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

APPLICATION OF 'ILLUDAS'

TO STORMWATER DRAINAGE DESIGN IN


SOUTH AFRICA

M. D. Watson

Report No. 1/81


Urban Hydrology Series

Hydrological Research Unit


University of the Witwatersrand
Jan Smuts Avenue
Johannesburg
April 1981
Director: Professor D. C. Midgley
i

PREFACE - SERIES IN URBAN HYDROLOGY

Considerable research is being conducted throughout the world on


stormwater hydrology and the effects of urbanization on runoff.
A group of researchers has been established within the Hydrological
Research Unit at the University of the Witwatersrand to conduct
research on these lines, bearing in mind the importance of flooding in
the urban environment. The s~udies embrace catchment monitoring for
runoff and pollution, mathematical modelling of urban catchments and
development of design aids for estimating floods in South Africa.

The mathematica~ modelling approach has proved very suitable for urban
drainage studies and challenging research has been commenced on these
lines. This particular report documents the adaptation and use of an
isochronal-type model for simulating the flow in storm-water drains.
ii

ABSTRACT

The Illinois Urban Drainage Area Simulator (ILLUDAS) is evaluated


for the purpose of stormwater drainage design in South Africa. The
historical development of the model is described and published
verification results are reviewed. The principles of the model are
described, as also are the features of the computer program which
was made available during the early part of 1980 by the Illinois
State Water Survey (ISWS).

ISWS recommendations for parameter estimation were tested with


rainfall-runoff data from two local catchments and in general
found to be adequate. Improvements to the method of accounting
for grassed area losses in sub-catchments having long flow paths
were found to be necessary.

Problems associated with design inputs are discussed and use of


the Chicago design storm is demonstrated. A modified metric
version of the program is presented and tentative recommendations
are made for design parameter estimation.
iii

TABLE OF CONTENTS
Page
Preface i
Abstract ii
Chapter 1 : Introduction 1

Chapter 2 : Development and testing


2.1 General 6
2.2 The TRRL hydrograph method 6
2.3 ILLUDAS 13
2.4 Discussion 16

Chapter 3 : Description of the model


3.1 Introduction 18
3.2 An overview of the model 18
3.3 Paved area runoff 21
3.4 Grassed area runoff 23
3.5 Entry time 25
3.6 Infiltration 26
3.7 Combining of hydrographs 28
3.8 Routing procedure 28
3.9 Flood control ponds 30
Chapter 4 : Local verification
4.1 General 33
4.2 Pinetown catchment 33
4.3 Kew catchment 42
4.4 Discussion 58

Chapter 5 : Storm inputs for design


5.1 General 59
5.2 The Chicago design storm 61
5.3 Evaluation of the Chicago storm 64
5.4 Estimation of parameters 74
5.5 Discussion of results 77
iv

Page

Chapter 6 : Modifications to ILDUDAS


6.1 General 79
6.2 Rainfall input 79
6.3 Grassed area losses 79
6.4 Sensitivity analysis 80
6.5 Pipe sizing 81
6.6 Further modifications 81
Chapter 7 : Conclusions 82

References 84

Appendix A Users Manual Al

Appendix B Program listing Bl

Appendix C List of variables Cl

Appendix D Pinetown rainfall and runoff data Dl

Appendix E Kew rainfall and runoff data El


Appendix F Norwood severe storms Fl
- 1 -

CHAPTER 1 INTRODUCTION

Changes from rural to urban land use have a dramatic influence on


runoff. Paving of large areas reduces infiltration and increases
the velocity of overland flow. Collection of runoff into gutters,
lined channels and pipes reduces the length of overland flow and
also increases flow velocities. The time required for water to
accumulate at any point is thus reduced, making discharge highly
sensitive to short-duration high-intensity rainfall. Analyses of
records collected before and after urbanization generally
indicate that urban flood peaks are three to four times as large
as the corresponding rural floods (Cordery, 1976).. The increase is
less significant, however,
for the more extreme storm events.

The primary functions of a stormwater drainage system are:

(i) the removal of street surface flows under short term


conditions to the extent required to avoid unreason-
able levels and frequencies of inconvenience to
pedestrians and traffic, and
(ii) the elimination or minimization of flood damages and
hazards under long-term storm conditions.

Design thus involves first the sizing of an underground conduit


network to remove flows exceeded at low recurrence interval
(eg. 2 years) and secondly the checking that damages and
hazards are acceptable for more infrequent storm events.

As new areas develop there is an increase in the volume of runoff


in downstream channels and it becomes necessary to ensure that
downstream development does not experience frequent or severe
inundation. Downstream culverts and bridges may have
to be enlarged or storage facilities incorporated into new
developments to cope with the increased flows.

Other factors to be considered in design are maintenance costs,


reliability of operation, inlet capacity, stream bank erosion,
- 2 -

adaptability of the drainage system to future expansion and


possible pollution of receiving waters.

At present in South Africa the Rational Method and Unit Hydro-


graph techniques are the most commonly used for design flood
determination. The Rational Method is employed to determine peak
discharges and the Unit Hydrograph method to establish the
temporal distribution of runoff.

The limitations of the Rational Method have been widely discussed


(Aitken, 1975; McPherson, 1969; J.P. MacLaren Ltd., 1975). The
method is recognised, however, as a valuable design technique for
small catchments (eg. where pipe sizes do not exceed 600 rom),
and as an approximate technique for larger catchments. The method
does not adequately account for pipe routing or variations in
rainfall intensity, contributing area and rate of contribution.
The lumping of all physical factors into two parameters (runoff
coefficient and time of concentration) makes parameter estimation
subjective and could be the source of serious errors.

Schaake et al (1967) showed for five catchments smaller than


20 ha that 20% of Rational Method estimates were in error by 25%
or more. Less encouraging information is given by Ardis et al
(1969) who show independent estimations of the Rational runoff
coefficient for the same area to vary between 0,2 and 0,5. A
similar local study (Watson and Miles, 1980) shows estimates
ranging from 0,23 to 0,78.

The parameters of the Rational Method cannot readily be verified


from short term rainfall-runoff records. Guidance provided in the
literature is rarely based on sufficient data and is more usually
of a subjective nature.

The Rational Method assumptions are generally found to be too


simplistic to permit actual discharge to be adequately predicted
from observed rainfall. A mean ratio of predicted to observed peak
discharge of 1,23 with a standard deviation of 0,50 is reported by
J.P. MacLaren Ltd (1975) for four catchments smaller
- 3 -

than 36 ha. Better results are reported by Watkins (1962) who


obtained a mean ratio of 1,12 with a standard deviation of 0,42
for storms observed on twelve catchments in the U.K. The better
results are due to the calibration of the runoff coefficient with
observed volume of runoff.

Watkins also evaluated the performance of the Unit Hydrograph


technique and found it adequate as long as sufficient runoff data
were available to determine the hydrograph shape. As data are
generally not available for urban areas, use of the unitgraph is
limited. Use of regional data from rural catchments is
theoretically unsound and cannot be recommended.

Due to the complexity of flood prediction as well as related


problems of stormwater pollution wide use is being made of
computer based techniques. These are generally mathematical models
of the runoff process and their use has many advantages.
Parameters have to be objectively estimated and this encourages an
understanding of the runoff process. The user can focus attention
on the actual problem rather than on the calculations. Sensitivity
analysis is desirable and can be easily accomplished using these
techniques. Tests using recorded data indicate
that mathematical models can accurately and reliably simulate
the runoff hydrograph.

The use of mathematical models for the assessment, planning,


design and control of storm and combined sewerage systems has
become widespread overseas. The potential user now has many well
documented and tested models at his disposal. Table 1.1 lists
the principal features of 24 such models as assessed by the U.s.
Environmental Protection Agency (EPA) in 1975 (Brandstetter et
al, 1976). Most of the programs are available at a nominal
charge to cover handling.

It should be noted that some of the models are continually


being developed and probably already have more capabilities
than indicated in the table. The U.S. EPA's Stormwater
Management Model (SWMM), for example, is now capable of con-
- 4 -

tinuous simulation. Program availability has also increased.


Much of the Water Resource Engineers program has been incor-
porated into SWMM which is freely available. The Hydrocomp
Simulation Program (HSP) is also freely available in a Fortran
version (HSPF) from the u.S. EPA.

Table 1.1 : Comparison of major model categories by the u.S.


Environmental Protection Agency (1975)

The present research is directed at evaluating the application of


one such mathematical model to local stormwater drainage design
practice. The Illinois Urban Drainage Area Simulator, ILLUDAS,
has been selected as being most appropriate for local design
needs. It is one of the few models that has pipe sizing as a
major objective; it is conceptually simple and has been widely
used and tested.

The major objectives of this report are to assess how well the
- 5 -

model will perform under local conditions and how the parameters
may be estimated for design purposes. The limited availability of
local urban rainfall and runoff data make this assessment
difficult and further data collection and subsequent assessment
are recommended. A modified metric version of the model is
presented and tentative recommendations for parameter estimation
are made. The successful performance of the model makes
it a promising design tool and it is therefore recommended for
local design use.
- 6 -

CHAPTER 2 DEVELOPMENT AND TESTING

2.1 GENERAL

ILLUDAS, which stands for Illinois Urban Drainage Area Simulator,


is a development of the popular British Transport and Road Research
Laboratory method, called the TRRL (previously RRL) hydrograph
method. The development of ILLUDAS was undertaken by Terstriep and
Stall of the Illinois State Water Survey (ISWS) with the aim of
providing engineers with an objective and reliable technique for
stormwater drainage design that would require little more input
than would be required for the Rational Method, and at the same
time would allow the user to examine various alternatives in
.arriving at a final design. Both ILLUDAS and the TRRL method have
had widespread use and have undergone considerable testing and
refinement.

2.2 THE TRRL HYDROGRAPH METHOD

During the early 1950's the United Kingdom Department of Scientific


and Industrial Research initiated a research program to examine the
current methodology in stormwater drainage design. The research was
undertaken by the Road Research Laboratory (now the Transport and
Road Research Laboratory). Eleven representative urban catchments
ranging in size from 0,5 to 250 ha and one mainly rural catchment
of 2100 ha were instrumented. In all,
286 storms on these catchments were analysed and a report was
published recommending what is now called the TRRL hydrograph
method (Watkins, 1962).

The report examined existing techniques (viz. the Rational and


Unit hydrograph methods) and compared them with the recommended
TRRL method. The comparisons showed the TRRL method to be the
most accurate and reliable in reproducing observed events, and on
this basis it was recommended for general design use.

As originally developed, the TRRL method is conceptually similar to


the well-known model proposed for flood estimation on rural
catchments by Clark (1943). In Clark's model the rainfall excess
- 7 -

is applied to the time-area diagram and then routed through a


linear storage at the outlet of the catchment. The TRRL method
also applies the rainfall excess to the catchment time-area
diagram. The routing, however, is through a slightly non-linear
storage at the outlet of the catchment. The method involves:

(a) computation of the catchment time-area diagram using


pipe-full velocities for all pipes in the catchment

(b) computation of the temporal pattern for the rainfall


excess hyetograph

(c) combination of the results of steps (a) and (b) to


yield the 'inflow hydrograph'

(d) computation of the storage-discharge curve based on the


assumption of uniform proportional depth. (This
assumption implies that at any time the proportional
depth of flow is the same throughout the stormwater
drainage system)

(e) routing of the inflow hydrograph through pipe storage


determined in (d) using a stage-discharge curve for the
outlet pipe.

In the tests by Watkins (1962) runoff from pervious areas was found
to be negligible and therefore ignored in the calculations. The
percentage of impervious area ranged from 2% to 100% and runoff
from such areas was determined using a percentage loss rate or in
some cases a uniform loss rate based on observations.
Storage:discharge relationships used in the calculations were
determined from the recession portions of observed hydrographs. The
mean ratio of computed to observed peak discharge for all areas was
0,98 with a standard deviation of 0,15.

The recommended assumption of uniform proportional depth in cal-


culating detention was found to be satisfactory for most areas. In
old development areas the presence of exceptionally large upstream
sewers, however, caused this assumption to lead to an over-
estimation of detention and therefore a serious underestimation of
peak discharge. The assumption was considered adequate only for
systems with a reasonable degree of taper
- 8 -

(i.e. a proportional reduction in pipe capacity with reduction


in area served) and with pipes of similar geometric properties.

It was also recommended that, subject to some qualifications in


exceptional circumstances, the whole area of paved surface should
be considered impermeable in sewer design calculations, and the
unpaved areas should be taken to be completely pervious.

The method established itself as a standard design method in the


U.K. Criticism of its one-step routing procedure led to a modi-
fication by Watkins and Young (1965). The routing was applied to
one pipe length at a time. This avoided the limitation of pipe
taper and enabled potentially surcharged pipes to be redesigned
before proceeding to the next reach.

Terstriep and Stall (1969) evaluated the performance of the orig-


inal method on three catchments in the USA. The catchment sizes
were 0.2, 5.2 and 930 ha and the percentages of paved area 100%, 45%
and 44% respectively. The average ratio of computed to observed
peak discharge for a total of 39 events was 0,94 with a standard
deviation of 0,15. In their calculations Terstriep and Stall used
Watkins' design recommendatian of 100% runoff from paved surfaces
directly connected to the sewer system. For the largest catchment
the storage:discharge relationship was obtained from observed
hydrographs. The smallest catchment, a parking let with no pipe
drainage did not require storage routing. For the remaining
catchment the uniform depth assumption was applied.

The authors concluded that the method provided a highly useful


tool far developing an understanding of storm runoff but recog-
nised the weakness of the one-step routing procedure.

In 1972 these authors reported a further evaluation of the method


using rainfall and runoff data for ten urban catchments ranging
in size from 6 to 2200 ha. Percentage of paved area ranged from
21 to 61%. In all 122 storm events were analysed. The mean
absolute error in the computation of peak discharge was 41%.
- 9 -

The principal findings of this study, as summarised by Aitken


(1975), were as follows:

(i) The model is accurate for computing runoff from paved


areas

(ii) The model is generally satisfactory providing:

 the area is less than 13 km2

• the directly connected impervious area exceeds 15%


of the catchment area, and

• the return period of the storm event is not greater


than 20 years

(iii) The model breaks down where there is significant runoff


from pervious areas; this occurs for pervious areas with
steep slopes, tight soils or high soil moisture
conditions

(iv) The model should not be promoted for general use (in the
USA) because as indicated in (ii) and (iii) under certain
circumstances the results are not satisfactory

(v) More good quality rainfall-runoff data are required

(vi) The model should be developed to include pervious runoff


and further use should be made of water balance models
such as the Hydrocomp Simulation Programme
(HSP)

Papadakis and Preul (1973) tested the original TRRL method for two
storms (three peaks) on Oakdale Avenue catchment in Chicago, USA.
The catchment has an area of 5,2 ha of which 45% is paved. The mean
ratio of computed to observed peak discharge was 1,15 with a
standard deviation of 0,31. Data from this area had been previously
used by Terstriep and Stall (1969) who noted that some observed
peaks were truncated due to problems with the recording device.
Further errors can partly be accounted for
in neglect of pervious runoff and initial losses.

Testing of the modified TRRL method in Australia was reported by


Aitken (1975) .The results of evaluations by Aitken(1968,1973), Read
- 10 -

(1971), and by Heeps and Mein (1973) are presented. The mean value
of the ratio of computed to observed peaks was 0,95 with a high
standard deviation of 0,45. Aitken concluded that the TRRL method
did not accurately simulate the runoff process. Examination of the
paper by Heeps and Mein (1974), however, shows that apart from the
sampling errors many of the errors can be ascribed to neglect of
pervious area runoff.

A further evaluation of the modified version for 57 events on four


catchments ranging from 5 to 36 ha in area is reported by J.F.
MacLaren Ltd. (1975). The percentage imperviousness of the
catchments ranged from 27 to 68%. The mean ratio of computed to
observed peak discharge is 1,13 with a standard deviation of 0,31.
One catchment (Northwood) showed large errors and after further
investigation was rejected as unreliable. The mean peak discharge
ratio for the remaining three subcatchments (43 events) was 1,ll
with a standard deviation of 0,22.

Another of the catchments studied was the Oakdale Avenue catch-


ment in Chicago. Thirteen storms were analysed including the two
used by Papadakis and Preul (1973). A mean peak discharge ratio
of 1,00 with a standard deviation of 0,23 was obtained. The
improved results achieved were possibly due to a change in
parameter estimation technique as well as to the larger data base
used.

Several models were tested in this study including the US EPA's


SWMM, the UCUR model of the University of Cincinnati, and the
TRRL model. It was concluded that SWMM gave the best overall
performance, but that the TRRL model had the advantage of
simplicity and, in general, performed as well as the other
methods in real event simulation. Its neglect of runoff from
pervious areas was considered a drawback and its routing compu-
tations were shown to be inadequate in two extreme examples. For
pipe slopes of 0,1% and lengths of 5500 m and 9100 m peak
discharges were shown to be in error by -37% and -50% respec-
tively.

Ford (1975) reports modifications of the method to allow for


runoff from unpaved areas. Losses were accounted for by a run-
- 11 -

off coefficient and runoff attenuated by a linear reservoir sub-


model. The modified method was calibrated and tested using 44
events on catchments in Nairobi and Kampala. The sizes
ranged from 35 to 63 ha with percentage paved area ranging from 16
to 29%. The mean ratio of computed to observed peak discharge was
1,00 with a standard deviation of 0,19.

The use of parameters derived from calibration accounts for the


high quality of results achieved. Tentative recommendations
for the estimation of pervious area parameters based on type of
development were made on the basis of the catchments studied. It
was recognised, however, that these parameters could not readily be
related to catchment characteristics, such as slope and drainage
density.

A recent study by Williams et al (1980) is of interest since it


shows that comparable results can be obtained with the TRRL
method using estimated parameters or with the Unit Hydrograph
method with calibrated parameters. Values derived by the two
methods were compared with observed values for one event on a
catchment of 107 ha, of which 35% was impervious. Both methods
overestimated discharges at the start of the storm and under-
estimated the peak discharge by approximately 30%. It was con-
cluded that both m0thods reproduced observed flows well and that
the introduction of a better rainfall loss model would improve
simulations.

The results of all these tests are summarised in Table 2.1. For
further information on performance of the TRRL method the reader
is referred to the original publications as well as an excellent
summary by Colyer and Pethick (1976).
- 12 -
- 13 -

2.3 ILLUDAS

In order to render the TRRL method generally applicable to urban


catchments in the USA, Terstriep and Stall (1974) incorporated an
improved method of accounting for losses. Each sub-catchment was
assumed to consist of three sub-areas: a paved area contributing
runoff directly to the drainage system, a grassed (or pervious)
area and a paved area which contributed runoff to the grassed area.
Depression storage losses were accounted for by
an initial abstraction from rainfall and infiltration losses were
accounted for by a time-decreasing loss rate on the grassed area.
The modified method was named ILLUDAS, and was tested with 246
storms on 21 urban catchments in the USA. Areas ranged from 0,2 ha
to over 2000 ha with percentage of paved area ranging from 21% to
100%. With minor exceptions the data originally measured on
topographic maps or obtained from the gauging agency were used
throughout the study. Infiltration parameters were estimated on the
basis of soil type and the depth of rainfall during the 5 days
preceding the storm. Overland flow travel times were computed using
an equation developed by Hicks (1944).

The results are summarised in Table 2.2 (catchments 1 to 21). The


mean ratio of computed to observed peak discharge is 1,04 with a
standard deviation of 0,37. Fairly high errors are evident for some
catchments and Terstriep and Stall concluded that the method
produced acceptable results for 12 of the catchments. Three other
catchments were considered marginal, 3 indeterminate and 3 were not
acceptable. In all cases results could have been improved by making
adjustments to the estimated catchment parameters such as
contributing areas and inlet times.

Data from two rural catchments were used to test the grassed
area runoff computations. The catchment sizes were 3 ha and
7 ha. Parameters were estimated in the same fashion as for
urban catchments. In all 12 events were tested. ILLUDAS predicted
zero runoff for four of these, three of which had an average
runoff of 0,4 mm and one of 15 mm. The mean ratio of computed to
observed discharge for the remaining G events was 0,65 with a
standard deviation of 0,20. The prediction of runoff volume was
slightly better with a ratio of computed to ob-
- 14 -
- 15 -

served of 0,84 with a standard deviation of 0,27. The results were


considered acceptable considering the increased sensitivity to
antecedent moisture conditions and the number of factors involved
in a complete analysis of pervious runoff not included
in ILLUDAS.

The presence of data errors for both urban rural catchments was
recognised but no attempt was made to eliminate them. This and
the use of estimated parameters prevents direct comparison of the
results with those from other verification studies.

Some of the data used to assess the performance of the TRRL method
were used in the assessment of ILLUDAS. A comparison of results
shows a decrease in the mean absolute error for the prediction of
peak discharge on 8 catchments from 42% to 37%. On another
catchment the error increased (20% to 49%). This was probably due
to inadequate consideration of antecedent moisture conditions.

Patry et al (1979) report the adaptation of ILLUDAS for the


HP9830 - 16k byte programmable calculator. This has been achieved
with no loss of flexibility: in fact, new features such as graph
plotting and hydraulic grade line analysis have been
incorporated. The model has been successfully applied to two
catchments in Canada, one of 23 ha and the other 862 ha.
Percentages of paved area were 34% and 49% respectively. The mean
ratio of computed to observed peak discharge for 22 events was
0,96 with a standard deviation of 0,16. The results are
summarised in Table 2.2 (catchments 22 and 23). The parameters
were obtained by calibration and some of the remaining error
could be attributed to spatial non-uniformity of rainfall.

Water quality algorithms of the US EPA's model SWMM have been


adapted for ILLUDAS in a version known as QUAL-ILLUDAS. This is
a single-event model and was reported on by Terstriep et al
(1978).

The routing computations of ILLUDAS have been improved since


its inception. Initially an explicit solution to the continuity
equation was used whereas a more accurate method using an
- 16 -

implicit solution is now available. For many reaches routing is


not significant and a simpler time-shift method is also available.

ILLUDAS has had fairly widespread use in the USA and Canada since
its inception. Its use in Canada has led to the following re-
commendations (Wisner et al, 1979) regarding practical application
to design:

Table 2.3 : Applicability of ILLUDAS

Applicability of
Use
ILLUDAS
Rural areas Fair
Preliminary analysis Very good to excellent
Detailed analysis of free surface peak flows Good
Detailed analysis of valley storage Fair
Surcharged flow Poor

2.4 DISCUSSION
ILLUDAS is a logical extension of the TRRL method to account for
pervious area runoff. Its use for design seems reasonable because
of its logical accounting for losses during severe events. The
test results presented in this chapter do not, however, show any
real improvement over the TRRL method in its capability of
reproducing observed events. This could be partly due to the fact
that most observed events are not sufficiently intense to produce
significant runoff from pervious areas. It does seem, however,
that the major cause could be the presence of errors
in the data and the different methods of assessment adopted in
the various studies.

Data errors can be due ,to many causes including spatial non-
uniformity of rainfall, discharge rating errors, drowned
weir, leakage or blockage of pipes and gauge malfunctions. Some
of these difficulties are reported both by Watkins (1962) and
Terstriep and Stall (1974) in the original assessments of the
two models. The different approaches adopted by these re-
- 17 -

searchers are significant when comparing the results. Watkins


corrected some observations and discarded others of dubious
quality. Terstriep and Stall, on the other hand, retained all
data uncorrected and merely acknowledged the possibility of data
errors.

The use of calibrated parameters in some studies also complicates


comparisons. Over half of the tests on the TRRL method were based
on the use of calibrated parameters. In contrast over
90% of the ILLUDAS tests were based on estimated parameters.

On average the models give good estimates of peak discharge.


The mean ratio of computed to observed peak discharge is 0,99
for the TRRL tests and 1,03 for ILLUDAS. The corresponding mean
absolute errors are 21% and 27% respectively, showing that in some
cases large errors are present. Many of these are probably due to
poor data rather than inherent inadequacies of the models.
- 18 -

CHAPTER 3 DESCRIPTION OF THE MODEL

3.1 INTRODUCTION

This chapter describes the principles of the ILLUDAS model as well


as the features of the version of the computer program made
available by the Illinois State Water Survey (ISWS) during the
early part of 1980. The description is based largely on the
original ILLUDAS report (Terstriep and Stall, 1974), while the
section on routing options is drawn from a note received from ISWS
with the program.

3.2 AN OVERVIEW OF THE MODEL

In order to apply ILLUDAS the catchment to be studied is broken


down into sub-catchments. A sub-catchment is generally defined as
a homogeneous portion of the catchment draining to a single inlet
or set of inlets that constitutes a specified node in the
drainage network.

Fig. 3.1 illustrates a typical subdivision. The points A through H


represent design points in the sy"stem. A to G. are groups of
inlets but could equally well be manholes, changes in grade or
changes in cross-section. Point H is the __ outfall. These points
are connected by reaches which may be circular pipes, closed
rectangular channels or open trapezoidal channels.

Inlet hydrographs are computed for each sub-catchment by con-


sidering the paved and unpaved (or grassed) areas separately.
The inlet hydrographs from each sub-catchment are accumulated
in downstream order proceeding through the· catchment. This
accumulation of inflow hydrographs is routed through each reach to
account for temporary storage. The result is a computed outflow
hydrograph from each reach, and ultimately a hydrograph at the
outlet of the entire catchment. The program can be used for the
hydrological design of new drainage systems or for the evaluation
of existing systems.

A flow chart for the program is given in Fig. 3.2.


- 19 -

Fig. 3.1 Typical subdivision


- 20 -

Fig. 3.2 Flow chart for ILLUDAS


- 21 -
3. 3 PAVED AREA RUNOFF

Paved areas are considered to be completely impervious and are


divided into two types; those which are directly connected to
the stormwater drainage system and those which drain onto
pervious areas. The first type generally includes roads,
paved parking areas and buildings that discharge runoff directly
onto the streets or into the stormwater drains. The second type
is assumed to supplement the runoff from pervious areas and is
considered in section 3.4.

The steps involved in developing a runoff hydrograph from the


directly connected paved area are illustrated in Fig. 3.3. Ex-
tending down the middle of the sub-catchment in Fig. 3.3a is
a street with a- pair of inlets at its lower end. Also shown
are rooftops and driveways. The directly connected area is
shown shaded.

Assuming runoff to flow at a constant velocity, travel times can


be determined for runoff from various parts of the paved area to
reach the inlets. These travel times are plotted on the paved
area, and by connecting points of equal travel time a series of
isochrones are drawn on the paved area, as shown in Fig. 3.3a.
The directly connected paved areas between these isochrones are
measured and designated areas PAl, PA2, PA3, PA4 and PAS. These
various areas are accumulated and plotted against travel time to
the inlet as shown in Fig. 3.3b. This time-area curve shows the
extent of paved area within the subcatchment that is contributing
water at the storm drain inlet at any time after the beginning of
runoff. In the computer program the time-area curve is assumed to
be a straight line connecting the origin and the end-point of the
curve.

Rainfall is input as a series of intensities of equal duration,


as shown in Fig. 3.3c. The time increment should be the same as
the time interval between the isochrones. Losses due to initial
wetting and depressions are combined and treated as an initial
loss to be subtracted from the beginning of the rainfall (Fig.
3.3d). The excess rainfall is referred to as the paved area
supply rate (PASR) and is shown in Fig. 3.3e.
- 22 -

Fig. 3.3 Development of the paved area hydrograph


- 23 -

The ordinates of the paved area hydrograph are computed by


applying the supply rate to the time-area diagram as shown in
Fig. 3.3f.

3.4 GRASSED AREA RUNOFF

Unpaved portions of the catchment are termed grassed areas.


Computations for grassed-area hydrographs closely parallel those
for paved-area hydrographs. Fig. 3.4a represents the same sub-
catchment used to illustrate the paved area runoff. The shaded
area represents the contributing grassed area which in this case
is assumed to include only front gardens. Additional grassed area
could contribute to runoff, but for the example in Fig. 3.4 it is
assumed that the contribution from the neglected areas would be
~ufficiently delayed so that its contribution would be
insignificant.

Construction of the time-area diagram is similar to that for


the paved area and is illustrated in Fig. 3.4b. In the com-
puter program the time-area diagram is again assumed to be a
straight line.

The rainfall input is the same as that for the paved area and
is shown in Fig. 3.4c. Runoff from the paved area draining onto
the grassed area is shown in Fig. 3.4d and is termed
supplementary paved area runoff (SPARO). This runoff is assumed
to be instantly and uniformly distributed over the grassed area
and is added to the rainfall input.

The losses illustrated in Fig. 3.4e include an initial loss to


account for depression storage plus infiltration. The grassed-
area supply rate in Fig. 3.4f is obtained by subtracting these
losses from the sum of rainfall and supplementary paved area
runoff.

The ordinates of the grassed-area hydrograph are computed by


applying the grassed area supply rate to the time-area curve
as shown in Fig. 3.4g.
- 24 -

Fig. 3.4 Development of the grassed area hydrograph


- 25 -

3.5 ENTRY TIME

The travel time for runoff from the hydraulically-most-remotepoint in


the sub-catchment to the entrance of the stormwa~er drainage system
is termed the entry time. This can either be supplied by the user or
can be determined by the program if it is sup-
plied with length and slope of the flow path. The methods used
by the program are simplistic but do give order of magnitude
estimates which are usually adequate for design.

For the paved area the program uses Manning's equation to compute
the travel time in the gutter assuming a hydraulic radius of 60
rom and a retardance coefficient, n, of 0,02. The n value of 0,02
is reasonable for street gutters while the assumed hydraulic
radius implies an average flow depth of about 70 rom for a 1 in
10 cross-slope gutter - a reasonable value for minor road drains.
The overland flow time for runoff to reach the gutter is assumed
equal to 2 minutes. This is added to the gutter travel time to
obtain the entry time.

For the grassed area an equation based on the empirical equation


by Izzard (1946) for time to equilibrium of overland flow is used.
In metric units the equation can be written as

where i = supply rate' (rom/h)


c = retardance coefficient
L = length of overland flow (m) S
= slope (%)

The program assumes a retardance coefficient for bluegrass turf


of 0,05 and a supply rate of 25 mm/h. Eq. (3.1) gives overland
flow travel time e~timates which are generally 30% smaller than
that from the equation of Izzard. These travel times are added to
the paved area entry time to obtain the grassed area entry time.
- 26 -

3.6 INFILTRATION

The runoff that is absorbed by the soil is termed infiltration.


This loss varies with the nature of the soil and its surface
cover as well as the soil moisture conditions. The U.S. Soil
Conservation Service describes the four hydrologic soil groups as
follows:

A - Low runoff potential, high infiltration rates


(consist of sand and gravel)

B - Moderate infiltration rates and moderately well


drained

C - Slow infiltration rates (may have layers that


impede, downward movement of water)

D - High runoff potential, very slow infiltration


rates, (consist of clays with a permanent high
water table and a high swelling potential).

Standard infiltration capacity curves have been devised for use


in ILLUDAS for soils of hydrologic groups A, B, C, and D with
lawn cover. These curves are based on the Horton equation

where fo = initial infiltration rate (rom/h) f

= final infiltration rate (rom/h)


c
k =a shape factor (h-l), selected as k =2 h-l
t = time from start of rainfall (minutes)

Corrections are made for the rainfall being less than the in-
filtration capacity and the equation is solved by the Newton-
Raphson technique.

Fig. 3.5 shows these infiltration capacity curves with initial


infiltration capacities assumed for different soil types and
antecedent moisture conditions.
- 27 -

Fig. 3.5 Infiltration capacity curves showing starting


points for each antecedent moisture condition
(after Wenzel and Terstriep, 1976)
In order to use the standard infiltration curves it is necessary
to evaluate the antecedent moisture conditions actually
prevailing at the time of a particular storrn~ An arbitrary
selection of antecedent moisture conditions (AMC) which were
used in the original testing of ILLUDAS is shown in Table 3.1.

Table 3.1 : Antecedent moisture conditions

Total rainfall during


ILLUDAS 5 days-preceding storm
number Description (mm)
1 Completely dry 0
2 Rather dry 0 to 12,5
3 Rather \vet 12,5 to 25
4 Saturated over 25
,
- 28 -

3.7 COMBINING OF HYDROGRAPHS

The paved and grassed area hydrographs determined for each sub-
catchment are combined to become an inlet hydrograph to the
drainage system. If the sub-catchment is at the uppermost end of
a series of pipes or open channels, the inlet hydrograph is
entered into the system by routing it downstream to the next
input point. If the sub-catchment occurs somewhere below the
upper end, its inlet hydrograph is combined with the upstream
hydrograph and the resulting combined hydrograph is routed
downstream to the next input point. If the sub-catchment is
located at the confluence of two or more pipes, the inlet
hydrograph is combined with the converging hydrographs before
routing downstream.

3.8 ROUTING PROCEDURE

A simple routing technique is used to transfer hydrographs from


one input point to the next. Two routing options are available:

(I) a lag (or time shift) of the entire hydrograph


without storage considerations, and
(2) storage routing using an implicit solution of
the continuity equation.

The first is the more economical of the two from the point of
view of computing time and is adequate in many instances. The
second is based on a sounder theoretical approach. Neither
method considers non-uniform flow, dynamic effects or the effect
of increased head due to surcharge.

Lag routing is performed by computing the velocity of the inflow


hydrograph peak (after discharge limiting due to pipe surcharging
and/or user-supplied discharge limiting for both pipes and
channels). This velocity is used to determine the time lag to be
applied to the inflow hydrograph.

In the implicit method lag routing is used to obtain an initial


estimate of outflows. The continuity equation is solved using a
linear approximation:
- 29 -

where 11 and 01 are the initial inflow and outflow; 12 and 02 are
the final inflow and outflow; CAREA is the cross-sectional flow-
area function in which the kinematic wave assumption that Sf=So
and Manning's equation are used, where Sf is the friction slope
and S is the reach slope; and L is the reach length. Eqs. (3.3)
and (3.4) can be combined to yield:

Eq. (3.6) must be solved for 02 and thus requires an implicit


solution.

The flow-area function (CAREA) is determined for each cross-


section by dimensionless flow-area versus dimensionless dis-
charge tabulated within the program. This non-dimensional
function is computed for the channel geometry. For the case of
trapezoidal open channels, a Newton-Raphson implicit solution
of the Manning's equation must be performed to develop this
dimensionless function. Parabolic interpolation between two-
percent intervals of the dimensionless area is performed by a
function sub-routine.

Eq. (3.6) is solved by an accelerated bracketing technique by


obtaining successive estimates of 02 by the equation
- 30 -

where 021 is the new estimate of 02 (the solution) and 020 is


the initial estimate; also

Eq. (3.8) is the zeroed continuity equation from eq. (3.6); F


equals zero at the solution for 02, and α is the acceleration/
deceleration factor.

In the implicit solution of eq. (3.6) certain conditions can


cause unreasonable solutions; these are corrected by the pro-
gram, for example as follows:

1) When 02 is greater than just-full pipe flow, 02


is set equal to just-full pipe flow.
2) When 02 is less than zero, 02 is set equal to
zero, with a subsequent change in routed hydrograph
volume.

If the computed travel time is less than the user-specified


time increment, the inflow hydrograph is interpolated at time
increments equal to the time of travel of the peak of the in-
flow hydrograph through the reach. This interpolation elimi-
nates instabilities in the hydrologic routing and obviates
"discharge lumping". After routing, the outflow hydrograph is
back-interpolated to the user-specified time increment
for subsequent computations by the program.

3.9 FLOOD CONTROL PONDS

One of the most promising methods of reducing urban drainage


costs embodies the use of storage ponds. Temporary detention of
runoff not only significantly reduces the cost of downstream
stormwater pipes and channels but also arrests sediment and
debris, thus helping to keep streams clear.

Depending on their function and location in the drainage system,


ponds are classified as either retention or detention, onstream
- 31 -

or offstream. Retention ponds differ from detention ponds in


having dead storage which is commonly used for recreation.
Onstream storage is where an embankment or wall is built across
the channel in contrast to off stream storage where the channel
overflows into storage during high discharge. A typical
onstream pond is illustrated in Fig. 3.6.

Fig. 3.6 Schematic diagram of a typical onstream


storage pond

ILLUDAS assists in the design of these ponds by providing


estimates of

(i) the volume of detention storage required to


ensure flows do not exceed pipe capacities or
user-specified discharge limits;

(ii) the outlet capacity to make effective use of


available storage.
The program operates with the storage equation:

∆s = I - 0 . . . . . . . . . . . . . . . .(3.9)

where ∆s = change in storage

I = inflow

O = outflow
- 32 -

The outflow rate is assumed to be constant and the storage


requirement for (i), above, is determined explicitly from eq.
(3.9). The outlet capacity in (ii) is determined by a trial and
error procedure using increasing estimates of outflow discharge.

The adequacy of the constant outflow assumption is dependent


on the predominant flow type. Orifice flow is proportional
to the square root of head, causing only moderate changes in
discharge as the pond fills. Weir flow is proportional to head
raised to the power of 1,5 (or 2,5 for a V-notch) causing
discharge to vary greatly with head. The constant outflow
assumption would be realistic for the former flow type but
not for the latter.

An orifice acts as a weir until it is submerged. Before sub-


mergence discharges are less than capacity and the constant
outflow assumption will underestimate storage requirement as
illustrated in Fig. 3.7. The significance of this error depends
on the proportion of pond volume filled before orifice flow
occurs. In SOIne cases it may be appropriate to assume this
volume to be dead storage.

Fig. 3.7 Hydrographs for storage pond computations


- 33 -

CHAPTER 4 LOCAL VERIFICATION

4.1 GENERAL

In order to ascertain whether reasonable simulations could be


achieved on local catchments using the ISWS recommendations for
parameter estimation (Terstriep and Stall, 1974) a study was
conducted using local data. Data for one catchment in Pinetown,
Natal, were obtained from the National Institute for Water
Research, Durban. Further data were obtained from
instrumentation by the author of a catchment in Johannesburg.

4.2 PINETOWN CATCHMENT

The Pinetown catchment (Fig. 4.1) is situated in the shopping


centre of Pinetown, which is approximately 20 km inland from
Durban. The catchment is monitored by the National Institute
for Water Research, Durban.

The catchment boundaries and the stormwater drainage system


are shown in Fig. 4.2. The total area is 11,9 ha of which
9,0 ha (75%) is directly-connected impervious surface, com-
prising roads, sidewalks, car parks, office blocks and shopping
complexes. The remaining area comprises lawns, unpaved parking
areas and small buildings that discharge onto pervious areas.
The ground slopes are moderately steep (up to 5%} and
approximate ground level contours are shown in Fig. 4.2. The
soils are sandy.

Rainfall was measured by two Casella siphon recorders, one


located within the catchment and the other immediately outside
the boundary near the outfall (Fig. 4.2). Water level was
measured in the outfall pipe by a Wesmar ultrasonic level
detector. The rainfall and runoff data at the outfall were
recorded on a punched tape. The raingauge within the catchment
recorded rainfall depth on a weekly drum chart and was used to
correct rainfall recorded at the outfall, the mean total depth
being used.
- 34 -
- 35 -

Fig. 4.2 Plan of the Pinetown catchment

Fig. 4.3 Discretization of the Pinetown catchment


- 36 -

The stage:discharge relationship was obtained by salt dilution


gauging (Simpson et al, 1980). A curve based on uniform flow
assumptions was fitted to the data and used in subsequent
calculations (Fig. 4.4).

Fig. 4.4 Rating curve for depth measurements in


the outfall pipe of the Pinetown catchment
- 37 -

For purposes of modelling the catchment was subdivided by


the author into 18 subcatchments as shown in Fig. 4.3.
The estimated subcatchment parameters are given in Table 4.1.
The pervious area depression storage was assumed to be 5 mm, as
recommended by ISWS, and the soils were classified as type B.
The impervious area depression storage was reduced from the
recommended 2,5 mm to 1,2 mm in order to improve agreement
between computed and observed hydrographs.

Pipe slopes were unavailable except at the outfall, and had to


be assumed on the basis of ground slope. The reach data used
for modelling are given in Table 4.2. A Manning roughness
coefficient, n, of 0,012 was assumed for all reaches. Time
steps of one and two minutes were used for simulation.

The three largest recorded storms were chosen to test the


model. A summary of the results is given in Table 4.3 and
observed and computed hydrographs are compared in Figs. 4.5,
4.6 and 4.7. Tabulations of the recorded data are given in
Appendix D.

The results are excellent considering that all parameters,


except paved area depression storage, were estimated. Adjustment
of the parameters would improve the results. In particular, an
increase in the subcatchment entry times and a decrease in
directly connected paved area would improve results by lagging
and attenuating the simulated hydrograph and by decreasing
runoff volume.

Due to the relatively steep catchment slopes pipe routing had


an insignificant influence on the computed hydrographs.
On this basis the assumed pipe slopes are considered adequate.
For larger discharges, however, further data would be required
to determine when the system surcharges.
- 38 -
Table 4.1 : Pinetown sub catchment data

Paved area Grassed area


% Supplemen-
Subcatchment % Paved % Grassed
Reach tary paved
area (ha) area Length Slope Entry time area Length Slope Entry time
area
(m) (%) (min) (m) (%) (min)

1 – 0 0,71 100 140 3 ,0 3,7


1 – 1 0,66 100 120 3,0 3,5
1 – 2 0,56 100 80 3,0 3,0
2 – 0 0,70 100 120 2,0 3,8
1 – 3 0,00
3 – 0 1,42 70 5 210 2,5 4,9 25 70 2,5 35,6
3 – 1 0,82 95 80 2,5 3,1 5 25 2,5 23,4
3 – 2 0,54 15 40 2,5 2,5 85 90 4,5 30,S
3 – 3 0,85 50 100 2,0 3,5 50 40 5,0 23,0
4 – 0 0,70 25 20 40 0,5 3,2 55 100 3,0 36,5
4 - 1 0,49 25 20 80 1,0 3,7 55 80 1,0 47,7
4 – 2 0,72 100 130 1,5 4;3
1 – 4 0,00
1 – 5 1,02 60 180 1,0 5,9 40 12 2,0 22,2
1 – 6 0,34 100 110 2,0 3,7
5 – 0 0,83 90 180 3,0 4,2 10 10 1,0 23,4
1 - 7 0,39 100 110 2,5 3,5
6 - 0 0,25 100 105 4,0 ].1
6 - 1 0,54 65 20 80 3,0 3,0 15 35 4,0 22,9
1 - 8 0,39 100 90 2,0 3,4

Table 4.2 : Pinetown reach data

Travel
Length Slope Diameter Height Width Lateral Discharge
Reach Section* time
(m) (%) (mm) (m) (m) slope capacity (m3/s)
(min)

1 – 0 50 3,5 C 457 0,6 0,2


1 – 1 70 2,5 C 457 0,5 0,4
1 – 2 40 2,5 C 457 0,5 0,2
2 – 0 40 2,0 C 457 0,5 0,2
1 – 3 90 1,2 C 610 0,8 0,6
3 – 0 60 3,0 C 450 0,5 0,3
3 – 1 75 3,0 C 610 1,2 0,3
3 – 2 60 1,8 C 610 0,9 0,3
3 – 3 40 1,8 C 610 0,9 0,2
4 – 0 50 2,2 C 375 0,3 0,3
4 – 1 70 1,8 C 381 0,3 0,5
4 – 2 30 1,5 C 381 0,2 0,2
1 – 4 90 2,9 C 610 1,2 0,4
1 – 5 45 2,9 C 610 1,2 0,2
1 – 6 45 3,2 C 610 1,2 0,2
5 – 0 90 4,0 T 0,3 0,5 0,1 5,2 0,3
1 - 7 45 3,2 C 610 1,2 0,2
6 - 0 40 2,5 C 300 0,2 0,3
6 - 1 70 2,5 C 457 0,5 0,4
1 - 8 45 3,73 C 762 2,4 0,1

* C = circular; T = trapezoidal
- 39 -
- 40 -
- 41 -

Fig. 4.6 Comparison of computed with observed hydrograph for


the storm of 29.9.79 on the Pinetown catchment

Fig. 4.7 Comparison of computed with observed hydrograph for


the storm of 4.11.79 on the Pinetown catchment
- 42 -

4.3 KEW CATCHMENT

The Kew catchment (Fig. 4.8), situated in the northern suburbs


of Johannesburg, has an area of 143 ha. The ground slopes are
moderately steep (up to 8%) and soils are residual granodiorite.
Although mainly residential, a significant part of the area is
occupied by industrial and commercial development. The
distribution of land use is shown in Fig. 4.9. The residential
sector occupies eighty percent of the area, industrial ten
percent, commercial five percent and open areas five percent.
The contrasting land uses are illustrated in Fig. 4.10 and 4.11.

The houses are generally on 0,2 ha lots. Most roofs drain on to


the gardens which are not graded and slope according to the
natural contours. Roads are paved and most have concrete kerbing.
The drainage system, consisting of concrete pipes, channels and a
natural stream, is shown in Fig. 4.9.

Rainfall and runoff were measured continuously from January to


May 1980. The rainfall recorder was a W. Lambrecht type 1509-20
with a 31-day strip chart propelled at 20 mm/h and recording
depth of rainfall at a scale of 1:0,125. Discharge was obtained
from stage measurements at a V-form Crump weir (Fig. 4.12)
placed in a culvert. Stage was measured by means of an Ott
pneumatic water level transducer and recorded by an Ott R20
strip chart recorder, with a 32-day chart propelled at 20 mm/h
and recording stage at a scale of 1:5.

The relationship between stage and discharge was affected by


large quantities of silt and debris washed down during storms.
Velocity-area measurements, recorded on two occasions, indicated
that the height of the crest above the upstream bed was
effectively zero. Crump weir rating curves are shown in Fig.
4.13(a) for no siltation and (b) for the case of zero upstream
depth. Also shown are the values obtained from velocity-area
measurements. It is evident that the latter rating would depart
from the measurements because of high approach velocities. A
means of extrapolation, therefore, had to be sought. For a given
discharge stage can be established if the relationship of depth
to critical depth is known. Critical depth can be
- 43 -

calculated for a given discharge for the channel geometry in


the vicinity of the control. Extrapolation based on a ratio
of depth to critical depth of 0,97 was found to fit the observed
data reasonably well. The extrapolated curve is also shown in Fig.
4.13.and the equations are summarised in Table 4.4.

Table 4.4 : Stage-discharge relationships

* (Ackers et al, 1978)


where Q = discharge (m3/s)
g = acceleration due to gravity (9,81 m/s2)
B = channel breadth (m)
Pv = difference between highest and lowest crest
levels (m)
H = total head (m)
= h + Q2
2
2gB (h+P)2
h = height of water surface above the crest (m)
P = upstream depth of channel bed below the crest (m)
hc = critical depth (m), and assumed equal to h/O,97

The runoff records were screened for large events and ten were
selected for further analysis. Three consecutive events had
hydrographs truncated at a discharge well below the maximum
observed. This was probably due to a temporary pipe blockage and
the records of these events had to be discarded. The remaining
seven events are tabulated in Appendix E.
- 44 -

Fig. 4.8 Aerial photograph of the Kew catchment


(Photo: Aircraft Operating Co., Jan. 1980)
Fig. 4.9 Kew catchment land use and stormwater
drainage system
- 46 -

Fig. 4.10 A view of the residential sector of


the Kew catchment

Fig. 4.11 A view of an industrial area in


the Kew catchment
Fig. 4.12 The V-form Crump weir at the outfall of
the Kew catchment (dimensions shown are in mm)

Fig. 4.13 Rating curve for the weir at the Kew


catchment
- 48 -

For the purpose of modelling, the catchment was subdivided


into 24 subcatchments as shown in Fig. 4.14. The subcatchment
parameters, shown in Table 4.5, were estimated largely from an
areal photograph and a topographic map. The subcatchment bound-
aries and the proportion of houses draining directly into the
streets were confirmed by a site inspection. The soils in the
industrial areas were generally well compacted and classified
as type D. For the remainder of the catchment the soils were
classified as type B. A figure of 5,0 mm was employed for
grassed area depression storage as recommended by the program
developers. Paved area depression storage was adjusted to
agree with observations and a value of 1,0 mm was used for all
except one storm. For the storm on the 15.3.80 rainfall in
the antecedent three hours was considered to have filled de-
pression storage. A one-minute time step was used for all
simulations.

Difficulties were experienced in distinguishing between contri-


buting and non-contributing grassed area. There seemed in fact to
be no clear distinction. In America where gardens are" generally
graded, one can consider the backyard to be noncontributing
because of the longer flow path. To avoid subjectivity in
estimating the contributing area the computations for the grassed
area hydrograph were modified, viz. losses (infiltration and
depression storage) were subtracted from the computed depth of
water on the surface. The original program computed excess
rainfall which it routed over the surface without loss. The
modification allowed all grassed areas to be considered as
contributing.

The verification results are summarised in Table 4.7, and com-


puted hydrographs are compared with observed hydrographs in
Figs. 4.15 to 4.21. The mean ratio of computed peak discharge to
observed is 1,09 with a standard deviation of 0,35. Exclusion of
three poorly simulated peaks (two on the 15.3.80 and the low
peak on the 17.3.80) lowers the mean to 1,02 and the standard
deviation to 0,13. The discrepancies for these events are most
likely due to spatial non-uniformity of the rainfall.
Fig. 1.14 Discretisation of Kew catchment
- 50 -
- 51 -
- 52 -
- 53 -

Runoff volumes were generally underestimated. The mean ratio


of predicted volume to observed was 0,83 with a standard
deviation of 0,13. Underestimation occurred mainly on the re-
cession limb of the hydrograph and could be due to subsurface
flow or runoff from some less pervious areas such as
driveways. The error is, however, not critical as far as
design calculations are concerned.

No checks could be made on the timing of peak discharge due


to difficulties in correlating times on the separate
recording charts for rainfall and runoff. The rain chart was
prone to slipping after adjustment, thus causing rainfall
records to lag behind those for runoff. When comparing
computed with observed hydrog~aphs time shifts were applied
to the computed hydrographs as indicated in the figures.

Only two storms showed significant runoff from grassed areas.


Hydrographs simulated with the new loss routine are compared
with those simulated by the original program in Fig. 4.15 and
4.16. It was assumed that all the grassed area was
contributing. The new routine is markedly better for one event
(18.3.80). Comparable results could, however, be achieved using
the original program with a calibrated contributing grassed
area.

The results are not as good as for the Pinetown catchment


partly due" to the increased complexity of the Kew catchment
and difficulties with estimation of the parameters. The
enlarged errors, however, seem to be due to spatial non-
uniformity of rainfall. The presence of this type of sampling
error is emphasised by the results for the storm of 15.3.80
for which no amount of adjusting of parameters could explain
the observed hydrograph.
- 54 -

Fig. 4.15 Comparison of computed with observed hydrograph


for the storm of 18.3.80 on the Kew catchment
- 55 -

Fig. 4.16 Comparison of computed with observed hydrograph


for the storm of 22.3.80 on the Kew catchment
- 56 -

Fig. 4.17 Comparison of computed with observed hydrograph


for the storm of 15.3.80 on the Kew catchment

Fig. 4.18 Comparison of computed with observed hydrograph


for the storm of 17.3.30 on the Kew catchment
- 57 -

Fig. 4.19 Comparison of computed with observed hydrograph


for the storm of 19.2.80 on the Kew catchment

Fig. 4.20 Comparison of computed with observed hydrograph


for the storm of 19.3.80 on the Kew catchment
- 58 -

Fig. 4.21 Comparison of computed with observed hydrograph


for the storm of 10.4.80 on the Kew catchment

4.4 DISCUSSION

The simulation results are highly satisfactory, considering


that all parameters except paved area depression storage were
estimated. The average ratio of computed to observed peak discharge
for both catchments is 1,08 (16 events) with a standard deviation
of 0,31. If at Kew three peak events that seem to be affected by
spatial non-uniformity of rainfall are neglected (i.e. two peaks on
15.3.80 and the first peak on 17.3.80), the ratio would be 1,04
with a standard deviation of 0,17. The calibrated depression
storage, though reasonably significant for the events considered,
would generally not be significant for design.

For the catchments and rainfall events considered the parameters


recommended by ISWS for entry time, grassed area depression
storage and infiltration capacity proved to be adequate. The
paved area depression storage (2,5 mm) was too large, and
a value closer to 1,0 mm was more appropriate. The routine
for determining grassed area losses was found to be inadequate
for one of the events. A more logical accounting for losses,
viz. subtracting from flow depth instead of from rainfall, was
found to markedly improve the simulation results.
- 59 -

CHAPTER 5 STORM INPUTS FOR DESIGN

5.1 GENERAL

It is a relatively straightforward task to predict runoff for a


specific storm. In practice, however, one is usually concerned
with the prediction of peak discharge or runoff volume with a
certain frequency of exceedance. Frequency analysis of a long
record of flows is the most rewarding approach, but for small
catchments adequate flow data are seldom available for this type
of analysis. In urban oatchments there is the added difficulty
of land use change which can invalidate a simple flow frequency
analysis. One must resort to other methods of frequency
estimation.

One approach is to simulate flows from a continuous record of


rainfall and perform a frequency analysis on the simulated
flows. The simulation model can be adjusted to cater for land
use change and if flow data are available they can be effec-
tively used to calibrate the model. For urban catchments it
is quite feasible to simplify the rainfall analysis by approxi-
mating the continuous record with a set of discrete events.
High costs and lack of readily available rainfall data, however,
make this approach unpopular.

A simpler approach, which is commonly used in practice, is to use


a design storm. From an analysis of rainfall records a
representative temporal distribution is developed and total storm
rainfall is varied with recurrence interval. The simulated flood
is assumed to have the same frequency of exceedance as that of
the rainfall. The storm duration is often chosen on the basis of
maximising the simulated flood. Implicit in this approach is the
assumption that the variation in total rainfall is more
significant than the variation in antecedent soil moisture (AMC)
or the variation in temporal distribution of rainfall. The
assumption is more likely to hold good for urban catchments than
for their rural counterparts because of their lower sensitivity
to AMC.
- 60 -

Many different types of design storm are currently in use.


Some require a trial-and-error approach to determine the critical
storm duration while others do not. The choice of a design storm
for local use depends mainly on the availability of data and its
ability to approximate flood flows.

The median first quartile distribution developed by Huff (1967)


was recommended by Terstriep and Stall (1974) for use with ILLUDAS
in Illinois, USA. This distribution is based on an analysis of US
rainfall data. Applicability to local conditions has not been
confirmed. Although a one-hour storm duration is suggested
(Terstriep and Stall, 1974) there is some doubt as to its
appropriateness. For a conservative design, the critical duration
should be determined by trial and error.

HRU Report 1/72 provides two techniques for synthesizing design


storms. One method (Fig. C8, HRU 1/72) gives mass plots of
lower envelopes of percentage rainfall versus percentage duration
for storms of various durations. This is based on an analysis of
hourly precipitation data (Wiederhold, 1969). The shorter critical
times for urban runoff makes these curves inapplicable to urban
catchments. The other method presented (Fig. C5, HRU 1/72) considers
only the most intense part of the storm and gives no guidance as to
the storm duration and quantity of rainfall antecedent to the peak
intensity. Though valuable in an analysis of probable maximum floods
this method is not readily applicable to the prediction of flood
frequencies.

Schulze (1979) has recommended the use of the US Soil Conser-


vation Service (SCS) design hyetographs for use on local rural
catchments. These hyetographs are based on US intensity--
duration-frequency (IDF) relationships and, although Schulze
indicates these to be reasonably applicable to South Africa a
similar but simpler approach is possible. The suggested approach
makes use of readily available IDF data and was reported by
Keifer and Chu (1957) for stormwater drainage design in the City
of Chicago, USA.

This synthetic distribution has had fairly widespread usage and has
become known as the Chicago design storm. It can read-
- 61 -

ily incorporate local data and application is simple. The


purpose of this chapter is to present the theory of this tech-
nique, to evaluate its adequacy and to provide a basis for its
local use.

5.2 THE CHICAGO DESIGN STORM

The Chicago storm is based on IDF curves and the distribution is


such that for any time interval the maximum average intensity is
equal to that from the IDF curves. This means that when one
applies the storm to a catchment the critical intensity for all
sub-catchments is used and the necessity of determining the
critical duration for each part of the catchment is eliminated.
The position of the peak intensity within the storm is based on
local storm characteristics.

Using an IDF equation of the form:

where I is the average rainfall intensity for duration t, and


a, b and c are parameters dependent upon the locality and de-
sign frequency, the equation for the Chicago design storm can
be derived as follows:

This is the equation for an advanced storm pattern, i.e. the


peak occurs at the beginning of the storm. If the peak occurs at
some later time, then the storm can be described by considering
the duration, t, as being composed of a time tb before, and a
time ta after the peak, i.e.
- 62 -

Fig. 5.1 The Chicago design storm


- 63 -

t = tb + ta

Now if r is the ratio of the time-to-peak, tp, to the total


duration of the storm, td, then

Substituting for t from eqs (5.4) and (5.5) in eq. (5.3) gives the
following relationships for intensities before and after the peak:

To use the Chicago design storm in ILLUDAS it is necessary to


reduce the storm-hyetograph to a set of discrete values. This
can be done as follows:

(i) Select the time step ∆t (eg. 5 minutes).

(ii) Compute the discrete point representing the peak rain-


fall from the equation:

(iii) Distribute the time interval selected (∆t) around the


peak as r∆t before the peak and (l-r) ∆t after the peak.

(iv) Compute the points before and after the peak by inte-
grating the design curve and calculating the discrete
- 64 -

intensity ordinate from the volumes for each increment


of t.

The general integral form of the hyetograph before the peak is


given by:

and after the peak by:

5.3 EVALUATION OF THE CHICAGO STORM

The Chicago storm has received the following criticism based


on theoretical considerations:
(i) It assumes the runoff frequency to be the same as the
A rainfall frequency.

(i) It is more peaked than real storms, i.e. a greater


proportion of rainfall occurs at peak intensity than
for real storms.

(ii) It makes incorrect use of IDF relationships. Maximum


intensities do not occur for all durations in real
storms.

(i) is a limitation of all design storms and (ii) and (iii)


supposedly result in overestimation of peak discharge.

To assess the adequacy of the Chicago storm for design calcula-


tions a comparison with flood frequency data is necessary. Syn-
thetic flood frequency data can be generated through real event
- 65 -

simulation. Marsalek (1978) used this approach and compared flood


frequency predictions for an actual catchment and several
hypothetical catchments patterned after typical urban develop-
ments in Canada. The comparisons indicated that the Chicago storm
overestimated peak discharges by approximately 75% and was
unsatisfactory for sizing of storage ponds.

Clarke and Bishop (J.F. MacLaren Ltd, 1979) conducted a similar


study, comparing simulation results for three interconnected
catchments in Edmonton, Canada. The Chicago storm predictions
agreed closely with those based on real storms. A similar study
was conducted in Winnipeg, Canada, (J.F. MacLaren Ltd, 1977) and
again it was concluded that the theoretical storms resulted in
runoff frequencies similar to those of the historical storms.

A further study was conducted by the author to assess the perfor-


mance of the method with local data. From rainfall data provided by
the Johannesburg City Engineer's Department for the Norwood rain
gauge, 54 events were selected and discretized at 5-minute
intervals from an effective record of 16 years. These events had
average intensities for durations of 5,,15, 30, 60 and 120 minutes
greater than arbitrarily chosen base levels and are listed in
Appendix F. The storms were then ranked according to maximum
average intensity for each duration. The top 16 storms for each
duration were selected for the analysis, making 28 storms in all.

In an IDF analysis the log-Gumbel frequency relationship was


found to fit the data adequately. An equation of the following
form was fitted to the log-Gumbel frequency estimates for re-
currence intervals between 2 and 100 years

where a, γ, b and c are coefficients and I is the maximum average


rainfall intensity (mm/h) for duration, t (minutes), and recurrence
interval, T (years). The coefficients obtained from this analysis
were:
- 66 -

a = 3930
γ = 0,12
b = 21
c = 1,1

The resulting equation gave predictions which differed from


those of the log-Gumbel analysis by less than 7%.

The time-to-peak ratio was determined in the manner described


in section 5.4. A 2-hour duration storm was chosen and a
corresponding r value of 0,28 resulted.

Antecedent moisture conditions (AMC) were established using the


ISWS recommendations outlined in chapter 3. An average AMC value
of 3 was used in conjunction with the Chicago storms.

Peak discharges were simulated using the models described in


chapter 4 for Pinetown and Kew. Design mode of the program was
used so that the program resized pipes such that they would be
adequate to accommodate the flows. The predicted peak
discharges thus represent the values one would adopt for design
if one did not want the system to be surcharged. Soil types
were varied for the Kew catchment to establish whether this
would influence results. Computations were conducted at a one-
minute time step, though rainfall data was averaged over 5-
minute increments.

Peak discharge predictions based on real and Chicago storms are


compared in Fig. 5.2 for the Pinetown catchment. A method-of-
moments fit to the peaks based on real storms is shown together
with 95% confidence bands, assuming normality of sampling
error (Adamson, 1978). The Chicago storm predictions agree
closely with those based on historical data.

Figs. 5.3 to 5.6 compare predictions of peak discharge from the


Kew catchment for different assumed soil types. The agreement
between predictions based on real and synthetic storms is not as
good as for the Pinetown catchment. They are nevertheless
satisfactory as indicated by the 95% confidence bands
- 67 -

embracing the predictions based on real storms. The greater


deviation of predictions is due to the high proportion of
grassed area runoff which makes discharge sensitive to assumed
AMC. The general under-prediction of the Chicago storm is
contrary to normal expectations but can be accounted for by the
occurrence of high AMC associated with some:of the more severe
real storms.

To assess whether all critical real storms had been considered,


flows were simulated for Kew (soil type C) for the 26 storms
eliminated earlier. The results were identical except for an
insignificant increase in predicted peak discharge for recurrence
intervals less than 1,3 years.

Fig. 5.2 Comparison of peak discharge estimates


for the Pinetown catchment
- 68 -

Fig. 5.3 Comparison of peak discharge estimates


for the Kew catchment assuming soils to
be type A

Fig. 5.4 Comparison of peak discharge estimates


for the Kew catchment assuming soils to
be type B
- 69 -

Fig. 5.5 Comparison of peak discharge estimates


for the Kew catchment assuming soils to
be type C

Fig. 5.6 Comparison of peak discharge estimates


for the Kew catchment assuming soils to
be type D
- 70 -

A comparison of synthetic with real storms shows the significance


of the assumed AMC. In Figs. 5.7 and 5.8 two real storms causing
maximum peak discharge at Kew are compared with the 20-year
Chicago storm that gave a similar peak discharge. The storm of
11.1.76 shown in Fig. 5.7 is very similar to the Chicago storm,
while that of 27.9.77 in Fig. 5.8 is much less peaked. The high
AMC value (=4) on 27.9.77 accounts for high peak discharge
prediction.

The sensitivity of the simplified pond sizing routine in ILLUDAS


to storm profiles was also tested. A similar analysis to
establish sensitivity of storage requirement was conducted for
the Kew catchment, assuming soils of type B; further tests were
omitted because of the similarity in the results for
peak discharge.

Peak discharge at the outfall was to be restricted by pondage


first to 4 m3/s and then to 8 m3/s. A frequency analysis of
storage requirement was carried out for real storms and com-
pared with values derived by Chicago predictions. Results are
shown in Figs. 5.9 and 5.10. Also shown is the Chicago
prediction based on the next higher AMC.

From the results it is evident that predictions of storage


requirement are highly sensitive to assumed AMC and the best
the Chicago storm can do in this situation is predict upper
and lower bounds.
- 71 -

Fig. 5.7 Comparison of storm recorded at Norwood on


11.1.76 with the 20-year Chicago design storm

Fig. 5.8 Comparison of storm recorded at Norwood on


27.9.77 with the 20-year Chicago design storm
- 72 -

Fig. 5.9 Comparison of storage requirement estimates


for an outflow rate of 4m3/s
- 73 -

Fig. 5.10 Comparison of storage requirement estimates


for an outflow rate of 8m3/s
- 74 -

5.4 ESTIMATION OF PARAMETERS

Many organizations have accumulated IDF relationships from which


the coefficients a, band c in eq. (5.1) can be evaluated by
regression analysis. For instance by rewriting eq. (5.1) as

and with an assumed value of c one can solve for the constants
(a1/c) and (-b) by linear regression using (I-l/c) as the abscissa
and t as the ordinate. This is repeated using a different value
of c until an acceptable fit has been achieved. Alternatively eq.
(5.1) can be written as

lnI = ln a - c ln (t+b) . . . . . . . . . (5.12)

In this case b is estimated, then the equation is solved for the


constants (-c) and (ln a), using ln(t+b) as abscissa and ln I as
ordinate.

Values of c usually fall between 0,5 and 1,5 and values of b


between 0 and 30 minutes. Pocket calculator programs for per-
forming this analysis are given by Watson (1980).

In the absence of local IDF relationships, the coefficients given


by Midgley and Pitman (1978) can be used. These coefficients, in
units compatible with those used in this work, are given in Table
5.1. The values of a are determined from Fig. 4 of HRU 2/78 by
rewriting eg. (5.1) as

For t equal to 60 minutes, the values of γ are given in Table 5.1.


- 75 -

Table 5.1 : Regional intensity-duration-frequency coefficients

Region b c γ60
Inland 14,4 0,883 44,9
Coastal 12,6 0,737 23,5

Alternatively, the following approximation has been found


satisfactory for mean annual precipitation (MAP) less than 1 000
mm:

where γR is a regional coefficient equal to 241 for


inland regions and to 84 for coastal regions
MAP = mean annual precipitation (mm)
T = recurrence interval (years)

For MAP greater than 1 000 mm, eq. (5.14) gives results differing by
up to 20% from those derived from the curves in Fig. 4 of
HRU 2/78. The maximum difference for MAP less than 1 000 mm
is 6%.

The time-to-peak ratio, r, determines the depletion of initial


losses and the depth of surface and pipe flow prior to the peak
intensity. The ratio can be determined from an analysis of local
storm distributions and depends on the depth of rainfall that can be
expected antecedent to the peak intensity. It can be calculated as
follows:

where d = depth of rainfall (mm), antecedent to the peak a


intensity, of duration t (minutes)
td = total storm duration (minutes)
It = maximum average intensity (mm/h) for duration t
Itd = average intensity (mm/h) for total storm duration
- 76 -

From eq. (5.1)

Eq. (5.16) is employed to determine r from the average ratio of


(da/dt) obtained from an analysis of significant storms. The
IDF coefficients band c should be those corresponding to the
recurrence interval of the average maximum depth dt.

Until local data have been analysed the time-to-peak ratios


obtained in other localities can provide guidance. Table 5.2
gives r values obtained in four overseas studies and from the
local analysis discussed in section 5.3.

Table 5.2 : Time-to-peak ratios


(td = 3 hours)

City Country r Reference


Chicago USA 0,386 Keifer and Chu, 1957
Winnipeg Canada 0,31 J.F. MacLaren Ltd, 1974
Burlington Canada 0,46 M.M. Dillon Ltdr 1977
Gauhati India 0,37 Bandyopadhyay, 1972
Norwood (Jhb) RSA 0,22 Section 5.3

Coogan (1951) analysed two Johannesburg rainfall records for a


period of 42 years and found that 64% of the severe storms had
- 77 -

their 30-minute peak intensities occurring during the first half


of the storm. Smith (1951) analysed 164 severe storms in Durban
and found that although the peak occurred randomly within the
storm, there was a tendency for more severe storms to peak
early.

From the above data it seems that values for r will generally be
less than 0,5. Higher values will probably be more appropriate for
coastal regions and lower values for inland regions. However, due
to the limited data it is suggested that a value of 0,4 be used
for all localities until further data become available.

In some cases computations will not be highly sensitive to


choice of r. An analysis of variations in peak discharge from
the Kew catchment for changes in the value of r showed a maximum
difference of 16% when r was varied from 0,4 to zero and then to
0,8.

The storm duration will depend on the longest time of concentration


to be considered for design and should be slightly longer than this.
Once a duration has been chosen it should be used for all areas.
Durations of 2 or 3 hours will generally be adequate.

5.5 DISCUSSION OF RESULTS

Design storms are merely approximations of actual rainfall and


cannot be expected to perform adequately in all situations.
In fact, in different situations different design storms may be
appropriate. The limited extent of testing so far performed
shows the Chicago storm to be an adequate technique for
predicting peak discharge for the range of rainfall data and
catchment characteristics considered. There seems to be no
reason why it should not be adequate for a wider range of
situations but further testing is recommended when data become
available.

For detention pond sizing the great importance of AMC must be


recognised. It is suggested that the use of the Chicago storm
- 78 -

with varying AMC might be used to generate envelopes. These


may, however, turn out to be very wide and serve merely to
emphasise the sensitivity of the results to a parameter that
cannot be accurately estimated.

Further analyses to evaluate storm parameters and AMC are


needed. In the absence of local data the regional parameters
provided in section 5.5 offer a reasonable basis for design.
- 82 -

CHAPTER 7 CONCLUSIONS

Techniques presently used in South Africa for urban flood pre-


diction are inadequate in many respects. The commonly used
techniques are unverified and make inadequate use of readily
available catchment and rainfall data. Parameter estimation, too,
cannot be other than rather subjective. ILLUDAS is a considerable
improvement in all respects and yet is not much more complicated
to apply than the commonly used Rational Method. In fact when
alternative design options have to be evaluated it can prove to
be a much simpler and more efficient method (eg. evaluation of
the use of detention ponds or testing of alternative design
frequencies require changes to only one parameter).

ILLUDAS is a logical model of the runoff process. Published


verifications as well as the local tests reported here show the
model to be adequate for simulation of urban runoff. Much of the
discrepancy between recorded and computed hydrographs can be
attributed to sampling and data errors. Results show further that
the parameters can be estimated with reasonable accuracy.
Furthermore, because of the deterministic nature of the model,
parameter estimation can be improved as additional runoff data
become available. Uncertainties associated with parameter
estimation can be largely resolved by means of simple sensitivity
analyses which incidentally form a sound basis for design.

Although ILLUDAS can handle most stormwater drainage design


problems it can not deal with pressure flows associated with
pipe surcharging or backwater effects. Furthermore routing in
reaches with very flat slopes (eg. 0,05%) will be inaccurate.
Where these aspects are important a more sophisticated model
such as SWMM or HVM should be used (See Table 1.1).

Application of ILLUDAS to small rural catchments seems logical


but has not been thoroughly verified. Discretization will
possibly be more complicated than for urban catchments and some
adjustment will have to be made to enable the model to simulate
- 83 -

overbank flows which are usually significant in rural catch-


ments. Moreover runoff from rural catchments is highly sensitive
to antecedent moisture conditions and therefore a continuous
simulation model may be more appropriate.

A major problem associated with single event models is the


selection of storm inputs. Use of a set of historically re-
corded severe storms is the most rewarding approach but is
likely to be costly and time-consuming. Use of synthetic design
storms is an alternative approach and the Chicago storm has
been shown to give satisfactory predictions of peak discharge.

Satisfactory performance of the model with local data has been


demonstrated and this commends it as a valuable design tool. The
model has been metricated and improved for local use.
- 84 -

REFERENCES:

1. ACKERS, P., WHITE, W.R., PERKINS, J.A. and HARRISON,


A.J.M. Weirs and flumes for flow measurement. Wiley,
1978.

2. ADAMSON, P.T. The statistics of extreme values and the


analysis of floods in South Africa. Department of Water
Affairs, TR86, 1978, pp84.

3. AITKEN, A.P. The application of storage routing methods to


urban hydrology. J. Inst. Engrs. Australia, Jan - Feb. 1968.

4. AITKEN, A.P. Hydrologic investigation and design in


urban areas - a review. Australian Water Resources
Council Tech. paper No.5, Australian' 'Government Pub-
lishing Service, Canberra, 1973.

5. AITKEN, A.P. Hydrologic investigation and design of urban


stormwater drainage systems. Australian Water Resources
Council Tech. Paper No. 10, Australian Government
Publishing Service, Canberra, 1975.

6. American Society of Civil Engineers. Hydrology Handbook.


ASCE - Manuals of Engineering practice - No. 28, Jan.
1949, pp 48-49.

7. ARDIS, C.V., DUEKER, K.J. and LENZ, A.J. Storm drainage


practice of thirty-two cities. J. Hyd. Div., Amer. Soc.
Civ. Engrs., Vol. 95, No. HYl, Jan. 1969, pp 383-408.

8. BANDYOPADHYAY, M. Synthetic storm pattern and runoff for


Gauhati, India. J. Hyd. Div., Amer. Soc. Civ. Engrs., Vol.
98, No. HY5, May 1972, pp 845-857.

9. BRANDSTETTER, A., FIELD, R. and TORNO, H.C. Evaluation of


mathematical models for the simulation of time varying
runoff and water quality in storm and combined
- 85 -

sewerage systems. Conf. on environmental modeling, U.S.


Environmental Protection Agency, Cincinnati, Ohio, USA,
April 1976.

10. CLARK, C.O. Storage and the unit hydrograph. Proc.


Amer. Soc. Civ. Engrs, Vol. 69, Nov. 1943.

11. COLYER, P.J. and PETHICK, R.W. Storm drainage design


methods - a literature review. INT 154, 3rd impression,
Hydraulics Research Station, Wallingford, UK, Dec. 1976.

12. COOGAN, J.M. Stormwater drainage in urban areas with


particular reference to Pretoria. J. Instn. Mun. Engrs. S.
Afr., Yolo I, No.4, Jan. 1951, pp 143-162.

13. CORDERY, I. Some effects of urbanisation on streams.


Trans. Instn. Engrs., Australia, 1976, pp 7-11.

14. M.M. DILLION LTD. Storm drainage criteria manual for the
city of Burlington. Ontario, Canada, April 1977.

15. FORD, W.G. The adaptation of the RRL hydrograph method for
tropical conditions. Proc. Nairobi flood hydrology
symposium, Oct. 1975, Transport and Road Research La-
boratory SR259, pp 409-455.

16. HEEPS, D.P. and MEIN, R.G. An independent evaluation of


three urban stormwater models. Research Report No.4, Dept.
Civ. Eng., Monash University, 1973.

17. HEEPS, D.P. and MEIN, R.G. Independent comparison of


three urban runoff models. J. Hydr. Div., Amer. Soc.
Civ. Engrs, Vol. 100, No. HY7, July 1974, pp 995-1009.

18. HICKS, W.I. A method of computing urban runoff. Trans.


Amer. Soc. Civ. Engrs., Vol. 109, 1944, pp 1217-1253.

19. HUFF, F.A. Time distribution of rainfall in heavy storms.


Water Resources Research, Vol. 3, No.4, 1967, pp 1007-
1019.
- 86 -

20. HYDROLOGICAL RESEARCH UNIT. Design flood determination


in South Africa. Report No. 1/72, Univ. of the Witwatersrand,
1972.

21. ILLINOIS STATE WATER SURVEY. Routing options for


ILLUDAS update - fa l l ~ 19? 9. Urbana, Illinois, USA"
1979.

22. IZZARD, C.F. Hydraulics of runoff from developed surfaces.


Proc. 26th Annual Meeting Highway Research Board, Vol.
26, 1946, pp 129-146.

23. KEIFER, C.J. and CHU, H.H. Synthetic storm pattern for
drainage-design. J. Hyd. Div., Amer. Soc. Civ. Engrs.,
Vol. 83, No. HY4, Aug. 1957, paper no. 1332.

24. J.F. MACLAREN LTD. Drainage criteria manual for the


city of Winnipeg~ Nov. 1974.

25. J.F. MACLAREN LTD. Review of Canadian design practice and


comparison of urban hydrological models. Research Report
No. 26, Ministry of the Environment, Ontario, Canada,
Oct. 1975.

26. J.F. MACLAREN LTD. Report on design storm selection.


City of Winnipeg, Canada, 1977.

27. J.F. MACLAREN LTD. Edmonton rainfall analysis; a com-


parison of historical and theoretical design storms for the
city of Edmonton. Canada, May 1979.

28. MARSALEK, J. Research on the design storm concept.


Amer. Soc. Civ. Engrs., Urban Water Resources Research
Program, Tech. Memo. No. 33, Sept. 1978.

29. McPHERSON, M.B. Some notes on the Rational Method of storm


drain design. Amer. Soc. Civ. Engrs., Urban Water
Resources Research Program, Tech. Memo. No.6, 1969.
- 87 -

30. MIDGLEY, D.C. and PITMAN, W.V. A depth-duration-frequency


diagram for point rainfall in Southern Africa. Univ. of the
Witwatersrand, Hydrological Research Unit, Report No.
2/78, Aug. 1978.

31. MUSGRAVE, G.W. and HOLTON, H.N. Infiltration. Section 12,


Handbook of Applied Hydrology, Ed. Ven te Chow, 1964, pp 12-
22 - 12-25.

32. PAPADAKIS, C.N. and PREUL, H.C. Testing of methods for


determination of urban runoff. J. Hyd. Div., Amer. Soc.
Civ. Engrs., Vol. 99, No. HY9, Sept. 1973, pp 1319-1335.

33. PATRY, G., RAYMOND, L. and MARCHI, G. Description and


application of an interactive mini-computer version of the
ILLUDAS model. Proc. SWMM Users' Group Meeting, May
1979, pp 242-274.

34. RAGAN, R.M. and DURU, J.O. Kinematic wave nomograph for
times of concentration. J. Hyd. Div., Amer. Soc. Civ.
Engrs., Vol. 98, No. HY10, Oct. 1972, pp 1765-1771.

35. READ, A.L. Hydraulic aspects of the West Lakes Development.


Instn. Engrs. Australia, Hydrology Papers, Adelaide
Symposium, 1971, pp15-25.

36. SCHAAKE, J.C., GEYER, J.C. and KNAPP, J.W. Experimental


examination of the Rational Method. J. Hyd. Div., Amer. Soc.
Civ. Engrs., Vol. 93, No. HY6, Nov. 1967, pp 353-370.

37. SCHULZE, R.E. and ARNOLD, H. Estimation of volume and rate


of runoff in small catchments in South Africa based on the
SCS Technique. Univ. of Natal, Agricultural Catchments
Research Unit Report No.8, 1979.

38. SIMPSON, D.E., STONE, V.C. and HEMENS, J. Water pollution


aspects of stormwater runoff from a commercial land-use
catchment in Pinetown, Natal. Inst. Wat. Poll. Control
(Southern African Branch), Pretoria, June 1980.
- 88 -

39. SMITH, H.A. Discussion of 'Stormwater drainage in Urban


areas with particular reference to Pretoria' by Coogan J.M.
J. Instn. Mun. Engrs. S. Afr., Vol. 1, No.4,
Jan. 1951, pp 156-157.

40. STALL, J.B. and TERSTRIEP, M.L. Storm sewer design - an


evaluation of the RRL method. US Environmental Protection
Agency, Tech. Series, EPA-R2-72-068, Oct. 1972.

41. TERSTRIEP, M.L. and STALL, J.B. Urban runoff by Road


Research Laboratory method. J. Hyd. Div., Amer. Soc. Civ.
Engrs., Vol. 95, No. HY6, Nov. 1969, pp 1809-1834.

42. TERSTRIEP, M.L. and STALL, J.B. The Illinois Urban


Drainage Area Simulator, ILLUDAS. Illinois State Water
Survey, Urbana, Bulletin 58, 1974.

43. TERSTRIEP, M.L., BENDER, G.M. and BENOIT, D.J. Buildup,


strength, and washoff of urban pollutants. Amer. Soc. Civ.
Engrs., Convention and Exposition, Chicago, Oct. 1978,
ASCE preprint 3439, 29 pp.

44. WATKINS, L.H. The design of urban sewer systems. Road


Research Technical Paper No. 55, HMSO; 1962.

45. WATKINS, L.H. and YOUNG, C.P. Developments in urban


hydrology in Great Britain. Road Research Laboratory, LN
885, Crowthorne, U.K., 1965 (unpublished).

46. WATSON, M.D. Sizing of urban flood control ponds. (In


press), 1980.

47. WATSON, M.D. and MILES, L.C. Stormwater drainage practice


in South Africa. (In press), 1980.

48. WENZEL, H.G. and TERSTRIEP, M.L. Sensitivity analysis of


selected ILLUDAS parameters. Illinois State Water Survey,
Contract Report 178, 1976.

49. WIEDERHOLD, J.F.A. Design storm determination in South


- 89 -

Africa. Hydrological Research Unit Report No. 1/69,


Univ. of the Witwatersrand, 1969.

50. WILLIAMS, D.W., CAMERON, R.J. and EVANS, G.P. TRRL and
Unit hydro graph simulations compared with measurements in
an urban catchment. J. Hydrol., Vol. 48, No. 1/2, Aug.
1980, pp 63-70.

51. WISNER, P., CHEUNG, B. and GUPTA, S. Introduction to


quantity modelling. Short course on Storm Water Management
for runoff control in new developments, Univ. of Ottawa,
Oct. 1979.

52. WOOLHISER, D.A. Simulation of unsteady overland flow.


Unsteady flow in open channels, Vol. II, Ed. by Mahmood
K. and Yevjevich V., Water Resources Publications, USA,
1975, pp 502.

You might also like