Vous êtes sur la page 1sur 332

POWER

SYSTEM
PLANNING
R. L. SULLIVAN
Associate Professor of Electrical Engineering
University of Florida

(
I

McGRAW-HILL INTERNATIONAL BOOK COMPANY


t

New York 11 St. Louis San Francisco Auckland Bogota Diisseldorf


Johann~sburg London Madrid Mexico Montreal New Delhi
Panama Paris Sao Paulo Singapore Sydney Tokyo Toronto
POWER
SYSTEM
PLANNING
R. L. SULLIVAN
Associate Professor of Electrical Engineering
University of Florida

(
McGRAW-HILL INTERNATIONAL BOOK COMPANY

New York~/ St. Louis San Francisco Auckland Bogota Dusseldorf


Johann~sburg London -Madrid Mexico Montreal New Delhi
Panama Paris Sao Paulo Singapore Sydney Tokyo Toronto

----------------------~~------ {
To Sandi, Todd, and Lee

Library of Congress Cataloging in Publication Data


Sullivan, Robert Lee.
Power system planning.

Includes bibliographies and index.


1. Electric power systems. I. Title.
TK1005.S76 621.31 76-27274
ISBN 0-07-061800-3

'0011315
POWER SYSTEM PLANNING
If
Copyright 1977 by McGraw-Hill, Inc. All rights reserved. Printed in the United States of America.
No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any
form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the
prior written permission of the publisher.

1234567890 MAMA 783210987

This book was set in Press Roman by Hemisphere Publishing Corporation. The editors were
Mary Dorfman and Mary A. Phillips; the production supervisor was Rebekah McKinney; and the
compositor was Wayne Hutchins.
The printer and binder was The Maple Press Company.
CONTENTS

-~' Preface xi
.J ',

Chapter 1 INTRODUCTION 1

1.1 The Electric Utility Industry 1


1.1.1 Growth Characteristics 1
1.1.2 Financing Energy Consumption Growth 5
1.2 Generation Systems 6
1.2.1 Fuels 9
1.2.2 Energy Generation and the Environment 9
Transmission Systems 17

LOAD FORECASTING 18

Classification and Characteristics of Loads 19


2.1.1 Classification 19
2.1.2 Characteristics 20
2.2 Approaches to Load Forecasting 22
2.3 Forecasting Methodology . 24
2.3.1 Extrapolation 24
2.3.2 Correlation 25
V
vi Contents

2.4 Energy Forecasting 25


2.4.1 Residential Sales Forecasts 26
2.4.2 Commercial Sales Forecasts 26
2.4.3 Industrial Sales Forecasts 26
2.5 Peak Demand Forecasting 27
2.5.1 Weather Load Model 28
2.5.2 Separating WeatherSensitive and Non-Weather-Sensitive
Components 29
2.6 Non-Weather-Sensitive Forecast (NWSF) 29
2.6.1 Discounted Multiple Regression 30
2.62 Mean and Variance of Coefficients 35
2.6.3 Forecast Variance 36
2. 7 Weather-Sensitive Forecast (WSF) 38
2] .1 Weather Variable 39
2.7.2 Weather Variable Transformation 41
2.7.3 WS Peak Demand Mean and Variance 42
2.7.4 Weekly WS Peak Load Probability Density Function 48
2.8 Total Forecast 51
2.8.1 Seasonal and Annual Forecasts 54
2.9 Annual and Monthly Peak Demand Forecasts 56
2.9.1 Annual Forecast 56
2.9.2 Monthly Forecasts 58
2.10 Summary 59
Problems 59
Bibliography 60

Chapter 3 GENERATION SYSTEM RELIABILITY ANALYSIS 61

3.1 Probabilistic Generating Unit Models 63


3.1.1 Mathematical Description of a Discrete State,
Continuous Transition Markov Process 65
3.1.2 Probability Density Function of Forced Outage
Capacity 67
3.1.3 Generating Unit Models 68
3.2 Probabilistic Load Models 68
3.2.1 Load Probability Distribution 69
3.2.2 Modeling Forecast Uncertainty 70
3.2.3 Expected Value of Demand and Energy 73
3.3 Effective Load 76
3.4 Reliability Analysis for an Isolated System 80
3.4.1 Weekly Effective Load Distribution 80
3.4.2 Reliability Evaluation 81
3.4.3 Unit Maintenance Schedules 83
3.4.4 Unit Effective Load-Carrying Capability 84
Contents vii

3. 5 Interconnected Systems 85
3.5.1 Load and Generator Models 86
3.5.2 Interconnected Effective Loads 86
3.5.3 Interconnected Effective Load Probability
Distributions 87
3.5.4 Reliability Analysis of Interconnected Areas 91
3.6 Summary 94 '
Problems 94
Bibliography 95

Chapter 4 GENERATION SYSTEM COST ANALYSIS 96


4.1 Cost Analysis 97
4. L 1 Capacity Cost 97
4.1.2 Production Costs 99
4.1.3 Plant Cost 100
4.1.4 Timing of Unit Additions 102
4.1.5 System Cost Analysis 104
4.2 Corporate Models 105
4.2.1 Input-Output Parameters 105
4.3 Production Analysis 108
4.3.1 Generation System Models 108
4.3.2 System Load Models 108
4.4 Production Costing 109
4.4.1 ECS Unit Models 109
4.4.2 Economic Commitment 110
4.4.3 Production Analysis Involving Nuclear Units 122
4.4.4 Production Analysis Involving Hydro Units 131
4.5 Fuel Inventories 134
4.6 Energy Transactions and Off-Peak Loading 135
4.6.1 Energy Purchases 136
4.6.2 Energy Sales 137
4.6.3 Off-Peak Energy Utilization 137
4.7 Environmental Cost 138
4.7 .1 Thermal Pollution 139
4.7.2 Air Pollution 140
4.7.3 Ambient Air Quality Assessment 141
4.8 Summary 148
Problems 148
Bibliography 150

Chapter 5 POWER-FLOW ANALYSIS 151


5.1 Corridor Selection 153
5.2 Steady-State Component Models 153

----~------------- -----------.----------.
viii Contents

5.2.1 The Synchronous Machine 154


5.2.2 The Power Transformer 154
5.2.3 The Transmission/Subtransmission/Distribution
Lines 156
5.3 Steady-State Power System Model 159
5.3.1 Variable and Parameter Constraints 159
5.3.2 State Variable Model of System 160
5.3.3 State Variable Model of Three-Bus System 166
5.3.4 State Variable Model for Two Interconnected Areas 170
5.4 Newton's Power-Flow Solution Method 172
5.4.1 Basic Solution Procedure 172
5.4.2 Gaussian Elimination and Optimal Ordering 177
5.4.3 Accelerated Newton's Method 188
5.4.4 Compact Storage Schemes 189
5.5 Sensitivity An~lysis 189
5.5.1 The Sensitivity Equation 190
5.5.2 Evaluating aF/iJU, oF/aD, aHtaX, aH/aU 191
5.5.3 The Two-Area System 195
5.6 Optimal Power Flows 198
5.6.1 Basic Problem Formulation 198
5.7 Reactive Power-Flow Analysis 204
5. 7 .1 Reactive Power and Voltage Considerations 205
5.7.2 Voltage Control 210
5.8 Stochastic Power Flows 214
5.8.1 Problem Formulation 214
5.8.2 Specifying the Covariance Matrix C17 217
5.8.3 Solution Procedure 219
5.9 Summary 219
Prob( ems 220
Bibliography 221

Chapter 6 TRANSMISSION SYSTEM RELIABILITY ANALYSIS 223

6.1 Deterministic Contingency Analysis 224


6.1.1 DC Power-Flow Contingency Analysis 224
6.1.2 AZ Matrix Method for Contingency Analysis 230
6.2 Probabilistic Transmission System Reliability Analysis 23t
6.2.1 Probabilistic Element Models 238
6.2.2 Basic Philosophy 238
6.3 Capacity State Classification by Subsets 241
6.3.1 Determining the Upper and Lower Critical Capacity
States x 0 and X 0 243
6.3.2 Exhausting Subset U 245
6.3.3 Maximum-Flow/Minimum-Cut Algorithm 246
Contents ix

6.3.4 Labeling Algorithm 247


6.3.5 The Augmentation Algorithm 249
6.4 Subset Decomposition for System LOLP Calculations 251
6.5 Subset Decomposition for Element LOLP and e{DNS)
Calculations 254
6.5.1 Expressions for e(DNS), LOLPm and Em (DNS) 255
6.5.2 Decomposition Rule for Obtaining Bn 258
6.6 Reliability Calculations for Single Areas 263
6.6.1 Solution Procedure Revisited 263
6.6.2 Completion of Sample System Analysis 264
6.6.3 Element LOLPs 269
6.6.4 Calculation of e(DNS) 271
6.6.5 Interpretation of Results 272
6.7 Multiarea Reliability Analysis 275
6.7.1 Element Models and Reliability Indices for Multiarea
Reliability Analysis 276
6.7.2 System Model 277
6.7 .3 General Solution Procedure 277
6.7.4 Interconnected System Reliability Analysis 282
6.8 __ Summary 286
Problems 287
Bibliography 288

Chapter 7 AUTOMATED TRANSMISSION SYSTEM EXPANSION PLANNING 289

7 .1 Automated Transmission System Planning Concepts 289


7 .1.1 Tellegen's Theorem 290
7 .1.2 Network Sensitivities 292
7 .2 Automated Network Design 298
7 .2.1 Problem Formulation 298
7.2.2 Determining the Gradient Vector 299
7.2.3 An Example 301
7.3 Automated Transmission Planning: A DC Method 304
7 .3.1 Problem Formulation 304
7.3.2 Determining the Gradient Vector 305
7.3.3 Including Practical Considerations 307
7.3.4 Expanding the Three-Bus System 309
7.4 Automated Transmission Planning Using Interactive Graphics 313
7.4.1 Characteristics of Interactive Graphic Systems 313
7.4.2 Interactive Graphic Power-Flow Analysis 315
7.4.3 Interactive Graphic Automated Transmission Design 318
7.5 Summary 318
Problems 318
Bibliography 319
Index 321

--------------.-----------------------------
PREFACE

A survey of textbooks available (1976) in the power system area reveals the lack
a
of any text devoted to power system planning. Rather, most books are composite of
notions used throughout the electric utility industry as a whole. Such books are
intended to introduce readers to all facets of the industry so that they may be
prepared to serve the industry in any of several areas. Typically, such texts are used in
an undergraduate program; as of this writing, almost a dozen books serve the needs of
the engineering community in this way. Books devoted to a particular area of the
utility industry are also available, but few, if any, concentrate on the methods that are
currently used in system planning.
This text, therefore, has been written to focus attention on ( l) the overall
process called power system planning and (2) the mathematical details of various
analytical tools. The nature of this book makes it a senior or graduate--level text, and it
has been written for those who already have a basic understanding of material that
appears in any undergraduate book on power system analysis.
The chapter sequence reflects, as much as possible, the natural progression of
events as they occur in industry, thus always providing the practitioner with a frame of
reference. At the same time, a reader unfamiliar with the industry obtains some insight
into the sequence of events that must take place in system planning. The text is
replete with built-in numerical examples to illustrate the concepts presented, with
xi
xii Preface

examples appearing as a natural extension of this material. Through the use of a


common sample system in all examples, the continuity of the material is further
enhanced and the amount of memory work required of the reader reduced.
As the contents indicates, the reader is given a grand tour of generation and
transmission system planning, starting with the trials and tribulations of forecasting
future requirements and moving through' the use of sophisticated computer-aided
design methods for automated transmission planning. Let it be emphasized that the
material presented should not be considered exhaustive or indicative of the way a
particular utility approaches system planning. There are perhaps as many different
approaches to identifying the ultimate expansion plan as there are utilities.- The
material presented here reflects an approach with characteristics not too different frorr..
those currently used.
In order to limit the length of thls text, distribudon system planning methods
are not specifically considered, although much of the material presented is applicable
to distribution planning. In addition, short circuit analysis and transient stability
analysis are omitted because these tools are used more in short-term system planning
to design system protection schemes and size-associated protection equipment.
Few books are written that include only the ideas of the author, and this book is
no exception. The education and support this author has received from the electric
utility industry in Florida is almost immeasurable, and I would like to acknowledge
their help here and now. Both Florida Power Corporation and Tampa Electric
Company deserve additional credit for their guidance. Mr. M. F. Hebb, Mr. H. B. Wells,
Mr. W. 0. May, and Mr. P. Dagastino, all with Florida Power Corporation,- have
provided information and on-the-job training that made this book possible. Similarly,
Mr. Lester tnm, Jr., Mr. R. E. Proctor, and Mr. A. N. Darlington, of Tampa Electric
Company, have contributed significantly to this project by sharing their time and
utility experience. It should be recorded that the material presented in no way reflects
either the position or philosophy of those companies that supported the author in this
endeavor. Finally, I acknowledge the help and encouragement received from my students
and the support of the University of Florida and, in particular, the assistance provided by
Ms. Edwina Huggins in typing early versions of the manuscript.
INTRODUCTION

The purpose of this introductory chapter is to familiarize the reader with past and
future trends in the electric utility industry. This qualitative information will, we hope,
put the industry in perspective before any attempt is made to present the detailed
analytical methods employed by the industry. Our emphasis will fall on those aspects
of the industry that directly relate to the topics discussed in the remaining six
chapters.

i.
.j <

1.1 THE ELECTRIC UTILITY INDUSTRY


By necessity all comments made here will be based on the industry as it exists in the
f ' United States. Fortunately, the technical characteristics of the industry in the United
States are similar to those of other countries; therefore much of what will be said in
this chapter applies equally well the world over.

1.1.1 Growth Characteristics


Figure 1.1 concisely demonstrates the role electric energy plays in maintaining a highly
developed industrialized country. Figure 1.2 projects that growth rate in generation
will drop off because of a leveling population growth rate and energy consumption per
1
N
250
240
230
220
20,000 210
20 ti 100 200
C:
19
10,000
;:;,O{:-
18 a:a'
:=ca
190
90 180
::..c
q,~
17 :E.5 170
5,000
~{:-
"/I. 16 00 80 160
.i:,~ .,.,.,. 15
150
<:~ .,
($' t; 14 350 70 140
C: 13 130
.,
)(

"tl 12 300 60 120


.E : 11
<( 110
1,000 10 250 10050
9 90
((\~,'.101\
8 200 40 80 .,
col\'>u
500
, el\eti;i
7 70 ., 5.
6 150 30 60
'1:0\~'
5 .,..."
50 C.
4 100 20 40 ~
3 30 -"
.,, "tl
C:
2 50 C: Ill:,
100 _ _...__ _ _ _ _ _ _...__ __._ __.__ _- 1 .2
0
1902 1910 1920 1930 1940 1950 1960 1970 ~ i=
Year 1900 1950 2000 2050 2100 2150 2200 2250
Year
Fig. 1.1 Indices of electricity generation and total energy use in the U.S.
1902-1972. Electric-energy growth rate exceeds that of total energy by Fig, 1.2 Growth patterns for electric energy consumption, generation,
almost 2 times during this century and will draw away. Reprinted from and population and generation growth rate. Reprinted from June I,
June 1, 1974 issue of Electrical World. Copyright 1974,McGraw-Hill, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc. All
Inc. All rights reserved. rights reserved.

_,
r ... - - --" ""

:! ..........A............... l .... .

1l g :,
10(1 ( I
1902 1910
I
1920
I
1930
I
1940
I
1950
I
1960
,I
1970
1900 1950 2000 2050 2100 2150 2200 2250
lLi 0
F
Year
Year
Fig. 1.1 Indices of electricity generation and total energy use in the U.S.
1902-1972. Electric-energy growth rate exceeds that of total energy by Fig. 1.2 Growth patterns for electric energy consumption, generation,
almost 2~ times during this century and will draw away. Reprinted from and population and generation growth rate. Reprinted from June 1,
June 1, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc. All
Inc. All rights reserved. rights reserved.

400,----------------~

1,700
1,600
1,500
1,400
300
1,300
1,200
;::
(!) .
C:
1,100
i
;:;
0
~ 1,000
.0
~
"'u 200 'O 900
$
.,
"C
.,e 800
~ .,
C:

.s "'
.r:.
700
s:
~ 600
500
100
400
300
200

o t::: ;r::::::::; :::::::-:<:: , ---,


1930 1940 1950 1960 1970 1920 1930 1940 1950 1960 1970
Year Year

Fig. 1.3 Growth in installed capacity by utility type. Reprinted from Fig. 1.4 Energy growth by type of utility. Reprinted from June 1,
June 1, 1974 issue of Electrical World. Copyright 1974, McGraw- 1974 issue of Electrical World. Copyright 1974, McGraw-Hill,
t.l Hill, Inc. All rights reserved. Inc. All rights reserved.
4 Power System Planning
t
capita. Obviously, if these projections do occur, the utility industry will eventually
reach a no-growth state-an interesting notion indeed. Of more immediate interest is
the fact that the growth patterns are dynamic and hence must be constantly
l
scrutinized by each utility to determine individual load characteristics and, finally,
expansion plans. Chapter 2 is devoted to a detailed discussion of load characteristics
and load forecasts.
12,--------------------7'1

11
(Estimated)

10

"'
C 8
-~
:.0
~
tJ)
:::> 7
.~..
1i
::,
a. 6
E
.......
0

'tJ
"'
,!!? 5
~
"'
'tJ
C
::,
u. 4

1887 1897 1907 1917 1927 1937 1947 1957 1967 '74
Year

Fig. 1.5 Public financing: Investor-owned electric utility industry


1918-1974 (includes refundings and excludes pollution control
bonds). Reprinted from June I, 1974 issue of Electrical World.
Copnight 1974, McGraw-Hill, Inc. All rights reserved.
Introduction 5

Of approximately 3,400 utilities in the United States, 500 are investor owned;
2,000 are municipals; the remaining 1,000 are cooperatives, public power districts, state
and federal organizations. Figures 1.3 and 1.4 depict the relative division of installed
capacity and energy generation between various utilities. Investor-owned utilities clearly
shoulder the major responsibility for producing electric energy; however, the coopera-
tives, public, power districts, etc., must be credited with making electric energy available
to rural America. In the early 1930s, prior to the passage of the Rural Electrification Act,
only two-thirds of the nation had access to electricity.

1.1.2 Financing Energy Consumption Growth


Financing tremendous growth in energy requirements has been of critical concern
to all utilities, both public and private, but certainly the investor-owned, publicly
regulated utilities have more difficulty walking the fine line between bankruptcy and
prosperity than the other utilities. Figure 1.5 shows public dollars acquired by the
investor-owned utilities to support the development of their systems through 1974. In
addition, Table 1.1 depicts the capitalization strategy employed by the industry for
two different periods: 1969-1973 and 1974-1978. It should be noted that approxi-
mately two-thirds of the capital was generated through the sale of bonds and of
preferred and common stock. Indeed the electric utility industry is a capital-intensive
iridustry, as these data make evident. It should also be emphasized that a mere 10% of
construction capital is obtained from retained earnings or profit, if you will.
Table 1.2 illustrates the financing strategy used in the industry for 1972 (actual)
and 1q7g (projected). More than half of the dollars needed to support construction
came from bonds alone, at a cost of US$6,200 million, or 7.16% of the amount

TABLE 1.1 Source of funds statement for total investor-owned


utility industry-1969-1973vs. 1974-1978 (US$ millions)a
1969-1973 1974-1978
Amount Percent Amount Percent
(USA) of total (USA) of total
External generation
Debt 23,833 60 34,200 61
Preferred 6,724 17 8,300 15
Common 9,052 23 13,800 24
Total 39,609 100 56,300
-
100
Internal generation
Retained earnings 5,884 28 10,000 28
Depreciation 13,345 64 21,600 60
Deferred taxes 1,726 8 4,600 12
Total
- --
20,955 100 36,200 100
Construction US$58,855 US$92,500
External generation/Construction 67.3% 60.9%
aReprinted from June 1, 1974 issue of Electrical World Copyright 1974, McGraw-Hill, Inc.
All rights reserved.
6 Power System Planning

TABLE 1.2 Capitalization and cost relationships for the investor-owned


electric utility industry (US$ millions)a
Average Cost Capitalization Capitalization
ca pitaliza tio n relationships ratios returnb

Year 19,78-projec ted


Debt US$ 86,600 US$6,200C 7.16% 52.0% 3.72%
Preferred 20,400 1,444d 7.08 12.3 0.87
Common 59,400 7,600e 12. 79 35.7 4.56
Total US$166,400 100.0 9.15
Year 1972-actual
Debt US$49,573 US$3,048 6.15% 54.2% 3.33%
Preferred 10,246 635 6.20 11.2 0.69
Common 31,687 3,721 11. 74 34.6 4.06
Total US$91,506 100.0 8.08
0
Reprinted from June 1, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill,Inc.
All rights reserved.
bThe percentage cost relationships of each category (debt, preferred, and common) multiplied
by its capitalization ratio.
CJnterest paid and as a percentage of average debt (long-term and bank loans outstanding).
dPreferred dividends paid and as a percentage of average preferred outstanding.
ecommon earnings and as a percentage of average common equity.

borrowed. Of the remaining capital needed, 11 % came from the sale of preferred stock
and 35% from the sale of common stock. In short, the cost of construction money for
1972 was 8.08% of that borrowed. We shall see in Chap. 4 that the capitalization
ratios and the cost of capital have a pronounced effect on the capacity cost of new
generation.

1.2 GENERATION SYSTEMS


Beginning with the Pearl Street Station, built in New York in 1881 , central generating
stations have spread across the entire country. To deliver energy consumed in a fashion
depicted by the typical load duration curve shown in Fig. 1.6, the utilities have relied
on everything from gas peaking units to large nuclear units. From Fig. 1.7 we can
easily see that fossil plants have dominated generation systems, followed by hydro and
then nuclear plants. Also quite evident is the growth in nuclear generation. We shall see
in Chap. 4 that this trend toward a greater percentage of nuclear generation is
motivated by favorable economics and by a concern that fossil fuel resources are
slowly dwindling.
As illustrated in Fig. 1.6, loads tend to have substantial short-duration peaks that
cannot be economically supplied using large base-load units. To supply these peaks, the
industry has relied on gas turbines, internal combustion engines, and pumped hydro
units. The role each has played in meeting peak demands is shown in Table 1.3. ,
Clearly, gas turbines provide the industry with most of its peaking capacity; however, -~
"-'-~~1,:'-> . ;~_r;~.'1,~ ,_ ,~-11,1; ,-:..., ,:., ...... ,!>.-;, ~..,,.-~ ~-"'~"-',' ___,,,,,,, l,.'

1,700
Reserve
1,600
1,500
100------------------
1,400
1,300
80 1,200
1,100
70
"' 1,000
C:
0
60 900
ii
"C 800

3"' 50 Intermediate load
~
.:,/. 700
~
600
401::------------------..
500
30
400
20 Base load 300
200
10
100
0'---''----''----''----'---'-....J.-....J.-___.._,_ __.__ _,. I::::
0 10 20 30 40 50 60 70 80 90 100 1920 1930 1940 1950 1960 1970
% Time, one year Year
Fig. 1.6 Annual load distribution curve. Reprinted from June 1, Fig. 1.7 Electric energy production by generation type. Reprinted from
1974 issue of Electrical World. Copyright 1974, McGraw-Hill, June 1, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill,
Inc. All rights reserved. Inc. All rights reserved .

...,
8 Power System Planning

TABLE 1.3 Electric power industry capacity, 1974 and beyond (MW) 0
Total capacity Capacity additions Present capacity
(Dec. 31, 1973) (1974 and beyond) plus additions
Fuel type MW % MW % MW %
Hydro 53,667 12.2 11,874 2.8 65,541 7.6
Pumped storage 7,613 1.7 16,807 4.0 24,420 2.8
Fossil fuel 318,357 72.7 164,006 39.2 482,363 56.3
Nuclear 21,070 4.8 199,253 47.6 220,323 25.7
Internal combustion 4,908 1.1 593 0.1 5,501 0.7
Gas turbine 32,877 7.5 26,371 6.3 59,248 6.9
Total industry 438,492 100.0 418,904 100.0 857,396 100.0
a Reprinted from June l, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc.
All rights reserved.

as Fig. 1.8 vividly displays, pumped hydro unit capacities have steadily increased since
their introduction.
Because of random outages in generation capacity, the utility industry attempts
to maintain sufficient reserves to insure a high degree of reliability. Planning a
generation system that is both economical and reliable is a major responsibility of the
system planner. Too much capacity or too little can have major consequences for the
financial integrity of a utility. More will be said of generation system reliability in
Chap. 3.

_,,,. .,,.
500
/
.,,..,,.,.,,.
400
.,,.
/ " Racoooa M,.
400MW
s:~
.; 300
:,
.,a. / Cornwall
:,
0
'=C
::,
200 LT,omSa,k

220MW
258 MW

100
Hiawassee
~ 62 MW
Flatiron 9 MW I
0
1950 1960 1970 1980
Year of order

Fig. 1.8 Growth in pumped hydro unit capacities. Reprinted from June 1,
1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc. All
rights reserved.

;,,
1
--------~---------------------------------------*
t
t
' Introduction 9

1.2.1 Fuels
The availability of fuels is a primary factor in determining the best generation mix.
Figure 1.9 depicts the extent to which each type of fuel has been used and will be used to
meet U.S. electric energy needs. A notable trend is the move away from a fossil-based fuel
supply toward a fuel supply of coal and uranium. The reason for this shift in fuel type is
best explained in Fig. 1.10, a graphic estimate of fuel resources in the United States.
Fuel prices are also important factors in determining the most economical
generation mix. In general it is felt that fuel prices will continue to rise at or near the
inflation rate. It goes without saying that fuel prices during the last five years
(1970-1975) have soared upward; oil prices, for instance, have increased threefold in
five short years. A similar story can be told of the other fuels, for which in some
instances the high prices are partly due to the need for cleaner fuels.

1.2.2 Energy Generation and the Environment


The impact the environmental movement is having on the industry, both public and
private, is substantial. In terms of capital expenditures alone, the industry has been
kWh, billions

Nuclear

1965 1970 1971 1972 1973 1975 1980


L- Forecast __J
Year
Fig. 1.9 Energy sources for total electric utility industry 1965-1980. Reprinted from June 1,
1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc. All rights reserved.
10 Power System Planning

Gas Oil
150 years 150 years

Fig. 1.10 United States gas, oil, coal, and nuclear fuel reserves
(excludes imports). Reprinted from June 1, 1974 issue of Electrical
World. Copyright 1974, McGraw-Hill, Inc. All rights reserved.

hard hit, as is evident in Table 1 .4. To put these costs into perspective, we must realize that
a 1000-MW nuclear generating unit costs about US$500-800 million (1975). Thus, the an-
nual operating expenses for the industry for environmental control have been approxi-
mately equal to the cost of a large nuclear unit, and the capital investment through 1973
for pollution control devices has been equivalent to the cost of ten large nuclear units.
Air Pollution Control
Table 1.5 shows the costs associated with air quality control for fossil plants in the
500- to I ,000-MW capacity range. On a per kilowatt basis these costs are I 0-20% of
Introduction 11

TABLE 1.4 Environmental control: How much utilities are


spending now (US$ thousands)a
Capital investment, total in service Operating expenses
Dec. 31, 1971 Dec.31,1973 1971 1973
Air quality $1,241,700 $1,800,000 $364,500 $629,000
Water quality 453,400 720,000 19,700 28,000
Undergrounding 3,690,300 4,320,000 12,400 22,126
Otherb 242,700 350,000 7,400 11,391
Total $5,628,100 $7,190,000 $404,000 $690,517
aReprinted from June 1, 1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc.
All rights reserved.
b Aesthetic, recreational, etc.

the capacity cost for fossil units. Considering that production costs for fossil units vary
between 15 and 30 mils/kWh, the operating costs for pollution control equipment,
again, are 10-20% of production costs.
Water Quality Control
Because of increased pressure on the industry to reduce thermal pollution, cooling
towers, ponds, and the like have become an essential part of the modern power plant.
TABLE 1.5 Air control costs for new units (500-1,000 MW)a
Particulate control
Precipitators
t---------'=~.:,;+~>%~
45 C::J Coal units
Venturi scrubbers 45
r------=-==~H'fi'FF""" 1:/{(::j Oil units
Bag filters 20
SO, control
Lime scrubbing
Limestone scrubbing 35
NOx control
t-----------='-'-'==~=""
1
Combustion 7
modification
0 10 20 30 40 50 60
US$ per kW
Particulate control
Precipitators
Venturi scrubbers
Bag filters
SO, control
Limestone scrubbing
Lime scrubbing
NOx control
Combustion
modification 0.06 0.2
0 0.5 1.0 1.5 2.0 2.5
Mills per k\1\/h
aReprinted from June 1, 1974 issue of Electrical World. Copyright 1974,
McGraw-Hill, Inc. All rights reserved.
170

1
160

150
r
140

,. 4Jlll--'v1""""~"'""~~.r----.~ ..
~,~ .

I JO,, Cooling tower

120

110

100

90

80

70

60

50

40

.
C
0
30

-E
20

Fig. 1.11 Estimated cooling tower sales to electric utilities.


Reprinted from June 1, 1974 issue of Electrical World. Copyright
1974, McGraw-Hill, Inc. All rights reserved.

12

. ,,,~ ,.....~,.:-;<~"''"'"'!l'""',l\'.:.:,...~.~~~~.~~--,,,,"1'111~- . . ........-.......


" .... 'l_l:l,;'r,,,(J> ijl,.,,.,.......,. --~,,~~-~~---. ,..,,..""".~...,.,,..... - ' ,. '. ~.... ' . ,

Off-shore subaqueous Fossil


discharge ~
Nuclear
'
100.--------------------, Cooling pond

Lakes and spray ponds


Natural draft
80 evaporated

Mechanical
60 draft evaporated
...
C

~ i

~ "' Parallel path
40 =0
0
(J

Fan assist
evaporated 8.90
20
Mechanical draft wet towers
28.90
Mechanical

0 .__....___.._ _.___._....____,__..,___,__.,___.__.__.___.
draft dry :::::::::::t<-:=~ .. ; :.::{;::-,- ::=:.,; -..::~ .::.~:.t:;:, :i1:~;f;J 45.10

1972 1974 1976 1978 1980 1982 1984 0 10 20 30 40 50


Year US$ per kW

Fig. 1.12 Percent of annual additions by type of supplementary Fig.1.13 Thermal discharge control costs for fossil and nuclear units.
cooling. Reprinted from June 1, 1974 issue of Electrical World. Reprinted from June 1, 1974 issue of Electrical World. Copyright
Copyright 1974, McGraw-Hill, Inc. All rights reserved . 1974, McGraw-Hill, Inc. All rights reserved.
...
w
-....
Approximate load concentrations ~
1,000 MW or greater 161-345 kV - - -
Q 300-1,000 MW 500 kV - - -
765kV - - - -
800 kV DC----

Back-bone transmission
(existing and authorized)
showing EHV and
representative portions
of 230-kV and 161-kV lines

Fig. 1.14 Major transmission in service or authorized. Reprinted from June 1, 1974 issue of
Electrical World. Copyright 1974, McGraw-Hill, Inc. All rights reserved.

~
., ..........,......<""...,.~"'-'r,. u-..-~- J:a" i ,, - .,,,_ ".,...,.,;.- ;o-t.~,';",o-,,., "' ... ,,,.,. _--.~ . ~ . ~- ~ ........... , , .,

'<11;,,P,,:~'Oi.i'#,:~~~~-- - ~..~
..

::::: ~i:ilit!:!::!) .;:::;:' :::,::-:: l


3,400
3,200


1ex1sung ana aumonzeai
showing EHV and
representative portions
of 230-kV and 161-kV lines

d-~ .. , . , ... ' - -


Fig. 1.14 Major transmission in service or authorized. Reprinted from June 1, 1974 issue of
Electrical World. Copyright 1974, McGraw-Hill, Inc. All rights reserved.
' .a.-~--~
, . , , ; : , ~ ; , i ~ ~ - ~ , l ~ ~..

,~'

3,800
3,600
3,400

3,200 I~
3,000
2,800

6,000 . 2,600 ~ I
5,600 ':t;';;/,<,.'L~ 2,400 \t
:"f:'.t)/::;;f:
5,200 1973 est. .,, 2,200
4,800
.
I
' '- -
~
'! 2,000
I '5
4,400 I ~ 1,800
/ ...... 1974 est. CJ
] 4,000 I 1,600

::E:,
I
'!5 3,600
~ 3,200
u
2,800

2,400 800

2,000
1,600
1,200
,/

1966 '67 '68 '69 '70 '71 '72 '73 '74 '75 '76 '77 1966 '67 '68 '69 '70 '71 '72 '73 '74 '75 '76 '77
Vear Vear

Fig. 1.15 Overhead transmission added. Reprinted from Fig. 1.16 Overhead EHV Transmission added. Reprinted from
...
UI
September l, 1974 issue of Electrical World. Copyright 1974, September l, 197 4 issue of Electrical World. Copyright
McGraw-Hill, Inc. All rights reserved. 1974, McGraw-Hill, Inc. All rights reserved.
16 Power System Planning
l

Figure 1.11 illustrates the extent to which cooling towers are being used to meet
thermal pollution standards. Figure 1.12 depicts the degree of use of various types of
cooling methods, with mechanical draft towers tending to be most prevalent. In terms
of increasing the capacity cost of new units, the capacity cost of various cooling
methods is given in Fig. 1.13. We should note that these costs are generally 2-10% of
the corresponding unit capacity cost.
Depending on the type of pollution abatement equipment used, we see from the
figures above that this equipment can increase the cost of modern plants by as much
as one-third. Whether this is too high a price to pay is beyond the scope of this book.

200
190
180
I
170 I
160 I
150 I
140 I
130 I
120
I
/"-.
"'"' 110
E
.,,.,,,
1974 est.

.<::: 100
::,
~
u 90
80
70
60
50
40
30
20
10 500 kV
:::::..-
1966 '67 '68 '69 '70 '71 '72 '73 '74 '75 '76 '77

Year
Fig. 1.17 Transmission cables added. Reprinted from September 1,
1974 issue of Electrical World. Copyright 1974, McGraw-Hill, Inc.
All rights reserved.
Introduction 17

1.3 TRANSMISSION SYSTEMS


Transmission networks span many countries from one end to the other, making
available the electric energy generated at central stations. The importance of a reliable
transmission sy~tem to deliver electric energy to load centers at the proper voltage level
and frequency is obvious. The growth of transmission networks in this country is
illustrated by Fig. 1.14. In this figure we see that the backbone of the network
consists of 161- to 765-kV AC lines with one 800-kV DC link. Figures 1.15 through
1.17 show the actual and projected circuit miles for each type of line, including
underground cables. It is important to note that underground transmission is presently
limited to 345 kV and lower and is used very sparingly. Unfortunately, underground
cables now have a cost disadvantage that ranges from 5: 1 to 20: 1 over comparable
overhead lines and, of course, arc limited in capacity to approximately 1,000 MW for
distances up to 50 miles. No doubt, major technological advances will bring
underground cables more in line with overhead transmission costs, but it is difficult
now to see how they can compete purely on a dollar-cost basis.
To plan a reliable transmission network requires the use of simulation tools such
as power-flow analysis and contingency analysis, both discussed in Chaps. 5 and 6.
With these analytical tools the planner can evaluate the capability of each alternative
expansion plan for providing the level of performance desired. Since lines must be
operated within their design capability, the network topology must be such as to avoid
any damaging overloads, even under single- and certain double-contingency situations.
In view of the relative difficulty of identifying the best expansion plans using
only power-flow analysis and modern batch processors, some utilities are finding the
use of interactive graphic systems and associated software to be convenient.
Considering this developing interest in interactive graphic planning, Chap. 7 is devoted
to automated network design and interactive graphic systems. Although the discussion
will scarcely be complete, the reader will be exposed at least to the basic ideas.
LOAD FORECASTING

,,

Power system expansion planning starts with a forecast of anticipated future load f
,,'
requirements. Estimates of both demand and energy requirements are crucial to
effective system planning. Demand forecasts are used to determine the capacity of
generation, transmission, and distribution system additions, and energy forecasts 1
determine the type of facilities required. For example, if a demand forecast stated that '
f
100 MW of capacity was needed and a corresponding annual energy forecast stated the t
need for only 200,000 MWh, in all probability a peaking generating unit would be l, ,
installed, instead of a base-load unit; the difference in cost between these two is often
substantial. Load forecasts are also used to establish procurement policies for
construction capital, where, for sound operation, a balance must be maintained in the
use of debt and equity capital. Further, energy forecasts are needed to determine
future fuel requirements and, if necessary when fuel prices soar, rate relief to maintain
an adequate rate of return. In summary, a good forecast reflecting current and future
trends, tempered with good judgment, is the key to all planning, indeed to financial
success.
Throughout this and subsequent chapters the words "load," "demand," energy,"
and "forecast" will be used often, and a few words clearly defining these terms are in
order. load is a general term meaning either demand or energy, where demand is the
18
I Load Forecasting 19

time rate of change of energy. The term forecast refers to projected load requirements
t determined using a systematic process of defining future loads in sufficient quantitative

I' detail to permit important system expansion decisions to be made.


The accuracy of a forecast is crucial to any electric utility, since it dictates the
timing and characteristics of major system additions. A forecast that is too low can
t easily result in lost revenue from sales to neighboring utilities or even in load
curtailment. On the other hand, forecasts that are too high can result in severe
financial problems due to excessive investment in an electric plant that is not fully
utilized or, equivalently, is operated at low capacity factors. Capacity factor is defined
as the ratio of average energy supplied to maximum energy capability. Unfortunately,
an accurate forecast depends on the judgment of the forecaster, and it is impossible to
rely strictly on analytical procedures to obtain an accurate forecast. Good judgment
cannot be emphasized enough in forecasting future requirements.

2.1 CLASSIFICATION AND CHARACTERISTICS OF LOADS


In some cases a total forecast is obtained by combining forecasts for various classes of
customers. Therefore, in this section we shall briefly discuss the load classifications
normally used as well as some important characteristics of various types of loads.

2.1.1 Classification
Loads may be classified broadly as residential, commercial and industrial, and other.
Residential customers use energy for domestic purposes, whereas commercial and
industrial customers obviously use energy for commercial and industrial purposes.
Other customers are municipalities or divisions of state and federal governments using
energy for street and highway lighting. In addition, sales to public authorities and to
railroads and railways, sales for resale, and interdepartmental sales also come under the
"other" classification.
Although these classifications may seem well defined, they may in fact be poorly
defined for forecasting purposes. For example, some utilities meter mobile homes as
individual residential customers; others meter the entire mobile home park as a
commercial customer. Unfortunately for forecasting purposes, these classifications
overlap in the sense that customers in a given class do not have characteristics unique
to that class; that is, the classifications are not mutually exclusive. Within the broad
classes mentioned, further subdivisions may be defined and in fact are defined by
many utilities. Residential customers may be subdivided into rural, urban, and other
subdivisions that may be valuable in some cases.
Perhaps more useful classifications of customers for forecasting purposes would
be made by type of use, level of use, rate schedule, or geographic area. Classification
by rate schedule is potentially a viable way to classify customers, because it tends to
lump similar types of customers into the same rate category and because the energy
data for each category are readily available. However, each utility has a different
philosophy in developing rate schedules, and care must be taken when forecasting load
requirements for pool~ consisting of several utilities.
20 Power System Planning

2 .1.2 Characteristics
Of the three ,;?~d classes of loads, residential loads have _t._he, mpst constant annual
growth rate and the most seasonal fluctuations. The ~~ variations of the
residential component ~many cases are responsible for the seasonal variations in
system peak, the ext&'it of the residential influence depending on the percentage of
total system load that is residential. Without a doubt, this characteristic is due to the
widespread use of weather-sensitive devices such as space heaters and air conditioners.
Other high-energy devices used by residential customers are water heaters, refrigerators,
and dryers. Refrigeration loads tend to have constant characteristics as compared to
the cyclical load characteristics of dryers and water heaters. Table 2.1, where numerical
entries are only representative, shows the extent to which residential customers depend
on electric energy. Barring a substantial leap forward in solar energy technology in the
next decade-which could indeed take place-the increase in per capita consumption
due to increases in weather-sensitive residential loads will play an even greater role in
determining system load patterns.
Commercial loads are also characterized by seasonal fluctuations, and again the
fluctuations are primarily due to the extensive use of air conditioning and space
heating. In the future it appears inevitable that electric heating and air conditioning
will be used even more, as will many appliances. The introduction of new devices-the
electric car, rapid transit, etc.-will certainly influence the characteristics of future
load.
Industrial loads are considered base loads that contain little weather-
dependent variation. However, depending on type of industry, these loads
may have unique characteristics because of shift operations, etc. Industries involved in
mining or any other high-demand, low-energy operation are usually unique enough to
warrant special consideration.
Other loads, as described- earlier, may have seasonal fluctuations depending on
specific cases. In most instances, however, the growth trend for this classification is
considered stable, although this does not mean it will remain stable. Electrification of
this country's railroads using advanced linear induction motor concepts could affect
the relative importance of the "other" classification.
As saturation level and per capita consumption increase, reflecting wide-
spread use of weather-sensitive devices, the need to include weather effects
in forecasting future requirements will become imperative. It is well docu-
mented that most system peak demands occur as a direct result of seasonal
weather extremes. "Heat storms" and "cold waves" are terms describing inclement
weather that adversely affects the ability of a utility to meet load requirements. Heat
storms are prolonged periods of high temperature, frequently aggravated by abnormal
humidity levels, extending over large areas. The impact of air conditioning and the
occurrence of heat storms is seen in an increasing summer peak and in more reliance
on emergency support from neighboring utilities. In many parts of the United States
the winter peak has usually been the highest peak, but this is no longer the case in
many instances. Cold waves are prolonged periods of abnormally low temperature.

J
('")~~w~~~g n " o ~ t:;
e= ~
~~ s lo.I ~ - ;:l
~ g g
u ;.:...
I;;
~1-'0::S~~~~~
::l O O ..- 0 Cl '"I c:,,
5 Jg <'D 0"'
~ :::<
....
0
e. ::i
'< 0
0
...., 01),)0..0.. o..~ O CT'"'
:::, ; - r-+
C
p., '<::ii:::S-
g- ~
p.. :::s

5 ~o::g.
C7(l
""1 ~
~ ~ e.;...g--g ~ ~
~
:::!". N oo" ~
g 5"' ...... ~
0 0 s- S. o-.- ~.... ..c
o..5.p'i~~~::;:co
::i ~ ..,, ....
C')
o

tT1 .=..
...,..
o
....
i:::
a.~~
.... ...., 0.. - ::i
~ ::i
0'"0.-+C'>c:,,,...O.. - r-+ 0 "'
< ....
~ Jg 0 i::::: ,....: ,__ - <
0 ..- =:; ~. r-+
(1)
rn (t) ~
~ CP ~ ~ ~ ~- ... ~
O "'
en (1) ~- 2. . . ::l - 0 C ..,,
OQs-g.~&~F~
0 ~ ::r' ::i ~ C')
-('t)
C') ~
......
.... ~ c::: a a. CT o ::i ~ o "'
(")
C')
~ ~ p.
......
C. ff~~Jg ;;:g53~n . . n
- 0
~ ::i
3 s-O 0g. r-+ < -"'....
s0..0 -::ti s ~::i
('t) C/)
- 'O
~
O::i-<3C7'lO~;:r'~
~ CJQ s C/) s ::r o "'o..o o..
~::io!)looa&.2
C/)

g- g ;
p:i 0
::i I),)
0 ~- &
(t) - 1-+i (1) ('t) 0 0
'"O
g"' ~ .... ::i < [ ....
~ 0..
.... I),);;?.. C7(l
g 8.. 0 ~
::: ~ a-: p.. ~-
I),) 0
.... "' C7(l "' ::i
i::: '< - 0..
i::: 'O C7(l - a:=-.0 "' ::i
g,.:::,:::,~og
5
C') ('t)

.... "' .... sCl.l (1) .... &.


"' o:=: o
_ < 0 8 3 - g. 3 ~ ~ 0 I),)
I),)

i:::
I),)

0 I),) o.. ~ ::i::


::is,Soo::ingo..
ao
' ::::i . ....,
o' ::i
::i
::i C7(l ~
0
0.. 0
...... I),)
....
::r'
s.t:: f g~ ::i ~~::l'"O!)l8,.~::i
....
0
....
I),)
C7(l
0
::i ...., ::i
"' .... ::i
~ ~ ~(t)e..~S.e.9~<:p ~ i:::
s ~ g, t;; g ....
o s
I),)
~ ~ .... r-+
::r' ::i
OOOQOO
-
C')
r-+
::r' - r-+ ('t) -
- 0 ~ :::s ('") r-+ 1-1 "'"I
::i::ioo..e.o.;"~ ::r' 0 0
0
....,
('t)
-
::i g- e.

\ ~]'',

- ~~--'

TABLE 2.1 Energy consumption of major household appliancesa,b


Est. kWh Est. kWh Est. kWh
Average consumed Average consumed Average consumed
wattage annually wattage annually wattage annually
Food preparation Refrigerator Heating pad 65 10
Blender (12 cu ft) 241 728 Humidifier 177 163
386 15
Broiler 1,436 100 Refrigerator Health and beauty
Carving knife 92 8 (frostless 12 cu ft) 321 1,217
Refrigerator/Freezer Germicidal lamp 20 141
Coffee maker 894 106 Hair dryer 381 14
Deep fryer 1,448 83
(14 cu ft) 326 1,137
(frostless 14 cu ft) 1,829 Heat lamp (infrared) 250 13
Dishwasher 1,201 363 615
Shaver 14 1.8
Egg cooker 516 14 Laundry Sun lamp 279 16
Frying pan 1,196 186 Toothbrush 7 0.5
Hot plate 1,257 90 Clothes dryer 4,856 993
Iron (hand) 1,008 144 Vibrator 40 2
Mixer 127 13
Oven, microwave (only) 1,450 190 Washing machine Home entertainment
Rangec (automatic) 512 103 Radio 71 86
With oven 12,200 1,175 (nonautomatic) 286 76 Radio/record player 109 109
With self-cleaning oven Water heaterc 2,475 4,219
12,200 1,205 Television
Roaster 1,333 ( quick-recovery) 4,474 4,811
205 Black and white
Sandwich grill 1,161 33 Comfort conditioning Tube type 160 350
Toaster 1,146 39 Solid state 55 120
Trash compactor 400 50 Air cleaner 50 216 Color
Waffle iron 1,116 22 Air conditioner (room) 860 860 Tube type 300 660
Waste disposer 445 30 Bed covering 177 147 Solid state 200 440
Dehumidifier 257 377
Food preservation Fan Housewares

Freezerc Attic 370 291 Clock 2 17


Circulating 88 43 Floor polisher 305 15
(15 cu ft) 341 1,195
Rollaway 171 138 Sewing machine 75 11
Freezer
Window 200 170 Vacuum cleaner 630 46
(frostless 15 cu ft) 440 1,761
Heater (portable) 1,322 176

aThe estimated annual kilowatt-hour consumption of the electric appliances listed in this handy reference is based on normal usage. When using these
~ figures for projections, such factors as the size of the specific appliance, the geographical area of use, and individual usage should be taken into consideration.
Please note that the wattages are not additive since all units are normally not in operation at the same time.
bsource: Edison Electric Institute, 90 Park Avenue, New York.
cBased on 1,000 hr of operation per year. This figure will vary widely depending on area and specific size of unit.
22 Power System Planning

Because of the relatively low saturation level of electric heating at this time, cold
waves do not create as serious problems as heat storms. This is not to say that, as cold
weather loads grow, dealing with winter peaks will always be less difficult. The
forecaster and planner must consider both cases and prepare for their effects as best he
or she can.
In addition to the tremendous impact of weather on residential and
commercial loads, other seasonal variations are caused by economic and demo-
graphic effects. In heavily industrialized areas many industrial customers are
characterized by cyclical variations in load requirement due to normal variations in
production level. Demographic effects also play a role in establishing load patterns and
vary with general migration patterns and other factors.
Let it be emphasized that solar energy technology may have a tremendous
impact on load patterns now experienced by electric utilities. It is a well-documented
fact that the use of solar energy technology for heating and air conditioning is well
within the state of the art. If widespread acceptance of solar conditioning occurs, it
will be incumbent on the forecaster to anticipate its impact on the utility industry.

2.2 APPROACHES TO LOAD FORECASTING


In preparing a forecast, the system planner is immediately confronted with the
following basic questions:
1. Should peak demand be forecast using forecasted energy and load factors, or
should it be forecast separately?
2. Should the total forecast be determined by combining forecasts of appropriate
load components, or should the total forecast be directly obtained from
historical total load data?
3. Should average or extreme weather conditions be used?
4. Should simple forecasting methods be used, or should more formal mathe-
matical procedures be investigated?
The answers to these fundamental questions set the direction to be taken by the
forecaster to determine future requirements. It is in our best interest to discuss each
question in some detail. Probably every utility answers these questions differently,
indicating that no one approach will be satisfactory in all cases. Depending on the
intended use of a forecast, the answers to these questions will differ, and rightly so.
For generation planning, for example, details describing minor differences of growth in
geographic areas may not be important; for site selection, however, an obvious
differential in load growth throughout a given service area may have some influence.
It is clear from previous discussions that both energy and demand forecasts are
usually needed. Thus, two options are open to the forecaster. He can forecast
energy-considered to be less difficult-and obtain the demand forecast from it by
using forecasted load factors (the demand forecast can be obtained by dividing the load
factor and time into the energy forecast). Unfortunately, forecasting load factors can
itself be difficult, for system load factors can vary erratically. The second option is to
Load Forecasting 23

forecast peak demand directly, although peak demand also can vary erratically. The
advantages of using energy as the primary forecast and obtaining the peak demand
forecast from it are that energy tends to be much less erratic, is considered a better
trend growth indicator, and is readily related to demographic and economic factors,
using energy data broken down into classes of services, areas, etc. The advantages
claimed for ,separately forecasting the peak demand are that it is a more direct method
and can be related directly to such weather variables as temperature. Because the
utilities have used the two basic approaches successfully, with an average error of
approximately 5%, no attempt will be made here to promote one approach or the
other. The approach is situation dependent.
The next question to be considered by a forecaster is whether to forecast the
total load directly or to assemble the total load forecast by appropriately combining
forecasts of certain components. Typically, the components consist of types of
customers, geographic areas, etc. An advantage claimed for the total load approach is
that it is easier to use and the totals are much smoother and more indicative of overall
growth trends. On the other hand, an advantage of the component approach is that
abnormal conditions in growth trends of a certain component can be detected, thus
preventing misleading forecast conclusions. Again, no one approach is consistently used
in the industry, and both are used successfully by different utilities.
Because weather has such an influence on load requirements (discussed in Sec.
2.1), it is necessary to reflect weather conditions in forecasts. In general, most utilities
construct scatter diagrams correlating certain weather variables, usually coincident
dry-bulb temperatures and humidity, with peak demands. From these data, future
weather-sensitive load components can be obtained by extrapolation. Other utilities use
historical weather data to develop normal probability distributions for weather
variables, which are then used in conjunction with a weather load model to determine
the probability distribution of the weather-sensitive component of future loads. In
most cases some adjustments made for weather conditions are considered necessary for
accurate forecasting. In some special cases additional care must be given to weather
models, particularly if extreme weather conditions are possible. Utilities that consider
extreme weather conditions often determine extreme peaks by using historical
differences between nominal and extreme weather peaks or by adjusting normal peak
forecasts for low-probability weather conditions. Since there is no well-defined
procedure to account for extreme weather. conditions, little will be said of these
matters.
Choosing a forecasting technique to use in establishing future load requirements
is a nontrivial task in itself. Depending on the nature of load variations, one particular
method may be superior to another. Before choosing a particular method, whether it
be simple curve fitting or stochastic modeling, a basic understanding of how a load
behaYes is essential. If, on the basis of historical data and good judgment, simple
extrapolation appears to suffice, it should be used. Unfortunately, since all electric
utilities are somewhat different, it is impossible to list all system attributes that may
be adequately modeled by a particular technique. Choosing the best technique for a
given utility, once again, requires good judgment and a knowledge of the advantages
em Planning

2.3 FORECASTING METHODOLOGY


Forecasting, as stated in the introduction to this chapter, is simply a systematic
procedure for quantitatively defining future loads. Depending on the time period of
interest, a specific forecasting procedure may be classified as a short-term, inter-
mediate, or long-term technique. Because system planning is our basic concern and
because planning for the addition of new generation, transmission, and distribution
facilities must begin 4-10 yr in advance of the actual in-service date, we shall be
concerned with the methodology of intermediate-range forecasting. For simplicity, the
word "forecast" will usually imply an intermediate-range forecast.
Forecasting techniques may be divided into three broad classes. Techniques may
be based on extrapolation or on correlation or on a combination of both. Techniques
may be further classified as either deterministic, probabilistic, or stochastic. We shall
explore briefly the fundamental ideas behind these techniques, with emphasis on
probabilistic extrapolation-correlation methods.

2.3.1 Extrapolation
Extrapolation techniques involve fitti;ttrend curves to basic historical data ~djusted to
reflect the growth trend itself. With a trend curve the forecast is obtained by
evaluati_ng the trend curve function at the crtsYreJ\uture point. Ait'toi.lgh a very simple
prbce'cture, it produces reasonable results in some instan,ces. Such a technique is to be
classified as a deterministic extrapolation, since no at't~fup{ is made to account for
- random errors in the data or in the analytical model. Some eight to ten standard
analytical functions are used in trend curve fitting, including:
1. Straight line Y = a +bx
2. Parabola Y =a+ bx+ cx 2
3. s curve Y = a + bx + cx 2 + dx 3
4. Exponential y = cedx
5. Gompertz Y = ln- 1 (a + cedx)
The most common curve-fitting technique for finding coefficients and exponents (a-d)
of a function in a given forecast is the method of least squares.
, ,. '- If the uncertainty of extrapolated results is to be quantified using statistical
' ~~tities such as mean and variance, the basic technique becomes probabilistic
extrapolation. The uncertainty arises ,from two sources: uncertainty in the historical
data and uncertainty in the analytical model chosen to describe the underlying growth
, in load. With regression analysis the best estimate of the model describing the trend
Load Forecasting 25

can be obtained and used to forecast the trend. We shall discuss this method in great
detail in the remaining sections.
The use of stochastic models to generate a forecast from random inputs derived
from historical data has been investigated, but in 197 5 the use of such techniques was
not widespread. The inputs to the model tend to be the random change in the trend
component, the random slope of the change in seasonal component and associated
weighting factor, and a general noise componenL The statistics of all the random
inputs plus the weighting factor are determined by matching the statistics of the
historical demand data with corresponding statistics for the output of the model. This
step can be performed only after a transformation of the basic model and the
historical data is made, since the data. and untransformed model are nonstationary-
that . is, their statistics are time dependent. With the model matched to fit historical
demand data, the forecast-is obtained by exciting the transformed model by random
inputs whose statistics are known. The resulting time series, after an inverse
transformation, is the forecast desired.

2.3.2 Correlation
Correlation techniques of forecasting relate system loads to various demographic and
economic factors. This approach is advantageous in forcing the forecaster to
understand clearly the interrelationship between load growth patterns and other
measurable factors. The most obvious disadvantage, however, results from the need to
forecast demographic and economic factors, which can be more difficult than
forecasting system load. Typically, such factors as population, employment, building
permits, appliance saturation, business indicators, weather data, and the like are used in
correlation techniques.
No one forecasting method, it must be emphasized, is effective in all situations.
The use of simple curve-fitting techniques is adequate for some utilities and completely
worthless for others. In any case, forecasting techniques must be used as tools to aid
the planner; good judgment and experience can never be completely replaced.

2.4 ENERGY FORECASTING

\ Since energy or sales forecasting at this time is highly subjective, we shall devote only
\\ modest space to the general philosophy prevalent in the industry. Again, every utility
has its own way of arriving at an energy forecast, and the approach we discuss here

i
may or may not agree with one used by a particular utility.

)
Energy forecasts tend to be developed using correlation and extrapolation
primarily, tempered with sound projections of future conditions. Generally, to arrive at
total energy forecast, the forecasts for the three major classes of customers
esidential, commercial, and industrial) are determined and then combined. Each class
~\ forecast separately because of different characteristics associated with these classes.
' , e shall briefly describe the basic philosophy used in determining these three fore-
\:~casts.

roo11315

~
26 Power System Planning

2.4.1 Residential Sales Forecasts


Residential energy requirements are dependent on many things, but the major factors
are
I. Residential customers
2. Population per customer
3. Per capita energy consumption
Clearly, if each factor were known for each week, month, etc., in the forecast period,
then the forecast of residential sales could be obtained simply by multiplying the three
factors. To obtain the forecast values of these three factors, either simple curve-fitting
methods or more sophisticated regression analyses can be used. Such an approach to
residential sales forecasting is usually referred to as the "population" method.
An alternate approach, known as the "synthetic" method, requires a more
detailed look at each customer. For this method the major factors are
I. Saturation level of major appliances
2. Average energy consumption per appliance
3. Residential customers
Once again, if these three factors are extrapolated into the future and multiplied
together, the result is the desired forecast of residential sales. Clearly, the synthetic
method requires a great deal of data collection to define the relative affluence of
customers.
The projection phase of residential forecasting involves determining from
information gathered from customers, builders, etc., how the forecast should be
modified, if at all, to include growth patterns not reflected in the factors considered.
In a locality of rapid growth this is an extremely important step.
2.4.2 Commercial Sales Forecasts
To forecast commercial sales, a host of different approaches may be taken. Perhaps the
simplest is based on the fact that commercial establishments are usually service
oriented; hence growth patterns are usually related closely to growth patterns in
residential sales. Specifically, one method for forecasting future commercial sales is to
extrapolate the ratio of commercial to residential sales into the future and then
multiply this forecast by the residential sales forecast to obtain the desired commercial
sales forecast. Another approach is to extrapolate historical commercial sales, since this
information is frequently available.

2.4.3 Industrial Sales Forecasts


Of all forecasts, obtaining an accurate industrial sales forecast is by far the most
difficult. Industrial sales are tied very closely to the overall economy, and we all know
how unpredictable the economy can be over selected periods. Even in view of these
difficulties, many approaches have been and are used by forecasters to develop
industrial sales forecasts. Two such approaches are
Load Forecasting 27

1. Multiply forecasted production levels by forecasted energy consumption per unit


of production.
2. Multiply forecasted number of industrial workers by forecasted energy consump-
tion per worker.

Whether these particular approaches work probably depends on the type and
location of the industry involved. In fact, in many instances the historical industrial
sales are decomposed into subclasses to facilitate an accurate forecast of each
component. This step is particularly important if a utility serves a broad spectrum of
types of industries. As in residential sales forecasting, the final step of tuning the
forecast is important. It requires projecting the way in which the forecast obtained
from historical data may be in error.
Even though we have decided not to describe concrete analytical methods for
energy or sales forecasting, the reader should not interpret this decision to mean that
sales forecasting is unimportant. No doubt an entire text might be devoted to this
topic alone. Indeed, accurate energy forecasts are essential for good planning, but
lacking a well-defined analytical approach accepted by the industry, we have presented
here only broad concepts. We shall proceed to a discussion of peak demand forecasting.

2.5 PEAK DEMAND FORECASTING


Unlike sales forecasting methods, some rather well-defined analytical methods have
been developed over the last 5 yr to aid the forecaster in determining a reasonably
accurate peak demand forecast. We shall discuss in Secs. 2.5 through 2.9 the details of
a representative method. In particular, we shall choose an approach to peak demand
forecasting that involves extrapolating historical demand data instead of calculating
forecasted sales and load factors, primarily because this approach permits weather
conditions to be included with ease in the forecast. In addition, we shall concentrate
on weekly peak demand forecasting as opposed to monthly or annual forecasting. The
difference is in the amount of historical data available, that is, the data-sampling rate.
Given weekly peak demand forecasts, we shall find it straightforward to obtain
monthly or annual forecasts.
The basic approach to be taken is
1. Determine seasonal weather load model.
2. Separate historical weather-sensitive and non-weather-sensitive components of
weekly peak demand, using weather load model.
3. Forecast mean and variance of non-weather-sensitive component of demand.
4. Extrapolate weather load model, and forecast mean and variance of
weather-sensitive component.
5. Determine mean, variance, and density function of total weekly forecast.
6. Calculate density function of monthly or annual forecast.
For our purposes we shall assume that the seasonal variations of the peak
demand are primarily due to weather. If not, then before step 3 can be undertaken,
28 Power System Planning

any additional seasonal variation remaining after weather-sensitive variations are


removed must also be removed. Several methods remove such seasonal variations; the
method proposed by Shiskin (6] has been successfully employed. To use the
forecasting method proposed, a data base of at least 12 yr is recommended. Further,
to develop the weather load models, daily peaks and coincident weather variable values
are needed.
2.5.1 Weather Load Model
To determine the weather load model based on historical data, it is common to plot a
scatter diagram of daily peaks versus an appropriate weather variable or variables.
Although there are many variables to choose from, dry-bulb temperature and humidity
are considered important. For simplicity we use only dry-bulb temperature as the
weather variable. Thus, the scatter diagram from which the load model is obtained will
be as shown in Fig. 2.1. Using linear regression analysis or curve-fitting techniques,
three straight line segments can be defined and used to describe the model; that is,
w = ks(T- Ts) if T> Ts (2.1)
= -kw(T- Tw) if T< Tw (2.2)
=0 if Tw ~ T ~ Ts (2.3)
The parameters of the model are the slopes ks and kw and the threshold temperatures
Ts and Tw

w=-kw (T- Twl

*/

** *
*

Temperature
Fig. 2.1 Weather load model.
Load Forecasting 29

Fig. 2.2 Three-bus sample system.

2.5.2 Separating Weather-Sensitive and


Non-Weather-Sensitive Components
With the weather load model known, the weather-sensitive (WS) component of weekly
peak demand data can be calculated knowing the weekly peak coincident dry-bulb
temperatures, and this component can be subtracted from the historical data to obtain
the non-weather-sensitive (NWS) component of peak demand. The NWS component is
used in step 3 to forecast the mean and the variance of the NWS component of future
weekly peak demands. For future examples, assume that Fig. 2.3a represents the NWS
component of historical peak demands for the three-bus system shown in Fig. 2.2.
Since the methods used to forecast the two components are very detailed, we
shall devote Sec. 2.6 and 2.7 to these important topics.

2.6 NON-WEATHER-SENSITIVE FORECAST (NWSF)


To obtain the NWSF, we must utilize a technique called discounted multiple regression
(DMR) to fit a time polynomial to the NWS component of historical weekly peak
demands. Once the time polynomial has been determined, it may be evaluated for each
week in the forecast period. Since most standard time polynomials do not fit all
weekly peak demand values, NWS weekly peak demands must be described by a
deterministic time polynomial plus a random variable. Fitting a time polynomial to the
NWS component using a DMR technique basically involves finding the coefficients of a
set of so-called "fitting functions" that best fit the NWS data. Fitting functions are
time functions that describe the underlying variation of the NWS component with
time. For example, the curve in Fig. 2.3a may represent 2 yr of NWS peak demand
data for our sample system. Clearly, the underlying variation is best described by
~(t) = a1 + a2t = a1f1(t) + a2f2(t)
30 Power System Planning

56.000

49.600
~
43.200 '
. If V
,
,
}. ,pf"'
.}j/ I!,,,
36.800

s:~
.,,,,-
30.400

_t
~"'J ~
"'
l1i_ 24.000
I,
'---
["-\ I $=a
>
:;,
QJ
QJ
17.600
~;)
.. #=b
'---
s: * =c

11.200
; :. ~
v.~ '

.~-
~~"
J A .
4.8000
,t1t_.\ .. a;~ " .. ~
,... .
. . .. ..
#,.
" ,, """"
*
...... .JI:;
*
-1.6000

-8.0000
' .

"
"'
. .....
"



..
'
"
.. .
;

'I .' I'

0.0 20.8 41.6 62.4 83.2 104.


10.4 31.2 52.0 72.8 93.6
Weeks

Fig. 2.3 (a) NWS peak demand data. (b) Trend component of NWS peak demand data. (c)
Gaussian noise component of NWS peak demand data.

where only two fitting functions in this case are required to describe the peak demand
variations:
f1(t) = 1.0
f2(t) = t
The DMR technique will enable us to calculate the values for a 1 and a2.
Knowing these coefficients, we can use Ht) to forecast the NWS component of peak
demand at some future week.

2.6.1 Discounted Multiple Regression


We may describe historical weekly peak demand data l(t) as
= Ht) + 71(t)
t(t)
= a'f(t) + 71(t) (2.4)
where a' = 1 n coefficient vector
f(t) = n 1 fitting function vector
Load Forecasting 31

and ri(t) is a random variable that describes the random variation of l(t) about Ht). It
is necessary to assume that ri(t) is a zero mean, Gaussian, white noise, random variable,
implying that
e(ri(t)) = fi(t) = 0 property 1 (2.5)

f(ri(t))= a
11
~ exp[- (T/~}) 2

] property 2 (2.6)

e(ri(t) ri(t + r)) = 0 T -=I= 0 property 3


= a11 2 T =0 (2.7)
Interpreting these three properties: Property 1 indicates that if we sum ri(t) over
all time, its average will be zero. For example, if we added all the points in Fig. 2.3c,
which is the random noise added to Fig. 2.3b to get Fig. 2.3a, and divided by the
number of points, then we would find its average close to zero. It is not exactly zero,
since we added only 104 points instead of an infinite number. Property 2 tells us that
the noise component is Gaussian distributed (bell-shaped) about its mean, as Fig. 2.4
shows. This probability density describes the probability that ri(t) will be in a range
ri(t) to ri(t) + dri(t). For example, the probability that ri(t) lies in a range from - 00 to
00 is

F(00 ) = f -oo
00

f(ri(t))dri(t) = 1.0

In other words, we are certain that ri(t) will have a value in the range - 00 to 00 ;
however, based on property 1, its average value is zero. Property 3 says that knowing
the noise at time t does not tell us what it will be at time t + r; it is serially
uncorrelated. However, for T = 0 we would interpret property 3 as the degree of
spread of the probability density function, where a11 2 is the variance. large variances
indicate that the noise varies widely about its mean. If the variance is zero, we are

.,.f"
>
+-
;;;
C
Cl>
"tJ
?;-
:.0
~
...
0
c..

Tl (t)
Fig. 2.4 Noise probability density function.
32 Power System Planning

certain that Tl= fJ. The standard deviation, Orp is a most useful parameter because it
defines the 99% confidence interval; that is, we can be 99% confident that a
Gaussian-distributed random variable lies in a range 2.57a71 about its mean. To
illustrate, consider Tl(t) shown in Fig. 2.3c again. It has a variance of 10 or a standard
deviation of 3.16, which can be obtained approximately by computing the average
squared value of T1(t) about fJ(t). This task is 'left to the student. The 99% confidence
interval in this case, then, is -8.12 to 8.12. Looking at Fig. 2.3c, we clearly see that
T1(t) does not fall out of this range. Further we see that most points are grouped
within a standard deviation of the mean.
In general, we want to find values of the coefficients on the fitting functions
that best describe the random weekly peak historical data. There are many ways of
defining a best fit, but we define the best fit to correspond to that value of a' that
minimizes the squared difference between the measured weekly peak demand l(t) and
the true peak demand computed using ~(t). We use the squared difference to give
negative errors-resulting when Ht) is greater than /(t)-the same weight as positive
errors. Clearly this procedure makes sense. Further, we associate with every l(t), Ht)
pair a weighting factor W(t) that will allow us to place more weight on newer data.
The function to be minimized thus becomes:

N
J = '\' W2 (t)[l(t) - Ht)]2
'-'
t= 1 (2.8)

W(t) = -vffN- t
where t is a discrete variable starting with week 1 and continuing to the last week N of
the available historical data.
Stated precisely, we want to find those values of a 1 and a2 that minimize Eq.
2.8. In carrying out the minimization procedure it is advantageous to rely on the
compactness of matrix notation and to express Eq. 2.8 in matrix form.
From Eq. 2.4 we see that for a given t

~(t) = (a 1 a2 ) fi(t)] = a'f()


t
Lflt)
l(t) = l(t)
W(t) = W(t)
From Eq. 2.8
J = W 2 (l)[l(l) - a'f(l)J 2 + W 2 (2)[/(2) - a'f(2)J 2 +
which, with some work, can be expressed as
J = [L(N) - a'R(N)JW(N)W'(N)[L(N) - a'R(N)J' (2.9)
Load Forecasting 33

where

L(N) = 1 N vector whose elements are l(t)


a' = 1 n fitting function coefficient vector
R(N) =,n N fitting function matrix whose columns aref(t)

Now we are ready to find the vector a' that minimizes Eq. 2.9. The minimization
process involves determining the value of a' that results in the first partial derivative of
Eq. 2.9 being equal to zero. Taking the first partial,

0
of= .,, [L(N) - a'R(N)]W(N)W'(N)[L(N) - a'R(N)]'
aa ua

= :a [L(N)W(N)W' (N)L' (N) - a'R(N)W(N)W' (N)L' (N)

- L(N)W(N)W'(N)R'(N)a + a'R(N)W(N)W'(N)R'(N)a] (2.10)

Note that a'R(N)W(N)W' (N)L' (N) = 1 X 1 = scalar, and the transpose of a scalar is a
scalar. Therefore

a'R(N)W(N)W'(N)L'(N) = L(N)W(N)W'(N)R'(N)a
and

:: = -L(N)W(N)W'(N)R'(N) + a'R(N)W(N)W'(N)R'(N) =0
which gives the best estimate:
a'(N) = [L(N)W(N)W'(N)R'(N)] [R(N)W(N)W'(N)R'(N)]- 1 (2.11)
Letting
F(N) = R(N)W(N)W'(N)R'(N) (2.12)
G(N) = L(N)W(N)W'(N)R'(N) (2.13)
we have
a'(N) = G(N)F(Nf 1 (2.14)
To find a'(N) computationally, we must process sequentially the historical NWS weekly
peak demand data to determine G(N) and F(N)- 1 as indicated by Eq. 2.14. From Eq.
2.12 with W(N) equal to an N X N diagonal matrix where the (N - t, N- t) diagonal
element equals v[JN-t (/3 is called the "discount factor" and is less than unity), we get
for the ikth element of F
34 Power System Planning

N
F;k(N) = l
t=1
1
[;(t)fiN- /k(t)

N
= 2/;(t)fiN-I-tfifk(t)
t=1
N-1
= /;(N)fk(N) + fi l
t=I
/;(t)fiN-l-tfk(t)

but

l
N-1

t= 1
[;(t)fiN-I - 1/k(t) = F;k(N - 1)

Thus we get a recursive relationship for the ikth element of F:


(2.15)
Using Eq. 2.15, we can easily process each historical weekly peak value to obtain
F;k(N), F(N), and finally F(N)- 1 The initial values of F;k, F;k(l), can be obtained
from Eq. 2.15.
(2.16)
For example, the initial values of F;k for fitting functions describing the sample
three-bus system are:
F 11 (I)= //(I)= 1.0
F12(l) =/1(1)/2(1) = 1.0
F21(l) =/2(1)/1(1) = 1.0
F22 (1) = f/(1) = 1.0
To build G(N) sequentially, we need a recursive formula similar to Eq. 2.15. We
easily obtain the recursive formula from Eq. 2.13, where we get for the nth element ofG
N

Gn(N) = L, l(t)fiN- 1
fn(t)
t=l
N
= L, f(t)fiN-l-tfifn(t)
t=I
N-1

= l(N)fn(N)+fi l
t=1
l(t){3N-J-tfn(t)
Load Forecasting 35

but

L
N-1 .
l(t)(3N-l-t(3fn(t) = f3Gn(N-1)
t=1

Thus we obtain the needed recursive formula:


(2.17)
Again, using Eq. 2.17, we can process the historical data sequentially to obtain
G(N). The initial values of Gn, Gn(l), can be obtained from Eq. 2.17.
Gn(l) = l(l)fn(l)
For the fitting functions for the sample system, we have the following initial values for G:
G 1 (1)= /(1)
G2 (1) = /(1)
In summary, to calculate the coefficients for a set of fitting functions that
describe the underlying variation of a set of NWS historical data points, we need only
process the historical data using the following recursive formulas:
Fik(N)= fi(N)fk(N) + f3Fik(,N - 1) (2.18)
Gn(N) = l(N)fn(N) + f3Gn(N - 1) (2.19)
a'(N) = G(N)F(Nr 1 (2.20)
F 1k(N) is the ikth element of F given historical data up to time N, and similarly Gn(N)
is the nth element of G given data up to time N.
Applying these equations to the NWS historical data shown in Fig. 2.3a, we get
a a
1 = 0.2326 and 2 = 0.510, giving us the following equation for Ht):

Ht)= o.2326 + o.stor


This last equation would be used, for instance, to forecast the NWS peak demand for
our three-bus system. In the general case we should of course use
i(t) = a'(N)f(t) (2.21)

2.6.2 Mean and Variance of Coefficients


The variance of aiCN) is an important quantity in that it quantifies the range over
which the elements of a(N) vary as data are sequentially processed.
To determine the coefficient variances, which are the diagonal elements of the
covariance matrix, we start with the standard definition of covariance:
cov(a(N)) le e([a(N) - ii(N)] [a(N) - ii (N)]') (2.22)
-
where a(N) is the mean value of a(N) given by
36 Power System Planning

fi'(N) = E(a'(N)) = E(G(N)F(Nf 1 )

But from Eqs. 2.12 and 2.13 we get

~'(N) = E(L(N)W(N)W'(N)R'(N))F(Nf 1
Realizing that W(N)W'(N)R'(N)F(N)- 1 .is'a deterministic quantity,
li'(N) = [E(L(N))] [W(N)W (N)R (N)F(Nf 1 ] 1 1

= [E(t(N) + 11(N)] [W(N)W (N)R'(N)F(Nr 1 ] 1

= ~(N)W(N)W'(N)R'(N)F(Nf 1 (2.23)
since E(77(N)) = 0. Therefore

fi(N) = F(Nr 1 R(N)W(N)W (N)f(N) 1

= a(N) (2.24)
Expanding Eq. 2.22,
cov(a(N)) = E(ii(N)ii'(N) - fi(N)ii'(N) - ii(N)1'(N) + ~(N)~'(N))

and substituting Eqs. 2.23, 2.24, and 2.20, we get

cov(a (N)) = a11 2 F- 1 (N)R(N)W(N)W (N)W(N)W (N)R (N)F(Nr 1 1 1 1


(2.25)

where a 11 2 is the variance of random component 77(t) of the NWS peak demand data.
As the variance of the noise is generally not known in advance, it can be shown
that an unbiased estimate of the variance a11 2 is given by
N N
a 11 2
= (N - n)- 1
l 2
r (t) or N
1
n l r2(t) (2.26)
t == 1 t == 1

where r(t) is the residual l(t) -ii'(N)f(t). For the sample system a11 2 = 10, therefore
= 0.3902 , var(82 ) = 0.0001
var(8 1 )
Clearly, the value of a1 varies a greatdeal about its mean as data are processed,
telling us not to have much confidence in a1 or, stated differently, not to use Ht) to
obtain the forecast demand for small values of t. This restriction does not bother us,
because we are interested only in using t(t) to forecast the NWS peak demand for the
sample system for times greater than t = 104. Similarly, since the variance of a2 is
0.0001, we can feel confident that a2 = 0.510 is a good estimate for this parameter.
Because the NWS peak demand data were artificially generated by adding 1/(_t) to
Ht)= 1 + 0.5t, we know that the conclusions above are correct; that is, a1 = 0.2326 is
not a good estimate of 1.0, but a2 = 0.510 is a good estimate of 0.5.
2.6.3 Forecast Variance
Forecast variance is an important quantity in that, as we now know, it quantifies the
uncertainty of the forecast. To get the forecast variance, we must start with

------------------
Load Forecasting 37

e((t)) = e(a'(N)f(t))
= e(a'(N))f(t)
or
{(t) -:- i'(N)f(t) (2.27)
Using Eq. 2.27, the variance of the best estimate of our forecast peak demand is easily
obtained as follows:
var({(t)) = e([{(t) - (t)] [(t) - {(t)])
= e([(t) i'(N)f(t)]'{(t) ft'(N)f(t)])
= r (t)e([il;(N) - i(N)] [a(N) - i(N)]')f(t)
= f(t)cov(a(N))f(t)
= al (2.28)
If we forecast the NWS component of peak load using the weekly peak load data
in Fig. 2.3a, we get, with {3 = I, the curve shown in Fig. 2.5. The curve in Fig. 2.6 is a
plot of the variance of forecasted NWS weekly peaks. Note that the variances increase
with time, reflecting the fact that the more futuristic the forecast, the less reliable it is,
as one would expect.
To conclude, we now have the basic understanding required to determine the
mean and variance of the NWS component of peak demand. Equations 2.18 through

80.26 -------------

77.18 - - - - - - - - - - -

54.10

105 150 156


t, wk

Fig. 2.5 Sample system NWS forecast (52 wk).


38 Power System Planning

1.239 -------------

1.110 - - - - - - - - - - - - - -

N
<=
b

0.39

105
T 5 150 156
t, wk

Fig. 2.6 NWS forecast variance (52 wk).

2.20 define the sequential algorithm to obtain the coefficients for f(t), and Eqs. 2.21
and 2.28 provide expressions for the forecast and forecast variance of any future week.

2.7 WEATHER-SENSITIVE FORECAST (WSF)


From Sec. 2.5 we have the form of the weather load model as defined by Eqs. 2.1
through 2.3. To forecast future WS weekly peak demands, we need
1. A weather load model that is valid during the time period for which we are
trying to forecast weekly peak demand, and
2. Information that will sufficiently describe our weather variable during a given
future week so that corresponding WS peak loads can be obtained.
Obtaining the weather load model that is valid during a given future week
involves somehow determining kw, ks, D 0 , Tw, and Ts that apply for that week.
Usually, a given set of these parameters is valid for much longer periods, such as a
season, since the growth in heating and air conditioning load varies an insignificant
amount on a daily or weekly basis. In essence this problem is a forecasting problem;
however, it is generally understood that extrapolating values of kw and ks can be
dangerous, because of saturation effects, different heating load and air conditioning
load growth rates, etc. In most cases, obtaining future values of these parameters is
accomplished using extrapolation tempered with good judgment.
Load Forecasting 39

2.7.1 Weather Variable


Knowing the future weather load models, it must be determined what the dry-bulb
temperature will be at the time of the forecast weekly peak. Unfortunately it is not
possible to say that during the 25th week 5 yr from now, the coincident dry-bulb
temperature, will be 85F. The best we can do is rely on historical data for this
information; that is, if we plot the historical daily coincident dry-bulb temperature for
a given week, we will find a minimum, maximum, and average temperature for that
week. Going one step further, using a standard numerical algorithm, we obtain an
approximate statistical description of the temperature data. If we assume that
temperature values are Gaussian distributed, then only mean and variance are required
to describe the random temperature variation. An idealized weather variable probability
density function is shown in Fig. 2.7. Normally it would be obtained, for example,
from a plot of historical daily peak coincident dry-bulb temperatures, as shown in Fig.
2.8 for a particular week for the last 15 yr. Since this curve was artifically generated
from a Gaussian random number generator to support our sample study, we know that

5.00000E-02

4.50000E-02

j
4.00000E-02

3.50000E-02
.
'\ft

-~
.n
3.00000E-02

co 2.50000E-02
; ~
.n
0
~

c..
.
2.00000E-02

1.50000E-02
IJ
:\
1.00000E-02
} .

5.00001 E-03
f-- ~
\
\ ~
3. 72529E-09
0.0 13.6 27.2 40.8 54.5 68.1
6.81 20.4 34.0 47.7 61.3
Temperature, F

Fig. 2.7 Winter temperature probability density function.


40 Power System Planning

69.000

66.300

63.600
. . ...
., If

lL
0

.."'...- 60.900
. -
. . -
......
:,
. ., . ... ~

... .
"'
"'Q.
E
58.200
..... . .. ~

......"' -- If
" - ,.c
C:
"'
55.500
<I
" ...
..
t ....
"tJ
c:; " .
ou
C:
52.800 . f
- .
>I-
. . .,
.?-
'iii
0 50.100

47.400
. .
,. .. ~ ~

.. ~.
"
44.700
. 1' " "
. .,.
42.000
0.0 21.0 42.0 63.0 84.0 105.
10.5 31.5 52.5 73.5 94.5
Days

Fig. 2.8 Winter peak temperature data (15 yr).

its mean is 55F and its variance is 25F or a standard deviation of 5F. Since our
weather variable is assumed Gaussian distributed, its probability density function is:

f(T, 55 , 5 ) o o 1
= a-Jfii exp - 1(T- 7')
2 a
2
(2.29)

where a= 5F and T= 55F.


It is instructive to compute the mean and variance of the sample historical daily
coincident dry-bulb temperature from data points in Fig. 2.8, using standard
definitions of sample mean and variance given by the following equations:
N
- 1 \"'
T= NL Ti (2.30)
i= 1
N
a
2
= Nl\"'
L (Tt- T)
-2
(2.31)
i= 1

--------~-----
Load Forecasting 41

The application of these two equations to the data in Fig. 2.8 is left to the
student; the results obtained for f'
and a2 will be close to 55 and 25F, respectively.
Since N is only 105, the exact values for f and a2 will not be obtained.
We are ready now to consider the actual WS forecasting phase since we have a
weather load model that is valid for the period in which we look for a weekly peak.
Further, we 'have a statistical description of our weather variable for the weeks in
question, as depicted by the Gaussian distribution in Fig. 2.7.
To obtain the weekly WS mean peak demand, variance, and density function, the
following steps are required:
1. The weather variable and associated probability density must be transformed
so that the new variable has a zero mean and unity variance (called a normal
distribution). As we shall see, this step greatly simplifies the problem of
obtaining the weekly WS peak demand mean and variance.
2. The weather load model must be expressed in terms of the new variable.
3. The weekly peak mean and variance must be determined.
4. The weekly peak demand probability density function must also be obtained.
Since weekly peak forecasts are required for perhaps 10 yr into the future, the
steps above must be repeated for each week, using appropriate weather load model and
weather variable distributions.

2.7.2 Weather Variable Transformation

To facilitate obtaining the weekly mean peak and variance for a future week, the following
transformation equation should be used, as suggested by Stanton and Gupta [9] :
T-f (2.32)
x=
Or
This transformation defines a zero mean, unity variance Gaussian process. To verify
this, we need only determine x and ax 2 Using basic definitions

x = e(x) = e(T;rf)
_ e(T) - T _ f - f =O
ar ar
and
a/ = e [(x - x)2] = e(x 2 )

= E
T-
'---'--
ar2
f')2] = ar2
-1 E
[(T - T)
- 2)
~
but
42 Power System Planning

therefore

With the transformation defined as in Eq. 2.32, all weather variable distributions
'
may be transformed into normal distributions, which will improve computational
efficiency. Given the normal distribution in Fig. 2.9, we may divide it into 99
segments, each having an area of 1/100. If each segment is defined by the variables
rn-l and rn, where n = 2, ... , 100 and the range of rn is from r 1 = -2.575 to
r 100 = 2.575, then the density function /(_x) between rn-l and rn may be expressed as
1
f(x) -
(2.33)
rn-1<x<rn
For convenience, the values of r;, i = l ... 100, are given in Table 2.2, where two
consecutive entries have the property that
1 .
f(r;-1 < x < r;) = 100
for all z

which is obvious from the way the normal distribution for x in Fig. 2.9 was
segmented.

2.7.J WS Peak Demand Mean and Variance


If we express the weather load model in terms of our new variable x, we get the
following. From Eq. 2.1
W = kw(Tw - T)
But from Eq. 2.32
T=xaT+ T

1
Area=
100

0.5
Area=
100

r, = -2.575 r 49 = -0.0125 r 50 = 0.0125 r 100 = 2.575


X ->

Fig. 2.9 Normal probability density f(_x).


TABLE 2.2 Tabulated normal
distribution
i r1< 0 i ri> 0
1 -2.57583 51 +0.01253
2 -2.17009 52 +0.03761
3 -1.95996 53 +0.06271
4 -1.81,191 54 +0.08784
5 -1.69540 55 +0.11304
6 -1.59819 56 +0.13830
7 -1.51410 57 +0.16366
8 -1.43953 58 +0.18912
9 -1.37220 59 +0.21470
10 -1.31058 60 +0.24043
11 -1.25357 61 +0.26631
12 -1.20036 62 +0.29237
13 -1.15035 63 +0.31864
14 -1.10306 64 +0.34513
15 -1.05812 65 +0.37186
16 -1.01522 66 +0.39886
17 -0.97411 67 +0.42615
18 -0.93458 68 +0.45376
19 -0.89647 69 +0.48173
20 -0.85962 70 +0.51007
21 -0.82390 71 +0.53884
22 -0.78919 72 +0.56805
23 -0.75542 73 +0.59776
24 -0.72248 74 +0.62801
25 -0.69031 75 +0.65884
26 -0.65884 76 +0.69031
27 -0.62801 77 +0.72248
28 -0.59776 78 +0.75542
29 -0.56805 79 +0.78919
30 -0.53884 80 +0.82390
31 -0.51007 81 +0.85962
32 -0.48173 82 +0.89647
33 -0.45376 83 +0.93458
34 -0.42615 84 +0.97411
35 -0.39886 85 +1.01522
36 -0.37186 86 +1.05812
37 -0.34513 87 +1.10306
38 -0.31864 88 +1.15035
39 -0.29237 89 +1.20036
40 -0.26631 90 +1.25357
41 -0.24043 91 +1.31058
42 -0.21470 92 +1.37220
43 -0.18912 93 +1.43953
44 -0.16366 94 +1.51410
45 -0.13830 95 +1.59819
46 -0.11304 96 +1.69540
47 -0.08784 97 +1.81191
48 -0.06271 98 +l.95996
49 -0.03761 99 +2.17009
50 -0.01253 100 +2.57583

43
44 Power System Planning
t
Therefore {
w = kw(Tw - xor - T) T < Tw (2.34) !

and

w=O TwTTs
To determine the mean and variance of w for any general temperature
distribution, three distinct intervals must be dealt with: the intervals for T w T Ts,
T < T w, and T> Ts In terms of the new variable x these intervals are
Tw-T
- Ts-T
--x _ __
or or
-'t
T_w_ _
X<- Or =xw (2.35)

-t
T __
x>-s =Xs
Or
Using Eqs. 2.33 to 2.3 5 and the standard definition of mean and variance, we may
easily obtain expressions for wand ow2 similar to those derived by Stanton and Gupta [9] .

<{w)- W - f- 00
w(Tw) wf(w)dw + f w(T,) wf(w)dw

w(rw)
+ J"
Ws(rs)
l-lf(w) dw (2.36)

Since w is a function of x, and xis always either increasing or decreasing depending on


the interval, it can be shown that
f(_w)dw = f(_x)dx
and Eq. 2.36 becomes:

-f X3
00

ks(Ts - xar - T)f(x) dx (2.37)

Consider the first term in detail.


Load Forecasting 45

xw Jxw
f-00
kw(Tw - xar - T)f(x) dx =
-00
- kwarxf(x)dx

Xw

+
!
-oo
kw(Tw - T)f(x) dx (2.38)

where

J
++
frw1
rw2 1
------xdx
100(rw2 - 'w1)

rw 2 and rw 1 are the values of ri whose range contains Xw- Upon integrating and
simplifying the last expression,

(2.39)

The second term in Eq. 2.38 may be similarly evaluated.

kw(Tw - T)
f-=
xw
f(x) dx = kw(Tw - T) I
[ r
I
2

lOO(,: _ ,
1
) dx

This term, upon integrating and simplifying, becomes:

kw(Tw -T) J Xw

-oo
f(x)dx = kw(Tw -T) ~~ (2.40)

Thus, for the first term in Eq. 2.37 we have

J
1
Xw
wf(x) dx =-
arkw (
200
r1 + rw2 + _l
w
2ri
)

+
kw(Tw - T)w1
100
-
(2.41)
-oo l =2

For the second term of Eq. 2.37 we get zero, and for the third we get, by the same
procedure used on the first term,
46

-
Power System Planning

~
-
IXs
ks(Ts - xor - T)f(x)dx =
0
;;; frs 1 + r100 +
\
.?
1-s2
2r;)

-ks(Ts - T)(IO0 - S 1 )
100 ( 2 .4 2)
where rs 1 and rs1, are the values of x whose range contains Xs,
Although the expression for the variance of WS peak demand can be obtained in
the same manner, starting from the definition of variance, the derivation is left as an
exercise. For future use, however, let us summarize the results:

ks(Ts - T)(l00 - S1)


100 (2 .4 3)

+
k/(T- Ts)2(100 - Si) . k/or(T- Ts)(.
100 + lOO \sl + r100 + ~ 2r; i~,
) _2
- w (2.44)

Table 2.2 provides the values for r;, rs;, and rwi When these values are
substituted into Eqs. 2.43 and 2.44 along with the weather load model and
temperature parameters, the desired WS peak demand mean and variance can be
determined.
To illustrate the use of mean and variance equations, consider determining the
mean of WS peak demand for a future week in which the weather load model and
temperature distribution are as shown in Fig. 2.10. For this week then,

kw= 0.5 Ts= 80


ks= 1.0 T= 10
Tw = 60 OT= 5
Load Forecasting 47

-0
"'0
.Y.
"'
Q)
0.
(/)
$'.

Tw=60 f T5 = 80
Temperature, F

Fig. 2.10 Weather load model and associated temperature density


function.

and
T -T
Xs = s
ay
= 2.0
T -T
Xw = w = -2.0
ar
Since r wi and r w 2 were defined as values of x whose range contains Xw, it is clear
from Table 2.2 that
rw 1 = -2.170 rw2 = -1.959
and W1 = 2, w 2 = 3. Similarly,
r5 1 = 1.959 rs2 - 2.170
and s 1 = 98, s 2 = 99. Substituting these values in Eq. 2.43, we get

w = ;()~ [2.575 + 1.959 + 2(2.170)]


5 5 10
+ 200 [1.959 + 2.575 + 2(2.170)] - 100(2) - 100(2)
= 0.203 MW
Since the temperature range fell mostly between the two threshold temperatures Ts
and T w, we expected the mean WS component to be quite small, as indeed it is.
48 Power System Planning

Now that we have discussed in depth how to calculate the means and the
variances of WS and NWS components of peak demand, to obtain the total forecast mean
and variance for a particular week, we need only add component means and variances.
In weekly forecasts (forecasts based on weekly data), however, the variances are of
little value in establishing confidence intervajs, because the density funcEon of the
total forecast is skewed. This feature of the total forecast density function is a direct
result of the weather load model being inherently nonlinear. To circumvent this
problem and to provide the system planner with meaningful results, it is necessary to
determine the total forecast probability density function in place of the variance.
If the variance of the NWS component is several times larger than that of the WS
component-usually the case with seasonal and annual forecasts, where data are
sampled on seasonal and annual bases-then the total forecast density function
becomes approximately Gaussian. Under these conditions the total forecast mean and
variance completely describe the statistics of the total forecast, and knowing the
variance is sufficient to calculate confidence intervals.

2.7 .4 Weekly WS Peak Load Probability


Density Function
Given that the WS component has been forecast on the basis of weekly historical data,
the nonlinearity of the weather load model coupled with the large variance in the
weather variable results in a truncated and skewed density function, as shown in Fig.
2.11. The density functions shown in Fig. 2.11 l,ave three distinct segments. The first
segment is a delta function at the origin, whose magnitude is the probability that
w = 0. The second segment is the probability density function associated with values
of w > 0 due to temperatures in excess of Ts The third and last segment is the
probability density function associated with the values of w > 0 due to temperatures
less than T w.
To obtain the WS peak load density function, segments of Gaussian distributions
must be combined with a delta function occurring at the origin. Figure 2.11, for
example, consists of a delta function and portions of two Gaussian distributions, where
the magnitude of the delta function is the probability that w = 0. For a given
temperature distribution, that probability coincides with the probability that
T w ..;; T..;; Ts. Since the temperatures are assumed Gaussian distributed with a mean and
variance of T and or, respectively,

F(Tw, f, or)=
f -~
rw 1
orffri
t1 T-
exp - -
2(
T
or )
J
2

dT

1 1 ~ (-l)"(Tw-f')2n+l
= 2 + v'rr L, (2n + I)v22n+loln+ln!
n=O
= probability that T..;; T w
Load Forecasting 49

w,MW
(a)

T< Tw

w,MW
(b)

Fig. 2.11 (a and b) Two sample WS peak load probability density functions.

and
00

1t(Ts - T) 2 n + 1
-
F(Ts, T, or)=
1
2+ yn
1

n=O
l( (-

2n + l)v'22n+l O
T
2n+

= probability that T ~ Ts

Thus, the probability density function for the first segment is the difference

1 (-It[(Ts - f)2n+l - (Tw - f)2n+l]


f1(w=O)=o(w)yn
n=O
l
oo

(
2n+I)y'22n+1 0 2n+1n!
T
(2.45)

where the delta function multiplier, cS(w), is used to indicate that w can assume only
one value, namely w = 0, in the first segment of the WS peak load density function.
Because w and T are linearly related for values of Tin excess of and less than Ts
and T w, respectively, the probability density functions for segments 2 and 3 defined
for values of w > 0 are:
50 Power System Planning
l
l
j
(2.46)
l
l'
and l
!
l

h(w, w, aw)= (w) k \ (2.47) i


wOT 12 rr
The step function (w) indicates that both [ 2 and [ 3 are nonzero only for values of
w > 0. The proof of Eqs. (2.46) and (2.47) will not be presented; suffice it to say that
since the weather load model is deterministic and a linear function of T in these
segments, the statistics of w will also be piecewise Gaussian.
Adding the density functions for the three segments, we obtain the density
function for the WS component of peak load.

f(w, w, Ow) = (w) ( 1


ksorvfii exp - 2 ksoT {[
1 w - ks(T - Ts) - ]2}
1
+----exp --
kwor.../iii 2 kwoT {~
1 w - kw(Tw - T)
-~2})
+ oW _1_ \' (-It[(Ts - f)2n+l -(Tw - f)2n+l]
( ) ../ii ~ n
(2n + 1)y'22n +1 o/n +1 n !
(2.48)

It is instructive to consider a simple example involving the determination of the


\VS peak load density function. For this purpose we choose the weather load model and
associated temperature distribution shown in Fig. 2.10, except

Substituting these relationships into Eq. 2.48, we obtain


2
_ 2 1(w+kor) ]
[ 2 \' (-It
f(w, w, Ow) = (w) kor../5:rr exp - 2 kor + o(w) ../ii f (2n + 1)n!y22n +1

As a quick check of the validity of this result, we make sure that the result does
indeed give us unity when integrated over all possible values w may assume; that is,

F(00 , W, Ow) = !~
-~
f(w, W, Ow) dw

-
2
- kor...fiii /
O
~ t ex
p
1 w+kor
- -
2 ( kar )
2] dw+- 2
../ii l (2n + l)n!y2 n
n
(-It
2 +l
Load Forecasting 51

'Jhe first terin tn the equation may be expressed as

.}; f
00

exp (-a?) dcx (2.49)


1/./2
upon making a change of variable

w+kor
ex ==
../2,kor
Since Eq. 2.49 is a standard error function integral, we get

-
vir
2 f 00
exp (-cx 2 ) dcx = 1 - erf - 1 = I - - 2
.../2 vir
l (-It
2
nh/f. n + 1(2n + I)
1/,Ji
Collecting our results, we have
_ 2 '\' (-It 2 '\' (-It
F(oo, w, or)= 1- yir Ln!...fj.2n+1(2n + 1) + yii Ln!y22n+1(2n + 1)
=1.0
as indeed it should. To get the probability that the WS component of peak load will
be less than a certain value, such as its mean value, Eq. 2.48 would be integrated over the
appropriate range.

2.8 TOTAL FORECAST

Knowing the means and variances of the NWS and WS components, we simply sum :'
them to obtain the total forecast mean and variance. Similarly, knowing f._ and f w, the
total forecast density function fo can be obtained. Since
0(t) = Ht)+ w(t) (2.50)
the probability density function fo may be obtained by convolving f._ with f w Because
t and w are assumed independent, the expression for fo is:

fo = f0
00 f._(0 - w)fw(w)dw (2.51)

With fe known, the system planner has at his disposal all the statistics required to
define quantitatively and judge future system weekly requirements.
Because of the computational ease of dealing with discretized quantities, the
actual convolution process required will be performed, not using Eq. 2.51, but the
discrete version:
52 Power System Planning

00

f0(81k) = ,L f1;(81k - wk)fw(wk) (2.52)


Wk=O

where ft and/w are the discretized density functions obtainable from Eq. 2.48 and the
Gaussian NWS peak load density function de.termined in Sec. 2.6. For example, if the
continuous density functions are as shown in Fig. 2.12, a and b, it will be a simple
matter, using standard numerical integration techniques, to create the discrete versions
shown in Fig. 2.13, a and b. Upon substituting the discrete values of wk into Eq. 2.52
for all values of 81k, the total forecast density function results.
To illustrate how / 0 may be obtained, consider the problem. of determining / 0
for the 46th future week for our sample three-bus system. From Figs. 2.5 and 2.6 we
found that
= 77.18 MW a/= 1.11 MW2
which completely describes the statistics of the NWS component for the 46th week,
because of the Gaussian characteristics of the NWS peak load probability density
function. Assume that the temperature distribution and weather load model for this
week are completely defined by the following:

w-->
(a)

(b)

Fig. 2.12 Continuous WS (a) and NWS (b) peak load probability
density functions.
Load Forecasting 53

Wk--+

(al

Fig. 2.13 Discrete WS (a) and NWS (b) peak load probability density
functions.

Weather load model: Temperature variable:


1. kw= ks= I MW/F = k I. T= 70F
2. Ts= 75F 2. or= 5F
3. Tw = 65F
Because Ts = T + or, T w = T- ar, and ks =kw, we may use the results of a previous
example for the assumed parameter values.

I 1(w+kor) ~
2
2 ] 2 (-It 0
fw=ko'I"fiirexpl-2 kor (w)+ ../ii nf;:on!(2n+I)y22n+I (w)

Upon discretizing f w, we get the discrete density function shown in Fig. 2.14a. An
interval of 2 MW was chosen to reduce the number of values wk could assume and
hence the computational work yet to be done. For example, the probability that
wk = 1.0 MW was assumed equal to the probability that O < w < 2 MW. The other
values were similarly determined. A numerical integration algorithm must be employed
to obtain the values depicted in Fig. 2.14a.
54 Power System Planning

0.6826

0.1563
i
fw
0.0946
0.0445
I
i:
i<
0.0187
0.0068

0 1 3 5 7 9
Wk-+
(a)

0.6826

'1;
0.1359 0.1359

0.0228
2a

74.55 75.60 77.18 78.76 79.81


/;k-+
(b)

Fig. 2.14 Sample system WS (a) and NWS (b) peak load probability
density functions.

Because the NWS component is Gaussian distributed, its discrete density function
may be represented as shown in Fig. 2.14b.
The next step is to convolve f w with f1; to obtain f 0 , where

81k =~I+ Wk (2.53) I'


and ~1 and wk are assumed independent, a simple procedure easily carried out by using 1
t
Eq. 2.52. The results of this convolution are tabulated in Table 2.3. Clearly, with the l
l
aid of Eqs. 2.50 and 2.52 to obtain 0 and fe as given in Table 2.3, the system planner 1
has a good quantitative description of the peak demand to expect during the 46th
week.
I
'
2.8.1 Seasonal and Annual Forecasts J
}

If the historical data base used is developed on a seasonal or annual basis, the steps !
1
required in determining future forecasts are simpler than the steps required for weekly

i
f
(
Load Forecasting 55

TABLE 2.3 Total peak demand weekly


forecast probability density function
0 fo (} /0
74.55 0.0156 81.55 0.0004
75.55 0.0036 81.76 0.0129
75.60 0.0928 82.18 0.0304
76.60 0.0212 82.60 0.0025
77.18 0.4659 82.81 0.0022
77.55 0.0022 83.55 0.00015
78.18 0.1067 83.76 0.0060
78.60 0.0129 84.18 0.0128
78.76 0.0928 84.60 0.00092
79.55 0.0010 84.81 0.0010
79.76 0.0212 85.76 0.0025
79.81 0.0156 86.18 0.00464
80.18 0.0646 86.81 0.0004
80.60 0.0060 87.76 0.00092
80.81 0.0036 88.81 0.00015

forecasts, primarily because the range over which the coincident dry-bulb temperature
varies usually is very much smaller or, stated differently, ar is quite small. The effect
of a small temperature variance is to locate the entire temperature density function in
a completely linear region of the weather load model, as depicted in Fig. 2.15. In this
case the resulting WS peak load density function is Gaussian since the temperature is
Gaussian distributed.

~
2
-ti
.,
C:

E
"'
"O
-"'
m
2l. Do
z,
.,
0

Ts
Temperature, F
Fig. 2.15 Seasonal temperature density function and weather load model.
56 Power System Planning

To obtain the mean and variance of w, we must start with


w = ks(T - Ts) or w = kw(Tw - T)
From these expressions wand Ow 2 can be easily obtained:
w = e(ks(T - Ts)]= ks(T - Ts) ., (2.54)
or

and

(2.55)
or
Ow2 = e{[kw(Tw - T)- w]2}
= kw2o/

The total seasonal or annual forecast probability density function is completely


specified as we know
e=l+w (2.56)
and
a02_
- 0t
2+ 0 2
w
where the appropriate wor Ow 2 is used, depending on the season in question.
Since the total seasonal forecast is Gaussian and its mean and variance are given
in Eq. 2.56, fo may be expressed as

oo -)2J
[-21 (0-0
- 1
fo(0, 0, oo) = oovfii exp (2.57)

2.9 ANNUAL AND MONTHL V PEAK DEMAND FORECASTS


From Secs. 2.7 and 2.8 we have the mean, 0, and variance, o0 2 , of the forecast on a
weekly basis. In addition we have the probability density function fo for the peak
demands, where fo is skewed and truncated for weekly forecasts. Here we discuss how
we may obtain an annual or monthly forecast given the mean, variance, and density
function for the weekly forecasts based on the work by Stanton and Gupta [9].

2.9.1 Annual Forecast


If we denote the weekly probability density function for the rth week as fr,
understanding that the f is really fo, then the corresponding probability distribution is:
Load Forecasting 57

Fr(0, 0r, Ur)= f


-00
8

fr d0

= probability that 0 .,;;; a particular 0 but> - 00


(2.58)

where

fr= fw
ft(0 -w)fw(w)dw (2.59)

which is the continuous convolution equation discussed earlier. The discrete versions of
Eqs. 2.58 and 2.59 are:
8

Fr(0, e,, Ur)= l fr(01k, Or, Ur) (2.60)

and

fr(01k, 0r, Ur)= l


Wk
ft(81k - Wk)fw(wk) (2.61)

Upon substituting Eq. 2.61 into Eq. 2.60, we have the probability distribution
for the peak load for the rth week.
8

Fr(0, Or, u,) = ll


-= Wk
ft(01k - wk)fw(wk) (2.62)

To find the annual peak, we must recognize that the probability that an annual
peak occurs in a given week r is equal to the probability that an annual peak occurs in
the rth week and not in the other 51 weeks. The probability that a peak greater than
A does not occur in the other weeks is equal to the probability that these weekly
peaks are less than or equal to the peak A; this probability is
52
fl FiCA,0i, ui)
i= I
(2.63)
i ,fo r

where Fi is the appropriate distribution as defined by Eq. 2.62. Since the probability
that the rth weekly peak is greater than A by an amount dA is
fr(A, Or, Ur)dA
then clearly

[~DI1,/a r
Fi(A, ei, Uj)lf,(A,
J
e,, u,)dA
58 Power System Planning

is the probability that an annual peak load greater than A but less than A + dA occurs
in week rand

f(r,A) = fr(A, 0,., Or) n


S2

i*r
F;(A, 0;, o;) (2.64)
i= I

is the joint probability density function for r and A. The occurrence of an annual peak
in week r is assumed to be independent of the nonoccurrence of an annual peak in any
other week. Stated differently, knowing that the annual peak did not occur in a
particular week is not enough to determine the week in which it will occur.
If we want the annual peak probability density function, we must sum Eq. 2.64
over all r; that is,
52
f(A) = '\"
rff I
fr(A' Br, Or) nFj(A' 8i, Oj)
i*r
(2.65)

Using the annual peak load probability density function in Eq. 2.65, we can
obtain the mean and variance of the annual peak load using

A=!~ Af(A)dA (2.66)


0

or
~

- LA;/i(A)
0

and

var(A) = !~(A-A)2f(A)dA (2.67)


0

or

-l
00

(A; -A)2f;(A)
0

2.9.2 Monthly Forecasts


A utility frequently finds it convenient to have monthly as well as annual forecast
information for use in corporate economic studies, for production analyses, and for
other purposes. To obtain the monthly forecast peak demand probability density
function, we need only reapply the concepts used to obtain annual forecast
information. Specifically, if Eq. 2.65 were summed over just those weeks in month k,
rE k, we have the desired density function. Thus, for month k
Load Forecasting 59

(2.68)

A and var(A) are given by Eqs. 2.66 and 2.67.

2.10 SUMMARY

ln this chapter we have briefly discussed the purposes of forecasting, the classification
and characteristics of loads, the approaches used in forecasting system requirements,
and the forecasting methodology. The point has been made that no particular method
or approach will work for all utilities. There are as many valid forecasting methods as
utilities; however, all methods are strung on a common thread, and that is the
judgment of the forecaster.
Because energy or sales forecasts are important in identifying the best expansion
plan, a discussion of considerations used in sales forecasting was presented. Presenta-
tion of analytical methods for sales forecasting was omitted because of the subjective
nature of forecasts dependent on broad den:iographic and economic indicators.
Of the basic methods presented for demand forecasting, the probabilistic
extrapolation-correlation method was favored, because it facilitates presentation of the
basics of extrapolation, correlation, weather load modeling, and so on, all fundamental
to intermediate-load forecasting. Further, through the use of a sample three-bus
system, the execution of details of weekly peak demand forecasting was illustrated.
Finally, the material was extended to include monthly as well as annual forecasting
procedures, for many utilities find monthly and annual forecasts more convenient for
certain applications.
In no way is the material in this chapter exhaustive. The intent has been to
introduce some ideas presently used in forecasting system load requirements and to
facilitate discussion in subsequent chapters of the methods and techniques used in
power system expansion planning.

PROBLEMS

2.1. Compile and plot monthly historical sales, peak demand, and load factor data for
a utility in your area. From these plots, describe the basic load characteristics for
the utility.
2.2. Compile and plot daily peak demand versus coincident dry-bulb temperature.
Superimpose on the plot a three-segment weather load model, and calculate D 0 ,
k 8 , kw, Ts, and T w
2.3. Show that Eq. 2.20 is an unbiased estimate of the coefficient vector. (An
unbiased estimate is an estimate whose expected value equals the true value.)
2.4. Prove that Eq. 2.26 is an unbiased estimate of a11 2
2.5. Using the recursive algorithm developed in Sec. 2.6 for determining Ht), calculate
the best estimate of the model that describes the following data:
60 Power System Planning

t l(t)

0 90
10 95
20 110
30 135
40 170

Also calculate 0 71 2 for the data above, using Eq. 2.26. Knowing 0 11 2 , determine
the confidence limits for each element of the coefficient vector.
2.6. Write a computer program to determine the model for sales data collected in
Prob. 2.1. Using the results of Prob. 2.2, separate the historical peak demand data
into its two components, and determine the model for the NWS component. Note
that if after removal of the WS component a significant seasonal variation
remains, it must be removed before attempting to obtain the model.
2.7. Assume that for a particular week the following data describe the weather load
model and temperature density function:
var(T) = oi
Determine the mean, variance, and probability density function for the corre-
sponding WS peak demand.
2.8. Consider the problem in which the total system peak demand is desired for a
particular season in which the NWS component was found to have a mean of {
and a variance of ot 2 The weather load model for the season is known, and the
temperature statistics are given by
var(T) = al
Determine the mean, variance, and density function for the total peak forecast.

Bl BLI OG RAPHY
1. Brown, R. G.: "Smoothing, Forecasting and Prediction of Discrete Time Series," Prentice-Hall,
Englewood Cliffs, N.J., 1963.
2. Davey, J.: Weather Sensitive Load Forecasting; Application to System Planning, IEEE Trans.,
vol. PAS92, pp. 971-977, May-June 1973.
3. Davies, M.: The Relationship Between Weather and Electricity Demand, Proc. IEEE, vol. 106C,
pp. 27-37, October 1958.
4. Gupta, P. C.: A Stochastic Approach to Peak Power Demand Forecasting in Electric Utility
Systems, IEEE Trans., vol. PAS71, pp. 824-832, March-April 1971.
5. Report to the Federal Power Commission on the Methodology of Load Forecasting, 1969.
6. Shiskin, J.: Electronic Computers and Business Indicators, National Bureau of Economic
Research, occasional paper,-1957.
7. Stanton, K. N.: Medium-Range, Weekly and Seasonal Peak Demand Forecasting by Probability
Methods, IEEE Trans., vol. PAS71, pp. 1183-1189, May 1971.
8. Stanton, K. N. and P. C. Gupta: Long Range Forecasting for Electric Utility Systems, Purdue
Energy Res. Educ. Cent. Rep. No. 31, August 1968. -
9. Stanton, K. N. and P. C. Gupta: Forecasting Annual or Seasonal Peak Demand in Electric Utility
Systems,IEEE Trans., vol. PAS70, pp. 951-959, May 1970.
GENERATION SYSTEM
RELIABILITY ANALYSIS

Generation system planning is one of the most crucial steps in planning the expansion
of a modern electric utility. Decisions and commitments made at this stage have a
tremendous effect on all other phases of system expansion and dictate the financial
posture a utility must assume. A sound plan for generation system expansion is crucial
to the success of any electric utility.
In broad terms a suitable generation expansion plan must provide the electric
utility with the capability of meeting customer needs for a reasonably priced, reliable,
quality electric energy source. Of course, every investor-owned utility must consider
only those expansion plans that will enable it to maintain a sound financial posture
and hence to attract new investors. The electric utility industry is a very capital-
intensive industry, a fact that continually requires the industry to seek externally
generated funds to finance construction.
Choosing a generation system expansion plan among the many available to an
electric utility is complicated, especially since all utilities must strive for the best
strategy in an environment of uncertainty. In addition to the uncertainty inherent in
forecasting future load requirements, the planner must deal with the uncertainties
associated with
I . Unit reliability and maintenance schedules
2. Fuel costs 61
62 Power System Planning

3. Pollution abatement legislation and costs


4. Construction costs
5. Start-up times
6. Availability and cost of capital

No system planner in 1971 would have projected a threefold increase in fuel


costs, pollution abatement equipment costs higher than small power plant costs, 10-yr
construction times for nuclear plants in the United States, increased forced outages for
1,000-MW plants, and astronomical increases in construction costs. These facts are a
reality today. System planners must prepare themselves to account for such
uncertainties if the electric utility industry is to survive in its present form.
Obviously we shall not be able to discuss comprehensively how all the
uncertainties mentioned may be accounted for. However, we shall concentrate in this
chapter on a technique used to determine whether a generation expansion plan satisfies
a desired level of reliability defined by two reliability indices: loss-of-load probability,
LOLP, and expected value of demand not served, E(DNS). Specifically we shall discuss
how to analyze the reliability of a proposed expansion strategy for isolated and
interconnected systems, considering

I. Forecast uncertainty
2. Unit availabilities
3. Unit maintenance schedules
4. Interconnection constraints

Although we choose the LOLP and E(DNS) as the reliability indices of interest,
the student should be aware of the existence of another approach to reliability
analysis, generally referred to as the frequency and duration (FD) method. In essence,
the FD method enables the system planner to determine the frequency of a particular
generation system outage as well as its expected duration. Also, the frequency and
duration of cumulative outages states may be determined using convenient recursive
relationships. The impetus for development of the FD method has come from the need
for. a generation system reliability technique compatible with transmission system
reliability methods and historical transmission system outage data. In fact, the Edison
Electric Institute has standardized. outage data into six basic indices, two of which are
average customer interruption frequency and duration per customer interruption.
However, the utility industry appears to prefer the LOLP and E(DNS) method to the
FD technique for generation system planning. This is not to say that, in certain
instances, more definitive outage frequency and duration information will not better
serve the planner. We do not categorically defend one method over the other; however,
for the sake of unifying the material and the ideas to be presented, the LOLP and
(DNS) methodology is used.
In practice, the way alternative generation system expansion plans are developed
is based on experience rather than on rigid analytical tools. Frequently the planning
staff will collectively define a number of feasible expansion alternatives on the basis of
Generation System Reliability Analysis 63

1. Load growth
2. Construction time
3. Availability of sites
4. Availability of fuel
Given these alternative plans, it is common to subject each to a detailed reliability
analysis to ensure that all satisfy the desired level of reliability. Obviously, plans that
do not meet the reliability criteria are eliminated or appropriately modified.
Another approach used to some extent in the industry is a semiautomated
procedure in which capacity requirements are determined over the horizon period using
analytical methods. With this information as a guide, various expansion plans can be
determined by varying the type and timing of unit additions. It is safe to say at this
point that few, if any, analytical methods exist for identifying plans that are equally
good in terms of reliability. Invariably, the final selection of acceptable plans is made
by the planning staff on the basis of information accumulated over years of living with
a system. This is not to say that a planner should ignore analytical methods that
provide additional insight into the advantages or disadvantages of various plans; he does
not and should not rely totally on such methods.
We choose here to discuss how expansion plans devised by the planner can be
analyzed in terms of reliability with the uncertainties associated with loads, unit
availabilities, etc., adequately included. This approach has been selected because of its
simplicity and flexibility. We shall leave it to the planner to define those plans to be
considered; indeed, he is best qualified to do so.

3.1 PROBABILISTIC GENERATING UNIT MODELS


Because the reliability of a system depends on the reliability of its generation system,
which often consists of many different types of units that are randomly forced off-line
by technical problems, we must discuss the method used to model the random
availability of a unit.
Today in the electric utility industry there are many different types of units, all
with their own peculiarities, but basically the units fall in one of the following classes:
1. Base load
2. Midrange
3. Peakers
As one might expect, the base-load units operate at very high capacity factors,
ideally at 90-95%, but often lower if the unit is immature or fraught with technical
problems. Units in the base-load class are usually large fossil-steam, nuclear, or to some
extent hydro units. Midrange units operate less than base7Ioad units but more than
peaking units; generally they operate at capacity factors of 30-75%. For midrange
operation, combined-cycle combustion turbines, hydro units, and small fossil-steam
units are usually employed. Peakers operate only during peak demand periods and
therefore are characterized by very low capacity factors; 5-10% is a typical range.
64 Power System Planning

Generally speaking, combustion turbines are used as peaking devices in thermal


systems, but hydro units are used in this manner as well. At one time it was safe to
assume that base-load units were the most efficient ones, and midrange and peaking
units the least efficient. Since the appearance of combined-cycle and regenerative
combustion turbines, the efficiency differential is now small. This is not to say that
midrange units and peakers should be base-lo'aded. These devices are designed for less
than full-time operation, and prolonged use beyond their design capability results in
high maintenance costs.
Although many different types of units are in use today, all units are randomly
forced off-line because of technical problems during a normal period of operation.
Hence, to account for the random outage or availability of a unit, it is necessary to
determine the probability density function that describes the probability that a unit
will be forced off-line or will be available during its normal period of operation. For
example, assume on the basis of historical data that the availability of the generating
capacity of a given unit may be graphically represented as shown in Fig. 3.1, which
conveys the idea that random failure and repair of a unit can be defined as a two-state
stochastic process, where a stochastic process is defined as a process that develops in
time in a manner controlled by probabilistic laws. The so-called "state-space diagram"
for the stochastic process, shown in Fig. 3.2, tells us that a unit may be in state 1 (the
up state, or the state corresponding to maximum available capacity) and then
randomly transfer to state 2 (the down state, or the state corresponding to no available
capacity) and vice versa. For simplicity, only two states are assumed.
Although several questions might be asked about the stochastic process described
in Fig. 3.2, we shall concentrate on one: What is the long-term average (steady-state)
probability of finding the unit up or down? The answer to this question defines two
important quantities:
1. Unit availability (denoted by the variable p): the long-term probability that
the generating capacity of a unit will be available
2. Unit forced outage rate (abbreviated FOR and denoted by the variable q): the

Up time
,,--A---._
Up state

J:7 R,[l
(or state 1)

Failure Failure Failu re

Down state
(or state 2) ----.---
Down time
Time
Fig. 3.1 Random unit performance record ignoring scheduled
outages.

----- -------- .. ------------'----


Generation System Reliability Analysis 65

Up state (or state 1 )

">-- (failure rate) (repair rate)

Fig. 3.2 Generating unit state-


space diagram.

Down state (or state 2)

long-term probability that the generating capacity of a unit will be


unavailable, or forced off-line
To obtain an expression for the long-term availability of the generating capacity
of a unit, it is first necessary to recognize the stochastic process we are considering is
a very special one, called a zero-order, discrete state, continuous transition Markov
process. Such a stochastic process has the following properties:
1. The system described (here the available capacity of a generating unit) can be
characterized as being in one of a set of mutually exclusive, collectively
exhaustive, discrete states S 1 , S 2 , , Sn at any time. Let us further
interpret this property. We assume that a generating unit can be in either the
up or the down state, but not in both simultaneously; thus the states are
mutually exclusive and discrete. Further, since we assume that the only
possible states for a generating unit are the up and the down states, then
these states define all the possible states we ever expect to find a unit in; thus
the states defined are collectively exhaustive.
2. Changes of state are possible at any time.
3. The probability of departure from a state depends only on the current state
and is independent of the independent variable time.
4. The probability of more than one change of state during an appropriate small
time interval t:.t is negligible.
There are two good reasons for modeling unit generating capacity availabilities as
Markov processes. First, a Markov process is usually a fairly accurate model of real life,
and second, the mathematical description of the process is simple.

3.1.1 Mathematical Description of a Discrete State,


Continuous Transition Markov Process
We have set the stage to define precisely the availability and FOR of a unit. Define
Pi(t) as the probability of finding a generating unit in state i (where i = 1 is the up
state and i = 2 the down state) at time t. Further, define X as the transition rate from
state 1 to state 2. This parameter can be obtained from information in Fig. 3.1; 1 /X is
66 Power System Planning

the average time a generating unit stays in the up state. Similarly, is defined as the
transition rate from state 2 to state I, and I/ is the average time a generating unit
stays in the down state. If we want to know the probability that a generating unit will
be in, say, the up state (state I) at time t + llt, then we must recognize that there are
only two ways a unit can reside in the up state at time t + llt. A unit could be in the
up state at time t + llt if it were in the up'state at time t and did not transfer to the I
I
down state in the time increment llt, or it would be in the up state at time t + llt if it -I
I
were in the down state at time t and then transferred to the up state in the time
increment llt. The probability that the first event will occur is P 1 (t)(I - "A.llt), where
P 1 (t) is the probability that a unit is in the up state in time t. It can be shown that
(I - "A.llt) is the probability that the unit does not fail in time llt if it is assumed that
the probability of a unit failure can be described by an exponential distribution:

F 1 (t) = e -'1'.t ~ probability of unit being available up to time t


2
=I- "A.llt + -"A 2llt
- +
"" I - "A.llt g probability of unit being available during time /1 t (3 .I)

where

"A.llt ~ probability of transferring in time llt


F 2 (t) = e -t ~ probability of unit being unavailable up to time t
:::::: I - llt ~ probability of unit being unavailable in time llt (3.2)

where
llt ~ probability of transferring in time llt
likewise, the probability that the second event occurs is P2 (t)llt, where P 2 (t) is
the probability of being in the down state at time t, and llt is the probability that a
transfer from the down state to the up state occurs in time increment llt. Thus
P 1 (t + llt) = P 1 (t)(I - "Allt) + P2 (t)llt (3.3)
and similarly
P 2 (t + llt) = P2 (t)(I - llt) +P1 (t)Mt
Upon rearranging these two equations, we have

P 1(t + 11~~ - P1(t) = -}.J\(t) + P2(t)

P 2 (t + 11~~ - P 2 (t) = "APi (t) _ P (t)


2
Generation System Reliability Analysis 67

or as /it-+ 0,

dt
dP (t)
= -).pl (t) + P2(t)

ad?)= )..Pl (t) - P2(t)

As we are interested only in the long-term (steady-state) probabilities of being in


either state 1 or 2, where P 1 + P2 must equal unity, we may set the derivatives to zero
and solve the resulting two equations for P 1 (t) and P2 (t). Thus

P 1 (t) = - - = p (3.4a)
A+
A (3.4b)
P2(t) = A+ = q
which are the expressions used to define the availability p and forced outage rate q of
a generating unit. Note that
A
P1 (t) + P2 (t) = A++ A+= I
or that
p+q=I

3.1.2 Probability Density Function of Forced Outage Capacity


It is common and convenient to work with the forced outage capacity probability
density, f 0 (L 0 ;), of a generating unit i. For example, if a 50-MW unit has a forced
outage rate q = 0.02 and an availability p = 0.98, then its forced outage capacity
density function will be as shown in Fig. 3.3. Interpreted correctly, Fig. 3.3 tells us that

-:;:,
C 0.98
:::!.
....,c

0.02 Fig. 3.3 Forced outage capacity


probability density function.
0 50
Outage capacity, MW
68 Power System Planning

:q: 0.98
:::!
,;:.

0.02 Fig. 3.4 Unit available capacity


probability density function.
0 50
Available capacity, MW

the probability of an outage of O MW is 0.98 (where a 0-MW outage is identical to


having 50 MW available) and that the probability of a 50-MW outage is 0.02. The
usefulness of unit outage capacity distributions will be demonstrated many times in
subsequent sections.
In some cases the outage capacity probabilities are expressed in terms of available
unit capacity. For example, instead of representing the outage capacity of a unit on
the abscissa, as in Fig. 3.3, if we represent its available capacity, then Fig. 3.4 results.
The information is clearly the same in both Figs. 3.3 and 3.4. We shall refer to Fig.
3.4 as the available capacity probability density, fA(LA;), of a unit.

3.1.3 Generating Unit Models


To facilitate introduction of the concept of an effective load in Sec. 3.3, we point out
now that a unit with capacity C; that is randomly available may be modeled as a
fictitious unit of capacity C; that is 100% reliable and a fictitious load of C; MW
whose availability is equal to the FOR of the actual unit. These two representations
are identical since, when the fictitious load is available, the net injected demand into
the system is zero, just as it is when the actual unit is forced off-line. This
representation of a unit will be used throughout this chapter and in Chap. 4.

3.2 PROBABILISTIC LOAD MODELS


The probabilistic load model that best suits our needs here describes the probability
that the load will exceed a certain value, for we are interested in determining the
probability that load exceeds installed capacity (IC) of a particular proposed generation
system (the LOLP). Fortunately, the data required to develop such a model are readily
available, since continuous recordings of system demand and energy are usually
obtained on a routine basis by electric utilities. If a recording of instantaneous
demands were plotted for a particular historical week, a curve such as depicted in Fig.
3.5 might result. From this curve the so-called "load duration curve" in Fig. 3.6 is
easily constructed. The load duration curve is created by determining what percentage
of time the demand exceeded a particular level, where percent time is shown on the

--- --------- -- ---- ------------------ -


Generation System Reliability Analysis 69

"O
C:

.,E"'
0

Time

Fig. 3.5 Instantaneous demand vs. time curve.

x-axis and demand level on the y-axis. The area under the load curve is the energy
in MWh consumed by the load and transmission system. Further, the area under
the curve defines the energy to be produced by the base-load units, peaking units,
and midrange units.

3.2.1 Load Probability Distribution


For generation system studies it is necessary to interchange the axis parameters in Fig.
3.6 and normalize time, producing the load probability distribution in Fig. 3.7, where
the y-axis shows the probability that the load exceeds the corresponding x-axis
megawatt value. This load distribution will be denoted generally by Fk(L), where k
indicates the time period ( usually a week) for which the distribution is appli-
cable.
The probability distribution for a particular historical week (like the
26th week of each year) when per-unitized (i.e., x-axis divided by weekly peak

..J
"O
C
"'
E
8 Base-load
Fig. 3.6 Typical load duration
curve.

0 100
Percent of time
70 Power System Planning

Fig. 3.7 Load probability distri-


bution for week k.

Demand, MW

demand) is essentially the same for each year. Thus each week is described by a basic
load shape such as depicted in Fig. 3.8. Using the load shapes for each week along
with the forecast mean weekly peak of Chap. 2, the load probability distribution for
each future week may be determined. In practical situations the load shapes may
change very little during a particular month; thus it may be necessary to define only
13 load shapes (12 monthly peak periods and I off-peak period) to describe 52 weeks
in a year.

3.2.2 Modeling Forecast Uncertainty


To include the uncertainty of what actual peak demand will be for a future week, as
defined by the weekly peak demand probability density function, / 0 , developed in
Chap. 2, we must recognize that the load density function including forecast
uncertainty is a joint probability density. The two random parameters described by the
joint density function are the random forecast peak demands defining the peak load
and the corresponding values of random system load levels. For example, for a
particular peak demand, the values of demand the system load could assume will
depend on the random load levels allowed, as defined by the load shape. As peak
demand randomly changes, demand values the system load can assume will change, the
amount of change again depending on allowable load levels. We shall work a numerical
example to clarify this difficult notion.

Fig. 3.8 Load shape for week k.


Ol.==0/8<1 8/8 == 1.0
Demand, per unit MW

1h
---------~-----------------------------11,.-.
Generation System Reliability Analysis 71

In order to construct the load probability distribution to include forecast


uncertainty, we must define the joint probability density function f(L, 0), where 0 is
the random peak demand and L denotes the random demand levels L = o:0(O < o: < 1)
that the system load assumes. We obtain the desired demand distribution F(L) as
follows:
f(L, 0) = f(L/O)f(O) (3.5)

f(L) = f-00
00

f(L/O)f(O)dO (3.6)

but

F*(L) = f-oo
L

f(L) dL = ff L

-0()
""

-00
f(L/0)f(O)d0dL

J J
"" L

= f (0) f(L/0)dLd8
-Q() -oo

= f-oo
00 f(0)F*(L/0)d8 (3.7)

since by definition the integral of a density function f(x) is the distribution function
F*(x ), where for a particular X, F*(X) is the probability that ----QC>< x < X. In
generation studies, however, F(X) is the probability that x > X, which is .1 - F*(X).
Although we might glide over this subtle point, it would no doubt create some
confusion later. To obtain F(L) from F*(L), given in Eq. 3.7, we need only subtract
both sides of Eq. 3.7 from 1.

I - F'(L) - F(L) - I - !-
-oo
f(O)F*(L/O)dO

since
72 Power System Planning

Therefore

F(L) = f -oo
00

f(0)F(L/0)d0 (3.8)

Since f(_O) is identical to the discrete forecast weekly peak demand density 1
function developed in Chap. 2, it may be expressed as
/(0) = fo(0)o(0) for all 0 = e,k (3.9)
Substituting Eq. 3.9 into Eq. 3.8,

F(L) = L
O[k
F(L/0)[0 (0) (3.10)

In the special case when forecast uncertainty is ignored,


e,k = 0 and f(0) = 1.0
Therefore F(L) including uncertainty= F(L) excluding uncertainty, as indeed it should.
To illustrate some of the ideas just discussed, assume that the simple load shape
shown in Fig. 3.9 adequately describes our three-bus system for the 46th future week.
If we now define the peak demand value to be equal to the forecast mean peak
computed in Chap. 2 and to be equal to 78.035 MW, the load probability distribution
in Fig. 3.10 describes the load. From this distribution the probability that the load will
exceed any specific value can easily be obtained; for example, the probability that the
load exceeds 60 MW is 0.385, etc. If we wish to include the uncertainty of our
forecast, then the forecast weekly peak demand probability density function tabulated
in Chap. 2 must be substituted into Eq. 3.10, and F(L) evaluated. F(L/0) is the
conditional demand probability distribution obtained for a particular value of 0 usi~g
the basic load shape in the same way that it was used to obtain Fig. 3.10 for 0 = 0.
Because a number of steps are required to determine F(L) completely, with certainty
included, we elect to leave the determination of F(L) including forecast uncertainty as
an exercise for the student.

F(LI = 1.0
1.0 1------.....

F/J.. I = -10
- 6- (L -0.4) + 1.0

0.0 ..___ _ _- - ; i - - - - - - - - . . . , _ _ -
0.4 1.0
L, per unit MW

Fig. 3.9 Load shape for sample system.


Generation System Reliability Analysis 73

F(L) = 1.0
1.0

F(L) =-~O (L -0.48) + 1.0


60

31.21 0 = 78.034
L,MW
Fig. 3.10 Load probability distribution for sample system.

3.2.3 Expected Value of Demand and Energy


In addition to the LOLP reliability index, we shall find it useful to determine the
expected value of the demand not served, 1:(DNS), associated with a given generation
system expansion plan. Using these two reliability indices, the planner can choose an
expansion plan that is a compromise between cost and reliability.
To initiate our discussion, we should for the moment consider only the basic
problem of calculating the expected value of the demand whose probability
distribution is F(L), as depicted in Fig. 3.7. By definition

L= f_,,,
"'Lf(L) dL

but

f (L) = .E,__ F*(L)


dL
Therefore

L= I 0
I

LdF*(L)

Making a substitution of variable

F*(L) = 1 - F(L)
dF*(L) = -dF(L)
then

L =
/
I
0

-L dF(L) = f
0
I

L dF(L) (3 .11)
74 Power System Planning

Although Eq. 3.11 is theoretically correct, it is not often convenient to express L as a


function of F(L), as required in the argument. On the other hand, to express F(L) as a
function of L is trivial. In the last example, the probability distribution shown in Fig.
3.10 could be defined as follows:
IO -
F(L) = --=(L - 0.40)+ 1.0 L >, 0.40
60
and
F(L) = 1.0 L <0.40
where 0 is the peak demand value assumed, with forecast uncertainty ignored.
To express Eq. 3.11 as an integral over L, we need only realize that
d(LF(L)) =LdF(L)+F(L)dL (3.12)
Therefore
L dF(L) = d(LF(L))- F(L)dL
and

f0
1

L dF(L) =1 0
8

F(L)dl

and hence

I= I0
8

F(L)dL (3.13)

To illustrate the use of Eq. 3.13, we shall evaluate l for the sample system load
distribution shown in Fig. 3.10.

I= f0
31.2 l

F(L)dL + f 78.034

31.21
F(L)dL

= f0
31.21

dl + I 78.034

31.21
[-0.02136(L - 31.21) + l]dL

4
= 31.21 + (-0.0IOE8L 2 + 0.666L + L)~~:~1
= 31.21 + (65.01 - 41.61) = 54.61 MW
Since the probability distribution describes a week, the expected value of the energy in
MWh is given by

--------- -------~----------------
Generation System Reliability Analysis 75

E= T f 0
8

F(L) dL = TL

= 9174.8 MWh (3.14)

for T= 168 h.
As we mentioned at the beginning of this section, e(DNS) is the expected value
of the demand not served in excess of the installed capacity of a system, which is
more difficult, conceptually anyway, to obtain than I. However, this quantity is
essential, so we must discuss theoretical points and develop analytical expressions.
Let
DNS=L-IC for 1c,,,;;;L,,,;;;0
(3.15)
DNS = 0 for L < IC
Then by definition

<(DNS)- !- -00
DNSf(DNS)dDNS

= 1 0
8

DNS f(DNS) dDNS

Based on Eq. 3.15, where IC is a constant


f(DNS) = f(L = DNS + IC) IC ,,,;;; L ,,,;;; 0 (3.16)
Substituting Eqs. 3.15 and 3.16 into e(DNS), we have

e(DNS) = J IC
8

(L - IC)f(L)dL

1.0

=
1 F*(IC)
(L - IC)dF*(L)

= 1 0

F(IC)
(IC - L)dF(L)

but
d(IC - L)F(L) = (IC - L)dF(L) - F(L)dL (3 .17)
76 Power System Planning

1.001-----~

8
LOLP = F(IC) 1 IC
F(L) = e(DNS)

0 I 0
IC
Demand L, MW
Fig. 3.11 Area definitions of load distributions.

Thus
o 0

E(DNS) =
J
0
d(IC - L)F(L),+.J F(L)dL
IC

= f
IC
0

F(L)dL
(3.18)

and

E(ENS) =T f IC
0 F(L)dL (3.19)

Pictorially the indices of reliability are shown in Fig. 3.11.


We shall use Eqs. 3.10, 3.13, 3.14, 3.18, and 3.19 in subsequent sections to
calculate
1. LOLP
2. Expected value of demand
3. Expected value of energy
4. Expected value of demand not served
5. Expected value of energy not served

3.3 EFFECTIVE LOAD


In Secs. 3.1 and 3.2 we developed probabilistic models for both the generating units
and the load, with and without forecast uncertainty included. Now we wish to
combine these two models to define the effective load of the system. Figure 3.12
Generation System Reliability Analysis 77

Capacity C 2 ( 100% reliable I


Capacity C 1 (100% reliable)

Lo, ={oc,
Random system load L

Fig. 3.12 Fictitious generating units and system load model.

depicts the relationship between the system load and generating units, where actual
units have been replaced by fictitious perfectly reliable units and fictitious random
loads, whose probability density functions are the outage capacity density functions of
the units.
From Fig. 3.12 the load, called the effective load, is defined by

(3.20)

where Loi is the random outage load, if you will, of the ith unit. As we discussed in
Sec. 3.1 .4, when Loi= C;, the net demand injected into the system for the ith unit is
zero, just as it would be if the actual unit of capacity Ci were forced off-line. Also
from Fig. 3.12 we note that the installed capacity of the system is given by

IC= IC;
i
(3.21)

In the special case where actual units are 100% reliable, Loi = 0, i = 1, ... , G,
and Le= L. Unfortunately, this case never occurs, so we are forced to obtain F(Le)
from F(L) and f 0 lL 0 ;). Since Le is the sum of the independent random variables, L
and L 01 , i = 1, ... , G, whose distributions are known, we can obtain F(Le) using the
recursive convolution equation

F 1(Le) = f
Loi
pi-l(Le - Lo;)foCLo;)dLoi (3.22)

where F 1(Le) is the effective load probability distribution with the outage capacity of the
first i units convolved in. Obviously
78 Power System Planning

for i =0
for i =G
Since [ 0 ; is a discrete density function as depicted in Fig. 3.3 for the sample system,
Eq. 3.22 becomes:

pi(Le) = ,l pi- l (Le - Lo;)fo;(Loi) (3.23)


Loi

Loi denotes the discrete values L 0 can assume.


Because the outage capacity of a unit was defined as a two-state stochastic
process where
foi(Lo; = 0) = Pi
foi(Loi = C;) = q;
Eq. 3.23 simplifies even further:
pi(Le) = pi- 1(Le)P; + pi- 1(Le - Ci)qi i=l, ... ,G (3.24)
The effect on F(L) of accounting for the random outages of the generating units
is to inflate it. This effect tends to increase the probability that the load will exceed a
given value, thus reflecting the fact that as units are randomly forced off-line,
remaining units see a larger effective load, for they must pick up the load not served
by units forced off-line.
For clarification of the use of Eq. 3.24, consider again the sample system whose
load probability distribution is shown in Fig. 3 .l O and whose unit outage capacity

p 1 = 0.98

q, = 0.02
fo,

0 50
L 01 , MW
(a)

p 2 = 0.97

Fig. 3.13 Forced outage capacity


q2 =0.03 probability density function of
fo, unit 1 (a) and unit 2 (b).
0 50
L 02 , MW
(b)

- - - ---------------- -------~---------------------
Generation System Reliability Analysis 79

10p1P2
Slope =- - = 0.020303
1.000 i - - - - ~
. . . ........ I
,.
68

10 l 0.0494
"---, Slope=--=- lp,q2 + q,P2)
60 I
F(L) ,.,; = 0.00104 ',
F(Le)
F(L)
F(L 6 ) '---, Slope= 1 ~ q 1 q 2 +o.0006
, 60
,/ = 0.000012a I
8 = 78.034 MW .,
.......
= C = 50 MW
C1 2 ,
,.
0 0.48 8 (C1 + 0.40) (C1 + O) (C 1 + C2 + 0.48) (ij + C1 + C2 = 178.034)
Land Le, MW

Fig. 3.14 F(L) and F(Le) probability distributions.

density functions are depicted in Fig. 3.13, a and b. From Eq. 3.24 we can easily
evaluate F 2 (Le = 60):

F 2(60) = F 1(60)p 2 + F 1(60 - 50)q2


= probability that effective load exceeds 60 MW
where
F 1 (60) = F(60)p 1 + F(60 - 50)q 1
and
F 1{10) = F(10}p 1 + F(l0 - 50)q 1
With numerical values from Figs. 3.10 and 3.13b substituted,
F 1 (60) = 0.385 (0.98) + 0.02 = 0.3973
F 1 (10) = o.98 + 0.02 = 1.000
Therefore
F 2 {60) = 0.3973(0.97) + 0.03 = 0.41538
With F(Le) evaluated for other values of Le in the same manner, the effective load
probability distribution in Fig. 3.14 is obtained. Note that the distribution is not
drawn to scale, because of the wide range of values of F(Le)-
Next we shall discuss the reliability of an isolated system using unit models,
load models, and the concept of effective load.
80 Power System Planning

3.4 RELIABILITY ANALYSIS FOR AN ISOLATED SYSTEM


We have discussed most of the basic tools needed to perform a reliability analysis of a
proposed generation system expansion strategy. Emphasis will now be placed on
determining LOLP and e(DNS) using material developed to this point. In addition, the
procedural details of including unit maintenance schedules in reliability analyses will be
presented, with a brief discussion of the concept of effective load-carrying capability
(ELCC) of a unit.
Even though we have discussed details pertaining to methods and procedures
used in generation system reliability analysis, we must remember that the goal of the
system planner is to select several expansion plans from perhaps a dozen feasible plans
that satisfy the desired reliability criteria established by an electric utility. Thus the
basic problem is to evaluate week by week the variation in reliability, as new units are
added to supply the growing load, to determine plans that have acceptable reliability
characteristics. For each week in the study period, care must be taken to simulate the
anticipated maintenance schedules, because these schedules drastically influence system
reliability. Carrying out the steps outlined enables the system planner to identify
quickly the expansion plans that are acceptable and should be evaluated and compared
on an economic basis.
In systems in which peak demand is very pronounced and peaking units are used
extensively during peak periods, it is usually necessary to define two load shapes for
each week or month. One load shape, for example, would define the so-called "peak
hours," in which peakers are needed, and the other load shape would define the
off-peak period, in which only midrange and base-load units are required. This
delineation of peak and off-peak periods usually results in a more accurate
representation of the system, desirable in all simulation and especially in studies
involving the generation system.
As indicated, the reliability analysis procedure is the same for every week in the
period studied; only the results differ. Therefore, in discussions to follow, we shall
single out a particular week to analyze, simplifying the task a great deal.

3.4.1 Weekly Effective Load Distribution

Using Eq. 2.52, repeated below,

I
00

f0(81k) = ft(81k - wk)fw(wk) (3.25)


wk=O

the probability density function describing all values the peak demand, 8, can assume
is determined (computation details in Sec. 2.8). Further, from historical data the basic
load shape or shapes (if both peak and off-peak are considered) for the week in
question can be determined and multiplied by 8 1k to produce the conditional
probability distribution F(L/8 1k) required to obtain F(_L), as given in Eq. 3.10, also
repeated below,

------- -----
Generation System Reliability Analysis 81

F(L) = l
8 1k
F(L/0)fe(0) (3.26)

When F(L) determined from Eq. 3.26 is substituted into the convolution
equation when,i = 0,

Fi+ 1(Le) = l Fi(Le - L 0 i)fj(L 0 i) i = 1, ... , G (3.27)


Loi

The result is the probability distribution for the effective load Le- With F(Le) known
and depicted as in Fig. 3.15, the reliability of the system for the week in question can
be evaluated.

3.4.2 Reliability Evaluation

We should recall that the load distribution in Fig. 3.15 tells us the probability that the
load will exceed a particular megawatt value given on the abscissa. Thus, since the
LOLP is by definition the probability that the load (now the effective load because of
unit random outages) will exceed the installed capacity of the system, we need only
evaluate F(Le) at the point Le = IC to get the LOLP for the week. The point
F(Le = IC) is shown in Fig. 3 .15.
By way of example we shall determine the LOLP for our sample system, in
which forecast uncertainty will again be neglected for simplicity. From Fig. 3 .10 and
Eq. 3.27,
F 2 (Le = 100) = F1(100)p 2 + F 1 (100 - 50)q 2
but
F 1 (Le = 100) = F(lO0)p 1 + F(l00 - 50)q 1

1.00r-----....

LOLP

e + IC
Le, MW Le= IC

Fig. 3.15 Typical effective load probability distribution.


82 Power System Planning

and from Fig. 3.10

F(lOO) = 0.0
F(50) = -0.02136(50 - 31.21) + I = 0.5986
F(O) = 1.0

Therefore

F 1 (100) = 0.5986(0.02) = 0.01197


F 1 (50) = 0.5986(0.98) + 0.02 = 0.6067
and finally, the LOLP is
F 2 (Le = 100) = 0.0119(0.97) + 0.6066(0.03) = 0.02981
Thus the probability that the load will exceed the installed capacity of 100 MW of the
sample generation system is 0.02981. Because the number 0.02981 is rather devoid of
meaning by itself, it is common to convert the LOLP into so many hours per week or
days per year. Since the LOLP can be interpreted as the percentage of time the load
exceeded the installed capacity,

LOLP (in h/wk) = 0.02981 X 168


= 5.008 h/wk
or
LOLP (in days/yr) = 10.88 days/yr
Expressing LOLP in hours per week does indeed shed some light on the relative
meaning of the reliability level, so we adopt hours per week as the LOLP index.
The other index of reliability, c(DNS), is beginning to see widespread use because
of its obvious physical significance, and since F(Le) is known, the extra effort required
to calculate the c(DNS) is minor. In Sec. 3.2 we argued that Eq. 3.18 was the
analytical expression for the expected value of demand in excess of IC, which by
definition is c(DNS). Of course, if all units are 100% reliable, c(DNS) = 0. However, if
the random outages of units are represented, the effective load distribution describes
the load, and Eq. 3.18 becomes:
0+IC

c(DNS) =
f
Le=IC
F(Le)dLe (3.28)

which is nothing more than the area under the curve in Fig. 3.15, from Le= IC to
8 + IC as indicated. If Eq. 3.28 is multiplied by T, the time period represented by
F(Le), the c(ENS) results:
Generation System Reliability Analysis 83

e(ENS) =T J 8+IC

Le=IC
F(Le) dLe (3.29)

To complete the analysis of the sample system for the week in question, we can

178.034 f
determine the (DNS) and e(ENS) using F(Le) in Fig. 3.14 and Eqs. 3.28 and 3.29.
128.034

e(DNS) =
f
Le=IOO
F(Le)dLe =
JOO
[0.0494 - 0.00104(Le - 81.21)]dLe

131.213 / 178.034

+
J
128.034
0.0006dLe +
131.213
[0.0006 - 0.0000128(Le - 131.2l)]dLe

= 0.4329 + 0.001905 + 0.0141 = 0.4489 MW


and
e(ENS) = Te(DNS) = 75.42 MWh

3.4.3 Unit Maintenance Schedules


Maintenance schedules can be easily incorporated in the basic reliability analysis
procedure, since a weekly time frame has been used. Generally speaking, normal
maintenance down times vary from 1 wk to a month, and in the case of nuclear
units, where new fuel batches must be loaded in addition to normal maintenance, the
down time can extend to 6 wk or more. During the down time of a unit its capacity is
not available, and the total installed capacity of the system is decreased. A reduction
in installed capacity usually results in a deterioration in reliability; in instances where
reliability drops too much, the system planner must either provide supplemental
capacity or, if possible, rearrange the maintenance schedule. In fact, through the use of
weekly LOlP calculations, it is possible to level the reliability over a particular year by
optimal scheduling of unit maintenance. Leveled reliability maintenance schedules are
attractive because the reliability of one week is not improved at the expense of
another.
One approach we could take to simulating maintenance outages for a particular
week or to analyzing the effect of various maintenance schedules on reliability is to
convolve all units into F(L) except those down for maintenance and evaluate the
resulting effective load distribution at the point corresponding to total capacity of
units not on maintenance. Although this is a straightforward approach, it is not
computationally feasible for systems with a large number of units. Fortunately there is
a way around the problem, and it involves deconvolving from F(Le) those units that
are to be considered down for maintenance. The deconvolution equation may be
obtained easily from Eq. 3.24. For instance, assume that all units are convolved into
F(L) to produce F(Le); then F(Le) may be written as
84 Power System Planning

F(Le) = F-i(Le)Pi + p-i(Le - C;)qi


where p-i(Le) is the load distribution with all units convolved in, except the ith unit.
Therefore
. F(Le) - p-i(Le - q)q;
F-'(Le) = -----
Pi
- - -'- (3.30)

Deconvolving from F(Le) the units to be considered on maintenance is clearly superior


to starting from F(L) and convolving in the units not on maintenance; usually only a
few units are maintained simultaneously.
In some respects, maintenance scheduling is another parameter the system
planner has at his disposal to effect improvements in system reliability for a given
expansion plan. Of course, since the planned timing of maintenance is subject to
changes due to unforeseen system conditions, using maintenance scheduling to
postpone capacity additions is not at all sound. Production planners commonly develop
a beautifully coordinated maintenance schedule, only to have it thrown awry in the
first week by the forced outage of a big unit. However, including maintenance
considerations is essential in developing the best generation system expansion plan, and
as the movement toward large nuclear units continues, it will become even more
important.
An example of the use of Eq. 3.30 would be beneficial at this point, but we
shall briefly postpone it until the concept of the ELCC of a unit has been discussed.
By doing so, we can illustrate both ELCC and maintenance calculations and
interpretations in one example.

3.4.4 Unit Effective Load-Carrying Capability


The ELCC is an interesting and useful idea that provides system planners with a
measure of the relative impact of new units in satisfying system load growth. In
essence, ELCC is that part of the capacity of a unit that is available to supply increases
in demand in order to maintain system reliability less than some desired level.
Typically, when a new unit of capacity C; is added to a system reliability is much
improved, but as demand grows reliability begins to decrease, since both LOLP and
c(DNS) increase. Unfortunately, the amount of load growth that can occur before
reliability returns to where it was before the new unit was installed is less than the
capacity C; of the new unit; this amount of load growth is the ELCC of the unit. The
remainder of the capacity of the unit is required to maintain the desired reliability
level.
As an example, consider the sample system before unit 2 was added, and assume
0 = lJ = 30 MW; the load distribution for this period is shown in Fig. 3.16. The LOLP in
this case is 0.02 (3.36 h/wk), and we assume this value as the upper limit of system
reliability. To improve reliability, the second unit of 50-MW capacity is added to the
system. From the new load distribution, also depicted in Fig. 3.16, we see that the
LOLP drops to an acceptable 0.0006 (0.1008 h/wk). With this new 50-MW unit
added to the system, we would think, a load growth of 50 MW could be experienced
Generation System Reliability Analysis 85

1.000

.::!"' '
I.I..
"O
C:
ro
0.02
' ' ' ' '---\
.::!"' F' (Lel--\
I.I..
'\
0.0006 0 =0 = 30 MW}\
.
C1 = 50 MW _ \
\
c. = 0 c,
0 c, + 0.40 c, + ii IC IC+ ii
c,
Le,MW

Fig. 3.16 F' (Le) and F(Le).

without exceeding the LOIP level of 3.36 h/wk. However, if 8 increases by only 35.66
MW, the LOIP reaches 3.36 h/wk, and hence unit 2 has an ELCC of only 35.66 MW,
not 50 MW; the difference, or 14.34 MW, must be held in reserve for reliability.
Percentagewise, only 71.32% of the new 50-MW unit can be used to supply new load.
If the analysis were repeated with the capacity of the new unit at 100 MW, the ELCC
would become 53 MW, or only 53% of the capacity of the new unit. Since unused
capacity is a luxury most utilities cannot afford, the system planner must avoid
expansion plans in which units cannot be effectively utilized to bring in operating
revenue and maintain the financial posture of the utility; that is, a compromise
between ELCC and unit construction schedules must be made.

3.5 INTERCONNECTED SYSTEMS

Rarely do electric utilities expand systems without eventually physically intercon-


necting with neighboring systems through tie lines of a given voltage level and capacity.
Because of these interconnections, the system planner must include in his reliability
analysis the anticipated changing characteristics of such systems, at least to the extent
that changes will influence the manner in which the system is to be expanded.
The rationale for interconnecting systems is based primarily on the fact that
interconnected utilities can help one another in times of emergency and hence improve
system reliability in effect, without investing in an additional electric plant; that is, the
interconnected utilities share reserves. In Secs. 3.5.1-3.5.4 we shall discuss how
interconnection benefits can be analyzed in terms of improved reliability for two
interconnected systems, starting with the analytical models needed to describe two
interconnected systems.
86 Power System Planning

3.5.1 load and Generator Models


The basic configuration is shown in Fig. 3.17, where LA, LeA, F(LA ), and F(LeA) are
the load, effective load, load distribution, and effective load distribution, respectively, for
company A, and Lr is the tie-line flow from A to B through a tie line whose capacity
is Cr MW. Similarly, LB, LeB, F(LB),'and F(LeB) are the load, effective load, load
distribution, and effective load distribution for system B. Of course, F(LA) and F(LB)
normally include the forecast uncertainty for each system, and F(LeA) and F(LeB)
should include the random outages of the appropriate units not on maintenance within
each system. Further, a weekly time frame is implied as before, thus facilitating the
inclusion of maintenance schedules as well as load diversity. Load diversity is an
important factor in interconnection studies because interconnected systems often have
noncoincident peak demands; thus reserves available for sharing may not coincide with
the timing of need. If monthly or seasonal analyses are performed, for example, it is
possible to incorrectly allow the reserves of system A in the first week of December to
supply the deficits of system B in the last week of that month-an obviously
unrealistic situation. By performing interconnection studies on a weekly basis, the load
diversity problem can be handled better, and the planner may be confident that a
generation system expansion plan is not chosen on the basis of using the December
reserves of a neighbor to supply February's deficiencies.

3.5.2 Interconnected Effective Loads


Once the interconnection between systems A and B is made, the effective loads for the
two become:

LeAr =LA+ Lj
LAoi + Lr
(331)

and

LeBr =LB+ li
LBoi - Lr

(3.32)
= LeB -Lr

Fig. 3.17 Sample intercor,:iected system.

-- -------- -~------
Generation System Reliability Analysis 87

where LeA r = the effective load of system A with the interconnection


Le Br = the effective load of system B with the interconnection
Lr = tie-line flow across the interconnection
We utilize these relationships to obtain the probability distributions F(LeA r) and
F(LeBr), from which LOI..Ps and e(DNS) can be calculated for a given interconnection
capacity, Cr; This procedure will of course require utilizing the joint probability
density function /(Lr, Le A, LeB, Cr). This joint density function reflects fae fact that
the values Lr can assume are dependent on the effective loads in the two interconnected
systems and the tie capacity Cr.

3.5.3 Interconnected Effective Load Probability Distributions


Since F(LeA) and F(LeB) are easily obtained using the ideas discussed in Sec. 3.3, we
must concentrate on the procedure for determining /(Lr, Le A, LeB, Cr) before we
can obtain distributions from which reliability information can. be extracted.
In many respects ~r, LeA, LeB, Cr) will and should reflect the interconnec-
tion policy to be adhered to by member companies. Typically, the following
no-load-loss policy is understood when financially independent utilities share reserves.
1. Company A assists B only to the extent that the reserves of A do not become
negative, resulting in a loss-of-load condition, and vice versa.
2. Company B accepts the emergency reserves of A only when loss of load will
result without them, and vice versa.
An additional factor governing Lr is:
3. The demand transmitted across the interconnection cannot exceed Cr.
From Eqs. 3.31 and 3.32 it is clear that Lr, LeA, and LeB are dependent
random variables whose joint probability density function j(Lr, LeA, LeB, Cr) can be
written as either
(3.33)
or
(3.34)
depending on which system is to be analyzed. Integrating Eqs. 3.33 and 3.34 with
respect to LeBr and LeAr, respectively, we get the probability that the effective load of
B exceeds LeBr given Cr and LeA to be:
F(LeBr, LeA, Cr)= F(LeArfLeA, Cr)f(LeA) (3.35)
F(LeAr,LeB, Cr)= F(LeAr/LeB, Cr)f(LeB) (3.36)
If Cr is assumed deterministic, then R.LeBr) and F(LeAr) are given by the following:

F(LeBr) = 1
LeA
F(LeBr/LeA, Cr)f(LeA)dLeA (3.37)
88 Power System Planning

F(LeAr) = fLeB
F(LeAr/LeB, Cr)f(LeB)dLeB (3.38)

Assuming we are only interested in area B, Eq. 3.37 becomes:

F(LeBr) = f
LeA
F(LeBr/LeA,Cr)f(L~A)dLeA

= f-oc
00 F(LeBr/LeA, Cr)f(ReA)dReA (3.39)

where ReA in Eq. 3.39 is


ReA = ICA - LeA
= effective reserve margin of area A (3.40)
From Eq. 3.40 and the fact that ICA is a constant, we note that
f(ReA) = f(LeA = ICA - ReA) (3.41)
To obtain ALeA ), it is only necessary to differentiate the effective load distribution
F(LeA ):

f(L eA ) = _ dF(L
dL
eA
)
(3.42)
eA

Equation 3.39 can be broken into three separate integrals:

F(LeBT) = Jo-oc
F(LeBr/LeA, Cr)f(ReA)dReA + !CrF(LeBr/LeA, Cr)f(ReA) dReA
O

+ f
Cr
00F(LeBr/LeA, Cr)f(ReA) dReA (3.43)

Since each integral has its own characteristics, we consider each one separately. Further
we need only evaluate F(Le8 r) for values of LeBr ~ IC8 to determine LOLP and
c(DNS). In the first integral it should be noted that because ReA ,;;;; 0, no flow can
occur on the interconnection. Thus

and
F(LeBT) = F(LeB) (3.44)
With Eq. 3.44 substituted into the first integral, it reduces to
Generation System Reliability Analysis 89

F(LeB = LeBT) f 0

f(ReA )dReA (3.45)

The result obtained in Eq. 3.45 could not be simpler, and if LeBr = IC8 , Eq. 3.45
reduces to. the product of LOLPs for areas A and B as determined assuming no
i.I'.terconnection.
The second integral of Eq. 3.43 is valid for values of O <Re A <:;;; Cr, indicating
that Lr= Re A and

and

(3.46)

Equation 3.46 is perhaps misleading: it appears that it should be developed by


convolving F(Len) with f(ReA), which is not the case. Rather, Eq. 3.46 was obtained
by recalling that F(LeBrlLe A, Cr) is the probability that the effective load for area B
interconnected with area_ A exceeds LeBr given (with absolute certainty) that the
effective load of area A is LeA. Again substituting Eq. 3.46 into Eq. 3.43, we have for
the second integral

(3.47)

In using Eq. 3.47, care must be taken to ensure that as LeBr varies over the range
from IC8 to 00 , the proper function F(LeB = LeBr + ReA) is used.
The third and final integral in Eq. 3.43 is valid for reserve margins ReA ;;;,, Cr;
thus

and

F(LeBr/LeA, Cr)= F(LeB = LeBr + Cr) (3.48)


Since Eq. 3.48 is independent of ReA, the third integral reduces to

F(LeB = LeBr + Cr) J"" f(ReA)dReA (3.49)


Cr
To summarize, the effective load probability distribution for area B connected to
area A for values of LeBr;;;,, ICB is:
90 Power System Planning

F(Lesr) = F(LeB = Lesr) f


-~
0

f(ReA)dReA

+ F(LeB = Lesr + Cr) !~


Cr
f(ReA) dReA (3 .50)

In the special case when Lesr = ICs, Eq. 3.50 gives the LOLP for area B connected to
area A; that is,

J
0

F(Lesr = I Cs) = LOLPsA = LOLPs f(ReA) dReA


-~

+F(Les = lCs+Cr) f~t(ReA)dReA (3.51)


Cr

Equation 3.51 is useful in showing the effects that interconnection capacity Cr can
have on reliability. For instance, consider two extreme situations: Cr= 0 and Cr= 00
Clearly from Eq. 3.51 when Cr= 0,
LOLPsA = LOLPsLOLPA + LOLPs(l - LOLPA)
= LOLPs
This result simply states that if there is no interconnection, the reliability of area B
reduces to the value obtained when area B is assumed isolated-not too surprising. In
the other extreme, when an infinite capacity interconnection exists,

LOLPsA = LOLPsLOLPA + !~0


F(LeB = ICs + ReA)f(ReA) dReA

The first term in this expression is the probability that both areas reside in negative
reserve margin states, clearly a loss-of-load situation for area B, and the second term is
the probability that area A resides in a positive reserve margin state but area B resides
in a negative reserve margin state even when assisted by area A.
Generation System Reliability Analysis 91

3.5.4 Reliability Analysis of Interconnected Areas


Because the material discussed in Sec. 3.5.3 was very detailed, we outline here how
reliability analysis of interconnected areas may be performed. Consider the following
procedure:
'
1. Develop the effective load distributions for the two areas considered: F(LeA)
and F(LeB)-
2. Modify F(LeA) and F(LeB) as necessary to reflect known maintenance
schedules.
3. Using Eq. 3.42, determine f(LeA ).
4. Using Eq. 3.41, determine f(ReA)-
5. Calculate LOLPBA using Eq. 3.51, where LOLPB = F(LeB = ICB)-
6. If e(DNS) is desired, first evaluate F(LeBT) using Eq. 3.50 for ICB,;;; Le BT,;;;
ICB + 0, and then determine the area under F(LeBT) from Le BT = ICB to
LeBT = ICB + 0.
Although this approach to interconnected area reliability analysis is a logical
extension of the single-area procedure, it is fair to say that this approach is
complicated and has inherent computational flaws that require some work to avoid.
Nevertheless, it does provide the planner with an approach to multiarea reliability
analysis. In Chap. 6 we shall present another, perhaps more promising, method, which
embeds generation reliability in the overall transmission reliability problem.
To illustrate how the steps above would be implemented in determining the
benefits of an interconnection between two areas, consider two identical areas, both
containing two 50-MW units having FORs q 1 = 0.02 and q 2 = 0.03. Assume that both
areas have load distributions F(L) and effective load distribution F(Le) as shown in
Figs. 3.10 and 3.14, respectively, and further, that they are interconnected via a
Cr= 20-MW tie line. Since these two areas duplicate the single area discussed earlier,
we already know that
LOLPA = LOLPB = 0.02981
Determining f(LeA) using Eq. 3.42, we get the result depicted in Fig. 3.18. In
addition, utilizing Eq. 3.41, we easily obtain the reserve margin density function
f(ReA) shown in Fig. 3.19. With this information and F(LeB), given in Fig. 3.14, the
first term of F(LeBT) becomes:

F(LeB = LeBT) J-~


0

f(ReA)dReA = F(LeB = LeBr)LOLPA

Similarly, the second term of F(LeBT) is:


Cy

1 0
F(LeB = LeBT + ReA )f(ReA )dReA
92 Power System Planning

10p,p,
60 = 1.633
--
.,
,,
<'{

...."'
:,!
- 10
~~ -=-q,q, =0.000128

-- 60

ICA = 100 MW
I I I
(31.210) (78.034) (81.210)_1 (128.034) (131.21) (178.034)
0.40 0 c, + 0.40 c, + o c, + c, + 0.40 c, + c, + ii
Lek MW

Fig. 3.18 Effective load probability density function of system A.

where from Fig. 3.14


F(LeB = Lenr + ReA) = -0.00l04(Lenr + ReA - 100) + 0.02981
100 <,_ L eBr <,_ l 08 .03

and from Fig. 3.19


f(ReA) = 0.00104 0 <,_ ReA <,. 18.78
=O 18.78 < ReA <,_ 20.0 = Cr
Substituting F(LeB = Lenr + ReA) and j(ReA) into the integral, we obtain the
second term for values of 100 <,.Lenr <,_ 108.03. Note that if F(Lenr) is desired for
values of L~nr> 108.03, a different expression for F(_Len =Lenr +ReA) must be
substituted into the integral.
Finally the third and last integral becomes:

F(L,s - L_,,T + Cr) 1-Cr=20


f(R,.)dR,A

or
l0P1P2 -
F(LeB = Lenr + Cr) (0.60) = F(LeB = LeBr + Cr)P1P2
60

------------~-------- -----
Generation System Reliability Analysis 93

Negative reserves Positive reserves

.,

,0q~q 2 = 0.0000128
68

(-78.034) (-31.21) (-28.03) (18.78) (21.96) (68.78) .


-0 -0.40 C2 - 8 c2 - 0.40 I cl + c,. - 0.40
c1 + c,. - o
ReA MW

Fig. 3.19 Effective reserve probability density function of system A.

since from Fig. 3.19

1
00

10P1P2 -
[(ReA) dReA = 60 (0.68) = P1P2
Cr=20

Collecting all the terms, we have


F(LeBT) = F(LeB = LeBT) LOLPA

+ f
0
18.78

0.00104[-0.00104(LeBT + ReA - 100) + 0.0298l]dReA

+ F(LeB = LeBT + Cr)P1P2


which is valid only for 100 <;,,LeBT <;,, 108.03. Since LOLPBA = F(LeBT = ICB) where
ICn = 100, we get
LOLPBA = LOLPBLOLPA + (-0.00052ReA 2 + 0.02981ReA)I (0.00104)
ReA = 18.78
+ F(LeB = 120)(0.98)(0.97)
= (0.02981)2 +0.000391 + 0.0085
= 0.0098 (1.65 h/wk)
94 Power System Planning

Obviously, the interconnection has drastically improved the reliability of area B,


since LOLPB = 5 .008 h/wk and LOLPBA = 1.65 h/wk.

3.6 SUMMARY
In this chapter we have discussed fn much detail the general ideas behind using
reliability analysis, for both isolated and interconnected systems, as a method for
enabling the system planner to determine whether a given expansion plan is
satisfactory in terms of reliability measured by LOLP and e(DNS).
By employing the ideas developed in Chap. 2 with regard to modeling forecast
uncertainty, we were able to describe, using the three-bus system, the importance of
including forecast uncertainty in the load probability distribution. We argued that
forecast uncertainty results in a deterioration in reliability because it tends to inflate the
load distribution.
The notion of an effective load as a means of incorporating the random
availabilities of generating units was presented, followed by the procedural details for
actually determining LOLP and e(DNS) from the effective load probability distribu-
tion. By choosing the effective load approach and a weekly time frame, we showed
that including unit maintenance schedules in analysis using the deconvolution equation
was trivial, although essential for an accurate representation of system conditions.
Further, we introduced the concept of the effective load-carrying capability (ELCC) of
a unit as a measure of the percentage of new unit capacity that could be used to
supply load growth and maintain system reliability below the predefined desired level.
Interestingly, we found that the larger the unit, the smaller its ELCC-a fact to be
considered in identifying the best expansion plan.
The reliability analysis technique used for interconnected systems was described
in depth, with the basic concepts of interconnected effective loads and tie-line
constraints as the foundation of the technique. Through the use of joint probability
density function f(_Lr, LeA, LeB, Cr), we were able to develop necessary analytical
expressions for the interconnected load probability distributions of the interconnected
systems and, from these, to obtain necessary equations for calculating LOLP and
e(DNS). Through the analysis of a detailed example, using identical three-bus systems,
the benefits to be derived from an interconnection were illustrated; the reduction in
LOLP was verification that sharing reserves is sound and effective.

PROBLEMS
3.1. A system contains three generating units having the following characteristics:
Unit 1 Unit 2 Unit 3
I. Oil 1. Coal 1. Hydro
2. C1 = 150 MW 2. C2=150MW2. C3=lO0MW
3. qi = 0.02 3. q2 = 0.03 3. q3 :::: 0
Calculate the outage capacity density function for the system, and determine the
probability of having more than 200 MW of capacity unavailable.

-
Generation System Reliability Analysis 95

3.2. Calculate the load shape, e(L) and e(E), for the load distribution given below:

FILI

210MW 350MW

3.3. Combine the unit and load models in Probs. 3.1 and 3.2, using the convolution
equation, and calculate LOLP and e(DNS) for the system.
3.4. Calculate the changes in LOLP and e(DNS) if unit 2 in Prob. 3.3 is removed for
maintenance. Use the deconvolution equation to remove unit 2 from F(Le).
3.5. Using LOLP calculated in Prob. 3.4 with unit 2 removed as a desired reliability
level, calculate how much load growth unit 2 can support without violating the
desired reliability. For this problem, assume that each week has the same load
shape determined in Prob. 3.2.
3.6. Rework the two-area problem in Sec. 3. S, assuming that Cr = 5, 10, and 15 MW.
Plot LOLPnA versus Cr. Calculate the value of Cr above which no further
improvement in reliability can be obtained.
3. 7. Develop a computer program to evaluate the reliability of isolated areas
containing up to five generating units. Check your program by reworking Probs.
3.1 to 3.3.

BIBLIOGRAPHY
1. AIEE Committee Report: Application of Probability Methods to Generating Capacity
Problems, AIEE Trans., vol. PAS-79, pp. 1165-1182, 1960.
2. Billinton, R.: "Power System Reliability Evaluation," Gordon and Breach Science Publishers,
New York, 1970.
3. Billinton, R.: Bibliography on the Application of Probability Methods in Power System
Reliability Evaluation, IEEE Trans., vol. PAS-91, pp. 649-660, March-April 1972.
4. Billinton, R. et al.: "Power System Reliability Calculations," MIT Press, Cambridge, Mass.,
1973.
5. Booth, R. R.: Power System Simulation Model Based on Probability Analysis, IEEE Trans.,
vol. PAS-91, pp. 62-69, 1971.
6. Booth, R.R.: Optimal Generation Planning Considering Uncertainty, IEEE Trans., vol. PAS-91,
pp. 70-77, 1971.
7. Cook, V. M._ et al.: Determination of Reserve Requirements of Two Interconnected Systems,
IEEE Trans., vol. PAS-82, pp. 18-33, 1963.
8. Day, J. T. et al.: Expected Value of Generation Deficit: A Supplemental Measure of Power
System Reliability, IEEE Trans., vol. PAS-91, pp. 2213-2223, 1972.
9. Garver, L. L.: Reserve Planning Using Outage Probabilities and Load Uncertainties, IEEE
Trans., vol. PAS-89, pp. 514-521, April 1970.
10. Garver, L. L.: Adjusting Maintenance Schedules To Levelize Risk, IEEE Trans., vol. PAS-91,
pp. 2057-2063, 1972.
11. Knight, V. G.: "Power Systems Engineering and Mathematics," Pergamon Press, Elmsford,
N.Y., 1973.
12. Pang, C. K. and A. J. Wood: Multi-area Generation System Reliability Calculations, IEEE
Trans., paper T74342-2, presented at Anaheim, Calif., 1974.
13. Probability Analysis of Power System Reliability, IEEE Spec. PubL 71M30PWR, 1971.
14. Vassel, G. S. and N. Tibberts: Analysis of Generating-Capacity Reserve Requirements for
Interconnected Power Systems, IEEE Trans., vol. PAS-91, pp. 638-649, March/April 1972.
GENERATION SYSTEM
COST ANALYSIS

In Chap. 3 the basic philosophy used in evaluating the reliability of a number of


alternative expansion plans was discussed. Using LOLP and c(DNS) as indices of
reliability, plans that did not comply with the desired reliability criteria were assumed
to be either modified or discarded. The remaining plans must be evaluated on the basis
of economics in order to identify the one plan that impacts on the utility as a whole
in the most favorable manner. We discuss in this chapter the concepts used in
analyzing the cost of a particular expansion plan, with most emphasis given to a
conventional present-worth arithmetic method. However, for completrness we include a
brief presentation of corporate models as they are used in making a final selection.
Since corporate models are computerized representations of the financial structure of a
utility, a discussion in depth is well beyond the scope of this book.
As we shall note in discussing cost analysis, the need for accurate production
cost information is essential in obtaining an accurate expansion plan cost result.
Because of this fact, we shall expend considerable effort discussing a probabilistic
production-costing procedure which easily handles forecast uncertainty, unit outages,
maintenance schedules, etc. Further, we shall emphasize how both nuclear and hydro
unit constraints may be simulated to portray actual operating conditions accurately.
96

,
--------- ---------- - - -- ---------------~--------------
Generation System Cost Analysis 97

Both nuclear and hydro units (conventional and pumped) create simulation difficulties,
primarily because of their somewhat rigid fuel cycle characteristics.
Because in many parts of the world energy transactions and off-peak loading are
common, a modest section is devoted to these topics. In particular we shall focus on
problems associated with simulating these two operating conditions. A presentation
describing the impact of environmental regulations on generation system costs
concludes the chapter. In addition, a brief look at a method for ambient air quality
assessment is included; the procedure used is a rather simple extension of that used in
production cost analysis.

4.1 COST ANALYSIS


In evaluating the cost associated with a particular expansion plan, it is essential that
the following be considered:
1. Capacity cost
2. Production cost
3. Timing of additions
These three quantities, more than any other, influence the overall expansion plan cost
and hence must be taken into account. As these quantities are so crucial, we shall
detail each of them.

4.1.1 Capacity Cost


The capacity cost of unit i, denoted by CC;, is usually defined as follows:
CC,= FCR; UC;C; (US$) ( 4.1)
where FCR; = fixed charge rate
UC;= unit capacity cost (US$/MW)
C; = capacity in MW
By definition, the fixed charge rate associated with investment in unit i is the annual
expense associated with the investment expressed as a percentage of initial investment.
Generally the fixed charges consist of
1. Depreciation
2. Rate of return
3. Taxes
4. Insurance
Although there are other charges, the ones cited are by far the major ones. The
importance of FCR warrants an example illustrating, albeit in oversimplified fashion,
how FCRs are determined.
Assume that we are thinking of building a new 2,000-MW nuclear plant
containing two identical 1,000-MW units and that the unit cost for each is
US$500/kW. Further assume that the plants are to be depreciated over 10 yr and that
the capital needed is to be obtained by selling US$500 million of common and
l8

TABLE 4.1 Analysis of fixed chargesa. b


Net Return
investment Depreciation PWFof Interest Tax plus After tax Income PWFof
Year (book value) charge Return return at 10% on debt depreciation depreciation income'" tax !Tat 10%
l 1,000 100 100 90.909 40 100 200 60 90 81.8
2 900 100 90 74.380 36 100 190 54 81 66.9
3 800 100 80 60.105 32 100 180 48 72 54.1
4 700 100 70 47.811 28 100 170 43 64.S 44.0
5 600 100 60 37.255 24 100 160 36 54 33.5
6 500 100 50 28.224 20 100 150 30 45 25.4
7 400 100 40 20.526 16 100 140 24 a 36 18.5
8 300 100 30 13.995 12 100 130 18 27 12.6
9 200 100 20 8.482 8 100 120 12 18 7.6
10 100 100 10 3.855 4 100 110 6 9 3.6
385.542d 348.oe
a Assumptions: (1) straight-line depreciation-book and tax; (2) life of plant-IO yr (for tax and book); (3) capitalization-SO% debt at 8%; (4)
income tax (IT) rate-60%; (5) zero net salvage; (6) rate of return-10%; (7) no investment tax credit; (8) local and state taxes-1.6%; and (9)
insurance-0 .6%.
bTotal fixed charge rate: return= 6.3%; tax (IT)= 5.7%; depreciation= 10.0%; local and state tax= 1.6%; insurance =0.6%; total= 24.2%.
c After tax income = equity cost= income - income tax; income tax= 0.6 income.
dEquivalent uniform annual amount: (CRF 10%, 10 yr)(385.542) = 62.745; percentage of total cost= 62.745/1000 6.2745%. =
eEquivalent uniform annual amount: (CRF 10%, 10 yr)(348) = 56.6; percentage of total cost= 56.6/1000 = 5.7%.

:~.;~~ .' ,v,t-.;~G.1;.:~1.:. .ff.~ ~.:i~

.:!!
'!' ...
o
N
o
US$/kW
w '" u, o,
al !:fl
:;:i.o
!J",
_.
~"'~8o'st-&.s~
3-~::,~1~
O Oc-.~-;_;o!!!.
-
s(l!I
<oEt-~~~Et-~-~~'t:I
o-o><.Ooo<.lO'l_,.,;:r,_,

... ~ ~ .l lWLi.t lJ. ~ l. li ~ ~-,M~~i4t! ~ p~ t


:.. ~ ~ ..,. C o ..., ~a w, o O
~gn~
,.,
I
0
~=c
"" I -
lff>>o< ..... =:,"";-.-,-"'OUJ""'a"'l0.
oa:..,
::,10C:::31><a; .... - - .
~
~ -, ~&a~~&oao-~~p,.-~1[-~
.2.
g
-
o 1'<-0~~1-::,..,.
..,. i"O ~'t:I -::r- ~ P . . l : : , ~ ~ p , .
-11~P...::,

" ~-~= -o-co1c~~-~ w


Generation System Cost Analysis 99

preferred stock (equity capital) and US$500 million of 8% bonds (debt capital), where
the return on equity must be 12% based on the projected appreciation of stock and
dividends. If the income tax rate is 60%, insurance is 0.6%, local and state taxes are
1.6%, and if straight-line depreciation and no salvage value are assumed, the analysis of
FCR is as shown in Table 4.1. All entries are in US$ millions, and PWF and CRF are
the standard' present worth and capital recovery factors. Table 4.1 is self-explanatory,
except that column 10 is obtained by realizing that column 9 is the return on equity
required after taxes; therefore, the entries in column 9 are 40% (assuming 60% income
tax rate) of the net taxable income. Multiplying net taxable income by 60% renders
the taxes in column 10. Summing depreciation, rate of return, taxes, and insurance, we
get that the FCR is 24.2%, meaning that the two 1,000-MW nuclear units represent an
investment of (0.242) (US$1 billion)= US$242 million per year over the next 10 yr.
Unit capacity cost is self-explanatory; however, it is important to be aware of the
relative capacity cost for various types of units. To emphasize the relative differences
in capacity cost for various types, Fig. 4.1 is provided. Not only is there a significant
differential in capacity cost, but economies of scale are evident. Note, for example,
that a 1,000-MW base-load fossil unit is 16.6% cheaper than two 500-MW base-load
fossil units. It is not surprising, given these facts, that the utility industry has
continually pushed for larger and larger units. If the basic. curves shown in Fig. 4.1
were escalated at a rate of 5% per year, then we would find the unit capacity cost for
a 1,000-MW nuclear unit in year 1990 at US$805 /kW, a figure that is probably
somewhat low.

4.1.2 Production Costs


Production costs, the second important component in evaluating the cost of a
particular expansion plan, can be accurately determined only if

700

600
'
"
Nuclear base load

-
..............
500
---...
s 400
.>t. ....... ~ b a s e load
~

-
(/)
:) 300
Midrange
200 ~

Peaking
100

0
500 1000 1500 2000
Unit size, MW
Fig. 4.1 Projected unit capacity costs (1982 operation).
100 Power System Planning

1. A realistic load model is known for each future week or month in the
planning period
2. The units are committed to supply load in a manner that reflects actual
operating procedures and conditions

On the basis of material presented in Sec. 3.2.2, it is obvious that including


forecast uncertainty is the key to having a realistic load model. Fortunately, we have
already discussed how a probabilistic load model can be developed, so we need not
pursue this further. In short, we shall use the load model described in Chap. 3. To
obtain a realistic commitment schedule for the generating units, it is necessary to
include not only the forced outages of a unit but also its scheduled outage for
maintenance-ideas penetrated in sufficient depth in Sec. 3.4.3. What we have not
discussed is how to commit units to simulate normal economic dispatching procedures;
that is, we must now concern ourselves with the order in which units are convolved
into the load distribution, for this order dictates the amount of energy each unit
produces. Much more will be said of this in Sec. 4.2; suffice it here to say that the
production cost associated with unit i is given by

EC; = FC; + OM; (4.2)


where EC; = production or energy cost in US$
FC; = fuel cost in US$
OM; = operations and maintenance cost in US$
Because future fuel costs are very difficult to predict, utilities must consider a
range of fuel costs in economically evaluating expansion alternatives. Table 4.2 shows
fuel cost ranges, by fuel type, that are considered to contain at least the fuel costs
utilities will have to pay in 1982. The costs shown may be drastically incorrect, as it
requires only a minor imbalance in supply and demand of any one fuel to irivalidate
the numbers given. For use later, typical operating and maintenance (O&M) costs for
various plant types are given in Table 4.3.

4.1.3 Plant Cost


As it is sometimes convenient to calculate the cost associated with a particular plant,
we can easily define an expression for plant cost (PC) in terms of Eqs. 4.1 and 4.2:

TABLE 4.2 1982 fuel price range for


the United States (unofficial estimates)
Type Range

Nuclear 0.2-0.4 US$/MBtu


Coal 36-84 US$/ton
Oil 15-27 US$/bbl
1. Cost of pumping if unit is
Hydro pumped-storage
2. Zero if no pumping required

,,.
Generation System Cost Analysis 101

TABLE 4.3 O&M costs for typical units


Fixed costa Variable cost
Type (mils/kWh) rated (mils/kWh)
Base load (- 1 ,000 MW)
Nuclear 0.60
Fossil coal 0.48
None
Fossil oil 0.34
Fossil gas 0.22
Midrange
Fossil steam l 0% less than base load
Combined cycle 0.56 0.5
Peaking
Simple cycle 0.64 0.75
Regenerative 0.6 0.85
a Assuming CF= 75% for base load and midrange and 500 h operation for
peaking units.

PC; = CC; + EC;


= CC; + FC; + OM; (4.3)

Further, if Eq. 4.3 were divided by the expected value of the annual energy produced
by unit i, denoted by e(E;), then the plant cost might be expressed in mils/kWh or
US$/MWh, which is often convenient when comparing different plant types. If PC; in
mils/kWh or US$/MWh is desired, then Eq. 4.3 can be used, but now

PC
PCi-+ e(E;) mils/kWh

CC; FCR;UC;
CC-+
1 - - - ---,---- mils/kWh
e(E;) - 8760e(CF;)
(4.4)
UFC-HR-e(E-)
1 1
FC; -+ e(E:) = IFC; mils/kWh

OM
OM;-+ e(E;) mils/kWh

We shall not define new variables for each of these new quantities; it should be
understood that Eq. 4.3 can be expressed either in dollars or in mils/kWh. For rough
approximations of PC; (mils/kWh) we should realize that
e(E;) = 8760e(CF;)C;
- 8760CF;C; (4.5)
The expression for the incremental fuel cost will be derived in Sec. 4.4.2.
IFC;:::: UFC;(US$/MBtu) HR;(MBtu/MWh) (4.6)
102 Power System Planning

where CF;~ assumed or representative capacity factor= e(E;)/8760C;


HR;~ heat rate in MBtu/MWh
Strictly speaking, Eq. 4.6 is correct only if the heat rate of the unit is constant, which
is never the case. However, for crude studies Eq. 4.6 is frequently used.
In summary, Eq. 4.3 can be used to obtain accurate plant costs expressed either
in dollars or in mils/kWh if the annual e(E;) produced by each unit is calculated using
a very detailed production analysis in which all normal operating conditions are
simulated. For quick but approximate plant cost results expressed in mils/kWh, Eq. 4.3
may be used, but each term must be replaced as defined in Eq. 4.4. Equations 4.5 and
4.6 are used to calculate e(E;) and UFC;, given an estimate of the assumed capacity
factor CF; and heat rate HR;.

4.1.4 Timing of Unit Additions

The third and final major factor in cost analysis is the timing of unit additions. Since
present-worth arithmetic is employed to determine the present worth of a given
expansion plan, it is obvious that timing the investment associated with adding a new
unit is important. Further, the effect of new unit additions on total system production
cost is a factor that should be reflected in the total present worth of a particular
expansion plan. Although this may not be obvious, we should realize that the addition
of a new unit can drastically change the operation of existing units and hence
production cost. For instance, if a base-load fossil unit were added prematurely to a
system containing base-load nuclears, the tendency to off-load the nuclear units could
result in substantially higher production costs.
We shall specifically include timing of unit additions in Sec. 4.1.5, when the
overall approach to a detailed system cost analysis is considered. When an approximate
cost result is desired for a particular plan, it is not uncommon to assume that all the
units are added simultaneously, in which case the system cost SC (mils/kWh) can be
calculated as follows:
SC = "i,ai PCi mils kWh (4.7)
t,
where ai = e(Ej)/e(Es)
e(Ej) = expected value of energy produced by unitj
~ 8760 CF;C;
e(Es) = expected value of energy consumed by the system
~ 8760 CFsCs
Substituting the approximate expressions for e(Ej) and E(Es) into Eq. 4.7, we get

\"' CFi q
SC= LCFs Cs PCi (4.8)
j

In choosing representative capacity factors for each unit it is necessary to satisfy the
constraint that
Generation System Cost Analysis 103

"L CF;;
C = CF3 (4.9)
. s
l

which comes from the fact that the sum of energies produced by the individual units
must equal the energy consumed by the system. Equation 4.8 is very convenient in
that given the plant costs, PC;, the effect of generation mix on system cost can be
easily investigated, where the ratio C;/C3 times 100 is the percentage of capacity of
type j being added to the system. To illustrate the use of Eq. 4.8, consider the
problem of determining the system cost for an expansion plan containing 75% existing
fossil, 20% new nuclear, and 5% new fossil peaking capacity. Further assume

CFp =? CFs = 0.765 CF1= 0.8


UCp = 150US$/kW CFn = 0.8 UCr = 300 US$/kW
FCRp = 0.17 UCn = 600 US$/kW FCR1 = 0.17
UFCp = 9 US$/bbl FCRn = 0.17 UFC1 = 40 US$/ton
HRp = 15 MBtu/MWh UFCn = 2 US$/MBtu HR1 = 12 MBtu/MWh
OMp = 1.39 mils/kWh HRn = 10 MBtu/MWh OM1 = 1.0 mils/kWh
OMn = 0.6 mils/kWh
From Eq. 4.9

0.05 CFp = 0.765 - 0.8(0.75) - 0.8(0.20)


CF = 0.765 - 0.6 - 0.16 = 0 l
P 0.05 .
Calculate PC; using Eq. 4.3:
p _ (0.17)(300)10 3 (40)(12)
Ct - 8760(0.8) + 24 + l .O
= 7.2 + 20 + 1.0
= 26.8 mils/kWh
PC _ 0.17(150)10 3 9(15)
P - 8760(0.1) + 6 + 139
= 29.1 + 22.5 + 1.39
= 53.0 mils/kWh
(0.17)(600)10 3
PCn = 8760(0.8) + 2(10) + 0.6
= 14.5 + 2.0 + 0.6 [:
j;!

= 22.5 mils/kWh

(!
104 Power System Planning

Finally, from Eq. 4.8


6
SC = ~:~; (53.0) + / 65 + 7 ~S (26.8)
= 26.07 mils/kWh
'
By repeating the example above with different percentages on the capacities, the
planner can gain some insight into the relative effects of generation mix on system
costs. If pumped hydro were to be considered, then the fuel cost would be determined
based on the cost of pumping. For conventional hydro plants, fuel costs are basically
zero, but because of the nature of the world's hydro resources, many industrailized
nations no longer have hydro resources to develop.
We leave this section by reiterating that for detailed, accurate cost analysis, Eq.
4.3 must be used in conjunction with good production analysis results. For
back-of-the-envelope analysis, ignoring timing of new additions and assuming reasonable
unit capacity factors produces useful results.

4.1.5 System Cost Analysis


In this section we discuss an approach to system cost analysis that relies on production
cost data determined using a simulation method that portrays the way units will
probably be operated if the plan is implemented. Although the details of the
probabilistic production cost simulation will be postponed until Sec. 4.2, we shall
assume that accurate production costs for each unit for each year in the planning
horizon can be obtained. The method to be discussed considers the time value of
money in that the timing of unit additions is taken directly into account using
conventional present-worth arithmetic. Also, the effect of new additions on total
system production cost is appropriately factored into the cost estimates.
Specifically, the method involves first calculating the annual capacity costs for
each year and then multiplying this result by the appropriate present-worth factor to
obtain the present worth of the annual capacity cost for each year. Summing the
present worth of the annual capacity cost over all years in the planning horizon gives
the present worth of the total annual capacity costs for the plan under study. To
quantify this procedure, let CC,{k) be the annual capacity cost for unit i in year k, and
let {3k be the present-worth factor for year k. The present worth of the total annual
capacity costs in dollars is given by
N G
CC = ll
k=I i=J
/3kCC;(k) (4.10)

If unit i =m is not installed by year k, then CCm(k) = O; otherwise CCm is given by


Eq. 4.1.
Similarly, let ECk be tht total system energy cost, as defined by Eq. 4.2, for
year k. Note that this cost in dollars includes the production cost, not only for new
units, but also for existing units. By incorporating the total system production cost,
Generation System Cost Analysis 105

the effects of n~w unit additions on the operation and production cost of existing
units are easily taken into account. By multiplying ECk by (3k, the present worth of
the annual system production cost is obtained, and the sum of all present-worth annual
production costs renders the present worth of the total production costs incurred over
the planning horizon:
.,

Of course ECk is determined using the probabilistic production-costing procedure


alluded to earlier.
Finally, with CC and EC added, the present worth of all new capacity and
system production costs incurred over the planning horizon is obtained for the
expansion plan being analyzed. It is this total present-worth cost that is used to
compare alternative expansion strategies for the purpose of reducing further the
number of plans to be subjected to full-scale corporate analysis.
As we have stated, there are nearly as many approaches to generation system
cost analysis as there are utilities. What we have done here is propose and discuss a
representative method. The intent has not been to discuss costing procedures
exhaustively so much as to unify the material and motivate the discussion of
probabilistic production cost analysis presented in Sec. 4.3.

4.2 CORPORATE MODELS


A corporate model is, as the name implies, a model of the financial structure of a
company. Through the use of such models a system planner can simulate the effect
that a major expansion effort will have on the financial status of his company.
Although such models are very detailed, since most reflect the accounting procedures
used by a utility, we shall gloss over many accounting procedures that are beyond the
scope of this work. We shall instead discuss in a simplified manner the basic structure
and input-output requirements.
If we refer to standard statistical data published by most electric utilities, a very
crude and oversimplified model of the cash flow in a utility can be obtained. For
instance, if the "summary of earnings" and "statement of sources of funds used for
construction" are translated into a standard block diagram, as used in classical control
theory, a diagram like Fig. 4.2 will result.

4.2.1 Input-Output Parameters


In effect we have in the corporate model an open-loop multiple input~multiple o~t
system where the major inputs are:
1. Forecast energy and customers by rate class for the months being simulated
2. Construction expenditures for new plants, transmission facilities, etc., for each
month
106 Power System Planning

***Carrying charges for nuclear


fuel recovered here
Preferred and
Other inc. ***Interest common dividends

Gross Retained
income earnings

Cl~ T & Dcust. * *Production cost and Earnings


Gen.units acc. sales& purchased energy
adm. Yield
1st cores
**Nuclear fuel costs excluding
carrying charges appear here FCC rate

E/S
Deferred income
Depreciation tax

Special project costs

Construction cost Retained earnings


project
costs
p referred l
Common5

i
Outside$ required Bonds
Short-term
loans

Fig. 4.2 Simplified corporate model.

3. Production costs for each month


4. Purchased energy costs
5. Costs associated with transmission and distribution, customer accounts, and
sales and administration
Similarly the major outputs are:
1. Gross income or operating income
2. Retained earnings
3. Earnings per share
4. Fixed charge coverage rate
5. Yield
6. Outside capitalization requirements
Clearly, such a planning tool is both convenient and necessary. To study the
impact of a given expansion plan on any one of the six outputs, it is only necessary to
input the initial cost of each plant and an expenditure-time curve, which defines the
rate at which money for the plant is to be spent, along with the other four inputs
listed above. Production costs and purchased energy costs for each month must be
determined, as we have said before, using detailed production analysis techniques.
Generation System Cost Analysis 107

The output of a generation system expansion study should provide the planner
with enough information to determine
1. How _much outside capital is needed and if it can be obtained
2. Whether the yield, earnings, etc., are maintained at sufficient levels or if rate
relief is necessary to make the company attractive to investors
Referring again to Fig. 4.2, we see that operating revenue is dependent on energy
and on customers in addition to rate structure. Since energy consumption and
customers are beyond the control of the utility, the only mechanism for adjusting
operating revenue is rate relief, in which the basic rate schedule is altered under the
regulatory controls of a public service commission. Altering operating revenue
obviously results in a change in gross income, since gross income is simply operating
revenue minus total operating expenses. The amount an electric utility is allowed to
increase its gross income is a direct function of its allowable rate of return, as specified
by a public service commission. Usually the upper level on gross income is determined
by multiplying allowable rate of return by the rate base of a company. The rate base
in essence is the total assets of the company committed to the production,
transmission, and distribution of electric energy. Allowable rate of return is that
percentage of the rate base that must be available for adequate payment of the cost of
capital made available by investors. It should be clear from Fig. 4.2 that excessive
interest charges without supporting operating revenue, a typical situation when plants
are being constructed and are not yet in the rate base, can result in low earnings and
perhaps no construction capital, and hence no expansion. The most common cure for
problems of this nature is rate relief. Low earnings can also be created by energy
conservation-minded customers as well as by excessively high production costs due to
supply and demand forces in the international fuel market. It is an understatement to
say that choosing the optimum expansion plan is not at all easy and is sometimes
based on educated guessing of what future conditions will be.
The major operating expenses, as shown in Fig. 4.2, are production and
purchased energy costs, followed by transmission and distribution and customer
accounts expenses, with sales and general administrative costs following close behind.
In addition, depreciation, taxes, and maintenance are also expense items covered by
gross income. Depreciation is not a real expense, except for tax purposes, since it is
subtracted to obtain (regulated) gross income, but added in again before outside
capitalization requirements are determined. By handling depreciation in this manner,
the utility is in effect able to establish rate schedules that will make it possible to
replace a plant after it is retired. Note that construction costs are not considered
operating expenses, with the exception of the construction of nuclear fuel cores; the
reason is that while they are under construction, new plants, transmission facilities, and
so on are not available for energy production and hence, by definition of rate base,
cannot be considered in determining allowable rates of return or in setting levels of
gross income.
Of all the expenses normally associated with an electric utility, we shall discuss
in more detail production costs as well as energy purchases and sales. The reason for
108 Power System Planning

singling out these particular expenses is that production and energy purchases and sales
analysis are more clearly the responsibility of the system planner, whereas depreciation,
taxes, sales, and administration are peripheral for us, although still important. Further,
production analysis will render information normally used in calculating S) stem costs
as well as future fuel requirements.

4.3 PRODUCTION ANALYSIS


The discussions pertaining to system costs and corporate modeling have clearly
indicated the role production analysis plays in generation system planning. In system
cost analysis it is necessary to know the expected value of the annual energy each unit
will produce, as well as the associated production costs. Similarly, in corporate model
studies performed on a monthly basis, a detailed account of monthly production costs
including energy purchases or sales is necessary. Further, because of sometimes lengthy
delays in fuel procurement, it is essential for an electric utility to have the capability
of determining anticipated fuel inventories. In the case of both nuclear and hydro
generation, fuel management is quite complex. Since these two sources of energy must
be utilized in accordance with rigid refueling and hydrological constraints, production
analysis of mixed generation systems is quite involved.
Because of the number of topics we must discuss pertaining to production
analysis, it is convenient to complete this section by briefly discussing the generator
and load models to be used in production analysis, including interconnection models
and load uncertainty. The remaining sections of this chapter will then be devoted to
production costing, fuel inventories, energy purchases and sales, and finally environ-
mental costs.

4.3.1 Generation System Models


We are fortunate at this juncture to have discussed in Chap. 3 how unit forced outages
can be modeled assuming two-state models for the units. All the ideas developed earlier
apply unchanged to production analysis; that is, a unit with an FOR of q can be
modeled by a fictitious unit that is always available in parallel with a fictitious load
whose availability is q. Since we have already argued the validity of this representation,
we need not dwell on the topic.

4.3.2 System load Models


We have already developed and discussed the methods and ideas associated with
modeling system loads in Secs. 3.2 and 3.3. Most important, we showed how the
probability distribution for the effective load of a system

Le= L + lLo;
i

can be obtained by convolving F(L) with f(L 0 ;) using the convolution Eq. 3.23.
Further, the technique was discussed for including forecast uncertainty, as defined by
Generation System Cost Analysis 109

the probability density function 1(8), to reflect more accurately the uncertainty
associated with the load forecast.
The system model with an interconnection to a neighboring company was
developed in Sec. 3.5. We shall use this model to simulate energy purchases and sales
transactions at a rate of Lr MW. With an interconnection of capacity Cr MW, the
effective load was defined as

Lesr = Les - Lr
When Lr;;,,, 0, energy is being purchased; when Lr,( 0, energy is obviously being sold.
The probability distribution for Lesr was developed in Sec. 3 .5 and is given in Eq.
3.39. In production analysis it is normally assumed that Lr is an independent variable
since interchange agreements are only honored when the reserves of the seller are
adequate to satisfy the firm agreement. Thus, LeA (A being the seller) and Lr are
independent, and hence LeA and Les are independent. With this assumption, Eq. 3.39
reduces to
F(Lesr) = F(Les = Lesr Lr) (4.1 I)
Based on Eq. 4.11, to obtain F(Lesr ), assuming system B is purchasing energy at the
rate of Lr MW, it is only necessary to shift F(Les) to the left by Lr MW. Similarly, if
B is selling energy at a rate of Lr MW, then F(Lesr) is obtained by shifting F(Les) to
the right by Lr MW. As we consider examples in subsequent sections, the validity of
the shifting of F(Les) will become intuitively clear.

4.4 PRODUCTION COSTING


As we discussed earlier, it is essential for corporate model studies as well as for system
cost studies to have for a particular expansion plan the. expected value of the energy
produced by each unit, (;), as well as associated production costs. Production costs
generally include both fuel, FC;, and O&M, OM;, costs. To determine e(EJ, FC;, and
OM;, it is necessary first to forecast what the load will be for each week in the period
being studied and to construct the load distributions F(L) using the forecast data and
load shapes as discussed in Chap. 3. With F(L) known for each week, F(Le) must be
determined. However, since we are now primarily interested in obtaining accurate
production data for each unit, the order in which the units are convolved into F(L)
must be such that the corresponding production cost is a minimum. This particular
order will be referred to as the economic commitment schedule (ECS). We must insist
on convolving units in this order if we are attempting to simulate the way units will
actually be used if the expansion plan is indeed implemented. Adhering to the ECS
will render reasonably accurate energy, production costs, and fuel inventories to be
used in selecting the best expansion plan and negotiating fuel contracts.

4.4.1 ECS Unit Models


In Chap. 3 the full capacity of each unit was convolved into F(L) because we were
only interested in LOLP and e(DNS), which are independent of the way units are
110 Power System Planning

convolved into F(L). Now we are interested in ordering the units to achieve the
smallest production cost, and therefore it is necessary to segment the capacity of each
unit into three segments (although more can be used, three has been shown to be
adequate) because in real life it is rarely economical to commit one unit completely
before calling on another. Typically, the ECS would commit lower segments of all
large units before commiting any particular unit' completely.
The reason for segmented commitments is the basic shape of heat rate curves; a
typical heat rate curve is shown in Fig. 4.3. Clearly on this curve the second segment
corresponds to higher efficiency, since fewer Btus are required for each MWh of energy
produced. From this basic HR curve the input/output, I/0, curve can be obtained by
multiplying every y-axis value by its corresponding x-axis value; a typical I/0 curve is
depicted in Fig. 4.4. Finally, differentiating the I/0 curve, we obtain the incremental
heat rate, IHR, curve in Fig. 4.5, which will be used a great deal in later discussions.
To quantify these relationships for future use, we have for unit i

and
dl/0;
IHR;(L;) = dL _
l

In the special case when the HR curve is assumed to be constant, HR;(L;) = HR;, then
dL;
lHRt(L;) = HR;dL-1
= HR;
This special case is important because in many economic studies the assumption that
HR;(L;) is constant is made to facilitate computing incremental fuel costs using Eq. 4.6.

4.4.2 Economic Commitment


In convolving units into F(L) to obtain the ECS, units are usually grouped into three
basic groups: base-load, midrange, and peaking. Because base-load nuclear units are

MBtu MBtu
MWh
h

Segment 1lsegment 2lsegment 3


I
MW MW

Fig. 4.3 Typical heat rate curve. Fig. 4.4 Typical input/output curve.
Generation System Cost Analysis 111

MBtu
MWh

Fig. 4.5 Typical incremental


heat rate curve.

MW

usually committed first, base-load fossil next, and so on, the grouping simulates a
coarse commitment schedule.- However, within a particular group the commitment of
the segments must be determined on an incremental cost basis; that is, segments with
the lowest incremental cost are committed first. We shall attempt to expose the
analytical details associated with determining ECS, e(Ei), and finally the production or
energy cost for each unit, e(ECJ.
If F(L) in Fig. 4.6 describes the load for a particular month in the period under
study, then after the ith segment of the nth unit is economically committed, the
distribution is depicted by the dotted curve pin(Le), in Fig. 4.6. The shaded area
indicates the position of the segment in pin(Le). In determining which segment should
be committed next, two basic possibilities exist: either the second or third segment of
one of the units already committed can be committed, or the first segment of a new
unit can be committed, where the segment chosen should be the one with the smallest
incremental cost. For purposes of illustration and for later use, we shall determine the
segment energies and incremental cost expressions for each of the two situations
described. By considering these two cases we can penetrate the concepts and introduce
the analytical details that the reader needs to understand.

Segment
in

C
Lor Le, MW

Fig. 4.6 F(L) and F(Le) after one segment is committed.


112 Power System Planning

First, consider determining the incremental cost, etc., for the first segment of
unit m, whose capacity is C1m, forced outage rate and availability are qm and Pm, and
incremental heat rate is IHRm(L 1m)- The effective load distribution after committing
segment Im, F 1m(Le), is

F'"' (L,) ~ 1- 0
F, m(L, - Lo, m)[(Lo,: )dL 0 <m ( 4.12)

where F 1111 is the load distribution before segment Im is committed andf(L 01 m) is the
outage load density function for the first segment. It is obvious that

Pm for L 01 m = 0
f(Loim) = C
qm for L 0 1m = Im
since a segment outage can occur only if the unit is forced off-line with the probability
q m. For convenience f(L 01 m) can be expressed as
f(L 01m) = O(Lo1m)Pm + O(Lolm - Cim)qm (4.13)
Upon substituting Eq. 4.13 into Eq. 4.12, we have

F'm (L,) ~ 1- 0
F,m (L, - Lo'"' )[6 (L O '"' )Pm +6 (L 0 ,m - C'"' )qm] dL 0 ,m

= PmF1m(Le)(Lo1m) + qmF1m(Le - C1m)(Lo1m - C1m)


= p111 F1m(Le)(Le - C) + QmF1m(Le - C1m)(Le - C- Cim) (4.14)

From our discussion in Chap. 3, we know that Eq. 3.28 defines the expected value of
the load not served for load levels in excess of IC but less than 0. Since we are seeking
the expected value of the load not served by the already committed capacity of C MW
but less than C+ C1111 MW, which indeed is the expected value of the demand to be
produced by segment l of unit m, then substituting Eq. 4.I4 into
C+C1m
(Lim) =
1 C
F 1m(Le)dLe

1
we get C+C1m
(4.15)
E(Lim)= Pm Fim(Le)dLe
C

Further, the expected value of energy produced by segment Im is


C+Ctm
E(Eim) = Tpm [ F1m(Le)dLe ( 4.16)
C
Generation System Cost Analysis 113

To develop the needed expression for fuel cost, we must recognize that each
incremental amount of energy produced by segment lm will have a different cost, since
the incremental heat rate is a function of unit load level. The cost of supplying dLe
MW of load from segment Im would be
(4.17)
where

(4.18)

As Le varies over the range C <Le< C + C 1m, the values of IHR 1m vary from
IHR 1m(O) to IHR 1m(C1m), as they should. Integrating Eq. 4.17 over the range C to
C + C 1m, we get the expected value of the fuel cost for the segment.
C+C1m
e(FC1m) = VFCmPmT
! C
lHR1m(Le -C)F1m(Le)dLe (4.19)

To obtain the expected value of the incremental cost for use in comparing segment lm
with the incremental cost of other segments, in order to determine which segment
should be committed to create the ECS, we need only divide Eq. 4.19 in dollars by
Eq. 4.16 in MWh to get the expected value of the incremental fuel cost, e(IFC 1m)-
C+C1m

e(IFC1m)
1 lHR1m(Le - C)F1m(Le)dLe

= VFCm -C - - - - - - - - - - - -
C+Ctm
(4.20)

!C
F1m(Le)dLe

In the special case where IHR is a constant and equal to HR, as assumed in Eq. 4.6,
e(IFC1m) - VFCmHRm US$/MWh or mils/kWh (4.21)
where again UFCm is in US$/MBtu. If UFCm is giv~n in US$/bbl, US$/ton, or
US$/Mcf, then Eq. 4.21 must be divided by the fuel heating value HVm
VFCmHRm
e(IFC 1m) - HVm US$/MWh or mils/MWh (4.22)

In the case where the segment to be committed, segment ii, is either the second
or third segment of a unit that has already been partially committed and hence is
already in Fu(Le), the segments already convolved in must be deconvolved out of
Fu(Le), then combined with the segment about to be committed, and the combination
convolved back into the load distribution to form Fi1(Le)- For instance, if the second
segment of unit / were the segment about to be convolved into F 21 (Le), it would be
necessary to deconvolve segment 1/ out of F 21 (Le), to combine the capacity C 11 with
114 Power System Planning

C21 , and to convolve the combined segments of capacity C 11 + C21 into the load
distribution to give F 21 (Le)- The reason for this cumbersome procedure is to avoid
making unit / look more reliable than it really is. If unit I is divided into three
segments, each having an availability of p 1 and FOR of q 1, then the probability of
losing the capacity C1 of the total unit is q/. But q/ < q 1, thus the unit looks more
reliable than it really is. To avoid this problem, all lower segments of the same unit
committed before committing a higher segment must be deconvolved out of the load
distribution and combined with the segment about to be committed. For example, if
we split the capacity of unit I in the three-bus system into two 25-MW segments, each
with a FOR of 0.02, then F 11 (Le) would be

F 11 (Le) = F(Le)P1 + F(Le - 25)q1

and

F 21 (Le) = F 11 (Le)P1 + F 11 (Le - 25)q1


= F(Le)P/ + F(Le - 25)2Q1P1 + F(Le - 50)q1 2

However, if we simply convolve in the entire unit without segmenting its capacity, then
F 21 (Le) = F(Le)P1 + F(Le - 50)q1
Clearly, the two expressions for F'2 1 (Le) are not the same, and in particular the
coefficients of the term F(Le - 50), which is the FOR of the two segments, are not
the same. In the first case, where the FOR= q 1 2 , the two segments look more reliable
than the unit itself, which is clearly absurd. As we stated, the solution is to deconvolve
out the first segment, combine its capacity with the second, and convolve the
combination back into F(L); that is, before convolving in the second segment, solve
the first equation for F(Le)-
F11(Le) - F(Le - 25)q1
F(Le) = Pt

which is the deconvolution equation. Next add C 11 to C 21 to get 50 MW, and convolve
this back into F(Le), which gives us the correct result.
With thes~ ideas in mind, the expression for (Eu) is:

(4.23)

where Fu- 1 is the load distribution before segment ii is committed, with all other lower
committed segments of unit / deconvolved out of Fu- In addition

(4.24)
Generation System Cost Analysis 115

and

(4.25)

After all segments have been committed, the expected value of energy produced
by each unit is simply the sum of the expected values of energy produced by each
segment of the unit. Further, the expected value of the annual energy for each unit is
the sum of the monthly or weekly energy produced by each unit. If we denote the
expected value of the energy produced by segment i of unit n for week k as Ek(Ein),
then

Ek(En) = l j
Ek(E;n) (4.26)

and the total energy produced by unit n for the year becomes

E(En) = lk
Ek(En) (4.27)

In addition, the production cost in dollars and mils/kWh is given below with O&M
costs for unit n week k expressed as a fixed cost, FOMkn US$/wk, plus a variable cost,
VOMkn, in US$/MWh or mils/kWh.

ck(ECn) = l
i
Ek(FC;n) + l
i
VOM;nEk(E;n) + FOMkn (4.28)

Again the total energy cost for unit n for the year is

E(ECn) = l k
Ek(ECn) (4.29)

On a mils/kWh basis, the energy cost expression becomes


Ek(ECn)
Ek(ECn) = Ek(En) (4.30)

The concepts and details introduced here are difficult to grasp because of the
basic complexity of the problem. We shall consider a rather detailed example involving
the three-bus system. For this example, assume two-segment generator models and the
following data, which have been intentionally simplified.
116 Power System Planning

Unit I - Fossil oil Unit 2 - Fossil coal


C 1 = 50 MW C2 = 50 MW
C 11 = 25 MW C12 = 25 MW
C 21 = 25 MW C 22 = 25 MW
Q1 = 0.02 Q2'= 0.03
P1 = 0.98 p2 = 0.97
UFC 1 = 15 US$/bbl UFC 2 = 36 US$/ton
HV 1 = 6 MBtu/bbl HV2 = 24 MBtu/ton
HR 11 = 10 MBtu/MWh HR 12 = 12 MBtu/MWh
HR 21 = 12 MBtu/MWh HR 22 = 15 MBtu/MWh
OM 1 =0 OM 2 =0
The load distributions for each week, T = 168 h, will be assumed identical to the load
distribution used in Chap. 3 with 0 = 78.034 MW and forecast uncertainty neglected.
For convenience, F(L) is again depicted in Fig. 4.7. The first step is to determine
which of the four segments of the two units should be committed first, choosing the
segment with the smallest incremental cost. Obviously, only two possibilities exist:
either segment 11 or segment 12 can be committed first, since second segments of a
unit cannot physically be committed before first segments. Clearly we need to
calculate E(E11 ), (12 ), E(IFC 11 ), and E(IFC 12 ) and to compare e(IFC 11 ) with
E(IFC 12 ) to ascertain which represents the smallest incremental cost. Starting with
segment 11, Eq. 4.16 would be used to calculate E(E 11 ) since no other segments of
unit I have been previously committed; that is,

e(E 11 ) = Tp 1 f
0
25

F(L)dL

1.0 1 - - - - - - r - - ,

I
I
Segment I
being I
committed I

0 ..__ _ __.__ _.__ _ _ _ _ _ _ _ _ __


25.00 31.21 78.034
l.MW
Fig. 4.7 Load duration curve.
Generation System Cost Analysis 117

But from Fig. 4.7, F(L) = 1.0 for 0 < L 25 MW, thus

(11) = (168)(0.98)(25) = 4116 MWh


From Eq. 4.19

25

E{FC11) =
UFC1P1 T
HVi
1 0
IHR 1 (L)F(L)dL

but IHR 11 (Le)= HR 11 = 10 MBtu/MWh. Therefore


25

E{FC 11 ) = 3 68
(1 5)(0.~ )(1 )
10
l0F(L)dL

(15)(0.98)(168)(250)
6
= US$102,900
Finally from Eq. 4.20 or 4.22

E{IFCu) = UFC 1HR11 = (15)(10) = 150 = 25 US$/MWh


HV 1 6 6
Repeating the same procedure, except for segment 1 of unit 2, we get

E{E12 ) = (168)(0.97)(25) = 4,074 MWh


E{FC )
12
= (36)(0.97)(168)(25)(12)
24
= US$73,332
and finally
73,332
E(IFC12) = 4 ,074
= 18 US$/MWh
With E(IFC 11 ) and E(IFC 12 ) compared, it is obvious that to minimize production
cost, segment 1 of unit 2 should be committed first. Therefore the load distribution
with this segment committed is shown in Fig. 4.8.
At this juncture we can commit either segment 22 or segment 11; to decide, we
must calculate E(IFC 11 ) and E(IFC 22 ). Using Eqs. 4.16, 4.19, and 4.20, we get that

(11) = (168)(0.98) f
25
so
Fu(Le)dLe
118 Power System Planning

1.00 - - - - - , - - 1()J,
Slope= - - - = 0.02071
60

., 0.482
:::!. 10
N
Slope=--::- = 0.021358
ii.. Segment 121 60

0.016
I 10
I Slope = -::- q 2 = 0.0006407
60
I
I
31.21 56.21 78.034
Le,MW

Fig. 4.8 Load distribution after segment 12 has been committed.

where from Fig. 4.8


Fu(Le) = 1.0 for 25 Le 31.21
F 11 (Le) = 1 - 0.02071 (Le - 31.21) for 31.21 <Le< 50
Thus

e(E 11 ) = (168)(0.98) 6.21 + J31.21


so
[l - 0.02071 (Le - 3l.2l)]dLe

= 1,022.4 + 164.64(Le - 0.01035L/ + 0.6463Le)~~- 21


= 3,505.25 MWh

e(FC11) =
(15)(0.98)(168)
6 f 0
31.21

l0dLe + f so

31.21
10(1 - 0.02071 (Le - 31.2l)]dLe

(15)(0 .98)(168)(213 .49)


-
6
= US$87,876.42
and
87 876 2

(IFC II ) = .4 = 25 US$/MWh
3,505.05
Similarly, using Eqs. 4.23, 4.24, and 4.25, we get for segment 22

e(E22) = (168)(0.97)
so
J 2S
2
F22 - (Le) dLe
Generation System Cost Analysis 119

F 22 - 2 is the load distribution before segment 22 is committed with all other


segments of unit 2 removed. Since segment 12 was committed in the first
step, we must, to calculate e(E22 ) etc., remove it by deconvolving it out of
F 12 (Le) shown in Fig. 4.8. Performing the deconvolution, we get F(_Le = L) as
depicted in Fig. 4.7. Thus,

.
e(E22 )
.,

= (168)(0.97)
{ / 31.21
F(L)dL +
f 50
[l - 0.021358(L - 31.2l)]dL
}

25 31~1

= 162.96[6.21 + (Le - 0.010679L/ + 0.666Le)~~- 21 ]


= 3,457.71 MWh
(36)(16)(3,457 .71)
e(FC22) = 24
= US$77 ,798.54
77,798.54
e(IFC 22 ) = 3,457.71
= 22.50 US$/MWh

With two segments out of four committed, it is obvious that unit 2, because of its
relatively low fuel cost, is completely committed. Further, it is obvious that the last
two steps in developing the ECS is to commit segments 11 and 21, in that order, after
the combined segments of unit 2 are first convolved back into F(L) to give F 22 (Le) as
needed to calculate the commitment information for unit 1. F 22 (Le) is as shown in
Fig. 4.9. Completing the problem,

e(E11) = (168)(0.98) f 50
75

F 22 (Le) dLe

= 164.64 f
50
75

[1 - 0.02071 (Le - 31.2l)]dLe

= 164.64(1.646Le - 0.01035L/)~~
= 1,449.65 MWh
e(FC ) = (15)(10)14,449.65
11
6
= US$36,241.37
(IFC I I ) = 36,241.37
1,449.65
= 25 US$/MWh
120 Power System Planning

1.00 ----...--------

= 0.02071
Segment 12 I Segment 22
I
0.03 I q, 10
Slope=--_ = 0.000640
I Segments 11 & 21 60
I
0 C 12 31.21 C2 78.034 81.21 c, + c, 128.03

Le, MW

Fig. 4.9 Load distribution before unit 1 is committed.

Continuing

E(E21) = 164.64 f 75
100

F21-
1
(Le)dLe

E(E21) = 164.64
{
f
75
78.034

[1 - 0.02071 (Le - 31.21)]dLe + f 81.21

78.034
0.03dLe

+ /'
81.21
0

[D.Q3 - D.DDD64(L, - 81.2I)]dL,}

E(E21 ) = 164.64 (1.646Le - 0.01035L/);: 034 + 0.03 (3 .176)


+ (0.0819Le - 0.00032L/)!~~21
= 119.27 MWh
E(FC
21
) = (15)(12)119.27
6
= US$3,578.I

(IFC 21 ) = 3,581.78
119.27
= 30 US$/MWh
Upon convolving unit I into F 22 (Le), we get the final load distribution as
Generation System Cost Analysis 121

Slope= l(p~p 2 = 0.020303


,.K../
60
........

F(L)
,_,~

/<
F(Lel
'\

'-
a
10
6' ~,q, + q,p,) a 0.00104
lo.0494

II
I
FIL,)

F(LI ' - - ............. 10 -t0.0006


" , Sia/= 60 q 1 q 2 I
ii= 78.034 MW ""<., = 0.00001281
C1 = C = 50 MW
2 ,
,. I
0 L------+-----'----'L---+----'---1------+----------""--'
0.40 ii C, + 0.40 C 1 + ii IC+ O.4ii IC+ ii= 178.034

L and Le, MW

Fig. 4.10 Final effective load probability distribution.

depicted in Fig. 4.10. In summary, the results of the example are depicted in Table 4.4.
Note, as a check of the validity of the results, that

(generation) = E(ioad) - E(load not served)


2

l
i= 1
E(E;) = TE(L) - TE(DNS)

TABLE 4.4 Results of three-bus sample system analysis'!


Segment, unit E;j (MWh) FCij (US$)
1, 2 4,074.00 73,332.00
2,2 3,457.71 77,798.54
1, 1 1,449.65 36,241.37
2, 1 119.27 3,581.78
e(ENS) 75.42b
Te(L) 9,174.48
LOLP 5.009 h/wk
a Average IFC = US$20/MWh.
bsee page 83.
122 Power System Planning

4.4.3 Production Analysis Involving Nuclear Units


Worldwide, the electric industry has recognized the benefits of nuclear generation.
Along with the United States, most industrialized countries are moving rapidly toward
the day when nuclear generation will dominate their generation mix. Right or wrong,
this trend appears to be well under way, and for this reason we elect to discuss the
impact of nuclear generation on the overall problem of production simulation. As the
nuclc:..r age is very much in its infancy, many utilities are faced with the problem of
evaluating expansion plans including only one nuclear unit, whose capacity is generally
less than minimum load. This situation is by far the easiest to discuss in terms of
production analysis. We shall also discuss the situation in which competitive nuclear
units with a combined capacity in excess of the minimum load of a system make up
part of the generation mix.
For utilities contemplating their first nuclear unit and hence the problem of
evaluating the cost of an expansion plan including a nuclear unit, a primary factor to
be dealt with is determining the nuclear fuel cycle cost to use in production analysis.
Typically, if only one unit is involved, the usual assumptions are:

1. The unit will be base-loaded.


2. The unit will be refueled annually.

If so, the problems associated with evaluating the nuclear fuel cycle cost are less
formidable, since the characteristics of each new batch (that fraction of the core
replaced annually-typically one-third) are identical. Because determining nuclear fuel
cycle costs is important to utilities moving into nuclear generation, a lengthy discussion
of a crude but simple method for determining fuel cycle costs follows.
The basic nuclear fuel cycle is depicted in Fig. 4.11, where 11 major phases of
the cycle are delineated for costing purposes. These are:

1. Mining and milling 7. Spent fuel recovery and shipping


2. Refining 8. Spent fuel reprocessing
3. Enrichment 9. Spent fuel enrichment
4. Conversion 10. Spent fuel conversion
5. Fabrication 11. Spent fuel fabrication
6. Use

The mining and milling of basic ore is required to produce the uranium concentrate,
U 3 0 8 , known as "yellowcake." Refining converts the basic yellowcake into gaseous
UF 6 , or so-called "uranium feed." Because natural uranium is only 0.711 wt%
enriched-too low for satisfactory use in large modern nuclear power plants-the
refined uranium feed must be enriched to levels in the neighborhood of 3 wt%. As the
enriched uranium then remains in the UF 6 form, it must be converted into solid
uranium dioxide, U0 2 , so that it may be fabricated into pellets used in producing fuel
rods and, finally, fuel assemblies. Once fuel assemblies are irradiated in the core,
usually for a period of 36 months, they are recovered and shipped to reprocessing
Generation System Cost Analysis 123

en
C:
c
~
u,o. UF 6 Enriched UF 6 uo.

---I I
Ore
en
-~
~
,' en
-~
C:
:;::
Q)
a:
....
...
C:

E
Q)

.r:.
<.)
;::
w
C:
I I C:

-~
0

Q)
>
C:
0
(.)
C:
0
.,,
.-::
.!2
.0.,,
u.

Plant

Plutonium
t en
-~
C:

Q)
<.)

Recovered
uranium
I '...
"'.,,
C.
Q)
a:
I
Spent fuel
s:"'

Fig. 4.11 Basic nuclear fuel cycle.

centers, where the fissil plutonium and the remammg enriched uranium may be
recovered. The recovered uranium is returned to the enrichment facility, where it starts
the cycle again.
The usual input data required to support steady-state nuclear fuel cycle cost
analysis are given in Tables 4.5, a and b, and 4.6. The input data shown in Table 4.5, a
and b, reflect the cost of uranium and of conversion, enrichment, and fabrication. In
addition, the cost of shipping and reprocessing the spent fuel is also given in terms of
recovery and reconversion costs. All the numerical data shown are merely representa-
tive of the orders of magnitude of costs associated with the various items given.
Obviously, these numbers are continually changing. To enrich natural uranium for use
in the production of electric energy, a certain number of separative work units, SWUs,
are required per kilogram of enriched uranium; each SWU has a basic cost as given in
Table 4.5a. The number of SWUs needed per kilogram of enriched uranium depends on
the level of enrichment desired, as shown in Table 4.5b. For instance, the number of
SWUs needed to enrich a quantity of uranium to a level of 0.711 wt% is zero since
natural uranium is already 0.711 wt% enriched.
The economic data and representative design parameters given in Table 4.6
describe the characteristics of fuel to be analyzed as well as the appropriate monthly
..
N
"" TABLE 4.5b Separative work units vs. enrichmentb
Output Output
enrichment enrichment
(wt% 135 U) swu (wt% 235 U) swu
0.5 -0.173 3.0 4.306
TABLE4.5a Sample economic nuclear fuel 0.6 -0.107 3.1 4.526
cycle parameters" 0.7 -0.012 3.2 4.746
0.8 0.104 3.3 4.968
Annual
Fuel cycle item reload 0.9 0.236 3.4 5.191
1.0 0.380 3.5 5.414
Initial uranium 1.1 0.535 3.6 5.638
Ore (U 3 0 8 ), US$/lb 7.50 1.2 0.698 3.7 5.864
Conversion of U3 0 8 to UF6 , US$/kg U 2.50 1.3 0.868 3.8 6.090
Costof separative work, US$/unit 26.00 1.4 1.045 3.9 6.316
Diffusion plant tails enrichment, wt% 0.20 1.5 1.227 4.0 6.544
1.6 1.413 4.1 6.772
Fabrication, USS/kg U 80.00
1.7 1.603 4.2 7.001
Capacity factor, % 85.00 1.8 1.797 4.3 7.230
1.9 1.994 4.4 7.460
Spent fuel shipping, US$/kg U initial 6.00 2.0 2.194 4.5 7.690
Reprocessing 2.1 2.397 4.6 7.922
Recovery of U and Pu, US$/kg U final 30.00 2.2 2.602 4.7 8.153
Reconversion of U nitrate to UF6 , 2.3 2.809 4.8 8.385
US$/kg U final 5.60 2.4 3.018 4.9 8.618
2.5 3.229 5.0 8.851
Final uranium 2.6 3.441 5.1 9.084
Ore (U 3 0 8 ), US$/lb 1.50
2.7 3.656 5.2 9.318
Conversion of U, 0 8 to UF6 , US$/kg U 2.50
2.8 3.871 5.3 9.553
Cost of separative work, US$/unit 26.00
2.9 4.088 5.4 9.787
Diffusion plant tails enrichment, wt% 0.20
by Power Engineering.
Copyright
Fissile plutonium, USS/kg Pu 8000.00
bSWU/kg enriched uranium. Natural concentration= 0.711 wt%
oCopyright by Power Engineering. 135 U; tails concentration= 0.200 wt% 135 U.

~.. st,,,f ,,. ,,- /#C~ ;,,,,~ ..


~

Iv-
;,g,
!:J-- ,....
o::s ::r'
O
~g'1;~'Tl2'"Sl~~
e!. ~ g_ f;; E: ~ g_ g. ~
;;,s
-g:O<l
i:: =
g'<
"O 'O i:: !' -
0 ~ ~ JJ n
"
~ qo !"'> !> S:: ?- P ?' I"' S:: ~
!-"
p- I"' !"
S:: ':"
g n n n "C g ~ ..... "C g >-:l t"' 'TJ 'TJ 'TJ - 'TJ - :=
n n n
goooi;. 0 0 0 3 5 - .. g-i?::i.;. ::r=:=~t.s=-sc!.o
I"'" ?' 'I' !"'> $1 p. P ?' I"' :"'
;;1
o'
<il
ji! CD
r
m

_ ~ ~~, pi(t ~, ~:~ t , a ltM ~ttW-ieiiiti, WW,W " it


Generation System Cost Analysis 125

TABLE 4.6 Sample nuclear fuel cycle design parameters0


Annual
Nuclear fuel cost estimate reload
1. Reference design parameters
a. Region power, MWt 800
b. Initial enrichment, wt% 3.2
c. Final enrichment, wt% 0.95
d. Initial uranium weight, kg U 24,820
e. Final uranium weight, kg U 23,820
f. Fissile plutonium produced, kg Pu 155
g. Discharge burnup, MWD/MTU 30,000
h. Full-power hours 22,338
i. Lifetime at 85% capacity factor, months 36.0
j. Thermal energy output, 10 11 Btu 609.92
2. Monthly carrying charge rates
a. Prior to commercial use, % 1.0
b. During and after commercial use, % 1.0
(both based on 12% per annum -i- 12)
3. Average time periods
Months prior to commercial use
a. Procure U 3 0 8 , months 12 == T,
b. Complete conversion, months 10 == T 2
c. Complete enriching, months 7 == T 3
d. Complete fabrication, months 2 == T 4
Months subsequent to refueling
e. Complete spent fuel cooling, months 4 == T 5
f. Complete spent fuel shipment, months 7 == T 6
g. Complete uranium and plutonium reprocessing,
months 10 == T 7
Copyright by Power Engineering.

carrying charges and the average time periods associated with each phase of the fuel
cycle. The reference design parameters are obtained, as discussed earlier, using
production and in-core analysis. Specifically, the thermal energy output of the batch,
full-power hours, and capacity factor are determined from production analysis.
Full-power hours is obtained by multiplying the CF by the lifetime of the batch,
usually 36 months. Initial and final enrichment, final uranium weight, fissile plutonium
produced, and burnup are obtained from in-core analysis. Burnup is in a sense the
heating value of nuclear fuel and, as such, it gives the amount of thermal energy
available in the batch when multiplied by the kilograms of enriched uranium.
In keeping with the approach used in discussing the corporate model, the details
of the average nuclear fuel cycle cost analysis will be presented in block diagram
fashion, as depicted in Fig. 4.12. The major inputs are
1. Initial kilograms of enriched uranium, kg EU0
2. Initial enrichment, wt%0
...,..,
CJ)

Enrichment
cc
u,o_!
cc
Conversion Net carrying charges
cc

O.ST%+CI Fabric.a1io~CC ) 1
UFC; US$/MBtu
Fabrication cost

wt%/

USS
kgEIJf I credit

Reprocessing and spent fuel ru "''"'-'" j 0.99US$ I kg Pu,

'--------;
0.99 US$
kg EU/
I
Conversion
cost
I
shipping costs
kg Pu(

Total displaced CC
Fuel shipping

"'"
Fig. 4.12 Block diagram of nuclear fuel cycle cost analysis.
Generation System Cost Analysis 127

3. Final enrichment, wt%/


4. Final kilograms of enriched uranium, kg EU/
5. Fuel shipping cost, US$
6. Carrying charges, %
7. Carrying charge periods, T 1 -T7
8. Plutonium produced, kg Puf
9. Thermal energy required, e(E 1)

The major output is the UFC for the batch considered.


To illustrate the details we have discussed, consider determining the UFC for one
batch of a 2,400-MW 1 unit. All the data needed are given in Tables 4.5 and 4.6; the
solution is provided in compact form in Table 4.7. Since additional discussion will not
clarify the details, the student is urged to spend time reviewing this example and
relating these steps to those in Fig. 4.12.
As long as nuclear units are assumed to be base-loaded and maintained and
refueled annually, the nuclear fuel cost analysis as discussed renders the information
needed to evaluate the production cost associated with the nuclear unit. In fact, under
the assumptions stated, the nuclear unit is handled in the same way as a fixed-energy
fossil unit, since the assumptions have effectively decoupled the fuel cycle from the
operation of the overall system.
For completeness we must consider a far more complicated situation-namely,
the one that arises when a utility must evaluate an expansion plan containing a
significant proportion of nuclear capacity, at least more than minimum load. This
situation creates very complex simulation problems. Specifically,

1. Nuclear refueling and maintenance schedules must be determined which


satisfy the reliability criteria at minimum cost.
2. Nuclear unit ECSs must be determined using an iterative process since nuclear
fuel cycle costs and hence incremental cost are not known a priori, because
they depend on the past, present, and future state and operating trajectory of
the nuclear core.

A broad view may help in orienting the discussions to follow. Figure 4.13
illustrates the major simulation blocks in determining the production costs needed in
order to evaluate economically an expansion plan dominated by nuclear generation. We
note that the initial step involves establishing maintenance and refueling schedules for
all units. Generally, if we were dealing with an all-fossil system, this step would already
be complete. However, with the relatively high replacement energy costs of nuclear
generation, a schedule based only on reliability may not be a minimum-cost schedule.
Strictly speaking, schedules developed using leveled LOLPs as the only criteria may
require modification before proceeding. Fortunately such schedules tend to be close to
minimum-cost schedules because, on the basis of the leveled LOLP criterion, nuclear
units tend to be refueled and maintained during low load periods of each year, when
the need for high-cost peaking energy is lowest. Thus, establishing refueling and
TABLE 4.7 Sample nuclear fuel cycle costs analysis"
Nuclear fuel Annual Nuclear fuel Annual
cost estimate reload cost estimate reload
4. Initial investments d. Reprocessing investment=
Uranium requirements Sb+ Sc, I0 3 US$ 847
a. Uranium as enriched UF 4 = Equivalent uranium requirements
ld kg U (sec Table 4.6) 24,820 for uranium credit
b. Uranium as natural UF 6 e. Uranium as reprocessed
feed to the enriching facil- UF 6 = le X 0.987,d kg U 23,510
ity = 4a X (lb - 0.2)/(0.711 f. Uranium as natural UF 6
-0.2), kg U 145,714 feed to the enriching facil-
C. Units of separative work per ity = Se X (le - 0.2)/
kg of enriched uranium as (0. 711 - 0.2), kg U 34,506
UF 6 (seeTable4.5b) 4.746 g. Units of separative work
d. Total units of separative per kg of enriched uranium
work= 4a X 4c 117,796 as UF 6 (Table 4.Sb) 0.308
e. Natural uranium feed for h. Total units of separative
U 3 0 8 to UF 6 conversion work= Se X Sg 7,241
facility= 4b/0. 995,b kg U 146,446 i. Natural uranium feed for
Uranium investment U,O 8 to UF 6 conversion
f. U 3 0 8 = 4e X 260 X (US$/ facility= Sf/0.995,e kg U 34,679
lb for U,O 8 ), 10 3 US$ 2,856 Uranium credit
g. Conversion = 4b X (US$/ j. Equivalent U 3 0 8 invest-
kg U for conversion), I0 3 ment = Si X 2.60 X (US$/
US$ 364 lbforU 3 O8 ), 10 3 US$ 676
h. Enriching= 4d X (US$/unit k. Equivalent conversion in-
for enriching), 10 3 US$ 3,063 vestment= Sf X (US$/kg U
i. Initial uranium investment for conversion), 10 3 US$ 86
= 4f + 4g + 4h, 10 3 US$ 6,283 L Equivalent enriching invest-
Fabrication investment ment = Sh X (US$/unit for
j. Unit fabrication price, enriching), 10 3 US$ 188
US$/kg U 80 m. Uranium credit = Sj + Sk +
k. Fabricatiort investment= 51, 10 3 US$ 950
Id X 4j, 10 3 US$ 1,986 Plutonium credit
I. Total ir1itial investment = n. Fissile plutonium credit =
4i + 4k, 10 3 US$ 8,269 1f X 0.99c X (US$/kg-Pu
for Pu-nitrate), 10 3 US$ 1,228
5. Final investments and credits 0. Total final investment=
Spent fuel shipping investment Sm + Sn - Sa - Sd, 10 3
_a Spent fuel1shipment invest- US$ 1,181
ment, 10 3 US$ 150
o. Uranium depletion and carrying
Reprocessing investment charges
b. Recovery of U and Pu in
a. Uranium depletion = 4i -
nitrate form= le X (US$/
Sm, 10 3 US$ 5,333
kg U for recovery), 10 3
US$ 715 Carrying charges prior to com-
c. Conversion of U-nitrate to mercial use
UF 6 = le X 0.99C X (US$/ b. On U 3 0 8 investment= 4f X
kg U for conversion), 10 3 US$ 132 3a X 2a/100, 10 3 US$ 343
acopyright by Power Engineering. dLoss during reprocessing assumed to be 1.3%.
bLoss during conversion assumed to be 0.5%. eLoss during conversion assumed to be 0.5%.
cLoss during recovery assumed to be 1%.
128
TABLE 4.7 (continued) Sample nuclear fuel cycle costs analysisa

Nuclear fuel Annual Nuclear fuel Annual


cost estimate reload cost estimate reload
c. On conversion investment= e. After refueling= [ ( Sa X 30 +
4g X 3b X 2a/100, 10 3 US$ 36 (Sb X 3g)] X 2b/100, 10 3 US$ 95
d On enriching investment= f. Total displaced carrying
4h X 3c X 2a/100, 10 3 US$ 214 charges = 8d + Se, 10 3 US$ 274
e. Subtotal = 6b + 6c + 6d,
10 3 US$ 593 9. Plutonium appreciation and
carrying charges
Carrying charges during com-
a. Plutonium appreciation =
mercial use
Sn, 10 3 US$ 1,228
f. On average uranium invest-
ment = [(4i + 5m)/2] X Carrying charges
2b/100, 10 3 US$ 1,302 b. During commercial use =
(9a/2) X Ii X 2b/100, 10 3 US$ 221
Carrying charges after refueling
C. After refueling= 9a X 3g X
g. On uranium credit = Sm X
2b/100, 10 3 US$ 123
3g X 2b/100, 10 3 US$ 95
d. Total plutonium carrying
h. Total uranium carrying
charges= 9b + 9c, 10 3 US$ 344
charges = 6e + 6f + 6g,
10 3 US$ 1,990 10. Summary
7. Fabrication depreciation and a. Uranium depletion = 6a,
carryitJg charges 10 3 US$ 5,333
b. Fabrication depreciation=
a. Fabrication depreciation =
7a, 10 3 US$ 1,986
4k, 10 3 US$ 1,986
C. Spent fuel shipping and re-
Carrying charges processing capital accumu-
b. Prior to commercial use per lation = Sc, 10 3 US$ 996
Westinghouse payment d. Plutonium appreciation =
schedule, 10 3 US$ 109 9a, 10 3 US$ (1,228)
C. During commercial use = e. Net capital depreciation=
(7a/2) X li X 2b/100, 10 3 357 10a + 10b + 10c - 10d,
US$ 357 10 3 US$ 7,087
d Total fabrication carrying f. Uranium carrying charges
charges = 7b + 7c, 10 3 US$ 466 = 6h, 10 3 US$ 1,990
8. Spent fuel shipping and repro- g. Fabrication carrying
cessing charges= 7d, 10 3 US$ 466
h. Spent fuel shipping and re-
Capital accumulations and car- processing carrying charges
rying charges
displaced = Sf, 10 3 US$ (274)
a. Spent fuel shipping accumu-
i. Plutonium carrying charges
lation = 5a, 10 3 US$ 149
= 9d, 10 3 US$ 344
b. Reprocessing accumulation j. Net carrying charges = 1 Of
= 5d, 10 3 US$ 847
+ 10g-10h + l0i, 10 3 US$ 2,526
C. Total accumulation= Sa +
Sb, 10 3 US$ 996 k. Total fuel cost= l0e + l0j,
3
10 US$ 9,613
Carrying charges displaced by
L Thermal energy output=
accumulations lj, 10" Btu 609.92
d. During commercial use =
m. Total unit energy fuel cost
(Sc/2) X li X 2b/100, 10 3 US$ 179
= lOk/101, /10 8 Btu 15.76
acopyright by Power Engineering. dLoss during reprocessing assumed to be 1. 3%.
bLoss during conversion assumed to be 0.5%. eLoss during conversion assumed to be 0.5%.
cLoss during recovery assumed to be 1%.
129
130 Power System Planning

Maintenance
and
refueling scheduler

Cycle lengths
maintenance and .,
refueling schedule
Unit da ta
' Initial UFC;
P robabi Ii sti c
production analysis

e(E;), CF;

Core simulator
and
fuel cycle optimizer

Enrichment
batch fraction
state of co re
Necessary Nuclear fuel cycle
data cost analysis

UFC;v Update
= UFC/- 1 UFC/- 1

Stop

Fig. 4.13 Flow chart of production analysis.

maintenance schedules with the ideas presented in Chap. 3 is quite satisfactory for
intermediate and long-range planning.
Assuming that we have several nuclear-dominated expansion candidates with
acceptable maintenance and refueling schedules, the cycle lengths (cycle length is the
time between refueling) for each unit over the planning horizon are known, as are the
times during which various units are maintained and/or refueled. With this information
the expected values of energy, fuel cost, etc., for each unit can be determined using
the concepts presented earlier, if the fuel cycle costs for each cycle of each nuclear
unit are known-but these costs are generally unknown at this point. The best that can
be done is to use representative fuel cycle costs in each cycle for each unit, as shown
in the second block of Fig. 4.13. Using representative fuel cycle costs, the ECS can be
determined. Knowing the energy produced as well as the capacity factors, in-core fuel
analysis can be employed to calculate the optimal (minimum cost) fuel characteristics
required to produce the energy needed. Finally, by recalculating fuel cycle costs using
the optimal fuel characteristics, the validity of the initial representative fuel costs and
associated ECS can be evaluated.
Generation System Cost Analysis 131

If the fuel cycle costs for each unit in every cycle, calculated using optimal fuel
characteristics (i.e., optimal batch fraction and state of the core), equal the values
assumed at the beginning of the analysis, we have the optimal solution; that is, we
have the ECS for each unit over the planning horizon. Of course it is this information
that is needed t? evaluate the expansion plan economically.
Given that the initial assumed nuclear fuel cycle costs are not the optimal costs,
we are faced with an iterative process depicted in Fig. 4.13. In essence, the initial
assumed fuel cycle costs must be updated so that when we resolve for the optimal fuel
cycle costs, the two agree.
The mathematical formulation of this problem is beyond the scope of this work;
the intent has been to familiarize the reader with an area of power system
simulation that, because of its complexity, has yet to take definite shape. Many seeds
have been planted, as the literature indicates; we have only to wait until the most
viable breaks ground.

4.4.4 Production Analysis Involving Hydro Units


In Sec. 4.4.3 we highlighted some factors that must be included in evaluating
expansion plans containing nuclear units. For completeness it is necessary to present
some basic characteristics of hydro units that must also be included in production
analysis to obtain accurate cost results. Although hydro resources are dwindling in the
more highly developed nations, many countries throughout the world are rapidly
tapping hydro resources to support industrialization. We shall dwell on conventional
reservoir and pumped hydro units, which tend to have control characteristics that
make them suitable for peak shaving-that is, supplying energy normally supplied by
high-cost fossil peaking units.
Using hydro units to peak shave is sound operational strategy because
conventional hydro units, which can be considered to have zero fuel cost, have a
negative replacement energy cost. Negative replacement energy costs are savings; hence it
is desirable to maximize these savings through their excellent controllability. Although
conventional units have zero fuel costs, pumped hydro units, as the name implies, must
consume energy to pump water back into the upper reservoir for future use, and hence
the cost of pumping must be considered as fuel cost.
Like nuclear units, hydro units create simulation problems, not only because of
the need to consider economics in scheduling their maintenance but, more important,
because of the nature of their fuel cycles. Conventional hydro units must be operated
along outflow trajectories adhering to the ECS as well as satisfying rather rigid
constraints on reservoir storage levels, which are determined by outflow from the
reservoir as well as by highly variable inflows. The major fuel cycle constraint is that
the equivalent energy of the reservoir or level must be returned to its original value at
the end of a water cycle. To determine the optimal outflow trajectory, given these
constraints, can be rather involved, especially for systems with interconnected
multireservoirs. In general, the optimal outflow trajectories as well as the system ECS
are determined with the help of dynamic programming, a procedure for selecting the
optimal path between two operating states interconnected by multiple paths, each
132 Power System Planning

having different costs. Because of the zero fuel cost associated with conventional
hydro, as well as its controllability, the tendency will be to peak shave as much hydro
capacity as possible and to use the remaining capacity to replace more expensive
base-load energy. Although we shall not pursue the details of determining optimal
outflow trajectories, the basic approach for a, single reservoir is presented below:
1. Construct the resen-oir water storage model;
Xn+1 = Xn + Yn - Un
where X 11 + 1 = storage level in equivalent MWh at the beginning of week n +1
Yn = forecast inflows in equivalent MWh for week n
= outflow in equivalent MWh during week n
U11
2. For each feasible decision Un, calculate the probabilistic production cost.
Repeat for every week in the water cycle being considered.
3. Using dynamic programming, determine the optimal decisions: the outflow
trajectory to be adhered to for the water cycle in question. Figure 4.14
depicts the structure of the dynamic programming problem to be solved.
Of course, if Y 11 is considered to be random-obviously it is, in reality-then the
stochastic version of dynamic programming must be used to determine the expected
values of the optimal decisions. We shall concentrate here on how to simulate the
utilization of the equivalent energy Un so that for week n the production cost needed
to determine the optimal outflow trajectory can be calculated for each feasible Un.
Assume we are interested in determining the ECS and corresponding production
cost for each value that Un can assume in week n. Consider determining the

x, x. Xn+l

Acceptable
storage
levels

Fig. 4.14 Decision tree for determining optimal outflows.


Generation System Cost Analysis 133

F(L)

L- 0

Fig. 4.15 Load distribution peak shaved by a conventional hydro unit.

production cost for a particular value of Un, say Un(I). From earlier discussions we
recognize the need to use as much of the equivalent energy Un(l) as necessary for
peak shaving, because replacing high-cost fossil peaking energy with zero-cost hydro
energy is highly desirable. If we assume that the hydro unit has capacity Ch, then Fig.
4.15 depicts the effect of peak shaving on the load distribution; for simplicity it has
been assumed that all of the energy Un(l) is used to peak shave. Further, the
peak-shaved energy produced by the hydro unit has been appropriately removed before
determining the ECS for thermal units. This step assumes that the hydro unit is 100%
reliable, hardly a limitation for planning studies, since hydro units are indeed reliable.
The resulting load distribution is then used as in Sec. 4.4.2, except that special
consideration must be given to pumped storage units when determining the ECS and
production cost.
Pumped hydro units, because system energy must be supplied for the necessary
pumping, are troublesome to simulate. We shall confront the simulation problem of
determining the least-cost generating and pumping strategy to satisfy the constraint
that the state of the upper reservoir be the same at the beginning of each week. Since
pumped storage units were conceived to replace expensive peaker energy during peak
periods and to be recharged, if you will, by inexpensive base-load energy during
off-peak periods, we may expect a pumped hydro unit of capacity Ch to alter the load
distribution as shown in Fig. 4.16. Obviously, the cost benefits of peak shaving must
be weighed against the cost of pumping. If the pumped energy had to be supplied by
peaking units instead of base-load units, the economic benefits of pumped storage
would quickly evaporate, since pumping efficiencies are considerably less than unit
efficiencies in generating energy. Thus, to determine how much energy should be used
for peak shaving, the point at which the incremental cost for pumping equals the
incremental cost for generating must be used as a boundary beyond which utilization
of pumped storage capacity is not economical. In general, system minimum loads are
approximately 50% of the peak, and since few utilities have generation mixes
consisting of 50% fossil peakers, base-load energy during off-peak periods is usually
available.
134 Power System Planning

1.0 1----------c:- - - --,


C h . . - ~ Pumping energy

I
I
I
F(L)

0
L-

Fig. 4.16 Load distribution peak shaved by pumped hydro unit.

Determining the optimal operating trajectory for pumped hydro units is


considered an operational planning problem because the cycle length is generally a
week. Although operational planning is beyond our scope, we should recognize that
dynamic programming could be used to develop optimal decisions w.r.t. the generation
and pumping of pumped storage units.

4.5 FUEL INVENTORIES

In the process of identifying the best expansion plan, some consideration must be
given to fuel availability and facilities in addition to cost. Since most utilities enter
into long-term contracts with fuel suppliers, they must know how much of a given
type will be needed, so that negotiations with suppliers can be undertaken. Since
storage and transmission facilities must also be constructed in advance, it is imperative
that future estimates of amount, type, etc., be available.
We are fortunate to have approached generation system expansion analysis as we
have, because obtaining fuel inventories involves only a simple extension of concepts
already presented. In Sec. 4.4.3 we developed expressions for fuel cost:
C+C1m

e(FC,m) = UFCmPmT [ IHRm(Le - C)F,m(f,e)dLe


C

and

i =I= 1

Since the coefficients of the UFCs are the expected value of segment energies in
MBtus, the expected value of the amount of fuel required for any unit, say m, is
Generation System Cost Analysis 135

e(FCm)
e(fuelm) = UFC (4.31)
m

or

(4.32)
UFCmHVm
where HVm is the heating value of the fuel. The dimensions of Eq. 4.31 will be tons,
Mcf, or barrels, depending on whether unit m is a coal-, gas-, or oil-fired unit.
In the case of a nuclear unit the amount of initial enriched uranium, etc., is
determined, as we discussed, through the use of detailed in-core analysis with unit
energy and cycle length used as inputs.
With Eq. 4.31 we can complete the example presented in Sec. 4.4.2 involving the
three-bus system. In the example we found that the cost of energy produced by oil
(unit 1) and coal (unit 2) units was

e(FC 2 ) = US$151,130.S4
e(FC 1 ) = US$39,823.15
Therefore, since UFC 1 = US$15/bbl and UFC 2 = US$36/ton
151,130.54
(fue12 ) = 36
tons

= 4,198.07 tons
e(fuel 1 ) = 39
,s:~ 15
bbl

= 2,654.87 bbl
This example nicely demonstrates that computing fuel inventories is a simple
process. We shall find that having this capability will facilitate evaluating the true
financial impact of using low-sulfur fuels to improve ambient air quality.

4.6 ENERGY TRANSACTIONS AND OFF-PEAK LOADING


In Chap. 3 we presented a few basic ideas pertaining to the operation of interconnected
systems when they are providing emergency support for one another during periods of
reserve deficiencies due to forced outages. We shall now address the problem of
. including firm interchange agreements in production analysis in order to simulate the
situation when a company is buying energy at a rate of Lr MW at a unit interchange
cost of UICr US$/MWh. We are interested in developing a procedure that will enable
the system planner to consider expansion plans in which purchased energy can be used
to satisfy reserve requirements on a firm basis during periods when, because of
maintenance or known construction delays, insufficient capacity exists. Using pur-
chased energy as an alternative to adding peaking or midrange capacity while
136 Power System Planning

completing construction of base-load units is a strategy that is not uncommon.


Analyzing the economic advantages and disadvantages of such strategies requires
determining the ECS for the alternative plans. The alternative with the most favorable
impact on the utility will be chosen to supplement existing system capacity until the
construction of base-load capacity is comp!ete. Generally, buying energy on a firm
basis is expensive; in most cases the buyer is purchasing the most expensive part of the
seller's energy source.
If a system planner intentionally overdevelops the system to exploit anticipated
deficiencies in other interconnected utilities-which probably does not happen
often-or if the concept of off-peak energy use can be successfully promoted, then
having the capability of analyzing the impact of energy sales is important. Energy sales
analysis involves basically the same procedures as purchases, at least from a production
analysis viewpoint. Therefore it is logical to include in our discussion both energy
purchases and sales.

4.6.1 Energy Purchases


Assume that two companies are interconnected as shown in Fig. 4.17 and that
company B, because of an anticipated delay in the completion of a new base-load unit,
wishes to buy on a weekly basis some Er MWh of energy at a rate of Lr MW and at a
unit interchange cost of UICr US$/MWh. The weekly cost of this energy would be
ECr = ErVICr US$ (4.33)
The energy cost for energy produced by units in system B is now changed, since
buying energy at a rate of Lr MW shifts the effective load distribution to the left. The
shift changes the amount of energy each unit in system B produces as well as the
corresponding production cost. Since

F(!,eBr) = F(LeB = LeBr + Lr) ( 4.34)


then

(4.35)

and
C+C1m

E(FC Im) = UFCmPm T


fC
IHR1m (LeBr - C)F1 m (LeBr )dLeBr ( 4.36)

Fig. 4.17 Two interconnected areas.


Generation System Cost Analysis 137

Similarly, for second and third segments


C+Cu

e(Eu) = P1T
J
C
1
F;'i (LeBT) dleBT (4.37)

(4.38)

Since O&M cost has a variable component depending on the energy produced by each
unit, the total production cost for each unit, ek(ECn), is altered as a result of the
purchase.

4.6.2 Energy Sales


Because systems frequently have capacity over and above that needed to satisfy reserve
requirements, as in the example in Sec. 4.6.1, selling energy to an interconnected
utility is a method of increasing operating revenues. In addition, it is an excellent way
to improve the system CF during periods immediately after a large unit addition.
Typically, system CFs are in the 50-60% range, indicating that 40-50% of system
energy capability remains idle.
Since the technique for including energy sales is the same technique used for
modeling energy purchases, we shall not pursue the analytical details further. Rather
we shall discuss the important concept of off-peak energy sales. During the off-peak
hours most utilities have excess capacity, and the CF of a system suffers most during
these hours. Since we are differentiating between peak and off-peak periods, any
gene~ation system expansion strategy that involves off-peak sales might, whether on a
weekly or a monthly basis, make use of two load probability distributions: one to
describe peak load hours when most of the system's capacity is committed with little
available for sale, and another to describe off-peak hours when excess system capacity
is available.

4.6.3 Off-Peak Energy Utilization


In many countries a sensible solution to improving poor CFs has been to encourage
through lower rates the use of off-peak energy. By defining two different load shapes
for peak and off-peak periods, as alluded to above, a utility can model off-peak energy
sales and evaluate the production costs of a particular expansion plan.
To demonstrate the effect off-peak sales might have on weekly load distributions,
Fig. 4.18 is presented. The benefits to be derived from off-peak sales can be
summarized as follows:
I. Improved CFs and hence lower capacity cost and better utilization of
investment
2. Reduced need for peaking devices, resulting in lower capacity, production,
and system maintenance costs for fossil peakers
138 Power System Planning

1.0 t--------~----
''
' \
\
F(L) \
\

0
L-
Fig. 4.18 Load distribution showing the effects of off-peak loading.

Since the shift to off-peak energy utilization changes only the load distribution, the
ECS for a particular expansion plan for a particular week is obtained in the manner
discussed in Sec. 4.4.2, with the exception that if different rate schedules are used, an
off-peak load distribution may be defined and used to obtain more accurate
production cost information.

4.7 ENVIRONMENTAL COST


It would be inappropriate to conclude the subject of generation system cost analysis
without discussing, although in some cases in a rather cursory manner, the impact of
environmental constraints on the modern electric utility. We shall concentrate on the
impact that air and water quality constraints have had, realizing that many other
constraints are perhaps equally as important.
Even though nuclear generation is increasing rapidly, many countries will
continue for some years to rely on fossil-fueled generators to supply the bulk of their
electric energy needs. In fact, with the apparently tight oil situation, a return to the
use of coal is being undertaken by many utilities. The point here is that fossil-fueled
plants will be around a good while, continuing to threaten the quality of the air we
breathe, until workable and economically feasible methods are found to clean up fossil
fuels. Because of the amount of concern for maintaining a desirable level of ambient
air quality, we shall discuss in detail a method for assessing the ambient air quality for
a given generation system expansion plan, with particular emphasis given to ground
level S0 2 and particulate concentrations. We shall discuss in some detail the cost of
reducing S0 2 and particulate emissions, using precipitators and scrubbers as well as
system design parameters such as stack heights, fuel sulfur contents, etc.
There is no doubt that thermally polluting rivers and lakes is having some effect
on ecosystems, but the effects have not yet been effectively quantified. What is clear is
that the water supply available for power plants is reaching a critical state in some
areas. This critical situation can be resolved only through the use of so-called dry
cooling methods that do not require makeup water and do not involve injecting
Generation System Cost Analysis 139

Stack losses, 0.9 X 109 Btu


t h

Turbine

Fuel, 9 X 109 B~u


Boiler
Condenser 430,000 gpm

In-plant losses, 0.4 x 109 _tu


h Makeup,
11,194 gpm
Fig. 4.19 Fossil steam system.

thermally polluted condenser water into natural bodies of water. We shall review a few
of the methods used in cooling modern power plants. The approximate costs
associated with the most common techniques are given in Chap. 1.

4.7.1 Thermal Pollution


To understand the thermal problems associated with large modern fossil fuel and
nuclear power plants, energy flow diagrams for both 1,000-MW nuclear and
fossil-fueled plants are shown in Fig.. 4.19 and 4.20. The primary inputs into these two
systems are fuel and water, and the primary outputs are energy, waste heat, and stack
gases associated with fossil-fueled plants. The amount of energy injected into the

Turbine 3.4 X 109 B~u Btu


_,,,-...._____J 6 .4 X 109 h

Fuel, 10.3 x 1 o Btu


h Nuclear
steam generator Condenser 475,000 gpm
Btu
In-plant losses, 0.5 X 109 h Makeup,
12,500 gpm

Fig. 4.20 Nuclear system.


140 Power System Planning

cooling system each hour is 4.3 MBtu for fossil plants and 6.4 MBtu for nuclear plants,
basically the amount of energy in one precious barrel of oil. To cool the two types of
plants properly, 430,000 gpm is needed for the fossil plant compared to 475,000 gpm
for the nuclear plant. To put this water requirement into perspective, the average
homeowner in the United States uses about 0.3 gpm.
If once-through cooling systems are used, conceivably we could deplete our water
resources and irrevocably damage our ecosystems. If once-through cooling from
artificial ponds is used, many acres of land would be required at a time when land is at
a premium. Typically, at least 1 acre/MW is needed to provide cooling water for
modern plants. The solution to these particular problems lies within the boundaries of
dry cooling technology. A partial solution, the one used most in 1976, is evaporative
cooling using either natural or mechanical draft towers. Because evaporative cooling
systems circulate water as shown in Fig. 4.19, a 95% reduction in water requirements
is realized, and of course the waste heat is removed by evaporation. As with most
solutions, the electric utility industry through its customers pays a price for these
improvements. Cooling system costs for fossil-fueled and nuclear units were given in
Chap. 1. In performing expansion plan analysis these costs should be added to the
initial capacity cost of the plant.

4. 7.2 Air Pollution


Referring to Figs. 4.19 and 4.20, we see that the major source of air pollution
involving nonradioactive pollutants is the fossil-fueled unit, from which the following
pollutants are emitted into the atmosphere from the stack:
I. SO 2 4. co
2. Particulates 5. HC
3. NOx
Table 4.8 indicates that power generation is responsible for nearly one-half of the SO 2
and one-quarter of the particulates and NOx emitted into the atmosphere. The Federal
Air Quality Standards, as yet not fully implemented, are given in Table 4.9. It is

TABLE 4.8 Source of estimated air pollution by weight (1968)


Pollutant (%)
Sulfur Nitrogen Carbon
Source oxides oxides monoxide Hydrocarbons Particulates Total
Transportation 0.7 4.2 46.5 8.5 0.7 60.6
Industry 6.3 1.4 1.4 2.8 4.2 16.1
Power generation 8.5 2.2 0.7 0.7 2.1 14.2
Space heating 2.1 0.7 1.4 0.7 0.7 5.6
Refuse burning 0.7 0.7 0.7 0.7 0.7 3.5
Totals 18.3 9.2 50.7 13.4 8.4 100.0
Generation System Cost Analysis 141

TABLE 4.9 Federal air quality standardsa


Standard
Pollutant Primary Secondary
Sulfur oxides
Annual arithmetic mean 80 g/m 3 (0.03 ppm) 60 g/m 3 (0.02 ppm)
Maximum 24-h concentration 365 g/m 3 (0.14 ppm) 260 g/m 3 (0.10 ppm)
Maximum 3-h concentration 1,300 g/m 3 (0.50 ppm)
Particulate matter
Annual geometric mean 75 g/m 3 60 g/m 3
Maximum 24-h concentration 260 g/m 3 150 g/m 3
Carbon monoxide
Maximum 8-h concentration 10 mg/m 3 (9 ppm)
Same as primary
Maximum 1-h concentration 40 mg/m 3 (35 ppm)
Photochemical oxidants
Maximum 1-h concentration 160 g/m 3 (0.08 ppm) Same. as primary
Hydrocarbons
Maximum 3-h concentration (6-9 a.m.) 160 g/m 3 (0.24 ppm) Same as primary
Nitrogen oxides
Annual arithmetic mean 100 g/m 3 (0.05 ppm) Same as primary
t:Emission rate standards: new plant; 0.8 lb/MBtu for liquid fuels or 1.2 lb/MBtu for solid
fuels; existing plant, 1.1 lb/MBtu for liquid fuels or 1.5 lb/MBtu for solid fuels.

interesting that the secondary standard on S0 2 concentrations is some 100 times


higher than the normal concentrations found in so-called clean air; in some respects
the standards are not very stringent. Still, the electric utility industry is having
difficulty meeting these standards with dirty fuels and abatement equipment that is
expensive and often less effective than desired.
Of all pollutants emitted from fossil-fueled power plants, particulates and S0 2
are considered the most troublesome. They are the two for which abatement
equipment exists. As in reducing thermal pollution using cooling towers, the use of
electrostatic precipitators and S0 2 scrubbers is expensive, and the expenses incurred
are eventually passed on to customers. Again, the cost of pollution abatement
equipment was given in Chap. 1.

4.7.3 Ambient Air Quality Assessment


As the pressure on the electric utility industry increases for the satisfaction of both
emission rate and ambient air quality standards, the system planner may find it
necessary to eliminate economically attractive expansion alternatives because of
noncompliance with air quality standards. We shall discuss S0 2 pollution further
because of its notoriety and because S0 2 abatement equipment has not yet been
widely accepted by the industry.
In essence there are two types of standards: emission rate standards and ambient
air quality standards. Emission rate standards define the maximum amount of
142 Power System Planning

pollutant, SO 2 in this case, that can be emitted into the atmosphere for every Btu/h
heat input. For a new oil-burning plant the maximum acceptable emission rate is 0.8
lb/MBtu, as compared to 1.2 lb/MBtu for coal-burning plants. These emission rates are
difficult to satisfy without using scrubbers because they require the use of hard-to-get
and expensive low-sulfur fuel; fuel with at most I% sulfur is usually required to satisfy
these standards. To illustrate the relationship between emission rate and percentage of
sulfur, we need only note that the emission of SO2 in pounds per hour, assuming the
mth unit is always available, is given below in Eq. 4.39. For probabilistic analysis Eq.
4.39 must be multiplied by Pm.
20S mLmHRm
Qm = HV lb/h (4.39)
m

Sm is the percentage of sulfur; HVm the fuel heating value in MBtu/lb; and Lm HRm
the input energy in MBtu/h, as discussed in Sec. 4.4.1. Emission rate, Q:n, is
the pounds of pollutant emitted for each MBtu/h heat input, or mathematically
I dQm 20Sm
Qm = dL m HRm = - - lb/MBtu (4.40)
HVm
Given approximately 18.5 MBtu/lb of oil, the emission rate for a unit burning 2.4%
sulfur fuel is
I 20(2.4)
Qm = _ = 2.59 lb/MBtu
18 5
Unfortunately this is three times the 0.8 lb/MBtu maximum. Working backward to find
the percentage of sulfur in fuel that could be used, we get
Q:nHVm
Sm = 20
= 0.74%
This simple example illuminates the position of the electric utility industry. It is
exceedingly expensive and difficult to even get I% sulfur fuel. The solution to this
problem is either to use scrubbers to remove SO 2 from stack gases or to accelerate
research and development of techniques to clean dirty fuels.
Even though emission rate standards cannot be met without the use of scrubbers,
ambient air quality standards can be met by judiciously choosing
I . Plant locations 3. Fuel sulfur content
2. Stack heights 4. Commitment schedules
This approach to ambient air quality control should be considered temporary, to be
superseded as soon as clean fuels or abatement equipment can be developed to meet
the more stringent emission rate standards.
Plant locations, stack heights, and commitment schedules can be used to help
maintain ambient air quality at a reasonable level because ground level pollutant
concentrations depend on meteorological conditions. The ability of the atmosphere to
disperse pollutants has been used since time began; however, we are stressing its
Generation System Cost Analysis 143

capability to handle the millions of tons of SO 2 emitted in the United States by power
plants. In order to evaluate how the ambient air quality can be improved using these
methods, an assessment technique is needed that will enable the system planner to
calculate SO 2 ground-level concentrations in predefined portions of the service area for
each week or month in the study period. Such a technique must facilitate the inclusion
of forecast. uncertainty, unit forced outages and maintenance schedules, energy
interchange .transactions, and meteorological data. In addition, the technique must
easily accept unit locations, stack heights, fuel type and sulfur content, and
interchange agreements as input parameters, and it must be able to determine
production costs.
The basic approach presented in Sec. 4.4 is suited to do this job, at least in
terms of providing the capability to determine the ECS for a given expansion
alternative, including all the necessary load and generator features over a predefined
study period. What we do not have is a pollutant dispersion model that can be used to
calculate ground level SO 2 concentrations, given the pollutant emissions, stack heights,
and locations of the units as well as the requisite amount of meteorological
information. The most common pollutant dispersion model is the Gaussian plume
model, which assumes that the plume containing the pollutants emitted from
fossil-fueled power plants is Gaussian distributed in both the y and z directions, as
shown in Fig. 4.21. The plume motion in the direction of the wind is assumed to be
due only to the wind and not to diffusion. Mathematically, the pollutant concentra-
tions in g/m 3 at any point x, y, z m from the stack is given by

Q) 1
x(x,y,z) = (u ...fiiioy exp -2 o/
( 1 y
2
){ 1 [ 1
...fiiioz exp L-2
(z -H)
Oz
2
]

+ 1 exp r_ .!. f.z + !f\2]} (4.41)


ffioz L Oz }
2\

where fl m/s is the mean wind speed at the top of the stack; Q g/s is the emission
from the stack; and the zero-mean Gaussian term in y models the diffusion in the y
z

/
/
T /
/

/
/

Fig. 4.21 Dispersion of stack gases in y and z directions.


144 Power System Planning

direction. The two Gaussian terms in z with means of H m, the effective stack height,
model the diffusion in the z direction. Two Gaussian terms in z appear because the
reflections of the plume from the ground are simulated using the concept of mirror
images: the actual plant whose plume is reflected by the ground can be modeled as
two plants with the ground removed, as shown in Fig. 4.22, a and b. Because the wind
does not always blow from the same direc'tion, the affected area around the polluting
source is normally divided into 16 radial sections, as shown in Fig. 4. 23, where the
width of each section, d m, x m downwind from the source is given by
2rr
= (4.42)
d
16 x
which we shall need later.
With the realization that we are interested in ground level concentrations, z =0
and Eq. 4.41 becomes

x(x,y, 0) = urrayaz
Q
exp
l\-21 H2) ( 1 y2 )
az2 exp -2 a} g/m3
(4.43)

Plume

I II I
I I I
I I I

(a)

Fig. 4.22 (a) Actual plume disper-


sion including ground reflections.
(b) Mathematically equivalent
// plume dispersion.
I/
I/
/// (b)
_,,,
_..,.
//
/

--- .......
'
'
Generation System Cost Analysis 145

Wind~

Fig. 4.23 Radial grid system used in ground-level ambient air


quality calculations.

From Eq. 4.43 the expected value of X, e(x), can be obtained for a given downwind
distance e(x) since both Oy and Oz depend only on x. Further, since y and Q are
independent random variables,

x(x,y,Q) = ff x(x,y,Q)f(y,Q)dQdy

= Jf x(x,y,Q)f(y)f(Q)dQdy (4.44)

where f(y, Q) = f(y)f(Q) is the joint probability distribution describing y and Q. With
Eq. 4.43 substituted into Eq. 4.44

x(x,y,Q) = un1zoy exp[-(~)] fexpt(t;)Jt(y) Qf(Q)dQdy

- _ Q
U1f0 2 0y
exp [- .!.(H)
2 Oz
2
] fexp [- .!. fy )~ f(y) dy
2 \ay J ( 4.4 5)

For values of x more than 0.6 miles downwind and y assumed to be uniformly
distributed with each of 16 radial sections d = x2n/16 m in width,

f(y) =~ for -! < < ! y


146 Power System Planning

and

f [ (y) 1
expl-2 ay
2
]
f(y)dy =
32ay
2../21tx (4.46)
y

Substituting Eq. 4.47 into Eq. 4.45 and multiplying by the probability, fn, that the
wind blows in the direction of the nth section being considered, we get that the SO 2
concentration at a point x m downwind from the mth unit is

{4.47)

where Om is the expected value of the emissions, E(Qm) in g/s, emitted by the mth
unit. Multiplying Eq. 4.39 by 12.65 gives the emissions in g/s. Further, by definition

(4.48)

but f(Qm)dQm = j(Lm)dLm because Qm increases monotonically with Lm. Thus

e(Qm) = f Qmf(Lm)dLm

{4.49)
= /F(Lm)dQm

But from Eq. 4.39


dQm 253SmPm dLmHRm
= HV dL g/s
(4.50)
dL m m m

From Sec. 4.4.1

dLmHRm = IHR
dL m m

Thus
253SmPm {4.51)
dQm = HVm IHRmdLm

Upon substituting Eq. 4.51 into Eq. 4.49,


Generation System Cost Analysis 147

253Sm
- UFC HV T e(FCm) g/s (4.52)
m m
Equation 4.52 is the expected value of the em1ss10ns needed to complete the
expression for e(x). Substituting Eq. 4.52 into Eq. 4.47, we finally have

513.59fnSme(FCm) exp [-(Hm/z)2] 3


Em(X) = UFC HV Ta xu g/m (4.53)
m m z n

where Hm is the effective stack height, which is the actual height of the plume as it
turns and becomes parallel with the ground, and Oz is the vertical dispersion constant,
which defines the plume spread in the vertical direction. The two commonly used
estimates of stack height and vertical diffusion are

(4.54)

(4.55)

where HA = actual stack height, m


u = mean wind speed at top of stack, m/s
g = acceleration of gravity, m/s 2
Vs = stack gas vertical velocity, /s
D = diameter of stack, m
Ts= gas exit temperature, K
TA = ambient temperature, K
a, b = vertical dispersion constants (a= 0.32, b = 0.78 for neutral stability
conditions)
Equations 4.53 through 4.55 define very nicely the expression for the expected
value of the SO2 concentration at a point x m downwind from a power plant. As can
be seen, the system planner has a great deal of flexibility in choosing and studying the
effects of stack heights, fuel sulfur contents, wind velocities and frequencies, as well as
of plant locations, on ambient air quality. Because the basic approach to production
costing was utilized in the development of the assessment procedure, production costs can
be directly related to changes in fuel type and sulfur content, unit maintenance schedules,
and interchange transactions made in the attempt to satisfy ambient air quality standards.
148 Power System Planning

By utilizing Eq. 4.53, the cumulative effect of all fossil-fueled units on the
ambient air quality in a given service area can be assessed and, if necessary, used to
eliminate or modify expansion alternatives.

4.8 SUMMARY
This chapter concludes the discussion of generation system expansion analysis;
emphasis was placed on evaluating the economic impact of a given expansion
alternative. Through the use of plant cost analysis including capacity cost, production
cost, and O&M cost, the basic philosophy of choosing the most economical expansion
plan that also satisfies the desired reliability criteria was presented. It was shown that
capacity cost depends on the unit capacity cost, the fixed charge rate, and the
expected value of the energy produced during the time period under study. To obtain
the expected value of energy to be produced by each generating unit as well as the
production costs, a detailed production analysis technique was proposed in which
forecast uncertainty, random unit outages, economic commitment, etc., can be
simulated to model the system realistically. Special emphasis was placed on developing
the ECS and corresponding production cost for expansion plans containing fossil,
nuclear, and conventional and pumped hydro units.
A brief discussion of corporate models was included to illustrate how and where
generation system costs impact on the financial structure of a modern utility. Although
only a cursory treatment was given to this topic, the student should appreciate the
need for a utility to maintain a financial posture that will support the cost of
construction, for internally generated funds are woefully inadequate to support the
construction activity of most utilities.
For completeness, an entire section was devoted to a discussion of how to treat
energy transactions and off-peak energy sales. It was found that both concepts can be
easily included in the simulation of an expansion plan. Off-peak energy sales, perhaps
stimulated by lower rates, can easily improve the overall operation of a modern electric
utility by improving capacity factors and by making more efficient use of installed
capacity.
The chapter was concluded with a few comments about the cost of complying
with air and water pollution standards now in force. Of the two, air pollution
compliance appears to be more difficult because of the nature of dirty fuels and the
lack of commercially and economically feasible sulfur removal, or desulfurization,
equipment as of this writing. Until such equipment is developed, some control over
ambient air quality can be affected by judiciously using low-sulfur fuels, stack heights,
etc.

PROBLEMS
4.1. To appreciate the impact cost of capital has on the capacity cost of generating
units, change the rate of return on the debt capital used in Table 4.1 to 12%, and
recompute the FCR for the two 1,000-MW nuclear power plants. Using these new
Generation System Cost Analysis 149

and old FCRs, calculate the change in capacity cost for the nuclear plant,
assuming
UC = 500 US$/kW CF= 0.8
4.2. Rework Table 4.1, but assume that the units are depreciated over 40 yr instead of
10 yr. By what percentage does the FCR change?
4.3. Rework the system cost example in Sec. 4.1.4, assuming the new capacity
mixture of 50% fossil, 45% nuclear, and 5% peaking. Compare this result with
that obtained in the example. Is the trend toward more nuclear capacity justified
according to your answer?
4.4. Using a discount factor (3 = 0.1, calculate the present worth of the annual
capacity costs for the following two expansion plans:

Plan 1 Year I 2 3
Unit p
type N F

Plan 2 Year 1 2 3
Unit p F N
type

Further, assume
UCP = 150 US$/kW VCn = 600 US$/kW VCt = 300 US$/kW
FCRp = 0.17 FCRn = o.17 FCRt= 0.17
CFp = 0.1 CFn = 0.8 CF1= 0.8
Cp =200MW Cn = 1,000 MW Ct= 1,000 MW
Based only on capacity cost and the 3-yr planning horizon, which plan is most
economical?
4.5. Determine the ECS for the three units described in Prob. 3.1 and the load
described in Prob. 3.2. To simplify the analysis, assume that each unit can be
committed in one segment and that the hydro unit is base-loaded. Further assume
HR1 = 12 MBtu/MW HR2 = 11 MBtu/MW
HV1 = 6 MBtu/bbl HV 2 = 24 MBtu/ton
UFC1 = 8 US$/bbl UFC 2 = 30 US$/ton
T = 168 h
In addition, calculate the E(Ei), E(E 2 ), E(L), and the total production cost.
Finally, detennine how much fuel each fossil unit will require.
4.6. Repeat Prob. 4.5, but now assume the hydro unit is committed to peak shaving
12,000 MWh during the week in question.
4. 7. Using the nuclear fuel cycle cost estimate, Table 4.4, and the design parameters
given in Table 4.3, calculate the nuclear fuel cycle cost for a batch (one-third of
the core) with the economic assumptions defined in Table 4.2a updated to reflect
current conditions. For instance, yellowcake is now US$25-35/lb instead of
US$7.50/lb.
150 Power System Planning

4.8. In Prob. 4.6 the hydro unit was assumed to be a conventional unit committed to
peak shaving. Now assume that the unit is a pumped hydro unit and hence the
12,000 MWh used to peak shave must be replenished by the fossil units. Calculate
again the ECS, unit energies, and production cost. In addition, compare the
capacity factors of the fossil units computed for this problem with those of Prob.
4.5. To simplify the analysis, assume that the pumping and generating efficiencies
are equal; that is, the energy consumed in pumping equals the energy produced to
peak shave.

BIBLIOGRAPHY
1. Bader, J. F. et al.: Estimating Average Nuclear Fuel Costs, Power Eng. December 1969.
2. Bibliography on Engineering Economics, IEEE Trans., vol. PAS-91, pp. 1976-1984, 1972.
3. Corbian, M. J.: Optimal Pumped Storate Operation with Interconnected Power System, IEEE
Trans., vol. PAS-90, pp. 1391-1399, 1971.
4. Federal Register: National Primary and Secondary Air Quality Standards, vol. 36, no. 84, p. 8186.
5. Grant, E. and W. G. Ireson: "Principles of Engineering Economy," 5th ed. Ronald Press, New
York, 1970.
6. Glossary of Electric Utility Terms, Financial and Technical, EE/ Pub!. No. 61-31, September
1961.
7. f'.oodrich, J. M. et al.: Stochastic Production Costing with Nuclear Unit Generation and Refueling
Constraints, Proc. Am Nucl. Soc. Conf, Ann Arbor, Mich., 197 3.
8. Jenkins, R. T. and D. S. Joy: An Electric Utility Optimal Generation Expansion Planning Code,
Oak Ridge Nat. Lab. Rep. ORNL-4945, July 1974.
9. Miles, W. T. and L. C. Markel: Simulation Methods for Nuclear Production Scheduling, 1975
Engineering Foundation Conference sponsored jointly by ERDA/EPRI, Henniker, N.H.
10. Power Special Report: Cooling Towers, Power, March 197 3.
11. Sager, M. A. and A. J. Wood: The Corporate Model Program for System Planning Evaluations,
IEEE Trans., vol. PAS-91, May-June 1972.
12. Sager, M. A. et al.: A New Generation Production Cost Program to Recognize Forced Outages,
IEEE Trans., vol. PAS-91, September/October 1972.
13. Sullivan, R. L. and D. W. Hilson: A Comprehensive Probabilistic Technique for Ambient Air
Quality Assessment, PICA Proc., 197 5.
14. Turner, D. B.: "Workbook of Atmospheric Dispersion Estimates," U.S. Department of Health,
Education, and Welfare, NTIS No. PB-191-482, 1970.
POWER-FLOW ANALYSIS

At this stage in the planning process it is necessary to investigate various alternative


ways in which the generation system can be interconnected to the load centers. In
effect, this amounts to developing a number of feasible transmission system expansion
plans and, with the help of a variety of analytical tools, identifying the one plan that
best satisfies the desired performance criteria. We might consider the task before us
impossible given the large number of ways to achieve the desired performance goals.
Although the complexity of the task cannot be minimized, the constraints that must
be respected in planning transmission systems drastically limit the feasible alternatives.
Not only do transmission systems have to satisfy rigid technical and financial
constraints, they must also comply with a host of environmental constraints that are
altered and increased almost daily.
The criteria that all transmission systems must satisfy can be defined by a rather
simple statement: A transmission system must be able to transmit electric energy
economically and reliably from generation centers to all load centers at an acceptable
voltage level and in the quantity desired through environmentally acceptable corridors
using environmentally acceptable transmission lines and towers. This broad statement
of performance is not quantitative at all; it is a qualitative guideline at best. Since a
transmission system must eventually be built, the guideline must be interpreted. For
151
152 Power System Planning

our purposes a reliable transmission system is a system that can withstand at least
single-contingency outages: the loss of a line, transformer, etc. Withstand here means
that electric energy can be made available to all load centers at the proper voltage level
in the quantity needed without overloading any system component. In some cases an
even more stringent constraint, whereby the system must withstand certain double
contingencies, is imposed when a higher reliability is essential because of the nature of
the load center in a given area. An acceptable voltage level is the voltage range that, if
maintained, enables voltage control devices in the system to regulate load center
voltages within an acceptable range determined by customer requirements. Environ-
mentally acceptable corridors can handle modern EHV lines without violating regional
land-use plans or unduly degrading the natural beauty of public properties. Finally,
environmentally acceptable lines and towers are not sources of visual pollution and
radio frequency noise.
Since a transmission system must first be able to satisfy the broad criteria stated
above when all components are in service, we focus in this chapter on the problems of
analyzing system expansion alternatives, assuming no contingencies. Our topics include
1. Transmission corridor selection
2. Steady-state component models
3. Power-flow analysis
4. Optimal power-flow analysis
5. Sensitivity analysis
6. Stochastic power-flow analysis
In Chap. 6 a lengthy discussion of contingency evaluation will be presented, including
both deterministic and probabilistic methods.
The first step in investigating a transmission system expansion plan is to define
the capacity and location of all existing and proposed generation and load centers as
well as the existing transmission system. Since plant-siting evaluations usually include
considerations related to load center sites and transmission corridors, information
about corridors is available. Corridor information, when superimposed on service area
maps, facilitates determining the physical and electrical characteristics of lines that may
be constructed within the corridor. In most cases the voltage level and line design are
dictated to a high degree by the amount of power to be transmitted, the characteristics
of the existing system, and the width of the corridor. For instance, for a particular
expansion plan a corridor may be required to transmit more power than is possible
with any available configuration of 230-kV lines, forcing the planner to consider 500
or 765 kV. The choice between 500 kV and 765 kV would be based, assuming no
corridor constraints, on compatibility with the existing system, anticipated growth, and
economics.

Making such decisions requires a full understanding of the utility involved, the
future plans of the utility, growth patterns, etc., all beyond strictly analytical
techniques. Hard engineering facts coupled with good judgment in using analytical
tools give the key to making proper decisions on questions of this nature. In effect,
given corridors and a basic understanding of the existing system as well as future
Power-Flow Analysis 153

needs, utility engineers can easily select a feasible initial transmission system and/or
feasible alternatives to be analyzed for compliance with the desired performance
criteria. This is not to say that certain plans will not be discarded early in the analysis
or that initial plans will not be modified in the process of seeking the best expansion
plan; indeed they will. This is the essence of system planning as we know it.

5.1 CORRIDOR SELECTION


Selecting corridors through which to route transmission lines begins transmission
system planning. The availability and characteristics of corridors often dictate voltage
level, power level, and line type. Since the industry is moving toward higher
transmission voltages, acquiring corridors will become increasingly difficult. In the
United States, where 765-kV AC is now the highest transmission voltage, corridors
must be used ranging in width from 200 to 300 ft. For a 100-mile 765-kV line, land
requirements soar to 3,659 acres. At an average cost of US$2,000 per acre the corridor
cost alone can exceed US$7 million.
Selecting transmission corridors in open country is difficult, but standard
overhead transmission technology can be used with such corridors. In highly urbanized
areas, routing transmission overhead is fast becoming an unacceptable alternative, visual
pollution being a major objection. Urban transmission corridors are going undergound,
necessitating a major deviation in transmission technology. Although underground
cables have been used for many years, the transmission capability now required to
serve major load centers demands cables with much higher capacity than presently
used. To increase the capability of underground transmission systems, forced cooling
using various gases is an alternative being investigated. The use of supercooled, so-called
"cryogenic cables" with little resistance and hence no heat dissipation at low
temperatures is being tested. Given the cost differential between underground and
overhead transmission systems and the problems inherent in the refrigeration system
needed for underground systems, the overhead line remains the industry's workhorse,
but the industry as a whole feels increasing pressure to move more rapidly to
underground systems.

5.2 STEADY-STATE COMPONENT MODELS


Let us briefly review the basic models used for each component in power-flow analysis.
To this end the mathematical models for the following will be presented:
I. The synchronous generator
2. The power transformer
3. The transmission/subtransmission/distribution system
Power-flow studies provide the planning engineer with information concerning the
system when it is operating under normal steady-state conditions; hence the
mathematical models we need are per phase models that apply under normal
steady-state conditions.
'
154 Power System Planning

5.2.1 The Synchronous Machine


The synchronous generator is the main energy converter in an electric power system. It
bears the major responsibility of converting mechanical turbine power into electrical
power at a desired frequency and voltage. In the United States the desired frequency is
60 Hz, and the desired voltage is usually 115, 230, 500, or 765 kV; these were the
most common transmission voltages in 1976.
Since in power-flow studies we are interested only in the steady-state power
output of the synchronous generator, obtained knowing merely terminal voltage and
current, we need only model the synchronous generator as a constant voltage source.
This is perhaps the simplest model of the generator, as shown in Fig. 5.1.

5.2.2 The Power Transformer


The power transformer is an important device in the overall makeup of a modern
energy system. Transformers make it possible for power to be transmitted at high
voltage and hence low losses to load centers, where it is transformed into power at a
voltage level compatible with customer requirements.
The per phase model of a standard voltage transformer is shown in Fig. 5.2.
However useful in other cases, this model does not suit our purposes. In Sec. 5.3.2 we
shall write the injected power equations in terms of the network bus admittance
matrix, which is only possible if all branch elements (including transformers) are
resolved into rr equivalents; that is, we need to transform the transformer model shown
in Fig. 5.2 into the model shown in Fig. 5.3.
To obtain the rr-equivalent model of the transformer, we identify the three new
elements in our rr equivalent in terms of actual transformer parameters using the model
in Fig. 5.2.
WI Lo

Fig. 5.1 Generator model.

l
--
Power-Flow Analysis 155

a:1

Vp,, ~ !>-----'.___z_L_ __.i----ls-ec--e Vsec

Fig. 5.2 Transformer model.

Vpri Vsec
/pri = a2zL - aZL

or in a more convenient form


1 1
lpri = a2zL Vpri - aZL Vsec

1 1
1sec =- -Z V pri
a L
+ z-Vsec
L

Similarly for the model in Fig. 5.3


I

/pri = (Y1 + Y3)Vpri - Y3Vsec


1sec = -Y3 Vpri + (Y2 + Y3)Vsec

Upon equating like coefficients, we obtain Y1, Y2 , and Y 3 in terms of ZL and a.

1
or Y3 = --
aZL

Fig. 5.3 1r-Equivalent model.


1 56 Power System Planning

or

Yi = a 2 ~i, - aiL = aiL ( - i)


I I I
Y2 = - - - = - ( a - I)
ZL aZL aZL
y - _I_
3 - aZL

With these equations and the model shown in Fig. 5.3, the transformer model needed
for power-flow analysis is complete and in the desired form.
In the case of phase-shifting transformers that cannot be expressed in terms of a
rr equivalent, it is common to use the model shown in Fig. 5.2. When the bus
admittance matrix is developed, buses connected to phase-shifting transformers will
have self-admittance terms of the form -
I '
Y-=
11 --+
aa*ZL
i = primary bus
or

y .. = _I + .. j = secondary bus
ll ZL
The transfer admittances become:

and
I
yji = aZL
Clearly, if a= a*, as is true for regular voltage transformers, the model shown in Fig.
5.3 results.

5.2.3 The Transmission/Subtransmission/Distribution Lines


The final model to be considered is that of the transmission/subtransmission/
distribution lines themselves. Since the models for each of these elements are identical,
except for the values of parameters used in the models, it is only necessary to consider
a general line X m long, with resistance, inductance, conductance, and capacitance per
meter given by R, L, G, and C, respecti..vely. These electrical parameters are calculated
using expressions determined from evaluating the fields, electric and magnetic, for the
line considered. Clearly, each of the parameters R, L, G, and C depends on the tower
configuration, spacing, and capacity of the line. We shall not dwell further on methods
for calculating the basic electrical line parameters, for it is an elementary topic covered
in many texts.
-
Power-Flow Analysis 157

Assuming we know the inductance, capacitance, and resistance per unit length of
line, and assuming we can neglect the coupling between phases (which is usually valid,
because either the coupling is negligible under normal steady-state conditions or the
lines are transposed to eliminate the coupling), we obtain the per phase 1r-equivalent
model of a line by solving the long-line partial differential equations that describe the
differential length of line depicted in Fig. 5.4.

ov(x,t) .( ) Loi(x,t)
--~=
ax -Rzx t -
, ---
at
ai(x,t) cav(x,t)
---'-~ = -Gv (x t) - --'----'-
ax ' ar
where v(x, t) and i(x, t) are the voltages and currents at time t and distance x along
the line.
Using the fact that we desire a sinusoidal steady-state model, we can assume

v(x,t) = V(x)eiwt
i(x,t) = I(x)eiwt
and hence remove the time dependence from the equations, leaving only two ordinary
coupled differential equations to solve. Upon substituting v(x, t) and i(x, t) into the
partial differential equations, we get

. dV(x) t
eJwt - ~ = -RI(x)e'wt - jwLe 1 w l(x)
dx
_
eJwt dl(x)
_ = -GV(x)e1w t - jwCe1w t V(x)
dx .

Dividing through each equation by eiwt, we have the coupled ordinary differential equations

R___E_ L-h-
v(x,t) meter meter v(x + dx, t)
- - - - - - W v - - - ~ - -.....-~,_____
i(x, t)

IF
lcm""
i(x + dx, t)

Fig. 5.4 Differential line model.


158 Power System Planning

dV(x) = -(R + jwL)/(x)


dx

di(x) = -(G + jwC)V(x)


dx

The proof is left to the reader, but it is simple to show that the solutions to these two
equations are

V(x) = V(0) cosh -yx - Zwl(0) sinh rx (5.1)

I(x) = /(0) cosh-yx - zV(0) sinh -yx


w
(5.2)

Zw is the characteristic impedance of the line, and -y is the propagation constant.


Both parameters are functions of R, L, G, and C. By definition

R +jwL
z w-
- G+jwC
-y = V(R + jwL)(G + jwC)

V(0) and /(0) are the peak values of the current and voltage at the point x = 0,
or at the beginning of the line. Llkewise, V(x) and /(x) are the peak values of voltage
and current at a distance x meters down the line.
For our purposes V(0) and /(0) are the voltage and current at the sending end of
the line, whereas V(x) and /(x) are the voltage and current at the receiving end of the
line. In effect, we have a lumped parameter mathematical model of a line x m long.
All we need do is use these equations to obtain the per phase 7T equivalent of the line,
beginning once again with the general 7T equivalent shown in Fig. 5.3 and replacing Vpri
and /pri with V(0) and /(0), Vsec and 1sec with V(x) and I(x). Writing equations for
V(x) and /(x) and equating coefficients as before, we finally obtain Fig. 5.5.

y == 1
3
Zw sinh -yx
V(O) ------...-----1,_______,,__________ V(x)

/(0) /(x)

1 -yx
Y, == Z- tanh - == Y 2
w 2

Fig. 5.5 1r-Equivalent line model.


Power-Flow Analysis 159

5.3 STEADY-STATE POWER SYSTEM MODEL


Given the models for each component normally found in a modern power system, a
single phase diagram of the system can be drawn, and the corresponding net injected
power equations as well as interchange equations can be defined. For an N-bus area the
net injected complex power equations are:
P; - jQ; = V{I; i=l, ... ,N (5.3)
For convenience Eq. 5.3 can be separated into its real and imaginary parts, giving
P; = Real(V;*/;) = PGi - PDi
(5.4)
Q; = -Im(V;*I;) = QGi - QDi i = 1, ... ,N

Pc; and QGi are the generated outputs at bus i, and PDi and QDi are the demands at
bus i. The net injected current into bus i, I;, can be expressed as

(5.5)

Y;j are the elements of the familiar Ybus matrix obtained from the electrical diagram
of the area being modeled.
For interconnected area studies in which each area must satisfy a net interchange
schedule, additional equations are needed. If lk is defined as the net MW interchange
for area k, then

(5.6)

where

lk = Real ( ~ v;(vp - Vq)Yp0


(5.7)
PEk
qff-k

Ypq is the line admittance; p Ek denotes that the sending end bus p is in area k, and
q $ k denotes that the receiving end bus q is not in area k.

5.3.1 Variable and Parameter Constraints


In addition to satisfying Eqs. 5 .4 and 5 .7 each area has certain parameter and variable
constraints that must be respected. We know, for instance, that
1. The MVA capacity of each component ISjl, j = 1, ... , E, must not be
exceeded if possible.
2. Voltage levels IV;I, i = 1, . .. , N, must not fall outside a given range.
160 Power System Planning

3. The tap setting at, t = 1, ... , T, of automatic tap-changing transformers must


not be allowed outside a given range.

These constraints may be stated more rigorously as follows:

ISi I ..;;; ~ i = 1, ... , E


lf;I ..;;; IV;I..;;; IV;I i = 1, ... ,N (5.8)
t=1, ... ,T

The bar subscripts and superscripts indicate lower and upper limiting values for each of
the parameters and variables, respectively. In the course of solving a particular
power-flow problem, these limiting values should not be violated. Provisions must be
made in solution procedures for efficiently dealing with these constraints.

5.3.2 State Variable Model of System


That the equations presented in Sec. 5.3 .1 are cumbersome is obvious. To circumvent
this problem and at the same time provide a methodical variable classification
technique, we shall resort to state variable notation that has been used successfully in
control theory. In essence there are basically three types of variables: dependent state
variables X;, independent control variables u;, and disturbance variables d;.
These variables can be compactly arranged in state, control, and disturbance
vectors; that is, for area k with Gk generators, Tk transformers. and Nk buses '

Xk ~ state vector=

Uk ~ control vector = (5.9)

Dk ~ disturbance vector =

In order to express Eqs. 5.4 and 5. 7 in state variable form, we must rearrange the
equations, starting by substituting Eq. 5.5 into Eq. 5.4 to give
Power-Flow Analysis 161

P; = Real ~;* l V;)


Y;;

Q; = -Im (vt lY;;V;)


and then defining
A O
VpglVple 1 P
Ypq ~ Gpq + jBpq = -(gpq + ibpq)
.i;. ~
P; = Real ( IV;I e-1 '~(G;; + jB;;)IV;le 1 'J
-6~

= Real ~V;I L(Gij + jB;;)1V;I exp [j(o; - o;)~

= L(IV;IIV;IG;; cos (o; - o;) - IV;IIV;IB;; sin (o; - o;)) (5.1 O)


j

Similarly

Q, - -[f IV1IIV;IB11 cos (6; - 61) + IV;IIVJIG1; ,m (6 1 - 6j (5.11)

lk = Real \' o -6 -6
~ IVp le-1 P(IVP le' P - IVqle+1 q)(gpq + jbpq))~
(
pEk
q(J.k

= l 2
(1Vpl gpq - IVpllVql(gpq cos (oq - op) - bpq sin (oq -op))) (5.12)
pq
pEk
qff.k
Moving all the terms in Eqs. 5.10 and 5.11 to the left-hand side of each of the
equations above, we obtain the 2Nk mismatch equations expressed in polar form,
where for simplicity the bus numbers are the area designations and not the absolute
interconnected system designations that will be used later.

= P1 - LIV1 IIV;IG,; cos (o; - 01)- IV1 IIV1 1B1; sin (o; - 01) = 0
j
(5.13)
fNk = PN - L'VNIIV;IGN; cos (o; - oN)- IVNIIV;IBN; sin (o; - oN) = 0
j
162 Power System Planning

~Nk+i = Q1 + _llVdlVjlBti cos (oi - 01 ) + IV1 11Vi/G1i sin (oi - 81 ) = O


j

f2Nk = QN + 2I j
VNIIVjfBNj cos (oj - oN) + IVNIIVj/GNj sin (oj - oN) =0

(5.13 continued)
To avoid perpetuating notation difficulties, only the mismatch quantities will reflect
the difference in area sizes through the use of the subscript Nk. It should be
understood that all variables defining the mismatches, such as PN, VN, etc., are really
PNk VNk Dropping the k designation will simplify the expressions and will not be
confusing. For single-area problems the control vector . U, state vector X, and
disturbance vector D for area k are defined in terms of the area bus designations as
follows:

/Vil
01
QDI
IV2/
Pn1
Pe2
QD2
6 Qee 6
Uk= xk~ Dk= Pn2 (5.14)
/Ve/ oe
Pee IVe+1 I
0e+1 QDN
ll I
PDN

As we all know, it is common practice to identify one bus as the reference bus;
all voltage angles are measured relative to that reference (any generator bus could act
as the reference bus). For our purposes we assume that the reference bus is bus I and
call the generator connected to that bus the slack generator. This generator balances
the injected power equations and hence is a dependent variable. All other generator
outputs PGi are independent and appear in U. We should note that all of the
independent variables in U can be controlled, in practice. The generator bus voltage
magnitudes can be controlled by adjusting generator excitation set points, and the
nonslack generator real power outputs can be controlled by adjusting steam valve set
points.
Little need be said of the validity of defining the state and disturbance variables
as we have. The definitions assumed are reasonable and compatible with the way
Power-Flow Analysis 163

operators view these variables. Using the word "disturbance" to describe the area loads
is intended to connote that those variables change at will and are not controllable like
state variables. Certainly there are exceptions to this statement, the most notable being
load shedding itself, but before we try to destroy the model proposed, we should keep
in mind that ac,cording to our assumption, the area is operating in a normal mode.
Load shedding is a control mechanism used when the area is no longer in the normal
operating mode.
Using definitions given for the state, control, and disturbance vectors, along with
Eq. 5.13, we may describe the area equations for area k compactly.

=O (5.15)

If we let Wk denote the vector of 2 line flows, E being the number of elements, and
Hk the vector of line flow equations for area k, and take X from Eq. 5.15, the line
flows Ware

(5.16)

The mth component of Hk, hm, is the line flow through line m of area k, and if the
sending end and receiving end buses are denoted by ij, then
2
hm = S;i = V;(V;*- V/)ylj + IV;l Ylo (5.17)
Equations 5.15 and 5.16 enable us to define the power-flow analysis problem: Given
the control and disturbance vectors u0 and D 0 , determine the state and output vectors
X and W that satisfy Eqs. 5.15 and 5.16. Not only can we use these two equations in
power-flow analysis, but also they facilitate discussing sensitivity analysis, the topic of
Sec. 5.5.
For interconnection studies the system model may be developed using Eqs. 5.15
and 5.16 in conjunction with Eq. 5.12. If we denote the mismatch equations for area
k by Fk(Xk, Uk, Dk), k = l, ... , A, where A is the number of areas, then we have a
set of mismatch vectors Fk(X, U, D), k = l, ... , A, to describe the interconnected
areas. In addition, since we have the requirement that each area meet its net
interchange schedule and that the net system interchange be zero, as defined by Eq.
5.6, the mismatch equations and the state and control vectors must be modified to
describe the interconnected areas accurately. In effect, the modifications required
are:

1. The net injected power equations for all buses to which an interconnec-
tion is made must be modified to reflect the admittance of the intercon-
nection.
2. The net interchange mismatch equation
164 Power System Planning

must be added to the bus mismatch equations.


3. In all but one area lk must be added to the control vector, and at the same
time a generator variable* say P a 2 , must be transferred from the control
vector to the state vector. In area A, IA must be added to the state vector
since it is a dependent variable.

For interconnected areas, then, we have the following model:

F1(X, U1, Di)


L 1 (X)
F 2 (X, U2 , D 2 )
F(X, U,D) = =0 (5.18)

We now have 2~Nk + A = 2Ns + A equations to solve, where Ns is the total number
of buses in the system of interconnected areas, and

IVd
01 IV1I
IV2I 81
IV3l IV2I
U1 Pa3 Pa2
U2 (5.19)
U= Uk - UA =
k*A IVal IVal
UA Paa Paa
a1 aI

*Note that any generator variable may be chosen (it is most common to choose 15 1 ), but to
simplify the numerical example in Sec. 5.4.2, Pa 2 is preferred here.
Power-Flow Analysis 165

Qe1 Qe1
Pei Pei
Qe2 Qe2
Pe2 02
02
X1
X2 Qee
X= xk - Qee XA = lie ( 5 .19 continued)
k*A
lie IVe+1 I
XA
IVe+1I 0e+1
0e+1

IVNI
IVNI ON
ON [A

It should be noted that in Eq. 5.18 each area mismatch vector Fk is now a function of
the system state vector X, which couples the area mismatch equations together.
As in the single-area case, once Eq. 5.18 is solved for the state vector, the flows
throughout the interconnected system become:

W= (5.20)

Stated simply, interconnected-area studies involve solving Eq. 5.18 for X and
calculating line flows from Eq. 5.20. Solving either single-area or interconnected-area
problems is conceptually the same; the dimensionality constitutes the only major
difference between the two. There are minor differences related to the structure of the
coupled equations, differences that are considered in developing computational solution
techniques to solve the mismatch equation in Sec. 5.4.
To summarize, we are faced with either solving 2Nk equations
Fk(Xk, Uk, Dk)= 0
and calculating the 2k flows
Wk = Hk(Xk, Uk)
or solving the 2Ns equations
F{X, U, D) =0
166 Power System Planning

and calculating the flows


W = H(X, U)
depending on whether we are evaluating the performance of the transmission system
for a single area or a system of interconnected areas. In Sec. 5.4 we discuss at length a
method for efficiently solving these mismatch 'equations.

5.3.3 State Variable Model of Three-Bus System

As we shall find it convenient to illustrate the ideas to be presented using a simple


example, consider developing the state variable model for the three-bus system
depicted in Fig. 5.6. For this sample system we have three buses (N = 3), three
elements (E = 3), and one area (A = 1). Assuming that the 77-equivalent models for
each transmission element are as shown in Fig. 5.7, then from Eq. 5.13 the mismatch
equation for the area is:

=0 (5.21)

PG, + iOG,

XFMR #1
m = 1
Po, + iOo,
S21
a, :1

XFMR #2

a 2 :1

#3

Po, + iOo,

Fig. 5.6 Three-bus sample system.

,,,.,,
Power-Flow Analysis 167

Y1
.......,1
'1
XFMR #1
m= 1

(a)

XFMR #2
m=2

(b)

Y1
f\/V\r--'.. ooQ.P PP

1 1
,.I
I'
LINE 1
m=3

t,
m =element# (c)

Fig. 5.7 1r-Equivalent models for (a) transformer 1, (b) transformer


2, and (c) line 1.

where
3

f1 = P1 - l
j=I
[IVillV;IG1; cos (o; - 01) - IV1IIV;IB1; sin (o; - 01)] =0
3

f2 = P2 - l
j=J
[IV2IIV;IG2; cos(o; - 02) - IV2IIV;IB 2; sin(o; - 02)] =O (5.22)

h = P3 - l
i=I
[IV3IIV;IG 3; cos(o; - 03) - JV3IIV;IB 3; sin(o; - 03)) = 0
168 Power System Planning

/4 =Q, + l j= 1
[IV1IIV;IB1; cos(o;-01) + IV1IIV;IG11sin(o;-0 1)] = 0
3

fs = Q2 + l [I V2 II V, IB2; cos (81 - 02) + I V2 II V; I G 2; sin (o; - o2)] = o


j= 1
3

/6 = Q3 + l [I V3 II V1IB3; cos (o; - 03) + IV311 Vi IG3; sin (8; - 83)] =o


i== 1
(5.22 continued)

From Eq. 5.14

V1 Qo1 Qv1
01 Poi Pv1
IV2I Qo2 Qv2 (5.23)
U1= X1= D1=
Po2 02 Pv2
a1 IV31 Qv3
a2 03 Pv3

Substituting the transmission element models into Fig. 5.6, we obtain Fig. 5.8, from
which

1 + Y2 + Y4 + Ys -y1
-y4 ]
= + Y3 + Y1 + Y9 -y (5.24)

L
Ybus -y1 Y1
-y4 -y7 Y4 + Y6 + ~7 + YB

Therefore

Gii = Real (Yii) (5.25)


Bi; = Im (Yi;)
Also from Fig. 5.8 and Eqs. 5.16 and 5.17 we can define the output vector Wk as

(5.26)
Power-Flow Analysis 169

Y,

Y2 Y,

Fig. 5.8 Ekment models and the three-bus system.

where for the sample system

h1(X1, U1) = Vi (Vt- V!)yt + IV11 2yi*~


element 1
h2(X1,U1) = V2(Vf- Vt)yt+IV212yt
2
h3(X1,U1) = V1(Vt- Vt)yt + 1Vil yf~
element 2 (5.27)
h4(X1, U1) = V3(Vt - Vt)yt + IV3fyt
hs(X1, U1) = V2(Vf - Vt)y.f + IV21 2yf ~
element 3
h6(X1, U1) = V3(Vt - V!)y.f + IV312yt

Not only will the equations developed for the sample system serve
us later, but their development has given us an opportunity to apply the
general equations to a specific case. The ease with which the state model
can be developed, not only for this system but for any system, has been demon-
strated.
170 Power System Planning

5.3.4 State Variable Model for


Two Interconnected Areas
For simplicity, consider developing the state model for the two identical areas depicted
in Fig. 5.9a, where each area is the same as the single area considered in Sec. 5.3.3. In
this case the mismatch equations are: '

F 1 (X, Ui,D 1 )
L1(X)
F(X,U,D) = =O (5.28)

/1(X1, U1,D1)
/2(X1,U1,D1)
/3 - Real (V3If6) (5.29)
F 1 (X, U 1 ,Di) = =O
/4(X1, U1,D1)
fs(X1, U1,Di)
/6 - Im (V3Jf6)
/1(X2,U2,D2)
/s(X2, U2, D2)
/ 9 - Real (V6It3)
F2(X,U2,D2) = =O
/1o(X2, U2,D2)
/11(X2, U2,D2)
/12 - Im(V6Jt3)

L 1(X) = / 1 - Real (V3Jf6) = 0 (5.30)


L 2(X) = /2 - Real (V6Jt3) = 0

From Eq. 5.19 the state, control, and disturbance vectors for the two-area
system are defined as follows:

(5.31)
Power-Flow Analysis 171

Area #1 Area #2
/ I

ypq

,,
(a)

Area #1 Area #2
v,
-
v.

-
v, (b)
v,
Fig. 5.9 (a) Two interconnected areas. (b) Three-bus interconnected areas.

where

Qc1 Qc4
IVil IV41
Pei Pc4
01 84
Qc2 Qcs
IV2I IVs I
U1= X1= Pc2 U2 = X2 = 85 (5.32)
a1 Pcs
82 IV6I
a2 a1
IV31 86
11 a2
03 /2

Note in Eqs. 5.29 through 5.32 that the system bus number designations, Fig. 5.9b,
have been used to define the system states. Also
172 Power System Planning

QDI QD4
Pn1 Pn4
QD2 QDS
D1= D2 =
Pn2 Pns
Qn3 QD6
Pn3 Pn6
Finally, the output vector equation is

j
W = H1(X1, U2)l
(5.33)
LH2(X2, U2)J
with H 1 (X1, U1) and H2(X 2 , U 2 ) defined by Eq. 5.27, except that for H 2 (X2 , U2 )
the bus numbers are 4, 5, and 6 instead of I, 2, and 3 as depicted in Fig. 5.9.
We shall return to the equations developed in Secs. 5.3.3 and 5.3.4 many times
during subsequent discussions. Although we might have chosen a more complicated
sample system that would more nicely demonstrate some ideas to come, the increased
complexity of a larger system would mask the points to be demonstrated; therefore we
have proposed the system as discussed.

5.4 NEWTON'S POWER-FLOW SOLUTION METHOD


The structure of the power-flow equations for single- and multiarea systems being
basically the same, we present a solution method for solving multiarea problems,
realizing full well that it can as easily be used to solve single-area problems. Stating
again the problem before us, we want to find the state vector X that satisfies
F(X, U, D0 ) = 0 (5.34)
given U and D 0 , and knowing X, to calculate
W= H(X, U0 ) (5.35)
The solution technique recognized as most efficient for both speed and storage is
the so-called "Newton's method," an iterative procedure for solving a large number of
nonlinear equations simultaneously, using optimally ordered elimination coupled with
compact storage and working row ideas. The acceleration mechanism used in Newton's
method requires LU factors associated with operations in the elimination process. Since
many readers may not be familiar with the details of Newton's method, further
discussion of working rows, elimination, etc., will be postponed until a few
fundamental ideas and definitions have been' provided.

5.4.1 Basic Solution Procedure


Newton's method uses an algorithm for updating the state vector, starting with some
initial vector, x 0 in such a way that after a finite number of iterations,
Power-Flow Analysis 173

F(X, U0 , D0 ) < o:. Each bus mismatch is less than or equal to some desired small mis-
match. To develop the basic algorithm, we must expand the mismatch vector in a Taylor
series about the initial point x O , uO , DO :

F(X, u , D = F(X,u,DO) + !~lx 0 8X + higher terms


O O
~ (5.36)

Dropping the nonlinear terms in AX and solving for AX, we get

aF
AXo=- ( dXxo I )- (F(Xo,u,Do)-F(Xs,U,Do)]
1
{5.37)

Since the solution sought is


F(X8, u, D0 ) =0
Eq. 5.37 becomes

-(!~I Xy
1
0 0 0
AX = F(X , u, D ) {5.38)

where the new state X 1 is


x 1 = x0 + AX0 (5.39)
The matrix 'iJF/aX is called the Jacobian matrix, and AX, the state update vector. For
a general iteration v, Eqs. 5.38 and 5.39 become:

8X" = -(!~I x"Y l F{X", U , Do) {5.40)

(5.41)
The last two equations define the Newton algorithm in its simplest form. Before we
attempt to solve large power-flow problems by applying the algorithm as stated, we
should realize that
I. oF/oX without any reduction in mismatch equations is a (2lls + A) X
(2Ns + A) matrix of first partial derivatives. For a modestly large intercon-
nected area, we could be faced with inverting a 1,000 X 1,000 matrix
involving a million elements. The problems of realizing such a feat are
obvious.
2. Since u 0 contains voltage control parameters that must be adjusted to satisfy
the voltage profile constraints discussed earlier, the power-flow problem
involves finding the solution to a changing problem, which in itself is
palatable but computationally expensive.
Although observation 2 is computationally unattractive, it can be dealt with in a
judicious manner. However, observation I, if not circumvented, restricts the algorithm
to solving relatively small problems, which might be solved more efficiently by other
174 Power System Planning

methods. Fortunately, the Jacobian matrix is a sparse matrix because the mismatch
equations have a sparse structure due to the sparsity of the Ybus matrix. It is perhaps
uncommon to have more than ten lines connected to a given bus; therefore, in a
600-bus interconnected network, each row of the Ybus matrix would have no more
than, say, 12 nonzero terms, which would give rise to no more than 24 terms in the
Jacobian. The sparsity of the Jacobian matrix for this hypothetical case would be
98-99%. The exact numbers are hardly important here, but the point is that most of
the elements in the Jacobian are zero and this sparsity must be exploited.
For illustration we develop the Jacobian matrix for the sample interconnected
system depicted in Fig. 5.9, using the state variable model defined in Sec. 5.3.4. From
Eqs. 5.28 and 5.31 we have

aFi1ax1 aFi1ax2 aFi1ax


aF oLi/oX 1 aL1/aX2 aLi1ax
-= - (5.42a)
ax aF2/ax1 aF2/aX2 aF2/ax
aL2 /aX1 aLz/aX2 aLz/ax
where

0 0

oh of3 of3 oh
- O---- 0
oFk oQGI 083 olV61 086
-- - (5.42b)
ax of4
-- of4
....... - 0 ...
k= 1, ... ,A
0
oQGI 083

Note that because of the interconnection, [ 3 and f 6 are now functions of 8 6 and V6 ;
hence the nonzero terms in the right-hand submatrix above. The expressions needed to
evaluate the partials in Eq. 5.42b are:

1. for all i and j

for all i =j (5.43a)


for all i =fa j
Power-Flow Analysis 175

at; = { 1.0 for all i = j


2.
aPa; o.o for all i =I= j
afi+Nk
= 0.0 for all i and j (5.43b)
aP0 ;

Ns
= a~_P; - IV;l G;; -
J
2
l~-
- p-rl
[IV;IIVplG;p cos (5p - 5;)

- IViii VP IB;p sin (5p - 5;)]


i =l=j { = IV;II V;I G;; sin (5; - 5;) + IViii V;IB;; cos (5; - 5;)
Ns
i=j =- I
p =i-i
[IV;IIVplG;p sin (5p - 5;) + IV;IIVplB;p cos (5p - 5;)]

Ns
= a~;Q;+ l
p=l
[IV;IIVplB;p cos(5p - 5;)+1V;IIVplG;p sin(5p - 5;)]

Ns
= a~_Q;+IV;i Bu+
J
2
l
p,foi
[IV;iiVplB;p cos(5p - 5;)

~ :j. {: -;V;II V;IB;; sin('.; - 5;) + IV;II V;IG;; cos (5; - 5;)
+ IV,.IIVplG;p sin (5p - 5i)]

1-J - I[IV;IIVplB;psm(5p-5;)-IV;IIVpiG;pcos(5p-5;)] (5.43c)


p=i-i

Ns
= a,t_,P; - IV;i2Gu -
J
I
p,foi
[IV;IIVplG;p cos (5p - 5;)

- IV;IIVplB;p sin (5p - 5;)]


i=l=j { = -IV;IG;; cos (5; - 5;) + IV;IB;; sin (5; - 5;)
Ns
i= j = -2IV,.IGu - l [I VplG;p cos (5p - 5;)- IVP IB;p sin (5p - 5;)] (5.43d)
p =pf
176 Power System Planning

i=l=j { = IV;IB;i cos (8i - 8;) + IV;IIG;i sin (8i - 8;)


Ns
i=j = 2IV;IB;; + l
piai
[IVplB;p cos (8p - 8;)+ IVplG;p sin (op - 8;)]

(5.43d continued)

5. (5.43e)

In general the areas are not identical, and the partial derivative expressions given
in Eq. 5.43 must be used to determine each of the submatrices oFk/oXk.
Continuing the process of establishing expressions for each of the submatrices of
Eq. 5.42a, we must now consider

aLk aLk aLk aLk aLk aLk aLk aLk


-------------
ax
where for tie-line buses p and q we have

for all k (5.43/)


j=l, ... ,G

7. :i~I = ol~il lk - l
pq
2
{IVpl Kpq - IVpllVqlfgpq cos (8q - op) - bpqsin (oq - op)]}

pEk
q(/.k

jE k (bus j is in area k)

j=l=p{ =u
j=p = 2
q(/.k
{-21VilKjq+IVql[gjqCOS(8q -8j)-bjqsin(oq -oj)]}

j./.k

(5.43g)
Power-Flow Analysis 177

8. l
pq
{IVpfgpq - IVpllVqlfgpq cos(liq - lip) - bpq sin(8q - lip)]}

pEk
qfi_k

jEk
j=l=p

j=p

and finally

j(t.k

~
aLk
=-
j=l=q
aoi = 0 (5.43h)

j=q t l
,
=
pEk
IVpllVil[-gpi sin (8i - lip) - bpi cos (8i - lip)]

For later use, and perhaps a little practice, consider constructing the Jacobian
matrix for our two-area system using Eqs. 5.42 and 5.43, where the state vector Xis
given in Eqs. 5.31 and 5.32 and Ns = 2N= 6. To simplify matters, assume that the
two areas are lossless; hence Gii = 0. The resulting submatrices are shown on pp. 178-180.
It is instructive to highlight the structure of the Jacobian matrix for the sample
system, as in Fig. 5.10. Note that the area equations are coupled; that is, the
off-diagonal submatrices are nonzero because of the interconnection between the two
areas. Further, we should recognize that for single-area studies the off-diagonal matrices
become null matrices, since setting b 36 to zero forces all the off-diagonal matrices to
zero. Clearly, the two sets of equations describing the two areas involved are now
independent; thus the two power-flow solutions may be obtained independently. This
obvious fact needs mentioning because studies involving a planner's own area are often
required and because reducing the size of the problem saves computer time.

5.4.2 Gaussian Elimination and Optimal Ordering


As stated in Sec. 5.4.1, the basic power-flow problem requires solving the system
mismatch equations for the states, given u and D0 , using the iterative algorithm
defined in Eqs. 5.40 and 5.41. As we pointed out, a straightforward application of the
iterative equations to large systems is not possible, and the sparsity of the Jacobian
must be dealt with effectively. To achieve the degree of effectiveness needed to make
the basic algorithm a viable solution procedure, it is necessary to perform the inversion
of the Jacobian using Gaussian elimination on an optimally ordered Jacobian. As
...
~

1.0 IVi IIV2IB12 cos (Iii - 61) IV1IB13 sin (63 - 61) IV1IIV3IB13 cos (63 - 61)
2N
1.0 - LiV2 IIVplB2p cos (lip -6 2) IV2IB23 sin {63 - 62) IV2IIV3IB23 cos (63 - 62)
pa#2
2N 2N

ilF1
1.0 IV311V2IB32 cos (02 - /i3) l
pa#3
83plVpl sin (lip - 63) -l
pa#3
IV3IIVplB3P cos (op - 63)

ilX1 -1V1IIV2IB12 sin (62 - 01) IV11B 13 cos (63 - 61) -IVi IIV3IB13 sin (03 -Ii,)

1.0 l
pa#2
IV2IIVplB2p sin (lip - Iii) IV2IB23 cos (63 - Iii) -1V2IIV3IB23 sin (63 - Iii)

2N
-1V3IIV2IB23 sin (62 - 63) 21V3IB33 + l
pa#3
B3plVpl cos (lip - 03) l
pa#3
IV3IIVplB3p sin (lip - 03)

(5.44a)

ilF 1 IV31B36 sin (66 - 63) IV3IIV6IB36 cos (06 - 63)


0
ilX2

IV3IB36 cos (66 - 63) -1V3IIV6IB36 sin (66 - 63)


(5.44b)

-:'~C.'1' ~--"

'

. I .
1_ 1.0 _ IV~.10~ "'' c,. _,.) . IV, ID% "" (6, _ ,.) IV,I IV.ID% ,os (6, _ 6,)
7
~ -l~l~B.,,c,.-t1 iI 'ir~ l~~"'a. 6
j ~ I

ax,l u -'\J J - ~ '~ V - ;;J.J

IV31B36 cos (06 - 03) -1V3I/V6IB36 sin (06 - o3) J


(5 .44b)

1.0 IV41/VslB4s cos (os - 04) IV41B46 sin (06 - 04) /V41/V61B46 cos (06 - 04)
2N
-Ip4'5
IVsl/VplBsp cos (op - o5) /V5IB 56 sin (0 6 - o5) JV5I/V6IB56 C(?S (06 - Os)

2N 2N

aF2
IV6I/V5IB6s cos (05 - 06) lB6p/Vpl sin (op - 06)
p4'6
-Ip4'6
w6IIVplB6p cos (op - 06)

ax2 I 1.0 -/V4IIV5IB45 sin (os - 04) /V41B46 cos (0 6 - 04) -/V4JIV61B46 sin (06 - 04)
2N
1.0 l
p4'5
IV5I/VplBsp sin (op - o5) IV5IB 56 cos (0 6 - o5) -/V5IIV6IB56 sin (06 - o5)

2N 2N
-/V6I/V5IB56 sin(o 5 - 06) 2IV61B66 + l
p4'6
B6p/Vpl cos (op - 06) l
p4'6
lV61/VplB6p sin (op - 06)

(5.44c)

aF2 IV61B63 sin (03 - 06) /V31/V61B63 cos (06 - 03)


0
. I ax1

IV6 IB63 cos (03 - 06) -/V31/V61B63 sin (06 - 03)


---'
(5.44d)
...
~
;,
~t{'
~-'
,-... ,-... ,-...
~
"Sf" $ ~
-::::
'ST
": ": ": ":
.._,,
V)
.._,,
V) .._,,
V) .._,,
V)
......,
,-...
'-0"'
.I
e"'
"'
-q
......, 0
u
"'
"'
.,:,
:::
~
I
......, ,-...
,-... '-0"' ,-...
'-0"' ,-... "'
'-0
'-0"'
"'
.._,,
'-0
"'
.._,,
'-0
e"'
"'
0 "'
Cl)
0
u .._,,
'-0
"'
0
u u
"'
"'
"'
.,:, .s "'
"' "' "'
.,:,
:::: ::: "'"'
.,:,
.,:,
::::
~
~ ::: ~
I I
,-... ,-... ,-...
"'
'-0 '-0"' "'
'-0
"'
.._,,
'-0 "'
.._,,
'-0 "'
.._,,
'-0
.s .s
en
.s
en
"'
"'
"'
.,:, "'"'
.,:, '"'
.,:,
::: ~ ~
I
II II II II
..r1x
1'0
..r1x- .s1x
l'0 l'0 l'0
.s1x-
l'0 l'0 l'0
1'0
180
Power-Flow Analysis 181

I
I aF, (2NX 2N+ 1)
I ax 2

I
I
I I
----_-_-_-_~_ +-~-=-------=---=- I

oF 2
ax, (2N X 2N + 1)  F 2 (2N X 2N + 1)
ax 2

a L2 1ax 2

Fig. 5.10 Structure of the Jacobian matrix for the two three-bus inter-
connected areas.

Gaussian elimination is no more than a method for solving linear equations, the state
update vector equation given by Eq. 5.40 must be written as follows:

i}F
ax IxvLlXv = -F(Xv .,o
,u
Do) (5.45)

Since at iteration v both the Jacobian and the mismatch vector are numerically
evaluated, Eq. 5.45 is indeed a linear equation with ti.Xv as the unknown vector. The
idea behind Gaussian elimination is to convert the Jacobian or optimally ordered
Jacobian matrix into an upper triangular matrix by carrying out a series of operations
on Eq. 5.45. Once the equation has been triangularized, the solution ti.Xv is easily
obtained by back-substitution. To ensure that we are clear on these basic ideas,
consider solving the linear matrix equation

'

L - - - - - - - - - - - --------.----------
182 Power System Planning

The procedure used in Gaussian elimination to create the triangularized equation is as


follows:
I. Divide the first row by the self-element of row I, in this case 3.
2. Eliminate the 21 element by multipl)'.ing the modified first row by -1 and
adding it to the second row.
3. Divide the modified second row by its self-element (4 - ~), and stop.
Doing so, we get the upper triangular equation

2 I
I 3 X1 3

= o-!3
0 I X2 --2
4--3

Upon back-substituting-that is, solving for X 2 and then X 1 -we find that

X1 = 3I - 3
2X2
I I 6 _ 2
=3+15=15-5
To check our results, we note that

3(X 1 = ) + 2(X2 = - ;0 ) = I

I(X 1 = )+4(X2 = - 1~) = 0


Although we have used a simple example to illustrate the use of the basic
Gaussian elimination and back-substitution procedure, the same procedure applies to
solving the state update equations for the state update vector.
For computational efficiency it is common to perform the elimination by rows;
that is, the elements of a given row to the left of the diagonal are all eliminated, and
the row is divided by its self-element before proceeding to the next row. The
advantage of this procedure is that each row may be subjected to the elimination
process in a low-storage, compact working area.
We have referred several times to the necessity of optimally ordering the
Jacobian before the elimination procedure is begun. Optimal ordering of the rows
before elimination achieves two things:
I. It insures that division by a zero self-flement is unlikely.
2. It minimizes the number of new nonzero elements created in the elimination
process.
Although there are several levels of ordering, a simple but still highly effective method
involves ordering the rows so that the rows with the largest number of zero elements to

---------------------- -
Power-Flow Analysis 183

the right of the diagonal are numbered first. As we shall see in a detailed example, this
procedure does constitute a viable method for holding storage requirements to a
minimum.
We have reached a point where demonstrating concepts requires a representative
numerical example. What better example can we use than the interconnected system
discussed in Sec. 5.3.4. To establish a numerical example, we describe the two
three-bus areas as follows:

1. YBus matrix for each area according to Eq. 5.24

YBUS =
j20 jlO
jl() . -j20
jlOJ
jl0 (5.46)
[
jlO jlO -j20

where Y;; = -jlO p.u. for all ij including the interconnection


Y;o = 0 p.u. for all buses i = 1, ... , N
2. Y BUS matrix for interconnected system

-j20 jl0 jlO


jlO -j20 jlO
jlO jl0 -j30 jlO
YBus = (5.47a)
-j20 jl0 jl0
jlO -j20 jlO
jl0 jlO jl0 -j30

3. Initial control and disturbance vectors

IV,1 1.0 QDI 0.0


8, 0.0 PD1 0.0
IV2l 1.0 Qv2 0.0
a, 1.0 PD2 0.0
a2 1.0 QD3 0.0
l, 0.0
Do= PD3 =
4.0 (5.47b)
U= =
iV41 1.0 QD4 0.0
04 0.0 PD4 0.0
IVs I 1.0 QDs 0.0
PGs 2.0 PDs 0.0
Q3 1.0 QD6 0.0
Q4 1.0 PD6 4.0

j
L
184 Power System Planning

4. Initial state and mismatch vectors


Qc1 0.0 2.0
Pei 2.0 2.0
Qc2 0.0 -4.0
Pc2 2.0
82 0.0
IV31 1.0 !6 0.0
83 0.0 L1 0.0
x= = 0.0
Fo(Xo' U' Do)= =
2.0
(5.47c)
Qc4 !1
Pc4 2.0 fs 2.0
Qcs 0.0 f9 -4.0
85 0.0
IV61 1.0
86 0.0 0.0
l2 0.0 0.0
Even with this simple system the level of sparsity of the Jacobian is pronounced, as
evidenced by Eq. 5.49, where for the initial iteration the Jacobian is about 85% sparse.
In addition to sparsity, Eq. 5.49 clearly shows that with the present row (?Ider, inver-
sion of the Jacobian is not possible; the existence of several zero diagonal elements would
induce an overflow condition during the solution process. Further, the row order does not
satisfy the conditions stated earlier to render an optimally ordered Jacobian. However, the
reordering process is easily implemented and involves searching each row for the row with
a nonzero diagonal element as well as the largest number of zero elements to the right of
the diagonal. Applying this ordering criterion to the initial Jacobian of the sample system,
we order the rows as shown in Eq. 5.50. The associated ordered mismatch vector becomes
f4 =0.0
f1 - 2.0
fs = 0.0
f2 = 2.0
f3 = -4.0 *
f6 - 0.0 *
L1 = 0.0 *
F(X0 ,u 0 ,D 0 )=
f10 = 0.0
(5.48)
f1 - 2.0
fu = 0.0
fs = 2.0 *
/12 = 0.0 *
f9 = -4.0 *
L2 = 0.0 *
The significance of the asterisks will be explained later.
Power-Flow Analysis 185

As the initial Jacobian is quite sparse and easily triangularized, the reader will be
treated to a look at the steps required to complete the initial elimination and
back-substitution step. To facilitate the task before us, we should note that in Eq. 5.50
rows 1-4 and 8-10 are in effect in eliminated form since only zeros are to the left of a
unity diagonal element.* This being the case, we simplify the example by tabulating in
Table 5.1 only those rows and corresponding mismatches indicated by asterisks in Eq.
5.48 and 5.50 that require triangularizing. The operations and results of operations at
each elimination step are indicated in Table 5.1. For instance, the first step, indicated
by CD immediately below element 55, is to divide element 55 by itself, resulting in
unity for the modified element. The second step requires dividing element 57 by
element 55, and so forth. Continuing, the fifth and sixth steps require dividing row 6
by element 66, as indicated. If this procedure is repeated until element 14, 14 is
reduced to unity, the elimination is finished, and the Jacobian is in upper triangular
form ready for the back-substitution process to proceed. Reinserting the modified rows
in Table 5.1 into Eq. 5.50, we have the triangularized state update equation shown in
Eq. 5.51.

*Note that this simplification resulted from transferring Pc2, instead of Ii,, into the state
vector when constructing the model of the interconnected system (see page 164 ).

TABLE 5.1 Rows 5-7 and 11-14 of the optimally ordered Jacobian
Jacobian submatrix ~ -F_::,.

Column"? 5 6 7 8 9 10 11 12 13 14
5 110.0 0.0 l-30.0 0.0 0.0 0.0 0.0 0.0 I +10.0 0.0 +4.0
CD 1.0 Q)-3.0 (l)+l.O 0.40

6 0.0 -30.0 0.0 0.0 0.0 0.0 0.0 +10.0 0.0 0.0 0.0
G) 1.0 @-0.333
7 0.0 0.0 -10.0 0.0 o.o 0.0 0.0 0.0 10.0 o.o 0.0
(j) 1.0 @-1.0
11 0.0 0.0 0.0 o.o 0.0 0.0 -20.0 0.0 I 10.0 o.o -2.0
9 1.0 @ --0.5 (ip 0.1
12 0.0 +10.0 0.0 0.0 0.0 0.0 0.0 -30.0 0.0 0.0 0.0
0.0 3-26.66 0.0
1.00
13 0.0 o.o +10.0 0.0 0.0 0.0 10.0 0.0 -30.0 o.o 4.0
0 0.0
0 0.0
;-20.0
-15.0
1.0
8 3.0
--0.2

I
14 0.0 0.0 10.0 o.o 0.0 o.o 0.0 0.0 I -10.0 1.0 o.o
0.0 @ 0.0 1.0
t t
Row Mismatches

L
186 Power System Planning

Row order
2 1.0 10.0 10.0
4 1.0 -20.0 10.0
5 10.0 -30.0 10.0
1.0 10.0
3 1.0 10.0
6 -30.0 10.0
-10.0 10.0
aF = 7 (5.49)
ax 1.0 10.0 10.0 9
-20.0 10.0 11
10.0 10.0 -30.0 13
1.0 10.0 8
1.0 10.0 JO
10.0 -30.0 12
10.0 -10.0 1.0 14
Row order

I 1.0 10.0
2 1.0 10.0 10.0
3 1.0 10.0
4 1.0 -20.0 10.0
s 10.0 -30.0 10.0
6* -30.0 10.0
-10.0 10.0
aF = 7* (5.50)
ax 8 1.0 10.0
9 1.0 10.0 10.0
IO 1.0 10.0
11 * -20.0 10.0
12* 10.0 -30.0
13* 10.0 10.0 -30.0
14* 10.0 -10.0 1.0

1.0 10.0 0.0


1.0 10.0 10.0 -2.0
1.0 10.0 0.0
1.0 -20.0 10.0 -2.0
1.0 -3.0 +1.0 +0.4
1.0 -0.33 0.0
1.0 -1.0 0.0
1.0 10.0
t.Xo=
0.0
(5.51)
1.0 10.0 IQ.O -2.0
1.0 10.0 0.0
1.0 -0.5 0.1
1.0 0.0
1.0 -0.2
1.0 0.0

Back-substituting the results shown in Eq. 5.51, we get


b.14 0 = = 0.0
/:::,./2
/:::,.)(1/ 1:::,.86 = --0.20 (-11.45)
=
/:::,.X12 =!:::,.IV6 1=0.0
(5.52)
!:::,.%11 = 1:::,.8 5 = 0.5.6.x-13 + 0.1 = 0.0
/:::,.%100 = !:::,.Qas = -I0.6.x-120 = 0.0

/:::,.J(9 = M G4 = -I 0/:::,.)(11 - I O/:::,.X13 o - 2 .0 = 0 .0


Power-flow Analysis 187

M 8 = AQc 4 =-lOM12 = 0.0


M1 = Ao 3 = AX13 = --0.20 (-11.45)
M6 = AIV3I = +0.333M12 = 0.0
M 5 = Ao 2 = 3.0M/ -M13 + 0.4 = 0.0
(5.52 continued)
M4 =.,Af'c 2 =20M5 - lOM1 -2.0 =0.0
M/ = AQc 2 = -l0M/ = 0.0
M2 =Af'c 1 =-10M7 -lOMs-2.0=0.0
M1 =AQc 1 =-10~6 =0.0
Adding these changes to the initial state vector x0 , we obtain X 1 . Substituting
X 1 into the output vector equation, Eq. 5.27, and displaying the results, we obtain the
state and output quantities shown in Fig. 5.11, where the initial states and outputs are

l
Pc 1 = 2.0 (2.0] Pc 2 = 2.0 (2.0]
ac 1 = o.o
--...---'--....-V 1
10.01

= 1.0 LO (1.0 LO]


j Oc 2 = 0.0 (0.0]
V = 1.0LO [1.0LO]
-r--''--~ 2

Mismatch summary
f, = -0.0266 [-4.0]
L 1 =L 2 =0.0 (0.0]
f 10 = -0.0266 (-4.0]
Po 6 = 4.0
Po, =4.0
V 6 = 1.0L-11.45 [1.0LO]

Area #2
s = 1.98 +j0.199/,
[O+iO] / /

__._.,......_.__ V 5 = 1.0 L O [ 1.0 LO]


V 4 = 1.0 LO [ 1.0 L 0]

)
Pc 5 = 2.0 (2.0]
OGs = 0.0 (0.0]
PG 4 = 2.0 [2.0]
Oc 4 = 0.0 (0.0]
1
Fig. 5.11 Results of first iteration of Newton power-flow analysis.
188 Power System Planning

given in brackets for easy comparison and accentuation of the fact that we are moving
toward the solution. Selected nonzero mismatches are also shown. In just one iteration
the real power mismatches at the load buses have been reduced by more than 99%,
and at the .same time the scheduled interchange and specified generation have been
maintained, as they should be. The sample problem is almost too unwieldy to present,
but it demonstrates the role sparse matrix methods play in viable power-flow solution
techniques based on Newton's method. The final solution obtained if additional
iterations are performed is depicted in Fig. 5.12.

5.4.3 Accelerated Newton's Method


Almost all iterative solution techniques have an accelerated version, and Newton's
method is no exception. The basis of its acceleration scheme is that the Jacobian

j SG, 0
2.0 + /0.4174 15G, c 2.0 + io.4174

~-~~-V, =1.0LO --.-~--..--V2 = 1 .0 L 0

2.0 + j 0.4174 l
0. +i0.
! 2.0 +i0.4174

o +j0.0

- , - - , - - ~ - ~ - V3 = 0.9789 L-11.78

So 3 = 4.0

0.0+i0.0

I
~ - ~ ~ - V5 = 1.0L 0 __,__-,-___.__ V 4 = 1.0 L 0

1SG, 0
2.0 + i 0.4174 SG, , 2.0 + j 0.4174

Fig. 5.12 Solved base case.


Power-Flow Analysis 189

matrix may remain relatively constant over several iterations, making it unnecessary to
perform the elimination step. The state update vector equation in such case becomes

aF I f::,J(v = -F(Xv, U, Do) (5.53)


ax xv-p

where the Jacobian is the same for iteration v - p, p = 0, ... , v. For instance, if
the Jacobian matrix does not change significantly from its initial value, which is not
likely, p = v, and the state update vector is calculated each time using the same initial
Jacobian matrix. In the other extreme, if the Jacobian must be reevaluated at each
iteration, then p = 0. When p = I, the interpretation is that the Jacobian evaluated at
iteration v - l can be used to calculate both ~xv- 1
and ~xv, after which time it
must be reevaluated. Although p may be fixed at the beginning of the solution cycle,
on the assumption that the same Jacobian can be used in two, three, and more
consecutive iterations, it is more effective to monitor the Jacobian at each iteration
and change p accordingly.

5.4.4 Compact Storage Schemes


The viability of the Newton power-flow method depends to a great extent on
conserving storage; thus in creating the upper triangular Jacobian matrix through the
use of Gaussian elimination, the lower triangular elements of the Jacobian must be
eliminated a row at a time using the concept of a compact working row, and the
modified upper triangular elements and mismatches must be stored in a compact and
convenient manner. Further, only the nonzero elements of the Jacobian must be
operated on in the compact working row, and only the nonzero modified upper
triangular elements and mismatches should be stored in the compact upper triangle
storage area. Obviously, there are as many compact working row and upper triangle
storage schemes as there are programmers. Therefore we shall not discuss their details;
several of the references in this chapter describe in detail some workable schemes for
triangularizing and storing the Jacobian. In regard to back-substitution, if the
triangularized Jacobian is conveniently stored in the compact upper triangle table,
back-substitution progresses, quite naturally, backward through the table.

5.5 SENSITIVITY ANALYSIS


Frequently the system planner is faced with an unacceptable power-flow solution
requiring a change in some independent input parameter, or the planner is interested in
knowing to what extent changes in either the control or disturbance variables will
affect the state of the system. In either case, the information needed is readily
available without actually making the change and rerunning the power flow. More
important, not only can the sensitivity of the state of the system to changes in one
control or disturbance variable be determined, but the sensitivities of the state of the
system with respect to all control and disturbance variables can be obtained by solving
190 Power System Planning

the so-called linear perturbation power-flow equations, which is no more difficult than
solving the power-flow equations themselves.

5.5.1 The Sensitivity Equation


The linear perturbation equation for any interconnected or single area is obtained by
expanding the appropriate mismatch vector in a Taylor series about the base-case
solution point xs, U, D 0 , where xs is the solution state corresponding to the
control and disturbance vectors u0 and D 0 :
F(X 3 + AX, U 0 + AU, D 0 + AD) = 0
= F(Xs, uo ,Do)+ aF/ AX
ax xS u n
' '
+ aF AU+ aF AD+ (5.54)
au aD
xs,u , n
0
xS,u ,n
0

Dropping the nonlinear terms and solving for AX, we get

AX = -(aF)-1 aF AU -(aF)I aF AD (5.55)


ax au ax aD
= SxuAU + SxnAD (5.56)
Similarly, expanding Eq. 5.20, we get

AW= aHI AX+ aHI AU


ax xS,u au xS,uo
aH aH (5.57)
= ax (SxuAU + SxnAD) + au AU
or
AW = SwuAU + Sw0 AD (5.58)
To solve for the matrices Sxu and Sxn we must solve the following linear equations:
aF aF (5.59)
axSxu = -au
aF aF
ax 8 xn = -aD (5.60)

However, these equations are not exactly in the same form as the state update equation
since Sxu and Sxn are matrices and not vectors. Therefore the elimination procedure
discussed in Sec. 5 .4 cannot be directly applied'. To remedy this situation so that we may
use the same solution procedure used to obtain the state update vector, we need only
multiply Eq. 5.59 by AU and Eq. 5.60 by AD and then solve for the unknown vectors
Sxu AU and Sx 0 AD in terms of AU and AD and equate like coefficients to identify Sxu
and Sxn For use later, Eqs. 5.59 and 5.60 become:

----------------- --- - --------------


Power-Flow Analysis 191

aF aF
axSxuAU = - auau (5.61)
aF aF
axSxnAD = - aoao (5.62)

5.5.2 Evaluating aF/au, aF/aD, aH/oX, oH/au


Before we can perform a sensitivity analysis study, we must define aF/aU, aF/aD,
aH/aX, and aH/oU more precisely. Starting with aF/aU, we easily see that
aF 1 aF 1 aF1 aF1
au1 au2 auk auA
aL1 aL1 aL1 aL1
- --
au1 au 2 auk auA
aF (5.63)
au
aFA aFA aFA aFA
-- --
au 1 au2 auk auA
aLA aLA aLA aLA
--
au1 au2 auk auA
The partial derivatives of the mismatches of area i w.r.t. the control variables of area k
in terms of the area bus designations for the two areas are:

of1 of1 ar1 of1 of1 0[1 0[1 0[1


- - -- - - - - -- ------
a1vi1 081 oiV2I aw31 aPa3 olVal aPaa oa1
ar2 of2
----
olV1I 081
(5.64a)

of2Nj
.............................
azk
If k =A, then
afi ofi
must be added after
OPG 2 olV2i
and
a/i
must be removed.
oZA
192 Power System Planning

Continuing,

aLi -[ ~ aLj aLj aLj ... aLj]


auk - a1vi1 ao1 a1v21 a1v31 azk (5.64b)
ka"'A
If k = A, again
aLj aL;
aPc must be added after a!V I
2 2

and

aLi
azA must be removed.

The expressions for the partial derivatives in Eqs. 5.63 and 5.64 were developed
in Sec. 5.4, Eq. 5.43. We should note that if the interconnection power flows between
the areas are functions of only the state variables, then

and
aLi
-=O for all i and k
auk
This is an important case, for it results in a very simple, very sparse diagonal aF/aU
matrix.
Now consider the matrix aF/aD.

aF (5.65)
ao
aFA aFA
ao1 aoA
aLA aLA
aoA aoA

, ----------'
Power-Flow Analysis 193

Clearly this is a particularly simple matrix since

1. :i: =0 for all i =I=- k

oFi
2. oD . o f -I ,s an d O's
= a matnx
1

04
3. =0 for alli and k
oDk
If we want to determine Swx and Swu, then, as stated before, we need oH/oX and
oH/oU given below:

oH
ax - (5.66)

Since Hi is the vector of flows only in area i

oh1 ah1 oh1 oh1 ah1 ah1 ah1 oh1 oh1 ah1 ah1
-- - ... - - ---
oQc 1 0Pc1 oQc 2 oPc 2 002 aQcc aoc olVc+1I aoG+I a1vN1 OON
oh 2 ohz
oQGI aoN

oh2E
.................................... - -
aoN
(5.67)

Although the output matrix is not so simple as the previous two, it is straightforward
to evaluate. The elements are defined as follows:
194 Power System Planning

for all m and j


(5.68)
m=l, ... -,2Ek
j= l, ... ,Nk

if line m is not connected to bus j

ahm a 2
3. al Vil = olVil [Vj(V/- Vt)Yi1 + IVil yfoJ j = sending end
= (21Vjl -IVilei(oro;~Y/1 + 21Vi1Yfo
0
= (2Vy~+
} JI 2Vy'f')e-i
} JO i - V'f'!
I
0
iy~
JI
= 21 Vil(Yii + Yi*o) - IV;lyj;/<oro;)
or

ahm _
a1v;1 - a ( 2
alV;I IVjl Yji - ,vi
* IIV * i(o-o;)
;IYjie J + Ivi 12 YJO"'-) i = receiving end
-- -IV-ly~ei<oro;)
J JI
= -V-y~e-J
J JI
"Ii
i

2
4. aaho; = aaoi [(1vi1 - IVillV;lei(oro;>)yJi + IVjl 2Y/o] i = sending end
= -jlVillV;lyj;ei(oro;)
ahm ahm
- - - - -
ao; -
-
aoi
i = receiving end

As before, when k = A, aHAfaXA given must be modified by removing ahm/aPc2 ,


m = I, ... , 2Ek, and adding ahm/a!A.
Finally, aH/aU is given by

an1 aH1 aH1


au1 au2 auA

(5.69)

au1 auA

where
aH;
-auk = 0 for all i =I= k (5.70a)

if the interconnection between area i and area k does not involve a tap-changing
transformer and if the two buses p E i and q Ek are not voltage-controlled buses. Further
Power-Flow Analysis 195

aH;
auk ~ (5.70b)
i=kif=A

If i = A, the partial involving Pc 2 must be added, and the partial involving l A ,


removed. The partial derivative expressions for ah;tau; are givenin Eq. 5.68.
5.5.3 The Two-Area System
For the two-area system the matrices defined in Sec. 5.5.2 become:

aF1 joF1
0 0
au1 aD1 oH 1 aH1
0 0 0
aF 0 0 oF
-=
0 oH ax1 aH = au1
au oF 2 an oF2 ax 0H2 au 0H2
0 0 0 0
au2 0D2 ax2 au2
0 0 0 0
(5.71a)
The submatrices

a/1 a/1 a/1 a/1 a/1 a/1


--
a1vi1 a.s1 a1V2I aa1 aa2 li
0/2 a/2
of 1 awi1 a11
- = (5.71b)
au1

-a/6- a16
olVil a11

-af1- af1 -at1


- at1 0/7 af1
a1v41 084 alVsl aPcs aa1 aa2
aF2
and - = (5.71c)
au2
a/12 a/12
a1v41 aa2
t
196 Power System Planning
$

0.0 -1.0 0.0 0.0 0.0 0.0


0.0 0.0 0.0 -1.0 0.0 0.0
aF 1 0.0 0.0 0.0 0.0 0.0 -1.0
-= (5.71d)
aD 1 -1.0 0.0 0.0 0.0 0.0 0.0
0.0 0.0 -1.0 0.0 0.0 o:o
0.0 0.0 0.0 0.0 -1.0 0.0
and
aF2 aF
- - -1 (5.7le)
aD2 aD1
Consider the problem of determining the effect that load changes have on the
state of the system. For this case Eq. 5.56 becomes
AX= Sx 0 AD
since AU= 0. The task before us is to solve the linear equation
aFj
ax aF
SxoAD = - an AD
xS,uo ,Do

for SxnAD, given the Jacobian evaluated at the solution state as well as aF/oD and
AD. From Fig. 5.12 we get the solution state

Qci = 0.4174
Pai = 2.0
Qc2 = 0.4174
Pa2 = 2.0
02 - 0.0
IV31 = 0.978
03 = -11.78
xs=
Qc4 = 0.4174
Pa4 = 2.0
Qcs = 0.4174
05 = 0.0
IV61 = 0.978
86 = -11.78
12 - 0.0

Therefore the Jacobian aF/oX is shown in Eq.' 5.72. Upon optimally ordering and
triangularizing, we obtain the upper triangular equation depicted in Eq. 5.73.
Back-substituting, the last column of the sensitivity matrix sought is defined by Eq. 5 .74.
With this result we can evaluate the effect of any change in load on the state of
the system. For instance, assume that the real load on bus 6 (or bus 3 in area 2) is in error
by an amount MD 6 That error will produce according to Eq. 5.74 errors in the system
states as follows:

'
j
..&.L
Power-Flow Analysis 197

AQ01 = 0.l 68APn6 AQc4 = 0.15Mn6


' Mc 1 = 2.0APn6 Mc4 = APn6
AQc2 = -0.4APn6 AQc 5 = 0.09APn 6
Mc 2 = -2.0APn 6 A85 = -0.03APn 6
A82 = -;-0.l 4APn6 AIV61 = -0.008APn6
AIV3I = -0.003APn6 A8 6 = -0.071 APn6
= -0.071 APn6 =
These results, although rough approximations, are feasible since we know that

1. Mc1 +Mc2 =O
2. Mc4= Mn6
3. Al2 = 0
as the data above verify. Relatively speaking, changes in the real demand tend to
change real power generation and bus voltage angles more than reactive generation and
bus voltage magnitudes.
1.00 10.00 - 2.04 9.58
1.00 -19.58 - 2.04 9.58
9.58 4.08 -28.74 9.58
1.00 9.78 1.99
1.00 - 4.04 9.78 1.99
- I.99 -29.36 3.99 9.58
9.58
aFI
ax x =
- 9.58
1.00 10.00 -2.04 9.58
-19.58 - 2.04 9.58
9.58 9.58 4.08 -28.74
I.00 9.78 1.99
1.00 - 4.04 9.78 1.99
9.58 - 1.99 -29.36 3.99
9.58 - 9.58 1.00
(5.72)
1.0 9.78 1.99 t.QD!
1.0 10.0 -2.04 9.58 Mm
1.0 - 4.04 9.78 1.99 6Qn2
1.0 -19.58 -2.04 9.58 Af'n2
1.0 0.42 -2.99 1.0 e.r;)J
1.0 -0.34 -0.33 -0.Q70 e.Q;)J
1.0 -1.0 0
1.0 9. 78 1.990
Sxn6D =
6Qn4
1.0 10.0 -2.04 9.580 Af'n4
1.0 - 4.04 9.78 1.990 6Qns
1.0 0.10 -0.490 Af'vs
1.0 0.266 e.Ql,6
1.0 Af't>6
1.0 0
aF '-'Ao af ter tnangu
ao . ti anza!Ion
. .

* Indicates elements modified during triangularization (5.73)

.....
1
198 Power System Planning

0.168 t.Qni
2.00 Mm
-0.40 t.QD2
-2.00 Mv2
-0.140 t.Qn3
-0.003 Mv3
Sxvt.D'=' -0.071 t.Qn4
(5.74)
0.15 Mv4
1.0 t.Qns
0.09 Mvs
-0.034 t.QD6
-0.008 Mv6
-0.071

5.6 OPTIMAL POWER FLOWS


In Sec. 5.5 we discussed in great detail the Newton power-flow solution method. We
noted that in specifying the power-flow problem, the independent control vector U
had to be given. Since the real power generations PGi, i = 2, ... , G, are control
variables and appear in U, they must be specified before a power-flow solution can be
obtained. Generally the specified values of Pc;, i = 2, ... , G, are arbitrary but
reasonable values defined by the planner. Unfortunately, these specified generations
rarely minimize production cost as determined by an economic load-dispatching
program used routinely in dispatch centers. Hence, the system planner often obtains
power-flow results for a generation schedule that would not be adhered to if the
expansion plan were implemented. It is not completely unreasonable to conclude,
therefore, that operating problems can arise from planning a system using generation
schedules different from those actually implemented. Although the industry has not
found, or at least recorded, any cases in which a poor design was implemented as a
result of using suboptimal generation schedules, it is important to consider an
alternative solution procedure that can be used to circumvent the problem if necessary.
In Sec. 5.6 the broad topic of optimal power flows is presented in enough detail
to familiarize the reaJer with the basic concepts.

5.6.1 Basic Problem Formulation


As the name implies, optimal power-flow methods are essentially static optimization
procedures in which the optimal generation schedule that satisfies the power-flow
equations and minimizes production cost C(X, U) is sought. Using the state space
notation developed earlier, we may state the prqblem for a system of A interconnected
areas as follows:

min C(X,U) = (5.75)


w.r.t.U
Power-Flow Analysis 199

subject to the constraints that

F(X, U,D 0 ) =0
x,<x,<x
u,<u,<u
This formulation would be appropriate for areas operating in a pool arrangement
whereby generation schedules are determined to minimize the production cost for the
entire system. If an area were not part of an overall system, its optimal generation
schedule would be determined by minimizing Ck(Xk, Uk) subject to the constraint
that F k(Xk, Uk, Dk 0 ) = 0, etc. Since the two problems are mathematically identical,
we shall not restrict ourselves by discussing a single area; rather we shall orient the
discussion to be valid for a system of A interconnected areas.
From Eq. 5.75 we note that basically three constraints must be satisfied by X
and U. The fir~t is the equality constraint that disallows any values of X and U that do
not satisfy the power-flow equations. The other constraints are inequality constraints
on X and U: that is, only those values of X and U within the ranges defined by Eq.
5. 75 and permissible. A constrained minimization problem like ours is solved by
transforming it into an unconstrained minimization problem. There are many
procedures for transforming the problem; perhaps the most elegant is the Kuhn-Tucker
method. Unfortunately its elegance is tarnished by extremely difficult computational
problems that arise in applying the method. In comparison, steepest descent, although
hardly elegant, is amenable to implementation on a digital computer. The source of the
computational problem in both cases is in the existence of inequality constraints on X:
It is usually difficult to choose successive values of U that produce smaller and smaller
values of cost C without violating the inequality constraints on X, and hence it is
difficult to foresee what effect a given change in U will have on X.
We elect to use the Kuhn-Tucker method to solve the optimal power-flow
problem, leaving steepest descent for later. In this way both solution methods will be
presented in an efficient manner.
The Kuhn-Tucker method involves adjoining the constraints to the cost function
with dual variables as follows:

C* = C + >..'F(X, U,D0 ) + '(U - U) + i(~ - U) + ff(X - X) + [Oi - X) (5.76)

where C* is now the unconstrained, augmented, cost function, and X, , and {3 are the
dual variables. The necessary conditions that the optimal value of U, U, must satisfy
are:

I ac* = = ac + .. , aF + ii'
ax 0 ax " ax "
2 ac* = = ac + .. , aF + -,
au 0 au "au (or-J!_) (5.77)

ac* 0
3. a>..' = 0 = F(X, U,D )
200 Power System Planning

4. '(U - U) = 0 or i N - U) =0
"if' (X - X) = 0 or fi. ~ - X) =0
(5.77 continued)
5. .' or i ~ 0
p' or fr~ 0
The first three necessary conditions are standard; that is, the optimal value of U
and associated X will be values that, when perturbed by amount AU and AX, will not
change the cost C* but will satisfy the power-flow equations. Condition 4 in Eq. 5.77
forces the solution X and O either to satisfy their constraints, in which case the
corresponding dual variables are zero, or to be at least on the boundary of the
constraint surface, in which case (U - U) or (!} - U) and (X - X) or (~ = X) are zero.
The last condition insures that no smaller value of cost can be obtained by changing X
and U in a direction away from their limiting values. This interpretation of the
dual-variable condition is easily demonstrated if we realize that at the optimum
dC* = 0 = dC + )..'dF + ji'd(U - U) + idN - U) + P'd(X - X) + {td~ - X) (5.78)
Thus
, ac - AC (5.79a)
= - a(u - U) = A(U - U)
or AC= -.'A(U - U)
Since Dis a constant, AU= 0, and
AC= -;i'AU (5.79b)
If U is decreased, then AU< 0, and if p.' > 0, the change in cost AC> 0 indicates an
increase in cost. In addition to explaining why condition 5 must be satisfied at the
optimum, we have uncovered the physical meaning of the dual variables: they define
the sensitivity of the cost to changes in the control vector limiting values. That is,
since the control vector U cannot exceed D, the only way U can be increased above its
optimal value is to relax D by allowing it to assume a larger value. The change in cost
resulting from relaxing tJ is given by Eq. 5.79b.
Solving Eq. 5. 77 for the optimal control vector is hardly straightforward, since it
is not usually known which inequality constraints must be included in the necessary
condition equations. Many possibilities exist because for the optimal 0, some control
and state variables may be forced to their upper limiting values while others are forced
to their lower values. Depending on which constraints are active, the necessary
condition equations change, and since the optimal U is needed to determine which
constraints are active, the problem becomes difficult to handle computationally. To
circumvent these problems, the following solution procedure can be used.
1. Choose a feasible control vector that satisfies all the constraints but is not
necessarily optimal.
2. Temporarily let D = !! = U, which forces U to be the optimal U as it is the
only U that satisfies all the constraints. For this case Eq. 5.76 becomes
Power-Flow Analysis 201

C* = C + }l.'F(X, U,D0 ) + (ii' - ,i)(U -1!)


I _, I
Lett mg= -I!
C* = C + >..'F(X, U,D0 )+ '(U -1!)
dC* = 0 = dC + >,.'dF + 'd(U - 1!)
Thus, ' = -ac/aU, which is the gradient of the cost C w.r.t U needed to
define the control vector update algorithm.
3. Using this last result, update uv as follows:
uv+l = uv - ex(-') = uv + ex' (5.80)
where' is calculated using Eq. 5.77. From Eq. 5.80
auv = uv+l - uv = ex(') (5.81)
This change in control must not produce a change in X that will result in a
violation of the state inequality constraints; that is,
~ ,;;;; xv + axv,;;;; x
where

(5.82)

With the step size ex adjustable, the change LlXv can be forced to satisfy the
condition above. If no change in U is possible without violating the state
inequality constraints, then ex= 0, the optimum value of U = uv, and the
process stops.
4. Update uv to uv+ 1, and return to step 1 if 1!,;;;; uv+ 1 ,;;;; 0. Otherwise stop
and calculate X corresponding to U.
Before proceeding further, an example is in order to clarify the material
discussed. Once again consider the three-bus system depicted in Fig. 5.13, which is
area 2 of the multiarea system considered in Sec. 5 .4. We want to
min C = ko 4 + 0.5PGi + kos + Pc/
w.r.t.PGs

subject to the constraints


F2(X2, U2,D2 ) =0
1.0 IV4 1 1.0
0.0 84 0.0
1.0 IVsl - 1.0
1!2 = ,;;;; U2 = ,;;;; U2 =
0.0 PGs PGs
1.0 1.0
1.0 1.0
202 Power System Planning

Po. + iOo. = 4.0 + i o.o

-.----._,._V6 =0.978L-11.78

V 4 = 1.0 L 0.0
~ ~ ~ - V5 = 1.0L 0.0

PG, +iDG, ~2.0+j0.4171

Fig. 5.13 Area 2 of the two interconnected areas.

-2.0 Qa4 2.0


0.0 Pa4 3.0
-2.0 Qas 2.0
X2 = ~X2 = ~X2 =
-0.35 05 0.35
0.95 1v61 1.05
-0.35 06 0.35
If we choose a feasible U2 to be the same as in Sec. 5.4 for area 2, then

1.0 0.417
0.0 2.0
1.0 0.417
uo- 2 - and X2 =
2.0 0.0
1.0 0.9789
1.0 -0.205
'
Clearly U 2 is feasible in that it results in no constraint violations. However, U 2 above
is not the optimal value which minimizes C2 Thus the next step is to update U2 as
in Eq. 5.80, where' must be determined from Eq. 5.77:
' = _ ac _ >..' aF
au au
Power-Flow Analysis 203

where from Eq. 5.77


1
)._' = _ ac (aF)-
ax ax

Thus
1
,= _ ac + ac (aF)- aF
,,_ au ax ax au
ac ac
=-au+ axSxu

where

0
:~ = (0 0 0 2Pcs 0 0)
= (0 0 0 4 0 0)

:~ = (0 Pe4 0 0 0 0)
= (0 2 0 0 0 0)
and Sxu is determined as discussed in Sec. 5.5 and defined by Eq. 5.59. Note that
because all the elements of 3C/aX are zero except the second, we need for this
example only the second row of Sxu, which can be shown to be
SpG4, u = (0 10 0 -1 0 0)
Therefore

' = -(0 0 0 4 0 0) + 2(0 10 0 -1 0 0)


= (0 20 0 -6 0 0)

With' substituted into Eq. 5.81,

0
-20
0
or
+6 Mes = o:6
0
0

Since Q4 = 84 = 0, 8 4 should not be updated. However


Pc/ =Pes0 -o:6
To obtain the optimal value of o:, we also need Pc 4 1 , since C is a function ofbothPe 4
and Pcs From Sxu we can show that
204 Power System Planning

Mes= SpcsuAU
= 10Ao4 - AI'c 4
but Ao 4 = 0. Thus
AI'c4 =-Mes= a6
or
Pc41 =Pc40 +a6
Upon substituting Pc/ and Pc/ into C and minimizing w.r.t. a, we find that the
optimal step size is O'. = . To insure that O'. = does not result in state violations, we
need only determine the changes in state using Eq. 5.82. If AX is too large, a can be
appropriately reduced to avoid violating the state constraints. In fact, this subproblem
is normally formulated as a linear programming problem, since we are seeking the value
of AU that maximizes AC subject to the linear constraints that AX :s;;; X - X or
AX>~ - X. We shall not pursue this subject here; suffice it to say that this extension
is clearly logical. Updating the generation as described results in the optimal power-flow
solution.

5.7 REACTIVE POWER-FLOW ANALYSIS


Historically, little attention has been given to the study of reactive power and its effect
on overall system operation. This basic lack of concern may be attributed to the way
in which electric utilities have billed their customers for energy. Only in a very few
instances does the electric utility industry, even today, charge for MVARHs; rather,
operating revenue is derived primarily from selling MWhs. The logic associated with this
procedure is well founded, since the source of energy is fuel and associated with each
unit of fuel is a well-defined cost. Extending this argument to MVARHs, we find that
the source of MVARHs is, in some sense, the excitation system associated with the
generator. The cost of producing MVARHs I is not so simple to define as the cost of
producing energy, since an excitation system is needed even for unity power factor
operation, and once installed is available at no additional cost to continue producing
MVARHs. What is easily defined is the differential cost associated with generating units
with larger reactive power capability. The lower the power factor of a generator, the
more costly the unit, because of its inherently larger physical size. -
Other reasons can be cited for the cursory evaluation reactive power has received
in the past, the most notable being that in low-voltage networks of yesterday, the
voltage profile problems were easily corrected by using shunt capacitors. With the
introduction of higher-voltage elements that produce very large amounts of reactive
power, a host of new control problems were created by the attempt to maintain the
'
reactive power balance needed to insure an acceptable voltage profile. In the in;tial
years of operation, before high-voltage elements are heavily loaded, and during off-peak
load periods in general, these elements generate very large amounts of reactive power
that, if not absorbed by other system elements, will produce excessively high voltages.
To compound the problem, the larger generating units being put in service today have

---------- - - ------- --------------


Power-Flow Analysis 205

smaller reactive power-handling capability; meaning that less capability is built into the
system to absorb the reactive power produced by the system itself.
We shall focus our attention on the voltage problems created in EHV systems by
using the results of our sample system. We shall investigate the effect of various voltage
control devices in maintaining a desired voltage profile, again using the sample system.
We conclude with a discussion of how to formulate the voltage control problem as a static
optimization problem.

5. 7.1 Reactive Power and Voltage Considerations


Before discussing voltage control techniques at length, we present a procedure for
determining the relative effects of the following on system voltages and reactive
generation.
1. Changing load
2. Llne charging
We must also consider how to determine the degree to which the following control
variables can influence the voltage profile.
1. TCUL transformer taps
2. Generator bus voltages
3. Reactor compensation
In Sec. 5.5 we developed the philosophy of sensitivity analysis, which can now
be used to obtain the information we desire. The sensitivity matrix Sxo, for instance,
provides the information we need to determine how load levels and line charging affect
voltages; for example,
LlX = Sx0 LlD (5.83)
For the sample system we obtain from Eq. 5.74 that
L\IV3I = -0.04L\Qv3 - 0.0lM'v3
If L\Qv 3 is unchanged from the base-case value, then
(5.84)
This result indicates that as the load level decreases, the voltage tends to rise, a
conclusion not too surprising. However, if we let
(5.85)
then as IV3 I changes from its base-case value, given in Fig. 5 .12, reactive power will be
generated and injected into the system, thus simulating the effect of line charging.
From Eq. 5.85 we get
L\Qv3 = -2WCIV3IL\IV3
but

206 Power System Planning

Therefore
-0.01Af>D3
3
l!.IV I = I - 0.08WCIV3I

Assume WC= 3.0 p.u., which is not unrealistic for some EHV lines. Then

(5.86)

Note in this case that the sensitivity of the load bus voltage is 30% higher with line
charging than without, so that the voltage will rise to higher levels for the same
decrease in demand. This result is important; it demonstrates why EHV lines usually
require compensation in order to hold voltage~ down during off-peak periods. Also
note that without line charging we see from Eq. 5.74 that

AQGJ = 0.4f!.QD3 + 0.08Af'D3


l!.Qc2 = 0.44!!.Qo3 + 0.5M'D3
and if AQD 3 = 0, that

flQc1 = 0.08Af'D3 (5.87)


AQc2 = 0.5Af'D3
These equations indicate that as the system load decreases, the generating units become
less excited; that is, they produce less reactive power in order to maintain base
voltages. If we include line charging as before, we get

l!.Qc 1 = 0.4(-2WCI V3 IA IV31) + 0.08APD3


LlQc 2 = 0.44(-2WCIV3!LllV31) + 0.5Af>D3
where from Eq. 5.86

AIV3l = -0.013Af'D3
Thus

AQGJ = 0.085Af'D3
(5.88)
AQc2 = 0.505Af'D3
The conclusion is obvious: Line charging tends to force generating units to operate at
even lower excitation levels, and in extreme cases can result in a given generator
operating on its minimum excitation limit.
With this modest background we can e.ntertain the second question: to determine
how sensitive system voltages may be to changes in various control elements. Again we
rely heavily on Sec. 5.5. From Eq. 5.56
dX = SxuAU + SxoAD (5.89)

~-------- - - -
t
Power-Flow Analysis 207

It can be shown that for the sample system

A IV3l =- ~1
llz
4
L2
IV312&z2 - 0.04AQn3 - 0.01LV'n 3 (5.90)

Substituting the base-case values for the parameters and variables given in Sec. 5.5, we
have '

AIV3l = -0.383&z2 - 0.04AQn 3 - 0.0lLV'n3


If no line charging exists, then
AIV3I = -0.383&z2 - 0.01LV'n 3 (5.91)
However, with line charging

A IV31 = -0.383&z2 + 0.04(2.0)(3)(0.9789).t.l V3 I - 0.01 b.Pn 3


= -0.50L\a 2 - 0.013LV'n3 (5.92)
Upon comparing Eqs. 5.91 and 5.92, we note that
1. TCUL transformers are effective voltage control devices. Changes in tap over a
0.10-p.u. range change the voltage over a 0.0383-p.u. range when no line
charging exists and over a 0.05-p.u. range when line charging is included.
2. Line charging increases the sensitivity of system voltage to changes in TCUL
tap positions.
Equation 5.89 also contains information needed to evaluate the effectiveness of
generator bus voltages as voltage control variables. Again, from Eq. 5.89 it can be
shown that
(5.93)
We note that generator bus voltages are effective control variables, as we would expect,
where the degree of effectiveness of TCUL transformers and generator bus voltages is of
the same order of magnitude.
This section would not be complete if we did not investigate the impact of
adding shunt reactors to the system. As indicated earlier, shunt reactors are used to
compensate EHV systems with excessively large amounts of line-charging reactive
power. We have seen how line charging results in higher voltages as loads drop off;
therefore we should demonstrate how reactors can reduce the sensitivity of load
voltages to decreasing loads. To do so, we must only modify Eq. 5.85 as follows. Let

Qv3 = (JL - wc)[lv31 2 - (0.9789)2] (5.94)

Thus
208 Power System Planning

and upon substituting into Eq. 5.93


0.363Ll IVil - 0.383lla 2 - 0.0IMD3
(5.95)
LllV3 I = I+ 0.0783(1/WL - WC)
Consider two cases: 1/WL = WC= 3.0 p.u,. and 1/WL = 0. In the first case, with
reactor compensation assumed,

The second case, without reactor compensation, produces

Adding reactor compensation


I. Decreases the sensitivity of the load bus voltage to changes in demand, which
means that as the load deer Jases, the voltage tends not to rise so high.
2. Decreases the sensitivity of the load bus voltage to changes in generator bus
voltages and TCUL tap settings, which means that to effect a given change in
load bus voltage, the TCUL tap and generator bus voltage must be changed by
a larger amount than is necessary without compensation.
To conclude, the three-bus system depicted in Fig. 5.14 was used to illustrate
again the effect of varying loads on the load bus voltage and reactive generation. To
accentuate the performance of EHV systems, a capacitive element was added to the

j0.04

j0.4 I Po,
a0 ,
= 10 + 6 sin
= o.5Po,
(45 Tl

Fig. 5.14 Three-bus system with sinusoidally varying load.


Power-Flow Analysis 209

1.0

iV,1

Peak Off-peak

0 2 3 4 5 6
T-

Otoss 1.0

2 3 4 5 6

Fig. 5.15 Effects of varying load on bus three.

load bus to simulate line charging. Figure 5 .15, a-c, depicts the variation in load bus
voltage, reactive losses, and reactive generation due to variations in load. We should
note that peak loads tend to

1. Create low voltages


2. Produce large reactive power losses
3. Force generators to operate highly overexcited

Off-peak loads in comparison are characterized by

1. High voltages
2. Negative reactive power losses
3. Low generator excitation levels

We have placed the voltage and reactive power problem in perspective, we hope,
and illustrated why compensation techniques are needed in planning the expansion of an
EHV system. Next we shall discuss a method for determining the range over which
control variables must vary for effective voltage control.
210 Power System Planning

5.7.2 Voltage Control


Based on discussions in Sec. 5.7 .1, we present here a method for determining the
TCUL tap positions, generator bus voltages, and reactive compensation needed to
maintain a desired voltage profile for a given load. By varying the load over the
anticipated load cycle and calculating the associated control variable values needed to
maintain a desired voltage profile, the planner can determine if adequate control
capability exists in the system. If the voltage control capability is inadequate, steps can
be taken to modify the expansion plan by including an appropriate amount of
compensation.
The approach proposed is a static optimization procedure in which the index of
performance, J, is given by
Ns
J = . ) 0
W;(IV;I - IV; 1)
2

1=~+1
(5.96)
= ex - X )"W(X -
0
X)
where W is an appropriate weighting matrix and Gs is the total number of system
generators. X is of course the system state vector as defined earlier, and x0 is the
desired value of the voltage states. The problem before us is to minimize Eq. 5.96
w.r.t. the control vector, subject to the constraints that
u,.;;u,.;;u
X ,.;; X ,.;; X (5.97)
Fex, U, D = 0
0
)

For our purposes the upper and lower limiting values for all control variables except
generator bus voltages and TCUL tap settings will be equal and equal to the base-case
values, in effect restricting the variables to be used in the optimization procedure to
TCUL taps and generator bus voltages. The state inequality constraint will disallow
control variable values that result in either extreme overexcitation or extreme
underexcitation of system generators; that is, it forces the optimization process to seek
a solution that falls within the reactive power capability of the generation system. If
we want to include reactive power compensation as a control variable, we need only
include it in U as a generator with zero real power output.
There are several ways to insure that the inequality constraints are satisfied. The
one we have chosen involves adding a penalty function G(X) to J to force the solution
away from the boundaries ~ and X. This addition changes Eq. 5.96 to
J = ex - X)'W(X-X) + cex) (5.98)
The inequality constraints on U are handled by not allowing U to take on a value
outside the range given by Eq. 5.97. The equality constraints, in this problem the
power-flow equations, are handled using the standard Lagrange multiplier method; that
is, equality-constrained optimization problems can be transformed into unconstrained
Power-Flow Analysis 211

problems by adjoining the equality constraint to the performance index with a


Lagrange multiplier }... Thus our problem reduces to

min J* = ex - X)'W(X -:x) + X'Fex, U,D0 ) + cex) (5.99)


w.r.t.U

The necessary conditions that must be satisfied by the optimal value of U are
that the first partial derivatives of the augmented performance index given in Eq. 5.99
must be zero:

aJ* = 0 = ~ + ac + A,aF
ax ax ax ax
31* = O = }..,aF (5.100)
au au
aJ*
a}.. = 0 = F(X,U,D
0
)

Solving for }._' in the first equation and substituting in the second, we get
1
a1* = -(aJ + aG)(aF5 aF (5.101)
au ax ax ax au
which is the gradient of the augmented performance index w.r.t. the control vector.
This information tells us how to change U to reduce J*; each successive value of U
should be chosen in the direction of steepest descent of the performance index. Using
Eq. 5.101, we can define the control vector update algorithm to determine successively
better values of U:

ui = u - o\(a1*"\
au / (5.102)

a is again the step-size parameter usually determined through an empirical procedure or


a rigorous one-dimensional optimization method, such as a Fibonacci search.
It is interesting that if we have a Newton power-flow and a sensitivity analysis
program, Eq. 5.101 is easily evaluated since

(5.103)

can be determined as described in Sec. 5.5. The remaining quantities can be evaluated
because both J and G are scalars, where

~=ex-
ax xo)'w (5.104)

and aG = (X- X)'B if X >X


ax
= (~-X)'B if X< ~
212 Power System Planning

since we define G(X) as follows


1
G(X) = (X - X) B(X - X) X>X (5.105)
or

= (~ -X)'B(~ - X) X<X

The steps in the solution process are:

I. Solve power flow.


2. If Eq. 5.97 is satisfied and X = X 0 , stop; otherwise proceed to step 3.
3. Construct and evaluate aG/aX, depending on which states violate their
constraints and on the nature of the violation. Note that load bus voltages are
not considered in G(X), since they are in J.
4. Evaluate Sxu, given in Eq. 5.103.
5. Evaluate aJ/aX, given in Eq. 5.104.
6. Update U using Eq. 5.102 but constraining U to the range given in Eq. 5.97,
and return to step I.

To illustrate how to determine the optimal control vector that produces the
voltage profile desired, we shall return to area 1 used in Secs. 5.4 and 5.5, since we
now have a great deal of information about the system. We shall assume that the
problem is to determine the value of a 2 that will result in a load bus voltage

I V3 I = 1.0. Generally we would want to include the other control variables in the
optimization procedure, but for simplicity we will restrict U to allow only a2 to vary.
Thus the problem is to

min (I V3 I - 1.0)2 = CX1 - X1 )'W(X 1 - X1 )


w.r.t. U 1

subject to the constraints that

~I X1 X1
'Q1 < U1 < U1
F(X1,U1,D1)=0
where

0
0 Q
0
W=
0
0 I
0
f
Power-Flow Analysis 213

Qa1 1.0 -1.0


Pai 2.0 2.0
Qa2 1.0 -1.0
X= = ~=
82 0.5 -0.5
lt'JI 1.0 1.0
83 0.5 -0.5
!Vil 1.0 1.0
81 0.0 0.0
IV2I 1.0 1.0
U= - !I=
P02 2.0 2.0
ii 1 1.0 1.0
ii2 1.15 0.85

From Eq. 5.101

aJ*
aa = -[0 0 0 0 (IV31- 1.0) (aF)- aF
1

0] ax aa
2 2
where from Eq. 5.59
'

aF)- aF
1

- (ax aa = SXa2 =
2
a1v31
aa2
a83
aa2
Since the multiplication of aJ/3X times Sxa 2 involves only the next-to-last row of
Sxa 2 , or a1v31/aa2, we can use the results obtained for AIV31 in Sec. 5.7.1, Eq.
5.92:

or
AIV31 a1v31
--=--=-05
&2 - aa2 .
Substituting this last result into the above expression for aJ*/aa2, we get
aJ*
-a-= -(IV3I - 1.0)0.50
a2

-------- --- ---- --- - - ---------


214 Power System Planning

But from Fig. 5.12, IV 3 I= 0.9789, thus

a/ = az + a(IV3 I - 1.0)0.5
= 1.0 - a'.(0.0105)

Letting a = 1.0, we get


a/ = 1 - 0.0105 = 0.989

Decreasing a 2 to 0.989 is certainly in the proper direction, and if the process were
continued, we would find the optimal value of a2 as well as the corresponding power
flows.

5.8 STOCHASTIC POWER FLOWS

We have discussed, in Secs. 5 .I to 5 .7, concepts that are applicable when the forecast
loads are assumed completely known. Obviously, as we learned in Chap. 2, there is
uncertainty associated with a demand forecast, and the degree of uncertainty is
quantified by the forecast variance. In Chaps. 3 and 4 we were concerned enough
about the forecast uncertainty and its effect on generation reliability and production
costs to include it by modifying the load probability distribution. Again, we must
recognize that the uncertainty in the forecast demands will result in uncertainty in
power flows as well as in voltage levels. In Sec. 5.8 it is our goal to consider a method
for including uncertainty in power-flow analysis in order to gain some insight into the
effects of system loads other than those normally chosen. Whether a given utility
should consider adding stochastic power-flow procedures to its arsenal of analytical
tools is a question not easily answered. However, if historical records indicate rather
large discrepancies between forecast and actual demands, a good case for stochastic
power flows may be made.

5.8.1 Problem Formulation


The method proposed here is almost identical to the method used in Chap. 2 for
demand forecasting. We know from Chap. 2 that the best estimate of the coefficient
vector a, given random historical data l(t), where
l(t) = a'f(t) + 77(t) (5.106)
or
L(N) = a'(N)R(N) + r,(N) (5.107)
is given by Eq. 2.1 I. The covariance of a is' given by Eq. 2.22. Since the stochastic
power-flow problem can be formulated in the same manner, the results obtained in
Chap. 2 can be applied here. We must obtain a linear model for the power system by
using either the DC power-flow equations or the Newton iterative equation, in which
the power-flow equations are expanded about an operating point X8 , U, D0 . The

- - - ~ - - - - - - - - - - - - - --- -- ---------~ -- -------- -- ---- --- -


Power-Flow Analysis 215

disadvantages of using the DC power-flow equations are obvious, so the Newton


iterative equation will be used to model the system about a point. Thus

AF = aF
aX
IxS,u,o AX + aDaF IxS,u,o AD = aF AX+ 1l
aX (5.108)

where 11 is a random noise vector whose covariance defines the variation of the forecast
demands about their base-case mean values D0 . X 8 is the corresponding base-case state
vector. Transposing Eq. 5.108, we get

AF' = AX' oF + 11 (5.109)


ax
Note the similarity between Eqs. 5.109 and 5.107. In stochastic power-flow problems
we are seeking the best estimate and covariance of AX, given noisy forecast data AF';
the noise vector 71' is now assumed to have the following characteristics:
1. (71) = 0. -
2. 11 is Gaussian.
3. [(71-77)] 2 = Cov (71) = i
In Chap. 2 we assumed that the noise covariance matrix C11 was diagonal, meaning that
the historical demands being processed were completely independent; that is, the
weekly peak demand levels were independent. In stochastic power flows it is not
reasonable to assume that the elements of F(X, U, D) are independent or that their
changes, AF', are independent. We know that all injections are correlated. The effect
of this correlation, or dependence, between the elements of AF' is seen in the noise
covariance matrix C11 ; it is not a diagonal matrix, since the off-diagonal elements
quantify the degree of correlation between elements of AF'.
If in Chap. 2 we had used as the weighting matrices W(T)W'(T) the matrix C11 - l ,
then the best estimate a' obtained would have been the so-called minimum variance
estimate instead of the least squares estimate. In stochastic power flows it is common
to replace W(T)W'(T) by c11 - 1 so that the correlation between elements of AF' can be
taken into account. On making this minor change, the best estimate AX' is given by
Eq. 2.11, with the variables changed to those given in Eq. 5.109. Thus
1
AX' = aF'c11_ aF /aF' c11 aF)-
1 -1
ax \ax ax
and
1
AX = (aF' c11_1 aF)- aF' c -1 AF (5.110)
ax ax ax 11
Equation 5.110 reduces to the state update equation if the Jacobian is constant,
the situation we assume here. Thus Eq. 5.110 becomes

AX =(!~) 1
AF (5.111)

----~------~----
'1 ~~ ----
216 Power System Planning

Since we prefer the best estimate of X instead of AX, we can see that
X=X8 +AX (5.112)
In addition to the best estimate of X, we would like to determine the degree of
spread of X about X8 . This information is embedded in the state covariance matrix Cx
or CAX CAx is, again, obtained from Eq. 2.22 of Chap. 2:

(5.113)

Since X 8 in Eq.5.112 is not random, then

(5.114)

Having the best estimate and covariance of X is important, but it is also


necessary to obtain the same information for the power flows so that we may see to
what effect uncertainty, if you will, in the forecast demands used in the base case will
have on system loading and hence system security. Resorting to Eq. 5.57, we have

Aw = aH AX + aH AU
ax au
Since AU= 0,

AW= aH! AX
ax xS,u,o

Thus

Aw = a HI AX =
1
aH (aF' c _1 aF)- aF' c _1L\F (5.115)
ax xs, u ,o ax ax fl ax ax fl

and
A A

W = W+AW (5.116)
Further, it can be shown that

(5.117)

Given Eqs. 5.113 and 5.117 we can calculate the 99% confidence limits for the
random variables AX and AW. Remembering ,that the diagonal elements of the
covariance matrices contain the variances, a/, of AX and AW, we have

-2.57aXj..;; Xj..;; 2.57aXj


-2.57aw; ..;; W;..;; 2.57aw;
Power-Flow Analysis 217

or

Xs - 2.57ax..;; X..;;
~
Xs + 2.57ax (5.118)
W8 -2.57aw..;; W..;; W8 + 2.57aw (5.119)

5.8.2 Specifying the Covariance Matrix CTI


It is clear from Eqs. 5.110 through 5.116 that the covariance matrix CTI, which defines
the uncertainty of the forecast demands, plays a major role in determining the
confidence limits of both the states and output power flows. In general, if we assume
that the net demands are Gaussian-distributed about their means such that the
uncertainty of the demands is defined by their variances, then for a multiarea system

CT11t CT112 CTJu

(5.120)

and

CTlkk =

where -1 ..;; i/Jii = correlation factor..;; 1


CT1kk is the covariance matrix for area k alone, and CT1kj is the covariance matrix of
areas k and j, which defines the degree of correlation between forecast net injections in
area j and area k. If the area net injections are assumed independent-which is not very
realistic-then Eq. 5 .120 becomes a diagonal matrix, with area covariance sub matrices
as the diagonal submatrices. In the same way, if the net power injections at each bus
within a given area are assumed independent, then the submatrices CTJkk become
diagonal matrices with the variances ap/ and aQ/ on the main diagonal. Of course,
the other extreme is to have all net power injections completely correlated.
The reason for concerning ourselves with the degree of correlation between net
power injections as expressed by the covariance matrix is that in real life loads and
generation throughout the system are somehow correlated, and this correlation
produces reasonable variances in both the state and output quantities. For instance, for
area 2 in the two-area interconnected system used before, we know that

-- --------~----------------- --------------------
218 Power System Planning

because there are no losses in the system. Since P4 is a dependent injection, we have

P4 = -P5 -P6
Therefore
t:J'4 = -t:.P5 - t:J'6
Assuming !l.P5 and !l.P6 are Gaussian-distributed random variables with zero mean and
variances a/ and al, then
Var (t:.1'4) = a4 2 = a/ + a/
Howev.::r, if we assume that !l.P5 and !l.P6 are correlated with a correlation factor of if;,
then
Var (t:.1'4) = a/+ a/+ 1/la5 a6
To illustrate the effect of correlation, let
a/= al 1/1 = -1 .0 and 0.0, respectively
In the case when if;= -1.0, we have
Var (t:.1'4) = a/ = a/
compared to
Var (t:.1'4) = a/ + al = 2a/
when if; = 0.0. The difference is obviously significant; ignoring the correlation results in
a 99% conf1dence interval for !l.P4 that is 40% larger. Stated differently, assuming the
net injected powers are independent produces state variances that are unrealistic.
Another way to justify the necessity and correctness of including the correlation
between the net injected powers is to note that peak bus loads are not necessarily
coincident, although we make this assumption quite often in power-flow analysis. This
assumption is by far the worst case, as we demonstrated above. The fact that peak bus
loads are not coincident loads is easily demonstrated using actual system data. If the
peak loads at each major substation in a system were simply summed, we would find
that this sum of the individual peaks is greater than the actual system peak; correlation
is responsible for this discrepancy. The problem of load correlation has been with
system planners a very long time; it tends to become formidable when forecast total
system peak demands must be distributed to each load center prior to performing
power-flow analysis. Simply proportioning the system peak to each load center on the
basis of relative noncoincident load center peaks is not, strictly speaking, correct. This
difficulty can be circumvented somewhat if, thr;ough the analysis of load center perfor-
mance, the system planner develops some insight into the nature of peak-load correlation.
In addition to the correlation that exists between peak bus loads, there is
obviously a strong correlation between net injected generation and loads. Realistically,
as the loads vary, the economic load-dispatching system distributes the change in load
to the individual generators. This correlation must also be taken into account.

------- .. ---~---------------------------------------- ------- ------------~~


Power-Flow Analysis 219

5.8.3 Solution Procedure


Typically, a stochastic power-flow problem is solved as follows:
1. Define the mean values X 8 , U, D 0 about which the power-flow equations are
to be expanded.
2. Defi~e the covariance matrix C11 as given in Eq. 5.120.
3. Calculate the covariance.matrices C~x and C~w using Eqs. 5.113 and 5.117.
4. Determine the confidence intervals for X and W using Eqs. 5.118 and 5.119.
We must forgo a lengthy example here. The reader should note that examples using
this method are given in the references [l, 3].

5.9 SUMMARY
In this chapter the topics included modeling, power-flow analysis, sensitivity analysis,
reactive power-flow control, optimal power-flow analysis, and finally stochastic
power-flow analysis. Because of the sheer amount of material concerning these topics,
a comprehensive presentation was impossible; however, an attempt was made to discuss
th(, salient topics and associated representative solution procedures. Beginning with a
cursory look at corridor selection and component models, we embarked on a lengthy
discussion of Newton's power-flow method as applied to both single and inter-
connected areas. Through the use of state-space notation, full advantage was taken of
the compactness of matrix notation. We found, in fact, that formulation of the
interconnected-area power-flow problem was a simple extension of the single-area
problem; only the additional constraints on net area interchange needed inclusion.
Using two interconnected three-bus areas, we were able to demonstrate the basic
Newton solution procedure as well as emphasize the necessity of exploiting the sparsity
of the Jacobian matrix.
With the state-space model of an interconnected system, the linear perturbation
equations used in sensitivity analysis were easily obtained. We showed how valuable
information can be obtained by the planner through the use of sensitivity analysis. The
sensitivity of system states and output quantities to changes in system demand was
found to be of particular interest in light of forecast uncertainties. Again, a numerical
example was presented to illustrate the procedure required to obtain the desired
information. In the example, it was noted, the sensitivity matrices were obtained using
the same elimination procedures as in power-flow analysis, a fact of practical
significance.
Because of the increasing amount of EIN transmission, it was necessary to
include a section on reactive power/voltage control. Using the sensitivity results of the
sample system, we demonstrated how the line-charging capacitance of EIN lines can
cause voltage and excitation problems unless properly compensated using reactors, etc.
A static optimization method, based on the Kuhn-Tucker procedure, was presented to
determine if the system control variables can control system voltages without
compensation over the entire load cycle. If voltage control is not possible, then the
220 Power System Planning

proposed technique can be used to determine how much compensation is required by


defining the compensator to be a generator whose scheduled real power output is
always zero.
Optimal power-flow analysis was presented as a method for solving the economic
dispatch and power-flow problems as one problem. Once again, the problem was
formulated as a static optimization problem in which the cost function was the total
system production cost and the equality constraint was the power-flow equation. The
inequality constraint on the states was handled using penalty functions. Although the
Kuhn-Tucker method handles such constraints more formally, penalty function
methods have been used with somewhat more success.
The final topic, stochastic power flows, was included to introduce the reader to
this base topic. At best, this area of power system analysis is yet in its infancy.
Because of the increasing uncertainty faced by system planners, stochastic power-flow
methods will no doubt receive a great deal of attention in the future. To date, the greatest
benefit to be derived from stochastic power flows lies in their inherent ability to
account for load and load-generation correlation. Through the noise covariance matrix
the degree of correlation can be quantified and included in the analysis.
We hope the reader will use the material presented to guide his work and fuel his
thoughts. Once again, all should recognize that no one set of analytical tools will work
satisfactorily for all utilities.

PROBLEMS

5.1. Consider the simple two-bus system shown below:

jX
SG2i

v. I
Q t S02

I 000 1

Develop the mismatch equations and the corresponding state update vector
equation to solve for the system states.
5.2. Solve for the system states in Prob. 1, assuming

V1 = 1.0 X= 0.1
IV2I = 1.05 0.0
Pc2 = 1.0
xo = 0.0
Pn 1 = 1.0 0.0
Pn 2 = 0.5 0.0

Note: Use the methods discussed in Chap. 5, including optimal ordering, to solve
the problem.
5.3. Calculate the sensitivity matrix Sxo, and investigate the variation in the
excitation levels of the two generators as the load profile varies about that given
in Prob. 2.

... . ,,,;,;.i
Power-Flow Analysis 221

5.4. Formulate the optimal power-flow problem for the system defined in Probs. 1, 2,
and 3, assuming

C1 = 2Pc1 + 2Pc1 2
C2 = Pc2 + Pc2 2
and

0 Pc 2 0.75
IVtl = 1.0
IV21=1.05
Solve for the optimal generation schedules and corresponding system state.
5.5. Assume that the reactive power loading on the two units in Prob. 2 is
unacceptable, that it is desired to determine the value of IV2 I that more equitably
loads the units. A possible problem formulation may be as follows:

min (Qc1 - Qc2l


w.r.t. IV2 I
s.t. F(X, U, D 0 ) =0
Using the notions developed in Sec. 5.7 .2, solve for the optimal value of I V 2 I.
5.6. If the load Pn 2 in Prob. 2 is Gaussian-distributed about Pn 2 = 0.5 and has a
variance of o 2 = 0.01, determine the mean and variance of the loading on unit 1.
5.7. Repeat Prob. 6, but assume Pn 1 has a 100% negative correlation with Pn 2 , where

OJ>n 2 = 0.01 OJ>ni = 0.01 t/; = -1 .o


Compare this result with the case in which Pn 1 and Pn 2 are assumed
independent, t/; = 0. Of particular interest is the effect of load correlation on the
loading of unit 1.
5.8. Develop a computer program to perform power-flow analysis using Newton's
method, and verify the results obtained by hand in Prob. 2.
5.9. Extend the basic power-flow program to calculate the confidence intervals for all
state variables, given the statistics of the bus loads. Verify the results obtained
for Pro bs. 6 and 7.

BIBLIOGRAPHY
1. Aboytes, F. and B. J. Cory: An Alternative Formulation of the Stochastic Load Flow Method,
PICA Proc. 1975.
2. Billington, R. and S. S. Sachdeva: Real and Reactive Power Optimization by Suboptimum
Techniques, IEEE conference paper CP596-PWR, 1971.
3. Borkowska, G.: Probabilistic Load Flow, IEEE Trans., vol. PAS-93, pp. 752-759, 1974.
4. Britton, J. 'r.: Improved Area Interchange Control for Newton's Method Load Flows, IEEE
Trans., vol. PAS-88, pp. 1577-1581, 1969.
5. Carpentier, J.: Contributions to the Economic Dispatch Problem, Bull. Soc. Fr. Electr. vol. 8,
pp. 431-447, 1962.
6. Dommel, H. W. and W. F. Tinney: Optimal Power-Flow Solutions, IEEE Trans., vol. PAS-87,
pp. 1866-1876, 1968.
222 Power System Planning

7. Dopazo, J. F. et al.: Stochastic Load Flows, IEEE Trans., vol. PAS-94, pp. 299-309, 1975.
8. Elgerd, O. I.: "Electric Energy Systems Theory: An Introduction," McGraw-Hill, 1971.
9. Heydt, G. T. and B. M. Katz: A Stochastic Model in Simultaneous Interchange Capacity
Calculations, IEEE Trans., vol. PAS-94, pp. 350-359, 1975.
10. Kuhn, H. W. and A. W. Tucker: Nonlinear Programming, Proc. 2nd Berkeley Symp. Math. Stat.
Probab., pp. 481-492, 1951.
II. Nagel, T. J. and G. S. Vassell: Basic Principles of Planning VAR Control on AEP's System,
IEEE Trans., vol. PAS-87, pp. 488-495, 1966.
12. Ogbuobiri, E. C. et al.: Sparsity-Directed Decomposition for Gaussian Elimination on Matrices,
PICA Proc., 1968.
13. Peschon, J. et al.: Optimum Control of Reactive Power-Flow, IEEE Trans., vol. PAS-87, pp.
40-48, 1968.
14. Peschon, J. et al.: Sensitivity in Power Systems, IEEE Trans., vol. PAS-87, pp. 1687-1696,
1968.
15. Sasson, A. M.: Non-linear Programming Solutions for Load-Flow, Minimum-Loss,andEconomic
Dispatching Problems, IEEE Trans., vol. PAS-88, pp. 399-408, 1969.
16. Stagg, G, W., and A. H. El-Abiad: "Computational Methods in Power Systems Analysis,"
McGraw-Hill, 1968.
17. Sullivan, R. L.: Controlling Generator MVAR Loadings Using a Static Optimization Technique,
IEEE Trans., vol. PAS-91, pp. 906-911, 1972.
18. Tinney, W. F. and C. E. Hart: Power-Flow Solutions by Newton's Method, IEEE Trans., vol.
PAS-86, pp. 1449-1460, 1967.
TRANSMISSION SYSTEM
RELIABILITY ANALYSIS

We have discussed some rather sophisticated analysis techniques for identifying plans
which, under normal operating conditions, meet desired performance criteria. We might
think that the planning process is complete, that the expansion plan that performs best
can be implemented. Upon further consideration we will realize that such expansions
are satisfactory only if the system operates free of disturbances in the normal mode.
Since no system ever operates in a contingency-free environment, we must test the
expansion plans to insure that the performance required is maintained in the face of
unavoidable disturbances. There are many factors to consider in protecting a system.
To maintain a high performance record as determined by voltage and frequency as well
as by frequency and duration of outages, a system must be able to survive abnormal
disturbances, such as faults and loss of system elements, created by equipment
malfunctions, switching surges, lightning strikes, etc. The protection problem can be
divided into subproblems because disturbances usually evolve in three phases: subcycle,
cycle, and steady state. Subcycle disturbances are associated with lightning strikes and
switching surges that create very large voltage waves on the transmission system. If
adequate protection by the use of lightning arresters does not exist, then short circuits
can be created, thus extending the disturbances into the cycle or transient range.
Transient disturbances are critical because their time duration is such that both the
223

I
I,,..__ ---------
224 Power System Planning

electrical and mechanical dynamics of the system are activated. If adequate protection
is not built into the system, large oscillations in power flow, abnormal voltages and
frequency, and system deterioration can result.
In this chapter we concentrate on the effects of a steady-state disturbance on the
system. In particular, we shall focus on techniques for determining if the state of a
system is acceptable after a disturbance has reached steady state. Generally, acceptable
will mean that no system component is overloaded and that all demands are met at
acceptable voltages.
Over the years steady-state contingency analysis has become more and more
sophisticated, primarily because of the development of the digital computer. In the
early days the network analyzer was used to simulate the effects of system element
outages on line loading, etc. Although we have efficient and effective AC power-flow
methods today, they are frequently computationally unattractive and too demanding
of engineering time for use in contingency studies. Rather, the industry prefers faster
approximate methods for locating potential trouble spots, to be studied in further
detail, if necessary, using the AC power flow. We begin our discussion of contingency
analysis by briefly considering methods that are strictly deterministic and proceed to a
more lengthy discussion of a method in which a probabilistic description of system
elements is used to determine loss-of-load probability, LOLP, and expected value of
demand not served, e(DNS). We shall discuss the application of the contingency
analysis methods for evaluating both isolated and interconnected areas, again using
the three-bus system of earlier chapters.

6.1 DETERMINISTIC CONTINGENCY ANALYSIS


Contingency analysis methods that do not model the outage probabilities of system
elements but rather simulate element outages by actual removal are considered to be
deterministic methods. The use of AC power-flow methods for contingency analysis
may be considered a deterministic approach, characterized by excellent accuracy but
otherwise expensive computationally and excessively demanding of engineering time.
Because contingency analysis is only a tool for detecting possible overloading
conditions requiring study by the planner, computational speed and ease of detection
are paramount cor.siderations; using AC power flows for this purpose is considered
unnecessary. Methods appropriate for contingency analysis are based on DC power-flow
equations and bus voltage equations. Each of these methods will be considered.

6.1.1 DC Power-Flow Contingency Analysis


The simplest, but perhaps most inaccurate, method for analyzing the effects of not
only single but multiple contingencies is a method based on the DC power-flow
equations. These equations, usually N - 1 in number, where N is the number of buses,
model only the real power flow and ignore the reactive flow. In many cases all line
resistances are neglected, and the line flows are assumed to be considerably smaller
than the steady-state stability limit of the line. The result of all these assumptions is
that we obtain a linear model of the network to facilitate performing

-------------~-----~---~-------
Transmission System Reliability Analysis 225

multiple-contingency outages using the principle of superposition. To develop the DC


power-flow equations, we must start with the standard AC power-flow equations:

P; = Real. (vt
'
l Y;i vi)
j
(6.1)
i= l, ... ,N
Q; = -Im (vt l Y;ivi) (6.2)
j

Assuming the voltage magnitudes remain unchanged and equal to their base-case values,
the N net injected reactive power equations can be ignored, leaving only the N real
power equations. Since we must always specify one of the angles as our reference bus,
we delete one of the equations in Eq. 6.1. Deleting the equation corresponding to the
reference bus, or slack bus, which we define as bus 1, we have

P; = Real (vt l Y;;V;) i = 2, ... ,N


j

= Real ~V;le -;o; i Y;ilei iil Vi


I
0 0
lei ; )
I

= Real (LIV; II V; IIY;;le;(e;; + 0i -o;))


j

With the line resistances neglected,

and

P; = LIV;IIV;IB;; sin (o; - o;) i = 2, ... ,N


j

Under the assumption that


IV;IIVilB;; sin (o; - o;) /4i; IV;IIVilB;;
o; - o;must be sufficiently small that the sine of the angle difference can be replaced
by the angle difference. Thus

P; = LIV;IIV;IB;;(o; - o;)
j
i-1 N

=-
j=I
I IV;IIV;IB;;o; + o; l
i*i
IV;IIV;IB;; - l
j=i+I
IV;IIV;IB;;o;
226 Power System Planning

Letting

K;; == l
j ,,t. j
IV;IIVilB;i

K;i == -IV;IIVilB;i
we get

i-1 N

P; == L
j=I
K;/.'i + Kuo; + l
j=i+I
K;ioi i==2, ... ,N

or in matrix form

P == Ko (6.3)

Oearly, if for a fixed set of power injections P, a line or lines are removed, both the K
matrix and the o vector will change from their base-case values K0 , os by an amount
LlK and Ao such that
po == (K 0 + AK)(os + Ao)
== K 0 os + A.Kos + K 0 Ao + AKAo
If the nonlinear term AKAo is dropped, then
po 2'c K0 os + AKos + K 0 Ao (6.4)
But from Eq. 6.3
po= Ko 0s

or
(6.5)
Equation 6.5 provides the changes in the bus voltage angles due to network changes.
Since the change in the power flow in line ij is
(6.6)
where Ao i and f:10 i are the ith and jth components of f:10, the effects of line outages
on line flows are readily available.
It is important to emphasize that (K 0 )- 1 need only be computed once, namely
at the beginning of the contingency studies. As long as the base case remains the same,
(K 0 )- 1 is valid. '
Using Eq. 6.5 is a simple matter: If single contingencies are to be studied, such
as the removal of line pq, then the only elements in K 0 that change are KPP' Kpq, and
Kqq, where
Transmission System Reliability Analysis 227

/:i..Kpp = Kpq
l:i..Kqq = Kpq
(6.7)
l:i..Kpq = -Kpq
l:i..Kqp = -Kpq

Substituting the information in Eq. 6.7 into Eq. 6.5, we have

0 0 0 0

0 Kpq -Kpq 0
!:i,.f,PQ = -(Ko)-1 r,s

0 -Kpq Kpq 0

0 0 0 0
t t
pth col. qth col.
1
= - (K f (o/!:i..Kp + o/.6.Kq)
0 (6.8)

where /:i..Kp and 1:i..Kq represent the pth and qth columns of the /:i..K matrix. The
changes in line flows due to the loss of line pq given are obtained by substituting the
appropriate elements of t:,.oPq into Eq. 6.6. Equation 6.8 may be used to obtain the
effects of any single contingency on line flows by substituting the appropriate data
into /:i..K. For instance, if we wanted to determine the effects of removing line mn,
then it would only be necessary to determine .6.Km, .6.Kn, om 0 , and Ono and to
substitute into Eqs. 6.6 and 6.8 to obtain
.1.r,mn = -(K0 r1com8 .\Km + On9 .6.Kn) (6.9)
and
!:i,.Pl/..mn= -K--(l::..Omn_!::,.oJ71n)
l/ 1 J
foralli1"..J..mn
-,...

In this way we easily obtain the effects of all single contingencies on system power
flows, and we can use this information to identify any overload conditions that must
be resolved.
To analyze the effects of double and higher-order contingencies is straightforward
since Eq. 6.5 is a linear function of the network changes .6.K. If we want the change in
line flows that will occur with the removal of both lines pq and mn, we should note
that no additional computational work is necessary, because
(6.10)
and
/:i..P-pqmn
l/
= !:i,.P.PQ
l/
+ !:i,.Pl/..mn iJ. =I= pq or mn
228 Power System Planning

This is the powerful feature of the DC power-flow method. The ease of computing the
effects of multiple-contingency outages from single-contingency information offsets the
inherently poor accuracy of the method.
It is easy to show why superposition is valid in this situation: Simply remove both
lines pq and mn simultaneously, calculate .D.K, and evaluate ..18 using Eq. 6.5. In this case

0 0 0 0 0 0

0 -Kpq 0 0 0

AK= 0 0 0 0

0 0 0 0

0 0 0 0

Therefore Eq. 6.5 becomes

Ali= -(K 0 )- 1 (8/..1Kp+ 8/AKq + lim8 AKm + lin8AKn)


= -(K0 )- 1 (8/..1Kp + 8/AKq) -(K0 f\lim8 AKm + lin8 AKn)
~

.D.lipq
'-~--~ ~--./
Alimn
= .,1[iPq + .D.limn
and
M';; = M';/q + M';;nn
To illustrate how this method works, consider the problem involving the
three-bus system shown in Fig. 6.1, in which it is desirable to determine the effect that
removing line 23 has on system flows. Given the base-case results,

J
~ ~ -~
~
P
0
= 0.5~
0.25 IVIS= .0 [jS = 0.0
0.0 ws = 12
P 13 = 0.5
= 0.0J
-0.75 .0 -0.025 Ll
P23 = 0.25

and given the DC power-flow equations,


P2 = 2082 + 10(82 - 83)
= 3002 -1083
P3 = 2003 + 10(03 - 82)
= -1082 + 3083
Transmission System Reliability Analysis 229

-,---'--,-- 1 2

j0.05

Su= 0.75 +iO.O

Fig. 6.1 Three-bus sample system.

Eq. 6.5 becomes

r J
lio = -

r
l 30 -1~-1 t-10
-10 3aj
107
o.o
IO -1aj t-0.025

= 0.006257
[=--o.0062sj

Thus, using Eq. 6.6,


M21 = -K21 D.fi 2 = 20(0.00625) = +0.125
M13 = +K13D.fJ3 = -20(-0.00625) = +0.125
and
M23 = -K23(D.fi2 - D.fJ3) - l1K 23(fi/ - fl/)
= 10(0.0125) - 10(0.025)
= -0.125
The conclusion to be drawn is that losing line 23 increases the loading on lines 21 and
13, where, according to the analysis, the loading on line 13 increased by 25%. Even a
novice can see that the results are in error by a substantial amount. We know that
losing line 23 should result, with the same generation, in an increase in flow of 0.25
p.u. on lines 21 and 13, instead of 0.125 p.u. In practice, results obtained from this

'
I

L
230 Power System Planning

method are better because there are usually more line connections per bus than in the
sample system and because the loss of a line affects a smaller percentage of the K matrix.

6.1.2 A Z Matrix Method for Contingency Analysis


Z matrix methods, as the name indicates, make use of the bus impedance matrix
associated with both the base-case system and the system modified by either line
removals or line additions [3, 4]. The Z matrix for a system can be obtained in several
ways. It can be obtained directly by inverting the bus admittance matrix, or it can be
constructed sequentially by using available algorithms. The latter approach to obtaining
the Z matrix for large systems is favored in most cases, because of the ease with which
line changes can be made without rebuilding the entire Z matrix-that is, computa-
tional efficiency and convenience. The fundamental approach to contingency analysis
using the Z matrix is to inject a fictitious current into one of the buses associated with
the element to be removed, of such value that the current flow through the element
equals the base-case flow; all other bus currents are set equal to zero. In effect, this
procedure creates throughout the system a current flow pattern that will change in the
same manner as the current flow pattern in the AC power-flow solution when the
element in question is removed. We must emphasize that only the changes in flows are
meaningful; the requirement that the fictitious injected current create the same current
flow in the element to be removed as in the base case is considered a constraint.
Implementing the ideas discussed, assume that we want to evaluate the effects of
an outage of line pq, given that the base-case system flows and voltages are available.
We may include all MVA loads in the Z matrix by first converting the MVA loads to
impedance loads using
2
IV;l (6.11)
Z1oad i = S="
l

and applying any of the Z matrix building algorithms. Knowing the Z matrix, we may
write the bus voltage equation:

= (6.12)

Since we do not yet know how much current to inject into bus p to create the flow
Ipq that equals the base-case flow Ip/, it is logical to start by letting Ip= 1.0 p.u.
Thus, from Eq. 6.12, injecting a unit current into bus p gives

-------~--~-~--------~
Transmission System Reliability Analysis 231

With the bus voltages known, we can calculate lpq:

Vp-Vq
Ipq = z line pq
Zpp -Zqp
= Zunepq
(6.13)

Obviously, lpq =t Ip/ the first time; therefore we need to calculate an adjustment
parameter, d, where
s
f:,. Ipq
d=- (6.14)
Ipq

Multiplying d by the initial unit injection results in a new injection, IP= d, and new
current flow,

(6.15)
Due to the injection IP = d, the current flows in all the other elements are

I--=---
vi- vi
1/ Z1me 11.. all ij =t pq

= (6.16)
Zune ij

The next step is to remove line pq and calculate the new current flow pattern /;;
in the new system. Once the new currents are available, the sought-after current flow
changes due to removing line pq are
Mii = l;j - Ii; for all ij (6.17)
Of course, when ij -
= pq, lpq = 0, therefore
AT
Ul.pq -
- - Ipq
s

With this result the need for adjusting the fictitious injected current Ip to create a flow
of I pq 8 in line 'PQ is seen to be crucial.
Calculating the current flow pattern
in the modified network Z, in which line
pq has been removed, requires only that we inject current Ip = d, as before, into the
modified network. The voltages that result from this injection can be used to
determine the needed flows:
232 Power System Planning

~\ zu Z12 Z1N 0 Z1pd

= d = (6.18)

VN ZN! ZN2 ZNN 0 ZNpd

Therefore

7ij =
v--v-
I I
Zune ij
d(Z;p - zjp)
= (6.19)
Zune ii
where 4,q = 0. Further, substituting Eqs. 6.19 and 6.16 into Eq. 6.17, we get that
M;i = z.d .. [(Z;p
- - -
Zjp) - (Z;p - Zjp)] (6.20)
lme 11

The modified network Z must be known before Eqs. 6.18 through 6.20 can be
used. It is determined using the building algorithms mentioned above. Fortunately, in
networks characterized as consisting of passive, bilateral elements that are not
magnetically coupled, the algorithm for removing and adding elements is straight-
forward. Line removals are simply special cases of line additions; that is, to remove a
line with impedance Zlinepq, it is only necessary to add, in parallel with the line to be
removed, a line of impedance -Zlinepq Thus, line removal algorithms are identical to
line addition algorithms, which we will discuss below.
Consider the addition of a line between buses p and q in the network shown in
Fig. 6.2. Initially we simply connect one end of the new element to bus p, with the

3



-Ziinepq

Fig. 6.2 Network in which line
q pq is being removed.

0
Transmission System Reliability Analysis 237

Af12 = li2 - 112 = -0.25 p.u.


Af13 = /,.3
- 113 = 0.25 p.u.
Af23 = J.23 - 12 3 = -0.25 p.u.
To faci,litate comparing these results with those obtained using the DC
power-flow approach, we note that

DP1 2 = Real (V1Af12)


= Real (1.0(-0.25)) = -0.25 (compared to -0.125 for the DC method)
N 13 = Real (Vi Af13)
= Real (1.0(0.25)) = 0.25 (compared to 0.125 for the DC method)
and finally

Af>23 + Af113 = Af13


but llP3 = 0; therefore
Af123 = -Af113 = -0.25
The accuracy of the Z matrix method is far better than the accuracy of the DC
power-flow method for the example worked. In fact, the results are excellent; they are
comparable to those obtained using an AC power flow.

6.2 PROBABILISTIC TRANSMISSION SYSTEM


RELIABILITY ANALYSIS
In Sec. 6.1 an attempt was made to introduce the reader to current thinking on
deterministic techniques for contingency evaluation. These methods, very useful in
their own right, do not render indices of reliability-figures of merit that facilitate
comparing alternative expansion plans or, when possible, associating a cost with
reliability. In the last few years new techniques have been developed that do provide
the planner with the capability to evaluate a transmission system in terms of two
indices, namely, LOLP and E(DNS). In particular, these methods enable the system
planner to describe ,the capacity of each element in the system to be evaluated,
including generating units, with a probability density function, in the manner in which
the outage capacity of generating units was described in Chap. 3. With this information
the LOLP and E(DNS) for the system can be determined. In addition, the LOLP and
E(DNS) associated with each element are easily obtained, such information being
indicative of the relative role each element plays in determining the reliability of the
system. Presumably, if the reliability of a given expansion plan is not sufficiently high,
those elements most responsible for poor reliability can be identified. Generating units
are included in the transmission system analysis because the outage characteristics of
generating units clearly influence the loading of the system and hence its reliability.
The probabilistic approach to reliability analysis is a powerful tool, but it must
be used properly. In no way is it suggested that these methods replace detailed analysis
238 Power System Planning

of the system using AC power-flow methods. Probabilistic methods should be used to


seek out and identify transmission bottlenecks and trouble spots. Used in this way,
such methods provide additional insight, reduce computing time, and, most important,
allow the planner to focus attention on areas most likely to create transmission
problems.

6.2.1 Probabilistic Element Models


In Chap. 3 we discussed in some detail the concept of forced outage rates for
generating units, which described the outage capacity density function for units when
two-state models were used. The same type of probabilistic description will be used for
other transmission elements in discussing transmission system reliability. We assume
that based on operating experience, each transmission element can be described
adequately by two-state models and associated FORs. Although the complexity of
methods discussed below would greatly increase if higher-order element models were
used, the basic philosophy and computational details would not change. However, for
simplicity two-state models will be used throughout.
6.2.2 Basic Philosophy
Given that each element, Vm, m = l, ... , E, in the system under study can reside in
either the "O" state, with probability qm, in which it has no capacity and is out of
service, or the "1" state, with probability Pm, in which it has capacity Cm and is in
service, the system will have 2 distinct capacity states X;, i = 1, ... , 2. For
instance, the three-bus system has five elements: two generators and three lines;
therefore, that system can reside in any one of 2 5 = 32 different capacity states X;.
Obviously, the upper and lower limiting states denoted by X and ~. within which all
other states exist, would be state X = {1, 1, 1, 1, 1) and state ~ = (0, 0, 0, 0, 0).
Associated with each of the 2 states is a probability /(Xi) that it will occur; for
example, the probability,/(X), that the three-bus system will reside in the upper limiting
state is
E
f(X) = fl Pm = P1P2P3P4Ps
m=l

Similarly, the probability that the system will reside in the lower limiting state would
be
E
/()5) = n
m=l
qm = Q1Q2Q3Q4Q5

In general, the probability that the system will, reside in any state Xi = (V1 , V2 , ,

VE) is

(6.32)
m=I
Transmission System Reliability Analysis 239

wheref(Vm)=Pm if Vm=l
f(Vm)=qm if Vm =0
As the ,definitions and symbology just introduced are fundamental to the method
being discussed, several readings of that section may be appropriate. In fact, this entire
chapter is fraught with a complex notational system not amenable to simplification.
The problem we face is conceptually simple but computationally tedious: We
must decompose the set Xi, i = 1, ... , 2E, of system capacity states into states that
are acceptable and states that are unacceptable. Unacceptable states are system
capacity states ~ for which the load, L, cannot be satisfied, either because of
insufficient generation capacity or because of insufficient transmission capacity.
Assuming we can single out the unacceptable states, the system LOLP is defined by

Xi = all unacceptable states (6.33)

In the same vein, if for each unacceptable state Xi, which has a probability /(Xi) of
occurring, the amount of load served is gi <L, then the E(DNS) would be the sum of
the products of demand not served and the probability that the associated state
occurred:

E (DNS) = l f(Xi)(L - gi) Xi = all unacceptable states (6.34)

Note that L - g; is the amount of demand not served because system capacity state Xi
is unacceptable:
DNS =L -gi (6.35)
We must agree that the conceptual implementation of these ideas is straightforward:
We need only determine which states of the 2E are unacceptable; calculate their
probabilities of occurrence, /(Xi), and corresponding flows, g;; and substitute this
information into Eqs. 6.33 and 6.34. Since the LOLP for each element is the
probability that it is among those elements (called minimum-cut elements) whose
individual capacity states are responsible for the insufficient flows g; when the system
resides in state Xi, we have that

LOLP m = l f(Xi) Xi = all unacceptable states for which


element m is a minimum-cut element
(6.36)

As we shall continually be using in this chapter the idea of network flows and flow
algorithms, it must be understood that these flows are not constrained by electrical
laws, except for conservation of flow at a bus. The constraints that the flow patterns
must satisfy fo be feasible are constraints due only to the capacity of the elements.
The electrical parameters of the system are not directly used as in DC power-flow and
Z matrix methods. For instance, the graph for our three-bus system is as shown in Fig.
6.3a, and the only characteristics it has in common with the actual electrical network
are the element capacities and the topology of the graph. Further, because many of
240 Power System Planning

the network algorithms we use require single source and demand buses, it is common
to represent all loads by perfectly reliable (q = 0) elements whose capacity equals the
load represented in Fig. 6.3a. An example of a feasible flow pattern corresponding to
L 1 = L 2 = 0 and L 3 = 4 p.u. and
c1 = 2 p.u. generation 1 capacity
c2 = 2 p.u. generation 2 capacity
c3 = 2 p.u. line 13 capacity
c4 = 3 p.u. line 12 capacity
c5 = 4 p.u. line 23 capacity
is depicted in Fig. 6.4. Note that feasible flow patterns are not unique; that is,
there is usually more than one flow pattern that will satisfy a given demand profile.
Agreeing that the approach is indeed conceptually simple, we might be tempted
to implement the procedure without further consideration. However, it becomes
painfully obvious that the procedural details of implementing the ideas for determining
the indices of reliability desired are impractical. For instance, if we were to attempt to
evaluate a relatively small system with just 100 elements, we would be faced with the
immediate problem of classifying 2 100 = 1 .267 X 10 30 capacity states. That complete
enumeration of all these states is not practical is obvious. We must look for methods

(a)

s L

(b)

Fig. 6.3 (a) Graph for three-bus system. (b) Graph of three-bus
system with a common demand and source bus.

- - - - - - - - ~ - - - - - - - - -----------
Transmission System Reliability Analysis 241

>----= 4.0 p.u.


3 L,

Fig. 6.4 Feasible flow pattern for three-bus system with L 3 = 4.0 p.u.

of grouping the capacity states into subsets, where each subset will contain capacity
states having common characteristics to facilitate determining the information desired.
Thus we must concentrate in subsequent sections on discussing such methods. As these
methods involve the use of concepts associated with network flow theory, it will be
necessary at times to digress and explain the bases of these notions.
Before proceeding to a detailed discussion of the major steps required in
determining the desired reliability indices, we shall outline the basic solution procedure
to be used. In this way the reader may digest more easily the material to come. In
broad terms the procedure for calculating the reliability indices is as follows:
1. Define the initial limiting capacity states X and ~-
2. Decompose the set of all capacity states into subsets of acceptable,
unacceptable, and unclassified states.
3. Determine the limiting capacity states of the unclassified states obtained in
step 2, and decompose these states into acceptable and unacceptable states.
Repeat if necessary.
4. Calculate the LOLP, which is the probability that the system resides in
unacceptable capacity states defined in steps 2 and 3.
5. Using network flow theory, calculate the maximum flow that can be
transmitted through the network to the loads for the set of unacceptable
states.
6. Determine the E(DNS) by summing the products of the demand not served
and associated probability. ~

6.3 CAPACITY STATE CLASSIFICATION BY SUBSETS [8]


. Consider the set of all 2E capacity states as depicted in Fig. 6.Sa. Initially they are all
unclassified; we must identify those states that are acceptable and those that are
unacceptable. For this purpose it is desirable to define capacity states, called upper and
lower critical capacity states and denoted by x0 and X 0 , such that any system state
x
Xi;;:,; 0 is acceptable and any capacity state X; < X0 is unacceptable. Those capacity
.~',
242 Power System Planning

x,
Xi+ 1
Xi

x 3

(a)

B xi< XO
Unacceptable

(b)

Fig. 6.5 (a) Venn diagram for system capacity states. (b) Major subsets
of system capacity states.

states X 0 .;;;; Xi< x 0 will be defined as unclassified, meaning that, as given, x 0 and X 0
are insufficient to ascertain the acceptability of these states. For ease of discussion,
assume that all acceptable states, based on X and X 0 , are collected into a subset
denoted by A. Similarly, assume that the unacceptable states, again based on x0 and
X 0 , are defined by subset B, Finally, let the subset of all remaining unclassified states
be denoted by U. Pictorially, we have decomposed the original set of unclassified states
into three new subsets, as shown in Fig. 6.5b. Since there still exists a subset of
unclassified states U, the process must be repeated until all unclassified sets are
exhausted. Given a methodical procedure for determining the critical capacity states
X and X 0 , the basic decomposition process is performed rather easily.
After the first classification step we have a situation as depicted in Fig. 6.5b.
Since all of the states are represented in the figure, we should note that to determine
the intermediate LOLP at this solution stage involves determining the probability that
states will reside in subset B, since
Transmission System Reliability Analysis 243

LOLP ~ f(B) (6.37)


In addition,
f(A) + f(B) + f(U) = 1 (6.38)
Equation 6.38 ~an be used as a stopping algorithm:
f(U) = 1 - f(A) - f(B) (6.39)
and if f(U) < a, then further classification may be halted on the basis that the
probability of the events defined by U taking place is too small to be considered in the
planning process.

6.3.1 Determining the Upper and Lower Critical


Capacity States X0 and X0

Given the guidelines above, we must entertain the problems associated with
determining the critical capacity states as well as probabilities f(A), f(B), and f(V) if U
is not exhausted. Consider the critical capacity states first. The following procedure is
suggested. For x 0 we must
1. Determine the limiting states X and ~ based on the unclassified set being
considered. Initially
x: = (1, 1, ... , 1)
and
~ = (0, 0, ... , 0)
since all states are unclassified.
2. Using the maximum-flow algorithm, obtain a feasible flow pattern for the
network and load profile being considered, where the flow, lm, through each
element must be retained.
3. Determine the state values for each element of X 0 (V 1 , V2, ... , vE 0 ) using
the feasible flows, lm, m = 1, . .. , E, as follows:
Vmo =0 if lm =0
(6.40)
Vm 0 =1 if0<lm <cm
The simplicity of this procedure is obvious from the steps just described, since all
we are saying is that element capacity states greater than the feasible element flows are
acceptable states. For the novice the computational details involved in implementing
the maximum-fiow algorithm may be of modest difficulty, but we can minimize these
difficulties by digressing in Sec. 6.3.3 to discuss the so-called maximum-flow/
minimum-cut algorithm.
To illustrate the procedure, consider determining x 0 for the three-bus system,
assuming all states are initially unclassified and hence belong to the subset U. Further

i
244 Power System Planning

assume that the flow pattern resulting from implementing the maximum-flow algorithm
is shown in Fig. 6.4. Therefore
11 = 2.0 p.u. C1 = 2.0 p.u.
/2 = 2.0 p.u. C2 = 2.0 p.u.
/3 = 2.0 p.u. C3 = 2.0 p.U.
/4 = 0.0 p.u. C4 = 3.0 p.U.
/5 = 2.0 p.u. C5 = 4.0 p.U.
From step 3 and Eq. 6.40 we get

Vs= 1
or x0 =(I, 1, I, 0, I) and all states Xi ;;;. x 0
are acceptable.
As we shall see, the procedure for determining the lower critical capacity state
X0 is as easy as that for determining X0 . Since X0 defines the boundary between
unclassified states and unacceptable states, we expect that it can be defined on the
b~sis of a flow pattern involving the network as it exists in the simplest possible
contingency states. Pondering this point, we might conclude that single contingencies,
the loss of one element, are the simplest contingencies. Thus X 0 is determined by
establishing a feasible flow pattern in the network for each different single-contingency
state, of which there will be E. For contingency i then, V; = 0, and the system resides
in state Xi=(V 1 , V 2 , , Vm,, VE) where all Vm =l=V; equal unity. Using this
philosophy, if with element i out, V; = 0, the maximum flow through the system is
less than the demand, then not only is that state unacceptable, but all states in which
V; = 0 will be unacceptable. By analyzing one single-contingency state, in effect we
classify a number of states and avoid complete enumeration. For use later, the
procedure for determining X 0 is as follows:
I. Remove element i from the network defined by the capacity state X, initially
the full net work since all states are initially unclassified.
2. Using the maximum-flow algorithm, establish a flow pattern in the network.
3. Set the ith element of X 0 = (V 10 , V 20 , , V; 0 , . , VEo) as follows:
V;o =I if g<L
{6.41)
V;o =0 if g=L
4. Augment i ~ i + 1 and return to step I. Repeat until i =E and stop, since X 0
is then completely known.
Using the three-bus system as an example, let i = 1, and determine V 10 , the first
component of X 0 Applying the maximum-flow algorithm with element I removed
results in the flow pattern shown in Fig. 6.7a. Obviously, since g = 2.0 p.u. is less than
the demand L = 4.0 p.u., this single-contingency state is unacceptable; therefore
V 10 = 1. Continuing in the same manner for other single contingencies gives
X0 =(1,I,0,0,l)
Transmission System Reliability Analysis 245

6.3.2 Exhausting Subset U


With the upper and lower critical capacity states known, the probability of the system
residing in the unclassified subset U,f(U), can be determined. As stated earlier, if f(U)
is sufficiently small, we may proceed to calculate the reliability indices. However, if
this is not the case, or if it is desired to exhaust U, in which case x 0 = X0 , and all the
states are classified as either acceptable or unacceptable, then further classification of
the states Xi E U is obviously necessary. To exhaust subset U it is first decomposed into
subsets Uj, where those states in Ui must satisfy the following:
(6.42)
if

Vio,.;; V;< V;
and
for m <i
for m >i
This decomposition rule creates nonoverlapping subsets with limiting states ~(Uj) and
X(U;). Given the limiting states for each subset U;, those states that are acceptable,
unacceptable, and again unclassified are determined by evaluating the critical capacity
states X 0 (Uj) and X 0 (Uj) and by applying the definitions of acceptability, etc., used
earlier.
Summarizing the results obtained for the sample system after the initial
classification process, we have
x = (1, 1, 1, 1, 1)
~ = (0,0,0,0,0)
x = (l,1,1,0,1)
0

Xo = (1,l,0,0,1)
and
25 = 32 states
We may conclude, then, that the initial classification results in
2 acceptable 1,l,1,1,1
states { 1,1,l,0,1
1,l,0,l,l
2 unclassified { 1, 1,0,0, 1 where X0 < Xi E U < X0
3

0,l,l,1,1
28 unacceptable { where Xi< Xo
o,o,o,o,o
32 states

L
246 Power System Planning

111e power of the classification procedure is evident from this example in that, of
the 32 states, only 2 states are unclassified after the initial classification. The two
unclassified states when further decomposed intp subsets fall into subset U 3 , in which
x>(U 3 )=X 0 (U 3 )=(1, 1,0, 1, 1). Thus the staie (I, 1,0, I, 1) is an acceptable state
belonging to subset A, and state (1, I, 0, 0, 1) is an unacceptable state belonging to B.
For the sample three-bus system we have 3 acceptable states and 29 unacceptable
states for the load profile and element capacities chosen.

6.3.3 Maximum-Flow/Minimum-Cut Algorithm


Since we shall frequently use the maximum-flow/minimum-cut algorithm, it behooves
us to set down a few pertinent definitions:
I. A graph consists of a set of buses and a set of pairs of buses called elements.
A graph is connected if there exists a path from any bus i to any bus j. In a
directed graph the pairs of buses are ordered, indicating the unidirectional
flow capability of the element.
2. A cut, ks-L, in a directed graph containing a source node, S, and a demand
node, L, is defined as a set of elements whose removal from the graph breaks
all directed paths from bus S to bus L. The set of all such S - L cuts is
denoted by K.
3. The capacity of a cut, c(ks-L), is the sum of the capacities (assumed to be
all positive) of all the elements defining the ks-L cut.
4. A minimum cut, C(M), is a cut M whose capacity is less than or equal to the
capacity of any other S - L cut. Note that it is possible to have more than
one minimum cut in a graph, a fact that will cause some difficulty for us
later.
5. The set of flows, ls-L, in a graph is said to be feasible if for each flow
pattern
a. The flow out of the source bus is positive.
b. The mismatch at each bus is zero.
c. The flow out of the load bus is negative.
d. The flow llm I on each element ~Cm.
We are now ready for the maximum-flow/minimum-cut theorem.
Theorem: For all feasible flow patterns

max Os-d = min (c(ks-L)) (6.43)


K

Although we shall not prove Eq. 6.43, a functional as well as convincing proof given
by Ford and Fulkerson [6] leads to a very efficient method for constructing flow patterns
computationally that will maximize ls-L. These methods are the foundation of the
basic algorithm as we use it.
Implementing the algorithm based on Eq. 6.43 enables us to determine not only
the maximum flow through a network but also the minimum-cut elements and
Transmission System Reliability Analysis 247

associated capacities. For us this is crucial because the minimum-cut elements are, in
some sense, responsible for not being able to satisfy the demand requirements
when the system or network resides in an unacceptable capacity state. More will be
said of this later, after we finish our discussion of the basic algorithm itself.
Computationally, to calculate the maximum flow through a graph three steps
are involved:
1. Using the labeling algodthm, the buses of the graph must be labeled in such a
way that an S - L path is identified through which flow can be increased, or
augmented, starting with a feasible flow pattern including zero. If it is not
possible to label the demand bus starting with the source bus, the algorithm
terminates, and the elements between labeled and unlabeled buses are the
minimum-cut elements.
2. Using the augmentation algonthm and the labeled graph, flow is augmented
along S - L paths until, because of element capacity constraints, the flow
cannot be augmented further.
3. Return to step 1.
As each of the two algorithms required to implement the maximum-flow/
minimum-cut procedure is quite detailed, we devote a separate section to each
algorithm and include examples utilizing the three-bus system.

6.3.4 Labeling Algorithm


The augmentation algorithm by itself does not provide a practical method for
maximizing, ls-L, since it does not contain a procedure for finding or identifying a
so-called "augmentation path." However, the labeling algorithm, as mentioned, does
provide us with this capability, and very efficiently, because in identifying an
augmentation path, each bus is to be examined at most once.
The basic idea behind the labeling algorithm is to identify an augmentation path
by associating with each particular bus a label that conveys three things:
1. The bus number of a labeled bus to which the bus being labeled is connected,
starting with the source bus, which is automatically (S, +, 00). The S identifies
the bus; the + indicates that flow into the bus can be increased; and the 00
indicates that the amount of flow available to bus S is infinite.
2. The flow available to the bus being labeled. The label contains a + if the flow
can be increased (as in the case of the source bus) and a - if the flow can
only be decreased.
3. The amount the flow into the bus being labeled can be either increased or
dccreas~d. Clearly, this amount depends on the amount of flow available at the
labeled bus and the flow-capacity relationship of the element connecting the
two buses. Two cases can occur:
a. If the flow through the element connecting the labeled and unlabeled
buses being considered, say element m, flows from the unlabeled bus j to
248 Power System Planning

the labeled bus i indicated by lm(j, i), then the amount the flow can be
decreased in element m, Alm, is

Alm = min (I;, Cm + lm(j, i)) (6.44)

I; is the flow available at bus i.


b. If the flow in element m, lm{i, j) is from bus i to bus i, the amount that
the flow can be increased in element m, t:..lm, is

t:..lm = min (I;,Cm - lm(i,j)) (6.45)

Given the rules for labeling a bus, the procedure for labeling a graph so as to
identify an augmentation path is as follows:

I. label bus S by (S, +, 00) . S is now labeled and unscanned (unscanned means
we have yet to extract any flow from the bus) and all other buses are
unlabeled and unscanned.
2. Select any labeled and unscanned bus i (initially S), then

a. For a11 unlabeled and unscanned buses j connected to i through element m


in which lm(j, i) > 0, label bus j by (i, -, t:..lm) where t:..lm is given by Eq.
6.44. Bus j is now labeled but unscanned.
b. For all unlabeled and unscanned buses j connected to i through element m
in which lm(i, j) > 0, label bus j by (i, +, Alm) where t:..lm is given by Eq.
6.45. Bus j is now labeled but unscanned.
c. Circle the index of bus i to indicate that it has been scanned.

3. Repeat step 2 until bus L is labeled or until no more labels can be assigned.

Before going any further, let us consider an example to illustrate the simplicity
of the procedure. Consider labeling the graph corresponding to the three-bus system
depicted in Fig. 6.6a, where the feasible flows are initially all zero and the source bus
is labeled (S, +, 00) , as shown in Fig. 6.6b. Bus I will be labeled (S, +, 2) since it is
connected to bus S (hence the S) and the flow through element I, 11 , can be increased
(hence the +) by an amount

A/ 1 = min (Is, c 1 - It)


= min ( 00 , 2.0 - 0.0)
= 2.0 p.u.
Similarly, bus 2 will be labeled (S, +, 2.0) sii;ice it is connected to bus Sand the flow
can be increased by an amount
D./2 = min (Is, C2 - 12)
= min ( 00 , 2.0)
= 2.0 p.u.
Transmission System Reliability Analysis 249

C5 = 4.0
' = 0.0
c. = 3.0
s ' = 0.0 3 L
m =1

c3 = 2.0
/3 = 0.0

(al

(S, +, 2l

(1,+,2)
(3, +, 2)
3 L

(S, +, 2l

(bl

Fig. 6.6 Unlabeled (a) and labeled (b) graph.

Following the labeling of bus 2 and bus 3, bus L will be labeled. Since the last bus
labeled was the demand bus, the algorithm terminates, producing the labeled graph
shown in Fig. 6.6b. Stated differently, the labeling process terminates because an
augmentation path has been identified. Circling the + signs in the labels for buses S, 1,
and 3 to indicate that these buses have been scanned in establishing a label for an
adjacent bus, we identify an augmentation path.
With an augmentation path established, the augmentation algorithm is employed
to augment the flow from bus S to bus 3 or to the dummy bus L, as discussed below.

6.3.5 The Augmentation Algorithm

Obviously from our previous discussion, the augmentation algorithm is a methodical


procedure using the labeled graph for actually augmenting the flow. The process is
started with the fictitious common demand bus L, and the logic for the algorithm is
250 Power System Planning

I. If the label on the bus is (i, +, D.lm ), then increase the flow through element
m by an amount D.lm. However, if the label is (i, -, D.lm), then decrease the
flow on element m by an amount D.lm.
2. If the next labeled bus considered is the source bus S, erase all labels and
return to the labeling algorithm. The feasible flow pattern used in the l:ibeling
algorithm will consist of the original plus augmented flows. However, if the
next labeled bus being considered is i -4= S, return to step 1.
For demonstration purposes, consider augmenting the flow through the labeled
network of Fig. 6.6b in accordance with these steps. Starting with the demand bus, we
increase the flow in elements 6, 3, and I by 2.0 p.u., 2.0 p.u., and 2.0 p.u.,
respectively, producing the new feasible flow pattern shown in Fig. 6.7a. Because the
label on bus I indicates that its source of flow was bus S, the augmentation procedure
terminates after flow is increased on element I. At this point the network and feasible
flow patterns shown in Fig. 6.7a are used in labeling the graph again. Remember, this
labeling/augmentation procedure only stops when labeling cannot proceed.

' == 2.0

3 L

(a)

(S, +, 2)
2

s (3, +, 2)

(b)

Fig. 6.7 (a) First augmented flow pattern. (b) Subsequent labeled graph.
Transmission System Reliability Analysis 251

s 3 L

,. = 2.0

(a)

' = 4.0
3 L

(b)

Fig. 6.8 (a) Second augmented flow pattern. (b) Total graph flow pattern.

Reapplying the labeling algorithm, we obtain the labeled graph shown in Fig.
6.7b. The reader is encouraged to verify the labeling using the step-by-step procedure
presented earlier. Proceeding to the augmentation algorithm with the labeled graph in
Fig. 6.7b, we obtain the augmented flow pattern given in Fig. 6.8a.
For the last and final time, the labeling algorithm, when reapplied to the graph,
and associated flow pattern in Fig. 6.8b, produces the labeled network in Fig. 6.9. The
elements linking labeled and unlabeled buses, as stated in Sec. 6.3.3, are the
minimum-cut elements. In this case, elements 1 and 2 are the minimum-cut elements,
and the capacity of the minimum cut is c 1 + c 2 = 4.0 p.u. No other cut will have a
capacity any smaller than c 1 + c2 , and based on Eq. 6.43 the maximum flow through
the graph is c 1 + c 2 = 4.0 p.u.

'
6.4 SUBSET DECOMPOSITION FOR SYSTEM
LOLP CALCULATIONS
In Sec. 6.2 it was made clear that one way of calculating the LOLP is to use Eq. 6.33.
Also made clear, however, was the fact that the inefficiency of calculating LOLP with

L ~~---~---- -
252 Power System Planning

schemes that require knowledge of each system capacity state Xi EB is intolerable.


Even for the three-bus problem the result of the first state-classification step produced
28 unacceptable states; hence, using Eq. 6.33 to calculate the intermediate LOlP for
this case would require 27 addition and 135 multiplication operations. Clearly, the
LOLP must be calculated using properties associated not with each XiEB, but with
subsets of B. Thus our interest now is to discuss a decomposition procedure that will
enable us to decompose B into nonoverlapping subsets (8). The requirement that the
subsets be nonoverlapping facilitates calculating the probability of existence of each
subset Bi in that

f(B) = lf(Bj) = LOLP (6.46)


i
if the nonoverlapping condition is forced. Stated in a more mathematical fashion, the
subsets Bi should be mutually exclusive such that no state Xi can satisfy the properties
of more than one subset. Diagrammatically, we wish to create a set structure as shown
in Fig. 6.10.
The decomposition algorithm to be used to create the subsets of B is

Xi= (V1, V2, ... , VE)EBi (6.47)

if Vj < Vi 0
for all m <i
Simply stated, this decomposition rule groups in a subset all those states that are
unacceptable because of the outage of a given element Vi. The condition that
Vm ;;. Vmo for all m < j ensures us that Xi is not unacceptable because of any elements
m<j.

s
3 L

Fig. 6.9 Final labeled graph showing minimum cut.


Transmission System Reliability Analysis 253

Bj-1

Fig. 6.10 Venn diagram showing mutually exclusive


subsetsBj

Since the subset Bi can occur only if Vi=0 (line j out) and if Vm;;;,, Vmo,
m <j, the probability f(Bj) is obtained as follows:

f(Vi = 0) = Q; (6.48)
and
f(Vm;;;,, Vmo) = 1.0 ifVmo=0
f(Vm;;;,, Vmo) = Pm if Vmo = 1.0
Therefore

f(Bj) = Qj IT f(Vm ;;;,, Vmo)


m<j
(6.49)

The implementation of Eq. 6.49 is trivial since Qj is known and Vmo is known, where
V mo is the mth component of X 0
Decomposing the 28 unacceptable states for the sample system resulting from the
initial classification step, we get the following subsets:

-[O,l,~,l,1]
B1 - .. total 16 points
9,0,o,o,o
l,0,1,l,l]
B2 = : total 8 pomts
[
1,0,0,0,0
254 Power System Planning

-[I, 1, ~, 1,0]
Bs - total 4 points

1, 1,0,0,0
For this case we should recall that
X0 =(1,l,O,O,l)=(V10, V20, V30, V40, Vso)
Using Eq. 6.49, we find that

f(Bi) = Q1
f(B2) = Q2f (V1 > V10)
= Q2f (V1 > 1.0)
= Q2P1
and

f(Bs) = Qsf(V1 > V1o)f(V2 > V20)f(V3 > V3o)f(V4 > V40)
= Qsf (V1 > l.0)f(V2 > l.O)f(V3 > 0.0)f(V4 > 0.0)
= QsP1P2(P3 + Q3)(p4 + Q4)
= QsP1P2
Therefore, the intermediate LOLP, using Eq. 6.46, becomes

LOLP = f(B1) + f(B2) + f(Bs)


= Q1 + P1Q2 + P1P2Qs
This value of LOLP is intermediate because subset Uc/= O; hence there can be
unacceptable states X;EU which will increase the LOLP.
The simplification resulting from the decomposition is evident even for the
simple example under consideration. By not settling for complete enumeration, we
have reduced the number of additions necessary to calculate the LOLP from 27 to 3
and the number of multiplications from 135 to 3.

6.5 SUBSET DECOMPOSITION FOR ELEMENT LOLP


AND e(DNS) CALCULATIONS
We might think that the decomposition rule discussed above should be sufficient to
calculate all the reliability indices desired, but as it turns out, the subsets Bi were not
constructed based on flow properties of the, states ~EB;. It is necessary to decompose
the subsets B; into new subsets in which the states within a given subset have the same
minimum cut; that is, the elements defining the minimum cut are the same for each
state ~ in a given subset. With this information the system c(DNS) and the LOLPm
and Em(DNS) for each element m = I, 2, ... , E, can be determined without resorting
to complete state enumeration.
Transminion System Reliability Analysis 255

The relationship between the (DNS) and the capacity of a minimum cut should
be emphasized before proceeding. As we know, for an unacceptable system state ~
the capacity of the minimum cut C(M) is equal to the maximum flow gi through the
graph, where L - gi is the demand not served as in the expression for (DNS) in Eq.
6.34. Thus we see that the maximum-flow/rninimum-<.:ut algorithm is crucial to the
evaluation of i:(DNS).

6.5.1 Expressions for (DNS), LOLPm and m(DNS)

If we decompose the B/s into subsets Bn in which each state Xi in the subset Bn has
the same minimum cut, and hence the same maximum flow, then starting with Eq.
6.34, we have

(DNS) = l
XiEBj
f(Xi)(L - gi)

=l lBj XiEBj
f(Xi)(L - gi)

=l l f(XiXL -- gi)
Bn XiEBn

= l l
L f(Xi) - l l f(Xi)gi
Bn Xj EBn Bn Xj EBn

= Lf(B)- l l f(Xi)gi
Bn XjEBn

Assuming f(U) = 0, that is, subset U is exhausted, from Eq. 6.38


f(B) = f(B) = I - f(A)
Therefore

(DNS) = L -(f(A)L +? l
Bn XjEl3n
A f(Xi)g,) (6.50)

The expression in parentheses is the expected value of the demand that can be
served, while the first term f(A )L is the expected value of the total demand served and
<
the second term is the expected value of the partial demand served (PDS). To
evaluate the e(PDS) using the expression given is inconvenient because, once again, it
involves almost complete enumeration of the states. Since gi is the maximum flow
through the network as the system resides in state Xi, which is the sum of the
capacities of the minimum-cut elements, the (PDS) may also be written as
256 Power System Planning

e (PDS) = _l _l f(Xi) _l Cm(Xi)


BnXiEBn mEM

= _l _l _l f(Xi)Cm(Xi) (6.Sla)
Bn mEMxiEBn
But using Eq. 6.32, we find that

_l f(Xi)Cm(Xi) = _l fiCVm)Cm(Xi) fl
ji:-m
fi(V;) (6.Slb)
XiEBn XiEBn
This last expression is nothing more than the expected value of the capacity of the
minimum-cut element Vm, where J;(Vm) IT;,;:.m/i(V;) is the probability that the system
resides in a capacity state XiEBn. Further, with the recognition that

(6.Slc)

then

Vm (6.Sld)
l
Vm=Ym
f(Vm)

Using Eqs. 6.51b and 6.Slc, we have

(6.Sle)

Substituting Eq. 6.Sld into Eq. 6.Sle,


Vm
l f(XDCm(Xi) = l Vm
XjEBn Vm=!:'.'m
l
Vm=!:'.'m
f(Vm)
Transmission System Reliability Analysis 257

Finally, substituting this result in Eq. 6.51a, we get

(PDS) = LL Cmf(~n)
Bn mEMf(Vm(Bn))
(6.52)

The E(DNS) expression most useful for solving actual problems becomes:

(DNS) = L(l - f(A)) - l l


A EM
Cmf(~n)
f(V m(Bn))
(6.53)
Bnm

where

Cm(Vm)f(Vm) = expected value of the capacity of the


minimum-cut element m

and

f(V m(B n )) = probability that the minimum-cut element


A

resides in a state V m E Bn

To obtain the element LOLPs, we need to determine the probability that each
element m is a minimum-cut element. Stated more formally

LOLP m = .l f(M) M = minimum cuts containing element m (6.54a)

where

f(M) = l f(Bn) n = subsets having m as a minimum-cut element (6.45b)

Equation 6.54a replaces Eq. 6.36 as a more convenient expression to use for
calculating LOLPm- Although enumeration of all subsets En is necessary, there are far
fewer En subsets than states XiEB. The element LOLPs can be used to identify weak
areas in a transmission system in that a high LOLP for a given element indicates that it
is somehow involved in not being able to satisfy the demand. More will be said of this
in a concluding section (Sec. 6.6.5).
In addition to LOLP m, we can associate with each element the expected value of
demand not served, Em (DNS), because of the element's being a minimum-cut element.
Using Eq. 6.53 ,as a guide, we should recognize that the quantity [I -f(A)] is the
probability that the system cannot supply a load of L MW, which in the case of an
individual element is its LOLPm. Further, the second term of Eq. 6.53 is the expected
value of the partial demand that can be served, which as we know depends on the
capacities of the minimum-cut elements. The expected value of the partial demand
served, in which a given element m was involved, is simply the summation of those
terms in whi~h mEM. Mathematically, we have

L
258 Power System Planning

\' \' cJ(Bn)


E,n(DNS) = LOLPmL - L L f(V(B )) (6.55)
BniEM z n

where Bn are those subsets in which element mEM.

6.5.2 Decomposition Rule for Obtaining Bn


The ~philosophy behind the decomposition of subsets Bj into subsets En in which each
X;EBn has the same minimum cut is that an upper critical state x0 (B;) exists that
defines the boundary between states in B; with the same minimum cut as the upper
limiting state X{B;). X 0 (B;) is defined by establishing a maximum-flow pattern through the
network as it resides in state X{B;); then for all mEM
0
Vm (Bj) = fm(Bj) {6.56)
and for all mrf.M
Vm 0 (Bj)=fm(Bj) iflm=O
Vm 0 (Bj)=l iflm>O
In effect, Eq. 6.56 states that any flow-carrying elements not part of the minimum cut
will have their critical capacities set to the maximum value as defined by X(B;), and
any element not carrying flow and all minimum-cut ele'. :1ents will have their critical
capacity values set as low as possible, as defined by the lower limiting capacity state
~(B;)- Defining X0 (B;) in this way insures that any states X; whose capacities equal or
exceed those defined by X0 {B;) will have the same minimum cut M:
(6.57)
if
Vm;;;,,vm 0 (Bj) m=l, ... ,E
For example, consider the subset B 1 as determined for the sample three-bus system.
We have

X(B 1) = (0, 1, 1, 1, 1)
~(Bi)= {0,0,0,0,0)
Obtaining the maximum-flow pattern with the system residing in state X(B 1 ) by
use of algorithms discussed earlier, we get the labeled network and associated flows
shown in Fig. 6.11, a and b. Based on Eq. 6.56, we find
0
Vi (Bi) = 0 }
minimum-cut elements
V2 (B2) =0
V/(Bi)=O}
zero flow elements
V4 (Bi) = 0
V5 (Bi) = 1.0} nonzero flow element

-. ,.c.'.X.
Transmission System Reliability Analysis 259

s \ m=6
I m=4
3 L
I
I
I
/ M12
1
(a)

~ ' = 2.0
s 3 L

(b)

Fig. 6.11 (a) Final labeled graph. (b) Final maximum-flow pattern.

Therefore
x 0 (Bi) = (0, 0, o, 0, 1)
and the eight states Xi for which Eq. 6.57 is satisfied are

A {O,l,l,1,1
E1 = 8 states
o,o,o,o, 1
These eight states have the same minimum cut involving elements 1 and 2, denoted by
Ml2, with a combined capacity of 2.0 p.u., the value of the maximum flow that can
be transmitted through the network.
The probability that En exists is easily determined using Eq. 6.57:

f(Bn) = flt(Vm;;;. Vm 0 (Bj)) (6.58)


m

J
~
260 Power System Planning

For our sample problem

As a result of decomposing set Bi, we shall always have a portion of the subset
that does not satisfy the conditions of En because of the classification rule originally
used to create the subsets Bi. The rules did not ensure that all states XiEBi had the
same minimum cut. For instance, in decomposing B 1 above we found that only half of
the states satisfied the condition for being in .B 1 The remaining states, XiE(B 1 -Bi),
must be further decomposed using the associated limiting capacity states, etc. In the
sample case

0, 1, I, 1,0
XiE(B1 -Bi) : 8 states

fo,o,o,o,o
To facilitate further decomposition of these states into subsets Bn, the states are
arranged into temporary subsets within B 1 In general, the states remaining in Bi after
each decomposition step will be temporarily placed in subsets BiP, in which these
states must satisfy the condition that
Xi= (V1, V2, ... , VE)EBip
if
(6.59)
and

m<p
If additional decomposition of the subsets BiP is necessary, the newly created subsets
will be denoted by Bjpq, where those XiEBjpq will be determined using Eq. 6.59, with
the subscripts denoting the generation of subsets being considered. For third generation
subsets Eq. 6.59 becomes
Xi= (V1, V2, ... , VE)EBjpq
if
(6.60)
and
m<q
As the decomposition process continues, the decomposition rule should be altered to
reflect the subset being decomposed through the use of subscripts.

' -~-" '<\f.f;,:l -


Transmission System Reliability Analysis 261

To illustrate how this process evolves, consider the further decomposition of the
remaining states in B 1 Apply Eq. 6.59, recalling that
:x(Bi) = (0, 0, 0, 0, 1)
We find only the subset

0,1,1_,1,o
Bis= 8 states
{
o,o,o,o,o
With a maximum feasible flow established with the network residing in the upper
limiting state of Bis
X(B 15 ) = (0, 1, 1, 1,0)
we would find that the upper critical capacity state
x 0 (B1s) = (0,0, 1, 1,0)
and the minimum cut Ml 2. Hence, the two states

X = ~ (0, 1, 1, 1,0)
1
l (0,0, 1, 1,0)
define the minimum-cut subset B15 having a minimum cut Ml 2. This leaves six states
to be further decomposed using Eq. 6.60 and x0 (B 15 ) just determined. We see that
Xi= (V1, V2, ... , VE)EB1s4
if

and if

And

if

and if

V m ;:;,, V m O (B 1s) for m <3
Therefore
o, 1, 1,0,0
X1 E B 154 ={ 2 states
0 ' 0 ' 1' 0 ,- 0

l --
262 Power System Planning

Although Eq. 6.60 did not specifically indicate to us that the fifth element, Vs, of
XiEB1s 4 could assume only the value of zero, the limiting capacity states for subset
Bis, namely X(B 15 ) and ~(B 1s ), tell us that element Vs = 0. If we allow element
V5 = 1, then these states fall in B1 , but we know that all XiEB 1 already have been
classified.

X (8 1 ) == (0, 1, 1, 1, 1)
X (8 1 I == (0, 0, 0, 0, 0)
0
X (8 1 ) == (0, 0, 0, 0, 1)

16 States

X (8 1 ) == (0, 1, 1, 1, 1)
X (8 1 I== (0, 0, 0, 0, 1)

8 states

x I~ 15 I == 10, 1, 1. 1, o)
X (8 15 ) == (0, 0, 1, 1, 0)

X (8.,3) == (0, 1, 0, 1, 0)
X (8 153 l == (o, o, o, o, o)

X;;;,, X 0 (8 154 )

8 154 -M145

X (8154) == X (8, .. )
X (8 154 ) == X 181541

Fig.6.12 SubsetsofB,.
Transmission System Reliability Analysis 263

Similarly

0,l,~,l,0
Xi E B1s3 = 4 states
{
' o,o,o,o,o
If the maximum feasible flow pattern for each of these subsets is determined, we find
X 0 (B 154 )= (0,0, 1,0,0)
and
X 0 (B 15 3) = (0, 0, 0, 0, 0)
where the minimum cuts for these two subsets are M145 and M35, respectively.
To summarize the results of decomposing B 1 into subsets Bn in which every
XiEBn has the same minimum cut, the diagram depicted in Fig. 6.12 is provided.
Calculating the probability that the system state Xi resides in a given subset En is
easily done using Eq. 6.58. For instance

= f(V1 ~ V/(Bi))f(V2 ~ V2(B1))f(V3 ~ V3(B1)) ...


= q1Ps
Similarly

f(B1s) = q1p3p4qs
f(B153) = q1q3qs
f(B1s4) = q1p3q4qs

6.6 RELIABILITY CALCULATIONS FOR SINGLE AREAS


The approach to be used in evaluating the reliability for single areas in terms of LOLP
and E(DNS) will be summarized as well as demonstrated by completing the analysis of
the sample system. In addition, comments will be made concerning the use of the
indices of reliability to provide insight into measures required to improve the reliability
level.
The general problem of evaluating the reliability of interconnected systems is
postponed until Sec. 6.7.

6.6.1 Solution Ptocedure Revisited


Although the basic approach to obtaining the indices of reliability was outlined earlier,
enough detail has since been discussed to warrant a look back. Given the system
structure, element capacities, element FORs, and bus demands, the LOLP and the
E(DNS) for a system and for each element are evaluated as follows:

l
264 Power System Planning

1. Determine X and ~ for the state space being considered in which initially all
Xi EU. Clearly, X = (1, 1, ... , 1), and~= (0, 0, ... , 0).
2. Determine x 0 and X0 using the flow algorithms in conjunction with Eqs.
6.40 and 6.41.
3. Classify as many Xi as possible as being in A(Xi > x 0 ), B(Xi < X0 ), or
U(Xo .,;; Xi < X0 ).
4. Calculate f(lf) using Eq. 6.39. If f(lf) is considered sufficiently small or
f(lf) = 0, continue to step 5; otherwise decompose U into subsets using Eq. 6.42,
and repeat steps 1 to 4 for each subset Ui.
5. Decompose B into subsets Bi using Eq. 6.47.
6. Decompose, using Eqs. 6.56 and 6.57, the subsets Bi into subsets Bn in which
each state XiEBn has the same minimum cut M.
7. Calculate LOLP and (DNS), LOLPm, and m(DNS), using Eqs. 6.46, 6.53,
and 6.55, respectively.

6.6.2 Completion o-f Sample System Analysis


We initially found that for the sample system

X= (1,1,1,1,1)
~= (0,0,0,0,0)
x0 =(l,1,1,0,1)
X0 = (l,l,0,0,1)
Using the definition of subsets A, B, and U, we were able to classify the 32
states shown in Fig. 6.13. From Fig. 6.13 and Eq. 6.46 the intermediate LOLP (before
U is completely exhausted) is:
5

LOLP = I
n=I
f(Bn)

= I n qn f(Vj;;;. Vjo)
n c.= 1 j 'F n

We must of course realize that since we still have some unclassified states, this
LOLP is but an intermediate value. If we arbitrarily assume the FORs q 1 = 0.02,
q 2 = 0.03, q 3 = 0.2, q4 = 0.1, and q 5 = 0.3, then LOLP = 0.3345. Also from Fig.
6.13 we have that
f(A) = Pt P2P3Ps
= 0.5323
and
f(U) = 0.1330
Transmission System Reliability Analysis 265

(1, 1, 1, 1, 1)

, (0, 0, 0, 0, 0)
2 States
82

B,
X (B2) = (1, 0, 1, 1, 1)
X (B 2 ) = (1, 0, 0, 0, 0)
X {B,) = (1, 1, 1, 1, 0)
X (B 5 ) = (1, 1, 0, 0, 0)

8 States
4 States

X (B 1 ) = (0, 1, 1, 1, 1)
X (B 1 ) = (0, 0, 0, 0, 0)

16 States

Fig. 6.13 Venn diagram for sample three-bus system.

Obviously, fCU) is not particularly small in this case, so we must repeat the process
above and classify the states XiEU3 , given the new limiting capacity state

X=(I,1,0,1,1)

and
~ = (1, 1,0,0, 1)
The new critical capacity states x 0 and Xo, obtained from reapplying the rules defined
earlier together with Eqs. 6.40 and 6.41, are

x 0 = (1,1,0,1,1)
E U3
X0 = (l,1,0,1,1)
Based on the state classification rule, we see that
Xi= (l,l,0,1,l)EA
266 Power System Planning

and
Xi= (l,l,0,0,l)EB 4
As a result of classifying these last two states, the subset U has been exhausted, or
stated more precisely, we have iteratively adjusted the critical capacity states until
X = X0 , for which U = 0 as shown in Fig. 6.14.
With the subset U exhausted, we can calculate the final LOLP:
LOLP = 0.3345 + P1P 2q3q4p 5 = 0.3478
In order to determine the element LOLPm and the Em(DNS), it is necessary to
decompose B 1 , B 2 , B 4 , and B 5 further, as indicated in step 6 in Sec. 6.6.1. As we
have done so to some extent in Sec. 6.5, we have only to complete the procedure. We
found, using the decomposition rule given by Eq. 6.57, that the subset B 1 decomposed
into subsets B1, B1s, B1s3, and B1 54 , as shown in Fig. 6.12. We must now decompose
B2, B4, and Bs.
Decomposing B 4 into En is trivial since B 4 contains only one state. The
maximum-flow algorithm with the network typology defined by X(B 4 ) in conjunction
with Eq. 6.56 gives
x 0 (B 4 ) = (I, I, 0, 0, I)

X (B.) = (1, 1, 1, 1, 0)
X (8 5 ) = (1, 1,0, o, 0I
4 States

8 States

X (B 1 I = (0, 1, 1, 1, 11
X (B 1 l = (0, 0, 0, 0, Ol

Fig. 6.14 Final Venn diagram with all states classified (U = 0).
Transmission System Reliability Analysis 267

x0 (B 2 ) = (1,0, 1,0,0)

x IB2,. l = X IB2, l
= (1,0,0,0,0)

Fig. 6.15 Decomposition of B 2 into subsets En.

as the critical capacity state of subset B 4 By use of the classification rule defined in
Eq. 6.57, the state XiEB 4 will decompose into subset B4 whose minimum cut is M234,
which has a ca?acity of 2.0 p.u.
Proceeding in the same manner but considering subsets B 2 and B 5 , the
decomposition process evolves as depicted in Figs. 6.15 and 6.16.
To calculate the probability f(M) that a given minimum cut exists, we need only
use Eqs. 6.54b and 6.58. For instance, in determining the probability that the
minimum cut Ml2 will occur, we have from Figs. 6.12, 6.15, and 6.16.

J
268 Power System Planning

The two terms involving Q1Qs and q2q 3 appear because subsets B15 and B23 also have
Ml2 as their minimum cut. Again, from Fig. 6.12, 6.15, and 6.16,

and

f(Ml45) = f(B1s4) = q1q5q4 fl f(Vm;;;,, Vm


m"1-1,4,5
0
(B154))

f(M234) = lt(B234) = q2q3q4


t n
ma"2,3,4
f(Vm;;;,, vm 0 (8234))7
J
+ lt(B4) = q4 fl f(Vm;;;,, Vm 0 (B4))7
L'. m*-4 :J

X (B,) = (1, 1, 1, 1, 0)
K (B,) = (1, 1,0, 0, 0)
4 States

x 1s,1 == 11. 1, 1. 1, 01
K IB,l == 11. 1. o. o. 01 Fig. 6.16 Decomposition of B 5
into subsets Bn.
4 States
M35

----- -----~
Transmission System Reliability Analysis 269

As all FORs are given for the sample problem and all critical capacity states x0
have been calculated, numerically we have

f(M12) = 0.04554 f(M145) = 0.00048


f(M35) = 0.28796 f(M234) = 0.01389

To check the validity of the decomposition, we know that

LOLP = Lf(B;)
i
= Lf(Bn) =
n
l
all M
f(M)

A quick check shows that we have indeed managed the decomposition properly.
For future use, the complete decomposition of B into B is shown in Fig. 6.17,
with all intermediate steps omitted for clarity.

6.6.3 Element LOLPs


Finally we are in a position to evaluate the LOLPs for each element as defined by Eqs.
6.54a and 6.54b; for example
LOLP1 = f(M12) + f(M145)
= 0.04602
LOLP 2 = f(M12) + f(M234)
= 0.05944
LOLP3 = f(M35) + f(M234)
= 0.30186
LOLP4 = f(M145) + f(M234)
= 0.01437
LOLP5 = f(M35) + f(M145)
= 0.28844
From this information we easily see that the high LOLP is due primarily to
elements 3 and 5. This does not mean that if we added capacity to elements 3 and 5
we would be assured of a significant improvement in reliability. What this information
tells us is simply that the LOLP is being adversely affected by the outage
characteristics of these elements. Some readers may question the value of performing so
the LOLP, such expenditures would best be directed toward improving the outage
characteristics of these elements. Some readers question the value of performing so
much detailed analysis merely to conclude that elements 3 and 5 are crucial elements,
which, after all, is obvious from the FOR data, since q 3 = 0.2 and q 5 = 0.3. By way
of response, the FOR data alone are insufficient information to determine the role a
given element plays in contributing to the LOLP. For example, looking at FORs, a
planner might conclude that elements m = 3, 4, and 5 would require additional capital
expenditures to improve their outage performance and hence maintain an acceptable
level of reliability. However, based on detailed reliability calculations, we see that
270 Power System Planning

:x = 11. o, o, 1, 1 l
K == (1, o, o, 1, 1l

x == 11, 1,0,0, 1l
K= (1, 1,0,0, 1)

><=11,1,1,1,ol X=(1,1,o,o,1)
K= (1, 1, o, o, ol K = (1, 1, o,
o, 1 l

Fig. 6.17 Summary of subset decomposition.

element m = 4 is not at all crucial; before money is spent to reduce its FOR, some
consideration should be given to elements m = I and 2. Herein lies the value of
reliability analysis: it provides the planner with additional insight of both a
quantitative and a qualitative nature, which can be used to guide the planner in
establishing the need for future detailed studies.
Transmission System Reliability Analysis 271

6.6.4 Calculation of (DNS}


Using Eq. 6.53 and the results of the decomposition as depicted in Fig. 6.17, we get

\" \" Cmf(Bn)


E(DNS),= 4(1.0 - 0.653) - ~ L f(Vm(Bn))
Bn mEM

The summation term contains ten parts, one for each subset Bn. For subset B1 the
summation gives

Cif(B1) C2f(B,)
f(V,(B,)) + f(V2(B1)) = C2P2Q1Ps
where
0

C, = l
v,=o
C 1(Vi)f(Vi) = 0

C2 = l
V 2 =0
C2(V2)f(V2) = 2.0f(V2 = 1) = 2.0p 2

f(V2(B1)) = P2 + Q2 = 1.0
The remaining nine parts produce the following results:

Bn = B,s ~ C2P2Q,p3p4q5
= B1s3 ~ o
= B,s4 ~ o
= B2 ~ C1P1Q2P3
= B23 ~ C1P1Q2Q3p4p5
= B234 ~ o
= B23s ~ o
= S4 ~ C2P2P1 Q3Q4p5
= Bs ~ C3p3p1p2Qs
Upon substituting the numerical values for the capacities and FORs assumed for
the three-bus system into the expression for e(DNS), we find that
e (DNS),= 0.81865 p.u.

or if 25 MVA is used as a base so that L = 100 MW, then


E (DNS) = 20.466 MW
272 Power System Planning

To describe completely and finally the reliability of the system, we can also
associate with each element Em (DNS), which, as discussed earlier, is a quantity that
indicates the degree to which a given element is responsible for the system being
unable to serve the load. Again, this information provides the planner with more data
on which to make future study decisions. Calculating the Em(DNS) for each of the
elements in our sample system using Eq. 6.55 and the results obtained earlier, we get

E1 (DNS) = L(LOLPi) - ll Cmf(~n)


mEMf(Vm(Bn))
t
Bn = B1,B1s,B2,B23,B154
= L(LOLP1) - C2P2Q1Ps - C2P2Q1P3P4Qs - C1P1Q2P3
- C1 P1 Q2 q3p4p5 - 0
= 0.09410 p.u. (or 2.3527 MW)
Similarly

2 (DNS) = L(LOLP2) - C2P2Q1Ps - C2P2Q1P3P4Qs - C1P1Q2P3 - C1P1Q2Q3p4p5


- 0 -c2P1P2Q3Q4Ps
= 0.121156 (3.0289 MW)
3 (DNS) ~ 0.72455 (18.1137 MW)
E4 (DNS) = 0.03088 (0.77222 MW)
Es (DNS) = 0.69750 (17.4375 MW)

6.6.5 Interpretation of Results

Because of the amount of information we now have, it is convenient to display this


information on the network, as shown in Fig. 6.18, along with the outage
characteristics for each element. The intent here is to explore the meaning of the
reliability indices defined as currently used by some utilities. As we shall see, proper
interpretation of the indices is essential.
From Fig. 6.18 we may conclude that elements 3 and 5 appear to be in some
sense responsible for the system's inability to serve the load during contingency
situations. If additional studies were considered to insure a good design, we would
closely scrutinize the performance and characteristics of these elements. We should not,
based on the information we have, make irrevocable decisions concerning additional
capacity. It is tempting to look at Fig. 6.18 and conclude that to reduce the E(DNS)
we should add capacity to elements 3 and 5, because these elements have very large
Em(DNS) indices. In fact, adding capacity to element 5 will do absolutely nothing to
reduce system E(DNS). '
Because proper interpretation of the indices developed is crucial, we should
explore the meaning of the indices using element 5 as an example. From Fig. 6.18 we
note that Es (DNS) = 0.697 p.u.; yet if we write out the expression for the E(DNS),
from Sec. 6.5
Transmission System Reliability Analysis 273

c1 = 2.0 p.u. = 2.0


c2
q,= 0.02 q, = 0.03
LOLP 1 = 0.046 LOLP 2 = 0.059
e 1 (DNS) = 0.094 e2 (DNS) = 0.121

4(DNS) = 0.031
LOLP 4 = 0.014
c4 = 3.0 p.u.
q. = 0.1

c5 = 4.0 p.u.
q, = 0.3
LOLP 3 = 0.301
e 3 (DNS) = 0.724

c3 = 2.0 p.u.
q, = 0.2

e (DNS) = 0.818 L = 4.0 p.u.


LOLP = 0.347
Fig. 6.18 Reliability data for three-bus system.

(DNS) = L(q1 + P1q2 + P1P2q3q4ps + P1P2qs) - .. ~-,2P3 + q2q3p4ps)


-c2P2(q1Ps + P1q3q4p5 + q1p3p4qs)-c3p3(p1p2qs)

we see that c 5 does not even appear in the expression, and neither does c4 The reason
c5 does not appear is that element 5 was only EM when its capacity c 5 = 0. Figure
6.17 shows this to be the case; that is, the upper and lower limiting states for each
subset En in which element 5 was a minimum-cut element all have V5 = fs = 0. For
instance, subset B154 is a subset in which element 5 is a minimum-cut element and
t\ (B 1s4) = fs (B 154 ) = 0. In effect, element 5 is a minimum-cut element only when it
is out of service. Hence, adding capacity to element 5 cannot improve the reliability.
Stated more simply, if

. _3_,(_DN_S-"-) = 0
acs

adding capacity to element 5 will not reduce the 1;(DNS).


274 Power System Planning

Continuing with our example, we see from the expression for c(DNS) that
ac(DNS)
:1. = L(p1P2 - P1P2q3q4) + C1P1(q2q3p4)- C2P2(q1p3p4 -qi -- Pi q3q4)
u(/5
- C3p3(p1P2)
This result, with our earlier findings, suggests that to reduce the c(DNS) it is more
prudent to improve the element's reliability than to add capacity. Because the
demonstration is instructive, the values of

ac(DNS)
m
= l ' ... , 5
aCm
ac}DNS)m = l, ... ' 5
qm

for several elements of the three-bus system are given below:

ac(DNS)
aq = L(l - q2 - P2q3q4p5 - P2qs) + C1(q2p3 + q2q3p4p5)
1
- C2P2(p5 + p3p4q5 -q3q4p5) + C3p3p2q5
= 1.43

ac(DNS)
aq = L(p1P2q4Ps) - c1P1(q2P4Ps - q2) + c2P2(q1p4qs -pi q4ps) + c3P1P2qs
3
= 0.7355
ac(DNS) =O
acs
ac(DNS)
a q5
= given above = 2 .243
So far it is obvious that information more useful than LOLPm and Em(DNS)
alone can be derived from the system c(DNS). Before specific action is taken to alter
an expansion plan under consideration, a thorough understanding of the results is
necessary. For our sample system we note that the information given in Fig. 6.18 is
helpful in identifying elements that are somehow responsible for the existing reliability
level. Under no circumstances is this information sufficient for making specific
improvements. Rather we have found that sensithity data can be determined which do
provide insight into the specific action that must be taken to improve reliability-that
is, whether improvements in element outage performance are required or whether
simply adding more capacity will suffice.
In addition to information available in the c(DNS), the LOLP for the system

"'*-
~-~
Transmission System Reliability Analysis 275

contains much useful information about the relationship between elements. From Sec.
6.6.2 we found that

LOLP =qi+ P1q2 + P1P2qs + P1P2q3q4p5


The sensitivities relating the changes in LOLP to changes in element outage
performance are ,

3LOLP
" = Pi -piqs -p1q3q4p5 = 0.6722
uq2
3LOLP .
" = P1P2q4p5 = 0.0665
uq3
3LOLP
- -q-- = P1P2q3p5 = 0.13308
0 4
3LOLP
oqs
= P1P2 - P1P2q3q4 = 0.9315
Although each of these is important, we see again how crucial the outage performance
of element 5 is, relative to the other elements. We should take note of the expressions
for elemrnts 3 and 4: if the outage performance of one element can be improved
substantially, the other element is essentially unnecessary. This is obvious in our simple
network; for large networks such observations may be not at all obvious. It goes
without saying that these and earlier comments are valid only if network topology
;emains unchanged. Changing the load profile by moving the load from bus 3 to bus 2
could alter drastically any and all comments with respect to the sample system.
Another use for the sensitivity information, especially involving element forced
outages, is to determine the effect of knowing imprecisely the performance data of
individual elements. It should be emphasized that the expressions for E(DNS) and
LOLP will not change if element FORs are changed; therefore, evaluating the effect of
FORs or element availability on system performance is easy.

6.7 MULTIAREA RELIABILITY ANALYSIS


In most transmission planning studies it is often necessary to evaluate not just the
reliability of an isolated area but the reliability of interconnected areas in a given
operational district. We know from Chap. 3 that interconnected areas derive substantial
benefits in terms of improved reliability through interconnections to neighboring
utilities. Further, the capacity and outage characteristics of the interconnections greatly
influence the degree of improvement in reliability resulting from the interconnection.
In planning interconnection expansions, due consideration using multiarea analysis
must be given to the benefits of each expansion alternative.
We might think that the probabilistic methods discussed earlier can be directly
applied to the multiarea problem. Unfortunately not, but the ideas presented earlier

L
276 Power System Planning

can help us explore various aspects of multiarea problems. The primary reasons that
single-area probabilistic analysis, as we have discussed it, cannot be directly applied are
I. We prefer to emphasize not the topological details within a given area but the
details of interconnections between areas that influence the ability of a
multiarea system to transfer reserves within the system.
2. The single area load was assumed deterministic, but in multiarea studies the
benefits of interconnections depend on the reserve-margin states of each area
as these states randomly vary due to unit outages and random load chan?es.
Note that we are leading up to a method for multiarea reliability analysis similar to the
method discussed earlier for single-area problems. We are interested in determining
reliability indices for each area and for each interconnection.

6.7.1 Element Models and Reliability Indices for


Multiarea Reliability Analysis
Typically we are faced with the problem of evaluating a given interconnected system as
depicted in Fig. 6.19, where each area, j, can reside in any reserve-margin state Rj with
probability f(R;) and each interconnection can reside in any capacity state V; with
probability f(V;). The possible reserve-margin states where a given area may reside were
discussed in Chap. 3; that is,
(6.61)

Fig. 6.19 Multiarea system.


Transmission System Reliability Analysis 277

and
dF (Lei= lCi - Rj)
= {6.62)
dRi
We should recall that Lei is the effective load for area j and that F(Rj) is the
cumulative reserve-margin distribution obtained by evaluating the effective load
cumulative probability distribution at the point ICi - Rj, ICi being the installed
capacity of area j.
The element models for each interconnection are identical to element models
used in earlier discussions; that is, it will be assumed that the interconnections may
reside in independent discrete capacity states. For convenience, we consider only
two-state models for the interconnections, realizing full well that high-order models
may be used if desired. We will continue to use integers to designate element states.
Thus, interconnections can reside in states V; = 0, l, and areas can reside in margin
states ri = 0, l, ... , nj, where state 'i = 0 will correspond to the lowest negative
reserve-margin state and 'i = nj, the highest positive state.

6.7.2 System Model


Since the areas randomly assume reserve-margin states rj, j = l, ... , number of areas
A, at the same time that the interconnection elements randomly assume capacity states
Vm, m = l, ... , number of interconnections /, the interconnected system will reside
in state

where

i = 1, ... , 21 n(ni + l)
jEA
{6.63)

For the multiarea system depicted in Fig. 6.19, a total of 248,832 states exist if/= 5
and ni = 5. The difficulties of complete enumeration are still with us, and we must be
clever to circumvent these difficulties again.
As in our analysis of a single area, the objective is to evaluate the system states
and determine
1. System LOLP
2. Area LOLPis,j = 1, ... ,A
3. Interconnection LOLPms, m = l, ... ,I

6.7 .3 General Solution Procedure


Since there is a ver'y large number of states Xi, we must continue to consider solution
procedures that do not involve complete enumeration. In this vein it is beneficial first
to decompose the state space into subsets of unclassified states Un, each subset
corresponding to a particular combination of deficient and self-sufficient areas. A
self-sufficient area is one in which the reserve-margin states Ri;;,. 0, and a deficient area
278 Power System Planning

i, one in which Rk < 0. As the reserve-margin states of a given area can fall only into
two broad classes, self-sufficient or deficient, there must be 2A combinations of
self-sufficient or deficient areas. Considering each subset Uk sequentially, the states
XiEUk are examined and classified using a procedure that attempts to avoid complete
enumeration. The flow chart in Fig. 6.20 depicts the basic solution process [9]. Although

Choose a combination of
self-sufficient and deficient
areas

..
Form an initial state
x 0
by setting all Rj = 0 ,
and all cm= 0


Establish maximum flow -
pattern between areas '

Yes

All loads satisfied
~
.

Generate new state
Store state and
by reducing rj of
minimum cut
a deficient area

Generating new state by


1. Increasing rj
or 2. Increasing Cm
or 3. Decreasing rk

~
I No_ Form new
I All X; e Uk examined I ~

state
t
ICompute reliability information I
No
' I
I k = -iA k->k+1I-

{ Yes

Fig. 6.20 Flow chart of multiarea reliability analysis.


Transmission System Reliability Analysis 279

it appears from Fig. 6.20 that each state requires examination, we must realize that
many states may be classified by examining just one state, because all the states ~EUk
must reside in one of the following three subsets of Uk:
I. Subset A contains all acceptable states.
2. Subsets En contain all states Xi that are interconnection unacceptable because
of interconnection capacity constraints.
3. Subsets Bi contain all states Xi that are reserve-margin unacceptable because
of reserve-margin constraints in the self-sufficient areas.
By examining a state Xi= X0 EBn (using the maximum-flow algorithm as in Sec.
6.6) a small subset of states reachable from X0 can also be classified without
examination, since any state in which the reserve margins of the deficient areas or the
minimum-cut interconnection capacities are reduced from the values defined by Xo are
unacceptable. Similarly, states in which the reserve margins of self-sufficient areas are
increased above the values defined by X0 are unacceptable. Stated more rigorously, if
X0 EBn, then any state satisfying the conditions below is also unacceptable. Further,
these states have the same minimum cut M.
!Jc rk rko k = deficient areas
'i;;;,, 'i;;;,, 'io = self-sufficient areas
j
(6.64)
!:'.'m Vm Vmo mEM
Vm;;;,, Vm;;;,, Vmo mff.M
To calculate the desired probability information, we need only store Xo and the
conditions implied by Eq. 6.64, including the number of states satisfying Eq. 6.64. The
probability associated with the classified states is

f(Bn) = fl f(rk rko)flt(ri;;;,, 'io) fl f(Vm Vmo) fl f(V m ;;;,, Vmo) (6.65)
k j mEM m"fcM

Similarly, any examination state X0 EBj usually defines a subset of states ~ EBi,
since if X0 is reserve-margin unacceptable, any state in which the reserve margins of
either defi~ient or self-sufficient areas are reduced will be unacceptable. So will be
states in which the interconnection capacities are increased. For later, if X0 EBj, then
any state satisfying the conditions that

!Jc rk rko k = deficient areas


!.j 'i 'io j = self-sufficient areas (6.66)
Vm ;;;,, Vm ;;;,, Vmo m = all interconnections
will be unacceptable. The probability of subset Bi existing is given by

f(Bj) = nf(rk
k
rko) nt(ri rjo)fl f(Vm;;;,, Vmo)
j m
(6.67)

where, again, only the point X0 and the conditions defined by Eq. 6.66 need be stored.

l.
280 Power System Planning

Once the state Xi= X0 has been examined and classified, along with all other
states satisfying the conditions given in Eqs. 6.64 and 6.66, a new examination state
Xi =X 0 EUk> in accordance with Fig. 6.20, must be generated. In every instance the
new state should be chosen, if possible, to be an unacceptable state, since we are
seeking information only about unacceptable states. If the last state examined was
unacceptable, then a new examination stateshould be chosen by making one of the
foll owing changes:

l. Increase ri; j = self-sufficient area


2. Increase Cm; mEM
3. Decrease rk; k = deficient area

If the last state examined was acceptable, then the new examination state should be
created by decreasing the reserve margin of a deficient area. Note that each of these
changes has the effect of creating new minimum-cut interconnections that amplify the
role of interconnections in the system.
If all states XiEUk have been examined and classified, the next step is to
calculate the desired reliability indices associated with the subset Uk. Since the
information defining Bi and B11 was <1 ,~termined at each examination step, the effort is
minimal to calculate the desired indices. With the information defined by Eqs. 6.64
and 6.66 for each examination state, the probability of residing in Bi and B11 is given
by Eqs. 6.65 and 6.67. Unfortunately, Bi and B11 need not be mutually exclusive, a
result of the crude classification mechanism employed; this fact is to be taken into
account when calculating the intermediate reliability indices, For instance, if all the sub-
sets of Uk are mutually exclusive, then the system, area, and interconnection LOLPs are

LOLPk = 2f(B 11 ) + 2f(Bj) (6.68)


11 j

and

LOLP/ = [lf(B 11 ) + ~f(Bj) i = deficient area


11 J

= 0 i i= deficient area (6.69)

and

LOLP m k = 2f(B 11 ) m EM (6.70)


11

However, if the subsets are overlapping, indicating that a given state actually resides in
more than one subset, before Eqs. 6.68 through 6.70 can be used, f(B 11 ) and f(Bj)
must be transformed into nonoverlapping subsets [9].
All subsets having been considered, we may calculate the long-sought indices

------ --- --- .....


Transmission System Reliability Analysis 281

LOLP =l k
LOLPk = system LOLP

LOLPi =l LOLP/ = area i's LOLP


'k

LOLPm = LLOLPmk = interconnection m's LOLP


k

A factor that will influence slightly the general solution procedure discussed is
whether the interconnected areas are to operate under a no-load-loss policy or a
load-loss policy, so called. For later we define these two operating policies as follows:
1. A no-load-loss policy implies that reserves are transferred only up to the limit
of the interconnection or available reserves, whichever is limiting.
2. A load-loss policy implies a complete sharing of available generation up to the
limit of the interconnections, even if sharing results in load curtailment in
otherwise self-sufficient areas.
The primary effect on the solution procedure when load Joss is permissible is that in
the maximum-flow algorithm, self-sufficient areas may transfer demand through the
interconnected network to deficient areas even though this transfer may result in loss
of load in the self-sufficient areas. Thus the interconnections are involved in
Joss-of-load situations more often; that is, f(Bn) will tend to increase for each subset
Uk examined. In addition, the area LOLPs will change in that self-sufficient areas
transferring reserves to aid deficient areas will create loss-of-load conditions in their
areas and simultaneously eliminate, possibly, loss of load in the deficient areas. We
shall investigate a simple situation in Sec. 6.7.4 and highlight these details, which are
conceptually and computationally not difficult to handle.
To illustrate the concepts being presented, we shall digress for a moment and
construct a simple example we can use. Because of the role the three-bus system has
played throughout our discussion, let us not abandon it but rather reconfigure it to
resemble two interconnected areas as we did in Chap. 3. It is not difficult to imagine
how the three-bus system may be reconfigured as shown in Fig. 6.21, where the load
on bus 3 has been absorbed in the correct proportions by the two areas and the line
interconnecting buses 1 and 2 is now designated the interconnection. Analyzing this
small two-area system will not demonstrate our method very convincingly, but the
reader will agree that previous discussions involving the three-bus system have been
fruitful and that obscuring the fundamental ideas by considering huge problems is not
in our best interest. Suffice it to say, once again, that the method proposed deals with
realistic problems quite adequately, as the reader will surely discover in implementing
the concepts to solve a practical problem.
For the sample two-area system we see that ni = 2, A = 2, I= I. Thus there are
2(3)(3) = 18 system states Xi and 2 2 = 4 combinations of self-sufficient and deficient
areas, as defined below:
282 Power System Planning

T1

Area data

Area 1 Area 2

r 1 = 2 (R, = 1 p.u.) r 2 = 2 (R 2 = 1 p.u.)


r 1 = 1 (R 1 =Op.u.) r 2 =1 (R 2 =Op.u.)
r 1 =O (R 1 =-1 p.u.) r 1 = 0 (R 2 = 1 p.u.)
f(r, =2)=p 1 =0.8 f(r 2 =2)=p, =0.9
f(r, = ll=s, =0.1 f (r2 = 1) = S 2 = 0.05

f (r I = Q) = qI = Q. 1 f (r2 = 0) = q 2 = 0.05

Interconnection data

V, = 1 (c, = 1 p.u.)
V, =O (c, =Op.u.)
f(V, =l)=p,=0.9
f(V, =O)=q 3 =0.1

Fig. 6.21 Sample multiarea system data.

1. Both areas deficient,Ri < O;j = 1, 2 (subset U 1 ).


2. Area 1 self-sufficient and area 2 deficient, R 1 > 0 and R 2 < 0 (subset U2 ).
3. Area 2 self-sufficient and area 1 deficient, R 2 > 0 but R 1 < 0 (subset U 3).
4. Both areas self-sufficient, Ri > O; j = l, 2 (subset U4 ).

6.7.4 Interconnected System Reliability Analysis


We shall consider the two-area problem depicted in Fig. 6.21 for both the no-load-loss
and the load-loss policy. From Fig. 6.22 the first subset to consider is U 1 , where both
areas are deficient. Choosing X 0 = (0, 0, 0) as the examination state, we see that it is a
reserve-margin unacceptable state since both areas reside in negative reserve-margin
states. Further, based on Eq. 6.66, any state in which the reserve margins rk rko is
unacceptable. Both states in U 1 belong to B 1 , and

LOLP 1 = f(r, r10)f(r2 r2o)f (V1 > V10 )


= f(B,) = Q1Q2 = f(Bi)
LOLPitz=t =0
LOLP}=1 = Q1Q2 = f(B,)
LOLP}=2 = Q1Q2 = f(Bi)
Considering U2 (area 1 self-sufficient and a;ea 2 deficient) and letting X 0 = (1, 0, 0),
we find from the corresponding maximum-flow pattern that this state is reserve-margin
unacceptable, because it is not possible even to label the source area or, stated
differently, the label is (S, +, 0), indicating no reserves available for transfer to the
deficient area. From Eq. 6.66 we find that any state Xi such that

.~'
"-
'
Transmission System Reliability Analysis 283

~ < r1 < r10 = I


[:). < r2 < r20 = 0
V1 ;;;;,, V1 ;;;;,, V1o = 0

is also reserv~-margin unacceptable and belongs to B 1 Clearly, two states in U2 belong


to B 1 , but only state X 0 and the condition above need be stored. Since we have
classified only two states in U2 , there remain two unclassified states. Forming the new
examination state by increasing r 1 to give X 0 = (2, 0, 0), we find that this state is
interconnection unacceptable and hence belongs to B1 , where interconnection m = I is
the minimum-cut element between the labeled and unlabeled areas depicted in Fig.
6.23a. Since there exist no other states that satisfy the condition defined by Eq.
6.64,

Xi= (r1;;;;,, 2, r2 < 0, V1 < 0) E B1


= (r1 = 2, r 2 = 0, V 1 = 0) E B1
In attempting to classify the remaining unacceptable states, the capacity of the
minimum-cut interconnection is increased to create the new and last examination state
Xo = (2, 0, I). With a maximum-flow pattern as depicted in Fig. 6.23b, the state X 0 is

X (U 1 ) = (0, 0, 1 )
K (U 1 ) = (0, 0, 0)

2 States
U2

X (U2) = (2, 0, 1)
KIU 2 )= (1,0,0)
u,
4 States
X (U 3) = (0, 2, 1)
~ (U 3 ) = (0, 1, 0)

4 States

8 States

Fig. 6.22 Venn diagram for multiarca system assuming no-load-loss


policy.
284 Power System Planning

r, =2 (R 1 = 1 p,u.) r2 = 0 (R 2 =-1 p.u.)

V 1 =0 (c, =0p.u.)

/1 = 0.0 p.u.
(S, +, oo)
(al

r, = 2 (R, = 1 p.u.) r2 =0 (R 2 = -1 p.u.)

V, =1 (c 1 = 1 p.u.)

,, = 1
(S, <l, oo) (S, +, 1.0)

(bl

Fig. 6.23 Labeled graph corresponding to (a) X 0 = (2, 0, 0) and (b) X = (2,0, 1).
0

obviously acceptable. The classification of U2 is completed as all four states XiEU2


have been classified. Summarizing, the results we have for U2 are
X;=(r 1 = I, r2 =0, V,;;;.0)EB 1 2 states
X; = (r 1 = 2, r2 = 0, V1 = 0) E B1 I state
Before calculating the final probabilities, we must check to see if these subsets are
mutually exclusive. In this simple case they are; therefore the reliability information
for the system, for each area, and for each interconnection is
LOLP 2 = f(B1) + f(Bi)
LOLPJ=i =0
LOLPJ=z = f(Bi) + f(B1)
LOLP;,,= 1 = f(B1)
where f(B 1) = f(r 1 = I )f(r2 = 0)f(V1 ;;;, 0)
= S1Q2
f(B1) = f(r1 = 2)f(r2 = 0)f(V, = O)
=p1Q2Q3
Upon classifying U 3 and U4 in the same manner, we find that for the sample system

LOLP = Q1Q2 + S1Q2 + P1Q2Q3 + S2Q1 + P2 Q1 q3 = 0.028

and LOLParea 2 = Q1Q2 + S2Q1 + P2Q1Q3 = 0.019


LOLParea I = Q1Q2 + S1Q2 + P1Q2Q3 = 0.014
LOLPm=I = P1Q2Q3 + P2Q1Q3 = 0.013
Transmission System Reliability Analysis 285

For the reader's benefit the reliability indices for the sample problem assuming
q3 = 0 to q 3 = 1 are given in Table 6.1. These conditions correspond to having no
interconnection and having an infinite, firm interconnection. The benefits of
interconnecting systems so that they share reserves are again obvious from the table, as
is the fact that little improvement in reliability is realized if excessive measures are
initiated to provide the perfect interconnection; that is, the cost/benefit ratio becomes
prohibitive as the performance of the interconnection is forced toward zero outages.
As we indicated at the beginning of this section, the problem will be repeated
assuming load-loss sharing is permissible. Starting with the states Xi E U1 , we see that
since both systems reside in negative but equal reserve-margin states, no transfer of
demand will take place. Hence the reliability information determined for U 1 remains
the same:

Proceeding to U2 , we find upon examining state X 0 = (1, 0, 0) that it is an


unacceptable state as before, but now it is interconnection unacceptable since area 1
can supply demand to area 2 even though supplying it causes area 1 to transfer to
state 0 if the interconnection capacity is not zero. Further, any state

The next state to examine is X 0 = (1, 0, 1), in which case area 1 can supply area 2
deficiencies by curtailing load in its own area. Hence this state is reserve-margin
unacceptable since area 1 has insufficient reserves to meet both its demand and area
2's demand. Clearly in this state area 2 is not experiencing loss of load, but area 1 is; a
fact to be properly noted. Further, any state
X; = (r 1 = 1, r 2 = 0, V 1 ;;i, 1) E B 1
Finally, state X 0 = (2, 0, 1) would be examined and found acceptable. For subset U2 ,

then, we have
f(Bi) = s1q 2p3 1 state
f(Bi) = (s1 + p 1)q 2 q3 2 states

TABLE 6.1 Reliability indices for sample system


assuming no-load-loss interconnection policy and
q 3 = 0, 0.1, and 1
System Area 1 Area 2 Interconnection
Q3 LOLP LOLP, LOLP 2 LOLP

0 0.02 0.01 0.01 0


0.1 0.028 0.019 0.014 0.013
1.0 0.15 0.1 0.05 0.013

,,:1,
l,11,

'11 1
1
Ii'
!<OJI
286 Power System Planning

TABLE 6.2 Reliability indices for sample system


assuming a load-loss policy and q3 = 0, 0.1, and 1
System Arca 1 Area 2 Interconnection
Q3 LOLP LOLP, LOLP 2 LOLP
0 0.02 0.01 0.01 p
0.1 0.028 0.019 0.014 0.014
1.0 0.15 0.1 0.05 0.014

Repeating the same procedure for subsets U 3 and U4 and combining these results with
those for U 1 and U2 , we have
LOLP = + (s1 + P1)q2Q3 + s1Q2P3 + (s2 + Q2)Q1Q3 + S2Q1Q3
q1q2 = 0.028
LOLParea 1 = Q1Q2 + s1Q2P3 + (s2 + P2)Q1Q3 = 0.019
LOLParea 2 = Q1Q2 + (s1 + pi)Q2Q3 + S2Q1P3 = 0.014
LOLPm=l = (s1 + P1)Q2Q3 + (s2 + P2)Q1Q3 = 0.014
To facilitate the comparison between load loss and no load loss, Table 6.2 is provided.
The major difference is the extent to which the interconnection is responsible for loss
of load. We see that in the load-loss situation the interconnection plays a more
important role because of the increased transfer of reserves through the intercon-
nection. As far as the other results are concerned, the reader may question why they
are the same, which in general they would not be. In answer to that question, for the
sample problem the data used were such that s 1 q 2 = s2 q 1 , thus making the numerical
results for the two cases basically the same. Since in general s 1 q 2 =I= s2 q 1 , the indices
will be different for the two cases. Taking advantage of the situation at hand, we note
that automatically going to a load-loss interconnection policy may place increasing
demands on the interconnection's performance without materially affecting the overall
interconnected system reliability.

6.8 SUMMARY

An attempt has been made in this chapter to present both deterministic and
probabilistic methods for evaluating the reliability of a transmission system so that
comparisons between alternative expansion plans can be made. The major differences
between the deterministic and probabilistic methods are that probabilistic methods
produce indices of reliability and use nonelectrical probabilistic models of the
transmission elements.
Two deterministic methods were discus~ed: one involving the DC power-flow
model, and the other the bus voltage equations. Although the DC power-flow method
is simple and lends itself to the use of superposition to evaluate the effects of multiple
contingencies, the Z matrix method appeared to give better accuracy; that is, the
results more closely agreed with those obtained using standard AC power-flow analysis.

---
__,_,_,.
Transmission System Reliability Analysis 287

Both methods suffer from the inability to provide voltage and var flow information. It
is felt that the Z matrix method can be easily extended to provide this information.
A probabilistic approach to reliability analysis was presented whereby the indices
of reliability were again LOLP, c(DNS), LOLPm, and Em(DNS). It was shown how
these indices may provide the planner with considerable insight into the reliability of
the system as well as the role of individual elements in determining system reliability.
A lengthy example involving the three-bus system was used to illustrate the concepts
presented.
In conclusion we presented a brief discussion of a reliability method applicable
for multiarea analysis. Although the evolution of the solution procedure differed
somewhat from that of isolated-area analysis, many concepts were found to be
common to both. In multiarea analysis we emphasized determining reliability indices
that describe the reliability of the overall system as well as the role of interconnections
in establishing these indices. Once again, the basic approach was illustrated by a simple
two-area interconnected system. The results obtained from the example vividly
demonstrated the benefits to be derived from an interconnection.

PROBLEMS

6.1. Using the DC power-flow method for contingency analysis, repeat the three-bus
example presented in Sec. 6.1.1, but now add a load Sn 2 = 0.5 to bus 2.
6.2. Repeat Prob. 6.1 using the Z matrix method, and compare the results.
6.3. Write a computer program to perform contingency analysis using both the DC and
Z matrix methods for systems up to ten buses and 15 lines. Construct a sample
system from either actual system data or fabricated data. Evaluate the accuracy of
the two methods relative to each other and to results obtained using an AC
power-flow method.
6.4. For the system below, determine the reliability indices LOLP and c(DNS).
Assume two-state models and FORs q 1 , q 2 , and q 3 .
c, == 4 p.u. c 2 == 3 p.u.

c3 == 2 p.u. L == 4 p.u.

6.5. For the system in Prob. 6.4, construct a Venn diagram defining the set A of
acceptable states and the subsets B; of unacceptable states.
6.6. Determine the expressions for LOLP and c(DNS) for the system in Prob. 6.4.
Also determine the indices of reliability for each of the three elements.
6. 7. Develop a computer program to check the reliability indices computed in the text
for the three-bus system. Resolve the problem for various load profiles, element
capacities, and element FORs, and note the changes in the reliability indices.
6.8. Calculate the reliability indices for the following interconnected system:
288 Power System Planning

Assume a no-load-loss policy and

Area 1 Area 2 Area 3

r1 = 2 (R 1 = 2 p.u.) r2 = 2 (R2 = 4 p.u.) r3 = 1 (R 3 =2p.u.)


r1 = 1 (R 1 = 0 p.u.) r2 = 1 (R2 = 0 p.u.) r3 =O(R 3 =Op.u.)
r1 = 0 (R 1 = -2 p.u.) r2 = 0 (R2 = -4 p.u.)
Interconnection
Tl n T3
V1 = 1 (c 1 = 2 p.u.) V2 = 1 (c2 = 2 p.u.) V3 = 1 (c 3 = 2 p.u.)
V 1 = 0 (c1 = 0 p.u.) V2 = 0 (c2 = 0 p.u.) V3 = 0 (c3 = 0 p.u.)

BIBLIOGRAPHY
1. Baleriaux, H. et al.: "An Original Method for Computing Shortfall in Power Systems," Cigre
Paper 32-09, 1974.
2. Billinton, R. et al.: "Power System Reliability Calculations," MIT Press, Cambridge, Mass.,
1973.
3. Brown, H. E.: Contingencies Evaluated by a Z Matrix Method, IEEE Trans., vol. PAS-88, pp.
409-412, April 1969.
4. Brown, H. E.: Interchange Capability and Contingency Evaluation by a Z Matrix Method, IEEE
Trans., vol. PAS-91, pp. 1827-1832, September/October 1972.
5. Cook, V. M. et al.: Determination of Reserve Requirements of Two Interconnected Systems,
IEEE Trans., vol. PAS-82, pp. 18-33, 1963.
6. Doulliez, P.: "Optimal Capacity Planning of Multi-Terminal Networks," Ph.D. thesis, Universite
Catho!ique de Louvain, Belgium, 1970.
7. Ford, L. R. and D. R. Fulkerson: "Flows in Networks," Princeton University Press, Princeton,
N.J., 1962.
8. Moeseke, P. V.: "Mathematical Programs for Activity Analysis," chapter 9, North Holland
Publishing Co., Amsterdam, and American Elsevier Publishing Co., New York, 1974.
9. Pang, C. K. and A. J. Wood: Multi-Area Generation System Reliability Calculations, IEEE
Trans., vol. PAS-94, pp. 508-517, March/April 1975.
10. Vassell, G. S. and N. Tibbets: Analysis of Generating Capacity Reserve Requirements for
Interconnected Power Systems, IEEE Trans., vol. PAS-91, pp. 638-649, 1972.
AUTO MA TED TRANSMISSION
SYSTEM EXPANSION
PLANNING

In this last chapter it is appropriate to discuss a method for automated transmission


system expansion planning. As in earlier chapters we restrict our attention to
transmission planning only, realizing that automated methods are being applied at the
distribution level. Automated transmission system expansion planning will be used to
describe a process whereby feasible expansion strategies are determined using analytical
methods. In most cases the results of automated methods are used as guidelines by the
system planner, the expansion plan actually implemented being usually a modification
of a plan obtained using analytical techniques. Obviously, no automated procedure can
replace the planner; factors that are not easily quantified often dictate the choice,
among feasible plans, of a particular plan that is optimal in a corporate sense.
The first three sections deal with automated methods for network design and
transmission system expansion planning. In the last section we discuss the use of
interactive graphics in transmission system planning.

7.1 AUTOMATED TRANSMISSION SYSTEM


PLANNING CONCEPTS
Historically, transmission system planning has followed generation system planning
because construction time for new transmission facilities is much shorter than for
289
290 Power System Planning

generation facilities and because transmission planning depends on knowledge of the


location and capacity of both generation and load centers. Automated methods for
transmission planning require knowledge of generation and load center locations and
capacities as well as of available transmission corridors. This requirement is not overly
restrictive; this information is generally available to the transmission planner. The
performance index often used in automated transmission planning measures the degree
of overloading in the p~oposed system; that is, if no lines are overloaded when the
system is required to transmit forecasted demands, then no expansion is necessary.
However, when new load centers develop, when new generators are added, they are in
effect connected to the existing network through zero-capacity elements, automatically
creating an overload situation resolved only by adding lines of appropriate capacity to
interconnect the new elements to the system or to reinforce the existing system. The
problem boils down to automatically determining the capacities and locations of
new lines or capacity additions to existing corridors, so that no overloaded
lines will be produced under normal steady-state conditions for some future forecasted
load level. Of course the expansion plan sought is the minimum cost plan, as always.
Several approaches have been proposed to solve this problem. For the most part
they are based on DC power-flow models. Since we have not tried to discuss every
method of solving a particular problem, we shall focus our attention on the one
method that appears to have most promise. This method has a mathematical base that
may be extendable to solve transmission problems using the standard AC power-flow
models. We call it the DC method for lack of a better name. The beauty of the basic
solution approach used in this method is its use of Tellegen's theorem and its
extensions to determine the sensitivity information needed to minimize the overload
performance index (PI) w.r.t. both existing and nonexisting elements. Since a
transmission network can be reinforced by reconductoring and adding new lines,
knowing the sensitivity of the PI w.r.t. these two different types of elements is
essential. Using Tellegen's theorem a methodical and efficient procedure for obtain-
ing this information is developed. The procedure involves solving two network problems:
the original network and its adjoint network. Since the use of Tellegen's theorem
is rather new in power engineering, we devote an entire section to this topic.

7.1.1 Tellegen's Theorem


We shall not offer any formal proofs of the basic theorem; rather we choose the more
dangerous method of proof by example. However, the reader is referred to the
references for an exhaustive proof [ 1 1] .

Tellegen 's Theorem: Given a physical network and any mathematical network ( called
the adjoint network) that have identical torological characteristics and that have element
voltages, Vm and Um, and currents, im and Tm, satisfying Kirchhoffs voltage and current
laws, then
E

I
m=I
Vmim = 0 (7.1)
Automated Transmission System Expansion Planning 291

and

(7.2)

In short, any two topologically equivalent networks containing those E elements


whose currents and voltages satisfy Kirchhoffs laws satisfy Eqs. 7. I and 7 .2. For
instance, consider two arbitrary networks that are topologically the same, as shown in
Fig. 7.1. From Eq. 7.1 we have
2

l
m=I
V mlm = V1l1 + V2l2

but from Kirchhoffs voltage law and Fig. 7.Ia, Vi - V2 = 0, thus

l Vmlm = V1(i1 +12)

From Kirchhoffs current law and Fig. 7.Ib, J;_ + Yi = 0, therefore

I-
!, 1, ! -1R,
v,

(a)

I I
l
+

l'
+
i?, ii,
v, T,

I-
(b)

Fig. 7.1 Two sample networks.


292 Power System Planning

The reader is encouraged to try this on any network to gain confidence in the
theorem. Better yet~ prove the theorem, which is easily done.
Although Tellegen's theorem as stated is interesting, it cannot be directly used in
automated transmission system expansion planning, because we need a methodical
procedure for determining sensitivity information. For instance, in the network of Fig.
7.1 it is edifying to know that Eqs. 7.1 and 7.2 hold, but a more useful result would
relate the sensitivity of the current / 2 to changes in the network parameter R 2 Such
information could be used to adjust R 2 to obtain a desired current or to determine the
range over which / 2 changes as R 2 varies with temperature. Typically, in automated
system design or planning, the problem is to minimize a performance index w.r.t.
independent parameters of the system, which requires sensitivity information. As an
example, assume we are seeking the value of R 2 , in the circuit shown in Fig. 7.la, that
results in a current / 2 = 1.0 amp. This is a trivial example, the answer being obvious by
inspection, but it illustrates the point. Mathematically, the problem is a minimization
problem:
minJ = (1 2 - 1.0)2
w.r.t.R,

The performance index here is again an error-squared index commonly used in


optimization problems. The necessary condition for a particular R 2 being the minimum is
dJ a12
dR = 0 = (1 2 - l.0)aR
2 2

Note the need for the sensitivity of the current / 2 w.r.t. R 2 . From this example, the moti-
vation for extending Tellegen's theorem to obtain sensitivity information should be obvious.

7.1.2 Network Sensitivities


The problem before us is to determine, using Tellegen's theorem, the sensitivity of
element variables Um and im describing a real physical network to changes in their
network parameters. To solve this problem, we must only realize that parameter
changes in the physical network cause changes ~um and ~im in the element variables.
Since the new element voltages and currents are now Um + ~um and im + ~im, Eqs.
7.1 and 7.2 become
E
l(
m=I
Um + ~Um )('l;,,,) = 0 (7.3)

l
m=I
Vm(im + ~im) = 0 (7.4)

With Eqs. 7.3 and 7.4 expanded,

l
m
Vmim + l
m
~Um 'l;,,, =0
Automated Transmission System Expansion Planning 293

and

_lvmim + _l 'vm!:l.im = 0
m m

but from Eqs. 7.1 and 7.2

Therefore

,l D..Um i,,, = l VmD..im = 0


m m

or

,l D..Um i,,, - VmD..im = 0 (7.5)


m

By using Eq. 7 .5, the sensitivity of any element variable to changes in either existing
or nonexisting element parameters can be obtained. Consider determining the
sensitivity of element voltage Up w .r.t. parameters Ym, m = l, ... , :
aup
m=l, ... ,E (7.6)
aym

If
im = f(um,Ym) m = l, ... ,E
then
.
D..lm
ar
= -a-D..Um + -a-D..Ym
ar (7.7)
Um Ym

With Eq. 7.7 substituted into Eq. 7.5

D..up Ip - vpf:l.ip + 10-:n -


mi=p
Um a~)!:l.um - l
mi=p
Vm a~ D..Ym = 0 (7 .8)

Since Eq. 7.8 is valid for any adjoint element values, we may arbitrarily define the ele-
ments of the adjoint network. In view of the information sought, the second term in Eq.
7.8 should be force,d to zero by defining the elements of the adjoint network to have
vm1m characteristics defined by af/avm; that is, for element m 1= pin the adjoint net-
work
, ., _Um -a-=
lm -
ar Q m =I= P (7.9a)
Um
294 Power System Planning

or

Tm
-=--
at m -=I= P (7 .9b)

If element m in the physical network is a constant voltage source, then at/aum = 00,
and from Eq. 7 .9b the corresponding adjoint element is a short circuit. Similarly, if
element m in the physical network is a constant current source, then at/aum = 0, and
the adjoint element is an open circuit. Finally, if element m is a constant admit-
tance, then from Eq. 1.9b, it also will be a constant admittance in the adjoint
network.
With the second term of Eq. 7 .8 set equal to zero, we have

(7.10)

Since

then, again,

D..1p
. == -a
at D..up + -aat-t,,.yp (7 .11)
Up Yp

Substituting Eq. 7.11 into Eq. 7.10, we have

(7.12)

Equation 7 .9 was used to define all the elements of the adjoint network except
element p. We now may define element p from Eq. 7.12. Considering that we want
sensitivity information about Up, the coefficient of D..uP in Eq. 7 .12 should be forced
to unity, a step that defines element p in the adjoint network:

or
,., at
ip = 1.0+ -a up (7.13)
up

This expression for adjoint element p defiqes a voltage-controlled current source.


Equation 7.12 reduces to

(7.14)
Automated Transmission System Expansion Planning 295

from which we obtain

(7J5a)

and
avp - at
--=v
aym
--
m aym m *p (7.15b)

Note that
1. To determine the desired sensitivities, we need only the element characteristics
evaluated at the solution point of the physical network and the corresponding
adjoint network voltages.
2. Equations 7.15a and 7.15b define the sensitivity ofup to changes in all network
elements y m .
Again, the procedure proposed allows us to calculate the network voltage
sensitivities by solving only two network problems. This feature of the method is
powerful; indeed it is the reason that the use of Tellegen's theorem in automated
network design is receiving considerable attention. In addition, the method applies
when sensitivities w.r.t. nonexisting elements are required, an important feature for
automated design considerations. This apparent flexibility is due to the fact that the
sensitivity expressions given in Eq. 7 .15 require only bus voltage information to
determine the element voltages and the associated element characteristics at/au;,
i = 1, ... , E; no information is needed about the parameter values themselves if tis a
linear function of Yp, which is usually the case for power systems.
Because of the number of unfamiliar concepts presented, consider the three-bus
system of Chap. 5, Fig. 5.16. For simplicity we shall combine the generators and lines
13 and 23, and replace the load by a constant admittance. The simplified network
model is depicted in Fig. 7 .2. Assuming that we wish to determine the sensitivity of

/1 V3 =0.9789L-11.78

+ v, -
t------+----i y 2 = -j20 ,___ _ __,

V 1 =1.0L0

y3 = 4.17
lI'

Fig. 7 .2 Simplified three-bus system.


296 Power System Planning

V3 to changes in the system parameters y 2 and y 3 , using Eq. 7.5, we get

t..VJ1 - J7iAf1 + t..V2f2 - V2t..f2 + t..V3]3 - V3t../3 = Q

Since clement I is a constant voltage source, t../1 = 00


t..Vi, and

71 - V1 = o
00

11 -
;;;; - V1 = 0

Therefore V1 = O; that is, element I in the adjoint network must be a short circuit.
Similarly, to eliminate the term containing t.. V2 ,

12 - V2Y2 = 0

or

V2
-::::;- =Y2 =y2
12
With two of the three elements in the adjoint network defined, Eq. 7.10 becomes
t..V3]3 - i\t..!3 - V2V2t..Y2 = Q
Substituting t../ 3 = y 3t..V3 + V3t..y 3 into the above, we obtain Eq. 7.12:
t..V3(]3 - V3y3) - V3V3t..y3 - V2V2t..Y2 = Q
With the condition defined by Eq. 7.13 forced,
73 - V3y3 = 1
The third adjoint element becomes a voltage-controlled current source in which
73 =I+ V3y3
We now have all the information to determine the desired sensitivities. The adjoint
quantities are calculated by solving the network equations for the adjoint network
shown in Fig. 7 .3:
av3 -
_a_~ V3V3
Y3
av3 -
-a-~
Y2
V2V2

From Fig. 7.2 we see that


V3 = 0.9789 L-11.78
V2 = 1.0L0 - 0.9789 L-I 1.78 = 0.205 L-84.1
Automated Transmission System Expansion Planning 297

Ii\= 0
+ v,
v, _,,

v,

Fig. 7 .3 Adjoint network for simplified three-bus system.

Also from Fig. 7.3

- -1 1 0
V3 = y2 +y 3 = 4 _17 _ j 2 0 = -0.049 L78.22
-
V 2 = -V3
-
= 0.049 L78.22 0

Thus

av3
-a- = -{0.049 L78.22)(0.979L-11.78)
Y3
= -0.048 L66.44
and

av3
-a- = (0.049 L78.22)(0.204 L-84.1 )
Y2
= 0.01 L-5.9

To check the validity of this information, assume lly 3 = 0.417. Using the
sensitivity information just calculated, we find

fl V3 = -(0.048 L66.44)(0.417)
= -0.02 L66.44 = 0.02 L-113.56

However, if we rework the problem, a simple voltage divider, with y 3 = 0.417 +


4.17 = 4.58, we find that
vnew 0.2183 = 0.2183
3
= 0.2183 + j0.05 0.2239Ll2.90
= 0.9749 L-12.90
298 Power System Planning

and
~ V3 = 0.9749 L-12.90 - 0.977 L-11.78
= 0.9502 - j0.2176 - 0.9564 + j0.1994
= -0.0062 - j0.0182
= 0.0192 L-108.81
The results for ~ V 3 are quite close, the error being simply inaccuracy in computation. It
can easily be shown that the sensitivity expressions are exact by differentiating the
voltage divider expression for V 3 w .r.t. y 3 ; this is left to the reader.
Armed with Tellegen's theorem, we are prepared to investigate its use in
automated network design. Although we present the basic notions in the context of
networks constrained only by voltage and current relationships, as we shall see later,
automated transmission system planning can be formi.1lated such that the notions
presented here may be directly applied.

7.2 AUTOMATED NETWORK DESIGN


We shall discuss in this section the basic problem of network design oriented toward
future application to transmission system design. In particular we show that by
appropriately exciting the adjoint network, the gradient of the PI w.r.t. the system
parameters can be directly obtained. This is an important idea, since the gradient is
needed to update the system parameters to optimize the PI.

7.2.1 Problem Formulation


Typically the network design problem involves determining network parameters Ym,
m = I, ... , L, that produce the desired dependent element voltages and currents. Here
L denotes the number of network elements with dependent voltages and currents, and
E the total number of network elements. To simplify the discussion, we assume here
that the PI is only a function of the dependent element voltages denoted by vector Xv
and that the problem can be stated as follows:

min J = G'(Xv)WG(Xv) (7.16)


w.r.t. Y

subject to the c0nstraint that


F(Xv, Xi, Uv, Y) = F(X, U, Y) = 0 (7 .17)
where

Xv -L]- L X 1 vocto, of depend:nt element voltages

L+] =
Uv =
[ u~ (E - L) X 1 vector of independent element voltages
Automated Transmission System Expansion Planning 299

X; ~ [ ] ~EX 1 vecto, of dependent element euuents

Y ~ [J ~ L X 1 vecto, of netwo,k parnmetm

To minimize J w.r.t. Y, we must obtain the value of Y that results in element voltages
and currents v and i, determined from Eq. 7.17, that force the gradient to zero. The
gradient is given by
aJ I axy
aY = G CXv)W ay (7.18)

The approach commonly used to choose successive values of Y is steepest descent; that
is, the solution algorithm is given by

yv+I = yv _ (X (!~J (7.19a)

= yv -
ax \"
ex ( ayv WG(Xv)j (7.19b)

By using Eq. 7 .19 the parameter vector is updated and used to solve for Xv and Xi in
Eq. 7 .1 7. If, upon substituting these quantities into Eq. 7.18, a zero vector results, the
optimization process is finished, and the optimal parameter vector is known.

7.2.2 Determining the Gradient Vector


Classically, the gradient vector given in Eq. 7.18 is determined in an analytical fashion,
as is done in creating the Jacobian matrix in power-flow analysis. However, there are
obvious disadvantages to obtaining the gradient in this manner, one being the need for
the analytical expressions relating vx and Y, and the other being the difficulty of
handling nonexisting elements. To circumvent these two problems, we look to
Tellegen's theorem. As we will show momentarily, the gradient information is easily
obtained from the solution of the original and adjoint networks, where the excitation
of the adjoint network must be the gradient of J w.r.t. the dependent element
voltages. All of this will become clearer as we develop the basic concept.
For use later, expand Eq. 7.18 so that the exact nature of the expression is clear.

(7 .20)
300 Power System Planning

We would like to show that Tellegen's theorem can be used to obtain information
defined by Eq. 7.20. Starting with Eq. 7.5,
E
l
m=I
t:..vmim - Vmt:..im =0

and substituting Eq. 7.27 into the above for all L-dependent elements, we get

(7.21)

The first term can be dropped by realizing that t:..vn = 0, n = L + 1, ... , , and by
defining Un = 0, n = L + 1, ... , . Thus Eq. 7 .21 reduces to

(7 .22)

Now, from Eq. 7 .20, define the coefficients of t:..vm in Eq. 7 .22 to be

(7 .23)

Therefore Eq. 7.22 becomes


L
at
L
m=I
Km(vm)Wmmt:..Vm =
L
"'Dm-a
~
m=I
-t:..ym
Ym
(7 .24)

Since we have control over which parameters change, let t:..ym = 0, m = 2, ... , L.
Clearly, Eq. 7 .24 after division by t:..y becomes
t:..u1 t:..v2 t:..vL _ at
K1(vi)W11 ~ + g2(V2)W22 ~ + ... + gL(vL)WLL ,:-
<->.Y1 uy1 uY1
= V1 - a
Y1
but

Thus
au1 avL at
K1(u1)W11 - a + + gL(vL)WLL - a
Y1 Y1
= 'i.J1 -
ay1
Comparing the left-hand side of this last equatidn to the right-hand side of Eq. 7 .20,
we note that
Automated Transmission System Expansion Planning 301

In general
a1
--=u--
_ a/
m=l, ... ,L (7 .25)
aym m aym
Further, in linear systems
at
--=v
aym m

and Eq. 7.25 becomes

a1
-a-= -VmVm m =I, ... ,L (7.26)
Ym
Interpreted correctly, Eq. 7 .26 says that the elements of the gradient vector needed to
minimize the PI can be obtained by knowing the voltages across the elements in both
the original and the adjoint network. The adjoint network is defined as having
short-circuited voltage sources and elements whose voltage-current relationship is
defined by Eq. 7.23. For linear systems Eq. 7.23 can be expressed as follows:
(7 .27)
From this last expression we note that the adjoint elements are admittances in parallel
with a voltage-dependent current source gm(vm)Wmm, which is the gradient of J w.r.t.
Um.

7.2.3 An Example
Consider the voltage divider network depicted in Fig. 7 .2. Assume we wish to alter y 3
in a way that will produce a voltage at bus 3 with a magnitude of 1.0 p.u. instead of
0.9789 p.u. The problem may be stated in the following manner:

minJ = (V3 - l.OL-11.78)2


w.r.t. Y,

subject to the constraint that


Vi
V3 = ----
1 + y3/Y2
In this case

The gradient ofJ from Eq. 7.18 is


302 Power System Planning

Fortunately, we have already developed Tellegen's equation for this example in Sec. 7 .1.2.
It is repeated for convenience:

In accordance with Eq. 7 .27


13 = y 3f\ + (V3 - 1.0 L-11.78)

Thus, Tellegen's equation becomes


(V3 - 1.0 L-11.78)b.V3 = J\V3b.y3
or

Finally

(V3 - 1.0 L-11.78)-a-


av3 = -a
aJ = -V3V3
Y3 Y3
To find V3 and V 3 , we must solve the adjoint and ,..9riginal networks for V3 and V 3 ,
respectively. As we have V 3 already, we can obtain V 3 by solving the adjoint network
defined below:
1. f\ = 0; that is, short-circuited
2. Yi= Yi
3. l; = y3 j:\ + (V3 - 1.0L-11.78)
The adjoint network described by the three elements above is shown in Fig. 7.4;
the original network, in Fig. 7.2. From Fig. 7.4 we see that
-V3 = - +1 (V3 0 - l.OL - 11.78)
Yi Y3
= -0.001 L246.48
Thus the gradient is

af -
-a = V3V3
Y3
= -0.000978 L234.70
and
y3 1 = y3 - o(-0.000978 L234.70)
At this point we need to obtain the optimal step size by using a one-dimensional
search procedure or by substituting y/ into the PI and minimizing w.r.t. Cl'..
A third and yet simpler alternative is to evaluate the PI and determine
how much it must be changed; then, using the gradient information, calculate the
Automated Transmission System Expansion Planning 303

V,=O

(ii, -1.0L-11.781 l
Fig. 7.4 Adjoint network for calculating gradient of sample system's
performance index.

appropriate change in lly 3 . Evaluating the PI at the initial solution point, we get

J -_ 2I ( V3-l.OL-ll. 78 0)21 v,=o.97S9L-11.?s 0

= 0.000223 L231.42
Thus
ll.J = -0.000223 L23 l.42
From the gradient expression

lly 0 = -M = -0.000223 L336.52


3
V3 V 3 -0.000978 L23 l .42
= 0.2278 LlOl.82
Therefore

y/ = 4) 7 + 0.2278 LlOl.82
= 4.123 + j0.222
To check the validity of this result, we must resolve the original and adjoint networks
for V/ and V/ and, theoretically speaking, evaluate the gradient to insure that it is
either zero or sufficiently small. Doing so, we find that

I V1 1~
3
V = 1 + y//y 2
= 1 + (4.123 + j0.222)/-j20
'
= 0.99L-11.77
1
and V3 1 =- (V/ - l.OL -11.78)
Y2 + Y3 1
"" 0
304 Power System Planning

TI1erefore

a1 - -v3 iv3 i
-.-- --- o
oy3
Note that as the optimal solution is approached, the excitation for the adjoint network
goes to zero. With no excitation, 'i\, like the gradi~nt, is zero.

7.3 AUTOMATED TRANSMISSION PLANNING:


A DC METHOD
TI1e approach to automated network design discussed in Sec. 7 .2 is obviously not
restricted to linear systems, since the functional relationships .between the problem
variables i and v were defined in terms of a general function f. However, if the
network is linear, the parameters of both original and adjoint networks that are not in
Y, the parameter vector, are constant throughout the optimization process-an
attractive characteristic from a computational point of view, and illustrative of the
special nature of linear networks. An example of this point is the parameter y 2 in the
problem presented at the end of Sec. 7 .2.3: it never changed.
It would be fortunate if we faced a strictly linear problem, but that is not the
case. Even though a power system under steady-state conditions is considered a linear
network, the power-flow equations used to describe the performance of the system are,
as we know, very nonlinear, Thus we must do one of two things: construct a
completely linear model of the system so that the networks solved in the optimization
process have constant elements, or linearize the nonlinear performance equations,
updating the two networks iteratively as we move toward the optimal solution.
Obviously the simplest but least accurate approach assumes a linear model of the
system. We shall discuss automated transmission planning using the DC power-flow
model as the linear system model [ 5]. As of this writing, no attempt has been made to use
the linearized AC power-flow model, and we must be satisfied with discussing the DC
method only.

7.3.1 Problem Formulation


In Chap. 6 we developed a linear model of the network, which we repeat below:
N
P; = ,l IV;IIVilB;i(8; - 8i) i = 1, ... ,N
j=I m = elements incident on bus i
jif:.I

(7.2&z)

where

Pm= IV;IIVilB;j(D; - 8i)


= B;j(D; - Dj) = Bm8m (7.28b)
Automated Transmission System Expansion Planning 305

since the voltage profile is normally assumed flat and equal to 1.0 p.u. and

Bm ~ B;;
(7.29)
Om~ 8; - 8;
With the variable definitions above and the linear network model we are prepared to
go into the basic approach of automated transmission planning using the DC
power-flow model. The problem may be stated, as indicated in Sec. 7.2, as a static
optimization problem:
minf = G'(X)WG(X) (7.30)
subject to the constraint that
F(X, U, Y) =0 (7 .31)
and
(7.32)
Defining the state, control, and output vectors in terms of element quantities, we have

X= U= (7 .33)

Pm, m = 2, ... ,L, are the dependent-element power flows including the slack generator
output P 1 , and Pm, m = L + 1, ... ,E, are the independent-element power flows, such as
loads and nonslack generator power output. Further

(7.34)

where Bm is the susceptance of element m. We must remember that both X and Y


may contain power flows and susceptances for elements that do not exist, for reasons
given in Sec. 7 .2. For simplicity we assume that the PI is a weighted error-squared-type
index in which the error is the difference between line loadings Pm and capacity Cm.

7.3.2 Determining the Gradient Vector


Applying Tellegen's theorem to this specific power problem, where the network is now
defined by Eq. 7 .28, we have
306 Power System Planning

f
n=L+I
/::,.onPn - bnMn +
L
l
m=2
/::,.omPm - f,mt:,.Pm + /::,.o 1P1 -'t{I !::i.P1 =0 (7.35)

From Eq. 7.28b


Mm= Bm!::i.om + 0 mMm m=2, .. . ,L
and /::,.o I = 0
Mn=O n=L+l, ... ,E
Equation 7.35 becomes

(7.36)

Setting Pn = 0, n = L + 1, . .. , E, and 81 = 0, we get an equation similar to Eq. 7.22. To


obtain the gradient, we must define the coefficients of Mm to be the aJ/aPm ; that is, in
accordance with Eq. 7.23
Pm - aJ
Bm - 0m = aP
m
= Km(Pm)Wmm (7.37a)

or
Pm= Bm8m + Bmgm(Pm)Wmm
(7.37b)
= Bm8m + Bm(Pm - Cm)Wmm
Substituting Eq. 7.37a into Eq. 7.36 with~ =Oand8 1 =0,weget
L L p
L
m=2
Km(Pm)WmmMm = l
m=2
O
;mm Mm

Using the procedure described in Sec. 7.2.2 and this last result, we get

aJ OmPm
= m =2, . .. ,L

om -
= B[Bmom +Bm(Pm - Cm)Wmm]
m

= Om8m + Om(Pm - Cm)Wmm (7.38)

From Eq. 7.38 we finally obtain the gradient of the PI w.r.t. the parameter vector Y.

(7.39)
Automated Transmission System Expansion Planning 307

The way an automated transmission design problem would be solved using the
equation above is to solve the power-flow problem for the original network and, with
the information obtained, to construct and solve the adjoint network. With the original
and adjoint network solutions, Eq. 7.39 can be evaluated and the parameter vector
updated. This procedure is repeated until the gradient is sufficiently small.

7.3.3 Including Practical Considerations

It is incumbent upon us to emphasize that the basic approach described has several
shortcomings:

1. It ignores the fact that as line susceptances change, the capacity Cm also
changes.
2. The PI is only a measure of network loading and does not reflect the
economic aspects of network design.
3. It results in an adjoint network that is not structured for easy solution by
Newton's method.

Let us discuss how these points may be resolved. Items 1 and 2 above are
handled basically the same way; that is, to deal with the fact that

(7.40)

we need only treat Cm as a variable in the derivation. The PI in Eq. 7.30 handles the
case where cm is not constant. However, Eq. 7 .37a must be changed, since it assumes
cm is a constant. Obviously, if Cm is not constant, then Eq. 7.37a becomes

Similarly, if the PI were to reflect the cost of line additions, not only would the Pl in
Eq. 7.30 require changing, but again Eq. 7.37a would change. If we assume a linear
cost, then the PI becomes
J = G'(X)WG(X) + K'X (7.41)
where K is the vector of unit line costs. Now Eq. 7.37a becomes

or finally

(7.42)
308 Power System Planning

Substituting Eq. 7.42 into 7.38, we obtain the new gradient vector.

a1
av = (7.43)

The last remaining point is computationally significant. To solve the adjoint network,
we are faced with solving a network whose elements satisfy the following relationships:

(7.44)
m=2, ... ,L

Unfortunately, these equations are not compatible with the standard DC power-flow
solution algorithms, since the m = 2, ... , L elements have dependent power-sources
in parallel with susceptances Bm. To solve this problem we need to define as inputs
the injected powers into the N buses. The question arises: How can we convert the prob-
lem before us into one that can be solved using a DC power-flow method? To answer this
question, consider Fig. 7 .5, which depicts the structure of the adjoint network. Since

P,
----bm - - -
Bm

Fig. 7.5 Structure of adjoint network used in automated power system design.

. ,,, ............,J
Automated Transmission System Expansion Planning 309

PG,= 4p.,. l c1 = 4.0 p.u.


PG2 = 4 p.u.j
C2 = 4.0 p,U,

-,--,--_,__-,- c5 ~ =0 -
c 12 -3 p,u. ~-'--~-
0~ = 0.072

B2, = 15 1.08
C 13 = 2 p,u.

~=,O

c 23 = 2 p.u.
c~ 4 = 4 p.u. I~ P~ 3 = 2.18
8~ 4 = 10 \ 2.9

\
\
II I
~--
C 34
-'-~--...L,- 0~

...--- ---; =
8~=15
2 p.u. So 3
=-0.218

= 4 p.u.

~.s~ ==-0.29

/ i
So4 =4 p.u.

New load center

Fig. 7.6 Three-bus system with new load bus.

the dependent-element sources are merely injections into and out of the sending and
receiving ends of each element, their combined effect is equivalent to injecting into
each bus the sum of the element injections of all elements incident on a particular bus.
Stated mathematically

l
N

P;= (7.45)
m=all elements
incident on bus i

7.3.4 Expanding the Three-Bus System


One final example will be presented, of course involving the three-bus system. To
demonstrate the mechanics of using the DC method for automated transmission system
planning, assume that because of load growth, a new load center is created as depicted
in Fig. 7.6. Further assume that the existing system cannot support new circuits, etc.,
and that the only feasible solution for serving new load is to add lines between buses I
and 4 or between 3 and 4, as indicated by the dotted lines in Fig. 7 .6. Given the initial
feasible line characteristics shown next to the line, the power-flow pattern in the network
is obtained by solving the following DC equations for the bus voltage angles and, finally,
the line flows:
310 Power System Planning

2501 - 1502 - 1003 - 1004== 4


-1501 + 2502 - 1003 == 4
(7.46)
-1001 - 1002 + 3503 + 1504 == -4

Since 8 1 == 0 (the reference bus), the DC equations can be written as

(7.4 7)

;oJving Eq. 7.47

0I == 0
02 == 0.072
03 == -0.218
04 == -0.29
/ith these angles substituted into Eq. 7 .28b the flow pattern shown in Fig. 7 .6 results.
fote that lines 13 and 23 are overloaded, as indicated in Fig. 7.7. The object, in view
f the overloaded lines, is to update the new line parameters in a way that reduces the
,ading on the overloaded lines. Stated differently, we want to minimize the

C 12 =3 P = 1.08 --,-.-L..----,,-
21
p.U. - - - - -

P 13 = 2.18 p.u.
l c 13 = 2 p,u.
= 4 p.u.
=2.9p.u.
11I elements

I
\
\

1 --~--
\ J

' c 34 - 2 p.u.
I I P 34 = 1.08 p.u.

~ .. ,
. 7.7 Expanded three-bus system depicting overloaded elements.

l
Automated Transmission System Expansion Planning 311

821 (P2, -c,.) = -15(1.92)

82, = 15
8 13
(Pl3 -c 13 ) = 10(0.18)
~

834 = 15

Fig. 7 .8 Adjoint network for expanded three-bus system.

performance index that reflects the degree of overloading in the system w.r.t. the new
line parameters. Assuming unit weighting factors, the Pl for the sample system is

J = [(Pu -c13)2 + (P23 -c23)2]


(7.48)
= [(0.18) 2 + (0.9)2] = 0.42
Since B 14 and B 34 are the only parameters that can be altered, the gradient
vector for this example becomes

(7.49)

All the quantities in the gradient vector are known except for the adjoint network bus
voltage angles Dm. These angles, as discussed in Sec. 7.3.1, must be determined from
the adjoint network described by Eq. 7.44. The adjoint network for this example is
shown in Fig. 7,8.
At first Fig. 7 .8 does not appear to have a structure similar to the actual
network, but based on our discussion at the end of Sec. 7 .3.3, we readily see that Fig.
7 .8 is equivalent to the network in Fig. 7 .9. Since the network in Fig. 7 .9 is
topologically the same as the original network shown in Fig. 7.6, we may employ the
same solution procedure to determine the adjoint bus voltage angles:
312 Power System Planning

25
-IO
[ 0
-10
35 -15 ~ u-~ = b ~3 19.~
24.6 (7.50)
-15 25 04 -24.8

from which
02 = 0.741
03 = -0.125 and
84 = -0.587
Upon substituting these angles and the overload information into the gradient
equation, Eq. 7.49, we get

aJ r-0.171]
as= L+o.028
Letting a= 20, the updated values of the new parameters are
, 0
B 14=B14 +a0.171=13.42
, 0
B 34 =B34 -a0.028 = 14.44

8 21 = 15

B' 3 IP 13 - C 13) + B,3 IP,. - c,,) - B 34 IP 34 - C 34) == 24.6


834 = 15

B 14 IP, 4 -c, 4 ) + 8 34 IP 34 -c 34 ) = -24.8

Fig. 7.9 Adjoint network with element injections rearranged.


Automated Transmission System Expansion Planning 313

Recalculating the bus voltage angles and line flows using the new line parameters
above, we find that

01 = 0 P 14 = 3.26 p.u. P2 3 = 2.76 p.u.


02 = 0.08 P 3 4 = 0.70 p.u. J = 0.29
03 = -0.194 P13 = 1.94 p.u.
04 = -0.243 P21 = 1.21 p.u.

Obviously, the procedure has properly shifted the flow to the new bus, away from the
overloaded existing system. Initially both lines 13 and 23 were overloaded, but after
only one iteration the overload on line 13 has been relieved and the PI reduced from
0.42 to 0.29. If this process were repeated, lines 14 and 12 would assume more
responsibility for serving the new load, and eventually the algorithm would remove line
34 entirely.
We close by cautioning the reader that blind use of the techniques presented will
always produce unacceptable plans. It is up to the planner to guide the algorithms to
seek and identify viable plans.

7.4 AUTOMATED TRANSMISSION PLANNING USING


INTERACTIVE GRAPHICS
We have explored, in great detail, methods used in power system expansion planning.
However, little has been said of a computational environment that may provide the
system planner with the capability for fast, efficient, and flexible analysis. Rather we
assumed that the methods discussed would be coded for use on batch processors. In
Sec. 7.4 we shall briefly discuss the use of interactive graphics in transmission system
planning. In addition, because some readers may be unfamiliar with graphic systems, an
overview of the hardware and software of an interactive graphic system is presented.
Finally we consider examples of the use of graphic systems in line design and
power-flow analysis.

7.4.1 Characteristics of Interactive Graphic Systems


Figure 7.10 illustrates the basic hardware configuration normally associated with small
interactive graphic systems. Typically, those systems built around minicomputers have
16 bit words with 32k words of core storage and 2 to 2.5 million words of disk
storage. The major components of such systems are
1. Central processor
2. Display processor
3. Disks
4. Teletype
5. Refresh CRT and light pen
The software normally supplied with such systems consists of the following
programs:

J
314 Power System Planning

Disk Disk

CPU

TTY

LP
Fig. 7 .10 Configuration of interactive graphic system.

1. Monitor or executor
2. File manager
3. Editor
4. Fortran graphics compiler
5. Linker and loader

Although this list is hardly complete, it does contain the major programs of interest.
Since the function of these programs is self-explanatory, we shall only comment on
features that are important in developing interactive transmission planning methods.

File Manager
This progr:1m is essential in transmission planning via graphic systems since the planner
will be continuously altering data files to create new files describing the various
alternative expansion plans or updating the topology of the network. A good example
of the use of the file manager occurs in contingency analysis; each time an element is
removed, the data files describing the network must be quickly and efficiently
updated.

Editor
The editor is important during the program development phase. Generally the function
of the editor is to allow the planner to insert source code into the text buffer, to
modify or correct the code as necessary and to transfer and store the source module in
a designated area on the disk.

Fortran
In addition to regular Fortran code the graphics' Fortran must be able to compile the
graphic subroutines listed below:
1. Scaler
2. Beam positioner

j
Automated Transmission System Expansion Planning 315

3. Vector generator
4. Text generator
5. Subpicture (a collection of dots, vectors, etc.)
6. Timer
7. Tracking cross

Again, the function of these subroutines is self-explanatory. What is perhaps not


obvious in the names is the fact that all dots, vectors, and subpictures can be made to
be light-pen sensitive. This feature enables a planner using a light pen to alter the status
of any light-pen-sensitive element. In addition, all dots, vectors, and subpictures can be
displayed using a multitude of intensity levels, or they can be made to appear blinking.

7.4.2 Interactive Graphic Power-Flow Analysis


For a sample of the use of an interactive graphic system in transmission planning,
consider the problem of determining the basic network changes needed to expand a
system to serve projected loads located at new load centers. The first step is to
determine the electrical parameters of the lines to be considered in interconnecting the
isolated new load center to the system. Specifically, this step includes
1. Defining the voltage level.
2. Defining the basic phase configuration.
3. Defining the insulator type.
4. Defining the tower configuration.
5. Calculating the line parameters.
Using a graphic-oriented line design program, the planner can draw a tower and phase
configuration for any preselected voltage level and insulator type and then calculate
the corresponding line parameters and surge impedance loading (SIL). With the SIL
known, a quick check ensures that the line in question has sufficient capacity to serve
the load center being considered.
To illustrate how a graphic line design program may be used, consider Fig. 7.11,
which is a picture of the initial line design grid used in a program developed on a
Digital Equipment Corporation Gt 44 by Dwight Odom, a graduate student in the
Department of Electrical Engineering at the University of Florida. The alphanumeric
text at the top of the picture provides the planner with several decision paths to create
a new line or to use a standard line, the data for which are stored on disk. Assuming a
new line is to be designed, the planner simply moves the light pen over the words
"new tower." Detecting a light-pen "hit," the program turns on a list of standard
voltage levels to the right of the screen. Hitting the voltage level desired, the program
enters the phase-building mode. Using the light pen to direct the tracking cross over
the desired location of each phase, the planner may construct many different phase
configurations; Fig. 7 .12 shows a common configuration. The dotted circles define the
minimum phase clearances for each conductor, and the boxes define the swing areas
for each conductor, as determined by the type of insulator chosen. Once finished with
the phase configuration, the planner simply hits the alpha text "finished" in the upper

J
316 Power System Planning

Fig. 7 .11 Initial transmission tower design grid.

left-hand corner of the screen, and the program immediately transfers to the
tower-building mode. Figure 7.13 depicts the final steps of the tower configuration.
Since every line is drawn by the planner using the light pen, any tower configuration
can be considered. Completing the tower-building phase, all the physical data needed
to calculate the electrical line parameters are known. A standard line parameter
program is used to calculate the line resistance, reactance, and shunt capacitance.
Figure 7.14 is a one-line diagram of a system in which the parameters of the new
lines were determined using the line design program just discussed. The interactive
power-flow program used to draw the one-line diagram was developed at the Swiss
Federal Institute of Technology in Lausanne by H. B. Puttgen, also a graduate student

Fig. 7 .12 Graphic results of phase design.


Automated Transmission System Expansion Planning 317

Fig. 7 .13 Graphic results of tower design.

in the Department of Electrical Engineering, University of Florida. Salient features of


the interactive power-flow program are
1. New lines and buses can be easily added using the light pen.
2. Existing lines and buses can be removed using the light pen.
3. Each bus can be interrogated.
4. Load level and generation schedule changes can be conveniently made using
the TTY.
5. Overloaded lines and low-voltage buses can be displayed blinking.
Although it is not obvious from Fig. 7.14, each time the topology of the network is
altered, the files containing the power-flow data are automatically restructured.

Fig. 7 .14 One-line diagram of a system being evaluated using an


interactive graphic power-flow program.
318 Power System Planning

7.4.3 Interactive Grap hie Automated Transmission Design


Based on the discussion in Sec. 7.4.2, it is not difficult to see how the use of graphics
can enhance the power of the automated transmission design procedure discussed in
Sec. 7.3, where we noted that the gradient vector contains the gradient of the PI w.r.t.
the parameters of both existing and nonexisting elements. Since it is unrealistic to
assume that new lines can be constructed between any two buses, the planner may
define with the light pen the receiving and sending ends of buses that, realistically, can
be connected with a line built through an available corridor. Further, since the PI
contains weighting factors for the relative importance of individual line overloads, the
system planner can alter these weighting factors according to flow patterns displayed
on the CRT after each iteration. In this way better control over the growth of the
system can be maintained.

7.5 SUMMARY

In this final chapter we discussed the application of Tellegen's theorem to the basic
automated network design problem. We developed sensitivity expressions describing the
sensitivity of dependent network variables to changes in network parameters, using the
solutions to the original and adjoint networks. It was shown that not only cari
sensitivity information for existing parameters be obtained, but sensitivities of
dependent network variables w.r.t. nonexisting elements can be computed as well.
These ideas were extended to produce a method for determining the gradient vector to
optimize the performance of a network by updating network parameters. It was also
shown that the gradient can be obtained knowing the original and adjoint network
solutions, where the adjoint sources are functions of the PI, a technique first proposed
by Director and Rohrer [4] .
The basic ideas for calculating the gradient vector for use in classical network
design were applied to transmission system design. In particular, we discussed how a
performance index reflecting the degree of system overloading can be minimized using
design ideas presented in conjunction with DC power-flow models. Through a rather
detailed example involving the three-bus system, the mechanics of applying the
technique to automated transmission system design were illustrated. Its usefulness, like
that of any analytical method, is highly dependent on the planner's ability to direct
the procedure toward a solution that is indeed viable.
In anticipation of the role that interactive graphic systems may play in planning,
a very brief discussion of the characteristics and use of such systems in transmission
system planning was presented.

PROBLEMS
7 .1. Prove that the relationships in Eqs. 7.1 and 7 .2 are correct.
7.2. Replace the current source and resistor in the adjoint network of Fig. 7.lb by

'

_ _J
Automated Transmission System Expansion Planning 319

v(t) =V sin wt and an inductor of inductance L henrys. Show that element


variables satisfy Tellegen's theorem.
7 .3. For the voltage divider network below, calculate the sensitivity of the output
voltage V 0 w.r.t. the addition of the dotted new element.
R2 2 R> 3

,J---=--
- 12 --,. v.
Vo

lo

l
'-v--'
v2
11
I !Rs
u
__L
--.,-,
v. R
r
7.4. For the network in Prob. 7 .3, calculate the value of R 5 that renders an output
voltage of V 0 = 6 volts. Assume V 1 = 10 volts, R 2 = l ohm, R 3 = 0 ohm, and
R 4 = 1 ohm. Use the method described in Sec. 7 .2.
7.5. Complete the automated transmission system design problem in Sec. 7.2.
7.6. Extend the automated transmission design technique to handle the AC power-flow
model. (This is a difficult task.)

BIBLIOGRAPHY
1. Benjamin, D. M.: "Mini-computer Aided Analysis of the Load-Flow and Economic Dispatching of
Electric Energy Systems," Master's thesis, University of Florida, 1973.
2. Carter, L. W. et al.: Interactive Load-Flow and Stability Computations for Power Systems Studies,
PICA Proc., pp. 89-95, 1975.
3. Coombe, L. W. et al.: Interactive Load-Flow System,PICA Proc., pp. 96-103, 1975.
4. Director, S. W. and R. A. Rohrer: The Generalized Adjoint Network and Network Sensitivities,
IEEE Trans. Circuit Theory,, vol. 16, pp. 318-323, August 1969.
5. Fischl, R. and W. R. Puntel: Computer Aided Design of Electric Power Transmission Networks,
IEEE Conference paper C72-168-8, 1972.
6. Garver, L. L.: Transmission Network Estimation Using Linear Programming, IEEE Trans., vol.
PAS-89, pp. 1088-1097, September/October 1970.
7. Odom, M. D.: "Interactive Computer Aided Transmission Line Design," Master's Thesis,
University of Florida, 1976.
8. Puntel, W. R. et al.: An Automated Method for Long Range Planning of Transmission Networks,
PICA Proc., pp. 38-46, 1973.
9. Ptittgen, H.B. et al.: A Graphical Interactive Load-Flow Technique, PICA Proc., pp. 105-113,
1975.
10. Piittgen, H. B.: "A User Oriented Method for Transmission System Planning Using Interactive
Graphics," Doctoral dissertation, University of Florida, 1975.
11. Tellegen, B. D. H.: A General Network Theorem, With Applications, Philips Res. Rep., pp.
259-269, 1952.

_j____ _
INDEX

Annual demand forecast, 56-58 interactive graphics: characteristics of, 313


Annual refueling, 122, 12 7 configuration of, 314
Automated circuit design: editing, 314
control vector for, 298 file management, 314
gradient vector: calculation of, 299-301 use of, 313, 315-316
definition of, 299 interactive line design, 316-317
example, 301-303 interactive power-flow analysis, 315
parameter vector for, 299 problem formulation, 304-305
performance index, 299 state vector for, 305
problem formulation, 298
sensitivity analysis, 292-295
state vector, 298 Base load units, 63
Tellegen's theorem, 290 Bus admittance matrix, 159, 168
update algorithm, 299
Automated transmission planning:
adjoint network, 308 Contingency analysis:
control vectoJ for, 305 de power-flow method: angle equation, 227
de method for, 304 de model of, 225
example, 309-313 example, 228, 229
gradient vector: adjoint element models, 306 multiple outages, 227, 228
calculation of, 305 philosophy of, 224
capacity considerations, 307-308 Z matrix method: basic procedure, 230,231
cost considerations, 307-308 current changes, 2 3 2

321

J
322 Index

Contingency analysis, Z matrix method Forced outage rate, 67


(continued): Forecast uncertainty, 70-72
example, 235-237 Frequency and duration, 62
modified Z, 233, 234
removing elements, 232
Convolution equation, 77, 78, 81,112 Generation cost analysis:
Corporate model, 105-108 capacity cost: definition of, 97
earnings per share, 106-107 depreciation, 97
fixed charge coverage rate, 106-107 fixed charge rate, 97, 98
outside capitalization, 106-107 insurance, 97
retained earnings, 106-107 rate of return, 9 7
yield, 106 unit capacity cost, 97
Corridor selection, 153 capacity factor, 102
Cost of capital, 5, 6, 97 heat rate, 102
plant cost, 101-103
production cost analysis: fuel cost, 100
Demand, 18 O&M, 100
Diffusion coefficients, 14 3, 14 7 system cost, 104-105
Discounted multiple regression, 30 timing of additions, 102
Dynamic programming, 131-133 Generation systems:
fossil, 6-9, 139
gas turbines, 6
Economic commitment, 110-116 generation mix, 103, 104
Effective load, 76-80, 86, 87 hydro, 6..:9
Effective load-carrying capability, 84, 85 nuclear, 6-9, 139
Ele"ctric utility industry: reliability: approach to, 61-63
energy consumption, 1 generator models, 63-65, 86
financing, 4, 6 indices of, 73-76
growth characteristics, 1 for interconnected areas, 85-92
Environmental considerations: for isolated areas, 80-85
air pollution, 10,140, 141-147 load models, 68-70
cooling methods, 13 United States: capacity of, 8
cooling towers, 11-12 energy production of, 7
cost, 11-13 fuels for, 9, 10
water pollution, 11, 139, 140
Environmental costs:
air pollution: pollutants, 140 Household appliances:
sources of, 140 average power for, 21
ambient air quality: assessment of, 141 energy consumption of, 21
effective stack heights, 14 7 saturation levels of, 26
emission rate, 142
emissions, 14 2, 146 Installed capacity, 75
fuel sulfur content, 142
ground-level concentrations, 145, 147
SO 2 dispersion model, 143-145 Load, 18,69
Expected value of: , characteristics of, 20
demand not served, 73, 239 residential, 21
energy not served, 76 Load duration curve, 6, 69
load, 73, 74, 121 Load factor, 22
Load forecasting:
approaches to, 22
Failure rate, 66 demand: method for, 27
Fitting functions, 24, 29, 30 weather model, 28
Index 323

energy: commercial, 26 reactive power, 15 9


industrial, 26 real power, 159
residential, 26 state variable: control vector, 160, 162, 165
methodology: correlation, 25 disturbance vector, 160, 162
extrapolation, 24 output vector, 163,165
weather effects, 23 state vector, 160, 162, 165
Load shape, 70 variable constraints, 160
Long-line equations, 157, 158 Production costing:
Loss-of-load probability, 68 commitment schedules, 109
energy transactions: off-peak loading, 13 7
purchases, 136, 137
sales, 137
Maintenance schedules, 83
example of, 116-121
Markov process, 66, 67
fuel costs, 113
Midrange units, 63
generation system model, 108
Monthly demand forecast, 58-59
hydro unit models: pumped, 133
storage type, 131, 132
incremental fuel costs, 113, 115
Newton power flow: incremental heat rates, 110,111
acceleration.of, 188-189 load model, 108
compact storage, 189 nuclear fuel cycle, 122
example, 183-188 cost of, 125
Gaussian elimination, 177,181, 182 example, 128
Jacobian matrix, 173-177, 181 flow chart of, 126
solution procedure, 172-173 parameters, 124
state update equation, 173 segment demands, 112
Nonweather component forecast: segment energies, 112
discounted regression, 30-35
method for, 29
model statistics, 35, 36 Reactive power flow:
variance of, 36-38 analysis of, 205
Normal distribution, 41 compensation, 205, 207
discretization of, 4 2 TCUL, 207
Nuclear fuel cycle, 122-123 voltage control: example, 212-214
gradient vector for, 211
problem formulation, 210
Optimal power flows: solution procedure, 212
Kuhn-Tucker method: example, 201-204 Repair rate, 66
solution, 200, 201 Reserve margin, 88, 277
theory of, 199-200
problem formulation, 198
Outage capacity, 67 Seasonal demand forecast, 54-5 6
Sensitivity analysis:
example of, 196-198
Peaking units, 63 sensitivity equation, 190
Population, 2 matrices of, 191-195
Power-flow equations: Steady-state component models:
area in ter~hange, 15 9 power transformer: phase-shifting, 156
area model examples: interconnected, 170- voltage, 154-156
172 synchronous generator, 154
isolated, 166-170 transmission lines, 156-159
mismatch equations: interconnected, 164 Stochastic power flows:
isolated, 161, 163 confidence limits, 216, 21 7
324 Index

Stochastic power flows (continued): feasible flow patterns, 240


correlation effects, 218 indices of, 237
demand covariance, 21 7 labeling algorithm, 24 7-249
output covariance, 216 maximum-flow/minimum-cut, 246-24 7
output vector, 216 minimum-cut subsets, 258-263
problem formulation, 214-215 of multiareas, 276-286
so Ju tion procedure, 219 of single areas, 263-272
state covariance, 216 state classifications, 241-243
state vector, 215 unclassified states, 245
United States: cable miles of, 16
diagram of, 14
Temperature, 22, 27, 28, 39-48 overhead miles of, 15
Total weekly forecast:
density function of, 48-51
example of, 52-53
procedure for, 51-52 Weather load model, 28, 42-48
Transmission systems: Weather-sensitive forecast:
reliability analysis: approach to, 237-239 density function of, 48-51
augmentation algorithm, 249-251 example of, 46-4 7
critical states, 243-244 mean of, 42-46
decomposition rules, 252-254 method for, 32
element indices, 239, 254-255 weather variable, 39-41
element models, 238 transformation for, 41

Vous aimerez peut-être aussi