2 designs with near maximal cover were produced for all networks
1
. Seven
computer days were required, and the network designs contained a substantial amount
of infrastructure. None of the network designs produced matched the target theoretical
optima (labelled A in Figure 7.9). However, the subscriber cover achieved using these
designs was high, and the computational investment required to obtain them was
reasonable. It may be concluded that ACP is capable of producing network designs with
near maximal cover in a reasonable time, but the amount of network infrastructure
required may be high.

1
Except for NetworkB1, the reasons for which were explained in that section.

174
7.1.2.2 Synthesising Networks with Minimal Infrastructure
The algorithms in Section 6.3 focused explicitly on the minimisation of network
infrastructure, and two distinct approaches were formulated. The first approach to
infrastructure minimisation (described in Section 6.3.1) relied on the modification of an
existing network design by the removal of redundant infrastructure. The average
reduction in the number of base stations was over 30% using seven computer days per
network. This approach is an extension of the previous design method (Section 7.1.2.1)
in so far as the target theoretical performance is that represented by A in Figure 7.9.
The second approach to infrastructure minimisation (described in Section 6.3.2)
produced network designs with a minimal number of base stations (as dictated by base
station capacity and network traffic) at the potential expense of subscriber cover.
Minimal infrastructure was achieved for all networks, but the average loss in cover was
approximately 30%. This approach implicitly targets point B in Figure 7.9.
It may be concluded that ACP is capable of significantly reducing network
infrastructure requirements, at least for synthetic network designs. Infrastructure
minimisation is most effective if the target subscriber cover does not need to be
guaranteed, however significant minimisation can be achieved regardless.
7.1.3 Separating the BSCP and the CAP
An important and novel feature of this thesis was the inclusion of channel assignment as
part of the whole wireless network design process, and a comparison was made with the
traditional two-stage approach. The network designs produced by the two-stage
approach had approximately 13% less traffic loss, and hence were superior. Note that
the efficiency of the two-stage approach is only available if a suitable interference
surrogate is used, as derived in this thesis.
Despite the superiority of the two-stage approach, the one-stage approach does have
certain advantages. Infrastructure minimisation algorithms, such as SearchPruneB, only
allow the removal of base stations if they do not effect subscriber cover, and this can
only be guaranteed if a full channel model is used to assess interference.

175
7.1.3.1 Signal Freedom
In Chapter 5 the interference surrogate Signal Freedom was developed. It was shown to
accurately predict relative levels of interference in candidate network designs. This
allowed proper consideration of the difficulty in assigning channels to any base stations
selected during the network design phase. Signal freedom was shown to significantly
outperform two current industrial interference surrogates: overlap and non-dominance.
7.1.4 Comparison of Manual and Automatic Cell Planning
There are many facets of cell planning that cannot be easily quantified and, hence,
automated. For instance, covering an area where a technology trade show is regularly
held (e.g. Earls Court in London), which is politically significant but not directly
economical.
ACP is intended to empower cell planners, not to replace them. ACP still requires the
cell planner to specify what they want (in terms of service provision and economics),
but not how to achieve it. This frees (valuable) cell planner time and allows them to
either tackle larger service areas or to concentrate in more detail on the issues arising in
the same service area (such as data transmission quality or network backbone
bottlenecks).
7.1.5 Network Economics
Consider a network with a fixed target subscriber cover (selected by the network
operator). If ACP techniques are used, the number of base stations required to meet the
target cover will decrease as the amount of computer power (number of computers
time) increases. This relationship is shown in Figure 7.10.

176

Computer Power (CP)
No. Base
Stations
Minmimum
no. BS for
required
cover

Figure 7.10: Number of base stations vs. computational investment for a fixed subscriber cover.
Now, the amount of money spent on network infrastructure will be proportional to the
number of base stations used as shown in Figure 7.11.

No. Base Stations
Cost

Figure 7.11: Infrastructure cost vs. number of base stations.
The operator would like to determine the optimal amount of computer power to devote
to ACP. This is the point where any additional investment in computers would cost

177
more than it would save by reducing infrastructure. This optimal point is illustrated in
Figure 7.12.

Computer
Power (CP)
Cost
Optimal CP (where total
cost is minimal)
Total cost
Computational
cost
BS cost

Figure 7.12: The determination of optimal computational investment by comparison of computational
cost and network infrastructure cost.
No method has been derived to allow an operator to determine this optimal
computational investment, at best it can be said that operators may currently be
investing too little. The illustration of computational economics (given above) is
simplistic, but it is included to try and place the thesis results in a commercial context.
If subscriber cover (and hence, revenue) is allowed to vary, the economic model
becomes more powerful. Further insights are gained in allowing the amount of required
radio spectrum to vary, as the efficiency conjecture suggests a trade-off between
required spectrum and required infrastructure. For example, an operator may save
money by renting less spectrum (an incredibly expensive commodity) and deploying
more base stations (cheap in comparison to spectrum). However, the validity of these
ideas is not considered in this thesis.

178
7.2 Network Model Enhancements
This section presents extensions to the network model. All these proposals are intended
to improve the quality of the predictions made by the model.
7.2.1 Handover
The mechanism of subscriber handover from one cell to another has not been explicitly
considered in this thesis. True mobility relies on the support of handover between cells,
and this becomes more important as cell sizes decrease, which is an essential feature of
high capacity networks. The model presented in this thesis is static it does not
consider the movement of subscribers. Some models, such as the connectivity model
in [2], have relied on approximate measures of a networks handover performance.
However, the quality of these approximations is unknown and, in pathological cases at
least, a predicted optimal performance would not guarantee handover support in a real
network.
A study of handover successes and failures in operational networks is required in order
to derive a complete model of handover behaviour. If the computational complexity of
this model prohibits its inclusion in an optimisation procedure, a surrogate function
should be derived (in a similar way to the interference surrogate derived in this thesis)
that has a known (or at least estimated) degree of accuracy.
7.2.2 Network Backbone Design
In this thesis no consideration is made of the networks backbone infrastructure (such as
the trunking elements, PSTN entry points, and subscriber database services). These
considerations were omitted because of a lack of readily available data, but their
inclusion in the network model would greatly enhance the quality of information that
could be extracted from simulations (for example, the latency of communication
between two points in the network). The layout of the backbone infrastructure as a
separate optimisation problem has been well studied, and examples can be found in [56]
and [57].

179
7.2.3 Three-Dimensional Modelling
The thesis network model assumes that all subscriber traffic will originate on the
ground, and no consideration is made of the vertical distribution of traffic that might
occur (for instance, in a building). It has been informally estimated that in the City of
London, only 40% of mobile traffic occurs at ground level. This clearly implies that a
two-dimensional model will be less accurate than a three-dimensional one.
As subscriber densities and bandwidth requirements increase, cell sizes must decrease.
The only way to achieve efficient spectrum re-use in a built-up environment is to
sectorise the building vertically; this may be achieved by pico-cellular architectures
(each room of a building is given its own base station) or by mounting directional
antennae on adjacent buildings at different heights. To improve prediction accuracy,
consideration of the impact to users in vertically adjacent cells must be made.
The computational requirements of a full three-dimensional model are greater than that
of the model presented in this thesis, in which a trade off has been made between
prediction quality and simulation run-time. However, as the price/performance ratio of
computers continues to improve, this trade off must be periodically re-assessed.
7.2.4 Temporal Variance
The inclusion of temporal variance to the model could give an improvement in
simulation accuracy. The busy-hour traffic assumption used in the thesis model is only
able to give a worst-case prediction of network performance. Gross temporal variances
could be considered as individual static cases, for instance the traffic pattern during
working hour differs significantly from the traffic pattern outside of those hours. The
current model makes a conservative estimate of network coverage performance based
on the peak traffic expected at each test point throughout a typical day, but this leads to
an over-estimation of interference and an under-estimation of network capacity; a more
realistic prediction would allow the deployment of less infrastructure. Finer temporal
variance characteristics, such as mobility, could also be included to improve predictions
of handover support (see Section 7.2.1)

180
7.2.5 Delay Spread
Path loss is not the only characteristic of a radio signal that affects reception quality;
another important factor is delay spread. Delay spread occurs when a transmitted signal
travels by more than one path to its intended receiver. In the GSM system, adaptive
channel equalization is used to recover the intended signal, whereas in IS-95 [9] Rake
receiver techniques are used to overcome the problem of delay spread. The problem of
optimising wireless networks to meet delay spread constraints is explicitly considered in
[3].
7.2.6 Quality of Service
The central goal of the network design problem outlined in this thesis was the provision
of a signal with a sufficient Carrier-to-Interference-Ratio (CIR) for each user.
Expressing the design requirement in terms of Quality of Service (QOS) measures (such
as data throughput, latency, and blocking probability) instead of technology specific
measures, such as CIR, has two distinct advantages:
1. These measures are closer to a users experience of the network, and this makes the
relative prioritisation of these quantities easier for the network designer. In addition
to simplifying the specification of network requirements, the specification becomes
much more portable between different technologies (such as GSM and UMTS).
2. Actual QOS statistics can be directly compared with other communication systems.
A user (or a users application) will choose a transport medium (such as wireless
LAN, wireline ISDN, or GPRS) based on a comparison of the offered QOS, and the
network designer will be best placed to win that competition (and hence transport
the users data) if QOS is given explicit consideration during network design.
Mobile telephone network operators are increasingly interested in the provision of a
variety of services, each of which may have different QOS requirements. Although
telephony is set to continue as the core service, its QOS requirements (for instance, low
latency and low bandwidth) are not the same as those for other services. The explicit

181
calculation of QOS statistics allows a range of services to be mapped onto a wireless
network and not simply the one-size-fits-all mapping that results from designing solely
for telephony.
7.2.7 Error Management Through Interval Arithmetic
The accuracy of predictions made by the model is dependent on the accuracy of input
data such as terrain topography, subscriber positions, and equipment operating
characteristics. Additional inaccuracies are introduced by computational operations such
as propagation prediction. The calculation of the inaccuracy (or tolerance) of the results
is non-trivial and the usual practice is to quote a standard, conservative, estimate of the
expected tolerance for each type of simulation result (for example, 4dB for propagation
predictions).
Interval arithmetic is a computational approach that may prove useful in improving the
quality of the simulation results (see [58] for background). Instead of input variables
being quoted as single figures (as in single point arithmetic) they are quoted as two
figures (an interval), and the true value is said to be contained within the interval.
These intervals are propagated through the calculation until they emerge as outputs. The
output is guaranteed to contain the correct answer, and the spread of the interval
indicates the certainty of the result. More conventional tolerance techniques usually
require an analytical analysis of the algorithms involved, but the numerical approach of
interval arithmetic delegates this book-keeping to the computer instead of the
programmer. Several disadvantages must be addressed if interval analysis is to be
included in network simulation:
1. Propagation prediction algorithms must be modified to include the interval concept,
but this is cannot be achieved by a simple extension of the existing algorithms.
2. No optimisation techniques have been developed that are designed to deal with
interval based objective functions.

182
These disadvantages do not represent shortcomings in the interval approach to
computation but are in fact implicit problems the single point approach ignores.
7.3 Improvements to Optimisation Methodology
A number of improvements to the optimisation methodology have been postulated, but
time constraints have prevented their implementation and testing. They are presented
here as a resource for future research on this problem as they represent an opportunity
for improving optimisation effectiveness and efficiency (this corresponds to Task T4 of
Objective O2 in Section 1.3).
7.3.1 Divide and Conquer
Divide and conquer is a classical approach to improving the efficiency of algorithms
that operate on large data sets (as exemplified in [51]). The underlying problem is
usually spacio-temporal in nature, or at least shares with these problems the following
property: adjacent data elements are more dependent on one another than non-
adjacent elements. It may be said, in general, that adjacent cells in a cellular network
have more influence on one another than non-adjacent cells (for instance, channel
selection), or that adjacent subscribers will affect each others quality of service more
than non-adjacent ones (for instance, base station availability).
When operators design networks manually (i.e. without automatic optimisation) they
invariably divide the network into manageable geographical regions. The automatic
division of the network area is non trivial, but it is easy to compare different approaches
by comparing the performance of the final network designs they produce.
7.3.2 Weighted Evaluation Function
Optimisation methods often rely on the use of a weighted evaluation function. This is
the combination of all the solution performance metrics into a single figure (by the
magnitude of which, each solution may be judged). In this thesis a novel approach was
taken based on the fact that the solutions are only ever compared with one another,
hence the evaluation function can be implemented as a binary comparison operator

183
(each network metric being compared in order of significance). This is the best one can
hope to achieve in the absence of a more complete network performance specification,
for instance, knowledge of the relative significance of traffic cover and test point cover.
This information could be obtained from network operators and embodied in a more
flexible evaluation function.
The ideal way to weight the evaluation function is to relate all network statistics to a
monetary (or dollar) value. A single figure could then be produced that represents the
networks profitability, which must be the principal concern if a network is to remain
competitive. The inclusion of economics necessitates the inclusion of economists, and
this crossing of (typical) departmental boundaries (within a network operating
company) may impede the implementation of this approach. For instance, the economic
analysis given in Section 2.3.1 may be deemed too simplistic, whereas a comprehensive
economic model may be prohibitively complex to understand and implement.

184
Acknowledgments
I would like to thank my supervisor, Steve Hurley, for his constant support and
constructive criticism. I am also indebted to the many practicing engineers who have
freely discussed their work and ideas with me and helped to improve the industrial
relevance of my research. Thanks also to my colleagues in Cardiff for their ideas and
friendship, in particular Simon Chapman and Steve Margetts with whom I have
thrashed out many of the central themes of my work. I would particularly like to thank
my mother, Elinor Kapp, for her practical support and encouragement, without which
none of this would have been possible.
I would like to dedicate this thesis to my wife Erika with all my love.
[1-92]

185
Bibliography
[1] ITU, World Telecommunication Development Report (Executive Summary),
International Telecommunication Union October 1999.
[2] A. Caminada and P. Reininger, Model for GSM Radio Network Optimisation,
presented at the 2nd International Workshop on Discrete Algorithms and Methods
for Mobile Computing and Communications, 1998.
[3] G. Athanasiadou, A. Molina, and A. Nix, The Effects of Control Node Density in
Cellular Network Planning Using the Combination Algorithm for Total
Optimisation (CAT), presented at the PIMRC2000, London, 2000.
[4] J. Button, K. Calderhead, I. Goetz, M. Hodges, R. Patient, R. Pilkington, R.
Reeve, and P. Tattersall, Mobile network design and optimisation, BT Technology
Journal, vol. 14, 1996.
[5] C. Shannon, A mathematical theory of communication, Bell System Technical
Journal, vol. 27, 1948.
[6] K. Holmberg and D. Yuan, A Langrangean Heuristic Based Branch-and-Bound
Approach for the Capacitated Network Design Problem, presented at the
Symposium on Operations Research, 1996.
[7] W. Hershberger, Principles of Communication Systems, Prentice-Hall, 1960.
[8] R. Steele, Mobile Radio Communications, Pentech Press, 1992.
[9] R. Macario, Personal & Mobile Radio Systems, Peter Peregrinus Ltd., 1991.
[10] W. Lee, Mobile Cellular Telecommunications: Analogue and Digital Systems,
McGraw-Hill, 1995.
[11] R. Steele, The Therapy of Percy Comms: A Dialogue on PCS Issues, International
Journal of Wireless Information Networks, vol. 2, pp. 123-132, 1995.
[12] D. Fogel, How to Solve It: Modern Heuristics, Springer, 2000.
[13] C. Papadimitriou and K. Steiglitz, Combinatorial Optimization: Algorithms and
Complexity, Dover Publications, 1998.

186
[14] W. Macready and D. Wolpert, No Free Lunch Theorems for Search, The Santa Fe
Institute 1995.
[15] M. Garey and D. Johnson, Computers and Intractability: A Guide to the Theory of
NP-Completeness, Bell Telephone Laboratories, 1979.
[16] D. Corne, M. Sinclair, and G. Smith, Evolutionary Telecommunications: Past,
Present and Future, presented at the GECCO'99, USA, 1999.
[17] A. Koster, Frequency Assignment - Models and Algorithms, PhD Thesis,
Maastricht University, Maastricht, 1999.
[18] J. Bater, Higher-order (non-binary) modelling and solution techniques for
frequency assignment in mobile communications networks, PhD Thesis, Royal
Holloway, University of London, 2000.
[19] R. Murphey, P. Pardalos, and M. Resende, Frequency Assignment Problems,
Handbook of Combinatorial Optimiziation, vol. 16, 1999.
[20] R. Beck, A. Gamst, R. Simon, and E. Zinn, An Integrated Approach to Cellular
Radio Network Planning, presented at the IEEE 35th Annual Vehicular
Technology Conference, 1985.
[21] S. Hurley, J. Last, D. Smith, and B. Turhan, Algorithms for interference limited
radiobeacon frequency assignment, presented at the Second International
Conference on Information Communications and Signal Processing, Singapore,
1999.
[22] S. Allen, S. Hurley, and D. Smith, Lower bounding techniques for frequency
assignment, Discrete Mathematics, vol. 197/198, 1999.
[23] R. Gower and R. Leese, The Sensitivity of Channel Assignments to Constraint
Specification, presented at the EMC97 Symposium, Zurich, 1997.
[24] S. Hurley, D. Smith, and W. Watkins, Evaluation of Models for Area Coverage,
Department of Computer Science, Cardiff University (Report No. 98003), 2 July
1998.
[25] P. Camarada, G. Schiraldi, F. Talucci, and R. Valla, Mobility and Performance
Modeling in Cellular Communication Networks, Mobile Computing and
Communications Review, vol. 1, 1997.
[26] L. Farkas and L. Nagy, Indoor Base Station Location Optimization using Genetic
Algorithms, presented at the PIMRC2000, London, 2000.
[27] G. Athanasiadou, A. Molina, and A. Nix, The Automatic Location of Base-
Stations for Optimised Cellular Coverage: A New Combinatorial Approach,
presented at the IEEE Conference on Vehicular Technology, 1999.
[28] F. Hillier and G. Liberman, Introduction to Operations Research, McGraw-Hill,
1995.

187
[29] H. Anderson and J. McGeehan, Optimizing Microcell Base Station Locations
Using Simulated Annealing Techniques, presented at the IEEE 44th Conference
on Vehicular Technology, 1994.
[30] C. Pendyala, T. Rappaport, and H. Sherali, Optimal Location of Transmitters for
Micro-Cellular Radio Communication System Design, IEEE Journal on Selected
Areas in Communications, vol. 14, 1996.
[31] M. Wright, Optimization Methods for Base Station Placement in Wireless
Applications, presented at the IEEE 48th Annual Vehicular Technology
Conference, Ottawa, 1998.
[32] S. Fortune, D. Gay, B. Kernighan, O. Landron, R. Valenzuela, and M. Wright,
WISE Design of Indoor Wireless Systems: Practical Computation and
Optimization, IEEE Computational Science & Engineering, pp. 58-68, 1995.
[33] P. Brisset, T. Fruhwirth, and J. Molwitz, Planning Cordless Business
Communication Systems, IEEE Expert/Intelligent Systems & Their Applications,
vol. 11, 1996.
[34] D. Agner, P. Calegari, B. Chamaret, F. Guidec, S. Josselin, P. Kuonen, M.
Pizarosso, and S. Ubeda, Radio Network Planning with Combinatorial
Optimisation Algorithms, presented at the ACTS Mobile Summit, 1996.
[35] R. Menolascino and M. Pizarroso, STORMS Project Final Report, ACTS, Public
A016/CSE/MRM/DR/P/091/a1, April 1999.
[36] C. Glaber, S. Reith, and H. Vollmer, The Complexity of Base Station Positioning
in Cellular Networks, presented at the Workshop on Approximation and
Randomized Algorithms in Communication Networks, Geneva, 2000.
[37] B. Krishnamachari and S. Wicker, Experimental Analysis of Local Search
Algorithms for Optimal Base Station Location, presented at the International
Conference on Evolutionary Computing for Computer, Communication, Control
and Power (ECCAP 2000), Chennai, India, 2000.
[38] Y. Baggi, B. Chopard, P. Luthi, and J. Wagen, Wave Propagation and Optimal
Antenna Layout using a Genetic Algorithm, SPEEDUP Journal, vol. 11, pp. 42-
46, 1998.
[39] R. Mathar, T. Niessen, and K. Tutschku, Interference Minimization in Wireless
Communication Systems by Optimal Cell Site Selection, presented at the 3rd
European Personal Mobile Communications Conference, Paris, 1999.
[40] J. Kwon, Y. Myung, and D. Tcha, Base Station Location in a Cellular CDMA
System, Telecommunication Systems, vol. 14, pp. 163-173, 2000.
[41] L. Correia, Traffic Modelling and its Influence in Planning Optimisation in Dense
Urban Networks, presented at the COST 252/259 Joint Workshop, Bradford,
1998.
[42] A. Ephremides and D. Stamatelos, Spectral Efficiency and Optimal Base
Placement for Indoor Wireless Networks, IEEE Journal on Selected Areas in
Communications, vol. 14, 1996.

188
[43] C. Esposito, E. Olivetan, D. Wagner, and C. Willard, Forward cell planning
networks, presented at the IEEE 44th Conference on Vehicular Technology,
Stockholm, 1994.
[44] L. Correia, L. Ferreira, V. Garcia, and L. Santos, Cellular Planning Optimization
for Non-Uniform Traffic Distributions in GSM, presented at the COST 259
TD(98), Bern,Switzerland, 1998.
[45] A. Caminada and P. Reininger, Genetic Algorithm for Multiperiod Optimisation
of Cellular Network, presented at the 2000 International Conference on Artificial
Intelligence, Las Vegas, 2000.
[46] A. Caminada and P. Reininger, Connectivity Management on Mobile Network
Design, presented at the 10th Conference of the European Consortium for
Mathematics in Industry, 1998.
[47] S. Sato and K. Takeo, Evaluation of a CDMA Cell Design Algorithm Considering
Non-uniformity of Traffic and Base Station Locations, IEICE Transactions, vol.
E81-A, pp. 1367-1377, 1998.
[48] H. Meunier, E. Talabi, and P. Reininer, A multiobjective genetic algorithm for
radio network optimization, presented at the Congress on Evolutionary
Computation, San Diego, USA, 2000.
[49] L. Floriani and G. Mateus, Optimization Models for Effective Cell Planning
Design, presented at the Discrete Algorithms and Methods for Mobile Computing
and Communications, Budapest, Hungary, 1997.
[50] U. Behr, X. Huang, and W. Wiesbeck, A New Approach to Automatic Base
Station Placement in Mobile Network, presented at the International Zurich
Seminar on Broadband Communications, Zurich, 2000.
[51] C. Valenzuela, Evolutionary Divide and Conquer: a novel genetic approach to the
TSP, PhD Thesis, Imperial College of Science, Technology and Medicine,
London, 1995.
[52] R. Barr, B. Golden, J. Kelly, M. Resende, and W. Stewart, Designing and
Reporting on Computational Experiments with Heuristic Methods, Journal of
Heuristics, vol. 1, pp. 33-42, 1995.
[53] E. Damosso (chairman), COST 231 Final Report 1997.
[54] N. Jones, Computability and Complexity : From a Programming Perspective,
MIT Press, 1997.
[55] C. Shannon, Communication in the Presence of Noise, Proceedings of the Institute
of Radio Engineers, vol. 37, pp. 10-21, 1949.
[56] R. Stanley, The Influence of MTSO Placement on Wireless Network Cost,
presented at the PIMRC2000, London, 2000.
[57] I. Godor, J. Harmatos, and A. Szentesi, Planning of Tree-Topology UMTS
Terrestrial Access Networks, presented at the PIMRC2000, London, 2000.

189
[58] G. Walster, Philosophy and Practicalities of Interval Arithmetic, in Reliability in
Computing:The Role of Interval Methods in Scientific Computing, Moore (Editor),
Academic Press Inc., 1988, pp. 309-323.
[59] S. Allen, S. Hurley, D. Smith, and W. Watkins, Frequency Assignment: Methods
and Algorithms, presented at the NATA RTA SET/ISET Symposium on
Frequency Assignment, Sharing and Conservation in Systems (Aerospace),
Aalborg, Denmark, 1998.
[60] M. Alouini and A. Goldsmith, Area Spectral Efficiency of Cellular Mobile Radio
Systems, IEEE Transactions on Vehicular Technology, vol. 48, pp. 1047, 1999.
[61] L. Ammeraal, STL for C++ Programmers, John Wiley & Sons, 1997.
[62] R. Battiti, A. Bertossi, and D. Cavallaro, A Randomized Saturation Degree
Heuristic for Channel Assignment in Cellular Radio Networks, presented at the
WSDAAL'99, Fonte Cerreto, l'Aquila, Italy, 1999.
[63] D. Beasley, D. Bull, and R. Martin, Reducing Epistasis in Combinatorial
Problems by Expansive Coding, presented at the Fifth International Conference on
Genetic Algorithms, University of Illinois, 1993.
[64] D. Beasley, D. Bull, and R. Martin, A Sequential Niche Technique for Multimodal
Function Optimization, Evolutionary Computation, vol. 1, pp. 101-125, 1993.
[65] N. Blaustein, Radio Propagation in Cellular Networks, Artech House, 2000.
[66] D. Bull, N. Canagarajah, and A. Nix, Insights into Mobile Multimedia
Communications, Academic Press, 1999.
[67] A. Caminada and D. Renaud, Evolutionary Methods and Operators for Frequency
Assignment Problem, SPEEDUP Journal (Proceedings of 22nd Workshop), vol.
11, 1998.
[68] A. Capone and M. Trubian, Channel Assignment Problem in Cellular Systems: a
new Model and a Tabu Search Algorithm, IEEE Transactions on Vehicular
Technology, vol. 48, pp. 1252-1260, 1999.
[69] P. Cherriman, L. Hanzo, and F. Romiti, Channel Allocation for Third-generation
Mobile Radio Systems, presented at the ACTS'98, Rhodes, Greece, 1998.
[70] D. Cox and D. Reudink, A Comparison of Some Channel Assignment Strategies in
Large-Scale Mobile Communications Systems, IEEE Transactions on
Communications, vol. COM-20, 1972.
[71] J. Doble, Introduction to Radio Propagation for Fixed and Mobile
Communications, Artech House, 1996.
[72] R. Dorne, P. Galinier, and J. Hao, Tabu Search for Frequency Assignment in
Mobile Radio Networks, Journal of Heuristics, vol. 4, pp. 47-62, 1998.
[73] A. Eisenblatter, A Frequency Assignment Problem in Cellular Phone Networks,
DIMACS series in discrete mathematics for theoretical computer science, vol. 40,
pp. 109-115, 1997.

190
[74] ESPRIT, ESPRIT0016: Algorithms for Radio Network Optimisation (ARNO),
1999.
[75] J. Flood, Telecommunication Networks, IEE, 1997.
[76] L. Floriani and G. Mateus, A Critic of the Graph Coloring Approach in the
Solution of Channel Assignment Problems, presented at the Optimization Methods
For Wireless Networks, Montreal, Canada, 1998.
[77] E. Gamma, R. Helm, R. Johnson, and J. Vlissides, Design Patterns: Elements of
Reusable Object-Oriented Software, Addison-Wesley, 1995.
[78] S. Hurley, D. Smith, and S. Thiel, FASoft: A system for discrete channel
frequency assignment, Radio Science, vol. 32, 1997.
[79] S. Hurley and D. Smith, Bounds for the Frequency Assignment Problem, Discrete
Mathematics 167-168, pp. 571-582, 1997.
[80] S. Hurley, D. Smith, and C. Valenzuela, A Permutation Based Genetic Algorithm
for Minimum Span Frequency Assignment, presented at the 5th International
Conference on Parallel Problem Solving from Nature, Amsterdam, 1998.
[81] A. Kershenbaum, Telecommunications Network Design Algorithms, McGraw-
Hill, 1993.
[82] T. Kurner, Analysis of Cell Assignment Probability Predictions, presented at the
COST 259 TD (98) 05, Berne, 1998.
[83] M. Matsumoto and T. Nishimura, Mersenne Twister: A 623-dimensionally
equidistributed uniform pseudorandom number generator, ACM Transactions on
Modeling and Computer Simulation, vol. 8, pp. 3-30, 1998.
[84] S. Meyers, More Effective C++, Addison Wesley Longman, 1996.
[85] S. Meyers, Effective C++ (2nd edition), Addison Wesley Longman, 1998.
[86] M. Mouly and M. Pautet, The GSM System for Mobile Communications,
published by the authors, 1992.
[87] Y. Myung, J. Kwon, and D. Tcha, Base Station Location in a Cellular CDMA
System, Telecommunication Systems, vol. 14, pp. 163-173, 2000.
[88] H. Niederreiter, Random Number Generation and Quasi-Monte Carlo Methods,
1992.
[89] S. Ubeda and J. Zerovnik, A Randomized Algorithm for a Channel Assignment
Problem, SPEEDUP Journal (Proceedings of 22nd Workshop), vol. 11, 1998.
[90] R. Wessaly, Dimensioning Survivable Capacitated Networks, Technische
Universitt Berlin, 2000.
[91] J. Zoellner, A Breakthrough in Spectrum Conserving Frequency Assignment
Technology, IEEE Transactions on Electromagnetic Compatibility, vol. EMC-19,
pp. 313-319, 1977.

191
[92] S. Zurbes, Frequency Assignment in Cellular Radio by Stochastic Optimization,
presented at the 2nd European Personal Mobile Communications Conference,
Bonn, 1997.

You might also like