Sei sulla pagina 1di 227

RESERVOIR CAPACITY

AND YIElD
DEVELOPMENTS IN WATER SCIENCE, 9

advisory editor

VEN TE CHOW

Professor of Hydraulic Engineering


Hydrosystems Laboratory
University of fllinois
Urbana, fll., U.S.A.

FUR THER TITLES IN THIS SERIES

1 G. BUGLIARELLO AND F. GUNTER


COMPUTER SYSTEMS AND WATER RESOURCES

2 H L. GOLTERMAN
PHYSIOLOGICAL LIMNOLOGY

3 Y. Y. HAIMES, W. A. HALL AND H. T. FREEDMAN


MULTI OBJECTIVE OPTIMIZATION IN WATER RESOURCES SYSTEMS:
THE SURROGATE WORTH TRADE-OFF METHOD

4 J. J. FRIED
GROUNDWATER POLLUTION

5 N. RAJARATNAM
TURBULENT JETS

6 D. STEPHENSON
PIPELINE DESIGN FOR WATER ENGINEERS

7 V. HALEK AND J. SVEC


GROUNDWATER HYDRAULICS

8 J. BALEK
HYDROLOGY AND WATER RESOURCES IN TROPICAL AFRICA
RESERVOIR CAPACITY
AND YIElD

THOMAS A. McMAHON & RUSSEL G. MEIN


Department of Civil Engineering,
Monash University,
Clayton, Vic., Australia

ELSEVIER SCIENTIFIC PUBLISHING COMPANY


Amsterdam - Oxford - New York 1978
ELSEVIER SCIENTIFIC PUBLISHING COMPANY
335 Jan van Galenstraat
P.O. Box 211, Amsterdam, The Netherlands

Distributors for the United States and Canada:

ELSEVIER NORTH-HOLLAND INC.


52, Vanderbilt Avenue
New York, N.Y. 10017

Lihr:uy of ('()ngn's~ Cataloging in Publication nata

HCHahon, Thomas ]\quinas.


Reservoir capacity and yield.
(Developments in water science; v. 9)
Bibliography: p.
Includes index.
1. Reservoirs. I. Mein, Russell G., joint
author. II. Title. III. Series.
TD395.H24 1978 628.1'3 77-18704
ISBN 0-444-41670-6

ISBN 0-444-41670-6 (Vol. 9)


ISBN 0-444-41669-2 (Series)

Elsevier Scientific Publishing Company, 1978


All rights reserved. No part of this pUblication may be reproduced, stored in a
retrieval system or transmitted in any form or by any means, electronic, mechan-
ical, photocopying, recording or otherwise, without the prior written permission
of the publisher, Elsevier Scientific Publishing Company, P.O. Box 330, Amsterdam,
The Netherlands

Printed in The Netherlands


PREFACE

The text for this book has evolved from the notes written for a work-
shop held at Monash University in May 1975. The format of the workshop, the
first of a series on specific topics in water engineering, was about one
half lectures supported by printed notes, and one half exercises involving
both manual and computer applications ot the theory.

For this text, the printed notes have been revised and expanded, and
the exercises have been replaced by worked examples. Most of the latter have
been worked using the streamflow data of one river, the Mitta Mitta
(Appendix E), chosen because of its median value of variability with respect
to other Australian streams; compared to North American and European data
it would be classed in the high range of variability.

The aim of the text is to provide a comprehensive review and classi-


fication of most of the currently used storage-estimation procedures. The
essential features of each method are presented and limitations inherent in
the assumptions are discussed. Recommendations based on the results of
considerable research effort in this department over a period of several
years are made. The book is written for the practising engineer involved
in storage estimation and for graduate level study in this field.

The authors are indebted to the contributions of several post-


graduate students who have worked, or are presently working in the
Department under the supervision of the senior author. These are Dr. C. Joy,
Dr. G. Codner, Mr. G. Philips, Mr. C. Teoh, Mr. S. Fletcher and
Mr. R. Srikanthan. Discussions with Dr. R. Phatarfod of the Mathematics
Department at Monash University have also provided a strong source of
stimulation. The authors' colleague, Professor E. M. Laurenson, the third
of the original workshop instructors, has given unfailing support and
assistance in this and many other areas. Finally, for the production of the
manuscript itself, many people have contributed; in particular, Mrs. J. Helm
typed the final draft most ably, Mr. R. Alexander drafted the diagrams, and
Mr. D. Holmes completed the photographic reproduction. We are very
appreciative of their efforts.

T. A. McMahon, R. G. Mein,
Department of Civil Engineering,
Monash University.
September, 1977.
This Page Intentionally Left Blank
CONTENTS

Chapter 1 INTRODUCTION

1.1 THE DESIGN PROCESS 2

1.2 CLASSIFICATION OF RESERVOIR CAPACITY-YIELD PROCEDURES 3

1.3 PROCEDURES IN CURRENT USE 5

Chapter 2 DEFINITION OF TERMS 6

2.1 TIME INTERVAL 6

2.2 INFLOW DATA 6

2.2.1 Measures of Central Tendency 7


2.2.2 Measures of Variability 9
2.2.3 Measures of Skewness 9
2.2.4 Measure of Persistence 10
2.2.5 Typical Parameter Values 11
2.2.6 Standard Errors of Parameters 11

2.3 STORAGE TERMS 14

2.3.1 Active Storage 14


2.3.2 Within-year Storage 14
2.3.3 Carryover Storage 14
2.3.4 Conceptual Storages 14

2.4 RELEASE 15

2.5 RELEASE RULE OR OPERATING RULE 16

2.6 PROBABILITY OF FAILURE AND RELIABILITY 16

2.7 NOTATION 18

Chapter 3 CRITICAL PERIOD TECHNIQUES 19

3.1 CRITICAL PERIOD 19

3.2 METHODS WHICH INDICATE RESERVOIR FULLNESS WITH TIME 20

3.2.1 Mass Curve Method (Rippl Diagram) 20


3.2.2 Residual Mass Curve Method 22
3.2.3 Behaviour (or Simulation) Analysis 24
3.2.4 Semi-Infinite Reservoir 27
3.3 METHODS BASED ON RANGE 30

3.3.1 Hurst's Procedure 31


3.3.2 Fathy and Shukry 33
3.3.3 Sequent Peak Algorithm 33

3.4 METHODS BASED ON LOW FLOW SEQUENCES 35

3.4.1 Minimum Flow Approach 35


3.4.2 Alexander's Method 38
3.4.3 Dincer's Method 46
3.4.4 Gould's Gamma Method 49
3.4.5 Carryover Frequency Mass Curve Analysis 52

3.4.5.1 Overlapping Sequence Approach 52


3.4.5.2 Independent Series Approach 55
3.4.5.3 Independent versus Overlapping Series 57

3.4.6 Wi thin-year Frequency Mass Curve Analysis 59


3.4.7 Regional Within-year Storage Estimates 63
3.4.8 Bias in Mass Curve Frequency Analysis 64
3.4.9 Combining Carryover and Seasonal Storages
- Hardison's Approach 65

3.5 OTHER CRITICAL PERIOD METHODS 67

3.6 SUMMARY 67

3.7 NOTATION 68

Chapter 4 PROBABILITY MATRIX METHODS 71

4.1 GENERAL CLASSIFICATION OF MORAN DERIVED METHODS 71

4.2 A SIMPLE MUTUALLY EXCLUSIVE MODEL 73

4.2.1 The Discrete Equations for the Mutually


Exclusive Model - General case 76

4.3 A SIMPLE SIMULTANEOUS MODEL 78

4.4 COMPUTATION OF STEADY STATE CONDITION 79

4.5 DISCUSSION - MORAN TYPE MODELS 81

4.5.1 Further Modifications 82


4.6 GOULD'S PROBABILITY MATRIX METHOD 83

4.6.1 Procedure 83
4.6.2 Practical Considerations 90

4.7 RELATED PROBABILITY MATRIX METHODS 93

4.7.1 McMahon's Empirical Equations 93


4.7.2 Probability Routing 96
4.7.3 Hardison's Generali zed Method 97

4.8 OTHER MODELS 101

4.8.1 Me1entijevich 101


4.8.2 K1emes 102
4.8.3 Phatarfod 102

4.9 SUMMARY IDS

4.10 NOTATION IDS

Chapter 5 USE OF STOCHASTICALLY GENERATED DATA 107

5.1 TIME-SERIES COMPONENTS 108

5.2 HISTORICAL DEVELOPMENTS TO 1960 110

5.3 ANNUAL MARKOV MODEL 111

5.3.1 Practical Considerations 112

5.4 THOMAS AND FIERING SEASONAL MODEL ll4

5.5 MODIFICATIONS FOR NON-NORMAL STREAMFLOWS 115

5.5.1 Modifying ti 115


5.5.2 Moment Transformation Equations ll7
5.5.3 Normalizing Flows 121

5.6 TWO TIER MODEL 124

5.7 OTHER CONSIDERATIONS 125

5.8 MODEL VERIFICATION AND PERFORMANCE 126

5.8.1 Unrepresentative Streamflow Data 135


5.9 SIMULATION 135

5.9.1 When and How to Use Generated Data 136

5.10 GENERALIZED RESERVOIR CAPACITY-YIELD


RELIABILITY RELATIONS 140

5.10.1 Gould's Synthetic Data Procedure 140


5.10.2 Gug1ij's and Svanidze's Synthetic Data
Procedures 144

5.11 NOTATION 144

Chapter 6 QUANTITATIVE ASSESSMENT OF CAPACITY-YIELD


TECHNIQUES FOR SINGLE RESERVOIRS 147

6.1 CRITICAL PERIOD AND PROBABILITY MATRIX METHODS 147

6.1.1 Mass Curves and Minimum Flow (Waitt) 147


6.1.2 Alexander's Method 149
6.1.3 Overlapping Series Frequency Mass Curve
Method (Thompson) 150
6.1. 4 Independent Series Frequency Mass Curve
Method (Stall) 151
6.1.5 Gould's Probability Matrix Method 152
6.1.6 Further Comparison of Gould and
Behaviour Methods 153
6.1. 7 Summary 154

6.2 CAPACITIES BASED ON STOCHASTIC DATA GENERATION 155

6.3 RAPID RESERVOIR CAPACITY-YIELD PROCEDURES 159

6.4 SAMPLING ERROR OF STORAGE AND DRAFT ESTIMATES 164

6.5 RECOMMENDATIONS 169

6.6 NOTATION 170

Chapter 7 MULTI-RESERVOIR SYSTEMS 171

7.1 A TYPICAL PROBLEM 171

7.1.1 Traditional Solution 172


7.1. 2 Recycled Historical Sequences 174
7.2 STOCHASTICALLY GENERATED FLOWS 175

7.2.1 Key Station Approach 175


7.2.2 Principal Component Approach 176
7.2.3 Regression Method 176
7.2.4 Residual Approach 176
7.2.5 Multi-site Model Performance 177
7.2.6 Use of Generated Data 177
7.2.7 Application to Multi-storage Systems 178

7.3 TRANSITION MATRIX APPROACH 179

7.4 OTHER ALTERNATIVES 180

7.5 NOTATION 181

REFERENCES 181

APPENDIX A Procedure to adjust Storage Estimate for


Net Evaporation Loss 190

APPENDIX B Adjustment for Assumption of Independence of


Annual Flows 193

APPENDIX C Theoretical Justification of a Non-Seasonal


Markov Model 199

APPENDIX D Newton-Raphson Method for solving an Inexplicit


Variable 202

APPENDIX E Flow Tables for Mitta Mitta River 203

AUTHOR INDEX 205

SUBJECT INDEX 207


This Page Intentionally Left Blank
CHAPTER 1

INTRODUCTION

The storage required on a river to meet a specific demand depends on


three factors; the variability of the river flows, the size of the demand,
and the degree of reliability of this demand being met. As this and sub-
sequent chapters will show a large number of procedures have been proposed
to estimate storage requirements. This text is concerned with examining
and classifying these procedures with the aim of recommending the ones most
suitable for particular requirements.

In its simplest form the problem being tackled is shown in Fig. 1.1.
It is required to divert water from the stream with flow sequence Q(t) to
meet the demand of perhaps an urban area or of a rural irrigation scheme.
Alternatively it may be necessary to augment the low flow periods of the
river. In any event, the question being posed is: "How large does the
reservoir capacity CC) need to be to provide for a given controlled release
or draft DCt) with an acceptable level of reliability?" Other variations
of this question are possible (such as determining release for a given
capacity) but the basic problem remains unaltered; the relationship between
inflow characteristics, reservoir capacity, controlled release, and relia-
bility must be found.

Following definition of terms in the next chapter, Chapters 3-5


examine in detail all the common and some relatively unknown procedures for

Stream flow sequence


QW
. . Demand area

Controlled release
sequence OW
Reservoir with active

SPil~
storage capacity C

FIG. 1.1 An idealized view of the


reservoir capacity-yield problem.
2

solution of the single reservoir problem. The performance of several of


the methods is assessed in Chapter 6, where recommendations for use of
particular procedures are made.

The use of more than one reservoir storage to satisfy the demand adds
a significant degree of complication to the problem. The reservoirs may be
on the same stream, different streams, or not on any stream (e.g. pumped
storage). Additional complexity may result from topographical or other
constraints which restrict flow between reservoirs and thus reduce system
flexibility. The multi-reservoir problem is discussed in Chapter 7.

1.1 THE DESIGN PROCESS

In the early analysis of a water supply development, a number of


alternative darn sites would be investigated, not only for the construction
requirements but also from the hydrologic point of view. For such studies
and for hydrologic reconnaissance or regional reviews, quick and relatively
simple techniques for estimating the reservoir capacity-yield relationship
are required.

The methods which can be used for rapid assessment are designated as
preZiminary design techniques. Simplifying assumptions are often made;
for example, releases may be assumed to be constant, evaporation and sedi-
mentation losses ignored, the probability of failure may not be considered,
and the seasonal characteristics of the river flows may not be taken into
account. For these preliminary methods, accuracy is reduced for ease of
application.

After using preliminary design techniques to eliminate unsuitable


reservoir sites from consideration, the remaining few should be evaluated
using a finaZ design technique. These techniques are often more compli-
cated because they take into account most, or all, of the factors which
influence storage. Thus, properties of the river inflOWS, variation of
releases with season, the possibility of water restrictions, the effect of
evaporation, and the probability of not being able to meet the demand must
be realistically treated.

In the text, recommended methods are designated as being suitable as


preliminary or as final design techniques.
3

1.2 CLASSIFICATION OF RESERVOIR CAPACITY-YIELD PROCEDURES

Reservoir capacity-yield procedures can be classified into three main


groups although the distinction between groups is not always clear-cut.
The first group (critical period techniques) includes methods in which a
sequence (or sequences) of flows for which demand exceeds inflows is used
to determine the storage size. Those methods related to Moran Darn Theory
or similar procedure are included in the second group, part of which is
grouped under a general umbrella of probability matrix methods. The third
group consists of those procedures which are based on generated data. A
detailed classification along with author references is given in Fig. 1.2.
The methods are discussed in detail under these groupings in Chapters 3, 4
and 5, respectively.

Briefly, critical period methods are those in which the required


reservoir capacity is equated to the difference between the water released
from an initially full reservoir and the inflows, for periods of low flow.
For the procedures designated as mass curve, minimum flow, or range, the
storage is normally associated with the severest drought sequence in the
historical record. If historical data is used with these procedures, an
estimate of the risk of being unable to meet the design releases (proba-
bility of failure) cannot be made. In contrast, other critical period
methods enable the reliability of the reservoir to meet the demand to be
estimated.

The second group of procedures is considered to be a development of


Moran's Theory of Storage (1954, 1955, 1959). In essence Moran derived an
integral equation relating inflow to reservoir capacity and releases such
that the probable state of the reservoir contents at any time could be
defined. However, except for idealized conditions the solution was
intractable. Subsequently Morml considered time and flow to be discon-
tinuous variables ffild showed how reservoir capacity, release and inflow
could be related to each other by a system of simultaneous equations, but
the method has several shortcomings. Gould (1961) modified Moran's approach
to a general procedure of direct practical use to the water engineer. In
this context it is worth noting that a Russian, Savarenskiy, published in
1938 similar ideas to those presented later by Moran and Gould but it is
only recently that his contributions have become known in English technical
literature.

Although procedures for estimating reservoir capacity-yield relation-


ships using streamflow data generated by stochastic methods were first used
RESERVOIR CAPACITY - YIELD ANALYSIS

PROCEDURES BASED
ON DATA GENERATION

GENERAL
OVER YEAR
WITH FREQUENCY FREQUENCY Hazen (1914) Gould (19611
PROBABILITY Sudler (1927) Maass (1962)
Barnes (1954) Guglij (1969)
Rippl (1883) Waitt (1945) Hurst 0951.56.57.6S} Alexander (1962) Thompson l1 950) U.S.G.S. Wilson (1949)
Hurst (1965) Svandze (1964)
King (1920) Fathy & Shukry C1956} Dincer (c1960) Stall (1962) Hardison ('965) Law (1953.55)
Thomas (1963) Gould (1964)
(Sequent peak)

MORAN THEORY
I Melentijevich (1966)
(1967)
CONTINUOUS DISCONTINUOUS
(1976)
TIME TIME

Moran (1956)
I
Gani (1965)
CONTINUOUS DISCONTINUOUS Langbein
Gant & Prabhu (1958.59)
Gani & Pyke (1960.52) ~ VOLUMES - - - - - - - - - - - - - - - - - - - Hardison (r965)

Gani &. Prabhu (1957)


I Gould 1960
Prabhu (1958)
I McMahon (1916)
NON-SEASONAL SEASONAL
Ghosal (1959.60> I ------r= I
INDEPENDENT CORRELATED INDEPENDENT CORRELATED

Moran (1954) Lloyd (1963) Moran (1955) Dearlove & Harris (1965)
Lloyd &. Odoom (1964) Venetis (1969)

FIG. 1.2 A classification of reservoir capacity-yield procedures


5

more than sixty years ago, it was not until the advent of high-speed
digital computers in the sixties that such procedures became established in
engineering hydrology. Stochastic data generation is the basis of the third
group of storage-yield procedures.

It should be noted that many of the methods shown in Fig. 1.2 are
included for only their historical importance in the development of a
particular technique or groups of techniques; they are often impractical
or use unacceptable assumptions in their derivation.

1.3 PROCEDURES IN CURRENT USE

Very little published information is available on techniques currently


in use by water authorities around the world. A questionnaire survey of
Australian water authorities by the senior author in 1974 showed that, in
general, storage capacity designs were based on mass curve or simulation
analyses using historical streamflows. These two methods were used both
for preliminary and final design calculations. In about one half of the
cases, the probability of the reservoir not being able to meet the demand
was computed, although it was never used as the sole design criterion.
Data generation techniques have been used by about one half of the
Australian water authorities although more than that indicated their belief
in the potential of the method.

There is no reason to assume that the methods in current use in


Australia are any different to those in current use overseas.
6

CHAPTER 2

DEFINITION OF TERMS

This chapter is concerned with defining and explaining several of the


terms used in reservoir capacity-yield analyses. The meaning of some of
these terms sometimes differs from one author to another; it is therefore
important that the reader be clear as to which interpretation is used in
this text.

The definitions given include those for several statistical measures.


These are necessary to specify the characteristics of the river inflows to
the reservoir because these characteristics have a major influence on
storage requirements. Other important terms discussed in this chapter
include several storage terms, release, release rule, and definition of
probability of failure.

2.1 TIME INTERVAL

The time interval required for the inflow data depends on the size of
the storage and on the degree of accuracy required. For small storages
designed to provide water in excess of the river flow for only a month or
two in the year, daily flow data are required. For larger storages, monthly
data are usually adequate to define the variations of streamflow with season
(seasonality), although annual data can often provide sufficiently accurate
results for preliminary design estimates.

As a general rule, monthly data are used for most studies. With this
time interval the data processing time is not excessive, variations in
streamflow and releases throughout the year are adequately accounted for,
and records are readily available. A minor drawback in dealing with monthly
flow volumes is that the calendar months are not equal in length; the effect
of this on storage is small, however, and is usually ignored.

2.2 INFLOW DATA

It is not possible to predict the future sequence of flows of a


natural stream. All methods therefore use historical flow data or parameters
derived from it, and thus implicitly assume that these data are represen-
tative of the true streamflow characteristics. Hence, any value of storage
(or draft) estimated using historical data has a sampling error inherent in
it (see Sec. 2.2.6).
7

In the analytical procedures that follow it is assumed that inflows


into the reservoir occur as daily, monthly, or annual discrete events.
These will be available as measured data at the site in question, or will
have been estimated by either regression analysis (see, for example, Searcy,
1966, or Brown, 1961) or with a deterministic process model such as the
Stanford Watershed Model (Crawford and Linsley, 1966).

It is also assumed that the data have been checked for homogeneity
and consistency. In this context homoger:eity requires that identical flow
events in a time series are equally likely to occur at all times. Consis-
tency requires that there has not been any physical change at the stream
gauging station that might affect the recorded flows. Searcy and Hardison
(1960) discuss this aspect in detail.

The inflow data can be represented as a frequency distribution of


flows, such as those plotted for several rivers in Fig. Z.1. Often these
distributions can be approximated by standard theoretical distributions such
as the Normal, log-Normal, Gamma, Weibull, Extreme Value Type I, and log-
Pearson Type III. These distributions are defined by parameters of the
flows, for example, mean, standard deviation,and coefficient of skewness.
Another important flow parameter is the lag-one serial correlation which
describes flow persistence.

If the flow volumes in successive time intervals are designated as


Xl' x z' ... ' Xi' ... , xn ' the parameters are defined as follows;

2.2.1 Measures of Central Tendency

* arithmetic mean
n
L x.
1 1
x (2.1)
n

* median
The median is the middle value or the variate which divides the
flow frequency distribution into two equal portions.

The armithmetic mean is more commonly used because of its compu-


tational simplicity. In extremely skewed distributions, however, the
median will provide a better indication of central tendency.
8

27
24
DIAMANTINA RIVER 21
WARRAGAMBA RIVER
6 ~18
~ 5 ~ 15
4
t:
V g-:J 12
5-3 L.t 9
~ 2 6
u. 1 3
O+------r--~~-L~--r_--~ O~--r_-r--~~~~~~~
o 1 2 3 4 0 2 4 6
Annual flow (xl09 m3 ) Annual flow (xl0 9 m 3 )

8 8
7
MITTA MITTA 7 MEKONG
>- 6 RIVER ~ 6 RIVER
g 5 ~ 5
~ 4 5- 4
g- 3 ~ 3
L.t 2 u. 2
1
04-~~---r--~----~~'_--~
1 2 3 o 100 120 140 160 180 200
Annual flow (xl0 9 m 3 ) Annual flow (xl0 9 m 3 )

18
16
14 YARRA RIVER 7 BATANG PADANG
~12 ~ 6 RIVER
~10 ~ 5
5- 8 :J 4
C'
~ 6 v 3
u. 4 L.t 2
2 1+----'
O+---r_----._----~~~~~ O+---r_--~--~~~~--~
o 0.1 0.2 0.3 0.4 o 0.5 0.6 0.7 0.8 0.9
Annual flow (x 10 9 m 3 )

8
7
>- 6
g5
~ 4
g- 3
L.t 2
1
O~----._--~----._--~~~
0.6 0.8 1.0 1.2 1.4

FIG. 2.1 Frequency distributions of annual


'flows of selected rivers.
9

2.2.2 Measures of Variability

* standard deviation
1 n
s [-1
n- L (x.1 (2.2)

1 n 2 1
\
[-1 ( L x.1 - n X-2)]2 (2.3)
n-

For computational convenience Eq. 2.3 is preferred. The standard deviation


is the basic measure of variability.

* variance is the square of the standard deviation.

* coefficient of variation

cv (2.4)

The coefficient of variation is a dimensionless measure of variability and


is widely used in hydrology.

* index of variaJ;iUty
1 n
[-1
n-
L (log 10 x.1 (2.5)

The index of variability is the standard deviation of logarithms of flows.

2.2.3 Measures of Skewness

The lack of symmetry of a distribution is called skewness.

* coefficient of skewness
a
Cs (2.6)
53
n
n
where a
(n-l) (n-2) L ex.1 - x) 3 (2.7)

x 3 _ 3x
(n-l~ (n-2) [L L x2 + 2nx 3] (2.8)

This dimensionless measure relates to the third moment of the data


and is one measure defining the shape of the distribution. Data with
positive skewness are skewed to the right (Fig. 2.2).

Another measure of skewness used in hydrology is given by:

* Pearson second coefficient of skewness, given by

3 (mean - median)
(2.9)
standard deviation
10

Median I
>-
+'
(/) Mean:
s:::
Q)
"'C

>-
:!: I
..0
~
..0
o
~
a..
Magnitude Magnitude
(a) (b)

FIG. 2.2 Skewed distributions,


(a) Positively skewed;
(b) Negatively skewed.

Typically, flow distributions have a positive skewness as shown in


Fig. 2.1. The degree of skewness generally decreases as the time interval
of the data increases. Thus, the distribution for annual flows will
normally be less skewed than the distribution for monthly flow of the same
river.

2.2.4 Measure of Persistenct

Persistence is the non-random characteristic of a hydrologic time-


series. For example, a month with high streamflow will tend to be followed
by another of high flow rather than by one of low flow. This feature, which
is important in storage-yield studies, but which is not a parameter that can
be included in theoretical distributions, is quantitatively characterized
by the serial correlation coefficient. It indicates how strongly one event
is affected by a previous event.

* serial correlation
1 n-k 1 n-k n-k
---k
n- \ x.l x.l + k - (----k)2
L n- \ x.l L\ x.l + k
L
(2.10)
n-k
~_1_
n-k I
x2
i
_

where lag k serial correlation coefficient, and


lag between flow events.
Except for some procedures using stochastic data generation, lag one
serial correlation (k = 1 in Eq. 2.10) is the only lag considered in later
chapters.
11

It should be noted that in addition to Eq. 2.10 there are several


other procedures for calculating serial correlation of time-series data.
The characteristics of each are discussed by Wallis and O'Connell (1972).
For reservoir capacity-yield analyses the differences among the procedures
are of little importance and Eq. 2.10 is recommended.

Serial correlation is usually significant for monthly flow data. For


annual flow data the majority of streams do not have a serial correlation
coefficient significantly different from zero; however, there are still a
large number of streams with significant coefficients.

2.2.5 Typical Parameter Values

In order to illustrate this discussion dealing with parameter values,


monthly and annual flow parameters for five Australian and two South-east
Asian streams are tabulated in Table 2.1. As well, Fig. 2.3 shows, for
156 Australian streams, frequency histograms of coefficient of variation,
coefficient of skewness, and serial correlation coefficient of annual flows.
Various continental values are superimposed for comparison along with
estimates of continental mean annual runoff. (The latter values are taken
from Australian Dept. of National Resources, 1976.)

2.2.6 Standard Errors of Parameters

It must be emphasized that the parameter values defined in the previous


section are no more than estimates of the population values. An indication
of the magnitude of the error of the estimate is given by the standard error
of the parameter. These are defined as follows:

standard error of mean (2.11)

1
standard error of standard (2.12)
s/(2n)"
deviation

standard error of coefficient


(2.13)
of variation

standard error of coefficient (2.14 )


of skewness
1
standard error of serial (n - k - 1)"
(2.15)
correlation coefficient n - k

where s standard deviation of flow volumes,


n number of items of data, and
k lag between flow events.
TABLE 2.1 Annual, monthly seasonal and non-seasonal parameters for selected Australian and South-east Asian streams.

River
(Country) Para- Annual All Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
meter Months
(Area km 2 )

Diamantina x 7
. - 4.5+ 10.1 28.7 '8.1 11.9 6.0 5.4 2.5 0.3 0.6 0.9 1.0
(Australia) C 1.19 2.80 1.82 0.97 1.74 1.62 1.62 2.78 3.21 2.36 1.84 1.96 1.98 2.07
V
(115 000) C 1. 85 4.71 2.63 0.69 2.51 1.67 1.86 3.83 3.97 2.56 1.47 3.06 2.76 2.10
s
r 0.11 0.53 0.33 0.62 0.59 0.91 0.18 0.97 0.71 0.66 0.01 0.40 0.27 -0.16
Warragamba x 122 - 5.2 8.5 10.3 6.8 10.7 15.3 14.0 9.7 5.7 6.8 3.2 3.8
(Aus t Tali a) CV 1.11 2.14 1.60 2.34 2.62 2.28 1. 81 1. 83 1. 58 1.85 1.35 1. 76 1.55 1.72
(8750) C 2.67 4.73 2.76 5.76 4.43 5.58 2.83 2.69 2.74 4.59 2.93 3.85 2.77 3.39
s
r 0.30 0.45 0.20 0.76 0.58 0.20 0.39 0.53 0.53 0.42 0.30 0.57 0.27 0.32

Mi tta Mitta x 270 - 3.1 2.1 2.4 3.7 4.9 8.1 12.0 15.6 15.6 16.2 10.6 5.7
(Aus t ra1 i a) C 0.57 1.06 0.66 0.58 1.03 1. 74 1.13 1.16 0.87 0.76 0.52 0.56 0.62 0.69
V
(4710) C 1.50 1.91 1.79 1.12 3.27 4.88 4.11 3.17 1.34 1.30 0.88 0.58 0.62 1.72
s 0.06
r 0.71 0.58 0.60 0.54 0.86 0.80 0.63 0.59 0.65 0.73 0.65 0.79 0.61

Mekong x 480 - 3.2 2.3 2.2 2.1 3.0 6.4 13.0 23.1 20.7 12.7 7.0 4.4
(Thailand/ C 0.17 0.90 0.19 0.17 0.16 0.19 0.30 0.32 0.25 0.22 0.22 0.28 0.30 0.22
V
Laos) C -0.08 1.27 -0.14 -0.18 -0.00 0.20 0.96 0.71 0.49 -0.03 0.53 0.47 1.26 -0.00
(299 000) s 0.45 0.95
T 0.74 0.89 0.91 0.88 0.86 0.69 0.56 0.62 0.49 0.52 0.50 0.85

Yarra x 538 - 3.3 1.9 2.0 3.1 5.4 9.0 13.3 15.8 15.6 14.3 9.9 6.4
(Australia) C 0.40 0.99 0.89 0.72 0.74 1.24 1.05 1.03 0.58 0.51 0.48 0.61 0.75 1.11
(334) CV 0.77 1. 78 5.15 3.89 2.29 3.31 2.23 3.52 0.91 0.67 0.83 1.48 1.53 4.11
s
r 0.12 0.62 0.53 0.35 0.36 0.51 0.38 0.59 0.49 0.46 0.44 0.52 0.61 0.36

Bat.ang Padang x 1700 - 9.6 6.9 6.9 8.3 9.3 7.1 5.9 5.6 6.9 9.43 12.4 11.7
(Malaysia) C 0.18 0.40 0.32 0.27 0.28 0.27 0.29 0.23 0.22 0.21 0.27 0.31 0.34 0.35
V
(378) C 0.60 1.46 1.63 0.70 0.66 0.30 0.32 0.31 0.54 0.35 1.98 0.59 0.77 0.79
rs 0.47 0.60 0.74 0.62 0.75 0.66 0.53 0.47 0.56 0.25 0.58 0.59 0.48 0.71

x
I King
(Australia)
(451)
C
Cs
V
2340
0.19
-0.31
-
0.63
1.64
4.4
0.74
1.19
4.2
0.70
0.84
4.6
0.69
1. 74
8.2
0.39
0.32
9.6
0.54
1.03
10.7
0.58
1.12
11.6
0.35
0.24
12.0
0.42
0.46
11.0
0.39
0.92
9.3
0.45
0.55
7.8
0.49
0.90
6.6
1.11
5.11
r -0.11 0.32 0.16 0.24 -0.35 0.03 0.18 0.01 0.44 0.35 0.36 0.41 -0.05 -0.11
. Mean is expressed as depth of runoff in mm. x is mean; C
v
is coefficient of variation;
tMonth1y flows are expressed as percentage of mean. Cs is coefficient of skewness; r is serial correlation.
13

EUROPE & ASIA

NORTH AMERICA
25
AUSTRALIA
,"",500 >-
E ~ 20
Q)
E :::J
.... 400 a
::t: ~ 15
o u.
300
~

10

OL..LJL..LJL..f-I-U.L....&.+...L.L.L..L+L~.&..,

o 0.5 1.0 1.5 2.0


Annual coefficient of variation
~OTHER
CONTINENTS

25 ~AUSTRAlIA "AUSTRAlIA
>-
u 1./ NORTHERN
c: I' HEMISPHERE
~ 20 >-
20
CT u
~ c:
Q)
U.
15
g. 15
Q)
~
u.
10
10
-
5 - 5

o -[ n-n,
2 3 -0.2 0 +0.2 +0.4 +0.6
Annual coefficient of skewness Annual serial correlation

FIG. 2.3 Annual flow parameters for Australian streams


compared with other continental values.
(Lines indicate average values for the
continents indicated.)
14

As a general rule, the interpretation of these errors may be likened


to the interpretation of the standard deviation of a variable. If we
assume the parameters values are normally distributed (this is a reasonable
approximation in most circumstances) two-thirds of the values will lie
within one standard error.

2.3 STORAGE TERMS

2.3.1 Active Storage

The active storage of a reservoir is the water stored above the level
of the lowest offtake. It is thus equal to the total volume of water stored
less the volume of "dead" storage (the volume below the level of the off-
take). Throughout this text the terms storage and active storage are used
synonymously.

2.3.2 Within-year storage

Many small reservoirs fill up and spill on the average several times
a year. These reservoirs are constructed to provide water over a short
drawdown period of only a month or two of low flows. The estimation of the
storage required in this case is termed a within-year storage analysis.

2.3.3 Carryover storage

Where the reservoir fills up and spills only every few years on the
average, the water stored at the end of one year is carried over to the
next. This is called carryover storage. On the other hand seasonal storage
results from the fluctuations of inflows and outflows during the year. In
procedures that utilize only annual data, the seasonal effects are not taken
into account. In this text such procedures are known as carryover pro-
cedures; those concerned only with seasonal storage are known as within-
year procedures. Figure 2.4 illustrates the difference between these two
components.

2.3.4 Conceptual Storages

A finite storage is a conventional storage which can spill and run dry.
Not all reservoir storage-yield procedures assume finite storages. A semi-
infinite storage is one that can spill but never run dry. It is a
conceptual tool and the consequences of using it are discussed in Chapter 3.
Another conceptual storage is the infinite storage which can empty but
never spill.
15

FULL

....C
(JI

....4lC Carryover storage


0
(J

...
'0
...>4l
(JI
4l
a:

EMPTY
n n+6 n+12
Time (months)

FIG. 2.4 Illustration of carryover and within-year


storages showing the increase of storage
necessary to cater for seasonal fluctuations.

2.4 RELEASE

Release is the volume of controlled water released from a reservoir


during a given time interval. The term release is used synonymously in
this text with the terms yield, draft, outflow and regulation and describes
regulated flow from the reservoir. Spill is regarded as uncontrolled flow
from the reservoir and will take place only when the water stored in the
reservoir is above full supply level.

Release is often expressed as a percentage of mean flow having values


generally around 50 - 70%, and because of net evaporation losses rarely
exceeds 90%. Data from Australian water authorities indicate that the median
regulation of Australian reservoirs is 65%. Potential regulation varies
markedly across a continent. Hardison (1972) has shown that for mainland
United States potential regulation varies from 57% in Lower Colorado to
95% in Tennessee. In Australia, potential values vary from 70% in the arid
zone to 95% in Tasmania (McMahon, 1977).

To estimate the design capacity of a reservoir it is necessary first


to estimate the demands which will be placed upon the storage at some time
(or times) in the future. This is a difficult and uncertain task which is
beyond the scope of this text. A general discussion of ways of tackling
this problem is given in Linsley and Franzini (1974).
16

2.5 RELEASE RULE OR OPERATING RULE

Usually the volume of water released from a reservoir is equal to the


volume of water required (or demanded) by the consumers. However, there
may be periods when either the reservoir level is so low that the water
required cannot be supplied, or that prudence dictates that only part of the
water demanded is released from storage (for example, water restrictions for
an urban centre). Another factor in the decision may be the time of the
year and the expected inflows for the subsequent period. The way in which
releases are controlled is called the release or operating rule.

The simplest release rule is to supply all of the water demanded


[Fig. 2.5(a)]. In this situation, the draft is independent of reservoir
content and season. If there is insufficient water in the reservoir to meet
the required draft, the storage empties.

Release
~ 100 ~ 100
"0 "0
c:
('iJ
c:
('iJ

E
Q)
E
Q)
Cl 0 Cl 0
0 C 0 C
Water stored Water stored
(al (b)

FIG. 2.5 Example of two operating rules: (a) Simple operating


rule; (b) Operating rule with restrictions.

The more complicated release rule shown in Fig. 2.5(b) is typical of


that used by a metropolitan water supply authority. As the volume of water
stored in the headwater reservoirs decreases, restrictions are placed on
users so that demand falls and releases are lowered. It will be noted in
later chapters that few procedures can accommodate such an operating rule.
In the majority of reservoir capacity-yield techniques, constant draft is
assumed, that is, seasonal fluctuations in demand are not considered.

2.6 PROBABILITY OF FAILURE AND RELIABILITY

A number of definitions of probability of failure of a reservoir are


given in the technical literature. Probably the most common one defines
probability of failure as the proportion of time units during which the
17

reservoir is empty to the total number of time units used in the analysis.
Hence,

P
e ~ (2.16)

where p the number of time units during which the storage


is empty, and
N the total number of time units in the streamflow
sequence.

The corresponding definition of reZiabiZity is defined as:

R 1 - P (2.17)
e e

These definitions of probability of failure and of reliability are


not very realistic for most situations. A city water supply reservoir, for
instance, would never be permitted to empty; restrictions on releases would
apply long beforehand. An alternative definition sometimes used is similar
to Eq. 2.16 but where p is taken as the number of months during which
restrictions are necessary, that is, months during which the reservoir
cannot meet the demand under the adopted operating rule.

Another definition of reliability, voZumetric reZiabiZity, is equi-


valent to Fiering's (1967) performance index; it relates the volume of
water supplied to the volume of water demanded for the study period as
follows:

actual supply
R (2.18)
v demand

This definition has merit for overall reservoir performance, but can mask
the severity of any restrictions imposed.

The definition of probability of failure used for the remainder of


this text is Eq. 2.16 unless otherwise indicated. Although it may be
somewhat unrealistic in practice, it enables comparisons to be made between
different methods. The reader can, of course, use an alternative definition
to suit his purpose for most of the methods recommended.
18

2.7 NOTATION

a (n-l~ (n-2) l: (xi-x) 3 (Eq. 2.7)


coefficient of variation

cs coefficient of skewness

I index of variability (standard deviation of logarithms


v
of flows)
k lag between flow events under analysis

n number of items of data

N number of time units

p number of months reservoir is empty

p probability of failure (emptiness)


e
r serial correlation coefficient

lag k serial correlation coefficient

R reliabilitv = 1 - Pe
e ~

R volumetric reliability defined as water supplied


v
divided by water demanded
s standard deviation

X. ' 1 th perlo
fl ow volumes d urlng . d
1

flow volumes in successive time intervals

x mean flow
19

CHAPTER 3

CRITICAL PERIOD TECHNIQUES

The methods presented in this chapter typify two general approaches


to the reservoir capacity-yield problem. The first group of methods all use
the historical inflows and projected demand to simulate the volumetric
behaviour of the reservoir, that is, the state of fullness versus time. The
second group of methods has in common that only the periods of low flow
(droughts) in the record are used in the analysis.

Some of the methods of each group provide a reservoir size that will
not fail for the historical inflow sequence; the remaining methods allow
the user to determine the storage size for a given probability of failure.
However, all methods base the estimate of required storage capacity on
sequences of low flows and hence can be placed under the general heading of
critical period techniques.

3.1 CRITICAL PERIOD

A criticaZ period is defined as a period during which a reservoir goes


from a full condition to an empty condition without spilling in the inter-
vening period. The start of a critical period is a full reservoir; the end
of the critical period is when the reservoir first empties. Thus, only one

I a critical I
~ period ~
FULL -----------c---
I
1
I
I
1
I
1
I
1
I
I
I
I
I
I
I
I
I
I
I
I
EMPTY~-+-1-9-40-4r-1-9-41~~19-4-2-r-1-94-3~--1-94-4-+~19~4-5~~1~94-6~
Years

FIG. 3.1 Behaviour diagram showing critical periods.


{Mitta Mitta River; draft = 75%)
20

faiZure can occur during a critical period. Figure 3.1 gives an example
where there are two critical periods. Note that the remaining failures
(empty condition) of the reservoir in years 1945 and 1946 are not included
in a critical period.

This definition is not universally accepted. For example, the


U.S. Army Corps of Engineers (1975) define the critical period from the full
condition through emptiness to the full condition again and use the term
criticaZ draw down period to apply from fullness to emptiness.

3.2 METHODS WHICH INDICATE RESERVOIR FULLNESS WITH TIME

3.2.1 Mass Curve Method (Rippl Diagram)

The mass curve technique (following Ripp1, 1883) would appear to be


the first known rational method for estimating the size of storage required
to meet a given draft (see classification, Fig. 1.2).

15000
M

E
'"o
<II 10000
5:
o
;;::

.,
.>
(I)

-3 5000
A

E
<.)"
I I I I I

1957 58 59 1960 61 62 63 64 1965 66 67 68 1969


Year

FIG. 3.2 Reservoir capacity-yield analysis


by mass curve (Mitta Mitta River; draft 75%) .

The steps in the procedure, as shown in Fig. 3.2, are as follows:

(i) For the proposed dam site in question, construct a mass


or cumulative curve of the historical streamflows. (As a
general rule this curve should be constructed using
monthly flows.)

(ii) Superimpose on the mass curve the cumulative draft line


for the reservoir such that it is tangential to each hump
of the mass inflow curve.

(iii) Measure the largest intercept between the mass inflow curve
and the cumulative draft line.
21

In the example shown in Fig. 3.2 the intercept C2 is greater than


intercept C and so the design capacity would be taken as C . (No allowance
l 2
has been made for evaporation.)

Thus, a reservoir of capacity C , full at time zero (Jan. 1957), and


2
experiencing the historical inflows and assumed drafts, will be full at A,
begin to empty from A to B, fill again from B to D, spill from D to
E, just empty when it reaches F and begin refilling again. Period EF is
the critical period in this example.

Assumptions:

(i) The reservoir is full at time zero and consequently at the


beginning of the critical period.

(ii) In using the historical flow data it is implicitly assumed


that future flow sequences will not contain a more severe
drought than the historical flow sequence.

Limitations:

(i) The draft is usually taken to be constant. Seasonality in


demand could be included but restrictions in releases (as a
function of storage content, for example) are difficult to
handle.

(ii) Storage capacities estimated by mass curve procedures


increase with increasing length of record. It is therefore
difficult to relate storage size to economic life.

(iii) It is not possible to compute a storage size for a given


probability of failure. On the other hand the rank 1
drought in a record of length N has an implied probability
of failure of P = 1/(N + 1) as shown in Sec. 3.4.5.

(iv) The method of analysis does not take into account net
evaporation losses. If required, an additional amount of
storage has to be added to the mass curve estimate to
cover this loss. Details of a procedure to do this are
given in Appendix A.

In this context it is of interest to note that King (1920) modified


Rippl's method to include evaporation losses and incident rainfall in the
mass balance of water over the critical period.

Attributes:

(i) The procedure is simple and it is widely understood.


22

(ii) It takes into account seasonality, serial correlation


and other flow parameters insofar as they are included
in the historical flows used in the analysis.

EXAMPLE 3.1

Using the mass curve method, find the storage required on the Mitta
Mitta River (Appendix E) to meet a draft of 75% of the mean flow.

* * * * *
Because the scale of the graph required to plot the entire 34 years
of data for the Mitta Mitta River is too large for this page size, only a
portion of the plot (the last 13 years) is shown in Fig. 3.2.

The slope of the draft line is 79.6 x 106m3jmonth determined as 75% of


the mean monthly flow (106.1 x 10 6m3). After drawing the draft line
tangential to major peaks, the storage is determine from the maximum inter-
cept between the draft line and the mass curve. From Fig. 3.2 the required
storage is 1110 x 10 6m3 , corresponding to a critical period of 3 years and
4 months. (Note that the storage has not reached a full condition at the
end of the period of record.)

3.2.2 Residual Mass Curve Method

The residual mass curve is a slightly more complicated version of the


mass curve. The steps in the procedure (see Fig. 3.3) are:

(i) Subtract the mean flow from each flow value of the record.
(If monthly data are used, subtract the monthly mean;
for annual data use the annual mean.) The resulting flows
are called residual values.

(ii) Plot the residual mass curve (cumulative values) and


superimpose the cumulative draft line (expressed as a
residual) such that the draft line is tangential to
each hump of the residual mass curve.

(iii) Measure the largest intercept between the mass inflow


curve and the draft lines.
23

3000
'"E
~ 2000 Draft-

~ 1000
:5
"'"'
III
1 mean flow
= 26.5 x 106 m 3
Imonth

E
_ -1000
III
:::
-0
-~ -2000
0::

-3000

1935 1940 1945 1950 1955 1960 1965 1969


Year

FIG. 3.3 Reservoir capacity-yield analysis by


residual mass curve (Mitta Mitta River;
draft = 75%).

Overall the assumptions, limitations and advantages of the residual


mass curve approach are similar to the mass curve procedure. However, it
does have an added disadvantage in that the method is not as easily under-
stood as the straightforward mass curve procedure. On the other hand, it
permits the use of a cumulative inflow scale many times larger than that
used for a mass curve of the same data and thus improves graphical accuracy.

EXAMPLE 3.2

Using the residual mass curve method, determine the storage required
on the Mitta Mitta River (Appendix E) to meet a draft of 79.6 m3/month
(75% of the mean flow).

* * * * *
The mean monthly flow (106.1 x 10 6m3) is subtracted from the monthly
flows to determine the "residual" flows. The cumulative residual flows
(residual mass curve) are plotted in Fig. 3.3.

To determine the slope of the draft line, the mean flow rate is
subtracted from the draft rate. Thus the slope of the draft line is
(79.6 - 106.1) x 106m3/month. Lines of this slope are drawn tangent to the
humps on the curve and the maximum deficit determined. The answer is
1110 x 10 6m3 as shown in Fig. 3.3. Note that the scale of the residual
mass curve allows more accuracy than a mass curve drawn on the same scale.
24

3.2.3 Behaviour (or Simulation) Analysis

In behaviour or simulation analysis, the changes in storage content


of a finite reservoir are calculated using a mass storage equation thus:

(3.1)

subject to 0 ~ Zt+l ~ C

where Zt+l storage at en d 0 f t h e t t htlme


. . d
perlo
(= storage at beginning of t+lth period),
storage at beginning of tth time period,
. fl ow d
ln '
urlng t t htlme
. . d,
perlo

re 1ease d urlng t t h . . d,
tlme perlo
net evaporation loss from reservoir during
tth time period,
L other losses, and
t
C acti ve storage capacity. (I f the active storage
is reduced by sedimentation during the life of the
reservoir this needs to be taken into account.)

The usual time period is one month, but other periods can be used.
The net evaporation loss is the difference between the evaporation from the
proposed reservoir and the evapotranspiration from the proposed reservoir
site and depends on the surface area of water in the reservoir. Other
losses are comparatively small and are usually neglected.

The steps in constructing a behaviour diagram are as follows:

(i) Arbitrarily choose a reservoir of active capacity C, and


assume that it is initially full, that is, Zo = C.
(ii) Apply Eq. 3.1 month by month for the whole historical
record. D can either be assumed constant or to vary
t
seasonally. It could also vary as a function of Zt'
(~Et may be included if necessary.)

(iii) Plot Zt+l against time on a monthly time scale (see Fig. 3.4).

(iv) Compute the probability of failure (Eq. 2.16) by dividing


the number of time periods for which the reservoir is
empty by the total number of time periods.

(v) If the probability of failure is unacceptable, choose a


new value of C and repeat the steps. (The process is
thus an iterative one to determine the storage size for
a particular design probability of failure.)
2S

~~
u
~
~
100

~ r ~
-
~
u 60
0
~
40
~
~
~
~
0
m20
0
1935
\ ~
1940 1945 1950 1955 1960 1965
LL
1969
Years

FIG. 3.4 Behaviour diagram for Mitta Mitta River


(draft = 75%) with a storage such that the
number of months empty is approximately 5%
of the total number of months.

It is possible to set a storage size for which the reservoir just


empties once for the period of historical data. This storage is the same as
that found using the mass curve and residual mass curve methods.

Assumptions:

(i) The reservoir is initially full.

(ii) The historical data sequence is representative of


future river flows.

Limitations:
(i) The reservoir is initially full. The significance of this
on storage size can be checked by examining a behaviour
diagram for various starting conditions. Analysis based on
generated data suggests that at least 100 years of streamflow
are required for some rivers before the effect of the
initially full assumption can be ignored (McMahon et aZ.,1972).

(ii) The analysis is based on the historical record. The


sequencing of flows may not be representative of the
population of flows.

(iii) Non-continuous records cannot be easily handled because of


difficulties of assigning the initial reservoir condition
after a break in the streamflow data.
26

(iv) Demands (and hence releases) that are related to growth


rates in time (for example, increased urban water demands
through population increases) are not easily taken into
account because of the difficulty of relating the demand
in a future year to a specific year of the historical flow
record. Some guidance in dealing with this problem is
given in Chapter 7.

Attributes:

(i) The historical behaviour analysis is a simple procedure


and displays clearly the behaviour of the stored water.
The behaviour diagram can be readily understood by
non-technical people.

(ii) The procedure takes into account serial correlation,


seasonality and other flow parameters insofar as they
are included in the historical flows used in the analysis.

(iii) The procedure can be applied to data based on any time


interval.

(iv) Not only can seasonal drafts be easily taken into account,
but also complicated operating policies can be modelled.
For example, there are real difficulties in including
release as a function of demand (itself a function of some
climatic variable) and the contents of the reservoir.

EXAMPLE 3.3

Use the behaviour analysis method to compute the storage required on


the Mitta Mitta River (Appendix E) to supply a uniform draft of 75% of the
mean flow with a probabi Ii ty of failure of 5%.

* * * * *
Mean flow 106.1 x 106m3/month (Appendix E).
Draft 79.6 x 106m3/month.

The method is an iterative one. A storage capacity is chosen, the


inflows are routed through the storage, and the probability of failure (the
number of months that the reservoir is empty divided by the total number of
months) is calculated using Eq. 2.16.
27

The progressive results are tabulated below.

Storage Corresponding probability


Trial Estimate of failure
(l06 m3) (%)
1 1000 0.73
2 800 4.17
3 700 6.37
4 750 5.39
5 760 4.90
6 755 5.39

Note: A probability of failure of exactly 5% cannot be obtained because


it is calculated on a monthly basis; 5% of 34 years is 20.4 months.

Thus, for a steady draft and storage initially full, a capacity of


760 x 10 6m3 is required. A plot of the behaviour diagram is given in
Fig. 3.4.' (One can note that if the same analysis is followed for an
initial condition of empty, the storage required is found to be 825 x 10 6m3 .)

3.2.4 Semi-Infinite Reservoir

The mass curve method does not enable a probability of storage failure
to be calculated. Perhaps the earliest method proposed to calculate failure
probabilities was to run the streamflow sequence through a semi-infinite
storage (Hazen, 1914). (As defined in Sec. 2.3.4., a semi-infinite storage
can spill but never run dry.) Unfortunately, this method leads to error in
calculation of the probability of failure of a finite reservoir as shown
below.

The depletion diagram for a semi-infinite reservoir is a behaviour


diagram without the lower boundary (empty) condition. It is shown as the
full line in Fig. 3.5. Drawdown or depletion of the reservoir content ABC
occurs because inflows are less than outflow. At C, inflows begin to
dominate and the reservoir depletion decreases until spill occurs at E.

A depletion or behaviour diagram for the finite reservoir is shown in


the same figure as a broken line. For this situation, a reservoir of some
finite capacity has been assumed and the lower boundary condition taken into
account. The effect of this is shown as BC'E'. The probability of failure
using the semi-infinite depletion diagram is given as:
28

II>
)(

<I:
2
c
;;::: .!!1
)(
c: <I:

E ~c:
II>
(/) u::
Full

.:: c
o 0
> '';:;
~ II>
11>-
II> 0-
Il> .,
0:0

FIG. 3.5 Comparison of behaviour diagrams using


finite and semi-finite storages.

p . (3.2)
seml

where i.
1
number of months of emptiness, and
N total in months of the record used in the analysis.

From the behaviour diagram of the finite reservoir the probability of


failure is found as:

(3.3)

where mi number of months of emptiness based on the


behaviour diagram for the finite reservoir.

From Fig. 3.5, it is seen that Eii > Lm . Thus the assumption of
i
semi-infinite storage always results in the overestimation of the probability
of failure, and therefore the overestimation of the required behaviour
(finite) storage for a given probability of failure. Barnes (1954) argued
that this overestimation provided a safety factor which was necessary for
urban water supply design. For six locations on Australian rivers, storage
estimates for two conditions of draft and one probability of failure are
compared in Table 3.1 using finite and semi-infinite behaviour analysis.
The table shows that this overestimation is about 20%.
29

TABLE 3.1 Percentage increase in required reservoir capacity


of semi-infinite storage over finite storage for 5%
probability of failure using monthly data.

Rivers
(Australian gauging 50% draft 90% draft
station no.)

Yarra
14 26
(229102)
Murrumbidgee
14 12
(410008)
Lachlan
( 412010) 17 14

Warragamb.a
41 21
(212240)
Namoi
14 27
( 419007)
Burdekin
37 6
(120090)
-- --
Mean 23 18

EXAMPLE 3.4

Use the semi-infinite storage procedure to calculate the storage


capacity required to supply 75% with no failures on the Mitta Mitta River
(Appendix E).

* * * * *
The procedure is the same as a behaviour analysis but with an
arbitrarily high storage capacity. The behaviour diagram for the semi-
infinite case is shown in Fig. 3.6. From the figure it can be seen that for
the period of record a storage of capacity 1110 x 10 6m3 or more will not
fail.

To calculate the storage size for a given, say 5%, probability of


failure with this method, a horizontal line is drawn such that for 5% of the
time period the reservoir contents are below it. The line is shown in
Fig. 3.6, giving the storage corresponding to 5% probability of failure as
900 x 10 6m3 . (Note that the storage determined in this way is an over-
estimate as shown in Sec. 3.2.4.)
30

1000

E 800
"'o
., 600
.
a
Cl

400
ci5 This line gives 5%
proba bility of failure
200

o~~~ __~=-~~__~~~~__~~
1935 1940 1945 1950 1955 1960 1965 1969
Year

FIG. 3.6 Behaviour diagram for semi-infinite


storage on Mitta Mitta River for 75%
Draft (Example 3.4).

3.3 METHODS BASED ON RANGE

In Sec. 3.2.2 we discussed how a residual mass curve is used to


determine storage size. Consider the four residual mass curve diagrams in
Fig. 3.7. In Figs. 3. 7(a), 3.7(b) and 3. 7(c), it is seen that the range (R)

( a) (b)

(e) (d)

FIG. 3.7 Definition of range (R). For (a), R = R1;


(b) R = Rl; (c) R = Rl + R2 ; (d) R = Rl + R2
31

of the cumulative sums of departures from the mean streamflow will be equal
to the critical period storage estimate assuming the draft to be equal to
the mean annual streamflow. However, in Fig. 3.7(d) the range R Rl + R2,
but the required storage will be equal to the larger of Rl or R2 (This
characteristic will be examined further in our discussion of the sequent
peak procedure.) The range defined as R in Fig. 3.7 is the basis of a
number of important reservoir capacity-yield procedures. These are now
discussed.

3.3.1 Hurst's Procedure

Hurst was concerned with storages on the River Nile. It was an


unusual problem in that he was dealing with large equatorial lakes where a
big increase in volume could be obtained by small structures at lake
outlets. Consequently, storage sizes capable of providing very high regu-
lation (approaching 100%) were being examined. Hurst's work (1951, 1956,
1965) is concerned more with mathemati cal experiment rather than theory,
but from it, he was able to formulate a general solution to the reservoir
capacity-yield problem.

From an analysis of some 700 natural time-series including stream-


flows, rainfalls, temperatures, pressures, tree rings and lake varves,
Hurst found that the range could be related to the length of record as
follows:

R/s (N/2) K (3.4)

where R range defined as the sum of the cumulative


departures from the mean,
s standard deviation of the time-series data,
N length of the time-series, and
K exponent, and was found by Hurst to have a mean
value of 0.72 and a standard deviation of 0.09.

As an aside it should be noted that for a purely random time-series


process K tends towards 0.5 (Feller, 1951). The fact that natural time-
series yield on the average values of K greater than 0.5 is known as the
Hurst phenomenon. This aspect is considered again in Sec. 5.7.

In developing his general solution to the storage-yield problem, Hurst


computed K and the storage size to "guarantee" a uniform draft Bless
than mean x using a mass curve analysis for records of natural phenomena.
The generalized storage relations (for all observations) are (Hurst, 1951):
32

(i) for draft mean (x)


C R (3.5)

(ii) for draft < mean


log) 0 (C/R) - O.OS - 1. 05 (x-B)/s (3.6)
,
or C/R 0.94 - 0.96 [(x_B)/S)2 (3.7)

where C required reservoir capacity,


R range given by Eq. 3.4,
B draft parameter which is defined as
x-B
draft (3. S)
s
s standard deviation of flow, and
x mean flow.

Examining these relationships using only Hurst's river flow data,


Joy (1970) found that the generalized curves were not lines of best fit and
that the results exhibited large scatter as shown in Fig. 3.S. Hurst
admi tted the generali zed nature of his results in replying to discussions
by Chow (Hurst, 1951, p. SOO). A further limitation of the procedure is
that only the rank 1 values of storage are considered and no other estimates
of risk of storage failure can be estimated. Nevertheless, Hurst's work is
of monumental significance because of the importance of the value of K in
stochastic data generation models (Sec. 5.7).

'/
CIA = 0.94 - 0.96 [G- B}/s] 2 log1Q(C/RJ = -0.08 -1.05 Cx-B}/s

1.0 1.0
0.8
0.8 0.6
a::
~ 0.6 ---
(J 0.4
(J
o o
o 00 00 o
0.4 8",8 o
0 0.2 o o
6' 0 80 o o o
o o
0.2 o
o '0 0
o
o o
o
0.2' 0.4 0.6 0.8 1.0 0.6 0.8 1.0
~x-BJ/sJ'/2 (x-BJ/s

FIG. 3.8 Comparisons of Hurst's generalized equations


with data for river flows used in their
derivation.
33

3.3.2 Fathy and Shukry

Fathy and Shukry (1956) agreed with Hurst that for 100% regulation
the equivalent mass curve reservoir capacity was given by Hurst's range R.
But for lower drafts they disagreed entirely with Hurst and developed their
own method. However, as shown by Joy (1970) their technique is no more than
a mathematical representation of the minimum flow approach of Waitt' (1945)
given in Sec. 3.4.1.

3.3.3 Sequent Peak Algorithm

Thomas (Thomas and Burden, 1963) proposed the sequent peak algorithm
as a method to circumvent the need to locate the correct starting storage
as is required in the mass curve procedure. In essence, the storage size
calculated using the sequent peak procedure is equal to the range of the net
cumulative inflow (inflow less draft) estimated for the historical record
concatenated with itself.

The sequent peak procedure is described by Fiering (1967) as follows.


Given an N year record of streamflow at the site of a proposed dam and the
desired drafts, it is required to find a reservoir of minimum capacity such
that the design draft can always be satisfied if the flows and drafts are
repeated in a cyclic progression of cycles of N periods each. The solution
requires only two cycles. The reason for the second period can be seen in
Fig. 3.7(d) where the storage by mass curve is max (R I , R2) but the range
equals RI + R2' This anomaly arises because the two periods in which there
are net outflows are broken by a net inflow period so that no continuous
depletion of the reservoir occurs as it does in the three other examples in
Fig. 3.7.

The steps in the sequent peak procedure are:

(i) Calculate X. - D, (that is, flow - draft) for all


l l
i = 1, 2, ... ,2N and calculate the net cumulative
t
inflow L (X.- D.) for t = 1, 2, ... ,2N. (Column 4 in Table 3.2).
i=l l l

(ii) Locate the first peak (local maximum), PI, in


the column of cumulative net inflows (column 4).

(iii) Locate the sequent peak, P2 , which is the next peak


of greater magnitude than the first.

(iv) Between this pair of peaks find the lowest trough


(local minimum), T I , and calculate PI - T I .
34

(v) Starting with P2' find the next sequent peak, P 3 ,


that has magnitude greater than P2

(vi) Find the lowest trough, T2 , between P2 and P3


and calculate P2 - T2'

(vii) Starting with P3' find P4 and T3, calculate P3 - T3'

(viii) Continue for all sequent peaks in the series for


2N periods.

(ix) The required reservoir capacity is:

C (3.9)

TABLE 3.2 Sequent peak procedure for Mitta Mitta River


(75% draft = 79.6 x 106m3/month).

Flow Flow l: (Flow


Month Remarks
-79.6 - 79.6)
10 6m3 10 6m3 10 6m3
( 1) (2) (3) (4) (5)

0 PI First peak
1 56 -23.6
-23.6
2 32 -47.6
-71. 2
3 32 -47.6
-llS. S
4 38 -41. 6
-160.4
5 31 -4S.6
-209.0 Tl Lowest point between
6 113 33.4 peaks
-175.6
7 lS9 109.4
-66.2
8 529 449.4
383.2
9 217 137.4
520.6
10 152 72.4
593.0
11 SO 0.4
593.4
12 S4 4.4
597.S P 2 Second peak higher
13 53 -26.6 than first
571. 2
14 27 -52.6
5lS.6
35

Limitations and Attributes

The sequent peak algorithm was designed for use with replicates of
generated data rather than with a single historical record and so several
of the limitations attributed to it here are a consequence of using a single
historical record.

Unlike the mass curve procedure it can handle variable drafts so long
as they can be specified without reference to reservoir content (for example,
seasonal drafts). But like the mass curve, the required storage is a
function of record length and we are unable to determine capacities other
than those that satisfy the rank 1 low flow sequence.

The use of generated flows allows a probability of failure to be


indirectly obtained. For a single sequence the method calculates the
storage required to meet the worst drought in the period of record. Thus,
for a record length of N years the implied probability of failure is
P = l/(N+l) as discussed in Sec. 3.4.5. N can therefore be chosen to give
the desired value of P. For replicated sequences, the required storage is
the average of the individual storage estimates.

Because the storage size is estimated from two cycles of a record of


length N years, the computed value is equal to or greater than the equi-
valent mass curve value. But this difference will only occur when case (d)
of Fig. 3.7 occurs. An analysis by Srikanthan (private communication, 1977)
of 16 Australian streams (discussed in Sec. 5.8) showed that there were
differences between the sequent peak and mass curve values in only two
streams. The differences were respectively 68% and 83%.

Computationally the algorithm is fast and in contrast to behaviour


analysis a trial and error solution is not required. Like the behaviour
analysis, it uses the historical data directly and consequently the effects
of seasonality, serial correlation and other flow parameters are taken
directly into account. A description of its use may be found in Fiering
(1967) and ~10ss (197l).
36

EXAMPLE 3.5

Use the sequent peak method to determine the storage capacity


required on the Mitta Mitta River (Appendix E) to provide a controlled
draft of 75% without failures.

* * * * *
The mean monthly flow for the Mitta Mitta is 101.1 x 10 6m3; therefore
75% draft is 79.6 x 106m3/month.

The first few months of the calculation are shown in Table 3.2. The
draft has been subtracted from each monthly flow and the residuals accumu-
lated. The first summed residual value is itself a peak (designated Pi)'
because the succeeding values decrease. The next peak is shown as P2 .
The storage required to cover this flow sequence is Pi Tl where Tl is the
lowest cumulative residual value between the peaks, that is,
o- (-209,0) = 209 x 10 6m3 .

After traversing through the historical data twice, the successive


P-T values are found to be:

209, 957, 960, 242, 1032, 227, 287, 104, 204, 220, 194, 197, 157,
16, 414, 323, 283, 682, 1107, 957, 960, 242, 1032, 227, 287, 104,
204, 220, 194, 197, 157, 16, 414, 323, 283, 682 (x 10 6m3 ).

The required storage capacity is the largest value, that is,


lllO x 10 6m3 .

This example shows the importance of running the historical flows


through twice. If only one sequence had been used, the maximum storage
capacity would not have been found.

3.4 METHODS BASED ON LOW FLOW SEQUENCES

3.4.1 Minimum Flow Approach

This approach was proposed by Waitt in 1945. From the historical


streamflow record, the lowest sub-sequences of flows of various durations
are selected and plotted as mass flow against corresponding duration.
Constant drafts are plotted as straight lines on this 50-called drought
curve graph. The critical storage is given by the ordinate at the origin
when the draft line is-tangential to the drought curve. The steps which
are related to Fig. 3.9 are as follows:
37

Drought curve
based on lowest
recorded flows

12 0 12 60
Time (months)
CDESIGN

FIG. 3.9 Reservoir capacity-yield analysis by


minimum flow approach.

(i) From the historical sequence, select the lowest (rank 1)


flows for various durations - these are usually taken as
12, 24, 36, 48, etc. consecutive months. Intermediate
values could also be selected which would allow for
seasonal inflow variations.

(ii) Plot only the rank 1 flow volumes against the corresponding
durations on arithmetic graph paper.

(iii) Superimpose the draft line such that it passes through


the origin.

(iv) Measure the largest intercept between the inflow curve


and draft line. This gives the storage estimate
denoted by CCRIT'

Waitt recommended that in practice one full year's supply be added


to the storage size as a safety fac,tor. This is illustrated in Fig. 3.9,
and the reservoir would be designed for a capacity equal to COESIGN'
In effect the procedure without the added safety factor gives the same
answer as the mass curve procedure.
38

EXAMPLE 3.6

Using the minimum flow procedure find the storage capacity required
to supply 75% draft (no failures) on the Mitta Mitta River (Appendix E).

* * * * *
A search of the flow records for the Mitta Mitta produces the
following:

Duration Lowest total flow for


(months) that period (l06 m3)

5 26
10 238
20 618
30 1356
40 2077
50 3341
60 4275
70 4925
80 5730
90 6615
100 7704

These are plotted on Fig. 3.10 with the draft line (75% draft 'is
79.6 x 106m3/month). The largest intercept between the two curves is the
storage requi red. From Fig. 3.10 this occurs for a duration of 40 months
(= critical period); the storage can be measured from the graph or
calcylated thus:

40 x 79.6 - 2077 1107


say 1110 (x 10 6m3).

Note: Waitt (1945) proposed a "safety factor" of 12 months draft. This


would give a storage capacity of 1107 + 12 x 79.6 = 2062 (x 10 6m3).

3.4.2 Alexander's Method

Neither the mass curve nor the minimum flow approach nor the range
provides an estimate of the probability of storage failure. Alexander (1962)
extended these earlier approaches by developing a series of drought curves
for different probabilities of occurrence and from these derived generalized
storage-regulation-probability curves.
39

8000

7000

;; 6000
E
<0
0 5000

~ 4000
0
;:
] 3000
0
I-
2000

1000

0 10 20 30 40 50 60 70 80 90 100
Months duration

FIG. 3.10 Storage capacity for the Mitta Mitta River


(75% Draft) determined using the minimum
flow procedure.

In effect Alexander's approach may be considered as a development of


Waitt's drought curve procedure. Alexander assumed that annual streamflow
could be represented by a Gamma distribution from which he specified the
mass inflows as shown in Fig. 3.11 for various recurrence intervals.
Recurrence interval is the reciprocal of probability thus:

T recurrence interval (years) (3. 10)


r
=
pro b a
b'l't
l l y of occurrence

'"'"
<1l
III
>-
III
>
.;:;
:J
<.>
III
'"c:
0 /
<.>
/
c: /
/
.E'" /
Historical flows
~ "",e .,,/
rank 1
<;)<~"'//////
0
;:
.S
''""
<1l
~ 0 2 4 5 6
Time (years)

FIG. 3.11 Introduction of recurrence interval


(and hence probability) to the minimum
flow approach (Alexander, 1962).
40

Draft lines were also superimposed on the drought curves as Waitt had done
and storage capacity estimated as the difference between cumulative inflows
and outflows.

It is necessary to understand a little about the Gamma distribution


and its properties.

Two Parameter Gamma Distribution

The probability function of the two parameter Gammadistribution is


defined as follows:

a-I e- x/S
f(x) = x (3.11)
Sa T (a)

where x the flow value,


a the shape parameter,
S the scale parameter, and
T (a) the Gamma function
a-l -x
I x e dx
0

To estimate the probability of a flow of magnitude x being less than


some value X given that the flows are Gamma distributed, one calculates:

X
Prob (x ~ X) I f(x) dx
o
a
1 IX x a-l e
-xiS
dx (3.12)
S T(a).O

As shvwn diagrammatically in Fig. 3.12, this equation estimates the area of


the shaded portion of the probability distribution.

FIG. 3.12 Probability of x being less than X.


41

To calculate the parameters a and B of the Gamma distribution, Thorn's


(1958) maximum likelihood method is used, thus:
1+(1+4A)~
3
& (3.13)
4A

where A ~ (that is, log of mean - mean of logs.).


e
There is a small correction that should be applied to a (see Thorn's
paper) but for Alexander's application it is not significant. The scale
parameter is calculated as follows:

X
B= (3.14)
a

where x = mean annual streamflow.

Detai ls of A lexa:nder IS Method

In developing his method, Alexander used a property of the Gamma


distribution which states that the sum of n independent Gamma variables
with parameters a and B is a Gamma variable with parameters:

an na (3.15)

and as x nx
n
xn nx
Sn S (3.16 )
an na

where an shape parameter of n consecutive years flow, and


Bn scale parameter of n consecutive years flow.

Thus assuming annual streamflow to be Gamma distributed and knowing


the shape parameter of the annual flows for a given stream, Eq. 3.15
allowed the shape parameter to be calculated for all n consecutive years
flow. As illustrated in Fig. 3.13, Alexander could then calculate the
drought curves for a'given probability using Eq. 3.12.

Using this concept he developed a series of curves relating constant


regulation or release to reservoir capacity as a ratio of a mean annual flow
and recurrence interval for a shape parameter of unity (Fig. 3.14).
Storage capacities for streams with shape parameters that are not unity are
obtained by dividing the estimated capacity for ~ =1 by the correct shape
parameter.
42

"'"
">
:g
<.J

"
VI
C
0
<.J
C

~
~
0

~
VI
VI

::2'"
0 1 2 3 4
Time (consecutive years)
ii 2ii 3ii 4&
Shape parameter of Gamma distribution

FIG. 3.13 Development of Alexander's drought curve.


(Shaded area indi cates probabi l i ty of n
consecutive years flow being less than
value indicated on ordinate.)

The computational steps are as follows:

(i) Decide on the percentage regulation with respect to


the mean flow (D). This must be constant.

(ii) Compute the shape parameter ~ of the Gamma distribution


for annual flows at the site in question using Thorn's
procedure Eq. 3.13.

(iii) Decide on the probability of failure or recurrence


interval CTr) of the critical design period.

(iv) Use Fig. 3.14 (Alexander, 1962, Figure 6) to estimate


'] and CPt knowing Tr and D where ,] reservoir capacity
as ratio of mean annual flow for ~ = 1, and CPj = critical
period (years) for a = 1.
(v) Compute required reservoir capacity (in volume units) thus,
'1
reservoir capacity C =~ x (3.17)
a
(vi) The corresponding critical period (in years) is
CP 1
critical period CP = (3.18)
a
43

0
I\.G~\)
1
JI
~O ~O ~~
",
1 :'1(
g0 ~'v -
\'~ I,<;, '1-2;..-X, f"-. ,0 I vi.s> 0-"-
~",v ~KI"\Y/Kl><~o ~o (~
,... 80 '?-JG:rK
~,1 1\ / K\V pj~Dr
.I~ k '~OOI
0'
,#_ " ...

-
Q
c-~l"> :(\ \V\ /K ~t?v~#
;1 \ V\ ) /V)~0~v
c
o
i
:::J
en
Q)
a:
70
V
')..
,
V
/
I
I
/\ /
/ / /.1 V/
/ //-~ / / ~~
r,
I J J j

..
Q)
en
cu
c
60
j
V J V/ Vj ~
VI 'i v,
/ V I V ~ v/VI
Q)
u
~
/ Il
cf 50
/ / 1/ ~
I I Regulation 0 with
capacity 1'1 giving I
J /
probability of failure'-
V~ 1/(h
V V
40 II 1/T of r for =1 a.

~
V /I
Y
f/ /
30
) ~ ~ / ~ ~I
0.5 1 2 3 4 5 6 7 8 g
Reservoir Capacity (1'1)

FIG. 3.14 Alexander's reservoir capacity-regulation-


probability curves (Alexander, 1962).
44

Assumptions:

(i) Annual flows are independent.

(ii) Annual flows are Gamma distributed.

(iii) The reservoir is full at the beginning of the


critical period.

Limitations:

(i) Regarding assumption (i), approximately 25% of Australian


streams exhibit lag one annual serial correlations that are
significantly different from zero at 5% level of significance.
Since the median value of Australian and Northern Hemisphere
serial correlations are similar (Fig. 2.3), it would be
expected that the same proportion of serial correlation values
would be significant in the Northern Hemisphere. If a
procedure has been used that is based on historical data
(for example, behaviour analysis) serial correJation effects
would have been implicitly taken into account. Therefore,
for consistency with procedures using the low flow periods
directly, an adjustment to the storage estimate should be
made to account for the assumption of zero serial correlation.
Unfortunately, no comprehensive study of the effect of the
independence assumption of annual flows on the reservoir
capacity-yield-probability relationship is available.
However, some guidance is given in Appendix B.

(ii) Annual flows for some streams may not be Gamma distributed
although studies of annual flow data for Australian streams
suggests that this distribution is generally suitable
(McMahon, 1977). As the shape parameter increases, the Gamma
distribution approaches the normal distribution. Thus for
the more normally distributed Northern Hemisphere streams,
the Gamma distribution should be satisfactory.

(iii) Because annual data are used, seasonal variations are


ignored and storage estimates will tend therefore to under-
estimate storages based on monthly data, particularly for
short duration critical periods.

(iv) Releases must be assumed constant.


45

(v) The reservoir must be assumed to be initially full and


no variation from this condition is possible.

(vi) The method of analysis does not take into account net
evaporation losses. If required, an additional amount
of storage has to be added to the computed value to cover
this loss (see Appendix A) .

Attributes:

(i) The procedure is rapid and is simple to perform.

(ii) It is sufficiently accurate to be used as a preliminary


design or reconnaissance technique, but not as a final
design tool.

(iii) Through estimating the recurrence interval of minimum


flows, failure probabilities are obtained.

EXAMPLE 3.7

Determine the storage required for 75% draft and 5% annual probability
of failure for the Mitta ~1itta River (Appendix E) using Alexander's
procedure.

* * * * *
From the data (Appendix E) and working in units of 10 6m3 ,

mean flow x 1274


loge x 7.1496

Sum of logs (base e) of annual flows = 238.042


From Eq. 3.13 A

7.1496 (238.042)/34

0.1484
4A ~
1 + (1 + 3)
4A
3.53.

From Fig. 3.14 using recurrence interval Tr 20 years (5% probability of


failure) and percentage draft D = 75%, we find
46

Tl 2.1, and

Storage capacity C [Eq. 3.17]

This capacity has to be adjusted to compensate for the annual serial


correlation of 0.06 (Appendix E). The adjustment factor, from Fig. B.l,
is approximately 1.06.

Required storage 758 x 1. 06


800 (x 10 6 m3 ).

Length of critical period

CP =~
3.53 = 3 . 4 years.

3.4.3 Dincer's Method

The derivation which follows is essentially due to Professor T. Dincer,


Middle East Technical University, Ankara (C.H. Hardison, personal communi-
cation, 1966).

Consider a sequence of annual flows, which are assumed independent,


wi th a mean x and standard devi ation s.

n-yearly mean x nx (3.19)


n

n-yearly standard deviation s


n
(3.20)

As a consequence of the Central Limit theorem, the distribution of


n consecutive annual flows approaches normality as n increases.
Therefore, the lower p-percentile flow is given by:
,
nx - z n ~s (3.21)
p
where n-year flow with a probability of occurrence of p%,
that is, for p% of time n-year flow ~ qn,p' and

z standardized normal variate at p%.


p
47

During a critical period,

required storage outflow - inflow

C D (3.22)
n,p n

where C depletion of an initially full storage at end of


n,p
an n-year period during which time the n-year flow
(q ) has a probability of occurrence of p%, and
n,p
Dn constant draft from the reservoir over n years
Dnx (3.23)
where D constant draft as ratio of mean annual flow.

substituting Eqs. 3.21 and 3.23 in Eq. 3.22,

C Dnx nx + z
n,p p
nx (D-l) + z (3.24)
p
To obtain the length of the critical period and the maximum required
storage differentiate Eq. 3.24 with respect to n and equate to zero giving
z 2
CP ~ C 2 (3.25)
4(l-D)2 v
z 2
and hence
C -E- C 2
T (3.26)
x 4(l-D) v

where CP length of the cri ti cal drawdown period in years,


C annual coefficient of variation,
v
C maximum required storage in volume units, and
T = maximum required storage expressed as a ratio
of mean annual flow.

Assumptions, Limitations and Attributes

In addition to the basic critical period assumptions (an initially


full reservoir and only one failure during the critical period), the above
derivation is based on a number of other assumptions, namely:

(i) that the critical period is large enough so that the n-year
flows will tend towards a Normal distribution;

Cii) that annual flows are independent; and

(iii) that the draft rate is uniform.

For non-normally distributed annual flows the procedure will tend to


overestimate storage need, but this is offset by a tendency for under-
estimation due to initially full conditions and the condition of no repeated
48

failures without spill. In addition, the procedure is based on annual data


and seasonal need is not taken into account. An adjustment to correct for
the independence assumption is given in Appendix B. Notwithstanding these
limitations, it has been found that the procedure provides reasonably
reliable storage estimates at high drafts (greater than 50%).

EXAMPLE 3.8

Use Dincer's method and the annual flow data for the Mitta Mitta
River (Appendix E) to estimate the storage capacity required for a
controlled release of 75% with a probability of failure of 5%.

* * * * *
From Eq. 3.26,
c
x

We have x 1274 (x 10 6 m3 ) (Appendix E)


z 1.65 from tables of the Normal distribution
p
for a one-tail (p = 5%)
D 0.75
731
cv - 1274
0.57 (Appendix E)
x

Therefore, storage capacity is:

c (1274) (1.65)2(0 57)2


4(1 - 0.75) .
1127 (x 10 6 m3 ).

The adjustment factor for annual serial correlation of 0.06


(Appendix E) is obtained from Fig. B.l as 1.06. Therefore required storage
is 1130 x 1.06 or 1200 (x 10 6m3).

The length of critical drawdown period is given by Eq. 3.25:


z 2
CP --L-
4(1-0)2
cv 2
(1.65)2 2
4(0.25)2 (0.57)

3.5 years.
49

3.4.4 Gould's Gamma Method

This method (Gould, 1964) can be thought of as a modification of


Dincer's method (Sec. 3.4.3) although it was derived independently. It uses
the fact that, while parameters for the Normal distribution are easy to
calculate and probability tables for it are readily available, the Gamma
distribution usually is a better approximation to the distribution of annual
flow data. The procedure is therefore to use the Normal dis!ribution for
the calculations, and then to apply a correction to approximate the Gamma
distribution.

The mean and variance of a one parameter Gamma distribution, G(c),


are equal and equivalent to the shape parameter, say c. It is possible to
convert the mean, X, and standard deviation, s, of a Normal distribution to
Gamma units by dividing them both by s2/x. The resultant Normal distri-
bution will have the same mean and variance, that is Gamma units.

For a Normal distribution Dincer (Sec. 3.4.3, Eq. 3.26) showed that
z 2
c -1'..- c 2 X
4(l-D) v

where C required storage in volume units,


D draft as ratio of mean annual flow,
C coefficient of variation,
v
x mean annual flow, and
z standardized Normal variate at p%.
p
Substituting for x and C 2 with Gamma units gives
v
z 2
cy -p-
4(l-D)
(3.27)

where C storage volume in' Gamma units.


Y

Gould argued that the difference d between the lower p percentile


flow of a Gamma c distribution and that of a Normal distribution with mean
and variance both equal to c is approximately constant for a given value
of p over a large range of shape parameter c. Values of d are given
in Table 3.3. Because of the constancy of d for a given p, according to
Gould the critical period for a Gamma distributed inflow is the same as that
for normally distributed inflow with the same mean and coefficient of
variation, and that the storage required for an inflow that has a Gamma
distribution is d Gamma units less than that required for a Normal
50

TABLE 3.3 Values of z and d.


p

Lower p percentile z d
value p

0.5 3.30 d not constant


1.0 2.33 1.5
2.0 2.05 1.1
3.0 1. 88 0.9
4.0 1. 75 0.8
5.0 1.64 0.6
7.5 1. 44 0.4 } Gould's Gamma method
not recommended for
10.0 1. 28 0.3 use in this range.

distribution. In other words, as shown diagrammatically in Fig. 3.15,


inflows for. a Gamma distribution will be greater than for the Normal case
and the storage should be decreased by d Gamma units.
Hence,
z 2
C ~ - d (3.28)
Y 4(l-D)
z 2
~ d
T (3.2fl)
Y 4c(l-D) c

where T required storage divided by mean annual flow


y
in Gamma units, that is, storage / c.

The required storage can be converted from Gamma units to units of


volume as a ratio of mean annual flow by multiplying the right-hand side of
s2 1
Eq. 3.28 by (~) -=-.
x x

Magnitude of n consecutive year flow.

Wd is approximately constant as c varies

FIG. 3.15 Diagrammatic illustration of the difference


between inflows for the Normal and Gamma
distributions.
S1

z 2
P C 2 _ dC 2 (3.30)
.4(T=D) v v

where T required storage divided by mean annual flow


in volume units per year.

Thus Eq. 3.30 is the same as Eq. 3.26 except for the correction
term dC 2.
v

Assumptions, Limitations and Attributes


By assuming flows to be Gamma distributed rather than normally
distributed, Gould overcomes one of Dincer's major assumptions, but in
doing so, introduces a small approximation relating to the difference
between the Normal and Gamma distributions. Other points concerning Dincer's
procedure also apply to Gould's Gamma method. It will be noted in
Chapter 6 that this procedure provides very good estimates of storage over
the whole range of practical interest.

EXAMPLE 3.9

Use Gould's Gamma procedure and the annual flows for the Mitta Mitta
River (Appendix E) to estimate the storage required on this river to meet
a draft of 75% with a probability of failure of 5%.

Equation 3.30
r4(6 z
*
2
* *

C
*

2
*

L ) v

We have z 1.64 from tables of the Normal distribution


p
for p = 5%,
D 0.75,
d 0.6 from Table 3.3 corresponding to z = 1.64,
P
= 731
:>... 0.57 (Appendix E).
1274 "
x
1. 64 2
T
[ 4(0.25)-
0.68.

Therefore storage capacity 0.68 x


(0.68 x 1274)
866 (x 10 6m3).
52

Adjustment for annual serial correlation of 0.06 (Appendix E) is obtained


from Fig. B.l as 1.06.

Required storage 866 x 1. 06

920 (x 10 6m3).

3.4.5 Carryover Frequency Mass Curve Analysis

Carryover frequency analysis is generally similar to the previous


techniques. Instead of using theoretical distributions like the Gamma dis-
tribution the inflow probability is defined by empirical procedures. Two
of these are now discussed. In both of them, the strearnflows are ranked;
in the first, overlapping sequences are adopted while independent sequences
are used in the other approach.

3.4.5.1 Overlapping Sequence Approach

In a flow sequence of N' months, there are (N' - n + 1) overlapping


sub-sequences of n consecutive months' duration. The absolute severity
(in a low flow sense) of any sub-sequence is measured by its rank. The
severest sequence of a particular duration, as measured in terms of minimum
flow, is ranked 1, the next worst sequence is ranked 2, etc. These
concepts are illustrated in Fig. 3.16. To obtain a relative measure of
severity, commonly rankings are translated into recurrence intervals
(return periods) or probabilities, using Eq. 3.31 .
.,
"C
2
""E'"
Cl

~
o
;;::

Tim e (months)

~ Ranked
~ 4 monthly

~
sub - sequences
from total

~ sequence of
24 months
~
.
~
FIG. 3.16 Rank and sub-sequence concepts (for a 4-monthly
sub-sequence, the lowest consecutive four months
of flow are ranked as no. 1).
53

p
m
probabi l i ty of occurrence (3.31)
N'+l
where m rank of the sub-sequence, and
N' length of record in months.

By considering a range of durations one can relate duration and


associated low flow values to probability of occurrence, and hence construct
drought curves (cumulative inflow curves) for given levels 9f severity
(probability of occurrence) as we have done in the previous procedures. A
draft line is then superimposed on the inflow curve and storage is computed
as in the minimum flow approach. This method was proposed by Thompson
(1950) .

Assumptions and Limitations

Being a critical period technique, two assumptions - an initially full


storage and no repeated failures - are implicitly made. These result in the
procedure tending to underestimate the number of behaviour failures (that is,
failures that would result if the same flows were run through a behaviour
analysis for a storage of similar capacity) and the reservoir capacity is
underestimated.

This underestimation is compensated to various degrees by the


"clustering" effect inherent in ranking overlapping sequences. Due to
overlap, any sequence tends to contain the last (n - 1) months of the
previous sequence. Thus, radically different values of mass flows cannot
be expected, especially for sequences of long duration. As a result,
sequences of similar rank tend to cluster around the same portion of the
streamflow record and define the same historical values. As a consequence,
the procedure tends to overestimate the number of behaviour failures in
certain portions of the streamflow-record and the reservoir capacity is
overes timated.

Thus there are two compensating discrepancies in this technique. It


underestimates the number of behaviour failures due to the neglect of
repeated failures. However, due to clustering it overestimates the
number of behaviour failures in certain portions of the streamflow record.

For short critical periods, clustering is minimal; consequently there


is little or no compensation for repeated failures, and the technique will
tend to underestimate the required behaviour storage.
54

On the other hand, for long critical periods, clustering is severe.


If the critical period is sufficient long, overestimation of the required
storage may be so severe that there is little difference between the rank
and higher ranked values.

EXAMPLE 3.10

Find the storage required on the Mitta Mitta River (Appendix E) to


provide for a 75% draft and a probability of failure of 5% using the over-
lapping sequence procedure.

* * * * *
The procedure requires searches through the data file (Appendix E) to
find the lowest, second lowest etc. cumulative flows corresponding to a
number of durations.

The iowest (that is, rank 1) sequence has a recurrence interval


(To liP).
NI + 1 409
From Eq. 3.31, T -1- 409 months
o m
34 years.

The second lowest (that is, rank 2) sequence has a recurrence interval
of 409 204.5 months
2
17 years.

Thus, to get a recurrence interval of 20 years, it is necessary to


plot these two values on log-probab.ili ty paper and interpolate.

The results of the searches for different durations are as follows:

Duration Lowest Total Second Lowes t


Months Flow (10 6m3) Total Flow (106m~

10 238 244
20 618 689
30 1356 1395
40 2077 2125
50 3341 3346
60 4275 4287
70 4925 4927
80 5730 5898
90 6615 6667
100 7704 7733
55

After interpolating between the two figures to get the flow corres-
ponding to a 20 year recurrence interval for each duration, the results are
plotted in Fig. 3.17.

The greatest intercept between the draft line (79.6 x 106m3/mont~ and
the 1 in 20 year flow line is the required storage, 1070 x 10 6m3 , and corres-
ponds to a critical period of 40 months. In practice, the storage estimate
needs to be increased by approximately 20% to account for bias due to cross-
nesting of flows (Sec. 3.4.8).

7000

6000

ME 5000
'"o
';; 4000
Storage required
~ 3000 = 1070
u::
1 in 20
2000

1000

I ! I I ! I I

o 10 30 40 50 60 70 80 90 100
Duration of sequence (months)

Fig. 3.17 Storage estimate for Mitta Mitta River


(75% Draft) using the overlapping series
approach (Example 3.10).

3.4.5.2 Independent Series Approach

Based on the work of Hudson and Roberts (1955), Stall and Neill (1961)
presented another method for estimating the frequency distribution of the
flow series. From the (N' - n + 1) overlapping series of n months
duration, the lowest flow is selected and ranked as 1. However, unlike the
overlapping series approach, the next step is to neglect all other sequences
which overlap the rank 1 sequence. The rank 2 sequence is then selected
from the remaining flow record. Higher rankings are determined in a similar
manner with none of the flows for any particular rank overlapping the
sequence of flows for any other rank.

Stall (1962) extended this technique to storage-yield analysis in a


similar manner to the previous techniques. The steps in his procedure
follow and can be related to Fig. 3.18.
56
8000

7000

;:; 6000
E
~ 5000
)(

; 4000
0
u:: 3000
Storage required
2000 = 980 (x 106 m 3 )
1000

0 10 70 80 90 100
Sequence duration in months

FIG. 3.18 Reservoir capacity-yield analysis by


Stall's procedure (Independent Series
Approach) for Mitta Mitta River (Draft 75%
and probability of failure = 5%).

(i) Convert the rank of each n month sub-sequence to


recurrence interval by Eq. 3.32.

N + 1
T. (3.32)
l m

where T. recurrence interval (in years) for an


l
independent series,
m the rank of the sub-sequence, and
N the length of the streamflow record in years.

The interpretation of this equation must be understood in the


light of the duration of the event under review (n months in
duration). The average recurrence interval of an n-month
event of rank m is defined as
l2N +
n
(3.33)
m

where l2N/m total possible number of independent events


in the sequence.

The recurrence interval Tn has units of n months and is


converted to years thus:
n
T. Tn
l 12
N + n/12
m
N + 1
(3.34)
m
57

An alternative to Eq. 3.32 is

N + 1 - 2c
T. (3.35)
1 m- c

where c is a function of the distribution of the data


under analysis, and in general is taken as c = 0.3 giving

T.
N + 0.4 (3.36)
1 m - 0.3

This latter equation is recommended for use by United States


Corps of Engineers, Hydrologic Engineering Center (see United
States Army Corps of Engineers, 1975 and Beard, 1962).

(ii) Inflow volumes (mass values) are plotted against duration


for design recurrence intervals.

(iii) The appropriate constant draft line is superimposed on the


inflow curves.

(iv) The required capacity is estimated as the largest interval


between the inflow curve and the demand line.

In Fig. 3.18 a storage of capacity of 980 x 10 6m3 will maintain the


constant 75% draft for inflows with recurrence intervals less than one in
20 years. In practice, the storage capacity would need to be increased to
cover evaporation losses and a bias resulting in part from cross-nesting of
the mass curves as explained in Section 3.4.7. Storages should be increased
by 20% to take this bias into account.

3.4.5.3 Independent versus Overlapping Series

From Eq. 3.31 the equivalent recurrence interval in years for a rank
m event by the overlapping procedure is given as follows:

NI + 1
T
o
---:i.2Iil year s (3.37)

where T recurrence interval (in years) for an overlapping series.


o
From Eqs. 3.32 and 3.37 it is seen that for a particular flow sequence
to have the same recurrence interval in both procedures, a value of given
rank in the independent series must be approximately 1/12 of the rank in
the overlapping series.

Recall that in the overlapping series clustering is more severe for


long duration sequences than for short ones. Thus, for long duration
sequences the effect of removing overlapping flows in the independent
58

series is to remove from further analysis a large number of flows of rank


similar to that of the ranked flow. In other words, for the independent
series and for long duration sequences, we are ranking from a decreasing
pool of flows, whose actual flow values tend to become rapidly high. This
effect is reduced for short duration sequences.

It is not possible to predict when the rank of independent series


(multiplied by 12) will underestimate the overlapping series rank. Because
of this factor the independent series will tend to have lower flows than
the overlapping series for the corresponding rank. On the other hand, the
effect of clustering results in the overlapping series having the lower
flows particularly for sub-sequences of long duration. Somewhere between
these two opposite effects, the results will be equivalent.

Thus for critical periods of short duration, the independent approach


will tend to overestimate the equivalent storage of the overlapping series.
On the other hand, for long duration critical periods the independent
approach underestimates the overlapping series storage estimates.

It is difficult theoretically to relate the independent approach to a


behaviour analysis. Empirically, it has been found that the independent
series overestimates the behaviour storage at low drafts and underestimates
it at high drafts. This conclusion fits the general observations made with
respect to the overlapping series.

From the above discussion it may be concluded that, at least from a


theoretical point of view, the carryover frequency mass curve techniques
described above are inadequate for several reasons. Unfortunately the mag-
nitudes of these inadequacies are not easily quantified and this fact needs
to be kept in mind if the procedures are used in design. The independent
series approach, slightly modified, is recommended by the United States Army
Corps of Engineers (1975).

EXAMPLE 3.11

Use the independent series method to determine the storage required


on the Mitta Mitta River (Appendix E) to meet a 75% draft with a failure
rate of once every 20 years on the average.

* * * * *
The procedure is to search through the data record to determine first
the lowest total flow for a given duration. Then, excluding those months
59

from the record and not allowing "straddling" of them by a sequence, the
remaining flow record is searched for the second lowest total flow for that
duration. This is repeated for several durations, as shown in the following
tab Ie.

Duration Lowest Flow Second Lowest Flow


(Months) (l06 m35 (10 6m3 )

10 238 244
20 618 837
30 1356 1480
40 2077 2367
50 3341 3763
60 4275 4301
70 4925 5386
80 5730 6030
90 6615 6827
100 7704 7798

From Eq. 3.32 the lowest flow for a given duration has a recurrence
interval of
34 + 1
--1- 35 years.

The second lowest flow has a recurrence interval of


34 +
--2- = 17.5 years.

Thus, to get a 20-year recurrence interval, interpolation (on log-


probability paper) is necessary. The resultant values are plotted on
Fig. 3.18 and the largest intercept between the draft line (79.6 x 10 6m3)
and the inflow curve gives the storage, 980 x 10 6m3 corresponding to a
critical period of 30 months.
In practice, the storage should be increased by 20% to account for bias due
to cross-nesting of flows (Sec. 3.4.8).

3.4.6 Within-year Frequency Mass Curve Analysis

Annual low flow frequency curves are the basis of the United States
Geological Survey approach to within-year storage requirements (see, for
example, Hardison and Martin, 1963). The steps in this procedure are as
follows:
60

(i) For duations 1, 7, 15, 30, 60, 120 and 183 consecutive
days, the mi.nimum flows for each year of the historical
record are determined.

(it) These flows are ranked (separately for each duration) in


order of magnitude, rank 1 being given to the lowest flow.

(iii) For each flow a recurrence interval (or probabili ty of


occurrence) is estimated by the formula:

1 N + 1
T (3.38)
r P m

where recurrence interval (years),


P probability, and
m rank of low flOl" for given duration.

(iv) Flows for each duration are plotted against recurrence


interval on Extreme Value probability paper, and smooth
eye-fitted curves are drahn (Fig. 3.19). (The scales are
chosen because they tend to lineari se the data and, as
well, Gumbel's (1954) low flow theoretical curves can be
superimposed on the diagram.)

sf-
I II>
5~
"'. 1
4
"i:'
'<.>"
'f
.t:.
!I>
183 ~

:B'"
'5 >-
I 120
"'" c
a;'" 2 SO
0

>
30 Cl
~
7

1.1 15 2 3 5 10 20 50
Recurrence interval (years)

FIG. 3.19 Annual low flow frequency curves for


Brandywine Creek at Chadds Ford, Pa., USA.
(U.S.G.S. 7600)
61

(v) Flows for a given recurrence interval are read from the
low flow frequency curves and replotted as inflows against
duration (like the drought curves in earlier approaches)
on arithmetic graph paper (Fig. 3.20).

50

M-

E 40 ./
./
'"0 ./
x Required reservoir ./
./
./
Q) 30 capacity ./
E ./
:l ./
"0
>
~ 20
0
;;::
.E

10

O~~~ ____ ~ ____- L____-L____ ~ ____ ~ ____L -_ _ ~L- __ ~ ____ ~

o 40 80 120 160 200


Duration (days)

FIG. 3.20 Mass inflow-duration curve for within-year


storage analysis (Data are for Brandywine
Creek at Chadds Ford, Pa., U.S.G.S. 7600).

(vi) Constant draft lines are superimposed on the diagram.

(vii) The largest intercept between the draft line and inflow curves
is taken as the reservoir capacity required to meet the draft
at the design level of reliability (or probability of failure).
For this situation the probability expresses the chance that
the reservoir, if operated under the design conditions, will
fail (empty) at least once within any year.

Asswrrptions:
(i) The reservoir is assumed to be initially full. Unlike
carryover storage situations this assumption will probably
be met in most situations, particularly as the levels of
regulation are usually very low.

(ii) Failures that occur after the end of the critical period
are neglected.
62

Limitations:

(il Variable draft conditions cannot be easily treated.

(ii) The use of frequency curves introduces a bias in the


computed storage estimates. This bias is due in part
to cross-nesting of the mass curves as explained in
Sec. 3.4.8. Computed estimates of capacity should be
increased by 10% to take this bias into account.

(iii) This method further underestimates the required storage


due to the method of establishing the low flows. As the
ranked values are necessarily from different years, there
is no allowance for two or more independent events in the
one year that are more severe than the next rank event
(from a different year). This effect would be small
except for low recurrence interval events (say less than
10 years).

(iv) The frequency analysis is based on daily flows. This


considerably increases the computational requirements
where these have not already been processed.

(v) The method of analysis does not take into account net
evaporation losses. If required, an additional amount
of storage has to be added to the computed value to
cover this loss (Appendix A).

Attributes:

(i) If low flow frequency curves are readily available


for the site in question, the method is quick and simple.

(ii) Notwithstanding the bias in the storage estimate~, the


within-year reservoir capacity determined using the
annual low flow frequency procedure gives satisfactory
estimates if compared with those determined using
annual mass curves (Hardison, 1965).
63

EXAMPLE 3.12

Using the annual low flow frequency curves for Brandywine Creek at
Chadds Ford, Pa., Fig. 3.19, determine the storage required to provide a
30 96 draft with a 5% annual probabi Ii ty of fai lure. (l>lean flow rate for
Brandywine Creek = 1046 m3/s).

* * * * *
The 5% annual probability of failure corresponds to a recurrence
interval of 20 years. From Fig. 3.19 the flow rates for durations of 7, 30,
60, 120 and 183 days corresponding to a 20 year recurrence interval can be
read off as follows:

Duration Flow Rate Flow Volume


( days) (m 3s- 1) (x 10 6m3)

7 1. 62 0.97
30 1. 89 4.67
60 2.29 11.4
120 2.70 28.0
183 3.25 50.6

These points are plotted on the graph (Fig. 3.20) and the storage
determined from the maximum intercept between the demand and mass inflow
curves, that is, 5.1 x 10 6m3 . In practice, the storage estimate needs to
be increased by 10% to account for the bias due to cross-nesting.

3.4.7 Regional Within-year Storage Estimates

Hardison (1965) also provides a technique for obtaining an approximate


estimate of seasonal storage requirements by using the median annual 7-day
flow as an index. Based on 72 streams in eastern United States, he related
seasonal storage need to the median annual 7-day low flow as shown in
Fig. 3.21. Storages taken from these curves are subject to the bias inherent
in using low flow frequency curves to compute storage-draft relations and
the storage estimates should be increased by 10% (see next section).
64

60

~
.2
Cii
::l
c:
c:
'"
c:
'"
Q)
E
'0
'EQ)
~
Q)

.3-
-
....
0'"
~
.l)
10
'"~
.2

0
0 0.3 0.4 0.5
Median Annual 7-day Low Flow.
(ratio to mean annual flow)

FIG. 3.21 Areal draft-reservoir capacity relationship


for 5% probability of failure as a function
of median annual fow flow. (Parameter is
storage capacity in percent of mean annual
flow.) (Hardison, 1965.)

3.4.8 Bias in Mass Curve Frequency Analysis

A further procedural error is associated with the use of frequency


curves to estimate reservoir capacity. This is evident in both the inde-
pendent and overlapping series, and the within-year frequen;y curves. This
effect, which results in under-estimation of the equivalent behaviour
capacity, has been attributed principally to the cross-nesting of the mass
curves of each record (Hardison, 1965) and can be illustrated as follows.

Consider the example given in Table 3.4 which is similar to that used
by Hardison to illustrate the bias. If the two years of low flow data are
time-wise ordered the rank 1 and rank 2 deficiences (or storage sizes) are
3000 and 2500 m3/s day respectively. Yet if the flows are ordered by rank
(which is the procedure in low flow frequency analysis) the rank 1 and 2
deficiences are 3000 and 2000 m3/s day. The difference in rank 2 estimates
results from the cross-nesting effects.
65

TABLE 3.4 Example of calculations to show bias in


frequency-mass curve storage procedure.

Deficiency in m3 js days
Flow in m3 js
for a draft of 600 m3 js
Ordering
5d~ 10 day 5d~ 10 day
period period period period

1929 200 300 2000 3000 1


By
year 1930 100 500 2500 2 1000

1 100 300 2500 3000 1


By
rMk 2 200 500 2000 2 1000

Hardison (unpublished paper, 1965) has examined this bias for the
independent series Md suggests that the under-estimation of storage is
approximately 20% for streams with an Mnual coefficient of variation less
than 0.4. For within-year analyses the degree of under-estimation is
about 10%.

3.4.9 Combining Carryover and Seasonal Storages - Hardison's Approach

In order to correct his carryover storage estimates based on annual


data for seasonal (or within-year) need, Hardison (1965) utilized the Mnual
low flow frequency mass-curve method (Sec. 3.4.6). He proposed two solutions
of which the one described below assumes that the probability distribution
of seasonal storage is independent of carryover storage. The alternative
but quicker procedure assumes that the average seasonal storage requirement
for 100% regulation is 0.4 times the mean annual flow.

Hardison's steps for combining seasonal and carryover storage are as


follows (Fig. 3.22):

(i) Divide the seasonal storage-probability curve (calculated


using the within-year frequency mass curve analysis) into
about eight segments Md compute the mean storage Md
corresponding incremental probability for each segment.

(ii) For a selected amount of total storage, the required


carryover storage for each segment of the seasonal curve
is computed by subtracting the seasonal amount from the
total amount.
66

1.2

\
1.0 \

Q)
:t:
0
\
\
,
\
Cl c: \ ,
tV
.... ::l 0.8 carryover
!II
....
iii
::l
, \
c: 0.6
".:;....
Q) c:
tV
\
\
cr- c:
tV
"\
"\ '.~ combined
Q)
a: Q)
~ 0.4
.....
"',
0.2

0 .0.1 95 98 99
Probability of failure ('Yo)

FIG. 3.22 Combination of seasonal and carryover


storage probabilities.

(iii) For each segment, the probability of the required


carryover storage taken from the carryover storage-
probabili ty curve is multiplied by the incremental
probability of the selected amount of total storage
in (ii). (If there are segments of the seasonal
storage curve which correspond to storage values equal
to or greater than the total storage, then the
probability for the corresponding carryover storage
for those segments is taken to be unity.)

(iv) This is repeated for other points of total storage


to give the total storage versus probability curve
as shown in Fig. 3.22.

In the above analysis, the carryover storage-probability curve must


be based on annual data. If monthly data were used the computed carryover
storage would automatically have taken the seasonal need into account.
67

3.5 OTHER CRITICAL PERIOD METHODS

Because of distinctive seasonal variations (four wet and eight dry


months each year) Wilson (1940) was able to define clearly the starting and
finishing months of critical periods and hence compute the storage size
necessary to offset specific inflow conditions. He evaluated the proba-
bility of failure associated with each initial period and summed these to
give the total probability of failure. The procedure is not of general
applicability.

Law (1953, 1955) developed massed curves of rainfall expressed as a


percentage of average rainfall for various durations and coefficients of
variation and for given probabilities of occurrence. Through regression
analysis, the rainfalls were converted to streamflow for the same durations.
Assuming an initally full reservoir, Law used a cumulative depletion diagram
to calculate the required reservoir capacity. The generalized procedure is
appli cab Ie to the Bri tish Isles for which the empirical massed rainfall
curves were derived. Outside this region, a complete analysis would need
to be carried out before the procedure could be applied. Other procedures
are less complex.

3.6 SUMMARY

Critical period procedures for estimating reservoir capacity-yield


relationships were reviewed under three main headings - methods which
indicate reservoir fullness with time, methods based on the range of flows
and methods based on low flow sequences. Another classification that could
have been used is based on whether the procedure allows storage to be
related to probability of failure. Those methods that do not consider
probability of failure - mass and residual mass curves, Hurst and sequent
peak, and Waitt's minimum flow approach - are considered to be inadequate.
However, the sequent peak algorithm, although reviewed as a technique to be
used with only an historical record, was developed as an efficient approach
to be used with generated sequences; in that context it is possible to
calculate the storage-probability relationship.

Examination of the assumptions and theoretical basis of the cri tical


period procedures not included above, showed that all are deficient in one way
or another. For example, all assume that the reservoir is initially full.
Alexander, Dincer and Gould Gamma procedures also assume that annual serial
correlation is zero. In addition, Dincer's procedure is based on the
68

assumption that n consecutive year flows are normally distributed whereas


Alexander and Gould assume that flows are Gamma distributed. Because of
cross-nesting of mass curves (Sec. 3.4.8) carryover and within-year
frequency analysis underestimates storage need. In addition the overlapping
carryover frequency procedure is inadequate because of effects of dependence.

From this review it is concluded at this stage that the Alexander and
Gould Gamma approaches appear to be suitable preliminary design procedures,
and that behaviour analysis of finite reservoirs is a useful technique to
display clearly the behaviour of the reservoir contents.

3.7 NOTATION

A loge x - loge x (Eq. 3.13) i.e. log of mean - mean of logs.

B draft parameter in Hurst's equations (Eqs. 3.6, 3.7)

c variable in Eq. 3.35.

C reservoir capacity

C1 ' C2 various reservoir capacity estimates

CCRIT reservoir capacity for critical drawdown using


the Minimum Flow approach (Sec. 3.4.1)

CDESIGN design reservoir capacity including safety factor

CP critical period

CP I critical period for a = 1

C coefficient of variation
V

C storage capacity for n year inflow with


n,p
p% probability of occurrence
C reservoir capacity in gamma units
y

d difference between the lower p percentile flow of Gamma


distribution Gec) and a Normal distribution N(c,c)

D draft as ratio of mean flow

D. draft
1

D constant draft over n year period


n
D draft during tth period
t

f(x) probability density function

G( c) Gamma dis tribution with mean and variance equal to c

K Hurs t exponen t
69

,1',. number of months of emptiness for semi-infinite storage


1

number of months of emptiness for semi-infinite storage


(Sec. 3.2.4)
water losses from the reservoir other than evaporation
during time t
m rank of event

m. number of months of emptiness for a finite storage


1

number of months of emptiness for a fini te storage

n 1 ength of a sub-sequence of monthly flows

n number of items of data

N number of flow events

N number of years of data

N' number of months of data

N( c, C) normal distribution with mean and variance equal to c

p percentage ch"ance of occurrence


p probability of failure or probability of occurrence

probability of failure using behaviour diagram

p . probabili ty of failure using semi-infinite depletion diagram


semI
sequent peaks

n year flow with a probability of occurrence of p%


. fl ow d
In '
urlng t th perla
. d

range of flows

range of flows
s standard deviation

s standard deviation of n year flows


n
t time

T. recurrence interval for independent partial duration series


1

T recurrence interval of an n-month event


n
T recurrence interval for overlapping partial duration series
a
T recurrence interval (years)
r
sequent troughs

x flow volume
70

x mean flow

x mean of n year flows


n
X value of flow

x.l flow

z
a'
z standardized normal variate
p
Zt,Zt+l reservoir storage contents at the beginning and the
end of tth time interval
a shape parameter in Gamma distribution

a n year flow Gamma shape parameter


n
a estimate of a CEq. 3.13)

S scale parameter in Gamma distribution

Sn n year flow Gamma scale parameter

B estimate of S CEq. 3.13)

~Et net evaporation loss during time t

T reservoir capacity divided by mean annual flow

Tj reservoir capacity divided by mean annual flow for = 1

T reservoir capacity divided by mean annual flow in Gamma units


Y
TCa) Gamma function
71

CHAPTER 4

PROBABILITY MATRIX METHODS

The second group in the classification of reservoir capacity-yield


procedures (Fig. l. 2) is headed "Moran Related and Other Techniques".
Virtually all of the methods shown are based on the theory presented by
Moran in his book Theory of Storage (1959).

In terms of practical usefulness, the most important methods in this


group are those described as probability matrix methods. The other
techniques are mainly of theoretical interest and are important only because
of their r6le in the development of the procedures which use historical data.
This chapter includes some of the theoretical development, but is mainly
devoted to details of selected probability matrix procedures.

4.1 GENERAL CLASSIFICATION OF MORAN DERIVED METHODS

In Fig. 1.2 we see that the Moran approach can be subdivided into
three groups:

(i) those in which both time and volume are considered as


continuous variables. The most common continuous time
model is the 'random buckets in the bath' model.
Moran's (1959, p. 79) description of the model is of
a man pouring buckets, at random instants of time,
into a bath which has no plug". Generally, the continuous
time model is the most complex and least realistic of the
various classes of techniques due to Moran [see, for example,
Gani (1955), Gani and Prabhu (1958, 1959) and Gani and Pyke
(1960, 1962)]. The assumed form of the inflow distribution
(often Poisson), the darn size (possibly infinite) and the
'buckets in the bath' approach all contribute to the
unrealistic nature of the solution. Thus, continuous
time solutions are of theoretical interest only;

(ii) those in which time is discontinuous but water volumes aPe


continuous. Moran (1955) derived the following integral
equations describing a mutually exclusive (see below)
situation.
72

For x ( C - D
D D+x
g(x) = f(x)! g(x) dx + ! f(x+D- t) get) dt (4.1)
o D

For x > C - D
D C
g(x) = f(x) ! g(x) dx + ! f(x + D - t) get) dt
o D

+ f(x + D - C) ! get) dt (4.2)


C

where x inflows,
C reservoir capacity,
D constant release during unit period,
f(x) inflow probability function, and
g(x) probability function of storage content
plus inflow during unit period.

Solutions for particular inflow distributions and release


rules have been obtained by Gani and Prabhu (1957),
Prabhu (1958a) and Ghosal (1959, 1960). Of these the most
potentially useful solution is that due to Prabhu (1958a)
in which the inflows were assumed Gamma distributed and
releases were assumed constant. However, evaluation of
this solution is very complex;

(iii) those in which time and water volumes are both discrete
variables. This approach by Moran (given in his 1954
paper) and followed by others (for example Ghosal, 1962
and Prabhu, 1958b) is the basis of the practical applications
of his work. Basically it involved sub-dividing the reservoir
volume into a number of parts, thus creating a system of
equations which approximate the integral equations
(Eqs. 4.1 and 4.2). This approximation primarily affects
the results at the storage boundaries (that is, full and
empty) but is satisfactory if the sub-division of the
storage volume is fine enough.

Two main assumptions can be made about the inflows and outflows,
which occur at discrete time intervals. The first, given by Moran (1954),
assumes that the inflows and outflows do not occur at the same time. In
this model, termed the "mutually exclusive" model, the unit period is
73

sub-divided into a wet season (all inflows and no outflows) followed by a


dry season (all releases but no inflows).

The alternative assumption, not given by Moran but which is only a


simple further development, is that inflows and outflows occur simultaneously
- the "simultaneous" model.

Both of these models are discussed in detail in the following sections.

4.2 A SIMPLE MUTUALLY EXCLUSIVE MODEL

It is convenient to choose the inflows, draft, and storage capacity as


integer multiples of some arbitrary volume unit. Consider the following
example:

Reservoir capacity: 2 units

Constant draft: unit per time period

Inflows: discrete and independent and distributed as in


Fig. 4.1. Note that the sum of the probabilities
equals unity.

Relative
Frequency
(Probability) 1/5

o 1 2 3
Units of flow

FIG. 4.1 Distribution of reservoir inflows.

For the mutually exclusive model we have:

Z
t+l o if Z + Xt
t
~ M (4.3)

Zt+l if M < Z + X < K (4.4)


t t
Zt+l ifK Zt + Xt
~ (4.5)

where
store d water at t h e b eglnnlng 0 f t h e t th perlo,
'd
Zt
Zt+l stored water at the end of the tth period or at
the beginning of the (t+l)th period,
K capacity of reservoir,
X . fl ow d
In '
urlng t th perlo,
. d an d
t
M constant volume released at the end of the unit
period.
74

Gi ven this information about capacity, draft and inflows, the firs t
step is to set up the "transition matrix" of the storage contents.
A transition matrix shows the probability of the storage finishing in any
particular state at the end of a time period for each possible initial
state at the beginning of that period. The transition matrix for the above
example is a (2 x 2) matrix representing an empty condition and a half full
condition as follows:
Initial State Zt

Empty Full
1 2

Finishing 1
1
Empty - + -2 -
State 5 5 5 (4.6)

1 -1 + 1 2
- + -1 + 1
Zt+l 5 5 5 5 5
Full 2
(always check)
I = 1 1

Each element of the transition matrix is found by applying Eqs. 4.3


to 4.5 to determine the inflows (and hence probability) of the storage
beginning and ending in the state corresponding to that element. In the
computations the boundary conditions (empty and full) must be considered
and, for the mutually exclusive model, the inflows must be considered
separately and prior to the outflows.

Consider the element (0,0) in Eq. 4.6 which represents a reservoir


starting empty and finishing empty. This can happen if there are no
inflows for the period (probability 1/5) or if there is one unit of inflow
(probability 2/5). In the latter case the release of one unit reduces the
reservoir contents back to zero. Hence, if the reservoir starts empty
there is a probability of 0.6 that it will still be empty at the end of the
time period.

Consider now the element (1,0) which represents a reservoir starting


empty and finishing half full. If there are two units of inflow
(probability 1/5) followed by one unit of release the reservoir will finish
half full. If there are three units of inflow (probability also 1/5) the
reservoir will spill because its capacity is only 2 units, then after 1 unit
of release, it will again finish half full. Thus the probability of going
from empty to half full is 2/5.
75

Note that the reservoir can never finish (and hence start) in the
full condition because of the mutually exclusive assumption about inflows
and outflows. Note also that the reservoir must finish in some condition
thus the sum of the probabilities in any column must be unity.

Let us now assume that the time unit is equal to one year and that
the reservoir of capacity 2 units is empty at the beginning of the year one,
that is, the initial probability distribution of storage contents is:

o
Storage
1 (4.7)
State
2

L=l
Since the transition matrix expresses the conditional probability of final
storage contents given the various values of initial contents, the proba-
bility distribution of final contents can be found by the matrix product of
the transition matrix and the probability distribution of initial contents.
Therefore, at the end of year one (or at the beginning of the year two) the
probability of storage content will be:

0.6 0.2J 0.6 x 1 + 0.2 x OJ = [0.6 ]


[ 0.4 0.8 [ 0.4 x 1 + 0.8 x 0 0.4
(4.8)
transition state state of storage L 1.0
matrix of at end of year one
storage
at
beginning
of year
one

The quantitative process in Eq. 4.8 may be described as follows. The


transition matrix shows the probability of the reservoir finishing in a
specific state, given an initial state. If the initial state is known in
terms of probability, then the joint probability will indicate the likeli-
hood of the storage ending in a specific state. In Eq. 4.8 the transition
matrix shows the probability of going from state 0 + state 0 as 0.6, and
the probability of being in state 0 at the beginning of year one is 1, thus
the probability of ending in state 0 is 0.6 x 1 = 0.6. But also it is
possible to arrive at state 0 from state 1 which from the transition matrix
has a probability of 0.2. The likelihood of being in state 1 at the
beginning of the year one is 0, thus the probability of ending in state 0
but beginning in state 1 is 0.2 x 0 = O. Hence the combined probability of
ending in state 0 at the end of the first year is 0.6 + 0 = 0.6. A similar
argument holds for state 1.
76

The process can now be repeated, using the state vector as the new
starting condition. Therefore, at the end of the second year, the proba-
bility of storage content will be:

[ 0.6 0.2J [0.6 ] [0.6 x 0.6 + 0.2 x 0.4 J [0.44 ]


0.4 0.8 0.4 0.4 x 0.6 + 0.8 x 0.4 0.56 (4.9)

transi tion state of state of storage at


matrix storage at end of year two L=1.00
end of year
one or
beginning
of year two

At the end of the third year, the probability of storage content will be:

[0.6 0.2J = [0.6 x 0.44 + 0.2 x 0.56] = [0.38 ]


[0.44]
0.4 0.8 0.56 0.4 x 0.44 (4.10)
+ 0.8 x 0.56 0.62
L=1.00

At the end of the fourth year, the probability of storage content will be:

0.6 0.2J 0.38J = [0.6 x 0.38 + 0.2 x 0.62] = [0.35] (4.11)


[ 0.4 0.8 [ 0.62 0.4 x 0.38 + 0.8 x 0.62 0.65
L=1. 00
At the end of the eighth year the probability of the storage content
will be:

0.33J (4.12)
[ 0.67

At the end of the ninth period it will be:

0.33J (4.13)
[ 0.67

It will be noticed that as successive years are consi~ered, the


probability vector of storage content becomes less affected by the initial
starting conditions(in this example, the reservoir was assumed empty) and
approaches a constant or steady state situation, which is independent of
the initial conditions. From the steady state vector (Eq. 4.13) it is seen
that there is a 1/3 chance that the reservoir will be empty at the end of
any year.

4.2.1 The Discrete Equations for the Mutually Exclusive Model -


General Case.

Consider a reservoir with discrete inflows, X , and a constant draft


t
of M during unit time period t. Zt is the stored content at the
77

beginning of time t. All volumes are multiples of a constant water volume.


The reservoir is divided into K-M+I discrete zones 0, 1, 2, ... ,K-M where
o is the empty zone.

From continuity it follows that

o Zt + X
t
:: ~1 (4.14)

M < Zt + Xt < K (4.15)

K :: Zt + Xt (4.16)

in which M is released at the end of the wet season. Let the probability
of the reservoir being in its empty zone at the end of a period be:
pI prob {Zt+l = o} (4.17)
o
prob {Zt + X :: M}. (4. 18)
t

Assuming that the X values are serially independent, then


t
Pb prob {Zt = olx t ' M}* + prob {Zt = llx t ' M - I}
+ + prob {Z = Mix = o} (4.19)
t t
M M-l
pI
o Po L qi + PI L q.+
l
... + PM qo (4.20)
i=O i=O

where P. probability that the content of the reservoir


l
is in zone i before the inflows arrive,
qi probabili ty of receiving an inflow of i units
during the wet season, and
Pi probability that the content of the reservoir
is in zone i at end of period.

In a similar manner equations can be derived for each of the K-M+l zones
that the reservoir contents can occupy. The resulting set of K-M+l equations
follow.
M M-l
pI
0 I q.
l L qi ... qo 0 0 0 P
0
i=O i=O
pI 0 0
1 qM+l qM ql qo PI
pI 0
2 = qM+2 qM+l q2 ql qo P2

I q.
i=K-l l

* prob {Zt olx :: M} means probability of Zt = 0 given that Xt :: M.


t
78

or in condensed form,

[P '] [T] [P] (4.21)

4.3 A SIMPLE S lMUL TANEOUS MODEL

Consider the previous example but now with inflows and outflows
occurring simultaneously.

Reservoir capacity: 2 units

Constant draft: unit per time period

Inflows: discrete and independent, and distributed


as in Fig. 4.1.

For the simultaneous model it follows that:

0, (4.22)

o M < K (4.23)

K M ;: K (4.24)

that is, X and M occur simultaneously.


t
Again the transition matrix is computed. In contrast to the mutually
exclusive case it is now possible for the reservoir to finish in a full
condition at the end of a period; the resulting matrix is:

Initial State Zt
Empty Full
a I 2

I 2 I
Empty 0 - + 0
Finishing 5 5 5
State
I 2 I
I (4.25)
Zt+l 5 5 5
1 I I 2 1 1
Full 2 - - + - +
5 5 5 5 5 + 5

I = I 1 1

the elements being determined in the same way as before. For example, to
start full and end full (with one unit of release) is possible for any
inflow except zero inflow. Hence the probability is the sum of the
probability of inflows of one, two, and three units of inflow, that is, O.S.
79

As before, if one examines the probability of reservoir contents in


successive years, one sees that the effect of initial conditions soon dies
out. For example, consider the storage initially empty.

.6 .4.2
.2 OJ [1] [0.60J
[.2.2 .4.8 0.20, [0.40J 0.25 , ~0'18J
0.24 , [0.29J 0.25l
0.25 ,.. [0.13 (4.26)
0.20 0.36 0.46 0.56 0.62J
End of End of End of End of steady
1st yr. 2nd yr. 3rd yr. 5th yr. state

Now consider the storage initially full.

6 .2 0] [OJ [O'0.20,0.24,0.25,0.25,
OOJ [0.04] [0.07J [o.l1J .. [0.13] (4.27)
[.2.4.8 1 0.80 0.72 0.68 0.64 0.25
.2.4.2
0.62
End of End of End of End of steady
1st yr. 2nd yr. 3rd yr. 5th yr. state

It is observed that steady state conditions are independent of either


initial condition.

4.4 COMPUTATION OF STEADY STATE CONDITION

Given the transition matrix of the storage content for a mutually


exclusive or simultaneous model, three methods are available to compute the
steady state condition.

(i) Use the method adopted in the previous numerical examples in


which the probability distribution of storage contents was
computed year by year using [Pl + = [Tl [Pl . At steady state
t l t
[Pl + = [Pl . This approach also allows us to determine year
t l t
by year the probable change in storage content, for example,
one is able to determine the probable rate of filling of a
newly constructed dam.

(ii) A second way to find the steady state condition of stored


contents is to "power-up" the transition matrix thus:

(4.28)
where [T] transition matrix, and
m positive integer large enough for the
resulting matrix to be equivalent
to steady state conditions.
80

For example, consider the transition matrix used previously.

[6.2
.2
.2
.4
.4
O~'
.8
[:
.2
.2
.4
.4
0:]
.8
x
['
.2
.Z
.2
.4
.4
:] = [40
.8
.24
.36
.20
.28
.52
OO:J
.24
.72

.20 .16
[40
.24 .28 OJ' [22
.24 .25 .25 09J
.25
.36 .52 .72 .53 .59 .66

[22
.25
.16
.Z5
o~'
.25 [14 .25
.13
.25 012J
.25
.53 .59 .66 .61 .62 .63

[14
.25
.13
.25
012J
.25
[13
.25
.13
.25 13J
.25
.61 .62 .63 .62 .62 .62 (4.29)

As a general rule squaring the transition matrix four or


five times is sufficient to determine the steady state
condition. This condition is reached when each column
contains the same set of elements.

(iii) The third approach is based on the fact that at steady


state

[T 1 [P 1t = [P 1t + 1

In the numerical example of the simultaneous model above


(Eq. 4.25), the equations become:

0.6 Po + 0.2 PI + 0 P Po (4.30)


z
0.2 Po + 0.4 P + 0.2 P PI (4.31)
1 2
0.2 Po + 0.4 PI + 0.8 P P (4.32)
2 2
I t wi 11 be found that these three equations are not
independent. To get a solution one of them (anyone) must
be replaced by the equation

(4.33)

Solving for PO' PI' P2 as a system of simultaneous


equations gives
81

Po 0.125

PI 0.250

P 0.625
2
These values are the same as those found by method (i)
(Eq. 4.27) and by method (ii) (Eq. 4.29).

4.5 DISCUSSION - MORAN TYPE MODELS

The mutually exclusive model (Sec. 4.2) overestimates both the


probability of failure and the probability of spill. This is because of
the assumption that inflows always precede outflows; thus the reservoir
can never be full at the end of a time period.

The simultaneous model, on the other hand, is more realistic in that


it is more representative of reservoir inflow and outflow conditions.

Limi tations:

For both models there are a number of limitations.

(i) Inflows are assumed to be independent. For a monthly time


period this assumption is not valid, and the required storage
capacity would be seriously underestimated. Annual flows are
independent for many streams, but this period is normally too
long for final design computations. If annual flows are not
independent (50% of Australian streams have serial correlation
coefficients greater than 0.1) the storage is underestimated.
This is discussed further in Appendix B.

(ii) As a general rule, seasonality cannot be taken into account


unless seasonal data is 'used and serial correlation assumed
to be zero. Note that if this were done, the definition of
probability of a given state would change. For rivers
exhibiting a distinctive wet and dry season, the mutually
exclusive approach could be adopted for annual data.

(iii) As described, the techniques assume constant draft, but


varying drafts can be accommodated so long as they are
stored-water dependent and not time dependent.

(iv) A computer is required for solution of realistic problems.


(The above examples were grossly simplified for illustrative
purposes. )
82
(v) An iterative solution is necessary if the storage capacity or
release for a given probability of failure is to be determined.
In the analysis the capacity and draft are assumed to be known
and probability of emptiness is computed. If capacity or draft
is unknown, a value is assumed and the computed probability
is compared with the design probability.

Attributes:

(i) The procedures sample all years of stream flow records without
reference to the historical sequencing (but serial correlation
is assumed zero).

(ii) Computed storage estimates are independent of the starting


conditions. This is in contrast to critical period techniques
and is one of the prime advantages of matrix methods.

(iii) N.on-continuous records can be handled. As annual flows are


assumed to be independent, the sequencing of flows becomes
unimportant and so records with missing data can be used as
effectively as continuous records.

(iv) Estimates of probability of the state of the reservoir can


be computed either at steady state or as a time dependent
function of starting conditions.

4.5.1 Further Modifications

An exampie of Moran's procedur~ is given by Jarvis (1964) who applied


the technique to the Ord River storage in Western Australia. His unit time
period was a year and he assumed that the annual flows were independent.
His release rule was more complex (and more realistic) than the constant
release adopted by Moran.

In 1955 Moran modified the discrete model to deal with seasonal flows.
Transi tion matrices were prepared for each season and were multiplied
together to yield an annual transition matrix. However, as before, the
seasonal flows must be assumed to be independent. Lloyd and Odoom (1964)
adopted a somewhat similar approach.

Harris (1965) gave a worked example of Moran's seasonal method applied


to a British catchment. He found the flows to be seasonal and independent,
and prepared wet and dry season transition matrices which were multiplied
together to give the ann~al transition matrix.
83

Lloyd (1963) partly overcame the independence assumption in the Moran


approach by assuming that the inflows were represented by a bivariate dis-
tribution rather than a simple histogram. In effect, this squares the
number of equations to be solved. Dearlove and Harris (1965) made the
technique more applicable by combining Lloyd's approach with Moran's
seasonal method, but computationally the problem is large and therefore
limited. However, Doran's recent work (Doran, 1975) on the divided interval
technique for solving the transition matrix may overcome tlris limitation.

Venetis (1969) developed monthly bivariate transition matrices from


generated flows using Roesner and Yevjevich's model (1966). Following Moran,
and Dearlove and Harris, he multiplied the matrices together to obtain an
annual transition matrix.

4.6 GOULD'S PROBABILITY ~MTRIX METHOD

Gould (1961) modified the simultaneous Moran-type model to account


for both the seasonality and serial correlation of inflows. He did this by
using the transition matrix with a yearly time period, but accounting for
wi thin-year flows by using behaviour analysis. Thus monthly flow variations,
monthly serial correlation and draft variations can be included.

In the approach storage capacity and draft are given, and probability
of failure of the reservoir is determined. If draft or storage size is to
be determined then a trial and error method needs to be utilised.

4.6.1 Procedure

The procedure to determine storage capacity for given draft and


probability of failure criteria is as follows:

(i) Decide on the releases lD ) required from the reservoir in


t
volume units per month and how they are to vary from month
to month throughout the year. It is usual that releases
are seasonal, although the method requires that annual
releases are not time variant. Operating policy may require
the releases to be reduced if the stored contents fall below
some predetermined values; thus the release rule needs to
be defined at this stage. In addition, net evaporation loss
from the reservoir which is a function of surface area
(hence stored content) needs to be defined.
84

(ii) Decide on the design probability of failure (P ) which


e
is defined as the probability of the reservoir running
dry in any month.

(iii) Assume a first trial, reservoir capacity e1 (volume units).

(iv) Set up a tally sheet (Fig. 4.2) to construct the transition


matrix. Figure 4.2 is shown here for descriptive purposes.
Wi th the computations being done on a computer, the "tally
sheet" is actually a computer matrix.

Zt (Beginning of period)
0 1 2 3 4 19
0
1 l

2
Zt+l 3 Hal Hbl
(End of 4
period)
16
17 HcJ
18 f,
19

FIG. 4.2 Example of tally sheet for transition


matrix in Gould's procedure.

Divide the trial reservoir capacity e into K zones.


As a general rule twenty is sufficient (see Sec. 4.6.2).
The volume of each zone including the top and bottom ones
is given as follows:

Volume (W) (4.34)

0, 1,2, 18, 19

t ~ --------------~V ~------------/
.....
t

empty zone 18 zones of full zone (e1)


of zero volume of equal volume of zero volume
(W)

(v) Apply the continuity equation (same as for behaviour


analysis) on a monthly basis taking one year of data
at a time as described below.

(4.35)
85

where Zt' Zt+l ilie storage content U ilie beginning


th
Md end of the t month,
X the inflow during the tth month,
t
D ilie release d '
urlng t h e t ili month, Md
t
6E the net evaporation loss during the
t
tth month.

The net evaporation loss is the difference between the


evaporation from the proposed reservoir and the evapo-
transpiration from the proposed reservoir site and will
be a function of water surface area which in turn is a
function of volume of water in active storage plus dead
storage capacity.

For each year of data (there are a total of N years),


Eq. 4.35 is applied month by month to determine the zone
in which the reservoir finishes at the end of the year.
This is done for each possible starting zone, and the
element corresponding to the starting and finishing zone
in the tally sheet (Fig. 4.2) is incremented by one. Any
failures (reservoir emptying) that occur during the year
are also noted.

In applying the continuity equation, the seasonal


variability of D should be taken into account in con-
t
junction with any restrictions imposed on the release as
a result of the present state (Zt) of the storage Md
also the 6E losses.
t
To illustrate the procedure, suppose the historical
inflow data for a reservoir site begins in 1929.
The steps are:

(a) Apply the continuity equation to 1929 (the first


year), assuming the reservoir contents are initially in
state O. Route the flows month by month observing the
operating rules associated with D , the losses 6E , and
t t
the boundary conditions of emptiness and spill. In the
example illustrated in Fig. 4.3, no restrictions are
applied to D and evaporation losses are zero. At the
t
beginning of the year Zt = 0 and the reservoir content
ends the year in zone 3, that is, Zt+l = 3, and the
corresponding element on the tally sheet is incremented.
It is noted as (a) in this case in Fig. 4.2.
86

19 19 19
18 18
17 17

A
1929 1929 1929
(a) (b) (c)

FIG. 4.3 Monthly routing example in Gould's Procedure.

(b) ~ext apply the continuity equation to 1929


beginning with the reservoir in state 1. Route
flows as before.

Hence Zt = 1, Zt+l = 3. See (b) on the tally sheet.


Other zones are completed in a similar manner.

(c) Finally, begin in zone 19 (that is, the full zone)


and then apply continuity equation.

Hence Zt = 19, Zt+ 1 = 17 and record. It is shown as (c).

~~en the routing is completed, the sum of the elements of


the tally sheet should be N x K (N years of flow x K zones)
with subtotals of N in each colurrrn of the matrix.

(vi) Divide each column by N to convert the element contents


to probability, thus forming the transition matrix which
expresses the state of the reservoir contents at the end
of a year as a probability relationship of the state of
the reservoir contents at the beginning of that year.

(vi i) In addition to the transi bon matrix, it is necessary to


estimate the probability of failure (that is reservoir
emptiness) as a function of the starting zone. For
example, consider Fig. 4.3:

in case (a) - reservoir started in zone 0 and failed


(emptied) in two months (July and August);

in case (b) - reservoir started in zone 1 and failed in


one month (August);

in case (c) - reservoir started in zone 19, and no


failures were recorded.
87

From this,another tally sheet (or vector in the computer)


is built up in which the probability of failure is related
to the contents of the reservoir at the beginning of the
year. The second tally sheet is shown in Fig. 4.4.

Zt ------ 19
No. of months
of failure

FIG. 4.4 Example of tally sheet for monthly failures in


Gould's Procedure.

As there is a total of 12 N months routed through each


zone, the values in Fig. 4.4 may be converted to probability
by dividing each entry by 12 N. This step is different to
that proposed by Gould in that he defined failure on an
annual basis, that is, one or more monthly failures within
a year constituted one annual failure. In effect, Gould's
measure overestimates the monthly behavioural capacities by
about 20%. (Based on an analysis of six Australian rivers
and two draft conditions, percentage increases are given in
Table 4.1.) For this reason the procedure recommended above,
that is defining monthly failures from monthly data, is
preferred to Gould's approach.

TABLE 4.1 Percentage increase in required capacity using


Gould's definition of failure over a behaviour
analysis using monthly data.

Ri vers
(Australian gauging 50% draft 90% draft
station no. )

Yarra
29 17
(229102)
Murrumbidgee
32 16
(410008)
Lachlan
15 15
(412010)
Warragamba
41 7
(212240)
Namoi
21 23
( 419007)
Burdekin
20 26
(120090)

Mean 26 17
88

(viii) Compute the steady state matrix by one of the three


procedures given in Sec. 4.4. (A computer is required
for this operation.) That is:
m
(a) Power up the transition matrix [T] or

(b) set up simultaneous equations and solve. Remember


to replace one equation by

Po + PI + . + PK- l =1 l4.36)

(see Eqs. 4.30 - 4.32 as an example); or

(c) compute the stored content year by year starting


initially with the reservoir in a known or assumed state.

(ix) Multiply the steady state probability of the reservoir


being in a particular zone [from step (viii)] with the
probability of failure for that zone computed in step (vii);
the resulting sum of the products is the probability of
failure of the reservoir. An example is shown in Table 4.2
where the probability of failure is computed to be 0.02.

TABLE 4.2 Example of combining steady state probability


with the conditional failure probability in
Gould's procedure.

Probability of Conditional
prob abi 1 i ty of Product of
starting in a
failure in any probabili ties
Zone particular zone
month within
(steady state)
any year for
that zone
(1) (2) (3) (2) x (3)

0 0.016 0.385 0.0062


1 0.017 0.154 0.0026
2 0.064 0.051 0.0033
3 0.094 0.026 0.0024
4 0.202 0.026 0.0053
5 0.216 0.000 0.0000

19 0.101 0.000 0.000

L= 0.0200
(x) If the computed probability of failure does not e~ual the
design probability, choose a new capacity and proceed from
step (iv) again.
89

Assu;nptions:

(i) Annual flows are assumed independent, that is, serial


correlation is zero.

Limitations:

(i) The analysis requires a computer for the computations.

(i i) The explici t result of the Gould analys is (like other


probability matrix methods) is probability of failure
for an assumed storage size. If storage size or draft
is required to be calculated, a trial and error approach
must be adopted. Usually three or four iterations are
sufficient.

(iii) Annual serial correlation is assumed to be zero. For


many streams this assumption is satisfactory but for
many others it is not.

Nevertheless, as recommended for other procedures in which


annual flows are assumed to be independent, storage estimates
should be corrected for the assumption of independence
irrespective of whether or note the serial correlation value
is statistically significant (see Appendix B for method).
Gould (1964) does provide a correction for storage
estimates on streams with a significant serial correlation
coefficient, but it is of limited value because it does not
cover the required range of design criteria.

AUributes:

(i) The procedure samples all years of data without reference


to the historical sequencing of annual flows. This is in
contrast to many critical period procedures which sample
only periods of critical duration.

(ii) Computed storage estimates are independent of the initial


reservoir conditions.

(iii) Non-continuous records can be handled. As annual flows are


assumed to be independent, the sequencing of flows is
unimportant and so records with missing data can be used as
effectively as continuous records.

(iv) No assumptions need to be made about the distribution of the


historical flows. Relatively speaking this is not a significant
advantage as many other storage procedures also do not require
90
an assumption about a theoretical distribution of flow for
their implementation.

(v) Monthly parameters and monthly serial correlations, except


that between the last month of year i and the first month
of year i+l, are automatically taken into account as yearly
flow sequences are routed on a monthly basis through the
reservoir.

(vi) Probability of failure is computed either at steady state


or as a time dependent function of the starting conditions.
The latter attribute is a most important characteristic of
probability matrix methods.

(vii) Varying drafts and complicated release rules can be handled


easily.

(viii) Net evaporation losses from the reservoir can be related


through the stored water content to the surface water area
during the analysis.
4.6.2 Practical Considerations

(i) To assist in determining a given reservoir capacity it has


been found useful to plot probability of failure (on a
Normal probability scale) against storage capacity (on a
logarithmic scale) as this tends to linearise the relationship
and reduce the number of iterations required to determine the
capaci ty for the given design probabi li ty.

(il) For more variable streams difficulties have been experienced


at high drafts (> 70%) with the limited accuracy (round-off
errors) associated with mini-computers in solving for steady
state using (20 x 20) or larger matrices. In these cases,
as a check to the solution of simultaneous equ~tions it is
recommended that the transition matrix be powered up and a
check made to ensure that the sum of each column equals unity.

(iii) Probability matrix solutions are affected by zone size and


hence the number of zones. Joy (1970, Appendix V) examined
this question using Moran's mutually exclusive model and
found that for streams with C < 0.5, 20 zones were required
v
to adequately define the storage size; for 0.5 < C < 0.85,
v
30 zones were required and for C ~ 0.85, 40-50 zones were
v
required.

Teoh (personal communication, 1977) analysed ten streams with


Cv varying between 0.19 and 1.79 using Gould's procedure. From
91

his results it is concluded that as a general rule:


for C < 0.5 usc 10 zones,
v
for 0.5 ~ C < 1.0 usc 20 zones,
v
for 1.0 ~ C < 1.5 use 30 zones, and
v
for C ;: 1.5 use 40 zones.
v
The differences between these and Joy's results (which were based
on Moran's rather than Gould's model) arc consistent with Doran's
divided interval approach (Doran, 1975) and Klemes' (1977)
recent analysis.

If an insufficient number of zones is used, sometimes a


hunting effect in the storage-probability relabon becomes
evident. An example is shown in Fig. 4.5 for the Nogoa River
in Queensland. Results arc plotted for 10, 20, 30 and 40 zones.

~ 10 zones
25 20 zones
+----+ 30 zones
./0. 40 zones
20

~
~
"
~
15

'0
.~ 10
:c
"'0
.0

ct
5

O+-----~----_r-----+----~~~
o 2 3 4
Reservoir capacity (xl0 6 m3 )

llG. ->'5 Effect of llllmber of zones on reservoir capaci ty -


probabili ty of failllre relation (Nogoa River -
Australian gauging station no. 13020J).

(iv) At this state little guidance can be given regarding the effect
of beginning a Gould analysis in different months because insuf-
ficient research information is available. It appears that, for
at least some rivers, the derived storage using a GOllld analysis
docs depend on the starting month if the draft is high. It is,
therefore recommendecl that before a fina 1 des i gn capaci ty is
chosen, four separate Gould analyses be carried out, each
heginning three months apart, to check the significance
of the starting month.
92

EXAMPLE 4.1

For the Mitta Mitta River (Appendix E) use Gould's probability matrix
procedure to determine the storage required to meet a draft of 75% of the
mean flow with a 5% probability of failure.
* * * * *
The Gould procedure requires a computer for efficient solution; for each
estimate of storage capacity the method requires each year of flow to be
routed through the storage for each possible starting condition. In
Sec. 4.6.2 it is shown that the number of zones required depends on the
coefficient of variation of annual flows, C ' For the Mitta Mitta River,
v
C is equal to 0.57; 15 zones should therefore suffice.
v
The procedure is an iterative one, the probability of failure being
calculated for the input draft and the storage capacity estimate. For the
Mitta Mitta at 75% draft and a storage capacity of 910 x 10 6 m3 the following
(15 x 15) transition matrix is obtained (terms are espressed as probability):

Starting Zone Zt

o 3 4 6 7 8 9 10 11 12 13 14
o .147 .147 .147 .147 .147 .118 .118 .088 .088 .000 .000 .000 .000 .000 .000
.118 .118 .118 .118 .118 .088 .029 .059 .000 .088 .000 .000 .000 .000 .000
.029 .029 .029 .029 .029 .088 .088 .000 .059 .000 .088 .000 .000 .000 .000
~ 3 .029 .029 .029 .029 .029 .029 .059 .088 .000 .059 .000 .088 .000 .000 .000
+
N~ 4 .029 .029 .029 .029 .029 .000 .029 .059 .088 .000 .059 .000 .088 .000 .000
~ 5 .118 .118 .118 .088 .088 .059 .000 .029 .059 .088 .000 .059 .000 .088 .029
~ 6 .059 .059 .059 .088 .059 .059 .059 .000 .029 .059 .088 .000 .059 .000 .059

....~ 7 .029 .029 .029 .029 .059 .059 .059 .059 .000 .029 .059 .088 .000 .059 .029

....~ 8 .029 .029 .029 .029 .029 .088 .059 .059 .059 .000 .029 .059 .088 .000 .029
~ 9 .029 .029 .029 .029 .029 .029 .088 .059 .059 .059 .000 .029 .059 .088 .029
10 .000 .000 .000 .000 .000 .000 .029 .088 .059 .059 .059 .aoo .020 .059 .088
11 .029 .029 .029 .029 .029 .000 .000 .029 .088 .059 .059 .059 .000 .029 .029
12 .059 .059 .059 .059 .029 .029 .000 .000 .029 .088 .059 .059 .059 .000 .029
13 .147 .147 .147 .147 .176 .llS .147 .147 .147 .176 .265 .265 .294 .353 .353
14 .147 .147 .147 .147 .147 .235 .235 .235 .235 .235 .235 .294 .324 .324 .324

To compute the steady state (or long term) probabilities of the


reservoir being in any particular zone the transition matrix can be powered
up or solved as a system of simultaneous equations (Sec. 4.4). The latter
option involves less computation and the result is tabulated below with
the respective probability of failure for each starting zone:
93

Steady State Probabi l i ty of Contribution to


Zone Probabi l i ty of Fai lure from Overall Probability
being in Zone Starting in Zone of Failure
(1) (2) (3) (2) x (3)

0 0.034 0.502 0.0171


1 0.026 0.480 0.0125
2 0.018 0.360 0.0065
3 0.017 0.260 0.0044
4 0.015 0.157 0.0024
5 0.057 0.083 0.0047
6 0.039 0.044 0.0017
7 0.044 0.017 0.0007
8 0.029 0.007 0.0002
9 0.051 0 0
10 0.055 0 0
11 0.031 0 0
12 0.028 0 0
13 0.276 0 0
14 0.279 0 0 I: = 0.0502

that is, the probability of failure of a 910 x 10 6m3 storage is 5.02%.

This estimate needs adjustment for the effects of the annual serial
correlation of 0.06. From Fig. B.l the adjustment factor is 1.06.

Therefore, the final answer by Gould method is 910 x 1.06 = 960


(x 10 m3).
6

4.7 RELATED PROBABILITY MATRIX METHODS

4.7.1 McMahon's Empirical Equations

For 156 Australian streams, McMahon (1976) used Gould's modified


procedure to estimate the theoretical storage capacities for four draft
conditions (90%, 70%, 50% and 30%) and three probability of failure values
(2%, 5% and 10%). These capacities were related by least squares analysis
to the appropriate coefficient of variation of annual flows by the following
simple relationship:
L)~

T C/x aC b (4.37)
v

where C storage capacity in volume units,


x mean annual flow in volume units,
T reservoir capacity divided by mean annual flow,
C coefficient of variation of annual flows, and
v
a, b empirically derived constants tabulated in Table 4.3.

In addition to a and b in Table 4.3, standard errors of estimate


in percent and coefficients of determination are shown. The constants were
not based on regional analysis and are considered to apply to the whole of
Australia.

TABLE 4.3 Reservoir capacity-yield equation coefficients,


standard errors and coefficients of determination
for Mc,lahon I s empi rical method. (e = standard error
of estimate, r2 = coefficient of determination)

Probabili ty of failure (% )
Draft
Parameter
(%) 2.5 5 10

a 7.50 5.07 3.08


b l. 86 l. 81 1. 82
90
e + 18, -15 +12, -11 +43, -30
r2 97 98 87

a 2.51 1. 81 1. 21
b 1. 83 1. 79 1. 74
70
e +21, -17 +25, -20 +29, -23
r2 96 94 92

a 0.98 0.75 0.51


b 1. 91 1. 93 1. 83
50
e +58, -36 +63, -39 +61, -38
r2 81 79 79

a 0.28 0.22 0.15


b 1. 53 1. 49 1. 79
30
e +44, -31 +61, -38 +64, -39
r2 82 72 77

Assumptions, Limitations and Attributes:

As the storage values used in the regression analysis are Gould


estimates, a major assumption relates to the neglect of annual serial
correlations. Corrections given in Appendix B should be made. Another
95

assumption is that capacity for given conditions is related only to the


coefficient of variation. The proportion of variance accounted for shown
in the table suggests that this is a reasonable assumption.
Because of the errors noted above and limitations due to constant oraft,
this procedure is regarded only as a preliminary procedure. However, as a
preliminary procedure it is based on monthly flol"s and on a large number of
well-distributed Australian streams and therefore should provide reasonable
estimates of storage at least within the Australian environment.

EXAMPLE 4.2

Compute the storage required on the Mitta Mitta River (Appendix E) to


meet a draft of 75% with a 5% probabi li ty of fai lure using McMahon's
Empirical equations.

* * * * *
From Eq. 4.37
b -
storage, C (aC ) x
v

From Appendix E for the Mitta Mitta

C 0.57
v

Table 4.3 gives values of a and b for various drafts and prob-
ability of failure. Since a draft of 75% is not mentioned specifically, it
is necessary to interpolate on a log-linear plot of draft versus storage
as follows:

Draft ae b C
a b (l06 m3)
(%) v

90 5.07 1. 81 1. 83 2331
70 1. 81 1. 79 0.66 841
50 0.75 1. 93 0.25 319
30 0.22 1. 49 0.10 127

Interpolation for 75% draft on the log-linear plot of storage versus


draft (Fig. 4.G) gives a storage estimate of 1090 (x 10 6 m3).
96

Adjust for annual serial correlation. From Fig. B.l for 75% draft
and annual serial correlation of 0.06 correction factor is 1.06 approxi-
mately. Thus the estimate of storage requirement by McMahon's procedure is:

1090 x 1.06

X
2000

1000
M X
800
E
~

0 600
~
500
400
~
CD
~ X
~ 300
0
~

00
200

X
100
0

FIG. 4.b Interpolation on log-linear plot for McMahon's


Empirical procedure (Example 4.2).

4.7.2 Probability Routing

Langbein's Probability Routing method (Langbein, 1958) is very similar


to Moran's (1954) probability matrix procedure except that Langbein modi-
fied his technique to deal with correlated annual inflows. Both the stream-
flow regime and reservoir storage were divided into low, medium and high
sub-regimes. By classifying each flow into the same streamflow regime as
its predecessor, three separate streamflow histograms were obtained. Thus
in setting up his system of equations describing the cumulative probability
97

of reservoir contents, Langbein used the inflow distribution appropriate to


the state of the reservoir. In this way he was able to take annual serial
correlation into account in an approximate way.

4.7.3 Hardison's Generalized Method

Hardison (1965) generalized Langbein's probability routing procedure


using theoretical distributions of annual inflow and assuming serial cor-
relation to be zero. This is equivalent to Moran's (1954) model except
that Hardison used a simultaneous model rather than the mutually exclusive
type adopted by Moran. The annual storage estimates are shown graphically
in Figs. 4.7, 4.8 and 4.9 for log-normal, Normal and Weibull t distributions
of annual flows. The percentage chance of deficiency shown in the figures
is defined by Hardison as the percentage of years that the indicated
storage capacity would be insufficient to supply the design draft.

In addition to the carryover storage based on annual data, Hardison


presented a procedure for determining the combined carryover plus seasonal
storage requirement. The latter procedure is discussed in Sec. 3.4.9.

Procedure:
(i) Compute the mean, standard deviation and coefficient of
skewness of both the annual flows and their common
logari thms.

(ii) The appropriate distribution depends on the parameters


computed in (i) as follows:

(a) Adopt a log-normal distribution if the coefficient


of skewness of the logarithms of flows is algebraically
greater than -0.2.

(b) Adopt a Normal distribution if the coefficient of


skewness of the absolute flows is algebraically less
than +0.2 or if the coefficient of variation of the flows
is less than 0.25.

tThe probability density function of the two parameter Weibu1l


distribution is:
11-1 11
f(x) 11 (~) exp [- (~e) 1
8 8
where 11 shape parameter, and
characteristic drought when Prob (x e) lie.
98

........,.... 10 PERCENT CHANCE


OF DEFICIENCY
5 PERCENT CHANCE
OF DEFICIENCY
0
c:
::t
\..
(ij
::t
c:
c:
<'G
c:
<'G
Q)
E
.....
0
0
:;::
<'G
\..
2 . PERCENT CHANCE . PERCENT CHANCE
<'G
1/1
<'G

3
.....>.
'v<'G
Co 2
<'G
()
\..

0
...
>
Q)
1/1
Q)
0:
Variability index Variability index

FIG. 4. Hardison's relationship of reservoir capaci ty-yield-


percentage chance of defi ciency for log-normal
distributions of inflow (Hardison, 1965).

eel Adopt a Weibull distribution if neither a log-normal


nor a Normal distribution is selected except when the
thms of flows have a negative skew coefficient
greater than 1.5.

(iii) Decide upon the reliabili ty (that is, Hardison' 5 percentage


chance of deficiency) and the required draft.

(iv) Based on (ii), enter appropriate Fig. 4.7,4,8 or 4.9 and


estimate reservoir capaci ty as a ratio of mean annual flow.
If the log-normal distribution is adopted use as the variability
index CEq, 2.5) as the measure of variability; otherwise
use the coefficient of variation,

(v) Compute the serial correlation of annual flows and adjust


appropriately the required storage capacity (Appendix B).
99

10 - PERCENT CHANCE 5 - PERCENT CHANCE


:;; 4 .--_......--O_F---,D_E_F_IC_I,.E_N_C_Y-.-_-. OF DEFICIENCY
o
c
2 3
"'iii
:J
C
C 2
('II

C
('II
Q)

-E
o
o
2 - PERCENT CHANCE 1 - PERCENT CHANCE

o
~
Q)
(/)
Q)
a: 0.1 0.2 0.3 0.4 0.5
Coefficient of variation Coefficient of variation

FIG. 4.8 Hardison's relationships of reservoir capacity-yield-


percentage chance of deficiency for Normal distribution
of inflows (Hardison, 1965).

Assumptions, Limitations and Attributes:

Like most probability matrix procedures the serial correlation of


annual flows is assumed independent. Adjustments are given in Appendix B.
As analysis is based on annual flows, seasonal variations are not accounted
for; Hardison provides a generalized procedure to correct for this, but it
is valid only for streams like those in the eastern United States where
variability is low (Sec. 3.4.7).

As a preliminary procedure for estimating carryover storage the


procedure is reasonably quick, theoretically sound (except for neglecting
annual serial correlation) and covers all but the most variable streams
(with C > 1.66) that one is likely to encounter in practice.
v
100

10-PERCENT CHANCE 5- PERCENT CHANCE


:::0 OF DEFICIENCY OF DEFICIENCY
c 4
...
:l

~ 3
:l
C
c 2
1'0
C
1'0
/I)
E
....0 0
0 2 - PERCENT CHANCE 1 - PERCENT' CHANCE
:p
...
1'0
4
OF DEFICIENCY OF DEFICIENCY
11/
1'0

<tJ
>- 3
u
1'0
Co 2
1'0
(.)

0
...
...>
/I)
11/
00 0 0.20.4
/I)
a: Coefficient of variation Coefficient of variation

FIG. 4.9 Hardison's relationship of reservoir capacity-yield-


percentage chance of deficiency for Weibull
distributions of inflows (Hardison, 1965).

EXAMPLE 4.3

Compute the storage required on the Mitta Mitta River (Appendix E) to


provide 75% draft with a 5% probability of failure using Hardison's
Generalized method.

* * * * *
As described in Sec. 4.7.3, it is necessary to compute the mean,
standard deviation, and skewness of both the annual flows and the common
logarithms of annual flows. Parameters for the flows are given in
Appendix E. These calculated annual parameters are:
101

Stmdard
~m Skew
Deviation

Flows 1274 731 1.50

Log 1O (flow) 3.04 0.24 -O.OOS

Using these parameters the choice of distribution indicated under


procedure step (ii) in Sec. 4.7.3 is the log-normal. Thus the curves in
Fig. 4.7 are used.

From the curve for 5% chmce of deficiency md a variability index


equal to the standard deviation of the logs (that is, 0.24) the reservoir
capacity for 75% draft is approximately 90% of the mean annual flow, that
is, 1150 x 10 6m3

After applying the correction for correlated annual flows (from


Fig. B.l the correction factor is 1.06 for a serial correlation of 0.06)
the estimated storage by Hardison's method is 1220 x 10 6m3 .

4.S OTHER MODELS

4.S.l Melentijevich

Assuming an infinite storage md independent normal inflows,


Melentijevich (1966) obtained expressions for both time dependent and steady
state distributions of reservoir content. In addition, he gave results for
the mean, varimce and skewness of the surplus, the deficit md the range
of storage. However, he made no mention of how they could be applied in a
design situation.

In considering finite reservoirs, Melentijevich used a random walk


model md a behaviour analysis of 100,000 rmdom normally distributed
numbers. From the analysis, he obtained an expression for the density
function of the stationary distribution of storage contents. The solution
is complex and limited in use because of the assumptions of normality,
independence md neglect of seasonality. The results are only of theo-
retical interest.
102

4.8.2 Klemes

By considering a reservoir to have only two states - wet and dry -


Klemes (1967) was able to reduce the probability of failure within a
limited period to the classical occupancy problem (that is, the determin-
ation of the number of arrangements of placing a number of identical balls
into a number of identical cells). His technique is restricted to a
uniform release or a randomly varying one. The major limitation relates to
the unit period being one year and, as a consequence, it is not possible to
discern a failure within a year. This could be overcome by working with
monthly rather than annual flows but the number of possible arrangements
becomes very large.

4.8.3 Phatarfod

Phatarfod's procedure (1976) is based on random walk theory and is


concerned"with finding the probability of the contents of a finite reservoir
being equal to or less than some value ~C where ~C > 0 and C is the
capacity of the reservoir. The physical process of dam fluctuations can be
likened to a random walk with impenetrable barriers at full supply and
empty conditions. Phatarfod used Wald's identity, an approximate technique,
to solve a problem with absorbing barriers and a relation connecting the
two kinds of random walks.

Steps in his method, which assumes the draft is the unit of measure-
ment, are:

(i) Assume a constant draft' D as a ratio of mean annual flow.

(ii) Calculate
4
h 2
(4.38)
Y

a Q::L (4.39)
2
20
e ).l - (4.40)
y

where ).l mean flow in draft units liD,


o = standard deviation of annual flow in
draft units, = ).lC '
v
C coefficient of variation of annual flow, and
v
y coefficient of skewness of Annual flow.

(iii) Decide on P and ~ where P is the probability of the


reservoir contents being less than C and C is the
total capacity.
103

(iv) Solve for y, the unique positive solution


(other than unity) of

Py v-I + Py v-2 + ... + Py - (l-P) o (4.41)

where l/t = v.
For example, if t = 1/3,
Y = H- 1 + [1 + 4(1~P)l\ (4.42)

(v) Use the Newton-Raphson iteration to solve for e, which is


the unique positive solution of:

e (1- e) =
+r) + ea(l-r) + [{il+r)+ eaCl-r)}2-4rl~J ( 4 . 4 3)
h loge [Cl---"----~~--"--'~~'----'-~-'-"--"-~

where r annual serial correlation coefficient.

(vi) Calculate required capacity of the reservoir as

C = v log y Dx (4.44)
e
where C capacity in volume units, and
x mean annual flow in volume units/year.

This model assumes that annual flows are Gamma distributed and is
based on a fixed draft. It is considered to be a preliminary design
procedure, although the solution of Eqs. 4.41 and 4.43 can be quite time-
consuming. The procedure is a useful preliminary way to determine the
likelihood of the reservoir falling below some level and the possibility of
restrictions in releases. Because of the approximation made, the procedure
is limited to v being less than or equal to about 5. (Phatarfod, personal
communication, 1977.)

EXAMPLE 4.4

Find the storage required on the Mitta Mitta River (Appendix E) to


provide for 75% draft with a probability of the reservoir being less than
one third full using Phatarfod's procedure.

* * * * *
For the Mitta Mitta River (Appendix E) annual flow parameters are:

coefficient of variation 0.57


coefficient of skewness 1.50.
104

Draft D 0.75

In draft units, mean flow lJ liD = 1. 33,


and standard deviation (J 11 Cv
(1.33) (0.57)
0.76
4
From Eq. 4.38, h
2y 1. 78

(0.76) (1.50)
From Eq. 4.39, a = 2
0.57

2a
From Eq. 4.40, e lJ -Y
1.33 _ 2 (0. 76)
--r:so
0.32.

Choose t = 1/3 so that the probability P of 0.05 gives the chance


of the reservoir being less than or equal to 1/3 full.

From Eq. 4.42, y H- 1 + [1 + 4(l-P)] \


P
+ 4(1 - 0.05)] }
!
H- 1 + [1
0.05

3.89.

Solve Eq. 4.43 for e:

eel-e) = h loge [(l+r)+ 8a(1-r)+ [{(~+r)+ 8a(1-r)}2 - 4r]!]

where r annual serial correlation = 0.06

8(1-0.32)= 1.78 loge [1.06+ 8(0.57) (0.94)+ [{1;06+ 8(0.57(0.94)}2 - 0.24]!]

Using the Newton-Raphson iteration procedure (Appendix D), the


required value of 8 is found to be 1.824.

From Eq. 4.44: Storage v log Y D x


8
3 loge 3.89
1. 824 Dx
2.23 (0.75)(1274)
2130 (x 10 6m3)

[Note that this is the reservoir size for which there is a probability of 5%
of being only one third (or less) full. Thus the figure of 2130 x 10 6m3 is
105

not directly comparable with reservoir sizes based on 5% probability of


failure. However, one can run the Gould procedure and compute the
probability of being in the lower one-third of the storage from 'the steady
state matrix. For 2130 x 10 6m3 storage the answer is 3.8%.]

4.9 SUMMARY

From a theoretical point of view the Gould procedure as described in


Sec. 4.6 stands out as the most acceptable reservoir capacity-yield
technique.

Essentially the procedure involves only one major assumption and


overcomes most of the disadvantages of other probability matrix procedures
and critical period approaches. Based on these reasons and the satisfactory
results of several extensive testing programs using Australian streamflow
data (reported in Chapter 6) the Gould procedure, modified as outlined in
Sec. 4.6.1 and corrected for annual serial correlation, is recommended as a
final design tool for establishing the single reservoir capacity-yield-
probability of failure relationship.

In this chapter it was also noted from a theoretical point of view


that there are two preliminary procedures which are suitable for storage
analysis - Hardison's (1965) Generalized carryover procedure and McMahon's
(1976) Empirical Equations. Both are based on results generalized from
applying probability matrix methods, but the empirical constants in the
latter procedure are based only on Australian streamflow data.

4.10 NOTATION

a variable in Phatarfod's procedure (Sec. 4.8.3.)


a variable in McMahon's Empirical procedure (Sec. 4.7.1)
b variable in McMahon's Empirical procedure (Sec. 4.7.1)
C reservoir capacity
C1, C2 various reservoir capacity estimates
C coefficient of variation
v
D draft as ratio of mean flow
D d
dra f turIng t th perlod
.
t
e variable in Phatarfod's procedure (Sec. 4.8.3)
f(x) probability density function
106

g(x) probability function of storage content plus inflow


during unit period
h variable in Phatarfod's procedure (Eq. 4.38)
K reservoir capacity as a multiple of constant water
volume in Moran analysis (Secs. 4.1-4.3)
K number of zones in Gould analysis (Sec. 4.6)
!C proportion of reservoir capacity being 1/2, 1/3,
1/4 or 1/5 (Eq. 4.41)
m positive integer exponent large enough so that resulting
matrix is equivalent to steady state (Secs. 4.4 and 4.6)
M release as a multiple of constant water volume in Moran
analysis (Secs. 4.1-4.3)
N number of years of data
P prohability of reservoir contents being less than
some amount (Sec. 4.8.3)
P. probability that the content of the reservoir is in
1
zone i at the heginning of the period
P! probability that the content of the reservoir is in
1
zone i at the end of the period
rPj probahi Ii ty vector
qi probability of receiving an inflow of units
r annual serial correlation coefficient
t time
[T] transition matrix of reservoir contents
v lit in Phatarfod's procedure (Sec. 4.8.3)
W zone volume in Gould's probability matrix procedure (Sec.4.6.1)
x flow volume
x mean flow
X inflow during tth period
t
y variahle in Phatarfod's procedure (Sec. 4.8.3)
Zt,Zt+l reservoir storage contents at the beginning and
the end of tth time interval
y coefficient of skewness of annual flows
LlE net evaporation loss during time t
t
~ shape parameter in Weibull distribution
e characteristic drought in Weibull distribution
e variable in Phatarfod's procedure (Sec. 4.8.3)
jJ mean annual flow in draft units
a standard deviaition of annual flow in draft units
T reservoir capacity divided by mean annual flow
107

CHAPTER 5

USE OF STOCHASTICALLY GENERATED DATA

The third grouping of storage estimation methods is based on the use


of generated or synthetic data. In essence, however, the methods are the
same as described previously; the difference is that the input streamflows
are changed. The technique involves using a stochastic generation model to
produce "streamflow" sequences with the same statistical properties as the
historical record. It is then possible to determine the storage capacity
(using some standard method) corresponding to each sequence, thus providing
a designer with a distribution of values. This in turn gives him an idea of
the confidence which can be placed on the adopted design value. "Synthetic
flows (or stochastic data) do not improve poor records but merely improve
the quality of designs made with whatever records are available." (Fiering
and Jackson, 1971, p.24.)

In this chapter we restrict our examination of data generation


processes to operational aspects of Markovian models that are used for
generating annual and monthly streamflows. Readers requiring more detail
are referred to the many excellent texts, reports and papers devoted to
stochastic processes. A selection of these is included in the references.
It should also be noted that data generation procedures will be dealt with
here not in terms of the physical mechanics underlying the streamflow
process but rather from an operational point of view. In this regard we
commend readers to Klemes' (1974) paper for some thought-provoking comments
on the relationships between physically based and operational models.

This chapter is divided into several distinct parts. The first


examines the time-series components making up the streamflow process. This
is followed by a review of historical developments in data generation
procedures up to 1960. Next, the methodology and performance of Markovian
data generation procedures are discussed in detail. Simulation and the use
of generated data are then considered and finally several procedures based
on generated data for making preliminary estimates of reservoir capacity
are reviewed.
108

5.1 TIME-SERIES COMPONENTS

From a stochastic point of view, streamflow data can be regarded as


consisting of four components (Kottegoda, 1970); trend (T ), periodic or
t
seasonal (St)' correlation (K ), and random (E ) components which can be
t t
combined simply as:

(5.1)

These components are represented pictorially in Fig. 5.1.

Time Time
FIG. 5.1 Time-series components of the
streamflow process.

To obtain representative stochastic data, it is necessary to identify


and measure the strength of each component. A fifth component not included
in Eq. 5.1 relates to catastrophic events. This aspect is beyond the scope
of this text and relates to the so-called 'Noah and Joseph' effects and
the 'Hurst' phenomenon. Data generation models accounting for these effects
are still at the research stage (Sec. 5.7).

A sequence of values arranged in order of their occurrence is called


a time-series. A time-series is considered to be stationary if the
statistical properties characterising i t are time invariant. In this
discussion it is assumed that the data are stationary or can be made so by
a simple transformation. For example, to partially eliminate the non-
stationary effect of seasonality, monthly data can be standardised by the
following equation:

x't (5.2)
109

where x monthly flows,


t
x, standardised monthly flows,
-
t
x. mean monthly flow for the j th month, and
J
s. standard deviation of monthly flows for the J.th month.
J
One characteristic of a time-series is persistence which relates to
the sequencing of the data. In Chapters 3 and 4 it was noted that this
property is very important in storage-yield analysis. In streamflow, per-
sistence arises from natural catchment storage effects which tend to
delay the runoff; over a short time period high flows in one interval will
tend to be followed by high flows in the following interval. The longer the
time period the lesser the effect and for many streams it is negligible for
annual flows.

The usual quantitative measure of persistence is serial correlation ..


Serial correlation coefficients may be calculated for the correlation
between the flow in any given time period (for example, month or year) and
the flow in k time periods earlier where k (= 1, 2, ... ) is called the
lag. In many studies only the lag one serial correlation is considered,
that is, the persistence between an event and the immediately preceding
event. Lag one models have been shown to be operationally satisfactory in
several studies (for example, Kottegoda, 1970; Philips, 1972; Wright,1975).
The algorithm to compute serial correlation is given as Eq. 2.10.

For a sample of finite size, computed values of serial correlation


(r , where k is the lag) may differ from zero because of sampling errors.
k
Thus it is necessary to test the values to determine if they are signifi-
cantly different from zero. Yevjevich (1972b) outlines a test for this
purpose. The confidence limits (eL) for a computed value of r are given
k
by:
-lz IN-k-l
Cl
N- k (5.3)

where z the standardized normal deviate corresponding


Cl
to the Cl level of significance, and
N number of flow events.

If falls outside the confidence limits, is considered to be


significantly different from zero at the Cl level of significance.
Equation 5.3 may be used to test the statistical significance of r for
k
k > 1 if k is small relative to N.
110

5.2 HISTORICAL DEVELOPMENTS TO 1960

Hazen (1914) is considered to be the first to recognise the


desirability of extending hydrolgic data. He combined standardised annual
flows for fourteen streams in the northwest of U.S.A. to produce a syn-
thesized record of 300 years. His procedure has a number of limitations.
The streams were geographically close, and more than half the records were
based on the period 1900-1910, so that records tended to be repeated. The
technique of combining the flows forces the residual massed curve to pass
through zero at least fifteen times thus restricting the range of the
combined data. This would result in an underestimation of the storage
requirement. But in Hazen's procedure this effect was compensated for
because his storage was determined on the basis of a semi-infinite reservoir.
Hazen's curves are still used today to assess preliminary storage sizes in
the eastern United States.

Sudler (1927) utilized historical and representative annual flows


which were entered on fifty cards. These cards were shuffled and dealt
without replacement to produce a sequence of 50 years. This was repeated
twenty times, producing in all 1000 years of data. The procedure is
limited by the process of non-replacement so that each 50 years is specified
by the same parameter set. The method assumes serial correlation to be zero.
Nevertheless, this was probably the first truly stochastic streamflow
generation model.

By dealing the cards without replacement, Sudler's mass curve passed


through zero at the end of every 50 years. Thus, as for Hazen's mass curve,
the range is curtailed. Because he dealt with finite storages, Sudler does
not have a compensating error as a result of his type of storage.
Consequently, Sudler's technique would tend to underestimate the required,
storage capacity.

In estimating the capacity of the Upper Yarra Dam in Victoria


(Australia), Barnes (1954) found that the annual flows were normally dis-
tributed and independent, and used a Monte Carlo approach to generate 1000
years of data. Historically, this approach contrasts with the earlier
procedures in that they were distribution free methods. Barnes adopted a
design criterion of probability of failure of 1 in 40 but used a semi-
infinite storage approach as an added safety factor.
III

From 1936 to the 1960's, Hurst studied the river Nile, and developed
various card sampling techniques to generate annual flows which were used
in simulated operational studies of the Aswan High Dam. Details can be
found in Hurst's text (1965), pp. 41-42.

5.3 ANNUAL MARKOV MODEL

The Russian mathematician Markov (1856-1922) introduced the concept of


a process in which the probabi I i ty distribution of the outcome of any trial
depends only on the outcome of the directly preceding trial and is inde-
pendent of the previous his tory of the process. In this case the "trial"
is the passage of one year and its "outcome" is the streamflow for that
year. If the probability distribution of annual streamflow is either
independent of previous streamflows or correlated only with the previous
years flow, we have a "simple" or "lag one" Markov process. The concept
has been extended to include cases of lag greater than one. The Markov
process was the basis of the developments at the Universities of Colorado
and Harvard in stochastic streamflow generation procedures during the early
1960's (Julian, 1961; Yevdjevich, 1961; Brittan, 1961; Maass et aL., 1962).

Brittan (1961) proposed the following Markov model to represent


actual streamflows:

(5.4)

where xi+l' xi annual runoffs for (i+l)th and ith years,


x mean annual historical flow,
s standard deviation of annual flows,
annual lag one serial correlation coefficient, and
normal random variate with a mean of zero and
a variance of unity.

This equation was adopted in order that the expected vaZues of the mean,
standard deviation and serial correlation of the computed xi+l's would be
equal to the respective values of those parameters derived from the
historical record and used in the right-hand side of the equation.
Moreover, if the xi values are normally distributed, then it follows
that the xi+l values will also be normally distributed. (Appendix C
shows theoretically that this algorithm does preserve the mean, standard
deviation and serial correlation of the flows.)
112

5.3.1 Practical Considerations

(i) The model CEq. 5.4) consists of two components:


a deterministic or correlation component [x + rl(x
i
- x)]
and a random component [tis (1 - rf)11.

If r = 0, the model is purely random. This sometimes


l
occurs with annual data, for example, as found by
Barnes (1954) in generating 1000 years of inflows for
the Upper Yarra Darn in Victoria. For the model as proposed,
r 1 cannot exceed unity and for annual data is generally
less than 0.4.

(ii) To use the model to generate annual flows, we need to compute


the mean, standard deviation and serial correlation of the
historical annual flows, and to assume that the flows are
normally distributed.

(iii) The normal random variate, ti' is generated by an appropriate


routine which is available for all computers. One method is
to generate pseudo-random numbers which are usually uniformly
distributed with a mean of 1 and variance of 1/12. If we
add 12 of these numbers together and subtract 6 the resulting
variate may be regarded as a normal random variate with a
mean of zero and a variance of unity - designated as N(O,I).

(iv) To initialize the model operation, Xl is set equal to X.


Consequently the first ten or so generated flows should be
discarded as they will be dependent on this initialisation
procedure. A similar initialisation procedure is used for
other variations of the Markov model.

(v) This and some other models can generate negative flows.
When this occurs the negative value is to calculate the
next flO\oJ, after which it is set to zero. Such a procedure
is acceptable so long as the proportion of negative flows
is not too high (say no more than 5%). In addition, one
should check the difference in mean flow of the generated
sequence with the negative values included and with them
set to zero. If the difference is greater than say one
percent, the model is probably unsatisfactory for that stream.

(vi) Sample calculations of annual generated flows for a stream


with the following historical parameters are given in
113

Table 5.1 for the Yarra River at Doctor's Creek (229103)


in Victoria, using:

N = 77 years, x = 180 10 m , s = 72 10 m , r l = 0.12


and a sequence of random numbers, ti' which are normally
distributed with a mean of zero and a variance of one.

TABLE 5.1 Sample calculation of annual Markov


streamflow model.

i
x.
1
X.-
1
x r l (xi- x) x+rl (xi- x) t.
1
\ s (l-rir xi + 1
deterministic random
component component

1 180 0 0 180 -0.52 -37 143


2 143 -37 -4 176 0.61 44 220
3 220 40 5 185 -0.36 -26 159
4 159 -21 -3 177 -0.39 -28 149
5 149 -31 -4 176 0.08 6 182
6 182 2 0 180 -0.93 -66 114
7 114 -66 -8 172 -0.03 -2 170
8 170 -10 -1 179 0.80 57 236
9 236 56 7 187 1.67 119 306

It illustrates very clearly the relative importance in the


model of the deterministic and random components. Even
though the serial correlation is about average, the
fluctuation in the deterministic component is small relative
to the fluctuations caused by the random component. Even
if the serial correlation were 0.5 (an approximate upper
limit for annual flows) the random component would still
contribute 75% of the variance in the generated flows.

(vii) In the above example using annual data, ti is defined as


a random normal variate, N(O,l). However, for many streams
this assumption is not acceptable. (For Australia, it is
valid for only about 20% of streams.) In order to provide
for this non-Gaussian situation, the model can be modified
in several ways which are outlined in Sec. 5.5.
114

In the annual Markov model as outlined above, only two of the four
components assumed to make up the streamflow process, as defined in Eq. 5.1,
are accounted for explicitly. Trend and periodicity are not considered.

How then do we treat trend? Unless there is an a priori reason for


knowing the type of trend, a non-parametric test such as Kendall's rank
correlation procedure (Kendall and Stuart, 1968) should be used to measure
its strength. Trends can be modelled by fitting either polynomials or
moving averages although there are difficulties with both approaches
(Tintner, 1968).

The most common form of periodicity relates to seasonality,


particularly with respect to monthly data generation. Here the most
appropriate practical model is the one proposed by Thomas and Fiering (1962).

5.4 THOMAS AND FIERING SEASONAL MODEL

The algorithm for the Thomas and Fiering seasonal model is as follows:

x '+ 1 + b. (x. - x.) + t. s. 1 (l-r~) ~ (5.5)


J J 1 J 1 J+ J

generated flows during the (i+l)th, ith seasons


reckoned from the start of the synthesized
sequences,
mean flows during (j+l)th, jth seasons within
a repetitive annual cycle of seasons (if months
are being used, then 1 ~ ~ 12),
b. least squares regression coefficient for
J
estimating (j+1)th flow from the jth flow

b. (5.6)
]

t. normal random variate with mean of zero


1
and variance of unity,
standard deviations of flows during the
. l)th , J.th seasons, an d
( J+
r. correlation coefficient between flows in
J
]. th an d (.J+l ) th seasons.

To use the model to generate monthly flows at a site, 36 parameters


monthly means, standard deviations and lag one serial correlations - are
required. These are obtained from analysis of monthly historical flows.
115

To run the model, set xl x


' and compute successively x , x , ...
JAN 2 3
where ti is the only unknown and for each step it is calculated as a
pseudo-random normal variate. Thus,

X
2
xFEB + bFEB/JAN(x l - xJAN ) + tl sFEB (1 - r2FEB/JAN)~ (S.7a)

x3 x
MAR
+ bMAR/FEB(x2- xFEB ) + t2 sMAR (1 - r2MAR/FEB)~ (5.7b)

x
13
= xJAN + bJAN/OEC(X12-XOEC) + t s
12 JAN
(l - r2JAN/DEC)~ (S.7c)

As defined above this model is restricted to normally distributed


flows, that is, ti is considered to be a Normal random variate. In order
to cater for non-normal streams the model can be modified as shown in the
next Section.

5.5 MODIFICATIONS FOR NON-NORMAL STREAMFLOWS

To cater for non-normal annual and monthly streamflows, three


alternatives are available:

(i) modify t. by an appropriate transformation;


1

(ii) modify the streamflow parameters and the model


algorithms such that the final generated data are
distributed like the historical flows upon which they
are based; and

(iii) generate normally distributed flows and apply inverse


normalizing equations.

5.5.1 Modifying tj

In dealing with the problem of skewed data, Thomas and Burden (1963)
transformed the Normal variate, t., to a skewed variate, t , with an
l Y
approximate Gamma distribution (designated as 'like Gamma' in the following
text), using the Wilson and Hilferty (1931) transformation thus:

Y t y2.~ 3
t 2 1 + t,~ i t,J (5.8)
Y [ - 36
Y t,j

where Yt,j coefficient of skewness of the like Gamma variate,

ti N (0,1)
t like G(O,l'Y t .), and
Y ,J
repetitive annual cycles of seasons usually 1 ~ j ~ 12.
116

In order to maintain the historical skewness in the generated flows,


the historical skewness is increased to account for the effect of serial
correlation. Using expectation theory Thomas and Burden derived the
following algorithm to do this.

(5.9)

seasonal coefficient of skewness for


(j+1) th and jth seasons.

To apply this method, called the like Gamma transformation, t. in


1
Eq. 5.5 is replaced by t from Eq. 5.8, y . being calculated with
Y t,]
Eq. 5.9.

The Wilson and Hilferty transformation is an approximation to the


Gamma distribution but it breaks down for large skews and serial calcula-
tions (McMahon and Miller, 1971; Phatarfod, 1976); the limits are given
in Fig. 5.2. In general these limits do not affect annual models but,
for monthly models restrict the use of this method to flow sequences with
low to medium variability.

0.6 ////
Wilson and Hilferty/
transformation /
unacceptable
0.4

0.2

0+-----r-----r----.,----1~~~
2 C 3
s

0.2

FIG 5.2 Limits of applicability of the Wilson and


Hilferty approximation.

Kirby (1972) provided an alternative transformation to Eq. 5.8 which


remains theoretically satisfactory over the whole range of hydrologic
interest.

(5.10)

where A, Band G are a function of skewness and given in


Kirby's paper, and
117

EXAMPLE 5.1

Over the period of record the January flows for the Mitta Mitta
River (Appendix E) exhibit a skewness of 1.8; for February, it is 1.1.
The serial correlation coefficient of monthly flows between January and
February is 0.58. Show how a random number from a Normal distribution
[N CO, 111 can be transformed to a like Gamma skewed variable [G(O,l,1 .)].
t, J
* * * * *
A random number taken from the Normal distribution [N(O,l)J is
- 0.4305.

- r~ I.
From Eq. 5.9: J ]

1.1 - (0.58)3(1.8)
(1 - 0.58 2 )3/2

= 1. 385

From Eq. 5.8: t


Y
-2 -
Yt,j
[1 +
It,j t.1
6 ~
'J
It ,]. 2
Yt, j

1. ~05 [
1 +
(1. 385) (- 0.4305)
6
_ (1.385)2] 3
36
2
1.385

=- 0.5655

That is, the corresponding number to the normally distributed - 0.4305 is


skewed gamma distributed - 0.5655. Random numbers transformed in this way
can be used directly in the Thomas and Fiering seasonal model (Eq. 5.7).

5.5.2 Moment Transformation Equations

Matalas (1967) presented moment transformation equations which


theoretically preserve the moments and the lag one serial correlation
coefficients. This method assumes that the Zogarithms of the flows are
normally distributed. Thus the procedure is first to calculate a series of
logarithms using a Normal model, and then obtain absolute flows by
exponentiation. The generating algorithms and the parameter estimation
equations for the three parameter log-normal model are as follows:
118

Generating Algorithm:

- j+l + Bj (X i - X-].) + t i Sj+l (1 - ]


X R2.)! (5.12)

(5.13)

where t. N(O,l),
l

Xi+l generated flow logarithms, and

xi+l generated flow in absolute units.

Other symbols are defined below.

Parameter Estimation:

x.
J
Aj + exp (0.5 sj + Xj ) (5.14)

s~ exp [2(S~ + it.)] - exp (S~ + 2X.) (5.15)


J J J J J
exp [3S~] - 3 exp [S~] + 2
g. J J (5.16 )
J {exp [S2] _ l} 3/2
J
exp [So S. 1 R.] -
] J+ J
r. (5.17)
J

where
Historical Log-transformed
data value
mean x.
J
standard deviation S.
J
coefficient of skewness gj
lag one serial correlation r.
J

and B. R. Sj+lJ
S (5.18)
J J [ j

To solve for A., X., S., R. begin with Eq. 5.16 and solve for S ..
J J J J J
This is not explicit in S.and an iterative solution is required. One fast
J
converging technique is the Newton-Raphson method providing a reasonable
initial guess is used. The procedure is given in Appendix D. Once S. has
J
been determined, then use Eqs. 5.14, 5.15 and 5.17 to obtain xJ., A. and R..
J J
119

EXAMPLE 5.2

Use moment transformation equations to transform the monthly para-


meters for the Mitta Mitta River (Appendix E) so as to preserve these
characteristics in a generation model using normally distributed random
numbers.
* * * * *
For January the historical parameters are (from Appendix E):
3
mean, x 39.9 (x 106)m
3
standard deviation, s 26.4 (x 106)m
serial correlation, r 0.58
skewness, g 1.8

From Eq. 5.16 :


exp 3S~ - 3 exp S2 + 2
J J
gj
{exp S~ - 1}3/2
J

This cannot be solved explicitly for Sj; a trial and error procedure is
required. If a reasonable first trial value is obtainable, the Newton-
Raphson procedure (Appendix 0) gives rapid convergence to the solution.

Using a starting value of 0.5 for Sl; that is, si = 0.25


3/2 (Eq. 05)
f(Sp [exp (3Sf) 3 exp (Sf) + 2] - gl (exp Sf - 1)
3/2
exp (0.75) - 3 exp (0.25) + 2 - 1.8 [exp (0.25) - 1]

0.0075.

3 exp 3Sf - 3 exp Sf - 1.5 gl(exp Sf - 1)0.5 exp S1 (Eq. D6)

3 exp (0.75) - 3 exp (0.25) - (1.5)(1.8)[exp (0.25)-1]0.5exp (0.25)

0.6513.

- 0.0075
Therefore, second estimate 0.25 - 0.6513

0.2615.

Similarly, after repeating the above, the third estimate is found to


be 0.260752 and the fourth, 0.260747, (that is, convergence).

Thus Sf 0.26075 and Sl 0.5106.


120

Equation 5.15 can be rearranged to give:

0.5 log [26.4 2 /(exp (2 x 0.2607) - exp (0.2607)]

3.748.

Equation 5.14 can be rearranged to give:

39.9 - exp (0.5 x 0.2607 + 3.748)

8.469.

Rl cannot be obtained from Eq. 5.17 until S2 is found from Eq. 5.16
(S2 0.3419) .

Equation 5.17 can be rearranged thus:

log [0.58,j{exp (0.5106)2 - l}{exp (0.3419 2 )- 1}+ 1]/(0.5106)(0.3419)

0.6054.

Similarly the computations can be done for the other 11 months; the
results are given in Table 5.2.

TABLE 5.2 Log-transformed parameters for the Mitta Mitta River.

Log-Transformed Parameters
Month
Mean Standard A. Serial
]
Deviation Correlation

1 3.748 0.5106 - 8.469 0.605


2 3.720 0.3419 - 16.83 0.658
3 3.323 0.7546 - 5.774 0.626
4 3.856 0.9124 - 24.88 0.892
5 3.879 0.8424 - 6.860 0.854
6 4.657 0.7420 - 36.25 0.684
7 5.697 0.3945 -169.3 0.609
8 5.833 0.3944 -170.2 0.665
9 5.842 0.2829 -159.7 0.737
10 6.390 0.1893 -399.9 0.654
11 5.990 0.2018 -273.0 0.805
12 4.445 0.4891 - 23.39 0.639
121

If a two parameter log-normal model is used, gj will be assumed zero


in which case A. = 0 and Eq. 5.13 becomes
J

(5.19)

and the model parameters for input into Eq. 5.12 can be determined
explicitly from Eqs. 5.14, 5.15 and 5.17. The model modified in this way
is based on a two parameter log-normal distribution rather than the three
parameter one.

For annual data generation, Eqs. 5.12 to 5.17, which are set down
above with monthly subscripts, are modified appropriately.

An important limitation arises in applying two parameter models to


streams with high coefficients of variation. For this distribution, the
coefficient of skewness (C ) is related to the coefficient of variation (C )
s v
in the following manner (Chow, 1964, p. 8-17):

= 3C v + C
v
3 (5.20)

In practice, the implied values of skewness are modified among other


things by the effect of serial correlation and so the generated value is
always less than the value given by Eq. 5.20.

5.5.3 Normalizing Flows

This procedure was proposed by Beard and has been adopted by the
United States Army Corps of Engineers (Beard,1972). The following equations
are for annual flows:

(i) Compute logarithms of all flows after a small increment


has been added to each in order to eliminate zero values.

(ii) Compute mean, standard deviation and coefficient of


skewness of log flows.

(iii) Standardize the log values

v L-=-.Z (5.21)
s
y

where v standardized log flow,


y loge (x + E)
y mean of loge (x + E) flows,
s standard deviation of loge (x + E) flows,
y
X historical flows, and
E small increment.
122

(iv) Normalize the standardized values, v, to eliminate


skewness using the inverse Wilson and Hilferty
transformation thus:

v = ~ {(~ v + 1)1/3_ I} + ~ (5.22)


g 2 6

where v normalized values, and


g coefficient of skewness of the log values.

(v) Compute serial correlation of normalized values.

(vi) Generate standardized variates by the Normal Markov process

(5.23)

generated standardized variates N(O,l),


lag one serial correlation of normalized
variates, and
ti = Normal random deviate N(O,l).

(vii) Apply inverse transformations as follows:

v = {[! (v ~) + 1]3 - l} -2 (5.24)


6 g
x = exp (y + vs ). (5.25)
Y
(viii) Subtract the small increment (E) added in step (i) . If negative
flows result, set them to zero.

Because of the initial decrease in skewness as a result of taking


logarithms, the procedure does not suffer from the Wilson-Hilferty limi-
tation (Fig. 5.2) but it has been found that the serial correlations of the
absolute flows are poorly :,reserved. These and other aspects are covered
more fully in Sec. 5.8 which deals with the performance of these models.

For monthly flows, the mean, standard deviation and toefficient of


skewness need to be modified. Details are given in Beard's (1972) report.

EXAMPLE 5.3

Apply the normalizing flow procedure to the annual flows for the
Mi tta Mi tta River (Appendix E) to demonstrate this method of flow
generation.
* * * * *
123

Following the steps listed in Sec. 5.5.3 and referring to Table 5.3:

(i) Add 0.01 (that is, s = 0.01) to all of the annual flows
in column (2) of the table, and enter the natural logarithm
of each in column (3).

(ii) Standardize the flows using Eq. 5.21. First, calculate:

mean of Column (3) 7.00121


standard deviation of Column (3) 0.559156
Column (3) - 7.0012
then Column (4)
0.5592

(iii) Use Eq. 5.22 to normalize the flows. First, calculate


skewness of Column (3) = - 0.0825

then Column (5) = _ 0.0825


6 [(-0.0 825 (Column (4))
2
+ 1) 1/3 - 1]
- 0.0825
+ 6
(iv) Use Eq. 5.23 to generate standardized variates. First,
calculate serial correlation of Column (5) =- 0.008.
Next, assume an initial value of vi; zero is appropriate.
Hence, for a random normal variate (say tl = 1.0752)

0.008(0) + 1.0752 (1 - 0.008 2)t

1.0752;

for t 0.0064, say,


z= -

0.008 (1.075Z) - 0.0064 (1 - 0.0082)~


0.0064.

Note: The very low serial correlation means that the vi+1 t ..
l

(v) Apply the inverse transformation Eq. 5.Z4.

V
z = {[- 0.~825 (v
2
- - 0.~825) + 1r - 1} _ 0.~825
- 1.1053.

From Eq. 5.25,

x exp (y + v
2
s)
y
exp (7.00121 - 1.1053 x 0.5592)

591.76
124

(vi) Hence the first generated flow in the sequence is 591.76 - 0.01
= 591.75 ~ 592. Subsequent flows are calculated in the same
way.

In normal application the first few generated flows would be dis-


carded to remove the effect of the starting condition applied in (iv).

TABLE 5.3 Derivation of normalized annual flows for the


Mitta Mitta River.

Year Flow Loge(Flow + 0.01) Standardized Normalized


x 10 6m3 Flows Flows
(1) (2) (3) (4) (5)

1936 1553 7.3480 0.6202 0.6120


1937 650 6.4770 - 0.9374 - 0.9392

1969 1010 6.9177 - 0.1493 - 0.1639

5.6 TWO TIER MODEL

Monthly models do not necessarily preserve the annual flow charac-


teristics. To overcome this deficiency, Harms and Campbell (1967) extended
the Thomas and Fiering model to constrain the annual and monthly flows
separately, and also to preserve the annual serial correlation. Annual
flows were generated by an annual Normal Markov model. A Thomas and
Fiering log-normal model was used for monthly generation, but the values
were adjusted to sum to the appropriate annual values by the following
algori thm:
x ..
x! . 1J
1] -12-- Qi (5.26)
I x ..
j =1 1J

where x'. adjusted monthly generated flow volumes,


1J
X
ij unadjusted monthly generated flow volumes,

Qi annual generated flow volumes,

i year, and

month.
125

Results presented in the Harms and Campbell paper suggest that the
model works well. In cases where the annual flows are not normally dis-
tributed, a skewed distribution could be used in place of the normality
assumption. One minor drawback with this approach is that the method of
adjusting monthly data does not allow the monthly serial correlation
coefficient from the end of one year to the beginning of the next to be
preserved.

5.7 OTHER CONSIDERATIONS

Many other considerations are involved in stochastic generation of


streamflow other than those aspects treated in this chapter, for example,
correlograms, partial correlation functions, spectral analysis, and daily
models. Nevertheless, the annual and monthly models outlined above do
provide a basis for the practical application of data generation techniques
to single site situations. In this text it would be out of place to
consider in detail multi-site models but some background material is given
in Chapter 7.

This discussion would be incomplete without a brief comment on the


use of so-called long memory models. Three data generation models, the
ARIMA, Broken Line and Fractional Gaussian Noise models fall into this
category. They have been proposed as replacements for Markovian schemes in
order that the Hurst phenomenon (see Sec. 3.3.1) is preserved in the stream-
flow sequence. In Markovian models the exponent K in Eq. 3.4 tends to
0.5, yet in reality its observed mean value is about 0.72.

But are these three models of importance to the practising water


engineer? High K values in Eq. 3.4 imply long-term persistence and result
in longer and more extreme events in the flow sequence. From Eq. 3.4, high
K values result in larger ranges than would otherwise occur, and con-
sequently relatively larger storage sizes. What is not clear is how
important are these effects. For example, Wallis and Matalas (1972)
observed that differences in storage estimates between the Markovian and
fast Fractional Gaussian Noise models occurred only for drafts greater than
80%. Yet Kottegoda (1970) found for British rivers that the Fractional
Gaussian Noise model gave unrealistically high estimates of storage if
compared with those estimated from historical or Markovian sequences. While
there are these and other differences in detail, the consensus of opinion in
the literature suggests that K should be preserved for high drafts and
high reliabilities.
126

5.S MODEL VERIFICATION AND PERFORMANCE

Before a data generation model is used in storage-yield analysis it


is necessary to check not only that it satisfactorily reproduces the main
statistical characteristics defining the streamflow process, but also that
critical periods are being satisfactorily generated. A validation procedure
for an annual or a monthly model might include the following tests:

(i) comparison of the mean and variability of various


statistics (annual and monthly means, standard deviations,
coefficients of skewness and serial correlations) computed
from many sets of generated data (each of length equal to
that of the historical record) with the actual values of
those statistics computed from the historical record;

(ii) comparison of flow duration and frequency curves based on


the generated data with the corresponding curves based on
the historical record;

(iii) comparison of correlograms based on monthly generated data


with that derived from the historical record; and

(iv) comparison of the mean and variability of reservoir storage


estimates based on replicated generated data compared with
estimates using historical data.

The number of replicates of generated data that are required will


vary with respect to streamflow variability. Twenty-five are generally
sufficient.

In the above tests, it should be noted that the historical values


from (i) and the flow distribution (used in the flow duration comparison)
in (ii) are part of the model structure, and, therefore, are more likely to
be satisfactorily modelled than the other factors listed.

To illustrate the level of performance achieved with the Thomas and


Fiering model, published results (McMahon et al., 1972a, 1972b, 1973) have
been included here as Tables 5.4, 5.5 and 5.6, In addition, results from
a recently completed evaluation of a number of annual Markov models are
available (R. Srikanthan, personal communication, 1977).

In Tables 5.4 to 5.6 historical monthly flows and storage-yield


results are compared with results from generated data using the Thomas and
Fiering monthly model incorporating three of the distributions discussed
TABLE 5.4 Comparison of historical and generated monthly and annual streamflow parameters

River M:JNTHLY ANNUAL


Australian Stream Mean Standard Standard
Gauging Station No. Distribution Seri al Serial
(Mm3) Deviation Skew Deviation Skew
(Years of Record) Correlation Correlation
(Mm3) (Mm 3 )
O'Shannassy Historical 8.8 5.7 1.2 0.76 31 0.7 0.04
229103
(59) LGLT 8.8 5.8 1.6 0.78 32 1.1 0.13
(0.4) (0.6) (0.7) (0.03) (6.4) (0.9) (0.13)
LN-2 8.9 5.9 1.S 0.78 32 1.0 0.06
(0.5) (0.5) (0.3) (0.02) (4.2) (0.4) (0.15)
LN-3 9.1 7.3 2.3 0.74 42 0.9 0.08
(0.5) (0.7) (0.5) (0.02) (6.0) (0. S) (0.15)
Gordon Historical 150 109 1.4 0.38 420 -0.1 0.06
308007
(36) LGLT 153 113 1.6 0.38 390 0.4 0.00
(5.4) (8.3) (0.5) (0.06) (66) (0.5) (0.18)
LN -2 152 III 1.9 0.39 390 0.7 -0.07
(5.8) (12.2) (0.8) (0.07) (79) (0.6) (0.14 )
LN-3 152 III 1.3 0.37 390 0.4 -0.01
(3.8) (4.6) (0.3) (0.05) (60) (0.6) (0.19)
Torrens Historical 4.2 7.8 3.1 0.57 39 1.5 0.02
504501
(72) LGLT 4.3 9.9 5.9 0.56 49 2.5 -0.02
(0.4) (1. 5) (2.0) (0.07) (9.6) (0.9) (0.07)
LN-2 4.4 9.0 4.9 0.61 43 2.4 -0.02
(0.4) (1.5) (1. 8) (0.05) (10) (1. 1) (0.10)
LN-3 4.7 9.3 3.6 0.60 43 1.5 0.00
(0.4) (0.9) (0.7) (0.04) (5.6) (0.8) (0.12)
Warragamba Historical 93 200 4.6 0.45 1200 2.7 0.29
212240
(72)
LGLT 110 560 15.0 0.34 2500 4.4 0.04
(35) (690) (8.0) (0.16) (2500) (2.0) (0.15)
LN-2 91 170 6.2 0.47 920 2.3 -0.03
(9.0) (35) (1. 7) (0.07) (230) (1.2) (0.08)
LN-3 110 190 4.2 0.43 970 1.7 0.00
(10) (32) (1.3) (0.06) (190) (0.7) (0.08)
Generated parameters are means of 20 replicates of length shown. Values in parentheses are standard deviations
of generated parameters.
TABLE 5.5 Comparison of historical and generated monthly parameters for O'Shannassy River in Victoria

Statistic Model J F M A M J J A S 0 N D

Historical 5.4 3.8 3.7 4.2 6.5 8.8 12.1 14.9 14.3 13.0 10.6 8.3

LGLT 5.4 3.8 3.7 4.2 6.7 8.8 12.1 14.9 14.3 13.0 10.5 8.1
Mean
(Mm3) LN-2 5.4 3.9 3.8 4.2 6.4 8.5 12 .5 15.3 14.7 13.1 10.6 8.0
LN-3 5.7 4.4 3.8 4.2 6.4 8.5 13.3 16.9 14.3 13.3 10.7 7.5

Historical 2.0 1.1 1.2 1.9 3.7 5.1 5.2 5.7 5.2 4.9 4.7 4.8

Standard LGLT 2.0 1.0 1.2 1.9 4.3 5.3 5.6 6.0 5.3 4.8 4.6 4.2
Deviation
CMm3) LN-2 2.3 1.5 1.4 1.7 3.3 4.6 6.3 6.8 6.2 5.1 4.3 3.9
LN-3 2.7 3.3 1.7 1.6 3.6 4.6 5.3 13.4 5.3 6.0 4.6 2.7

Historical 1.2 0.5 0.8 1.6 1.7 2.0 0.9 0.4 1.0 0.7 0.9 3.4
Skew
LGLT 1.3 0.5 0.8 1.4 2.4 1.9 1.3 1.0 1.1 0.5 l.0 1.9
LN-2 l.2 l.0 0.8 1.1 1.3 1.4 1.4 1.2 l.2 0.9 1.0 1.3
LN-3 1.3 1.2 0.9 1.0 1.3 1.4 1.3 1.1 0.9 0.7 0.8 l.4

Historical 0.86 0.64 0.56 0.44 0.64 0.82 0.68 0.73 0.61 0.76 0.66 0.70

LGLT 0.86 0.69 0.59 0.39 0.77 0.87 0.67 0.72 0.77 0.77 0.73 0.90
Serial
LN-2 (j.92 0.68 0.51 0.33 0.62 0.87 0.79 0.80 0.66 0.75 0.50 0.80
Correlation
LN-3 0.98 0.81 0.42 0.40 0.61 0.94 0.94 0.73 0.75 0.74 0.33 0.85

Note: Generated parameters are mean values of fifty replicates.


129

TABLE 5.6 Reservoir capacity estimates based on generated


flows compared with historical values.

River Annual Model


Australian Stream coefficient
Gauging Station No. of LGLT LN-2 LN-3
(Years of Record) variation

O'Shannassy
a a a
229103 0.29 94 97 l57
(SO)

Gordon
0.23 90 74 92
308007
( 7) (7) (10)
(36 )

Yarra
0.44 86 85 116
229103 (10) (8) (12)
(50)

Torrens
0.77 94 59 89
504501
(19) (13) (13)
(72)

Warragamba
1. 07 107 53 60
212240
(24 ) (20) (12)
(72)

aMedian value based on ten replicates.

Note: Storage estimates for conditions of 1% probability of failure


and 50% draft are expressed as percentages of the long-term
historical Gould values. Generated storage estimates are
means of 20 replicates of length shown. Values in parentheses
are standard deviations expressed as percentages of appropriate
mean value.

earlier - two and three parameter log-normal (denoted by LN-2 and LN-3) and
the like Gamma distribution (LGLT). In using the latter distribution a
logarithmic transformation was initially applied to the data. These tables
highlight a number of points:

(i) In Tables 5.4 and 5.5, except for LGLT for Warragamba and the
LN-3 August standard deviation, the annual and monthly means
and standard deviations compare well with the historical
estimates.

(ii) Coefficients of skewness are generally too high but are


considered satisfactory except for the LGLT monthly estimate
for Warragamba. Seasonal historical skews are erratic and
the variations are poorly modelled (Table 5.5). Some workers
recommend smoothing seasonal coefficients of skewness (and
other moments) prior to analysis (Beard, 1965).
130

(iii) Monthly serial correlations are well modelled (Tables 5.4


and 5.5) yet high annual serial correlations are poorly
simulated, for example, Warragamba in Table 5.4. This
inadequacy is typical of monthly Markov models and can
be overcome by using the two tier approach outlined in
Sec. 5.6.

(iv) In Table 5.6 reservoir capacity estimates using Gould's


procedure (Sec. 4.6) and based on generated flows are
compared with historical values. The results show large
discrepancies. Overall,the storage values from the LGLT
model compare most favourably with the historical estimates.
At least two of the storage estimates for the LN-3 model
deviate considerably from their historical values. The
LN-2 model shows even greater discrepancies and is
considered satisfactory for only one river.

From this analysis, it can be seen that using the historical values
as a basis of comparison, the LGLT model shows least variations yet exhibi ts
some unsatisfactory parameter estimates. On the other hand, the LN-2 and
LN-3 models reproduce the parameter values, but deviate markedly from the
historical storage estimates. Thus no model is wholly satisfactory.

Tables 5.7, 5.8 and 5.9 deal with a more detailed evaluation of
Markov models than that discussed above, although the evaluation was
restricted to the annual lag one type (R. Srikanthan, personal communication,
1977). In all, 16 rivers * which represent the range of streamflow varia-
bility. encountered across the Australian continent were examined. Up to
seven variations of model distributions were considered and for each case
5000 years of data were generated. In addition to the parameters and
characteristics listed at the beginning of this section results were
examined for the range, Hurst's H and K exponents, run lengths, extreme
events, spectral values and distribution types.

* In Table 5.7, 5.8 and 5.9, the names and national stream gauging numbers
refer to the following Australian rivers as follows: (1) King (309001),
(2) Wilmot (315003), (3) South Johnstone (112101), (4) Yarra (229103),
(5) Murray (401201), (6) South Esk (318001), (7) Wungong (615071),
(8) Serpentine (615074), (9) Loddon (407203), (10) Torrens (504501),
(11) Ord, (809302), (12) Peel (419004), (13) Warragamba (212240),
(14) Burnett (136001), (15) Wide Bay Creek (138002) and (16) Goulburn
(21006) .
131

TABLE 5.7 Comparison of historical parameters with generated


parameters based on annual flows generated using
an annual Markov model.

Parameter Di5tr1- f-_____________Ri_v_e_'_'_e_fe_'_en_'_e_n_o_._ _ _ _ _ _ _ _ _ _ _ _ _ _~


and
button
10 11 12 13 14 15 16

Hist. 2230 1620 1710 540 )60 238 214 122 61 144 94 209 120 44 126 27
::nm)
2350 1610 1720
530 61 205 121 43 124 26
2320 1620 1710 540 370 236 210 119 61 132 96 210 123 44 127 26
119 43 124 26
LN-2 540 )70 239 216 123 61 145 96 209 123 45 127 27
LN-3 540
540
)70
370
239
240
217
216
123
122
62
63
145
146
97
97
213
214
124
124
46
46
132
131
028

Hist. 0.18 0.22 0.38 0.40 0.47 0.47 0.50 0.57 0.78 0.78 0.80 0.82 1.11 1.14 1.26 1.79

0.18 0.22 0.37

0.18 0.22 0.38


0.39
0.40 0.46 0.47 0.49 0.56
0.77
0.74 0.78 0.79
0.81
0.79
1.11
1.09
loll
1.11
1, 15
1. 22
8
1. 71
1.07 1.13 1.23 1.64
LN-2 0.40 0.46 0.47 0.49 0.56 D,75 0.76 0.76 0.77 1.02 88~
LN-J 0.40 0,46 0.47 0.49 0.56 0.74 0.76 0.76 0.78 1.05 1.07 1.158
0.40 0.46 0.48 0.46 D.578 0.79 0.86 0 1 . 1 1 1.24 1.33 1.80

Hist. -0.2 0.3 0.2 0.8 1.3 0.9 0.6 1.0 0.9 1.5 1.2 1.1 2.8 2.4 2.5 3.6

C":\r,:-...-0
-0.2 0.4 0.2
0.7
0.7 1.2 0.9 0.6 0.8
0.9
0.9 1.4 1.2 ::: ~~~:: ~
2.32.0 2.6 ~
LN-2
LN-3
-9 0.7
1.]
1.2
GC08C0GG@
A
0.8 @ 0.9 0.9 1.3 1.1 1.1
2.7 2.6
1.9
@)~
2.5 1.9
0.8 1.] 1.2 ~ee 1.5 @8 2.9 2.9 2.8 'd
Hist. -0.09 -0.06 -0.10 0.12 0.11 0.01 0.21 0.21 0.14 0.02 0.07 0.24 0.30 0.20 0.28 0.18

-0.10 -0.06 -0.10


, 0.13 0.14 0.25 0.31 0.1'.1 0.29 0.16
80,22
G LN-2
-0.088-0,12 0.13

0.11
0,10 8 0 , 2 1

0.09 0.00 0.20


0,22

0,19
0,16

0.13
0.01

0.01 0.06 0.22


0,30
0.30
0,]0
0,20
0.22
0.20
0,27
0,27
0,27
0.18
0.16
0.19

wh~r~
LN-3

__ -

~~ ~_-=--_~- _98 o,o~~~~_~~~ S88888


,0[ .. normal distributlon
0,11 0,09

W - Weibull
0,00 0.20 0,19 0.13 0.01 0.06 0.22 0.29 0.19

G .. like Gamma distribution


0.26 0,16

K .. Kirby's transformation LN-2 ,. T.... o parameter log normal LN-3 = Three parameter log nonnal
B .. 8eard's method
1.)Z

TABLE 5.8 Percentage of negative flows in generated sequences


from annual Markov model and their effect on mean
flow estimate.

Parameter Distri- Ri ver reference no.


and bution*
error 9 10 11 lZ 13 14 15 16

N - - - - - - - -

%
W @]) - - Ci]) CQ)~@
e g C8) 0.9 6.5 ~
~
negative G 1.6
flows
K - - - - o ~ 16. 3l.
8 LN-2 0 0 0 0 0 0 0 0
LN-3 g 3.3
80 QJ) cQ)@ @
B 0 0 0 0 0 0 0

N - - - - - - - -

9 g
~~
W 0.9 - - 0.6
%
increase
G C[D 0.1 0.7
CO> 0.1 0.4 1.6 5.8
in x K - - - - 0 0.6 Z.O 3.8

0 LN-Z
LN-3
0
C[D
0
0.4
0
0.8
0

cQ)
0
0.9
0 0
cQ) CQ) ~
0

B 0 0 0 0 0 0 0 0

* For definition of symbols, see Table 5.7.

The characteristics compared in Tables 5.7 to 5.9 include the mean


flow (x), the annual coefficients of. variation (C ) and skewness (C )' the
v s
annual serial correlation coefficient (r), storage capacities for 50% and
90% draft rates (Sso and S90) using the Sequent Peak method, the rank one
two-year and ten-year consecutive low flows (1:2 and 1:10), the percentage of
generated negative flows and the percentage increase in mean in setting them
to zero. In all, seven distributions or variations are examined - Normal
eN), two and three parameter log-normal (LN-Z and LN-3), like Gamma (G);
Kirby's modification (K), Beard's normalizing procedure (B) and a Weibull
distribution (W). The values given in Tables 5.7 to 5.9 are averages based
on replicates in length equal to the historical records and equivalent to
5000 years of flow. To assist in evaluating the performance of the various
distributions, generated values outside 5%, 10% or 25% from the
appropriate historical values are circled in three tables. These limits
indicate an arbitrary level of performance.
133

TABLE 5.9 Comparison of historical low flow and storage


characteristics with those based on annual flows
generated using an annual Markov model
(in units of mean annual flow) .

Parameter River reference no.


Distri-
and
bution
10 II 12 13 14 15 16

Hist. 0.00 0.00 0,23 0.18 0.20 0.23 0.50 0,50 1.1 0.67 l.2 1.2 1.5 1.5 1.1 2.4

0, 01 88 - - - - - -

1.~ 8~
~
0.18 3.1
5" 8 0.01 0. 23 8.21 GO.57 o l.~ ~:~ ~: 85
1.4 8 .. 5.1
4.5

G LN-2
LN-3
0.14 8~ @ 0.56@8
9880.58
0.21 0.20
e@ 0.69 1.3 1.00 1.1 1.4
1.7 1.4

e
2.6

0.20 0.22 0.24 0.50 0,58 1.2 0.86 1.0 l.0 1.6 1.6 2.1

Hist. 0.43 0.48 0.94 1.8 2.2 1.7 2.2 3.3 5.2 3.5 5.9 5.9 13.7 8.8 11.1 13.9

(!.43~-
6.2
8 14 . 2 10.2 10.1 IS.1

5" 0 : 9 0 . ; 7 8 8 5 2.5
eG@ 6.3
8 4.9 6.2 12.8
13.78
10.7 12.7
8
G
12,0 17.1
LN-2 2.0 2.5 4.7 6.8 12.2 10.0 11.6 15,1

LN3 2.2 2.6 5.1 6.9 12.5 9.8 11.0 12.4


2.0 2.4 6.1 6.5 11.7 9.7 10.4 13.4

His!. 1.4 1.3 1.1 1.10 0.94 0.77 0.63 0,58 0.23 0.33 0.32 0.28 0.21 0.23 0.11 0.30

1.5 1.3 0.9 - - - - - -

8 8 8 '~0.02
0.88 0.02

"'
1.4 J.4 0.9
8 O,8S D,79 0.56 0.08
.
0.26
.
0.01 0.11
0.10

G LN-Z

LN'
0.97
0.84
0,86
0.82
0. 90
0.75
8
0.54
0 . 72
0.46 ~~(B@5
0.11 0.23 0,19 0.11
0.27
0.03
0.88 0.86 0.86 0.60 0.54 0.23 0.36 0.36 0.32 0.26

Hist. '.9 9.2 8.9 7.9 7.2 7.7 7.7 6.3 '.6 6.0 5.4 4.8 4.0 4.' '.7 2.9

9.1 8.9 8.2


7.4 2.7

'10 9.0 8.9 8.2 7.3 7.2 7.2


8 6.3

G LN-2
LN,
7.7
7.6
7.4
7.4
7.6
7.5
7.1

@
6.8
6.5
7.7 7.' 7.6 7.1 0.8

For definition of symbols, see Table 5.7.


134

Overall the models preserved the mean and coefficient of variation


satisfactorily but the coefficient of skewness was poorly modelled for the
two parameter log-normal distribution and Beard's normalizing procedure.
The overestimation of LN-2 is expected because the implied skewness given
in Eq. 5.20 is always considerably greater than the observed skewness for
C greater than unity (McMahon, 1975, p. 384). The fourth input parameter
v
to the models - serial correlation - is well preserved except for Beard's
model which underestimates all but one historical value. A further relevant
factor relates to the proportion of negative generated flows and their
effect on the mean. This is shown in Table 5.8 for the eight streams with
a coefficient of variation greater than 0.7. Those with a coefficient of
variation less than 0.7 generate less than one percent negative flows, and
were not tabulated. The table shows that except for the two parameter log-
normal distribution and Beard's model, all others generate too many negative
flows and are regarded as being unsatisfactory. From Tables 5.7 and 5.8 we
conclude that for annual data generation streams with an annual coefficient
of skewness of less than 0.3, the most suitable model is like Gamma; for
other streams with an annual coefficient of variation of less than 0.7, the
three parameter log-normal or like Gamma are suitable; for streams with
larger variability, the two parameter log-normal model is recommended.

Unlike the generated parameter values which should approach the


historical values input into the model, the correct generated values for
other characteristics like low flow values and storage capacities are
.unknown. Nevertheless, if one takes several sets of flows drawn from a
range of geographic, climatic and hydrologic regions as in Table 5.9,
estimates of generated characteristics based on historical input parameters
should be a reasonable approximation overall of the historically observed
characteristics. Accepting this approach and examining the two storage
values (S50 and S90) and the 2 year and 10 year low flow sums (~2and ~10)

in Table 5.9, it is concluded that Beard's procedure performed the most


satisfactorily, followed by the two and three parameter log-normal distri-
butions. Generally all the models overestimated the historical storage
values but Beard's estimates are on the average no more than 10% different
to the historical estimates. This is considered to be very satisfactory.
For streams with low skewness (say Cs < 0.3) the Gamma model is recommended.
135

This detailed review confirms our observations regarding the results


in Tables 5.4, 5.5 and 5.6, namely that no model 'is wholly satisfactory for
all purposes. Consequently, if one is using a data generation model in
practice, one needs to understand clearly the objectives of the study which
should influence model choice.

5.8.1 Unrepresentative Streamflow Data

Input parameters to generating models are based on historical data.


If the parameters are not representative of the flow population, generating
more data will not improve the relevant information. For this reason it is
important to use the "best" estimates of the parameters. Modellers should
be wary of bias in historical data. If suspected, it may be necessary to
employ regional techniques to estimate the parameters (Benson and Matalas,
1967). A similar approach is recommended for ungauged catchments.

5.9 SIMULATION

Simulation analysis is defined as " a process which duplicates the


essence of a sys tern or acti vi ty wi thout actually attaining reality i tsel f"
(Hufschmidt and Fiering, 1967). This involves developing an algebraic model
of all the inherent characteristics and probable responses of the system to
an operating rule. Simul ation analysis has been described as a "brute
force" fitting technique (Fiering, 1961); however, planners are often
forced to use the method to effectively deal with large and complex systems
that become intractable with analytical techniques.

Digital simulation does have several limitations which need to be


recognised.

(i) Where a large number of variables has to be optimised, it


may become infeasible to examine all possible combinations
(Dorfman, 1965).

(ii) It is a trial and error approach that does not necessarily


lead to an optimal solution (Maass et al., 1962).

Notwithstanding these limitations simulation analysis using sto-


chastically generated data as input is a very powerful tool particularly
for a large complex system; aspects relating to multi-reservoir systems
are dealt with in Chapter 7.
136

5.9.1 When and How to use Generated Data

In reservoir capacity-yield analysis data generation procedures should


be used to provide alternative yet equally likely flow sequences to the
historical one. In a large number of generated sequences, some will contain
less severe droughts than the historical record and some more severe. When
the sequences are used in simulation or behaviour studies with a range of
assumed storages and demand values a quantitative picture of the probability
of failure-storage-yield relation can be built up.

The point is best illustrated by an example. If a behaviour analysis


(see Sec. 3.2.3) is carried out on the Mitta Mitta River (Appendix E) for
various combinations of draft and probability of failure a series of curves
will be obtained (Fig. 5.3). From such a diagram the storage required for
any given draft and probability of failure can be obtained; for example,
for 75% draft and 5% probability of failure, the storage required is
760 x 10 6m3 .
100~---------------------------------- ____________________,
Probability of failure
10% 5% 2%
90

80

70

~ 60

.:::<II
5 50

40

30

20

10~----'-----'-----r----Lr-----r-----r---~r---~----~----~~
o 200 400 600 800 1000 1200 1400 1600 1800 2000
Storage (106m 3)
FIG. 5.3 Relationship of draft and probability of
failure to reservoir capacity for the
Mitta Mitta River.
137

This single estimate of the storage required (that is, 760 x 10 6m3)
is based on the historical record which itself is only a sample of the total
flow record for the river. It is probable that other flow sequences, with
the same mean, standard deviation, and serial correlations, and equally
likely to occur could result in different storage estimates.

Table 5.10 shows the storage estimated from flow sequences generated
using a three parameter log-normal distribution. The historical parameters
were converted to the equivalent log-normal parameters using moment trans-
formation equations (Sec. 5.5.2) and each sequence generated was of the
same length as the historical to avoid the effect of storage estimate
increasing with record length (Sec. 3.2.1). To estimate the storage a
fixed release (75%) of the historical mean flow was used.

It will be noted from the table that the mean flows of the generated
sequences vary, but the overall average (1282) is consistent with the
historical mean flow (1274). The storage estimates vary widely, however,
being quite sensitive to the mean flow (and other parameters) of the
corresponding sequence.

The values of storage should be Extreme Value distributed (Burges and


Linsley, 1971) and are shown plotted on Extreme Value paper in Fig. 5.4.
1400

1300

1200
x
x
1100 x

1000

;; 900 x
E
<0
0 600 x

Gl
700

'"~ 600
0
x
iii
500

400 x

300

200
099 0.95 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.15 O. I 0.07 0.05 0.03 0.02
Risk that storage estimate is too small

FIG. 5.4 Risk that indicated storage is too small for


design conditions on Mitta Mitta River.
138

TABLE 5.10 Storage estimated for generated flow


sequences for the Mitta Mitta River
(Appendix E). (Draft = 79.6 x 1Q6 m3/ mon th,
probability of failure = 5%, sequence
length = 34 years.)

Probability of
Mean annual Storage estimate being
Sequence flow estimate too low

(l06 m3 ) (l06 m3) (= rank)


25

1 1205 575 0.68


2 1186 620 0.56
3 1404 891 0.24
4 1213 745 0.40
5 1253 666 0.52

6 1244 783 0.36


7 1093 1181 0.08
8 1214 801 0.28
9 1508 401 0.96
10 1261 1109 0.12

11 1460 501 0.80


12 1031 1229 0.04
13 1296 439 0.88
14 1152 724 0.44
15 1242 702 0.48

16 1291 617 0.60


17 1404 454 0.84
18 1409 608 0.64
19 1311 902 ~. 20
20 1230 516 0.76

21 1219 1001 0.16


22 1518 547 0.72
23 1225 801 0.28
24 1426 411 0.92

Mean 1282 718 -


Standard 125 238
deviation
139

From such a graph the risk of the storage estimate being too low is clearly
shown. For example, using historical data a storage of 760 x 10 6m3 is
required, but there is a 20% chance that a storage greater than 930 x 10 6m3
is necessary to meet the design requirements, that is 75% draft,
5% probability of failure.

Thus the use of generated data gives the designer a clearer picture
of the variability of his estimate and allows him to choose a value
dependent on the risk acceptable to him.

Stochastic data generation processes cannot be used to provide


estimates of historical flows. Data estimation or extension should be
based on deterministic methods. In addition, if the objectives of the
study can be achieved using analytical procedures with historical flows,
there is no need for analysis to be based on generated data. However,
where the problem cannot be solved analytically, simulation based on gene-
rated data is appropriate. Beard (1972) considers the problem in the
following manner.

"The design of water resource projects is commonly based on


assumed recurrence of past hydrologic events. By generating
a number of hydrologic sequences, each of a specified desired
length, it is possible to create a much broader base for
hydrologic design. While i t is not possible to create information
that is not already in the record, it is possible to use the
information more systematically and more effectively."

In storage-yield analysis, for cases in which a data generation


approach is adopted, simulation of the system is carried out either by
behaviour type analysis or sequent peak algorithm for each generated
sequence or replicate. The behaviour approach is preferred to the sequent
peak algorithm which is computationally more efficient but cannot be used
for problems in which draft is related to water in storage and where evapo-
ration losses need to be directly taken into account.

Twenty-five replicates are sufficient to indicate the variability in


the objective being examined. In this context, our studies suggest that
if it is necessary to negate the initial starting conditions of the
storages, for highly variable streams and high drafts at least 100 years of
records should be used in each replicate. On the other hand, if starting
conditions are known the length of each replicate should be related to the
objective of the analysis.
140

5.10 GENERALIZED RESERVOIR CAPACITY-YIELD-RELIABILITY RELATIONS

Gould (1964), Guglij (c. 1959) * and Svanidze (1964) provide curves
relating carryover storage to draft, flow variability, flow distribution
shape, persistence and probability of failure. To establish these
relationships, the authors stochastically generated annual flow data and
analysed these for storage need.

5.10.1 Gould's Synthetic Data Procedure

Using a three parameter log-normal model, Gould (1964) generated


240 sets of data, each equivalent to 10,000 years of independent flow events.
The 240 sets were analysed for different combinations of eight draft rates,
five storage sizes and six values of skewness. The results are presented in
Figs. 5.5, 5.6 and 5.7 where

mean flow - draft


and (5.27)
standard deviation

reservoir capacity
standard deviation' (5.28)

Lines for three probability of failure values are also shown on the
figures. Gould also provides a correction factor (Eq. 5.29 below) for
assuming annual serial correlation is zero. This equation was based on
analysis of additional synthesized data.

Procedure:

The procedure for estimating the required carryover storage for given
values of constant draft and probability is as follows:

(i) Compute the mean, standard deviation, coefficient of


skewness and serial correlation of annual flows.

(ii) To allow for annual serial correlation, modify the


design probability by the fOllowing equation:

P (P' - l2r)/(1.7r + 1) (5.29)

where P equivalent probability of failure after


adjusting for serial correlation,
P' design probability of failure, and
r annual serial correlation coefficient.

* A specific reference to Guglij's analysis is unavailable. Curves developed


by him were provided by Dr. S. Selvalingam, Asian Institute of Technology,
Bangkok.
141

O+-~~-'------.------r------r------
5

-2~--------------------------------~

FIG. 5.5 Storage and draft ratios based on Gould's synthetic


data for coefficient of skewness equalling zero (Gould, 1964).

10.----------------------,~------r_--_,

O+---~~------._----_r----_,------~
2 3 4 5

k1 +O.15
-2

-4~---------------------------------

FIG. 5.6 Storage and draft ratios based on Gould's synthetic


data for coefficient of skewness equalling one (Gould, 1964).
142

10~--------------------------~------~----

-2

-41-______________________________________ ~

FIG. 5.7 Storage and draft ratios based on Gould's synthetic


data for coefficient of skewness equalling two (Gould, 1964).

(iii) Compute kl from Eq. 5.27.

(iv) According to the coefficient of skewness enter Fig. 5.5,


5.6 or 5.7 and estimate k2 for the appropriate probability
of failure. (Interpolate between lines or between figures
if necessary.)

(v) Compute carryover reservoir storage from Eq. 5.28.

Comments:

The procedure utilizes the results of an analysis of synthesized


flows from a data generation model using a three parameter log-normal
distribution. It was noted in Sec. 5.8 that this distribution does not
model the time-series parameters nor the low flow values particularly well.
From Fig. 5.7 it will be seen that for kl > 0.85, which can occur at low
drafts, k2 becomes negative. This unrealistic result can occur for other
skewnesses as well. As will be noted in Sec. 6.3, the procedure tends to
underestimate capacity for the smaller reservoirs.
143

EXA~lPLE 5.4

Use Gould's Synthetic Data procedure to estimate the storage required


to meet 75% draft with a 5% probability of failure for the Mitta Mitta
River (Appendix E).
* * * * *
Following the steps given in Sec. 5.10.1:

(i) From Table E3 for annual flows:

!vIe an 1274 (x 10 6 m3 )
Standard Deviation 731 (x 10 6 m3 )
Coeff. of Skewness 1.5
Serial correlation 0.06

(ii) Modify probability:

From Eq. 5.29: P (P' - 12r)/(1.7r + 1)


(5 - 12 x 0.06)/(1.7 x 0.06 + 1)
4.28/1.1
3.9%
mean flow - draft
(iii) From Eq. 5.27: kl
standard deviation
1274 - 0.75 x 1274
731
0.44

(i v) From Fig. 5.6: k2 (P 3.9%, skew 1. 0) 2.0

From Fig. 5.7: k2 (P 3.9%, skew 2.0) 1.3


2.0 + 1.3
Interpol ating: k2 (P 3.9%, skew 1. 5)
2
1. 65

(v) From Eq. 5.28: Reservoir capacity standard deviation x k2

731 x 10 6 x 1.65
144

5.10.2 Guglij's and Svanidze's Synthetic Data Procedures

Guglij and Svanidze used Monte Carlo models and a modification of the
Pearson Type III distribution known as the Kritskii-Menkel curve
(Kartvelishvili, 1969, p. 35) to generate long sequences of annual flows
(1000 years or more) for various combinations of coefficients of variation,
skewness and serial correlation. These data sequences were used to deter-
mine finite storage capacities (as ratios of mean annual flow) for given
values of draft and reliability. Guglij results are summarized in 66 graphs
(provided by S. Selvalingam, personal communication, 1976). Svanidze
includes more details in his 120 graphs (Kartvelishvili, 1969, Appendix VII).
But both sets of results are of use only for streams with low variability -
less than 0.4 or 0.8 for Svanidze's curves and 1.4 for Guglij's relations.
Moreover, their usefulness for high drafts is limited because the computed
storage sizes are limited to values less than 2.8. These relationships
include annual serial correlation as a parameter which is not available in
Gould's Synthetic Data curves.

It wi 11 be noted in Sec. 6.3 that reservoi r capacity estimates from


these procedures are not as satisfactory as other estimates; consequently
other procedures are recommended before these.

5.11 NOTATION

a. number of days in month i


1
A variahle in Kirby's transformation (Eq. 5.10)
A. location parameter in three parameter log-normal distribution
J
b. regression coefficient for estimating Cj+l)th flow
J
B variable in Kirby's transformation CEq. 5.10)
B Beard's procedure in Tables 5.7, 5.8 and 5.9
B. log transformed regression coefficient
J
C coefficient of variation
v
C coefficient of skewness
s
CL confidence limits
g coefficient of skewness of standardized log flows (Sec. 5.5.3)

coe ff lClent a f s k ewness of f lows during J. th season
gj
G like-Gamma distribution in Tables 5.7, 5.8 and 5.9
G variable in Kirby's transformation CEq. 5.10)
GCO,I,y) Gamma distribution with zero mean, unite variance
and skew = y
H variable in Kirby's transformation (Eq. 5.11)
k lag between flow events under analysis
145

draft ratio defined as mean less draft, divided by


standard deviation of flow
storage ratio defined as reservoir capacity divided by
standard deviation of flow
K Kirby's procedure in Tables 5.7, 5.8 and 5.9
K correlation component of the streamflow process
t
LGLT like-Gamma Logarithmic Transformation model (Sec. 5.5.1)
LN-2 2 parameter Log-normal model
LN- 3 3 parameter Log-normal model
N Normal distribution in Tables 5.7, 5.8 and 5.9
N number of flow events
N (0,1) normal distribution with mean zero and variance equal to one
P probability of failure or probability of occurrence
P' design probability of failure in Gould's synthetic data
procedure (Sec. 5.10.1)
annual generated flow volumes
annual serial correlation coefficient
r lag one serial correlation of normalized variates
(Sec. 5.5.3)
serial correlation between (j+l)th and jth months
lag k serial correlation coefficient
lag one serial correlation coefficient
log transformed serial correlation coefficient
s standard deviation
s. . .
stan d ar d d eV1at1on orJ th month
of mont h ly flows f
J
s standard deviation of loge (x + ) flows
Y
S. . tlon of flows for J.th season
1og t rans f orme d s t an d ar d d eVla
J
St seasonal component of the streamflow process
SSO ,S90 storage capacities for 50 and 90% draft rate us.ing
the sequent peak procedure
t time
t. normal random variate
1
G(O,l,y .) like-Gamma variate
t,J
Kirby's modified like Gamma variate CEq. 5.10)
trend component of the streamflow process
standardized log flow values (Sec. 5.5.3)
W Weibull distribution in Tables 5.7, 5.8 and 5.9
x .. unadjusted monthly generated flow volumes CEq. 5.26)
1J
x. . .th perlo
f low volumes durlng 1
. d
1
x monthly flow
t
146

x! . adjusted monthly generated flmv volumes CEq. 5.26)


lJ
Xl standardized monthly flow
t
X mean flow
Xj ,X j +l .
mean mont hI y fl ow durlng J' th and CJ' +1) th season
.
f low ln t
th perla
. d
\
\+1'\ generated flow logarithms for (i+l)th and ith seasons
.th
X. 1og t rans f orme d mean flow for J season
J
Y loge ex + E) flow

Y mean of loge ex + E) flows


za' zp standardized normal variate
Yj 'Y j +l seasonal coefficients of skewness for jth and
U+l) th seasons

Yt,j coefficient of skewness of the like -Gamma variate


E small increment
E random component in streamflow process
t
v normalized flow values
1: 7 ,L)0 sum of the rank one 2 and 10 year consecutive low
flows respectively
147

CHAPTER 6

QUANTITATIVE ASSESSMENT OF CAPACITY YIELD

TECHNIQUES FOR SINGLE RESERVOIRS

Over the past several years, the authors have been involved in four
comparative studies using Australian streams in which reservoir capacity-
yield techniques have been assessed. Results of three of the studies have
been published (Joy, 1970; Joy and McMahon, 1972; ~jcMahon and Codner, 1973;
Codner and McMahon, 1973; McMahon, Codner and Joy, 1973); the fourth will
be published shortly (C.H. Teoh, personal communication, 1977).

In all of the above studies the storage estimates provided by the


behaviour analysis method (with the reservoir initially full) were taken as
the benchmark to which other methods could be compared. It was selected
for this purpose because of its current widespread use by water authorities
as a final design technique, and because it suffers from only one major
drawback - that of dependence of storage es timates on the initial condi tion
chosen.

6. I CRITICAL PERIOD AND PROBABILITY MATRIX METHODS

The first study was concerned with illustrating quantitatively the


inadequacies in some of the techniques discussed in Chapters 3 and 4 by
applying them to six Australian rivers on which major dams already existed
or were proposed. The rivers and their flow characteristics are given in
Table 6.1.

In Figs. 6.1 and 6.2 storage estimates at 50% and 90% drafts
respectively for the six rivers are compared among the critical period
techniques - minimum flow (Waitt), Alexander's method corrected for the
assumption of zero serial correlation, overlapping series mass curve
frequency method (Thompson), independent series mass curve frequency method
(Stall) and behaviour estimates for 5% probability of failure. (Note that
to obtain a suitable scale, storages have been standardised by dividing by
the standard deviation of monthly flows. Reference numbers refer to the
rivers listed in Table 6.1.)

6 .1.1 ~lass Curves and Minimum Flow (Waitt)

From Figs. 6.1 and 6.2 it is seen that the Waitt technique (excluding
the additional one year's supply) which is equivalent to a mass curve
TABLE 6.1 Rivers investigated in evaluation of critical period and probability matrix methods.

Annual Monthly
No. River Area Period of
(Australian Streamgauging (sq. km) Record Mean Coeff. Coeff. Serial Coeff. Coeff. Serial
Station Reference Number) (mm) of of Corre1. of of Corre1.
Var. Skew Var. Skew
1 Yarra 334 1892-1968 540 0.40 0.77 0.12 l.00 1.77 0.62*
(229103)

2 Murrumbidgee 13000 1927-1960 128 0.85 l. 99 0.08 l.39 3.66 0.60*


( 410008)

3 Lachlan 8290 1931-1967 110 l.09 l. 89 0.14 l. 79 3.42 0.61*


( 412010)

4 Warragamba 8750 1881-1959 122 1.11 2.67 0.30* 2.14 4.71 0.45*
(212240)

5 Namoi 5700 1924-1960 76 1. 06 l. 96 0.23 l.97 4.22 0.42*


( 419007)

6 Burdekin 114000 1915-1950 57 l.00 l.72 0.07 2.86 5.05 0.41*


(120090)

* Values significantly different from zero at the 5% level.


149

0
10
(4)
Ii
(3)

0 Waitt
6 Thompson

8 Stall

Alexander

(5)

0
6

SI
(f (6)
"
0 I

4
(2)
0
(1)
0


,,/ "
2 "

"

O~---------.----------r----------'~
o 2 4 6
Sbl (f

FIG. 6.1 Comparison of critical period storage estimates


(ordinate) with historical behaviour estimates
(abscissa) at 50% draft and 5% probability of failure.

analysis, overestimates the behaviour storage values computed using all


other critical period techniques. This is to be expected as the approach
allows only one failure whereas in the other techniques we have adopted a
5% probability of failure criterion.

6.1.2 Alexander's Method

It was suggested earlier (Sec. 3.4.2) that Alexander tends to under-


estimate the behaviour storage for short critical periods (neglect of
seasonality) and to approximate it for long critical periods. From Fig. 6.1
it is seen that the Alexander method underestimates the behaviour storage
for rivers 1, 3, 4, 5 and 6, but overestimates it for river 2, possibly
because the flows for 2 are not adequately described by a two-parameter
Gamma distribution. For long critical periods, as expressed in Table 6.2,
the behaviour storages are closely approximated except for river 4.
150

80 o
(4)
o Waitt
6. Thompson
Stall

60
Alexander

(3)

(2)

40 0
(5)

e

(1) (6)
0 0
20



o 20 40 60

FIG. 6.2 Comparison of critical period storage estimates


(ordinate) with historical behaviour estimates
(abscissa) at 90% draft and 5% probability of failure.

6.1.3 Overlapping Series Frequency Mass Curve Method (Thompson)

Earlier it was shown (Sec. 3.4.5) that the relationship of the


Thomp~on storage estimate to the behaviour estimate depends on the
clustering effect of sequences of the same duration, which leads to over-
estimation of the behaviour storage results. This is significant for long
critical periods. At short critical periods, due to the neglect of repeated
failures, it can be shown theoretically that Thompson's method tends to
underestimate the required behaviour storage.

These views are confirmed in practice by Fig. 6.1 which shows that
the Thompson procedure underestimates the behaviour storage for rivers 1,
2, 5 and 6 and overestimates it for rivers 3 and 4. From Table 6.2 it is
seen that rivers 1, 2, 5 and 6 have short critical periods whereas rivers
3 and 4 have long ones.

For 90% draft, the duration of Thompson's critical period approaches


that of Waitt for rivers 2, 3, 4, 5 and 6. In this investigation, the
151

TABLE 6.2 Duration of critical periods for storages


designed for 5% probability of failure (years).

River*
Method
1 2 3 4 5 6

50?, Draft
Mass curve and Waitt 1.5 2.0 10 8.3 5.0 6.7
~linimum flow with
0.5 2.2 3.1 3.2 3.0 2.7
probabi li ty (Alexander)
Overlapping series
1.5 1.5 8.3 5.0 2.0 1.5
(Thompson)
Independent series
0.8 2.0 3.0 3.0 0.5 1.5
(Stall)

90% Draft
Mass curve and Waitt 7.5 20 18 15 15 15
Minimum flow with
10 + greater than 50 ->
probability (Alexander)
Ove r 1 app ing series
+ greater than 12 ->
(Thompson)
Independent series
(Stall)
1.5 5.0 I 6.7
I 8.3 I 6.7 I 8.3

* As listed in Table 6.1.

maximum duration of critical periods examined for Thompson's procedure was


12 years. This was insufficient to define Thompson's storage for these
ri vers . Hotvever, for river 1 repeated fai lure compens ates the cl us tering
effect and so the Thompson method approximates the behaviour storage.

In view of these inadequacies, we recommend that this procedure


should not be used.

6.1.4 Independent Series Frequency Mass Curve Method (Stall)

It was shown earlier that Stall's method tends to overestimate the


required behaviour storage for short critical periods and to underestimate
it for long cri tical periods. This is confi rmed by the empirical results
of Figs. 6.1 and 6.2. Table 6.2 contains the durations for Stall's critical
periods.

From Fig. 6.1 it is seen that Stall's procedure overestimates the


required storage for rivers 1 and 5, and underestimates it for rivers 2, 3,
4 and 6. From Table 6.1, rivers 1 and 5 have short critical periods while
rivers 2, 3, 4 and 6 have longer critical periods.
152

At 90% draft, Stall's procedure underestimates all the behaviour


storages. From Table 6.1 we see that all rivers have long critical periods,
except for river I in which Stall and behaviour estimates are very similar.

Because of these limitations we recommend that Stall's procedure


should not be used.

6.l.5 Gould's Probability ~latrix Method

This is the only method in the second group of storage-yield tech-


niques for which there are comparative results. The limitations of Moran's
discrete probability matrix models were discussed in detail in Chapter 4.
In Figs. 6.3 and 6.4 results from Gould's probability matrix method (as
proposed by Gould and also as modified to cater for monthly failures rather
than annual failures) are compared with behaviour estimates for conditions
of 50% and 90% drafts and 5% probability of failure.

In both Figs. 6.3 and 6.4 the original Gould method overestimates the
behaviour storage. On the other hand, the modified Gould method fits the
behaviour relationship satisfactorily.

6 Gould
Modified Gould
(4)

6.
6

(6) (5)
6.
6.

4

(2)
6.

(ll

./
6.

O~---------.----------.----------'r---------~
o 2 4
Sb/
6 8
(f

FIG. 6.3 Comparison of Gould's probability matrix storage


estimates (ordinate) with historical behaviour estimates
(abscissa) at 50% draft and 5% probability of failure.
153

(4)
to.
(3)

to.

60 (5) (2)
to.
to.


40

(6)
to.
s/ ()

20 (ll

6 6. Gould
Modified Gould

O~-----------'~-----------r------------~----~
o 20 40 60

FIG. 6.4 Comparison of Gould's probability matrix storage


estimates (ordinate) with historical behaviour estimates
(abscissa) at 90% draft and 5% probability of failure.

6.1.6 Further Comparison of Gould and Behaviour Methods

A second study, which was aimed at establishing empirical equations


relating storage to other flow parameters, involved 156 Australian streams
(Sec. 4.7.1, McMahon, 1977). In the analysis for each stream, storage
sizes required to meet conditions of 5% probability of failure and 50% and
90% drafts were computed using Gould's modified approach (20 zones) and two
behaviour approaches, one assuming the reservoir to be initially full and
the other initially empty. Storage estimates expressed as ratios of the
mean annual flow are plotted in Figs. 6.S and 6.6 which represent
respectively 50% and 90% drafts.

Overall, the fit at 50% draft is satisfactory with the behaviour


empty case tending to overestimate the steady state storage requirement for
the more variable streams. In other words, the initial emptiness state is
significant for the conditions examined. At 90% draft, there is a tendency
for steady state storage need to be underestimated for the behaviour full
case as a result of the initially full assumption.
154

2.5 /

,!,V/

1.0

'i
0.5
"cc'"
'"c
.,'" 0.25
E X
--.,
0>
::::
0

0.1
.&til
E
:;:;
~ 0.05
:;
o
.;; a-Behaviour Method: Initially empty
til
6-Behaviour Method: Initially full
~ 0.025
III

0.01
/
/
/
o;y
/

0.005~/~__~~______~~~__~~____~~______~~____~~____-L________~
0.005 0.01 0.025 0.05 0.1 0.25 0.5 1.0 2.5
Gould Estimate (Storage/mean annual flowl

FIG. 6.5 Comparison of behaviour results (initially full and


empty) with modified Gould estimates for 156 Australian
streams at 50% draft and 5% probability of failure.

It is believed that in both cases the effect of ignoring annual


serial correlation in Gould's procedure is masked by the various inadequacies
in the behaviour approaches. The results confirm the modified Gould pro-
cedure as being analytically superior to the behaviour approaches.

6.1.7 Summary

From these two studies we conclude that the modified Gould technique
is a suitable analytical storage-yield procedure for final design. In
addition, it is noted that the behaviour analysis, which was used as a basis
for comparison, gave results consistent with theoretically acceptable
procedures so long as the effect of initial conditions is recognized.
155

25 I I I I I I /
/
/
/
/
/
/
/

! /

lOt-
/
/
/
! -

5.0 t-
~! ~~{~<!~ !
!
-

2.5 -
B fA!:& 1, -

d~
!

1.0 r -
g1 !

~lr
0.5 t- -

/
~
0.25 t- / -
/ 0- Behaviour Method: Initially em pty
:or 6- Behaviour Method: Initially full
/
/

O.l~____~~I__~~I~__~IL-____~~l ____~I~____L-I______~
0.1 0.25 0.5 1.0 2.5 5.0 10 25
Gould Estimate {Storage/mean annual flowl
FIG. 6.6 Comparison of behaviour results (initially full and empty)
with modified Gould estimates for 156 Australian streams
at 90% draft and 5% probability of failure.

5.2 CAPACITIES BASED ON STOCHASTIC DATA GENERATION

By generating several streamflow sequences, all with the same


statistical characteristics and all assumed to be equally likely to occur,
the designer is able to determine something about the precision of his
storage estimate. This result contrasts strongly with methods like the
)ehaviour analysis using historical data in which only one value is
)btained; nothing can be deduced about the distribution of this single
,stimate.

A study of three rivers (numbers 1,4 and 5 in Table 6.1) involved


:omparing historical behaviour and modified Gould estimates with those
f-'
Uo
1.0
4 '"
Range~ 5.1
1.14
80% bond ~ Range~

0.8
/
/ ", \
/ \ M
I
... ...
/
/ \
\ E 3 I ... ...
/
/
\
\ '"0 I
I ... ...
\ .... ....
M
\ "",,,,,,,"" .... .... X 80% band'\.
I
E .... .... I
I
'"0 ....
X
0.6
Q)
01 I

'"
.8
'-
2
<Il
01
'"
'-
0

I
(j)
, ....
"-
,- " / .....
ci) 0.4 I ,- / .... ...
,, --- "" /
" .- .-
,,
/

0.2 ,"
, /

-.- I
Historical Behaviour
Historical Gould
--- I
Historical Behaviour
- i-Historical Gould
x Generated Behaviour
X Generated Behaviour
0 Generated Gould
0 Generated Gould
o 0
50 100 200 500 20 50 100 200
Generated Period (years) Generated Period (years)

FIG. 6.7 Behaviour and modified Gould storage estimates FIG. 6.8 Behaviour and modified Gould storage esti-
based on historical and generated data for mates based on historical and generated
conditions of 50% draft and 1% probability of data for conditions of 90% draft and 1%
failure [Namoi River at Keepit No. 2 probability of failure [Namoi River at
(419007) New South Wales]. Keepit No.2 (419007) New South Wales].
TABLE 6.3 Reservoir capacltles estimated by behaviour and *
Gould techniques using historical and generated data.

River t Behaviour Gould

Historical 50 years 200 years Historical 50 years 200 years

50% draft, 1% ~_robabi1ity of failure


Yarra 0.22 0.18 0.19 0.21 0.17 0.19
(+41%, -10%) ( +4%, -6%) ( +10% , -10%) (+7% , -3%)

Warragamba 1.5 1.0 1.2 1.2 0.9 1.2


(+25%, -26%) (+22%, -9%) ( +43%, -34%) (+23%, -6%)

Namoi 0.9 1.0 1.2 1.1 l.2 1.0


(+83% , -19%) (+26% , -14%) (+28% , -28%) (+8% , -18%)
:----
90% draft, 5% probabi1i ty of failure
Yarra 0.89 0.94 0.89 0.94 0.94 0.89
(+35%, -43%) (+26% , -31%) (+47%, -23%) (+22% , -15%)

Warragamba 11 4 8 9 5 7
(+110%, -30%) (+93% , -34%) (n. a.) (n. a. , -22%)

Namoi 4.6 2.5 5.1 5.3 4.2 5.1


(+70%, -15%) (+22% , -38%) (+34% , -52%) (+33%, -23%)

* Storage estimates are expressed as ratios of mean annual flow. Synthetic estimates are the median
values of ten replicates. Values in brackets represent the variation of storage sizes within which
eight of the ten replicates lie.

tPor details see Table 6.1.


158

obtained using generated data of length 20-500 years. Results for one
river are shown in Figs. 6.7 and 6.8 and for all three rivers in Table 6.3.
A Thomas and Fiering monthly model using the like Gamma distribution and an
initial logarithmic transformation (LGLT) was adopted as the data generation
procedure. Assuming that the monthly generated flows are representative of
alternative historical sequences, a number of observations follow:

(i) It is observed that for the behaviour analysis a wide range


of storage values results from the different flow sequences.
This variation is indicative of the sampling error associated
wi th a hydrologic sequence. Sampling error of the storage
estimate is discussed in Sec. 6.4.

(ii) By increasing the record length, sampling variability is


lowered. Generally, too, the storage variations reSUlting
from the modified Gould analysis were smaller than those
found using the behaviour technique.

(iii) For the behaviour technique as typified in Figs. 6.7 and


6.8, average storage estimates increase to a maximum value
with increasing record length. This feature is a consequence
of critical period approaches in which the storage is assumed
initially full so that the failure in the sequence is dependent
on the initial storage condition, whereas additional failures
can result without the storage refilling. Thus proportionally
less failures will occur during the earlier part of the
sequence as compared with those occurring in the later part.
Results suggest that for highly variable streams and high
drafts at least 100 years of streamflows are required before
the effect of the initially full assumption can be ignored.
Because the Gould procedure is not dependent on the initial
conditions of the storage, this limitation does not occur
with it.

(iv) The results also confirm the consistency of the Gould


procedure. For each river the Gould results show less
variability between the historical and generated flow
estimates and within the generated flow estimates than do
the corresponding behaviour values. This mainly arises because
of the problem of the initial starting conditions.
159

The results shown in Table 5.9 are also pertinent to this evaluation.
There it was found that the storage estimates based on annual generated data
using Beard's technique overestimated the historical values by about 10%.
This result is not inconsistent with the results in Table 6.3 where the
estimates based on replicates of 200 years of generated data are within
10% of the historical Gould value.

6.3 RAPID RESERVOIR CAPACITY-YIELD PROCEDURES

Rapid storage-yield procedures suitable for preliminary reservoir


capacity-yield estimation were evaluated in the fourth study (C.H. Teoh,
personal communication, 1977). Rapid procedures are considered to include
any method that would allow a storage-yield estimation to be made quickly
(say within an hour) without the aid of a digital computer. Eight pro-
cedures were evaluated: three critical period approaches - Alexander's,
Dincer's and Gould's Gamma; two probabi Ii ty matrix methods - Hardison's
Probabi lity Routing and McMahon's Empiri cal; and three based on generated
data - Gould's, Guglij's and Svanidze's. In the evaluation, four combi-
nations of draft and reliability were adopted (30% draft, 95% reliability;
50/90; 70/98 and 90/95) and reservoir capacities were computed for
36 Australian and 12 Malaysian streams with annual coefficients of variation
ranging from 0.12 to 1.8 and annual coefficients of skewness ranging from
-0.8 to 3.7. Storage estimates using Gould's transition matrix procedure
and the behaviour method were used to determine datum values against which
the estimates from the other procedures could be tested. Comparisons are
shown in Figs. 6.9 to 6.15.

From these figures a number of points should be noted.

(i) (a) With regard to Alexander's procedure (Fig. 3.14),


for drafts less than 40%, recurrence intervals are
restricting; no values could be estimated for 30% draft.

(b) There is a tendency as seen in Fig. 6.9 for


Alexander's storage estimates based on the same draft
values to be generally overestimating or underestimating
what are expected to be the correct estimates.

(c) For drafts of 85% and above, Alexander's curves


(Fig. 3.14) require extrapolation for use.
160

20

,0

"
"Qj I
""C
c:
'"
~

<'"
30% draft. 95% reliability
50% draft. 90% reliability
70% draft, 98% reliability
90% <:iraft. 95% reliability
05

05
2 5 1 2 5 10 20
Gould's probability matrix estimate
(reservoir capacity / mean annual flow)

FIG. 6.9 Comparison of Alexander's estimates with


Gould's probability matrix values.

20 ++

'0 .*

'"
0; 2
.,E
'""
"
~E 5
c:
o x 30% draft. 95% reliability
o 50% draft. 90% reliability
c, 70% draft, 98% reliability
+ 90% draft. 95% reliability

05

05 2 5 1 2 5 10 20
Gould's probability matrix estimate
( reservoir capacity I mean annual flow)

FIG. 6.10 Comparison of Dincer's estimates with


Gould's probability matrix values.
161

20

+
+
+
+
it

..E j:
2 .0Y'J
.
E ;(~<&
""
'0
"S .5
.
. ;:;''1
/x
0
:t4'l
o
o XX Xx.,p>.0
x ~~..

/
<o:;~~.: x
o
30%
50%
draft,
draft.
95%
90%
reliability
reJiabiIJty
n 70% draft. 98% reliability
.0. 00'" + + 90% draft. 95% reliability
05 x 00

.)( 0 '"

05 2 5 1 2 5 10 20
Gould's probability matrix estimate
(reservoir capacity / mean annual flow)

FIG. 6.11 Comparison of Gould's Gamma estimates with


Gould's probability matrix values.

//
/

~
20
/
.~
r::+:
10

'"'"
, "
~'"
:0
o~ 2 0.0;;/.
..
ii
.0
o
1
,~,
~ 5
"'~~~

~ /~.:x~o
2
x

A
30%
50%
70%
draft,
draft,
draft.
95%
90%
98%
reliability
reliability
reliability
:x: 1 b.-<.~
+ 90% draft, 95% reIJability

05 '" 4)(

IAli
05 2 5 1 2 5 10 20
Gould's probability matrix estimate
( reservoir capacity / mean annual flow)

FIG. 6.12 Comparison of Hardison's probability routing


estimates with Gould's probability matrix values.
162

20

10
~
'"
E 5
:;:;

"'"

'"c:
o
.c
'"
::;
u
2 x 30% draft, 95% reliability
::; o 50% draft, 90% reliability
l:!. 70% draft, 98% reliability
+ 90% draft. 95% reliability

05 2 5 1 2 5 10 20
Gould's probability matrix estimate
(reservoir capacity / mean annual flow)

FIG. 6.13 Comparison of McMahon's empirical estimates


with Gould's probability matrix values.

20

! 10

'"
E
~
~
"
~ 2
u
+f'
~ 1 ,
-;; h
c:
>..
'" 5 +-t ,..
+ +
"
'"
"0
*
+
30% draft, 95% reliability
:;
o 2 50% draft, 90% reliability
Cl 70% draft. 98% reliability
+ 90% draft, 95% reliability

05

05 1 2 5 1 2 5 10 20
Gould's probability matrix estimate
(reservoir capacity / mean annual flow)

FIG. 6.14 Comparison of Gould's synthetic data estimates with


Gould's probability matrix values.
163

20

10

x 30% draft, 95% reliiibility


o 50% draft. 90% reliability
6 70% draft. 98"10 reliability
+ 90% draft, 95% reliability

05

05 1 5 1 10 20
Gould's probability matrix estimate
( reservoir capacity / mean annual flow)

FIG. 6.15 Comparison of Guglij's synthetic data


estimates with Gould's probability matrix values.

(ii) Overall, Dincer's values overestimate the Gould storage


estimates (Fig. 6.10). The magnitude of overestimate is
large and unacceptable for 30% drafts, but for storage sizes
greater than unity the overestimation is less than 10%.

(iii) For Gould's Gamma procedure, the results are satisfactory


although a slight bias is evident for some drafts (Fig. 6.11).
The increase in variability for storages less than unity is
expected, as reservoir capacities based on annual flows are
compared with results determined from monthly data.

(iv) (a) Hardison's procedure tends to slightly overestimate


storages greater than unity and severely underestimate
very small ones (Fig. 6.12).

(b) Some values are missing in the upper range because


the curves (Fig. 4.8 to 4.10) at high drafts and high
variability are restricting.

(v) As expected, results using McMahon's procedure fi t well


because 36 of the 48 streams shown in Fig. 6.13 were used
to develop the empirical relations. Nevertheless, for the
Malaysian streams which were not l:sed to develop the empirical
equations no anomalies are evident although some of the streams
have coefficients of variation outside the range of data used
to develop the relationships.
164

(vi) Gould's Synthetic Data procedure performs poorly for


smaller storages but satisfacorily for storages greater
than unity (Fig. 6.14).

(vii) Only Malaysian data were used to test Guglij's method


mainly because most of the Australian streams were outside
the range of parameters given in the original paper
(Sec. 5.10.2). Guglij was used in preference to Svanidze
because he covers a wider range of condi tions. The results
in Fig. 6.15 are very similar to the smaller storage
estimates from Gould's Synthetic Data procedure (Fig. 6.14)
where for storages less than unity the estimates are
unacceptable.

From this evaluation three procedures appear to be satisfactory.


Gould's Gamma and Mc~1ahon' s Empiri cal methods are s lightly superior to
Alexander, but it must be noted that the regression equations for McMahon's
method were based on Australian data and until it is thoroughly tested using
data from other continents, its use should be restricted to Australian
streams.

6.4 SAMPLING ERROR OF STORAGE AND DRAFT ESTIMATES

Few papers provide an estimate of the sampling error of storage size.


Some procedures are given here:

(i) Gould (1964) assumed that the draft is near-normally


distributed and used the results of his Synthetic Data
procedure (Sec. 5.10.1) in the estimation of sampling
error. Based on the results of that procedure, it is
recommended that the following error analysis be applied
to storage sizes (in terms of mean annual flow) greater
than unity. Gould showed that there is an ct% chance that
the draft which would result in a p% probability of
failure is less than D', where

D' D - t C [lIn + 0.5 (EL + FL)J" (6.1 )


a,n-l v

where D design draft (as percentage of mean annual flow),


single-tailed Student's t for a% and
n-l degrees of freedom,
165

n years of recorded flow,


C coeffi cient of variation of annual flows,
v
c(2k + b)
1 2
E [ QlS (6.2)
rn
_ ck2
(k2 + b)2 J
F 2 (6.3)
n(k + b)3

k2 reservoir capacity/ standard deviation

Variables b and c are tabulated in Table 6.4.

TABLE 6.4 Values of b and c in Eqs. 6.2 and


6.3 for different values of probability
of failure and coefficients of skewness.

Probabi Ii ty of Coeffi cient of Skewness


Parameter
Failure (p)
0 1 2 3
(%)
0.5 1.5 2.8 4 5.5
b 2 1 2.1 3.1 4
5 0.5 1.7 2.5 3

0.5 3.3 3.8 4 4.2


c 2 2.4 2.7 2.9 3
5 1.7 2 2.1 2.2

(ii) Alternatively, one could determine the magnitude of sampling


error from one of the theoretical methods - the most suitable
being Gould's Gamma approach (Sec. 3.4.4). Equation 3.30 gives

T [,(1 'P'D)- d] C/ (6.4)

where T required storage divided by mean annual


flow in volume units per year.

Differentiating with respect to Cv leads to


z 2
IH
2 [ 4 (1 - D)
P
d] C i'lC
v V
(6.5)

i'lT i'lC
2 v
(6.6)
T CV
166

where the standard error of the estimate of C '


v
denoted by ~Cv' is given by Eq. 2.13. Equation 6.6
gives the sampling error with respect to the storage
estimate.
Again using Eq. 6.4 as the basis of the storage. draft relation,
the sampling error with respect to the draft, may also be
derived by differentiating with respect to D, leading to:

D' D - (6.7)

Since (from Eq. 6.4)


2
1 - ~
1
D
4
(;- + dj
Cv
2 2T ~C
~ v
D' -
4(~2 + d)
[1 +
C3(~2 +
dJJ
(6.8)

V C
v v

which is an alternative equation to Eq. 6.1, but with


a much simpler derivation.

EXAMPLE 6.1

Previous examples have used various procedures to estimate the


storage requi rement on the Mi tta Mi tta River (Appendix E) to meet a 75%
draft with a 5% probability of failure. However, due to sampling errors,
it is possible that the realizable draft may be lower (or higher) than the
design value. Use each of the alternative Eqs. 6.1 and 6.8 to determine
the magnitude of realizable draft which has a 90% chance of being equalled
or exceeded.
* * * * *
(i) For the variables in Eq. 6.1:

o design draft 0.75


n number of years of record 34
C( 0.10
from Student- t distribution 1. 3.
167

C coefficient of variation of annual flows 0.57


v
(Appendix E)
b 2.1
} from Table 6.4 (for skewness 1.5, P 5%)
c 2.05
k2 1. 65 (from Example 5.4)

Equation 6.2 gives:

E _1_ [0.15 2.05 (2 x 1.65 + 2.1) ]


!34 (1.65 + 2.1)2
- 0.11

Equation 6.3 gives:

2.05 (1.65)2
F 3
34(1.65 + 2.1)
0.003

Substitute in Eq. 6.1 to obtain:

D' 0.75 - (1.31) (0.57) "'1/1/34 + 0.5 (0.1l2 + 0.003 2 )


0.61

That is, Eq. 6.1 predicts that, for a probability of failure of 5%,
there is a 90% chance that the realizable draft from the storage obtained
using the design draft will be 61% or more.

(ii) For the variables in Eq. 6.8 and results from Example 3.9,

z 1.64 (from Normal distribution, p = 5%)


P
T 0.68
C 0.57
v
d 0.6
Equation 2.13 gives one standard error of C
v
1 + 2C
( 2n V)
1 + 2 x 0.57, t
( 2 x 34 )

0.18

Now, one standard error is equivalent to 67% confidence(Normal


distribution). For 90% confidence multiply 0.18 by 1.28.

6e v 0.23
168

Substitute in Eq. 6.8 to get:


2

~
l.64 2(0.68) (0.23) ]
D'
0.68
4 - - + 0.6 0.57 3 0.68 + 0.6
2
0.57 0.57 2

0.59

That is, Eq. 6.S predicts that the realizable draft which one can be
90% sure of being met is 59% or more.

(iii) An alternative approach to sample error analysis is an


empirical examination of the effect of changes in parameter
values on storage size using stochastic data generation models.
Results of such an analysis are shown in Table 6.5. A monthly
log-normal Markov model was used to generate sequences of
flows for four rivers in which random errors were superimposed
on. the input his torical parameters of the model. Storage
estimates for each replicate were calculated using Gould's
procedure.

TABLE 6.5 Effect of Flow Parameters on Reservoir Size *

River Mean Coeff. of Skewness Seri al Correlation


(Annual
Coeff. of l.0 0.9 0.8 Ignore 1.0 2.0 1.0 1.1 1.2
Skewness) Skew
LN2 Model LN2 LN3 Model LN2 Model

Gordon 100 130 167 73 100 SO 100 lOS 105


(0.24) (17) (21) (19) (17) (17) (15) (17) (8) (15)
Yarra 100 127 187 76 100 76 100 106 114
(0.40) (10) (13) (20) (10) (26) (12) (10) (14) (16)
Torrens 100 147 219 66 100 86 100 110 122
(0.78) (23) (31) (34) (23) (25) (38) (23) (22) (22)
Warragamba 100 140 288 100 123 125
Not avail ab Ie
(1.11) (29) (30) (51) (29) (38) (36)

* Storage estimates are mean values of 20 replicates and are expressed


as percentages of the long term Gould values for the appropriate
parameter. Values in brackets are standard deviations expressed as
percentages of corresponding mean value.
169

In Table 6.5 it is seen that with a decrease in mean flow,


storages increase markedly to compensate for the more severe
flow conditions. A 10% decrease in mean results in a 30-40%
increase in storage requirement. A further decrease in the
mean results in a relatively larger increase in storage.
As the variability of the streamflow increases, this effect
becomes more significant (for example, the storage increase
is larger for the Warragamba river than for the Gordon river).
For the case of serial correlation the magnitude of errors
in storage is considerably less than for errors in the mean.
Increasing the coefficients of skewness in the three parameter
log-normal model results in an increase in the means and
hence lower storage requirements. This effect is even more
marked for the two parameter model where the implied skewness
(Eq. 5.20) is large, resulting in less storage than required
by the three parameter model.

6.5 RECOMMENDATIONS

Based on the theoretical considerations presented in earlier chapters


and on the assessment of the methods when applied to a wide range of rivers,
certain recommendations can be made. These are:

Within-year Storage (Critical period less than say 6 months):

If low flow frequency curves are available, use the within-year


frequency-massed curve procedure (Sec. 3.4.6) and increase the estimated
storage capacity by 10%. If frequency curves are not available, use a
historical behaviour analysis.

Preliminary Design of Large Storage (Critical period say


equal to or greater than 6 months):

Use Gould's Gamma procedure (Sec. 3.4.4) or McMahon's Empirical


equations (Sec. 4.7.1) if Australian streams are being considered. Adjust
the storage estimate to account for annual serial correlation (Appendix B).
Alexander's (Sec. 3.4.2) and Dincer's (Sec. 3.4.3) procedures can be used
to estimate the length of the critical period.

Final Design of Large Storage:

Use Gould's Probability Matrix procedure modified as recommended to


define monthly failures (Sec. 4.6) and adjust the storage estimates to
account for annual serial correlation (Appendix B).
170

Alternatively, a behaviour (simulation) analysis using stochastic


data could be adopted (Sec. 5.9).

In either situation as a check, an historical behaviour analysis with


empty and full initial conditions should be carried out (Sec. 3.2.3).

6.6 NOTATION

b variable in Eqs. 6.2 and 6.3


c variable in Eqs. 6.2 and 6.3
C coefficient of variation
v
d difference hetween the lower p percentile flow of Gamma
distribution G(c) and a Normal distribution N(c,c)
D draft as ratio of mean flow
0' minimum draft for a% chance based on sampling error analysis
(Sec. 6.4)
E variable in Gould's sampling error analysis (Eq. 6.1)
F variable in Gould's sampling error analysis CEq. 6.1)
k2 storage ratio defined as reservoir capacity divided by
standard deviation of flow
n number of items of data
p percentage chance of occurrence
S storage estimates
Sb storage estimate based on behaviour
single tailed student's t for a% and n-l degrees of freedom
standardized normal variate
percentage change of occurrence
lIC incremental change in C
V v
LIT incremental change in T
a standard deviation of monthly flows (Figs. 6.1 to 6.4)
T reservoir capacity divided by mean annual flow
171

CHAPTER 7

MULTI - RESERVOIR SYSTEMS

So far this text has been concerned wi th the storage capaci ty-draft-
probability relationship for a single reservoir. Many water supply systems
however consist of more than one reservoir and for these cases the single
reservoir technique is only a very approximate guide to the capacity-yield-
probability relationship of either an independent reservoir or the whole
system. In most cases of multi-reservoir systems, the extent of the
reservoir interconnection either among the storages themselves or at the
point of demand determines whether a single or a multi-reservoir analysis
is most appropriate.

7.1 A TYPICAL PROBLEM

The general problem can best be illustrated by reference to a real


situation. Consider, for example, the Canberra water supply system in
Australia.

Canberra, wi th a population of 161 000 in 1971 and with an annual


growth rate in population of 9.5 percent, was served by three headwater
storages located in series along the nearby Cotter River. The locations of
the three reservoirs, the major supply lines and stream gauging stations in
relation to the city are shown in Fig. 7.1. Water from the Cotter reservoir,
the most downstream storage, can only be supplied to the city by pumping.
Releases from Corin reservoir supplement the Bendora reservoir, water from

CANBERRA

Gl
Cotter Dam , 9
4.7 Mm 3 \ \ "
480 km 2 G4<~;~
Proposed "I ~
Googong Dam\ ';
118 Mm 3 ~
Bendora Dam
'!.
10.7 Mm 3 Stream gauging
290 km 2 Station

Carin Dam
75Mm 3
200 km 2

FIG. 7.1 Canberra water supply system.


172

which feeds by gravity through a conduit to the city service basins.


Investigation reveals that annual per capita consumption is constant, but
seasonal variations from 370 t/capita day during winter to about 1150 t/
capita day during summer can be expected.

Relative to many water resources engineering studies, adequate hydro-


logic data are available. At gauging station Gl, there are 57 years of
continuous monthly data; seven and eight years of data for concurrent
periods are available at sites G2 and G3 respectively. Monthly rainfall and
pan evaporation data for the urban area are also available for the 57 year
period. On the adjacent Queanbeyan river, 57 years of monthly streamflow
data are also available (site G4 in Fig. 7.1).

A preliminary hydrologic and topographic survey indicated that a


possible dam site existed at Googong (site G4 in Fig. 7.1). The designers
then require answers to questions such as the following:

(i) For the projected demand increase with time how long would
the system with the added reservoir operate at a given
level of reliability?

(ii) Alternatively to (i), what extra releases can be safely


made with the new reservoir in the system?

(iii) Given the~esent system and demands, at what time will its
reliability fall below an accepted design level? In other
words, when does the new reservoir need to be ready to add
to the system?

(iv) How large a reservoir needs to be added to the system to


ensure reliability in a given period?

Like any water resources project, an important initial step is to


define clearly the objectives and the design criteria. In this example, the
objectives are engineering ones, namely the design capacity of the reservoir
and construction commencement date. Often the design criteria will be
related to economic factors modified by social and environmental aspects,
but for our purposes we assume the design criteria are related to the
reliability of supplying the predicted demand given a specific operating
policy for the system.

7.1.1 Traditional Solution

Traditionally, this problem would have been tackled by mathematical


simulation of the whole system. The model used would have included the
following components:
173

(i) the streamflow inputs to each reservoir including the


proposed one. If historical data for particular sites
are inadequate then some estimation procedure (for
example, regression of streamflows) would be necessary;

(ii) the transfer of water between the reservoirs, and the


supply to the point of the demand. This component would
include as constraints the conduit capacities of the
transfer links and pump capacities where appropriate;

(iii) the operating rule for the system (Sec. 2.5) which
governs the proportion of the demand to be supplied;

(iv) a constant annual demand (probably taken as an extrapolated


population estimate times a per capita demand, plus an
allowance for evaporation losses) broken down into
seasonal components.

All these components would be used in the model to simulate the entire
system on a monthly basis; that is, for each reservoir the operation would
be equivalent to a behaviour analysis.

The objectives of the simulation study could have been subject to some
constraints. For example, for a given maximum capacity of the proposed
reservoir determined by the site, the system could be modelled to determine
the level of demand which could be supplied without failure occurring. This
in turn would lead to an estimate of the design life of the system from the
population growth predictions.

If storage capacity was not a limitation, the capacity of the proposed


reservoir would have been adjusted until the system could just meet the
design demand without failure,that is, the same criterion as used in a mass
curve analysis.

The commencement date of construction of the proposed reservoir would


be estimated from an examination of the growth trend in demand and the
potential yield of the original reservoir system, taking due cognizance of
the possibility of some severe future drought, the time taken for the
reservoir to fill, and of the estimated construction time.

Probability of failure.
The description above has given the historical approach to analysis
of a reservoir system. In recent years many water authorities have
accepted the concept of probability of failure as a more useful design
standard. The definition used for failure is not normally that of the
174

system storage being empty (because it is unrealistic), but is more likely


to correspond to Eq. 2.18 which is a volumetric reliability (ratio of water
supplied to water demanded). Another definition in use is the proportion
of time that water restrictions are necessary, weighted if desired by the
severity of restriction imposed.

To simulate the system and compute the probability of failure is


similar to the analysis described above. The only extra computational step
is the keeping of a record of the number of failures (according to the
definition adopted) during the simulation.

For a demand that is growing with time it is of particular interest


to compute the probability of failure versus time relationship. This would
indicate the point in time where extra storage is going to be needed to
maintain the required design reliability of the system. Since the simulation
or behaviour analysis requires a constant annual demand, it is necessary to
repeat the entire simulation for different demand levels and to calculate
the probability of failure corresponding to each level. Then a plot of the
expected system reliability (probability of failure) versus time can be made.

7.1.2 Recycled Historical Sequences

The traditional approach does not give the water planner or decision
maker any idea of the level of risk t of the system being unable to meet the
design demand during a future drought. Nor do they get any idea of the
likelihood of the new reservoir being required before the proposed
commissioning date.

One approach that has been used in recent years to answer these
questions is to carry out N simulation analyses as described above using
the N year historical sequence concatenated with itself and beginning each
separate analysis in a new year of the historical record and continuing for
N successive years. In this way the N analyses based on N years of
historical record provide N estimates of storage size. In addition, if
demand is superimposed in the form of an annual trend rather than being
considered constant at an extrapolated value, N estimates of the time

tIt is important that the reader not confuse risk with probability of
failure. Risk is related to the error in the estimate of the parameter
concerned, be it storage capacity for a given draft, or draft for a given
storage capacity. Probability of failure can be defined in several ways,
but in all of them it is a measure of the proportion of system failures
which can be expected over an extended period.
175

period from year one to the commissioning date of the proposed dam could be
made. The analysis is completed by ranking the N values of storage
capacity or time to commissioning, plotting them on probability paper and
reading off the value of storage or time to commissioning equivalent to
some probability of occurrence. A major limitation with this type of
approach is that the storage estimates and time values are not independent
and the probability associated with any estimate could be in (serious) error.
This difficulty is similar to that observed in the overlapping sequence
approach as discussed in Sec. 3.4.5.1.

Where annual flows are independent, a valid variation of this pro-


cedure is to choose at random yearly flows (but at the same time maintaining
for each reservoir the same flows in the one year) from the N years of
data to make up synthetic flow sequences \'ihich are then used as described
above. Such a procedure eliminates dependence but suffers from other limi-
tations; for example, the synthesized flows are restricted to those
observed in the historical record, and the inflows to each reservoir (and
the climatic variable upon which a variable demand would be based) are
totally dependent.

7.2 STOCHASTI CALL Y GENERATED FLOWS

The inadequacies of the above methods may be overcome by using sto-


chastically generated data for both the inflows and the climatic variables
in a manner to preserve the time-series structure of all variables and their
cross-correlations. At least four approaches are available.

7.2.1 Key Station Approach

In a key station approach (Thomas and Fiering, 1962), flows are


generated at a key station (chosen because of its record length, reliability
or relative significance of a reservoir located on it) by a single site
procedure (Chapter 5). Using this synthetic record as an independent
variable, regression equations are used to estimate flows at all other sites.
The regression algorithms include a random term which accounts for the
variability between the key and secondary stations. A basic limitation with
this approach is that it ignores the serial correlation at the secondary
station and introduces a spurious serial correlation of value r r2
x y,x
where r is the serial correlation at the key station and r is the
x y,x
cross-correlation between the key station (y) and secondary site (x)
(Fiering 1964). Relative to other limitations in the data generation
process, for example the selection of distribution types (Sec. 5.5), this
inadequacy may not necessarily be serious.
176

7.2.2 Principal Component Approach

A principal component approach (Fie:ring, 1964) allows gauging station


flows to be linearly combined to represent fictitious gauging station
variables. These variables are orthogonal and consequently the generation
of data at the fictitious sites may proceed independently using single site
methods. After generation the synthetic data are recombined through an
inverse transformation to form generated flows at the original gauging
stations. Matalas (1967) noted two related errors in this technique but
Fiering states that they are only of minor significance (Codner, 1974).

7.2.3 Regression Method

The Hydrologic Engineering Center, U.S. Army Corps of Engineers,


recommends a regression method (Beard, 1972). A separate regression
equation is used to generate streamflows for each calendar month at each
gauging station. Each equation incorporates either a lag zero or a lag one
component (but not both) between the site in question and all other sites.
In addition, a random term dependent on the amount of variance not accounted
for in the deterministic component is included. Thus, the generated data
for some stations are mainly dependent on current flows, while those for
other stations are primarily related to the previous month's flow. Other
aspects of the procedure are identical to Beard's single site method
discussed in Sec. 5.5.3.

7.2.4 Residual Approach

In his residual approach, Matalas (1967) adopted the following


algorithm to generate flows at m multi-sites:

(7.1)

(m x 1) matrices whose elements are


standardized flow residuals for (i+l)th
and ith consecutive periods,
(m x 1) matrix whose elements are random
normal deviates, and
[AJ, [BJ (m x m) matrices whose elements consist
of a combination of the lag zero and
lag one cross-correlations between sites.

Each flow value is related to the flow at all other sites in both the
present and preceding time intervals. The elements of [AJ and [BJ must be
177

defined in such a way that the generated multi-variate sequences resemble


the historical multi-variate sequences in terms of mean, variance, skew,
serial correlation and lag zero and lag one cross-correlations for all
sites. The derivation of [A] and [B] is given by Matalas. The solution of
[A) is straightforward but [B] poses difficulties sometimes (Fiering, 1964;
Young and Pisano, 1968; Crosby and Maddock, 1970) .

To apply the Matalas residual method to a seasonal (monthly) problem,


two methods are available. Young and Pisano (1968) first transformed all
flows to a Normal distribution by either taking logarithms or square roots,
and removing the seasonal effect by standardizing the transformed flows
with respect to the monthly means and standard deviations. The data are
then used to calculate [A) and [B). Next, flows are generated by Eq. 7.1
and finally, inverse transformations applied. Theoretically, the method
preserves only the transformed flows but this is not serious.

An alternative, theoretically more correct, approach utilizes Matalas


moment transformation equations. The steps in this procedure, which is
more complex than the method of Young and Pisano, are given in McMahon,
Codner and Philips (1973).

7.2.5 Multi-site Model Performance

~lul ti-si te model performances as presented in the literature


(Beard, 1965; Codner, 1974; Doran et al., 1973; Fiering, 1964; Thomas
and Fiering, 1962; McMahon, 1971; McMahon et al., 1973; Young and
Pisano, 1968) suggest that models generate flows which are of the same order
of variability as that found for single-site models. Generally too, as for
the single site, the theoretically correct models preserve the parameters
slightly better than the less rigorous procedures. However, no detailed
studies have been made comparing and evaluating the various approaches.

7.2.6 Use of Generated Data

By replacing the climatic and streamflow variables in the


Recycled Historical Sequence approach (outlined in Sec. 7.1.2) with sto-
chastic data generated by one of the procedures described above, one has
available a powerful technique for simulating a water resources system.

Wi th each set of data the required information is found (for example,


required capacity of added reservoir, time before new reservoir is required)
by simulating the multi-reservoir system, accounting for all flow constraints
between reservoirs and for seasonal variations in demand.
l7R

At the end of this process there are thus twenty to thirty estimates
of the variable in question, enough for a designer to plot a distribution of
them and to calculate the confidence limits which surround the chosen value
(be it the mean value, or any other). The answer may be expressed in the
following way: for example, a new reservoir must be added to the existing
system within 2 years 3 months to maintain the design reliability of
supply. Contrast this answer wi th a behaviour analysis using historical
data which can only give a single estimate of each variable.

It is also worth noting that the problem of initial conditions


inherent in the behaviour analysis technique is less important in multi-
reservoir systems. This is because most applications involve the addition
of a reservoir to an existing system and the actual initial conditions of
the system can be specified.

7.2.7 Application to Multi-storage Systems

Consider the problem of deciding the construction commencement date


of a headwater storage as outlined at the beginning of this chapter. In
effect, a multi-site monthly data generation model is required; four stream-
gauging sites (for which flow volume needs to he adjusted to he equivalent to
reservoir inflows) and one climatic variable (monthly rainfall), the latter
being considered as a fifth gauging station for the purposes of the data
generation procedure. In the original study upon which this discussion is
based (HcMahon, 1971), the key station concept was adopted because at that
time the data at two sites were considered to be too short to establish
stable seasonal parameters.

The system outlined in Sec. 7.1 was simulated using a behaviour


approach for each reservoir; monthly inflows and design drafts were gene-
rated, and apportionment of flows between reservoirs and drafts to the city
were all based on specific operating rules. Evaporation losses in the
model were taken into account by an equivalent population release. Given
initial water contents of the three original reservoirs, and some criterion
for introduction of another headwater storage (in the original study three
criteria were used: no restriction policy, reliability approach and a
minimum cost function) the time from year zero to when the proposed darn was
required could be estimated from the simulation. In the original analysis
10 and 18 replicates were adopted, but 25 is considered to be a more
desirable number. The results of one analysis are shown in Fig. 7.2. This
brief illustration emphasizes the usefulness of simulation analysis using
stochastically generated data in providing useful design information not
possible by traditional procedures.
179

1.0

"~
'5
0- 08
v
'-
C>~
,S ca
v
V '-
..c 0.6
E .-
!/)

."'=
"'"
'0
v
5 0.4
>. Gi )(

~.D
:0
'"
..c 0.2
~
0..

2 4 6 8 10
Time to when Googong Dam is ready for use
(years)

FIG. 7.2 Simulation result using stochastically


generated data. (Curve gives designer a
picture of the risks associated with any
given completion date).

7.3 TRANSITION MATRIX APPROACH

An alternative concept to the multi-reservoir system approaches noted


above is based on Gould's probability matrix method (Sec. 4.6). For a
single reservoir system, it is possible to construct the transition matrix
of stored contents by considering mass continuity on a monthly basis. From
the transition matrix the time variant and steady state conditions of the
reservoir can be determined.

This concept can be extended to a multi-reservoir system in which the


transition matrix is developed for the system as a whole by setting up each
element of the transition matrix conditional on the state of all the other
reservoirs in the system. This approach is still being investigated
(Fletcher, personal communication, 1977) but it appears to offer a viable
alternative for multi-reservoir analysis to procedures involving data
generation.
180

7.4 OTHER ALTERNATIVES

Mathematical programming and primarily dynamic programming are


alternatives to simulation, but for complex systems it is necessary to
introduce simplifying assumptions often negating the validity of the
results for final design, but give useful preliminary estimates. Readers
interested in this field are referred to Hall and Dracup (1970), Hall and
Howell (1970), Asfur and Yeh (1971) and Manoel and Schonfeldt (1977).

7.5 NOTATION

[AJ (m x m) matrix whose elements consist of a combination of


lag-zero and lag-one cross correlations between m sites
[BJ (m x m) matrix whose elements consist of a combination of
lag-zero and lag-one cross correlations between m sites
(m x 1) matrices whose elements are standardized flow
reS1. duals f or (.1+1 ) th an d'1 th consecut1ve
. .
per10ds
(m x 1) matrix whose elements are random normal deviates
181

REFERENCES

Alexander, G.N., 1962. The use of the Gamma distribution in estimating


regulated output from storages. Civil Engineering Transactions, The
Institution of Engineers, Australia, CE4(1) : 29-34.

Alexander, G.N., 1964. Effect of variability of streamflow on the optimum


storage capacity. In Water Resources Use and Management. Melbourne
University Press, Melbourne.

Anderson, R.L., 1941. Distribution of the serial correlation coefficient.


Annals of Mathematical Statistics, 8 : 1-13.

Asfur, H. and Yeh, W. W-G., 1971. An annotated bibliography on the design


of water resources systems. University of California Water Resources
Center, No. 134.

Australia, Department of National Resources, 1976. Review of Australia's


Water Resources, 1975. Australian Government Printing Service,
Canberra.

Barnes, F.B., 1954. Storage required for a city water supply. Journal of
the Institution of Engineers, Australia, 26 : 198.

Beard, L.R., 1965. Use of interrelated records to simulate streamflow.


Journal of the Hydraulics Division, ASCE, 91 (HY5) : 13-22.
Beard, L.R., 1962. Statistical methods in hydrology. U.S. Army Engineer
District, Corps of Engineers, Sacramento California.

Beard, L.R., 1972. Hydrologic data management. In Hydrologic Engineering


Methods for Water Resources Development, Vol. 2, Corps of Engineers,
U.S. Army.

Benson, M.A., and Matalas, N.C., 1967. Synthetic hydrology based on


regional statistical parameters. Water Resources Research, 3 : 931-935.

Bibra, E.E. and Riggs, H.C.W., 1971. Victorian River Gaugings to 1969.
State Rivers and Water Supply Commission.

Brittan, M.R., 1961. Probability analysis to the development of a synthetic


hydrology for the Colorado River. In Part IV of Past and Probable
Future Variations in Streamflow in the Upper Colorado River.
University of Colorado.
182

Brown, J.A.E., 1961. Streamflow correlation in the Snowy Mountains area.


,Journal of the Institution of Engineers, Australia, 33 (3) : 85-95.
Burges, S.J., 1972. So~e problems with log-normal Markov runoff models.
Journal of the HydI'aulics Division, ASCE, 98 (HY9): 1487-1496.

Burges, S.3. and Linsley, R.K., 1971. Some factors influencing required
reservoir storage. Journal of the HydI'aulics Division, ASCE,
97 (7) : 977-991.

Chow, V.T. (ed.), 1964. Handbook of Applied HydI'ology. McGraw Hill,


New York.

Codner, G.P., 1974. Simulation of Melbourne water supply scheme on a


digital computer. Thesis (Ph.D.), Monash University.

codner, G.P. and McMahon, T.A., 1973. Log-normal streamflow generation


models re-examined. Journal of the HydI'aulics Division, ASCE.
99 (HY9): 1421-1431.

Crawford, N.H. and Linsley, R.K., 1966. Digital simulation in hydrology:


Stanford Watershed Model IV. Department of Civil Engineering,
Stanford University, Technical Report No. 39.
Crosby, D.S. and Maddock, T., 1970. Estimating coefficients of a flow
generator for monotone samples of data. Water Resources Research,
6 (4) : 1079-1086.

Dear10ve, R.E. and Harris, R.A., 1965. Probability of emptiness III. In


Proc. Reservoir Yield Symposium. (Water Research Association).

Doran, D.G., 1975. An efficient transition definition for discrete state


reservoir analysis: The divided interval technique. Water Resources
Research, 11 (6) : 867-873.

Doran, D.G., Lindner, M.A., and Wright, G.L., 1973. Spatially and serially
correlated streamflow synthesis. Hydrology Symposium 19?3, The
Institution of Engineers, Australia.
Dorfman, R., 1965. Formal models in the design of water resource systems.
Water Resources Research, 1 (3) : 329-336.
Fathy, A. and Shukry, A.S., 1956. The problem of reservoir capacity for
long-term storage, Journal of the Hydraulics Division, ASCE.
82 (HY5) , Paper 1082.

Feller, W., 1951. The asymptotic distribution of the range of sums of


independent random variables. Annals of Mathematical Statistics,
22 : 427-432.
183

Fiering, M.B., 1961. Queueing theory and simulation in reservoir design.


Journal of the Hydraulics Division, ASCE. 87 (HY6): 39-69.
Fiering, M.B., 1964. Multivariate technique for synthetic hydrology.
Journal of the Hydraulics Division, ASCE, 90 (HY5): 43-60.

Fiering, M.B., 1967. Streamflow Synthesis, Macmillan, London.

Fiering, M.B., and Jackson, B.B., 1971. Synthetic Streamflows, American


Geophysical Union, Water Resources Monograph 1.

Gani, J., 1955. Some problems in the theory of provisioning and of dams.
Biometrica, 42 : 179.

Gani, J. and Prabhu, N.U., 1957. Stationary distributions of the negative


exponential type for the infinite dam. Journal of the Royal
Statistical Society, Series B, 19 342.

Gani, J. and Prabhu, N.U., 1958. Remarks on a dam with Poisson type inputs.
Australian JournaZ of AppZied Science, 10 : 113.
Gani, J. and Prabhu, N.U., 1959. The time dependent solutions for a
storage problem with Poisson inputs. Journal of Mathematics and
Mechanics, 8 : 653.
Gani, J. and Pyke, R., 1960. The content of a dam as the supremum of an
infinitely divisible process. Journal of Mathematics and Mechanics,
9 : 639.

Gani, J. and Pyke, R., 1962. Inequalities for first emptiness


probabilities of a dam with ordered inputs. Journal of the Royal
Statistical Society, Series B, 24 : 102.
Ghosal, A., 1959. On the continuous analogue of Holdaway's problem for
the finite dam. Australian Journal of Applied Science, 10 : 365.

Ghosal, A., 1960. Emptiness in the finite dam. Annals of Mathematical


Statistics, 31 : 803.

Ghosal, A., 1962. Finite dam with negative binomial inputs. Australian
Journal of Applied Science, 13 : 71.
Gould, B.W., 1961. Statistical methods for estimating the design capacity
of dams. Journal of the Institution of Engineers, Australia,
33 (12) : 405-416.

Gould, B.W., 1964. Statistical methods for reservoir yield estimation.


Water Research Foundation of Australia, Report No.8.
184

Gould, B.W., 1964. Discussion of paper Alexander, 1964. In Water


Resources Use and Management, Melbourne University Press, Melbourne.
Gumbel, E.J., 1954. Statistical theory of droughts. Proceedings ASCE,
80 : 439.

Hall, W.A. and Dracup, J.A., 1970. Water Resources Systems Engineering,
McGraw-Hill, N.Y.

Hall, W.A. and Howell, D.T., 1970. Optimal allocation of stochastic water
supply. Journal of Irrigation and Drainage Division, ASCE,
96 : 395-402.

Hardison, C.H., 1965. Storage to augment low flows. In Proceedings of


Reservoir Yield Symposium. (Water Research Association).

Hardison, C.H., 1972. Potential United States water-supply development.


Journal of the Irrigation and Drainage Division, ASCE, IR3 : 479-492.

Hardison, C.H., and Martin, R.O.R., 1963. Water-supply characteristics of


streams in the Delaware River basin and in southern New Jersey.
Geological Survey, Water Supply Paper 1669-N.

Harms, A.G., and Campbell, T.H., 1967. An extension to the Thomas-Fiering


model for the sequential generation of streamflow. Water Resources
Research, 3 : 653.

Harris, R.A., 1965. Probability of reservoir yield failure using Moran's


steady state probability method and Gould's probability routing
method. Journal of the Institution of Water Engineers, 19 : 302.
Hazen, ~., 1914. Storage to be provided in impounding reservoirs for
municipal water supply. Transactions, ASCE, 77 : 1539.

Hudson, H.E., and Roberts, W.J., 1955. 1952-1955 Illinois drought with
special reference to impounding reservoir design. Illinois State
Water Survey Bulletin, No. 43.

Hufschmidt, M.M., and Fiering, M.B., 1967. Simulation Techniques for the
Design of Water-Resource Systems. Macmillan, London.

Hurst, H.E., 1951. Long term storage capacity of reservoirs.


Transactions, ASCE, 116 : 770-799.

Hurst, H.E., 1956. Methods of using long term storage in reservoirs.


Proceedings of the Institution of Civil Engineers, Paper 6059, 5 519.
185

Hurst, H.E., Black, R.P. and Simaika, Y.M., 1965. Long Tepm Stopage.
Constable, London.

Jarvis, C.L., 1964. Application of Moran's theory of dams to the Ord


River project. JouPnal of HydI'ology, 2 : 232.

Joy, C.S., 1970. On the relationship between reservoir capacity and yield.
Thesis, (M. Eng. Sc.), Monash University.

Joy, C.S. and McMahon, T.A., 1972. Reservoir-yield estimation procedures.


Civil Engineeping Tpansactions, The Institution of EngineeI's,
Australia, CE14 (1) : 28-36.

Julian, P.R., 1961. A study of statistical predictability of stream runoff


in the Upper Colorado River basin. In Part II of Past and ppobable
Futupe VaY'iations in Stpeamflow in the Uppep Colopado RiVeI'.
University of Colorado.

Kartvelishvili, N.A., 1969. Theory of stochastic processes. In HydI'ology


and RiVer Runoff Regulation, Israel Prog. for Scientific
Translations.

Kendall, M.G. and Stuart, A., 1968. The Advanced Theory of Statistics,
Vol. 3, Charles Griffin and Co.

King, C.W., 1920. Supply of water for towns in New South Wales.
Transactions, The Institution of Engineers, Australia, 1 : 262.

Kirby, W., 1972. Computer-orientated Wilson-Hilferty transformation that


preserves the first three moments and the lower-bound of the Pearson
Ty e 3 distribution. Watep Resources ReseaI'ch, 8 (5) : 1251.

Klemes, V., 1967. Reliability of water supply performed by means of a


storage reservoir within a limited period of time. JouPnal of
Hydrology, 5 : 70.

Klemes, V., 1974. The Hurst phenomenon a puzzle? Watep Resources


ReseaI'ch, 10 (4) : 675-688.

Klemes, V., 1977. Discrete representation of storage for stochastic


reservoir optimization. Mss.

Kottegoda, K.T., 1970. Statistical methods of river flow synthesis for


water resources assessment. Institution of Civil EngineeI's,
Supplement 18, Paper 7339S.

Langbein, W.E., 1958. Queueing theory and water storage. ppoceedings,


JouPnal of the Hydraulics Division, ASCE, 84 : Paper 1811.
186

Law, F., 1953. The estimation of the reliable yield of a catchment by


correlation of rainfall and runoff. Journal of the Institution of
Water Engineers, 7 (3) : 273.

Law, F., 1955. Estimation of the yield of reservoired catchments.


Journal of the Institution of Water Engineers, 9 467.
Lins ley, R. K. and Franzini, J. B., 1972. Water Resources
McGraw-Hi 11.

Lloyd, E.H., 1963. A probability theory of reservoirs with serially


correlated inputs. Journal of Hydrology, 1 (2) : 99-128.

Lloyd, E.H. and Odoom, S., 1964. Probability theory of reservoirs with
seasonal input. Journal of Hydrology, 2 : 1.

Maass, A., Hufschmidt, M.~l., Dorfman, R., Thomas, H.A., Marglin, S.A.,
and Fair, G.M., 1962. Design of Water Resources Systems.
MacMillan and Co., 620 pp.

McMahon, T.A., 1971. Simulation Analysis of a water supply system using


generated data. Proceedings, Fourth Australasian Conference on
Hydraulics and Fluid Mechanics, 188-195.

McMahon, T.A., 1975. Predictive use of catchment outputs. In


Prediction in Catchment Hydrology, (T.G. Chapman and F.X. Dunin, Eds.)
Australian Academy of Science, 371-425.

McMahon, T .A., 1976. Preliminary estimation of reservoir storage for


Australian streams. Civil Engineering Transactions, The Institution
of Engineers, Australia, CE18 (2) : 55-59.

McManon, T.A., 1977. Some statistical characteristics of annual stream-


flows of Northern Australia. Hydrology Symposium, The Institution
of Engineers, Australia, National Conference Publication No. 77/5,
131-135.

McMahon, T.A., 1977. Australia's surface water resources I - Potential


development based on hydrologic factors. Mss.

McMahon, T.A. and Codner, G.P., 1973. Inadequate hydrologic data and
reservoir capacity. In Decisions with Inadequate Hydrologic Data
(D.A. Woolhiser, Ed.) Proceedings, Second International Hydrology
symposium, Fort Collins.

McMahon, T.A. and Miller, A.J., 1971. Application of the Thomas and
Fiering model to skewed hydrologic data. Water Resources Research,
7 (5) : 1338.
187

~IcMahon, LA., Codner, G.P. and Joy, C.S., 1972. Reservoirstorage-yield


estimates based on historical and generated streamflows. Civil
Engineering Transactions, The Institution of Engineers, Australia,
CE14 (2) : 147-152.

McMahon, T.A., Codner, G.P. and Philips, C., 1972. Single and multisite
operational hydrology. Nordic Hydrology, 3 : 214-238.

McMahon, T.A., Codner, G.P. and Philips, C., 1973. A note on single and
mu1tisite operational hydrology. Nordic Hydrology, 4 : 54-55.

Manoe1, P.J. and Schonfe1dt, C.B., 1977. Economic optimization of an


expanding water supply. HydroZogy Symposium, The Institution of
Engineers, Australia, National Conference Publication No. 77/5,
88-92.

Mata1as, N.C., 1967. Mathematical assessment of synthetic hydrology.


Water Resources Research, 3 (4) : 937.
Matalas, N.C. and Benson, M.A., 1968. Note on the standard error of the
coefficient of skewness. Water Resources Research, 4 (1) : 204-205.
Melentijevich, M.J., 1966. Storage equations for linear flow regulations,
Journal of Hydrohogy, 4 : 201-223.

Moran, P.A.P., 1954. A probability theory for dams and storage systems.
Australian Journal of Applied Science, 6 : 116.

Moran, P.A.P., 1955. A probability theory of dams and storage systems:


modifications of the release rules. Australian Journal of Applied
Science, 6 : 117.

Moran, P.A.P., 1959. The Theory of Storage. Methuen, London.

Moss, M.E., 1971. Stochastic methods in urban water systems. In


Treatise on Urban Water Systems (Albertson, M.L., Tucker, L.S. and
Taylor, D.C., Eds.) Colorado State University, Fort Collins,
384-401.

Perrins, S.J. and Howell, D.T., 1971. Neglect of streamflow serial


correlation as a source of error in reservoir yield estimation.
The Institution of Engineers, Australia, Hydrology Papers 1971,
66-70.

Phatarfod, R. M. 1976. Some aspects of stochastic reservoir theory.


Journal of HydroZogy, 30 : 199-217.
188

Philips, C., 1972. Stochastic analyses and generation of streamflow data.


Thesis (M. Eng. Sc.), Monash University.

Prabhu, N.U., 1958 (a). On the integral equations for the finite dam.
Quarterly Journal of Mathematics, Series 2, 9 : 183.

Prabhu, N.U., 1958 (b). Some exact results for the finite dam. Annals
of Mathematical Statistics, 29 : 1234.

Rippl, W., 1883. Capacity of storage reservoirs for water supply.


Minutes of Proc., Institution of civil Engineers, 71 : 270-278.

Roesner, L.A. and Yevjevich, V.M., 1966. Mathematical models for time
series of monthly precipitation and monthly runoff. Hydrology
Paper 15, Colorado State University, Fort Collins, Colorado.

Savarenskiy, A. D., 1940. Metod Rascheta Regulirovaniya Stoka,


Gidrotekhnicheskoe Stroitelstvo 2 : 24-28.

Searcy, J.K., 1960. Graphical correlation of gauging - station records.


Geological Survey, Water Supply Paper l54l-C.

Stall, J.B., 1962. Reservoir mass analysis by a low-flow series.


Journal of the Sanitary Engineering Division, ASCE, 88 (SA5): 21-40.

Stall, J.B. and Neill, J.C., 1961. A partial duration series for low
flow analysis. Journal of Geophysical Research, 66 (12) : 4119-4125.

Sudler, C.H., 1927. Storage required for regulation of streamflow.


Transactions, ASCE, 61 622.

Svanidze, G. G., 1964. Osnovy rascr.eta regulirovaniya rechnogo stoka


metodom Monte-Karlo (Elements of river runoff regulation computation
by Monte Carlo method), Tbilisi, Izdatel'stro Metsniereba.

Thorn, H. S. C., 1958. A note on the gamma distribution. Monthly Weather


Review, 86 (4) : 117-122.

Thomas, H.A. and Burden, R.P., 1963. Operations Research in Water Quality
Management. Harvard Water Resources Group.

Thomas, H.A. and Fiering, M. B., 1962. Mathematical synthesis of stream-


flow sequences for the analysis of river basins by simulations.
In Design of Water Resources Systems (A.Maass et al., Eds.) Harvard
University.

Thompson, R. W. S., 1950. The application of statistical methods in the


determination of the yield of a catchment from runoff data.
Journal of the Institution of Watel' Engineers, 4 : 394-428.
189

Tintner, G., 1968. Econometrics. John Wiley &Sons.


United States Corps of Engineers, 1975. Hydrologic Engineering Methods
for Water Resources Development, Vol. 8 Reservoir Yield. Hydrologic
Engineering Center, Davis, California.

Venetis, C., 1969. A stochastic model of monthly storage. Water Resources


Research, 5 (3) : 729.

Waitt, F.W.F., 1945. Studies of droughts in the Sydney catchment areas.


Journal of the Institution of Engineers, Australia, 17 (4-5) : 90-97.

Wallis, J.R. and O'Connell, P.E.,1972. Small sample estimation of p.


Water Resources Research, e (3) : 707-712.

Wilson, H. McL., 1940. Safe yield from Mundaring reservoir. Journal


of the Institution of Eng1:neers, Australia, 12 (2) : 39.

Wilson, E.B. and Hi1ferty, M.M., 1931. Distribution of Chi-square.


Proceedings National Academy of Science, 17 : 684-688.

Wright, G.L., 1975. Mu1ti1ag Markov models for eastern Australian streams.
Hydrology Symposium 1975, The Institution of Engineers, Australia.
National Conference Publication 75/3.

Yevdjevich, V.M., 1961. Some general aspects of fluctuations of annual


runoff in the upper Colorado River basin. In Part II of Past and
Probable Future Variations in Streamflow in the Upper Colorado
River, University of Colorado.

Yevjevich, V.M., 1972 (a). Probability and Statistics in Hydrology.


Water Resources Publications, Fort Collins.

Yevjevich, V.M., 1972 (b). Stochastic Processes in Hydrology,


Water Resources Publications, Fort Collins.

Young, G.K. and Pisano, W.C., 1968. Operational hydrology using


residuals. Journal of the Hydraulics Division, ASCE, 94 (HY4)
909-923.
190

APPENDIX A

PROCEDURE TO ADJUST STORAGE ESTIMATE FOR NET EVAPORATION LOSS

Prior to dam construction, the long term rainfall-runoff relation for


the area which will be flooded by the proposed reservoir can be expressed
as follows:

RB P - ETB (Al)
B

where RB mean annual runoff before reservoir inundation,

P mean annual rainfall, and


B
ETB mean annual evapotranspiration.

After the reservoir is filled, the relation can be expressed as:

RA PA - EO A (A2)

where RA mean annual runoff after reservoir is filled,

PA mean annual rainfall, and

EO mean annual evaporation from the water surface.


A
Assuming that the mean annual rainfall before and after construction
remains approximately equal, that is, P
B
= PA, then
(A3)

(A4)

noting that nE ~ 0, where nE = mean annual net evaporation loss.


Lake or open surface water evaporation can be estimated by one of the
recognized theoretical or empirical procedures (see Chow, 1964, Section 11)
or by applying an annual pan coefficient (p) to tank evaporation data (Ep) ,
thus:

P E (AS)
P
Pre-dam evapotranspiration estimates are difficult to determine.
One apporach is through Eq. Al, thus:

(AI)
191

Estimates of mean annual rainfall (P ) are readily available. Mean annual


B
runoff (R ) can be calculated using data for the catchment or estimated
B
from regional runoff maps.

Another factor that is required is the length of the critical draw-


down period. It can be found from Fig. Al in which the critical period in
years is related to the draft and to the annual coefficient of variation.

100~----~-------+--~~~------~----~

50~----~----~4-------~-----+-

70

0.5~~~+4~----4-------~-----+----~

0.1 L-_ _ _ _L-._ _--'_ _ _-L_ _ _--L_ _....J


o 2
Coefficient of Variation

FIG. Al Critical period in years versus annual coefficient


of variation and draft (Curves are based on Eq. 3.25
and points on Alexander's method - Sec. 3.4.2)
192

This figure was derived from Eq. 3.25 and computations were carried out for
a probability of failure of 5%. As an approximate check on the relationship,
values for 50% and 70% drafts are superimposed using Alexander's procedure
(Sec. 3.4.2). The fit is satisfactory. It should be noted that estimates
of critical periods of one year or less will not be very reliable.

Thus the final adjustment factor for net reservoir evaporation loss
is given by combining mean annual net evaporation loss, mean surface
reservoir area and drawdown period as follows:

0.7 A LIE CP (A6)

where LISE amount that computed storage needs to be increased


to account for the net reservoir evaporation loss, (m 3)
A surface area of the reservoir at full supply level, (m 2)
LIE is determined from Eq. A4, (m/year)
CP is estimated from Fig. AI. (years).

The factor 0.7 relates the mean reservoir surface area exposed to eva-
poration during a critical drawdown period to the total area at full supply
level. This factor is based on an analysis of the storage capacity-surface
area curves for six Australian dam sites: Bendora, Corin and Cotter dams
on the Cotter River (N.S.W.), proposed Sites 10 and 15 on the Mitchell
River (Vic.), and the Talbot dam site on the Thomson River (Vic.).

NOTATION

A surface area of reservoi.r at full supply level


CP critical period
E tank evaporation data
P
EO mean annual evaporation from reservoir water surface
A
ETB mean annual evapotranspiration
P pan coefficient
PA' PB mean annual precipitation
RA equivalent mean annual runoff from area inundated by reservoir
RB mean annual runoff from area of proposed reservoir
fiE mean annual evaporation loss
LISE increase to account for the net reservoir evaporation loss
193

APPENDIX B

ADJUSTMENT FOR ASSUMPTION OF INDEPENDENCE OF ANNUAL FLOWS

No comprehensive study of the effect on the reservoir capacity-yield


relationship of the assumption of independence of annual flows is generally
avai lab Ie. However, several incomp lete studies (in the sense of covering
the needs of users of this text) are available and the results of these are
summarized below under two headings - the effect of neglecting serial
correlation on the required storage capacity and the effect on yield.

Increase in storage required to compensate for the neglect of serial


correlation

The results of five studies (Thomas and Burden 1963, Gould 1964,
Joy 1970, Perrins and Howell 1971 and McMahon and Codner 1973) which
examine the quantitative increase in storage required to compensate for the
independence assumption are summarized in Tables Bl, B2 and B3. Results are
for three levels of serial correlation 0.1, 0.2 and 0.3 and for a range of
models. For the studies of Thomas and Burden, Gould, and Perris and Howell
results have been generalized.

In looking over the tables, it can be seen that there are discre-
pancies betl"een studies (and wi thin studies). Al though these are probably
attributable to the various distributions, parameters, and definition of
probability (or reliability) used in the analyses, it is difficult for the
user to isolate the adjustment factor for a particular river.

A recommended alternative to Tables Bl to B3 is Fig. B.l which shows


the increase in storage relative to the increase in serial correlation.

The lines in this figure are based on the average reservoir capacity
computed for 156 Australian streams using an historical behaviour analysis
assuming the reservoir to be alternatively initially empty and initially
full. This average value was divided by the equivalent Gould estimate
modified for monthly failures and the resulting ratio, which is the required
correction factor, was plotted against the annual serial correlation.
Separate analyses were carried out for 50% and 90% drafts but only one
probability of failure condition of 5% was examined (McMahon, 1976).
194

TABLE Bl. Storage adjustment factors (percentage increase required)


to account for assumption of independence in reservoir
capacity-yield analysis for serial correlation of 0.1.

Draft (%)
Probabili ty Streamflow
of failure characteris ti cs
100 90 80 70 60 50

Thomas and Burden (1963) - theoretical simul ation study


1/25 years Normal 8 0* 0*

1/50 years
distribution
Cv
= 0.25
} 5 13* 0*

1/25 years Gamma 8 0* 0*

1/50 years
distribution
Cs = 1
} 11 17* 0*

Mc~lahon and Codner (1973) ** - simulation of four Australian rivers


5% probability 0.7 8
(monthly)
LN -2
distri-
bution
['"
Cv = 1.0
Cv = 1.9
Cv = 2.1
6
10
23

Gould (1964) - theoreti cal simulation study (results generali zed


with respect to skewness)
5% probability
(annual) LN-3
distri-
bution r<= 1
0.5

Cv = 2
I

* The results marked by an asterisk are based on storage estimates


I
20
I
I
reported only to one significant figure. In these cases the
adjustment factor could be in error by up to 25%.

**
This study was based on a seasonal monthly model and adjustments
were made to the monthly serial correlations rather than the
annual one.
195

TABLE B2. Storage adjustment factors (percentage increase required)


to account for assumption of independence in reservoir
capacity-yield analysis for serial correlation of 0.2.

Draft (%)
Probabi l i ty of Streamflow
of failure characteristics
100 90 80 70 60 50

Thomas and Burden (1963) - theoretical simulation study

1/25 years Normal 17 17* 0*

1/50 years
distribution
C
v
= 0.25 } 10 25* 0*

1/25 years Gamma 25 0* 0*


distribution }
s =
1/50 years C 1 16 33 0*

McMahon and Codner (1973)** - simul ation of four Australian rivers

r
5% probabi Ii ty 0.7 5
(monthly) v
LN -2 C = 1.0 14
. . v
dls~n- C = 1.9 22
butlon v
C = 2.1 25
v

r
Gould (1964) = theoretical simulation study (results generalized
--- with respect to skewness)

5% probability
LN-3 v ' 0.5 I
(annual)
distri- C'v = I
I 45
bution
C = 2
v I I
Joy (1970) - theoretical simulation study using
Lloyd's model (1963 )

S% probabi l i ty Normal
distribution 22
C = 0.4
v

* The results marked by an asterisk are based on storage estimates


reported only to one significant figure. In these cases the
adjustment factor could be in error by up to 25%.

**
This study was based on a seasonal monthly model and adjustments
were made to the monthly serial correlations rather than the
annual one.
196

TABLE B3. Storage adjustment factors (percentage increase required)


to account for assumption of independence in reservoir
capacity-yield analysis for serial correlation of 0.3.

Draft (%)
Streamflow
Reliability
characteristics 100 90 80 70 60 50

Pcrrins and Howell (1971) - theoretical simulation study


80% Normal 33
95% distribution 58 57
99% C = 0.5 48
v

lbe fit between the results for the five studies and the lines in
Fig. B.l is poor. This is because the individual studies are for specific
cond; tions whereas the lines arc generalized results from a range of
streams and conditions.

,.....
Draft (,)

90% 50%
...
...- _-Appendix B 1.8 0
+' 90%
* Gould 1964
(,)

A
* Thomas &
Burden 1963 1.6- c:
ell
u..
0
0 181 McMahon & '';:;
Codner 1973 (,)
Q)

Joy 1970
Perrins &
1.4-:
0
0 Howell 1971 ()

1.2
*
0.1 0.2 0.3 0.4 0.5

0.8
I
--11-+---+---;
Serial Correlation

FIG. B.l Reservoir-capacity correction factor for annual


serial correlation.
197

Decrease in yield required to compensate for the neglect in serial


correlation

Hardison (1965) carried out a simulation analysis using a two-


parameter log-normal distribution to assess the effect of neglecting serial
correlation on yield. Figure B.2 shows for a 2% chance of deficiency the
adjustment that should be subtracted from the allowable draft computed with
no serial correlation between years.

Perrins and Howell (1971) provide some data but it is limited


because the simulation runs are based on a normally distributed Markov
model with C 0.5.
v

~j
I
,,

~~
0.6
Index of Variability

o 023 0.49 0.78 1.16 1.66


Coefficient of Variation

FIG. B. 2 Effect of serial correlation on draft at 2% chance of


deficiency for log-normal distributions of annual
discharge for indicated serial correlation (Hardison, 1965).
(The parameter shown is the length of the critical period
in years. The adjustment should be subtracted from the
allowable draft computed with no serial correlation between
years. )
198

NOTATION

C coeffjcient of variation
v
C coefficient of skewness
s
LN-2 2 parameter Log-normal model
LN-3 3 parameter Log-normal model
19!

APPENDIX C

THEORETICAL JUSTIFICATION 01' A NON-SEASONAL MA.RKOV MODEL

Expectation theory is used to justify the form of the algorithm of


the Markov model.

(i) Consider a model of the form:

X+b(X.-X) (Cl)
1

where \+1' Xi generated flows,

X mean flow, and

b regression coefficient between ith and (i+l)th flows.

The expected value or mean of \+1' E(\+l)' is given by:

E[X + b(\- X)]

E(X) + b E(X ) b E (X)


i
X+bX-bX

X (C2)

Therefore, the mean is preserved by this type of generation model.

Consider now the variance.

E(X?l+ 1) - [E(X.1+ 1)]2 (C3)

E[ (X + b (X. - X)) 2] F
1

E(F + 2Xb (X. X) + b 2 (X. - X) 2) X2


1 1

E(X2 + 2bXX. - 2bX 2 + b 2X?- 2b 2 X.X + b 2X2) F


1 1 1

E (F) + 2bXE(\) 2bE (X2) + b 2 E (Xf) - 2b 2X E (Xi)

+ b 2E (F) - X2

X2 + 2bX2

(C4)

si+l
but b r -s-.- and si+l s.
1
= 5
1
200

standard deviation of (i+l)th and ith flows, and

r = correlation coefficient between ith and (i+l)th flows.

(e5)

Therefore, the model preserves variance when r 1, but not when


r f 1. Consequently, the model is unsatisfactory.

(ii) Consider now a model of the form:

x + b(X i - X) + S tiel (C6)

where t N(O,l) and


i
s standard deviation of flows.

This is Brittan's model as given in Sec. 5.3.

The expected value of Xi+l is


,
E (Xi +1) E[X + b(\- X) + s ti (1 r2) 2] (C7)
,
E (X) + b E(X ) b E eX) + s (1 - r 2)2 E(t.)
i 1

X (C8)

Therefore, mean is again preserved.

The variance is again computed as:

E(X?l+ 1) - [E(x.l+ 1)]2 (C9)

Let c

E[X X) + c t
+ b(\-
i
]2 - X2
2
E[X2+ b [X. - X)2 + c 2t 2 + 2bX(X.- X)
1 1 1
+ 2c X t. + 2bc (X.- X) t.] - x2
1 1 1

X2 + b 2E (X?) - 2b 2 X E (X.) + b 2E(X)2 + c 2 E(t?)


1 1 1
+ 2bXE[X - X) + 2cX E(t ) + 2bc E(t ) E(X )
i i i i
- 2bc E(t ) E(X) - X2
i
b 2E (X?) - b 2X2 + s2 [1
1

b 2 [E(x2) - X2] + s2[l (ClO)


1

[Recall that s 2 [[(X2) - X2) and that b


1
= r] (ell)
r 2s2 + s2 _ s2r2
52 (C12)

Therefore, the variance is preserved.


201

Now check that dependence is preserved as lag one serial


correlation (r). By definition,

COY (Xi+l \)
r (C13)
(VAR X.1+ 1) ~ (VAR Xi)
~
2

COY (X i + Xi)
l
2 (C14)
s
where COY (X i +l Xi) = covariance of Xi and Xi+l

s2r COY (X i +l Xi) (CIS)

E[(\+l- X) (X.- X) ] (C16)


1
E(X i +l Xi) - X E(X i +l ) X E (Xi) + X2
E[X (X + b(X.- X) + ct.)] - x2
i 1 1
E(X X. + bX.1 - bX X. + c X t i ) - x2
1 1
x2 + b E(X?) - b X2 - x2
1
b [E eX?) - x2]
1
b s2 (C17)
r b (C18)

Therefore, dependence is preserved.

Thus the model preserves the three parameters: mean, standard


deviation and serial correlation and is therefore in theory a satisfactory
generation model.

NOTATION

b regression coeffici~nt
1

c s(l - r2)2
COV(\+l'\) covariance between X th and x.th flows
i+l 1
E (x) expected value of x
N(O,l) normal distribution with mean zero and variance equal to one
r serial correlation coefficient
S. standard deviation of flows in ith time period
1
t. normal random variate
1
VAR(X ) variance of flow sequence X.
i 1
X mean flow
generated flows
202

APPENDIX D

NEWTON-RAPHSON METHOD FOR SOLVING FOR AN INEXPLICIT VARIABLE

Given f(x) = 0, to solve for x.

Assume first trial solution is x ..


1

Application of the Newton-Raphson method gives an estimate of a


better value of xi' Call it x i +l '

f(xl

f(Xjl I------~

FIG. D.I Illustration of Newton-Raphson Solution.

d ~;x) = slope, of tangent at x X. (DI)


1

(Fig. D.1) (D2)

(D3)

f(x i )
Thus xi+l = Xi - ~ (D4)
1

If xi+l - Xi exceeds the allowable error, then use xi+1 to find


another estimate xi+2'
Equation 5.16 can be treated as a function of S thus:
f(S2) = exp (3S 2) - 3 exp (S2) + 2 - g[exp (S2)_ 1]3/2 (DS)

and
20:

APPENDIX E

TABLE E1. Monthly flows for ~1itta Mitta River at Ta11andoon


(Station 4012(4), Victoria, Australia, for period
1936-1939 inclusive. Units are 10 6m3 .

Year Jan. Feb. Mar. Apr. May June July Aug. Sep. Oct. Nov. Dec. Total

1936 56 32 32 38 31 113 189 529 217 152 80 84 1553


1937 53 27 26 20 27 28 32 54 171 125 56 31 650
1938 16 16 15 20 26 44 47 58 91 52 19 9 413
1939 6 44 179 130 94 183 179 395 318 363 276 99 2266

1940 43 19 14 33 44 44 42 60 93 58 31 28 509
1941 88 22 46 27 20 32 101 63 100 136 52 23 710
1942 14 12 12 12 112 149 347 215 316 232 149 64 1634
1943 37 20 15 76 51 52 110 139 201 241 113 52 1107
1944 22 12 12 17 64 39 64 43 39 46 26 17 401

1945 14 16 7 14 16 56 42 154 146 101 89 30 685


1946 14 44 69 47 44 91 444 302 164 162 109 58 1548
1947 30 22 35 30 36 80 253 237 276 300 185 94 1578
1948 57 33 22 23 67 69 59 81 126 158 252 65 1012
1949 39 21 33 28 32 52 95 117 174 236 238 86 1151

1950 35 49 67 130 44 49 79 113 164 220 167 73 1190


1951 41 25 19 36 100 159 297 321 250 253 126 63 1690
1952 28 16 22 49 120 534 312 207 472 260 349 241 2610
1953 84 48 28 26 44 58 158 253 297 338 195 84 1613
1954 59 65 28 30 46 64 68 149 122 80 249 153 1113

1955 53 56 53 31 48 121 180 638 417 449 241 123 2410


1956 139 64 88 481 414 548 513 456 402 382 231 116 3834
1957 54 36 38 32 42 65 117 69 69 132 60 43 757
1958 43 22 17 21 89 105 191 471 165 426 154 72 1776
1959 32 26 36 44 23 39 42 96 245 211 96 46 936

1960 23 15 11 20 148 112 217 279 223 218 132 75 1473


1961 37 17 22 33 32 41 78 111 139 95 58 54 717
1962 32 19 12 16 41 139 86 144 127 169 90 53 928
1963 44 28 16 15 49 46 60 141 163 137 105 46 850
1964 17 14 12 22 26 80 451 271 305 421 178 91 1888

1965 32 15 14 16 19 22 28 67 152 78 62 48 553


1966 15 14 15 12 25 44 68 136 212 242 152 204 1139
1967 58 22 16 15 15 15 20 35 52 91 20 10 369
1968 7 2 1 6 80 128 51 222 155 342 163 73 1230
1969 35 20 27 42 43 84 178 132 197 115 75 62 1010

Extracted from Bibra, E.E. and Riggs, H.C.W. (1971): Victorian River
Gaugings to 1969. State Rivers and Water Supply Commission.
204

TABLE E2. Monthly flow parameters for Mitta Mitta River


at Tallandoon (1936-69).

Mean Standard Coefficient Coeffi cien t


Month Flow Deviation of of Serial
(lOD m3) (l06 m3) Variation Skewness Correlation

January 39.9 26.4 0.66 1.8 0.58


February 26.9 15.4 0.57 1.1 0.60
March 31.1 32.3 1.03 3.3 0.54
April 46.8 81. 7 1. 74 4.9 0.85
May 62.1 70.1 1.13 4.1 0.81
June 102.5 118.9 1.16 3.2 0.63

July 152.9 132.2 0.86 1.3 0.59


August 198.8 151. 4 0.76 1.3 0.65
September 198.8 103.5 0.52 0.89 0.73
October 206.5 115.8 0.56 0.58 0.65
November 134.6 83.1 0.62 0.62 0.78
December 72.6 49.9 0.69 1.7 0.61

TABLE E3. Annual flow parameters for Mitta Mitta River


at Tal1andoon (1936-69).

Mean Standard Coe f fi ci en t Coefficient


Flow Deviation of of Serial
(l06 m3) (l06 m3) Variation Skewness Correlation

1274 731 0.57 1. 50 0.06


205

AUTHOR INDEX

Alexander, G.N. 39-45, 67, 147, 149, Franzini, J.B. 15, 186
151, 159-60, 169, 181, 191
Gani, J. 71,72,183
Anderson, R.L. 118
Ghosa1, A. 72, IS3
Asfur, H. 180, 181
Gould, B.W. 3, 49-50, 67, 83-91, 105,
Australia, Department of National
140-2, 152-169, 179, lS3-4,
Resources. 11, lSI
193-6
Barnes, F.B. 28, 110, 112, 181 Guglij, G. 140, 144, 159, 163-4
Beard, L.R. 57, 121-2, 129, 132-4, Gumbel, E.J. 60, lS4
139, 159, 176-7, 181
Hall, W.A. 180, IS4
Benson, M.A. 135, 181, 187
Hardison, G.H. 7, 15, 46, 59, 62-6,
Bibra, E.E. 181, 203
97-100, 105, 159, 161, 163,
Black, R.P. 185
184, 197
Brittan, M.R. Ill, 181
Harms, A.G. 124-5, 184
Brown, J.A.E. 7, lS2
Harris, R.A. 82, 83, 182, 184
Burden, R.P. 33, 115-6, 188, 193-5
Hazen, A. 27, 110, 184
Burges, S.J. 137, 182
Hilferty, M.M. 115, 122
Campbell, T.H. 124-5, 184 Howell, D.T. ISO, 184, 187, 192,
Chow, V.T. 121, 182, 190 196- 7
Codner, G.P. 147, 176-7, 182, 186, Hudson, H.E. 55, 184
187, 192-5 Hufschmidt, M.M. 135, 184, 186
Crawford, N.H. 7, 182 Hurst, H.E. 31-32, Ill, 184-5
Crosby, D.S. 177, 182
Jackson, B.B. 107, 183
Dear1ove, R.E. 83, 182 Jarvis, C.L. 82, 185
Dincer, T. 46-7, 49, 51, 67, 159-60, Joy, C.S. 32, 33,90, 91, 147, 185,
163, 169 lS7, 193-6
Doran, D.G. 83, 91, 177, 182 Julian, P.R. Ill, ISS
Dorfman, R. 135, IS2, 186
Kartvelishvili, N.A. 144, ISS
Dracup, J .A. ISO, lS4
Kendall, M.G. 114, ISS
Fair, G.M. 186 King, C.W. 21, 185
Fathy, A. 33, IS2 Kirby, W. 116, 132, 185
Feller, W. 31, lS2 K1emes, V. 91, 102, 107, 185
Fletcher, S.J. 179 Kottegoda, K. T. 108, 109, 125, 185
Fiering, M.B. 17, 33, 35, 107, 114-5,
Langbein, W.B. 96-97, 185
124, 126, 135, 158, 176, 176-7,
Law, F. 67, 186
183, 184, 188
Lindner, M.A. 182
206

Linsley, R.K. 7, 15, 137, 182, 186 Stall, J.B. 55, 151-2, 188
Lloyd, E.H. 82, 83, 186 Stuart, A. 114, 185
Sudler, C.H. 110, 188
Maas, A. Ill, 135, 186
Svanidze, G.G. 140, 144, 159, 188
McMahon, T.A. 15, 25, 44, 93-5, 105,
116, 126, 134, 147, 153, 159, Teoh, C.T. 90, 147, 159
162-4, 169, 177, 178, 182, 185, Thorn, II.S.C. 41, 188
186, 187, 193-6 Thomas, H.A. 33, 114-6, 124, 126,
Maddock, T. 177, 182 158, 175, 177, 186, 188, 193-5
Manoe1, P.J. 180, 187 Thompson, R.W.S. 53, 147, 150-1, 188
Marglin, S.A. 186 Tintner, G. 114, 189
Martin, R.O.R. 59, 184
United States Corps of Engineers,
Matalas, N.C. 117-8, 125, 135, 176,
20, 57, 58, 121, 176, 189
181, 187
United States Geological Survey,
Me1entijcvich, M.J. 101, 187
60, 61
Miller, A.J. 116, 186
Venetis, C. 83, 189
Moran, P.A.P. 3, 71-3, 81-3, 90,
96, 187 Waitt, F.W.F. 33, 36-8,147,150-1,
Moss, M.E. 35, 187 189

Neill, J .C. 55, 188 Wallis, J .R. 11, 125, 189


Wilson, E.B. 115-6, 122, 189
O'Connell, P.E. 11, 189
Wilson, H. McL. 67, 189
Odoom, S. 82, 186
Wright, G.L. 109, 182, 189
Pcrrins, S.J. 187, 193, 196-:
Yeh, W. W-G. 180, 181
Pisano, W.C. 177, 189
Yevjevich, V.M. 83, 109, 111, 188,
Phatarfod, R.M. 102-3, 116, 187
189
Philips, C. 109, 177, 187, 188
Young, G.K. 177, 189
Prabhu, N.U. 71, 72, 183, 188
Pyke, R. 71, 183

Rippl, W. 20, 188


Riggs, H.C.W. 181, 203
Roberts, W.J. 55, 184
Roesner, L.A. 83, 188

Savarenskiy, A.D. 3, 188


Schonfeldt, C.B. 180, 187
Searcy, J.K. 7, 188
Se1va1ingam, S. 140, 144
Shukry, A.S. 33, 182
Simaika, Y.M. 185
Srikanthan, R. 35, 126, 130
207

SUBJECT INDEX

Active storage 14
Alexander's method 38, 67, 169
evaluation of 149, 159
Annual flows
parameters 13
generating model for III
ARIMA model 125

Beard's normalizing procedure 121, 132


Behaviour analysis 24, 169
comparison with Gould method 153
for multi-reservoir systems 172, 174, 178
use with generated data 155, 170
Broken line model 125

Carryover frequency mass curve analysis 52


Carryover storage 14
and seasonal storage 65
Classification of reservoir capacity-yield procedures 3
Moran derived methods 71
Coefficient of skewness 9
Coefficient of variation 9
effect on zone size (Gould) 91
Conceptual storages 14
Critical drawdown period 20
Critical period 19
and evaporation loss 191
Critical period techniques 3, 19
evaluation of 147
Cross-nesting of mass curves 64, 68

Design
preliminary 2, 169
procedures in current use 5, 173
final 2, 169
Design process 2
Demand 15
allowance for increase in 174
208

Dincer's method 46, 67, 169


evaluation of 159
Distribution of flows 7, 44
Gamma 40, 49, 72, 103
Gumbel (Extreme-value) 60, 137
like Gamma 115,129,132
Log-normal 97, 117, 129, 132, 142, 168
Normal 46, 49, 97, 110, 132
Weibu11 97, 132
Draft 15
sampling error of 164
Drought curves 36, 42
Dynamic programming 180

Evaporation loss, adjustment for 190


Extreme-value distribution 60, 137

Failure, probability of (see Probability of failure)


Final des ign 2
recommended procedures 169
Finite storage 14
Flow parameters (see Inflows, Distribution of flows)
Fractional Gaussian Noise model 117
Frequency distributions of parameters 7

Gamma distribution 40, 49, 72, 103


shape parameter for 41, 49
Gamma units 49
Generated data (see also Stochastic data generation)
comparison with historical data 126
negative flows 112
when and how to use
single reservoir 136
multi-reservoir systems 177
Gould's Gamma method 49, 67, 169
evaluation of 159
use for determination of sampling error 165
Gould's Probability Matrix method 83, lOS, 169
effect of starting month 91
effect of zone size 90
evaluation of 152
failure vector for 87
209

for multi-reservoir systems 179


transition matrix for 84
use with generated data 155
Gould's Synthetic Data procedure 140
evaluation of 159
use for determination of sampling error 164
Gug1ij 's method 144
evaluation of 159
Gumbel distribution 60, 137

Hardison's method
for combination of carryover and seasonal storage 65
for probability routing 97, 105
evaluation of 159
Hurst phenomenon 31, 108
importance of 125
Hurst's storage procedure 31, 67

Independent series 55
evaluation of 151
vs. overlapping series 57
Index of variability 9, 98
Infinite storage 14
Inflows 6
and outflows, occurrence of 72, 78
distribution of 7 (see also Distribution of flows)
estimated 7
generated 107, 111, 114
sampling error of 6, 11
statistical parameters of 7, 11, 108
Initial conditions
effect of full assumption 25
independence of effect of 82

Kirby's transformation 116, 132

Langbein's Probability Routing method 96


Like Gamma distribution 115, 132
Log-normal distribution 97, 117, 129, 132, 142, 160
Low flow frequency curves 59, 65, 169
bias in 62, 64
210

Low flow sequences


independent series 55
methods based on 36
minimum flow approach 36
overlapping series 52

~!C~lahon I s empirical equations 93, 105, 169


evaluation of 159
Markov model Ill, 130, 199
~lass curve method 20, 67
evaluation of 147
frequency analysis 52, 59
bias 64
cross-nesting 64, 68
res idual 22, 67
Matalas transformation equations 117
multi-site 176
Mean of flows 7
~ledian of flows 7
Moment transformation equations 117
Monte Carlo model, use of 110, 144
Moran probabi Ii ty matrix procedures 3
general classification of 71
mutually exclusive model 73
equations for general case 76
transition matrix for 74
simultaneous model 78
Mul ti-reservoir probl em 2, 171
flow generation for 175
transition matrix approach 179
system reliability 172
Mutually exclusive model (Moran) 73

Negative generated flows


Newton-Raphson method 117, 201
Noah and Joseph effects 108
Normal distribution 46, 49, 97, 110, 132
Normalizing flows 121, 132

Operating rule 16
for multi-reservoir systems 178
Outflows 15
211

Overlapping series 52, 68


clustering 53
evaluation of ISO
vs. independent series 57

Performance index (Fiering's) 17


Persistence 109
measure of 10, 109
Phatarfod's procedure 102
Preliminary design 2
recommended procedures for 169
Principal component method 176
Probabi 1 i ty mat rix methods 71
evaluation of 147
Probability of failure
and risk 174
annual vs. monthly 87
definition of 16
for multi-reservoir systems 173
Probabili ty routing 96 (see llardison' s method)

Random walk 101, 102


Range, definitions of 30
Recycled historical sequences 174
Regional within-year storage estimates 63
Regression method for multi-site generation 176
Regulation 15
Release 15
rule 16
Reliability 17, 172
volumetric 17
Residual mass curve 22
Restrictions 16
Risk
definition of 174
of storage estimate being low 137

Samp Ii ng error
of draft estimate 164
of flows 6, 11
of storage estimate 164
Seasonal storage 14, 65
212

Semi-infinite storage 14, 27


Sequent peak algorithm 33
use of 67, 132
Serial correlation
adjustment for effect of 169, 193
confidence limits for 109
definition of 10
preservation by generating models 130
significance 44, 109
Sequences
low flow 36
independent 55, 151
overlapping 52, ISO
recycled historical 174
Shape parameter, Gamma distribution 41, 49
Simulation 135 (see Behaviour analysis)
Simultaneous model (Moran) 78
Skewness
measures of 9
modifications to cater for 115, 117, 121
preservation by generating models 129
Spill IS
Standard error of parameters 11
Stochastic data generation
multi-site models 175
key station method 175
performance of 177
principal component method 176
regression method 176
residual approach 176
single site models
annual flows 110
evaluation of storage estimates using ISS
generalized results using 140
modifications for non-normal flows 115, 117, 121
monthly 114
number of sequences required 139
theoretical justification of 199
two tier model 124
verification and performance 126
213

Steady state probability 79


Storage, definitions of 14
Storage estimates
effect of record length 158
effect of serial correlation 169, 193
rapid procedures for 159
sampling error of 164
Svanidze's method 144
evaluation of 159
Thomas and Fiering mode 1 114, 126
theoretical justification of 199
Time interval 6
Time series components of flows 108
Transition matrix
Gould's method 84
Moran model 74, 78
steady state 76, 79
multi-reservoir systems 179
Two tier model 124

Variability, measures of 9

Waitt's procedure 36, 67


evaluation of 147
Weibull distribution 97, 132
Wilson and Hilferty transformation 115
Within-year frequency mass curve analysis 59
Within-year storage 14
regional estimates of 63

Yield 15
adjustments for serial correlation 197
This Page Intentionally Left Blank

Potrebbero piacerti anche