Sunteți pe pagina 1din 201

Water Hammer Simulations

WITPRESS
WIT Press publishes leading books in Science and Technology.
Visit our website for the current list of titles.
www.witpress.com

WITeLibrary
Home of the Transactions of the Wessex Institute, the WIT electronic-library
provides the international scientific community with immediate and permanent
access to individual papers presented at WIT conferences.
Visit the WIT eLibrary athttp://library.witpress.com
This page intentionally left blank
Water Hammer
Simulations

S. Mambretti
Universidade Estadual de Campinas, Brazil
S. Mambretti
Universidade Estadual de Campinas, Brazil

Published by
WIT Press
Ashurst Lodge, Ashurst, Southampton, SO40 7AA, UK
Tel: 44 (0) 238 029 3223; Fax: 44 (0) 238 029 2853
E-Mail: witpress@witpress.com
http://www.witpress.com

For USA, Canada and Mexico

WIT Press
25 Bridge Street, Billerica, MA 01821, USA
Tel: 978 667 5841; Fax: 978 667 7582
E-Mail: infousa@witpress.com
http://www.witpress.com

British Library Cataloguing-in-Publication Data


A Catalogue record for this book is available
from the British Library

ISBN: 978-1-84564-680-6
eISBN: 978-1-84564-681-3

Library of Congress Catalog Card Number: 2013938170

No responsibility is assumed by the Publisher, the Editors and Authors for any
injury and/or damage to persons or property as a matter of products liability,
negligence or otherwise, or from any use or operation of any methods, products,
instructions or ideas contained in the material herein. The Publisher does not
necessarily endorse the ideas held, or views expressed by the Editors or Authors of
the material contained in its publications.

WIT Press 2014

Printed by Lightning Source, UK.

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,
mechanical, photocopying, recording, or otherwise, without the prior written
permission of the Publisher.
Preface

Waterhammer is a pressure surge or wave caused when a fluid (usually a


liquid but sometimes also a gas) in motion is forced to change its velocity.
It commonly occurs when a valve closes or opens at the end of a pipeline
system, or a pump starts or stops; as a consequence, a pressure wave
propagates in the pipe.

This pressure wave can cause problems, and normally the efforts of the
designers are aimed at reducing its effects. However, in some cases the
pressure pulses are deliberately caused in order to achieve particular results,
such as the pumping of a fluid or the mapping of a network.

In all cases, the study of the consequences of this phenomenon is required,


and this might be performed with rough formulas as those developed
between the end of 19th Century and the beginning of the 20th, or through
more advanced models. Nevertheless, all the models have limitations and
might be criticized.

During the years I spent as a professional consultant and researcher in the


field, the need for a reference able to guide through the several models
that have been developed and the analysis of their results clearly emerged.
Hence the reason to write this book.

In the book, chapter 2 presents the governing equations and the hypotheses
involved; chapter 3 shows the simplified solutions that have been developed
and used before the arrival of computers; in chapters 4 and 5 two numerical
integration methods are shown. In chapter 6 a number of devices are
presented with the methods that can be used for their modeling. In chapter 7,
problems related to the implementation of the model are discussed, giving
some suggestions for their solution. In chapter 8 an important phenomenon,
the presence of air and cavitation, is analyzed, while in chapter 9 the most
advanced models are presented and discussed. Chapter 10 presents some
actual hydraulic plants whose behavior has been analyzed and studied with
the methods and models presented in the book.

Writing a book is always a challenging task, and many people contributed


in some way to making this volume a reality.

All the case studies have been carried out with the invaluable help of
my colleague and friend Dr. Paola Pianta, with whom I discussed many
developments and applications of the models presented in this book. Most
of my research has been carried out at Politecnico di Milano, where all
the experimental tests have been performed, thanks to Prof. Enrico Orsi,
head of the Laboratory of Hydraulics, and Prof. Enrico Larcan, head of the
Department of Hydraulics. After moving to Brazil, invited by Prof. Paulo
Barbosa, head of the Faculty of Civil Engineering, I found an excellent
welcome and all the freedom and help I needed to finish the book, thanks
to the kindness of Prof. Jos Geraldo Pena de Andrade, head of the School
of Technology, University of Campinas. I cannot forget Prof. Daniele De
Wrachien, who has always been a support and a help in my academic life,
and Prof. Carlos Brebbia, who not only is the Director of WIT, but he is also
an example to be followed both from the scientific and the human point of
view. Finally, I have to thank my wife Grazia, both because she allowed me
to work overtime and because she drew all the figures for the book.

All the mistakes that can be found in the book have to be attributed to my
limitations. I would be grateful if the Readers would report any errors and
suggest any improvements for the subsequent editions of the book.

Stefano Mambretti
So Paulo, 2013
Contents

1 An old topic still not completely solved 1

1.1 An old topic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Applications and problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 This book . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 The computer programs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Compressible flow theory: basic concepts 9

2.1 Instantaneous operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9


2.2 Wave celerity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Velocity of operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Non-negligible headlosses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . 17
2.5.1 Continuity equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5.2 Momentum equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.5.3 The governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3 Simplified solutions 25

3.1 The Allievis method (1913) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25


3.2 The non-elastic hypothesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.1 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.2.2 Comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.3 Graphical method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4 Numerical solution of the governing equations:
The method of characteristics 41

4.1 Numerical solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41


4.2 Initial and boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.2.1 Reservoir . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
4.2.2 Valve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.2.3 Junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . 47
4.3 The computer code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.4 First simple application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

5 Numerical solution of the governing equations:


finite difference methods 55

5.1 The CourantFriedrichsLevy stability condition . . . . . . . . . . . . . . . . . . 57


5.2 The LaxWendroff method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.3 Solving the governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.4 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4.1 Asymmetrical schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.4.2 Ghost cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.5 The computer code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.6 Again the simple application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

6 Devices Boundary conditions 71

6.1 Surge tanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71


6.1.1 Simple surge tanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.1.2 Different types of surge tanks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.2 Air chambers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.3 Relief valves and rupture disks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.4 Centrifugal pumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.5 Other methods for controlling the pressures . . . . . . . . . . . . . . . . . . . . . . . 84
6.6 The computer codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.6.1 The simple surge tank . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
6.6.2 The simple air chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
6.6.3 Air chamber with headlosses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
6.6.4 Air chamber and valve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
6.6.5 Valve modelling: an example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
6.6.6 Pumps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

7 Instabilities 101

7.1 Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101


7.1.1 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
7.1.2 Computer program for oscillating velocity . . . . . . . . . . . . . . . . 103
7.2 Transfer matrix method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.2.1 General remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
7.2.2 Application to hydraulic transients . . . . . . . . . . . . . . . . . . . . . . . 106
7.2.3 Description of simple system: pipes . . . . . . . . . . . . . . . . .. . . . . 108
7.2.4 Description of simple system: valves and effluxes . . . . . . . . . 109
7.2.5 Global matrix of a system . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 111
7.2.6 A simple application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
7.3 Numerical instabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.3.1 Changing CFL number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
7.3.2 First order methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
7.3.3 Flux-limiters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
7.3.4 Artificial dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 122

8 Effects of air and cavitation 127

8.1 Cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127


8.1.1 Formation of the cavities . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 127
8.1.2 Collapse of the cavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
8.1.3 Description of the motion in the presence of cavities . .. . . . . 130
8.2 Changing of celerity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
8.3 Water column separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.4 Additional resistance terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
8.4.1 Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
8.4.2 Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.5 Laboratory experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.5.1 Experimental set-up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.5.2 Experimental tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.6 Computer code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

9 Advanced models 145

9.1 2D models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145


9.1.1 Continuity equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
9.1.2 Momentum equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
9.2 Headlossess . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
9.2.1 Pezzingas model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 147
9.2.2 k model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 149
9.3 Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 150
9.3.1 Release gaseous cavitation model . . . . . . . . . . . . . . . . .. . . . . 150
9.3.2 Second viscosity cavitation model . . . . . . . . . . . . . . . . . . . . . 152
9.4 Numerical schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
9.5 Further problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154

10 Case studies 157

10.1 Simple pressure pipe for petroleum products in Djibouti . . . . . . . . . 157


10.1.1 Plant characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
10.1.2 Expected scenarios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
10.1.3 Case 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
10.1.4 Case 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
10.1.5 Case 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
10.1.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
10.2 A more complex example for seawater treatment plant in
Tanzania . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
10.2.1 Plant characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
10.2.2 Unsteady flow simulations: existing plant . . . . . . . . . .. . . . . 162
10.2.3 Plant to be designed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
10.2.4 No air chambers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
10.2.5 Air chamber 3 m3 volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
10.2.6 Air chamber 5 m3 volume . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
10.2.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
10.3 A very complex example for seawater treatment plant in
Algeria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
10.3.1 The plant to be modelled . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 169
10.3.2 A peculiar device: energy recovery PX . . . . . . . . . . . . . . . . . 172
10.3.3 Laboratory plant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
10.3.4 Laboratory tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . 173
10.3.5 Model of the seawater plant in Algeria . . . . . . . . . . . . . . . . . 175
10.3.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
10.4 Final remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

References 181
This page intentionally left blank
Chapter 1
An old topic still not completely solved

1.1 An old topic

The high pressures generated by a sudden variation of the velocity in closed pipes
have been studied for centuries. Italian books of the eighteenth century describe
the problem of percossa idraulica (hydraulic stroke) even if without being able
to quantify its value.
The pressures generated gave the idea that they might be used to produce work:
in 1772 John Whitehurst, in the United Kingdom, invented a manually controlled
precursor of the hydraulic ram called the pulsation engine, which in its first instal-
lation raised water to a height of 4.9 m. The inventor did not patent his idea and the
details are obscure. The first self-acting ram pump was invented by the Frenchman
Joseph Michel Montgolfier (best known as a co-inventor of the hot air balloon) in
1796.

The principle is quite simple, although ingenious; a simplified hydraulic ram


is shown in Figure 1.1. The device has to be positioned where fluent water is
available, as the ram will use its energy. Initially, the waste valve (3) is open,
and the delivery valve (4) is closed. The water in the drive pipe (1) starts to
flow under the force of gravity and picks up speed and kinetic energy until the
increasing drag force closes the waste valve. The momentum of the water flow in
the supply pipe against the now closed waste valve causes a water hammer that
raises the pressure in the pump, opens the delivery valve (4), and forces some
water to flow into the delivery pipe (2). Because this water is being forced uphill
through the delivery pipe farther than it is falling downhill from the source, the
flow slows; when the flow reverses, the delivery check valve closes. If the water
flow stops, the loaded waste valve reopens against the now static head, which
allows the process to begin again. A pressure vessel (5) containing air cushions
the hydraulic pressure shock when the waste valve closes, and it also improves
the pumping efficiency by allowing a more constant flow through the delivery
pipe. Although, in theory, the pump could work without it, the efficiency would
drop drastically and the pump would be subject to extraordinary stresses that
could shorten its life considerably. One problem is that the pressurized air will
gradually dissolve into the water until none remains. One solution to this prob-
lem is to have the air separated from the water by an elastic diaphragm (similar
to an expansion tank); however, this solution can be problematic in developing
2 WATER HAMMER SIMULATIONS

Air (5)

(2)
Outlet-delivery
Weight
pipe

(3)
(1)
Waste
valve

v
(4) Waste
Delivery valve
check valve A

Figure 1.1: Hydraulic ram.

Countries where replacements are difficult to procure. Another solution is to


have a mechanism such as a snifting valve that automatically inserts a small
bubble of air with each pump cycle. Yet another solution is to insert an inner
tube of a car or bicycle tyre into the pressure vessel with some air in it and the
valve closed. This tube is in effect the same as the diaphragm, but it is imple-
mented with more widely available materials. The air in the tube cushions the
shock of the water the same as the air in other configurations does.
The efficiency of this device is obviously quite low, but it has the advantage to
be independent from any external source of energy, using only that of the inflowing
water, and for this reason it is still used in small plants installed to provide water in
villages in the developing countries.
The hydraulic ram was developed on the basis of empirical observations, as
analytical evaluations of the pressures during transients have been carried out at
the end of the nineteenth century. The first equations which provided a quantitative
assessment of the pressures were produced by Menabrea in 1858 [1] and Michaud
in 1878 [2]; afterwards, Joukowsky in 1898 [3], and, independently, Allievi in
1903 [4] completed Michauds work, correcting his results and developing a more
comprehensive theory. The paper of Joukowsky was the first that introduced the term
waterhammer, which later on is used worldwide to refer only to the elastic model,
while the phenomenon related to the rigid column model, i.e., mass oscillations,
took the name surge. Both the Italian Colpo dAriete and French terms Coup
de Blier mean literally Stroke of a Ram, and probably they have beed derived
from the above-described device hydraulic ram.
These theories allowed to compute only the maximum value obtained at the
beginning of the transients; thenAllievi in 1913 [5] was able to compute the different
phases, under simplifying assumptions, with the so-called chained equations.
AN OLD TOPIC STILL NOT COMPLETELY SOLVED 3

Evangelisti, in the sixties of twentieth century [6, 7], developed a numerical


method based on the characteristics, which is still used worldwide and implemented
in computer programs.
In the thirties of the same century graphical methods appeared [8, 9] that allowed,
still under simplifying hypotheses, to compute the value of the pressures during the
development of the transient.
Since then, a really large number of books and papers have appeared on the
scientific literature, and among them the works of Streeter and Wylie [10, 11] and
Chaudry [12] have to be highlighted.
Moreover, many computer programs are already available to study this
phenomenon.

1.2 Applications and problems


The phenomenon is typically destructive, but some authors [1316] tried to use the
impulsive wave generated during transients to perform network analysis studying
the subsequent reflections and in particular to detect the position of the leakages in
the water supply systems. As the pressure impulse is reflected by any variation in
diameter, material or in presence of a leakage, the analysis of the response to that
impulse can show the dissimilarities between the expected and the real networks.
As mere evaluation of the position of the leak is not sufficient to evaluate the
efficiency of the system, this method has also been developed in order to allow the
determination of the amount of the leakage. These authors agree on the following:

An higher leakage implies a greater decrease of the pressure;


The variation of the pressure does not significantly depend on the shape of the
orifice or on the outflow conditions;
The friction plays a significant role on the deformation of the pressure wave.

In Figure 1.2, a very simple network is shown, where water is provided from
reservoir A and discharged to the reservoirs B and C; the leakage is positioned
25 m upstream the reservoir B (position L) and the discharge is abruptly inter-
rupted at the valve V, where the pressures are also recorded during the transients.
Actually, the analysis has been performed only with numerical methods and not
on physical models or real cases. The comparison between the responses with
and without the leakage is reported in Figure 1.3. As the celerity of the wave
has been imposed equal to 1000 m/s, the first reflection is recorded after 0.18 s
because of the junction J; for the second reflection, the two cases with and with-
out leakage differ: if there are no leakages, the reflection is recorded after 0.28 s
while, with the leakage, it is recorded after 0.23 s.

The subsequent oscillations could be followed, but they become confuse and lose
significance. Moreover, the recorded reflection at 0.23 s means that at a distance
equal to 115 m from the valve there is a dissimilarity from the expected network,
4 WATER HAMMER SIMULATIONS

m
70
A V B
J L
90 m 25 m 25 m

Figure 1.2: Simple network: losses are positioned 25 m upstream the reservoir B.

625

No leakage
620 Leakage

615
Head (m asl)

610

605

600

595
1.8 2 2.2 2.4 2.6 2.8 3
Time (s)

Figure 1.3: Response of the simple network comparison between the responses
with and without leakages.

but, for instance, this piece of information alone does not allow to detect if this
dissimilarity is located on the pipe JC or on the pipe JB. In depth analysis is
required for a better understanding of the actual system.

This analysis is surely valid and effective in the localization and assessment of
small leaks, but it is sometimes difficult to apply in real cases since it is difficult
to estimate the initial conditions and the values of some parameters of the model.
Moreover, the reverse procedure that is used to evaluate the extent of the leakages is
often burdensome. Finally, this is surely an interesting field of research in the years
to come, especially considering the importance of reducing the losses from the
water supply networks, both economically and ecologically. However, the methods
of modelling the networks and analysing its response need a great improvement.
This is not the only field where this phenomenon needs further research: in
fact, despite the efforts made to explain the phenomena involved in the transients,
many aspects are still to be clarified, and some of them will be discussed in this
AN OLD TOPIC STILL NOT COMPLETELY SOLVED 5

book. For instance, another important problem still unresolved is represented by


headlosses: in the simplified models that have been developed during the twentieth
century these have often been neglected; this assumption is surely on the safe
side, but does not allow to follow the evolution of the different phases during the
transient. Even more complex models available nowadays normally evaluate the
headlosses with the expressions carried out for steady flow; as it will be shown
in the appropriate chapters of this book, the use of steady flow models does not
allow a good description of the unsteady flow phenomena. Therefore, many authors
dedicated their time in studying the best methods to improve the models; however,
when air is mixed with water, and especially when cavitation occurs, the behaviour
of the mixture is very difficult to predict.
Finally, as it has been mentioned that the phenomenon is destructive, methods
and devices to reduce its effects have been studied and modelled. For instance,
surge tanks are very large structures designed since the end of nineteenth century
to protect the first hydropower plants, and their dimension together with their costs
and their strategic importance justified a very accurate design. However, as will
be shown in this book, it is very difficult to accurately model the behaviour of
even the simplest devices, and some of them can even produce instabilities that
worsen the condition of the plant where they are installed. Pumps, on the other
hand, are particular devices that generate unsteady flow and are affected by the
additional pressures obtained during transients; their behaviour during transients is
very difficult to be reproduced. Again, for the interested researchers there is a lot
of work to do.

1.3 This book

As the literature on this topic is very extensive, before starting this job the question
was whether a new book was really newsworthy and what it would have added to
the general knowledge in this area.
The main target of this book is providing the methods for modelling transients
in closed pipes to the reader. When is a computer model, i.e., a deep analysis nec-
essary? ASCE [17, 18] developed guidelines to identify pipelines at risk; however,
Ellis [19] correctly stated that caution is exercised in all cases where some form of
study is carried out, even if not a detailed computer analysis.
In the first chapters the old formulae are shown, because they are normally used
for first evaluations, as a rule of thumb; moreover, some ancient methods are also
presented to give an idea of the simplifications they required and the differences that
can be expected using more advanced models. The governing equations are also
derived, because an expert engineer should not use the computer programs as simple
tools, always remember in a computer the old adage trash in trash out applies;
therefore, the assumptions underlying the equations should be known as well.
This book is also devoted to the computer model development. To help the reader
to build a model, a methodology is shown. To this end, not all the devices that can
be found to reduce the effects of waterhammer have been described; instead, are
6 WATER HAMMER SIMULATIONS

described, the most widely used devices and especially the assumptions and the
methods to implement them in a computer program in addition to the problems
found in their implementation and the reliability of the carried out results. Also,
methods available to integrate the governing equations are nowadays in a large
number: in the book only the method of characteristics and one based on finite
differences (LaxWendroff) are shown. However, the methodologies shown in this
book can be used as a guide in case the reader is interested in the implementation
of a different scheme.
In most cases, when a method or a model is described, it is also implemented in
a computer program, enclosed with the book: this means seven computer programs
are enclosed. The programs developed are not meant to be used for professional
purposes: they have been tested and all the attention has been placed in their devel-
opment, and they can be used for educational purposes or even for small real plants.
However, these programs are very simple and do not allow to model complex plants,
for which the reader should turn to commercial codes. Anyway, the presence of a
running computer program shows that all the methods described have been tested
and they work.
Both the executable file and the source codes are enclosed, therefore the pro-
grams can be used as they are, in order to reproduce the examples presented in the
book or to try new cases; but the main purpose is for the reader to study the list
code to see how the different methods have been implemented, learn the technique
and, if possible, improve the outcome.

1.4 The computer programs


The computer programs enclosed with the book have been developed in Delphi [20],
which is an object-oriented Pascal. Most of the code is devoted to input/output of
data and for managing the hydraulic engine which is constituted by:

setting initial conditions;


starting the loop and call alternatively the procedures to solve:
internal points;
boundary conditions.

The procedures to compute internal points and boundary conditions will be


discussed in each case through the book, while the initial conditions are normally
those of steady flow conditions.
The steady flow conditions are computed using the DarcyWeisbach for-
mula, i.e.:
V |V |
J = (1.1)
2gD
where J is the distributed headlosses per unit length (or slope of the hydraulic
grade line), is a dimensionless coefficient called the Darcy friction factor, V is
the flow velocity, D is the pipe diameter and g is the acceleration of gravity. As
AN OLD TOPIC STILL NOT COMPLETELY SOLVED 7

normally the roughness of the conduit is expressed with parameters as Mannings


or Hazen-Willimans, in the programs developed for the book it has been decided to
let the User introduce the former, which is easily found in all engineering practice
manuals (e.g., Stephenson [21], who also compare the different formulae available
in the literature).
The Mannings formula is:
1 2/3 1/2
V = R J (1.2)
n
where n is the Mannings coefficient dependent on the pipe characteristics, and R is
the hydraulic radius.
As the pipe and flow characteristics are expressed in diameter D and discharge
Q, in the computer programs the headlosses J are computed rearranging eqn (1.2)
and obtaining:
n2 Q 2
J = 10.29 16/3 (1.3)
D
Combining eqns (1.1) and (1.3), the values of can be computed from the
Mannings coefficient with the formula:
n2
= 124.528 (1.4)
D1/3
All the units are expressed in SI. The use of practical formulae like Mannings
implicitly assume the flow is turbulent (while the formula of DarcyWeisbach can
be used in all flow regimes), but this hypothesis is strictly verified in most practical
cases, and however acceptable.
Obviously the application of the formula in the cases presented in the book is
trivial, as in the examples there will be one pipe only and because a reservoir is
always present (upstream or downstream), the head in each required point in the pipe
is computed starting from the reservoir and adding or subtracting the headlosses.
For complex networks, a nonlinear system has to be written and many methods have
been developed for its solution, which are not part of the topic of this book (for
instance see Ref. [22] for a solid handbook or Ref. [23] for a scientific approach
and a selected list of references).
Results are printed in two text files: one reports the heads and the other the
velocities computed during the simulations; these files can be imported for analysis
in any spreadsheet. Each of the file has data printed in seven columns: the first
column reports the time of simulation, the other six columns the required data
(heads or velocities) located at the upstream of the pipe (second column) and at
20%, 40%, 60%, 80% and 100% of the length of the pipe the last column is
obviously related to the downstream condition.
As the programs have been developed only for demonstration purposes, they
have no checks or alerts in case of overflows or errors occurred during simulations:
that mean they can abruptly crash without notice. On the other hand, they are
sufficiently light and fast to be used practically in all computers and require few
seconds to accomplish the average simulations.
This page intentionally left blank
Chapter 2
Compressible flow theory: basic concepts

Transients in pipes could be analyzed with two different approaches, depending on


the hypothesis related to the flow compressibility.
In the elastic theory approach (waterhammer) the liquid compressibility is
computed, while in the anelastic approach (surge) the liquid compressibility is
neglected.
Although in most of the hydraulic phenomena liquid compressibility may be
neglected, in the analysis of transients this is often a key parameter which cannot
be ignored, as its mistreatment would bring completely mistaken results, except in
very peculiar cases.
On the other hand, a model which take into account the liquid compressibility is
more general and obviously allows the analysis of the phenomena that could have
it ignored. Therefore, in this book the liquid is considered compressible, even if in
few cases comparisons with the anaelastic approach will be performed.

2.1 Instantaneous operations


Let us consider the simple sketch shown in Figure 2.1, where for the time being the
kinetic energy and the headlosses are neglected. The initial velocity in the pipe of
area A is equal to V0 and L is the length of the pipe.
At the upstream boundary there is a reservoir, which can be considered having a
constant head, while at downstream boundary there is a valve, open in steady flow
conditions and instantaneously closed at the beginning of the operations.
When the velocity is stopped at the downstream end, under the hypothesis of
uncompressible liquid, the whole column stops: the momentum of the water column
(a finite quantity) goes to zero in an infinitesimal time, and therefore the force
exerted by the liquid on the pipe is infinitive, which is neither true nor acceptable
from a designer point of view.
Thus, the hypothesis of uncompressible flow has to be discarded, accepting
its elasticity: in this case, when the valve closes, the liquid is still flowing in the
upstream part of the pipe, and, therefore, the quantity of water which actually
stops in the infinitesimal time dt is an infinitesimal volume of length ds, while the
remainder of the water column moves with the initial velocity.
As the stopped volume is equal to A ds, the stopped mass is equal to A ds,
and therefore, the variation of the momentum is equal to A ds V0 . This
10 WATER HAMMER SIMULATIONS

Upstream reservoir

Downstream valve

Figure 2.1: Sketch of a simple plant.

variation must be equilibrated from the impulse of the forces operating on the
same mass; these forces are given by the increase of the pressure p which is
found at the downstream boundary, and therefore the force is A p dt.
As a consequence, the equilibrium can be expressed by:

A ds V0 = A p dt (2.1)

and therefore the increase of pressure due to the abrupt closure of a downstream
valve is:
ds
p = V0 (2.2)
dt
Let c = ds/dt be the wave celerity1 , eqn (2.2) can be written as:

p = c V0 (2.3)

See Figure 2.2 for a better understanding of the event and of the adopted symbols.
The increment of pressure given by eqn (2.3) can be expressed in terms of water
column, as:
c
h = V0 (2.4)
g
Equation (2.3) or (2.4) is known as AllieviJoukowskys formula, even if the
literature shows Menabrea [1] to have made calculations with these equations.
In the following instants, the subsequent volumes get stopped. The pressure
wave propagates with celerity c towards the upstream reservoir, which is reached

1
In this book, the word velocity refers to the mass transportation, while the word celerity
refers to the velocity of propagation of a wave.
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 11

ds
V0
A

V
P0 0

A
c  ds
dt
P0 
p

Figure 2.2: Forces equilibrium sketch and symbols.

at time:

L
t1 = (2.5)
c

When t < t1 , the pipe is divided in two halves: the upstream part where the liquid
is still moving and the downstream part where the liquid is stopped and pressures
have increased of the quantity given by eqn (2.3) (Figure 2.3).
As mentioned, when the time is equal to t1 the pressure wave reaches the
upstream tank where the head is constant. As a consequence, at time t1 there is
a difference in the pressures of the liquid at the upstream end of the pipeline, while
the velocity of the liquid is null everywhere. Due to this pressure differences, water
starts to flow from the pipeline towards the tank, with velocity V0 , while pressures
are back to those of steady flow. This second phase lasts until time t2 = 2L/c when
the pressure wave reaches again the downstream valve. In the time t1 < t < t2 , the
pipeline is still divided in two halves: in the upstream part, the liquid is still moving
with velocity V0 and pressure is equal to that of steady flow; in the downstream
part, the liquid is stopped and there is high pressure.
At time t2 , which is the end of the second phase, the pressure wave reaches again
the downstream valve where obviously the velocity must be zero. With the same rea-
soning carried for the initial phase, but considering that now in the pipe the velocity
is equal to V0 , the wave reflection induces pressures equal to p = c V0 .
The third and fourth phases are very similar to the first and second, respectively,
but in this case the pressures have opposite sign.
At time t4 = 4L/c, the initial conditions are reached again and, therefore, the
cycle starts again. These four phases reproduce continuously and, theoretically,
never end; in the real world, the waves smooth because of the headlosses and
therefore the phenomenon ends in a finite time, which can be very short.
12 WATER HAMMER SIMULATIONS

h  c V0
g h  c V0
g

First phase Second phase

h  c V0
g h  c V0
g

Third phase Fourth phase

Figure 2.3: Four phases of transient.

2.2 Wave celerity


As it has been observed, the characteristics of the event depend substantially on the
wave celerity. In the following, the expression of the celerity is reported only in the
case of non-deformable pipes, while in the Section 2.5.1 this value will be carried
out for more general cases.
The bulk modulus of elasticity of the fluid is defined as:

dp
= (2.6)
d

Let us consider again Figure 2.2 which reproduce the conditions in a generic
time t in a pipe after the instantaneous closure of the downstream valve. In the sub-
sequent time dt the section that separates the stopped fluid (downstream) from that
in motion (upstream) moves upstream from AA , together with the pressure wave;
instead, the section AA itself moves downstream of a distance V0 dt because
the volume W between that section and the downstream valve is compressed; the
volume decreases of a value equal to:

dW = A V0 dt (2.7)
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 13

At the same time on the same volume W acts an increased pressure given by
eqn (2.3) to which, because of eqn (2.6), corresponds a decrement in volume:

W p A c V0
dW = = ds (2.8)

Equalling eqns (2.7) and (2.8) and reminding that c = ds/dt, it follows that:



c= (2.9)

This is the value of sound celerity in the fluid. For water with a temperature
equal to 8 C, this value is equal to c
= 1425 m/s, which is an increase of about
3 m/s for each increased degree of temperature.
If the liquid is more compressible, as it happens for certain oils, the celerity c
decreases.
The wave celerity also decreases when the elasticity of the pipeline wall is
considered. For a pipe with diameter D, thickness e and modulus of elasticity E,
eqn (2.9) becomes (see Section 2.5.1):




c=  (2.10)
D
1+
Ee

being a coefficient which considers the effects of the junctions between pipes and
can be assumed equal to 1 for a single pipe.
For steel pipes, a widespread assumption is c = 1000 m/s, while for plastic
pipes this value is lower because their elasticity is higher. As will be discussed in
Chapter 8, the presence of air may have a significant effect on this parameter.

Example 1
Let us compute the wave celerity in a steel pipe with diameter D = 500 mm
and thickness e = 6.3 mm, knowing that the steel elasticity modulus is equal
to E = 2.0 1011 N/m2 , the bulk elasticity modulus for water is equal to
= 2.14 109 N/m2 and water density is = 1000 kg/m3 . Wave celerity is equal to:
 
2.14 109
1000 1462.87
c=  = = = 1075 m/s
D 500 2.14 10 1.85
9
1+ 1+
eE 6.3 2.0 1011
14 WATER HAMMER SIMULATIONS

About the dimensions, it is to be noted that:


 
kg
  
m s2 m2 m
  = 2
kg s s

m3

while:

D
1+
eE
is non-dimensional.

2.3 Velocity of operations


Let us consider now, instead of an instantaneous closure of the valve, the case when
a valve is closed in a finite time Tc > 0.
For the sake of simplicity let us evaluate the velocity outflowing from the valve
with the following linear expression, which, however, can be easily generalized:


V (t) = V0 1 t when t Tc
Tc (2.11)


V (t) = 0 when t > Tc

In the first instants of the transient, downstream pressures can be computed


by the already-known formula p = c V . As the velocity linearly decreases
because of eqn (2.11), pressure linearly increases.
Let Tc the time when the valve is completely closed, for a given time t > Tc , in
the section where the valve is positioned the pressure due to the transient can still
be computed by eqn (2.3); the front of the wave which moves upstream is inclined,
as shown in Figure 2.4.
As already mentioned, when the wave reaches the upstream reservoir, it is
reflected downstream, where it propagates as descendant perturbation. If the pipe
is relatively short, the valve may close in such a long time that the closure is still
in progress when the reflected wave is back; this time 0 is called critical phase
duration and it is equal to:
2L
0 = (2.12)
c
where L is the length of the pipe. When Tc > 0 , the pressure in the section of the
valve cannot completely develop and the pressure in the section where the valve is
positioned is equal to:
p = c (V0 Vf ) (2.13)
where Vf is the velocity at the valve when t = 0 .
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 15

Tc

c
h  V
g 0

Figure 2.4: Pressure wave moving upstream when the downstream valve closes
linearly.
The maximum value of pressure that is reached in this case can be computed
considering Figure 2.5 and applying simple geometric considerations, obtaining:
c V0
g p
= (2.14)
Tc 2 L
c

Rearranging and simplifying the equation, the so-called Michauds formula [1]
is carried out:
2 L V0
p = (2.15)
Tc
The increment of pressure due to the transient can again be expressed as water
column and, therefore:
2 L V0
h = (2.16)
g Tc

Example 2
Let us consider a pipe with length L = 1000 m where the wave celerity is equal to
c = 1000 m/s and the flow has a velocity equal to V0 = 5 m/s. Pressures have to be
computed in both cases of valve closure with time Tc = 0.8 s and 5.0 s.
The phase duration of the system is = 2L/c = 2 s.
In the former case, we have 0 < Tc and, therefore, pressures have to be computed
with the AllieviJoukowski formula, obtaining:
p = c V0 = 1000 kg/m3 1000 m/s 5 m/s
= 5 106 kg/m2 m/s2 = 5 106 Pa
16 WATER HAMMER SIMULATIONS

cV0

2L Tc t
c

Figure 2.5: Computation of waterhammer pressure for long operations.

or, in terms of water column:

c 1000 m/s
h = V0 = 5 m/s
= 510 m
g 9.806 m/s2

In the latter case, instead, we have 0 > Tc and, therefore, pressures have to be
computed with the Michaud formula:

2 L V0 2 1000 kg/m3 1000 m 5 m/s


p = = = 2 106 Pa
Tc 5s

or, in terms of water column:

2 L V0 2 1000 m 5 m/s
h = = = 204 m
g Tc 9.806 m/s2 5 s

2.4 Non-negligible headlosses


Now let us consider a pipe where headlosses cannot be neglected, as in the case
shown in Figure 2.6.
In the time interval t1 , the mass 1 stops and the water column increases to
a quantity equal to c V0 /g. The increased head propagates to the mass 2, which,
however, has an initial (steady flow) pressure larger than the mass 1, the difference
being equal to J x. For this reason, the mass 2 is pressed by a downstream head
equal to cV0 /g J x which is obviously smaller than the value c V0 /g and
is necessary to stop mass 2 completely.
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 17

J x

c V0
g
Steady flow
Hydraulic grade line

3 2 1

Valve completely and


instantaneously closing

Figure 2.6: Pressures due to transients when headlosses cannot be neglected.

In other words, the pressure generated at the valve section is not sufficient to
stop the mass 2, which, therefore, has a residual velocity equal to V = J x g/c.
Because of this residual velocity, mass 2 compresses again the mass 1, further
increasing its pressure, until the mass 2 stops completely; this happens when for
that mass the increased head is equal to c V0 /g. The increased head on the mass
1 is, therefore, equal to c V0 /g + J x.
Repeating this procedure for the subsequent temporal steps (and analogously
for the following masses), it can be shown that the increased head at the downstream
valve is equal to c V0 /g, to be added to the static head.

2.5 Governing equations


The governing equations are the classic continuity and momentum; they are usu-
ally written in one dimension, along the pipe, and in this form they will be
derived. In the following, the flow will be characterized by its average velocity
V (x, t) = Q(x, t)/A(x, t) and pressure p(x, t) computed in the centre of mass of the
pipe cross section.

2.5.1 Continuity equation

The continuity equation has to be written considering that during transients the
pipeline wall deforms because of the pressure variations. This deformation is a
18 WATER HAMMER SIMULATIONS

V
V x

V u
x
u
x

Figure 2.7: Control volume for determining the continuity equation.

quite complex phenomenon, as it consists not only of radial shape variations, but
also elongations or shortening of the pipeline in the longitudinal direction.
The continuity equation will be carried out referring to a control volume con-
sidered fixed with respect to the pipe, which therefore follows all deformations.
The sketch of the phenomenon and the adopted symbols are reported in Figure 2.7.
Let u be the velocity of the pipe wall in the direction of the pipe axes; the mass
that enters the control volume is equal to A(V u), while the mass outflowing
downstream is (except for infinitesimal of higher order):

[ A (V u)]
A (V u) + x (2.17)
x

The difference between the masses inflowing and outflowing the control volume
must be equal to the increase of the mass inside the volume itself; therefore, still
neglecting infinitesimal of higher order:

[ A (V u)] D
x = ( A x) (2.18)
x Dt

where the symbol D /Dt denotes the total derivative2 with respect to the axial
motion of the pipe.

2
The total derivative of a function f with respect to the pipe motion, which has velocity u,
is f
t
+ u f
x
.
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 19
D (x) u
Developing and keeping into account that = x, it yields:
Dt x
( A V ) ( A)
+ =0 (2.19)
x t
Rearranging again and denoting with a dot positioned above the variable the
substantial derivative3 with respect to the average flow velocity, it yields:

A V
+ + =0 (2.20)
A x

So far the most general case has been handled without making any hypothesis
neither on the shape or the material of the pipe nor on the transported fluid. Now
few hypothesis must be introduced, which are quite common in the practical cases
of hydraulic engineering.

Hypothesis 1 (on the fluid). The fluid is Newtonian, i.e., denoted with the bulk
elasticity modulus, its rehology can be described by:

p
= (2.21)

Hypothesis 2 (on the pipe geometry). The pipe is cylindrical with circular base.
Let D be the specific extension of the diameter, it yields:

A
D
= 2 = 2 D (2.22)
A D
Hypothesis 3 (on the material of the pipe). The material that constitutes the pipe
is mechanically homogeneous and isotropic, with linear elastic behaviour and the
wall thickness e is negligible when compared with the diameter. Therefore, the
Hooks law applies for two dimensions stress states, i.e.:

C x
C = D = (2.23)
E

x C
x = (2.24)
E

where C is the specific extension of the wall in the radial direction, x is the
specific extension of the wall in the axial direction, C is the normal stress in the

3
The substantial derivative of a function f with respect to the average flow velocity V is
f
t
+ V f
x
.
20 WATER HAMMER SIMULATIONS

radial direction, x is the normal stress in the axial direction, is the Poissons
coefficient (0 0.5) and E is the bulk elasticity modulus of the material of the
pipe.
If the pressure is large enough, its variations on the generic section of the pipe
can be neglected, i.e., the pressure p can be assumed as constant along the wall of
the pipe and therefore the Mariottes4 formula applies:

pD
c = (2.25)
2e

The axial stress depends on the constraints to the pipe. If the pipe is free to
expand or reduce in the axial direction, then the axial stress is always zero (this is
the case of a pipe with many expansion joints) so that x = 0. On the contrary, if
the pipe is constrained so that ant axial deformation is prevented, it is obviously
x = 0.
Therefore, the expression (2.22) for the former case (axial stress equal to zero)
becomes:

 
A pD
= (2.26)
A Ee
while in the latter case (axial deformation equal to zero) it becomes:

A D pD  
= 1 2 (2.27)
A Dt Ee

Both eqns (2.26) and (2.27) can be written in the general case as:


 
A pD
= (2.28)
A Ee

where depends on the constraint conditions and varies within the range 0.75 and 1.
This parameter depends only on that is constant in space and time because of
the hypothesis on the pipe material. As mentioned, the wall thickness is negligible
if compared with the diameter, and because the stress can be considered two-
dimensional, the thickness e can also be considered constant, both in space and
time. As a consequence, eqn (2.28) can be simplified as:

A
= (p D + p D) (2.29)
A Ee

4
Mariotte [24] was able to show the strength of the weakest material element in a structure
is likely to decrease with increasing structure size.
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 21

= D A finally yields:
Considering D , substituting and extracting A/A
2 A
p D
A
= Ee (2.30)
A pD
1
2Ee
In the usual applications, the denominator on the right hand equation is very
close to one, which allows a drastic simplification of eqn (2.30)5 . This simplified
expression, substituted in the continuity equation (2.20) and also considering the
eqn (2.221) allows to obtain the following:

p D p V
+ + =0 (2.31)
Ee x
that can also be written as:
D
p 1 + E e V
+ =0 (2.32)
x

Keeping in account the celerity term defined in eqn (2.10) in Section 2.2,
eqn (2.32) becomes:
V
p + c2 =0 (2.33)
x
and finally, developing the substantial derivative, the expression of the continuity
equation is:
p p V
+V + c2 =0 (2.34)
t x x

2.5.2 Momentum equation

The momentum equation can be written balancing all the forces applied to a part
of the flow between two pipe sections that have distance x. In Figure 2.8, those
sections are shown with all the forces applied. The different forces are as follows.
Forces due to the pressures on the upstream and downstream areas are,
respectively:
m = p A (2.35)

5
In a normal steel pipe, the ratio between the diameter and the wall thickness is D/e =
50 100 and the elasticity modulus of steel is = 2 1011 N/m2 . In this case with a pressure
equal to a water column of 1000 m the mentioned denominator is equal to 0.997.
22 WATER HAMMER SIMULATIONS

l
v

Mv
I
m

Mm
x
x

Figure 2.8: Control volume to carry out the momentum equation.

(p A) p A
v = p A + x + x2 (2.36)
x x x
The difference between eqns (2.36) and (2.35), neglecting higher infinitesimals, is:

(p A)
x (2.37)
x

The axial component of the force applied by the pipe wall is:

p x A
l = p + x (2.38)
x 2 x

The axial component of the weight of the fluid volume is:

G sin = A x sin (2.39)

where sin = dz/dx.


For the Newtons second law the addition of all the forces acting on a volume
is equal to the inertia force due to the acceleration of the mass in the volume.
Accelerating, the mass moves in the axis direction, and, therefore, the acceleration
has to be computed as substantial derivative of the velocity V . Therefore:

V V
I = A x +V (2.40)
t x
COMPRESSIBLE FLOW THEORY: BASIC CONCEPTS 23

Note that, because of the assumption of the control volume, in this term
the differences between the incoming and outcoming momentum flux is already
computed6 .
The resulting shear force, being the shear stress at the pipe wall, is equal to
D x. Under the hypothesis that the headlosses in unsteady flow conditions
are similar to those in steady flow, it yields:
DJ
= (2.41)
4
J being the slope of the hydraulic grade line. Adding all the terms:

(p A) A dz V
x p x + A x + A x
x x dx t

V DJ
+ A x V + D x = 0 (2.42)
x 4
which can be simplified as:

p V V dz
+ +V + J + =0 (2.43)
x t x dx

2.5.3 The governing equations

Finally, the governing equations can be written together as:

p p V
+V + c2 =0 (2.34)
t x x

p V V dz
+ +V + J + =0 (2.43)
x t x dx
where eqn (2.34) is the continuity equation and eqn (2.43) is the momentum
equation.
Being two partial differential equations, these must be integrated to both initial
and boundary conditions.

6
Differently, the local inertias have to be computed, which have to be written as I =
(V ) (V )
t
dV = t
A dx and the momentum flux are m = V 2 A and v = V 2
V
2 A)
A + (Vx
dx + A
t
V dx. In the latter equation, the second term is the momentum
flux through the plane section, while the third term is the momentum flux through the lat-
eral surface. Adding up and simplifying these terms, keeping in account that A t
V =V
(AV )
(A)
t
V A
t
and that (because of the continuity equation) x
+ (A)
t
= 0, the same
result is achieved.
24 WATER HAMMER SIMULATIONS

These equations have been carried out using the unknown parameters the pres-
sure p and the velocity V . It is quite obvious that similar results could have
been carried out using unknowns the head h and the discharge Q (or any of their
combination) being, respectively:
p
h=z+ (2.44)

Q =V A (2.45)

Independent variables are time t and the longitudinal abscissa x.


As mentioned earlier, the other variables c, (or ), D and the pipe roughness f
do not vary significantly during the transient and therefore they can be let constants.
Rigorously, the hydraulic grade line does not depend only on the roughness f
(which in the following will be let constant), but also from the Reynolds number
Re; however, in practical cases these variations are negligible.
In matrix form, the governing equations are written as:
 
  V c2     

  
    0
p p
     
 +1   =  (2.46)
t  V   dz 
 V  x V  g gJ
dx
or:

U 
 U
+B 
=E (2.47)
t x
where:
   
  V c2   
p  
0

 =  
U  =  1

B

  =
E  (2.48)
V    dz 
 V  g gJ 
dx
 can be found solving the equation (in the
The eigenvalues of the matrix B
unknown ):
(V )2 = c2 =V c (2.49)
As both eigenvalues are real and different, the system of differential equation
is said to be hyperbolic.
The governing equations can be solved in different ways, as will be seen in
Chapters 4 and 5, but some preliminary consideration has to be drawn beforehand
and these will be presented in Chapter 3 together with few simplified methods to
carry out approximate solutions.
Chapter 3
Simplified solutions

With the availability of computer the governing equations carried out in Chapter 2
are quite easily numerically integrated. In the past, if only the peak of pressure
was needed the AllieviJoukowsky formula (2.3) or (2.4) or the Michaud formula
(2.15) or (2.16) were used.
If the values of pressure and velocity were required for the whole transient,
simplified approaches were used.
In this chapter, three methods are presented. The first is the Allievis method,
which uses simplified equations and finds the solution iteratively; the second is
based on the hypothesis that the fluid is non-elastic, again to find a simplified
solution and the third is the graphical method.

3.1 The Allievis method (1913)


This method has been presented in 1913 [5] and the equations obtained are known
as the Allievi chained equations. The governing equations are simplified as follows.
Continuity equation is written as:

p V
+ c2 =0 (3.1)
t x
while momentum equation is:

p V
+ =0 (3.2)
x t
These equations are valid under the following hypothesis:

1. velocity V is negligible when compared with the wave celerity c and, there-
fore, also terms like |V F/x| are negligible when compared with terms like
|F/t|;
2. headlosses are negligible;
3. density variations, due for instance to elevation variations, are negligible.

The usual sketch of simple plant is considered (Figure 3.1). Having the abscissa
starting from the downstream valve (x = 0) and assuming the x-axis positive in the
26 WATER HAMMER SIMULATIONS

Upstream reservoir

Downstream valve
x

x
0

Figure 3.1: Sketch of a simplified plant.

upstream direction, the governing equations can be written as:


p V
= c2 (3.3)
t x
p V
= (3.4)
x t
where eqn (3.3) is the continuity equation and eqn (3.4) is the momentum equation.
The initial conditions, which define the pressures and the velocities before the
beginning of the transient, are:

p(x, t = 0) = p0 (3.5)

V (x, t = 0) = V0 (3.6)

Boundary conditions are given at the upstream reservoir:

p(x = L, t) = p0 (3.7)

and at the downstream valve, where the law of velocity variation is known:

V = V (x = 0, t) (3.8)

Deriving eqn (3.3) with respect to t and eqn (3.4) with respect to x, and eliminating
the term in V , a single second-order differential equation can be carried out:

2p 2
2 p
= c (3.9)
t 2 x2
Similarly, deriving eqn (3.3) with respect to x and eqn (3.4) with respect to t and
eliminating the term in p, it yields:
2V 2
2 V
= c (3.10)
t 2 x2
SIMPLIFIED SOLUTIONS 27

Equations (3.9) and (3.10) are the DAlembert equations [25] and their solution
is well-known in Mathematical Analysis [26]. The solution of eqn (3.9) can be
written as:
p = p p0 = F (x c t) + f (x + c t) (3.11)
where F and f are functions to be defined on the basis of boundary conditions. This
equation describes the trend of the pressures during transients.
With regards to the velocity, its derivative can be computed deriving eqn
(3.11) with respect to x and substituting the result in the continuity equation (3.3),
obtaining1 :
V F(x c t) f (x + c t)
c = (3.12)
t t t
This equation, integrated with respect of time, yields:

c V = c (V V0 ) = F(x c t) f (x + c t) (3.13)

With eqns (3.11) and (3.13), it is possible to compute velocities and pressures in
the pipe.
For each couple of values (xi , ti ) for which the following relation is valid:

x1 c t1 = x2 c t2 (3.14)

the function F must have the same value. The same is for the function f , for which
the relation that ties the couple of coordinates is:

x1 + c t1 = x2 + c t2 (3.15)

That means the values the function F assumes at time t1 and at time t2 given by:

x2 x1
t2 = t1 + (3.16)
c

must be equal or, in other words, that the function F describes a wave which
propagates with celerity c. As mentioned, for the function f the same reasoning can
be done.
Function F is, therefore, related to a perturbation that propagates from the valve
to the reservoir, while function f is a wave which has opposite direction (i.e., from
the reservoir towards the valve).
Few simple considerations allow to further simplify the problem, letting it to be
solved with only one equation.

 
1
It can be demonstrated that: p
x
= F
x
+ f
x
= 1c F
t
f
t
.
28 WATER HAMMER SIMULATIONS

At the section located near the reservoir, i.e., x = L, the pressure p must be
constant, and, therefore:

p = F (L c t) + f (L + c t) = 0 (3.17)

At the junction between the reservoir and the pipe, the two functions have the
same value but opposite sign. In other words, when from the downstream pipe a
wave arrives to the reservoir, a new wave is generated, with opposite sign, which
propagates downstream. This result has been already found in Section 2.1.
Therefore, if at time t1 a disturbance f (L c t1 ) is generated at the reservoir,
it propagates downstream with opposite sign; at a subsequent time t > t1 , the dis-
turbance is located in the section x, according to the expression t = t1 + L c x and
due to the relation L + c t1 = x + c t. This relation together with eqn (3.17) of the
reservoir, it yields:

f (x + c t) = f (L + c t1 ) = F(L c t1 ) (3.18)

A more general expression can be written as follows:


 
2 (L x)
f (x + c t) = F x c t (3.19)
c

which connects the two functions. Expressions (3.11) and (3.13) then become much
simpler:
 
2 (L x)
p = p p0 = F(x c t) F x c t (3.11 )
c

 
2 (L x)
c V = c (V V0 ) = F(x c t) + F x c t
c
(3.13 )

As in the first phase of the transient only the ascending wave is present, the equations
simplify to:
p = p p0 = F(x c t) (3.11 )

c V = c (V V0 ) = F(x c t) (3.13 )
Removing F from both, the well-known formula p = c V is derived again.
Let = 2L
c
, for the section x = 0 the eqns (3.11 ) and (3.13 ) can be finally
written as:
p = p p0 = F(t) F(t ) (3.11 )

c V = c (V V0 ) = F(t) + F(t ) (3.13 )


SIMPLIFIED SOLUTIONS 29

In order to study the second phase of the transient, when the ascending waves adds
with those descending, eqns (3.11 ) and (3.13 ) are used, choosing a sequence of
times ti that:

0 < t1 <
t2 = t1 +
t3 = t2 +
..
.
ti = ti1 +

so that, for t = t1 :

p(t1 ) p0 = F(t1 ) c [V0 V (t1 )] = F(t1 ) (3.20)

being still in the first phase. For the following instants:

t = t2 :
p(t2 ) p0 = F(t2 ) F(t1 ) c [V0 V (t2 )] = F(t2 ) + F(t1 ) (3.21)
..
.
t = ti :
p(ti ) p0 = F(ti ) F(ti1 ) c [V0 V (ti )] = F(ti ) + F(ti1 ) (3.22)

Here again, removing F from the two series of equations, it yields:

p(t1 ) p0 = c [V0 V (t1 )]


p(t2 ) + p(t1 ) 2 p0 = c [V (t1 ) V (t2 )]
(3.23)
..
.
p(ti ) + p(ti1 ) 2 p0 = c [V (ti1 ) V (ti )]

With eqns (3.23), the study of the pressures at the valve during the transient is
possible, at any time. In these equations, in fact, the time step has to be equal to :
but, as the initial time t1 is chosen arbitrarily, under the only condition t1 < , the
values can be actually computed for any time.
For a generic section with abscissa x the procedure is similar; the computations
start from the values p and V at the downstream valve. Neglecting all the passage
for the sake of brevity, the final equations are:
  x  x 
x p = F0 tn F0 tn1 + (3.24)
c c
30 WATER HAMMER SIMULATIONS
g   x  x 
x V = F0 tn + F0 tn1 + (3.25)
c c c
In eqns (3.24) and (3.25), x p and x V refers to pressure and velocity variations,
respectively, at the section x. With F0 (t) the value that the function F has at the
section x = 0 at time t is indicated. Values x p and x V can be iteratively computed
with the following simple procedure:

1. pressure and velocity values of at the downstream valve are computed with
eqn (3.23) at different times;
2. values of the function F0 (t) are computed iteratively from: F0 (ti ) = F0 (ti1 ) +
[p(ti ) p0 ];
3. values of x p and x V are computed with eqns (3.24) and (3.25).

3.2 The non-elastic hypothesis


3.2.1 Governing equations

In Figure 3.2, the classic scheme of an hydropower plant is shown.


This is the simplest scheme of an hydropower plant, with an upstream reservoir;
a tunnel, which is usually quite large and where velocities are relatively low, in order
to reduce headlosses; a surge tank, the use of which will be discussed in Chapter 6;
a penstock, where velocities are very high, as at the end there is a turbine and the
target is the production of the maximum possible amount of electricity. This plant
can be studied as a whole with the mathematical methods described later in this
book.
However, in the past this was not possible and the only way to study the behaviour
of the plant was splitting it in two parts: first the penstock, which was studied with
theAllievi or Michauds formula or theAllievis method; then the tunnel was studied,
pretending the downstream pipeline was nonexistent and therefore positioning the
downstream valve immediately downstream the surge tank; moreover, the valve
was supposed to close instantaneously.

A1  A
R
A B
Downstream
Tunnel valve
Penstock
Reservoir
Turbine
T

Figure 3.2: Sketch of a classic hydropower plant.


SIMPLIFIED SOLUTIONS 31

These hypotheses mean the two phenomena (in the tunnel and in the penstock)
are not connected or, better, that they can be studied separately; this is not so
senseless, as the phenomenon in the penstock depends on wave celerity (around
1000 m/s) and it is so fast that it can be considered finished when the transient in the
tunnel starts; the latter depends on the flow velocity (few meters per second), there-
fore being much slower. However, in this section the accuracy of these hypotheses
will be tested.
In this simplified system, where only the tunnel is considered, the hypotheses of
non-elasticity of the fluid and of the non-deformability of the pipe can be accepted.
In essence, the water flowing in the tunnel cannot enter the penstock and therefore
flows in the surge tank, rising its water level and consequently the piezometric head
at the bottom of the tank itself.
The discharge reduces until the maximum level in the tank is reached, which is
higher than the level in the upstream reservoir because of the water inertia; when
that maximum point is reached, the velocity of the water in all points of the system
is zero. At this point, the system has not reached the equilibrium, and therefore the
level in the tank decreases while the flow in the pipe has negative velocity, which
becomes zero when the level in the tank is minimum.
The transient continues with a series of oscillations that reduce their amplitude
because of the headlosses. This phenomenon can be someway considered similar
to the oscillations that can be recorded in an U-shaped pipe, shown in Figure 3.3.
As the fluid is considered uncompressible and without headlosses, Bernoullis
theorem is:

H V
+ =0 (3.26)
s g t

where the Coriolis coefficient 1. Because of the non-compressibility of the


fluid and the non-deformability of the pipe walls, the continuity equation becomes
V (s) = constant, and therefore:

V (s) dV
= (3.27)
t dt

As a consequence, the Bernoullis theorem can be integrated, resulting:

 2
1 dV L dV
H2 H1 = ds = (3.28)
g dt 1 g dt

Let z the elevation of the free surface, because of the continuity equation, the
following equation can be written:

dz d dV 1 d2 
A dz = Q dt = V = 2 V = 2 (3.29)
dt dt dt 2 dt
32 WATER HAMMER SIMULATIONS

Difference
in level
Area A

Length L

Figure 3.3: Mass oscillation in a U-shaped pipe.

 being the difference in elevation between the two free surfaces, as shown in
Figure 3.3. Finally, Bernoullis theorem becomes:

d2 
+ 2  = 0 (3.30)
dt 2
where:

2g
= (3.31)
L
Imposing the initial conditions:

(t = 0) = 0
V (t = 0) = 0 (3.32)

a periodical motion is obtained, i.e., oscillations, described by:

 = 0 cos( t) (3.33)

If the headlosses cannot be neglected, Bernoullis theorem has to be modified as:


H V
+ +J =0 (3.34)
s g t
which, upon integration, yields:
L dV
+LJ =0 (3.35)
g dt
SIMPLIFIED SOLUTIONS 33

This equation, together with the continuity equation, form a set which in general
does not allow analytical integration. It is not necessary to describe too much in
detail these equations, but it has to be said that in turbulent flow and in some cases
in laminar flow, the motion is oscillatory and it is smoothed because of the flow
resistances. Particular cases may be faced when the flow is laminar and the viscosity
of the fluid is very high: in these cases, the initial different elevation asymptotically
tends to zero with no changes in sign.
However, generally speaking and assuming the presence of concentrated head-
losses in the connection between the tunnel and the surge tank (see Chapter 6), the
governing equations are:

L dV
+Z H K =0
g dt (3.36)

Asurge tank dZ = Q Qds
dt
where Z is the water elevation in the surge tank, computed from the static level;
H is the value of the headlosses in the tunnel; K the concentrated headlosses in the
connection between the tunnel and the surge tank; Q is the discharge in the tunnel
at time t and Qds is the discharge at the same time in the penstock.

3.2.2 Comparisons

It is now interesting to compare the results that can be obtained with this simplified
model with those that can be carried out with a numerical model that uses the
complete equations described in Chapter 2.
To do that, let us consider a plant with the following characteristics:
Q0 = 15 m3 /s discharge;
Dtunnel = 2.80 m diameter of the tunnel;
L = 4200 m length of the tunnel;
Dsurge tank = 11.50 m diameter of the surge tank.
The surge tank has circular cross section and no concentrated headlosses at the
junction with the tunnel. With the mentioned hypotheses and integrating with finite
differences, the governing equations (3.36) become:

L V
+ Z V 2 (t) = 0
g t (3.37)

Asurge tank Z = Q(t) = Atunnel V (t)
t
In eqn (3.37), the discharge in the penstock is Qds = 0, that means the transient
is finished. Distributed headlosses in the tunnel are computed as:

L 0.018 4200
= = = 1.377
2gD 2 9.806 4200
34 WATER HAMMER SIMULATIONS

In this case, the resistance term is = 0.018, which means full turbulent flow
and a Manning resistance term equal to n = 0.143 s m1/3 . As the initial veloc-
ity in the pipe is equal to 2.44 m/s, the steady flow headlosses are equal to
H = V 2 = 8.17 m = Z0 .
The method of integration is explicit and really simple: given a timestep t, from
the continuity equation, the difference in the water elevation z is computed by:

Atunnel
Z = Vi t (3.38)
Asurge tank

and therefore the elevation at the new time ti+1 is given by:

Zi+1 = Zi + Z (3.39)

Then the velocity variation is computed:

g
V = (Zi+1 Vi2 ) t (3.40)
L

and finally the velocity at the new time:

Vi+1 = Vi + V (3.41)

Results of the computations, performed with a spreadsheet (but they can be carried
out by hand), are reported in Figure 3.4 and in Table 3.1.

10

0
Velocity (m/s)
Elevation (m)

5

10

15 Water elevation in the surge tank


Velocity in the tunnel
20
0 100 200 300 400 500 600
Time (s)

Figure 3.4: Oscillations in the pipe and in the surge tank, computed with the non-
elastic model.
SIMPLIFIED SOLUTIONS 35

Figure 3.5 reports the water elevations in the surge tank carried out with this
model and those obtained with the more complex simulation models described in
the following.
As can be seen, the agreement between the two sets of data is very good, at least
at the beginning, while in the subsequent instants of the simulation, the anelastic
model tends to underestimate the headlosses, thus being on the safe side.

Table 3.1: Numerical results of the non-elastic model: water levels and velocities
in the surge tank.

t Z Z V V t Z Z V V
(s) (m) (m) (m/s) (m/s) (s) (m) (m) (m/s) (m/s)
0 8.17 2.44
10 1.44 6.73 0.034 2.40 310 1.09 0.42 0.098 1.93
20 1.42 5.30 0.062 2.34 320 1.15 1.56 0.084 2.02
30 1.39 3.92 0.085 2.26 330 1.19 2.76 0.066 2.08
40 1.34 2.58 0.103 2.15 340 1.23 3.99 0.046 2.13
50 1.28 1.30 0.119 2.03 350 1.26 5.25 0.023 2.15
60 1.21 0.10 0.131 1.90 360 1.28 6.53 0.004 2.15
70 1.13 1.03 0.141 1.76 370 1.27 7.80 0.034 2.11
80 1.04 2.08 0.148 1.61 380 1.25 9.05 0.068 2.05
90 0.96 3.03 0.155 1.46 390 1.21 10.26 0.105 1.94
100 0.87 3.90 0.160 1.30 400 1.15 11.42 0.145 1.80
110 0.77 4.67 0.163 1.14 410 1.06 12.48 0.188 1.61
120 0.67 5.34 0.166 0.97 420 0.95 13.43 0.231 1.38
130 0.58 5.92 0.168 0.80 430 0.82 14.25 0.272 1.11
140 0.48 6.39 0.170 0.63 440 0.66 14.90 0.309 0.80
150 0.37 6.77 0.171 0.46 450 0.47 15.38 0.339 0.46
160 0.27 7.04 0.171 0.29 460 0.27 15.65 0.359 0.10
170 0.17 7.21 0.171 0.12 470 0.06 15.71 0.366 0.27
180 0.07 7.28 0.171 0.05 480 0.16 15.55 0.361 0.63
190 0.03 7.25 0.169 0.22 490 0.37 15.18 0.342 0.97
200 0.13 7.12 0.168 0.39 500 0.57 14.60 0.311 1.28
210 0.23 6.89 0.166 0.55 510 0.76 13.84 0.271 1.55
220 0.33 6.56 0.163 0.72 520 0.92 12.92 0.224 1.78
230 0.43 6.14 0.160 0.88 530 1.05 11.87 0.176 1.95
240 0.52 5.62 0.156 1.03 540 1.16 10.71 0.128 2.08
250 0.61 5.00 0.151 1.18 550 1.23 9.48 0.082 2.16
260 0.70 4.30 0.146 1.33 560 1.28 8.20 0.041 2.20
270 0.79 3.51 0.139 1.47 570 1.31 6.90 0.005 2.21
280 0.87 2.64 0.131 1.60 580 1.31 5.59 0.026 2.18
290 0.95 1.69 0.122 1.72 590 1.29 4.29 0.053 2.13
300 1.02 0.67 0.111 1.83 600 1.26 3.03 0.075 2.05
36 WATER HAMMER SIMULATIONS

10

0
Elevation (m)

5

10

15
Anelastic model
Elastic model
20
0 100 200 300 400 500 600
Time (s)

Figure 3.5: Water elevation in the surge tank: comparison between elastic and non-
elastic model results.

Actually, the hypothesis of non-compressibility of water moving in the


tunnel-surge tank is surely acceptable; however, the differences in the effort of
implementation of the two models is considerable.
Design of simple surge tanks is surely possible even with the simplified model.
The disadvantage of these models is not given by their precision, which is proven to
be sufficiently good, but instead on their poor versatility. Simplified models cannot
keep in due account the devices that can be installed in the surge tanks, or the
ramification the pipelines may have, or other problems. With complex models, all
these devices can be implemented, even if no model is exempt from errors, being
always necessary to critically evaluate the coherence of the carried out results and
to contrast the model instabilities.

3.3 Graphical method


This is probably the most visual representation of the effects of waterhammer and
it should be studied for a deeper understanding of the phenomenon, even if it is
not used anymore for the evaluation of real cases [8, 9, 27, 28]. As the theoretical
description might be confused, the method will be explained solving a simple
problem.
Again, the usual very simple plant reproduced in Figure 3.6 is considered, the
characteristics are those used in the example in Chapter 2: pipe length L = 1000 m
where the wave celerity is equal to c = 1000 m/s and the flow has a velocity equal to
V0 = 5 m/s. Pressures have to be computed in both cases of valve closure with
SIMPLIFIED SOLUTIONS 37

Upstream reservoir

Downstream valve

Figure 3.6: A simple plant to be solved with the graphical method.

6 V1

2
R2
1

0
0 0.2 0.4 0.6 0.8 1

V3

Figure 3.7: Graphical solution for sudden valve closure.

time Tc = 0.0 s (sudden valve closure) and Tc = 5.0 s. Headlosses are considered
negligible.
The dimensionless variables are defined as: h = H /H0 , v = V /V0 , t = T /Tc .
For the sudden valve closure, the following points have to be computed and
placed on a graph which x-axis is v and y-axis is h, as reported in Figure 3.7.
The point V0 refers to the steady flow conditions (time t = 0) at the valve V ,
and therefore has co-ordinates (v = 1; h = 1). Then we have to move on that graph
38 WATER HAMMER SIMULATIONS

following two lines whose slope S is equal to those of the characteristics, i.e.:

c V0 1000 5
S= = = 5.1 (3.42)
g H0 9.81 100

Starting from the point at the valve V0 and moving with slope S (negative
characteristic line) we touch the axis v = 0 (which is reached at time t = 1) with the
co-ordinate h = 6.1; this is the point V1 (0, 6.1) related to the time when the valve
is closed. Because of the hypothesis of negligible headlosses, the point R0 at the
reservoir R at time t = 0 is V0 R0 . The next point is related to the upstream
reservoir: starting from V1 and moving on the positive characteristic line, i.e.
with slope S, until the intersection h = 1, because the reservoir keeps the head
constant; therefore we reach the point R2 (1, 1). Continuing this procedure the
co-ordinates of V3 (0, 4.1) and V4 V0 are computed. As the headlosses are consid-
ered negligible, the oscillations do not reduce and the cycle continues as it has been
computed so far.
With regards to the slow closure, a timestep of 1 s is chosen, which, being
Tc = 5.0 s, in dimensionless form is t = 0.2. In the case of slow closure, the law to
be used is:

v = (1 t)n (3.43)

where n is an exponent that can be assumed equal to 1 for simplicity, but reminding
that this choice does not change the application of the procedure. The method is
reported in Figure 3.8: again the starting point is V0 (1, 1), and again we have to move
on the negative characteristic line to find the point V1 . This point is found at the
position computed with the (3.43) and therefore v = 0.8. From the point V1 , moving
on the negative characteristic we arrive to the point V2 to be reached at time t = 0.4
and therefore v = 0.6. At the same time, moving on the positive characteristic, the
point R2 is found, which has also to be at the intersection of h = 1 because the point
R is at the reservoir, and therefore, as known, with constant head.
Continuing this procedure, point V3 can be positioned and so on until the
requested time of simulation is reached. The points computed with this procedure
are shown in Figure 3.8.
This solution indicates that the maximum peak of pressure is reached at time
t = 0.4 (in the dimensional values this is equal to 2 s) and it has a value that can be
estimated equal to 3; dimensionally, this means that the peak has to be expected
around 300 m. After this high peak, the oscillations are repeated with smaller values,
graphically estimated equal to 2, and therefore around 200 m.
Moreover, after the complete closure of the valve, t > 1 and the values of R are
on the horizontal line h = 1 as usual, and oscillate between v = 0.2 and 0.2; the
values of V are on the vertical line v = 0 and oscillate between h = 0.0 and 2.0.
With regard to the effect of the headlosses, they can be considered but using
an expedient, which consists in applying them at the upstream end of the pipe.
Therefore, the boundary upstream condition is not represented by an horizontal
SIMPLIFIED SOLUTIONS 39

7
t  1.0 t  0.8 t  0.6 t  0.4 t  0.2 t  0.0

Ch
5 ara
cte
rist
ic l
ine
4 c-
V2
3
h

V5  V6
V3 V1
2 R R
6 7

1
V4  R3  R4  R5 h  1 R2 V0  R0  R1
0
V7
1
0.2 0 0.2 0.4 0.6 0.8 1
v

Figure 3.8: Graphical solution for slow valve closure.

straight line h = 1.0 but by a parabola given by:



L V02 v 2
h=1 1+ (3.44)
D 2 g H0

Under all the other aspects, the procedure remains unchanged.


This page intentionally left blank
Chapter 4
Numerical solution of the governing equations:
The method of characteristics

There are many methods to numerically integrate differential equations, which can
be roughly divided in three main families: finite differences, finite elements and
characteristics. One of the first integration methods used to solve waterhammer
problems is the method of characteristics applied by Evangelisti to the case of fluid
transients [5, 6] following a more general development [29, 30]. This method has
proved to be easy, fast and reliable. Actually, many problems cannot be faced with
this integration method, but it surely deserves a description, especially because its
worldwide diffusion.
The method of characteristics is a technique for solving partial differential equa-
tions. Typically, it is applied to solve first-order equations, although more generally
it is valid for any hyperbolic partial differential equation. The method is to reduce
a partial differential equation to a family of ordinary differential equations along
which the solution can be integrated from some initial data given on a suitable
hypersurface.

4.1 Numerical solution

A linear combination of the partial derivative of a given function u (x, t), for
instance:
u u
a +b (4.1)
x t
is the derivative of the function in the direction:

dx a
= (4.2)
dt b

The problem is the identification, on the plane (x, t), of particular direc-
tions along which in the governing equations (continuity: eqn (2.34) and motion:
eqn (2.43)) the functions p(x, t) and V (x, t) are derived, in order to reduce the partial
derivative differential equations to ordinary differential equations. These directions
are called characteristics and, if they are real and distinct, the system is said to be
hyperbolic.
42 WATER HAMMER SIMULATIONS

To identify these directions, the two equations are linearly combined, multi-
plying the momentum equation by a constant K and summing the result to the
continuity equation [10, 11]. The result is:

p V V dz
K +K +K V +K J +K
x t x dx
p V
+V + c2 =0 (4.3)
x x

upon collecting the common terms, it yields:



p p V V dz
(K + V ) + + (K V + c2 ) +K + K J + =0
x t x t dx
(4.4)

The coefficients of the derivative of p and V with respect to x and t are:

p
1 (=a) (4.5)
t
p
K +V (=b) (4.6)
x
V
K (=c) (4.7)
t
V
K V + c2 (=d) (4.8)
x

Therefore, p and V are derivative along the same direction if a/b = c/d, and
consequently if:

1 K c2
= K +V =V + (4.9)
K +V (K V + c2 ) K

i.e., if K = c.
As mentioned, the system is hyperbolic because the two characteristics are real
and distinct, and they are:
dx
=V +c (4.10)
dt

dx
=V c (4.11)
dt
It can be noted that the velocity V is normally much lower than 10 m/s, while
the wave celerity is normally in the range 800 1200 m/s. As a consequence, an
acceptable approximation is V + c c and V c c.
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 43

The governing equations can be rewritten from eqn (4.4), inserting K = +c


and c. Reminding that:
p p
dp = dt + dx (4.12)
t x

V V
dV = dt + dx (4.13)
t x
and being in this case:
 
p p
dp = c dx (4.14)
t x
 
V V
dV = c dx (4.15)
t x
Writing again eqn (4.4) yields:
   
p p V V dz
c c c c J c =0 (4.16)
t x t x dx
or:
dp dV dz
c c J c =0 (4.17)
dt dt dx
Writing the piezometric head h instead of the pressure p, being:
p
h=z+ (4.18)

it finally results:

p h z
= + (h z) (4.19)
x x x x
In eqn (4.19) a good approximation is:

0 (4.20)
x
Moreover:
p h
= (4.21)
t t
Inserting these expressions in eqn (4.16), the following system of two ordinary
differential equations is obtained, which is equivalent to eqns (2.34) and (2.43):
dh c dV

dt + g dt + c J = 0

(4.22)


dh c dV c J = 0
dt g dt
44 WATER HAMMER SIMULATIONS

where the first equation is valid along the positive characteristic line (4.10) and the
second is valid along the negative characteristic line (4.11).
Expressing the headlosses with the uniform flow formula:

V |V |
J = (4.23)
2gD

and writing eqn (4.22) with finite differences, it yields:



h c V V |V |

t + g t + c 2 g D = 0

(4.24)

h c V V |V |

c =0
t g t 2gD

which can be written using an explicit method:


 
hi, j hi1, j1 c Vi, j Vi1, j1 Vi1, j1 Vi1, j1 

+ +c =0
t g t 2gD
  (4.25)

h h c V V V Vi+1, j1 

i, j i+1, j1

i, j i+1, j1
c
i+1, j1
=0
t g t 2gD

This is an algebraic system of two equations in the two unknown hi, j and Vi, j
that can be easily solved, giving:

1 c  
hi, j = hi+1, j1 + hi1, j1 Vi+1, j1 Vi1, j1 +
2 g

c     
   
Vi1, j1 Vi1, j1 Vi+1, j1 Vi+1, j1 t
2gD

g c  
Vi, j = hi+1, j1 + hi1, j1 + Vi+1, j1 + Vi1, j1 + (4.26)
2c g

c     
   
Vi1, j1 Vi1, j1 + Vi+1, j1 Vi+1, j1 t
2gD

Graphically, the characteristic lines may be represented on a (x, t) plane, where


a point has co-ordinated x = i x and t = j t and the variables h and V in this
point are indicated as hi, j and Vi, j . In Figure 4.1 the (x, t) plane with the symbols
mentioned is shown.

4.2 Initial and boundary conditions

With eqns (4.26) it is possible to compute the values of the required parameters in
the internal points of the field, when the initial and boundary conditions are known.
As a differential equations problem, in fact, these conditions must be defined.
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 45

Figure 4.1: (x, t) plane with the representation of a point.

As mentioned in Chapter 1, initial conditions are normally computed with the


usual methods of steady flow in water supply networks, under the hypothesis that
before the transient these methods can be applied.
With regard to the boundary conditions, these depend on the device positioned
at the extreme ends of the pipe, which determine the behaviour of the system.
A specific boundary condition settles one of the parameters (head h or velocity
V ), or ties these two parameters with a definite function. The two parameters h
and V at each end of the pipe are computed solving the equation(s) which describe
the boundary condition together with the characteristic line, for downstream or
upstream conditions, respectively.

4.2.1 Reservoir

The simplest case that can be handled is probably the reservoir, under the hypothesis
that it is large enough to keep its level unchanged during the transient: in this case,
the head can be considered constant. To this very simple condition, a characteristic
equation has to be associated.
Let us assume the pipe is divided into n + 1 sections, from 0 to n and the reservoir
is positioned upstream (i.e., in the section 0), it yields:

h(0, t) = h0 t the head in the reservoir is constant (4.27)

g V (1, t) |V (1, t)|


V (0, t + 1) = V (1, t) + [h0 h(1, t)] t (4.28)
c 2D

Very similar is the case of a reservoir positioned downstream (i.e., in the


section n), so that to the condition that impose the constant head has to be associated
the positive characteristic line; it yields:

h(n, t) = h0 t the head in the reservoir is constant (4.29)


46 WATER HAMMER SIMULATIONS

c c
Reservoir Reservoir
upstream downstream

Figure 4.2: Scheme for the reservoir (upstream and downstream) on the (x, t) plane.

g V (n 1, t) |V (n 1, t)|
V (n, t +1)=V (n1, t) [h0 h(n 1, t)] t
c 2D
(4.30)
The scheme of the method is reported in Figure 4.2.

4.2.2 Valve

Valves are very complex devices, but in this section the simplest description is
given and implemented: the hypothesis is that to a certain opening of the valve
corresponds an univocal value of the flow velocity. In this case, eqn (2.11) seen in
Section 2.3 can be used again, even if given more generally as follows, to allow its
use even for nonlinear operations:

t n
V (0, t) = V0 1 when t Tc
Tc (4.31)

V (0, t) = 0 when t > Tc

In this case, the valve is positioned upstream, i.e., in the section 0. Therefore,
the characteristic equation to be associated is negative and yields:

c c V (1, t) |V (1, t)|


h(0, t + 1) = h(1, t) + [V (0, t + 1) V (1, t)] + t
g g 2D
(4.32)
Again, very similar is the case of the valve positioned downstream, where the
condition on the velocity has to be associated to the positive characteristic line.

t n
V (n, t) = V0 1 when t Tc
Tc (4.33)

V (n, t) = 0 when t > Tc
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 47

c
h(n, t + 1) = h(n 1, t) [V (n, t + 1) V (n 1, t)]
g
c V (n 1, t) |V (n 1, t)|
t (4.34)
g 2D

4.2.3 Junction

This case is quite interesting as it allows the understanding of the characteristic


lines, visualizing schematically the computing diagram.
Let us suppose the simplest case, when two pipelines meet in a junction. The
indexes 1 and 2 will be used for the upstream and downstream pipes, respectively.
The upstream pipe is divided into n + 1 sections, from 0 (upstream) to n (down-
stream); the downstream pipe is divided in m + 1 sections numbered from n to
n + m. Obviously, the section n, where the junction is positioned, belongs to both
the upstream and downstream pipes; however, it will be divided in nup and ndown as
the flow parameters h and V are, generally speaking, different.
Should the concentrated headlosses in the junction be negligible, the heads have
to be equal, and therefore:

h(nup , t + 1) = h(ndown , t + 1) (4.35)

In general cases, headlosses are not negligible, and therefore a further term is
to be added to eqn (4.35), normally a function of the velocity V .

The continuity equation is:

Dup
2
Ddown
2
V (nup , t + 1) = V (ndown , t + 1) (4.36)
4 4

To these two equations, the two characteristic lines (one positive for the down-
stream pipe and one negative for the upstream) have to be associated, forming a
system of four equations in the four unknown flow parameters hup , hdown , Vup and
Vdown .
As working with this numerical method the condition t = x/c must be sat-
isfied, the assumption of generic spatial discretization steps xup and xdown is
shown in Figure 4.3. The use of arbitrary spatial steps brings to different time steps,
which force the user to interpolate the carried out values at the junction to compute
the flow parameters h and V at the same time. This is possible, but it introduces
approximation errors that decrease the precision of the results; moreover, interpo-
lation algorithms must be implemented in the computer code, slowing the solution:
with this solution, the advantages given by the method of characteristics, precision
and velocity, have to be partially discarded.
48 WATER HAMMER SIMULATIONS

t2
t1

x
0
n nm

Figure 4.3: xt plane for two pipes with different characteristics and problem
at the junction.

As a consequence, working with the method of characteristics the condition that


the timestep t be the same for all the pipes is normally imposed. That means:

x1 x2
t = = (4.37)
c1 c2

which is:
l1 l2
= (4.38)
n c1 m c2
Again a new problem arises: as the parameters n and m must be integer, it may
be necessary to slightly change the geometric characteristics of one of the pipes
(upstream or downstream). If a new length for the downstream pipe l2 is to be
computed, that can be done with the following:

l1
l2 = m c2 (4.39)
n c1

This new value is obviously different from the real length, but simple computa-
tions show that if the pipelines are long few thousands meters, and if their lengths
do not differ too much, the difference between the original and the new computed
lengths is equal to few tens of centimeters, and therefore completely insignificant
on the computed results.
A non-negligible problem can arise when a network with very different pipe
lengths has to be simulated. In this case, the condition (4.38) may drive to a very
small timestep and a huge number of sections in the longest pipes, which are not
necessary and that highly increase the computational effort. In these cases, the use
of an interpolation algorithm can be suggested, because the slowing down of the
program is matched by the possibility to have an higher timestep and a reduced
section number in the longest pipes.
The generalization for a higher number of pipes is possible, as for each pipe two
new unknowns are set up (the head and the velocity in the junction for the pipeline
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 49

ReadNetworkCharacteristics;
ReadInitialConditions;
t=0
while t < tmax do
begin
t: = t + dt;
ComputeInternalPoints;
ComputeBoundaryConditions;
end;

Figure 4.4: Main procedure of the solution program.

Procedure ComputeInternalPoints
for i:=1 to n-1 do
begin
vp[i]:=0.5*(v[i-1]+v[i+1]+(g/cel)*(h[i-1]-h[i+1])-
(Lambda*dt/(2*diam))*(v[i-1]*abs(v[i-1])+v[i+1]*abs(v[i+1])));
hp[i]:=0.5*(h[i-1]+h[i+1]+(v[i-1]-v[i+1])/(g/cel)-
(Lambda*dt/(2*diam))*(v[i-1]*abs(v[i-1])-v[i+1]*abs(v[i+1]))/(g/cel));
end;

Figure 4.5: Core procedure of the solution program: calculation of the unknowns
(head and velocity) for the points internal to the pipe.

end) and two new equations are added to the system: the equality of the heads,
similar to eqn (4.35) and the characteristic line; moreover, the continuity equation
(4.36) must be changed accordingly.

4.3 The computer code


In this chapter, a simple computer program is shown and applied to simulate a pipe
with a reservoir positioned upstream and a simple valve positioned downstream;
the program has the simple structure reported in Figure 4.4.
The core procedures have the structures reported in Figures 4.5 and 4.6; as it
can be seen, they exactly match eqns (4.26) for the internal points (Figure 4.4),
eqns (4.27) and (4.28) for the upstream reservoir and eqns (4.33) and (4.34) for the
downstream valve (Figure 4.5).
The parameters to be used (heads and velocities) are stored in different arrays:
h[1 . . . NumberOfSections] and v[1 . . . NumberOfSections]. The unknowns, i.e.,
the parameters at time t + dt, are stored in the arrays hp[1 . . . NumberOfSections]
and vp[1 . . . NumberOfSections].
50 WATER HAMMER SIMULATIONS

Procedure ComputeBoundaryConditions;
// Reservoir positioned upstream
hp[0]:=h0;
vp[0]:=v[1]+g*(h[0]-hp[1])/cel-Lambda*v[1]*abs(v[1])*dt/(2*diam);
// Valve positioned downstream
if t<tc then
begin
tau:=power((1-t/tc),m);
vp[n]:=v0*tau;
hp[n]:=h[n-1]-cel*(vp[n]-v[n-1])/g-Lambda*cel*v[n-1]*
abs(v[n-1])*dt/(2*g*diam);
end else
begin
tau:=0;
vp[n]:=0;
hp[n]:=h[n-1]+cel*v[n-1]/g-Lambda*cel*v[n-1]*
abs(v[n-1])*dt/(2*g*diam);
end;
end;

Figure 4.6: Core procedure of the solution program: calculation of the unknowns
(head and velocity) for the points at the boundary of the pipe.

4.4 First simple application


The simplest example is given by a pipe which connects an upstream reservoir with
a downstream valve. The pipe has a diameter equal to 75 mm, length of 800 m
and the headlosses are negligible. The wave celerity has been computed equal to
1000 m/s and the initial discharge is equal to 2.5 l/s. The downstream valve has a
sudden closure.
Using the described computer program, the inserted parameters are presented
in Figure 4.7, which shows the users interface.
As for the simulation parameters, the number of computational sections has
been let equal to 1000, which is very high, and the effect will be discussed later;
the checkbox Ignore Headlosses has been thicked and the computation time set
equal to 15 s.
Results are printed in two text files and imported in a spreadsheet. As the program
can produce a very large amount of data, they might be reduced by printing them
less frequently, fixing a larger time step, in the example let equal of 0.1 s; moreover,
results are printed only in few sections, as mentioned in Chapter 1, at the upstream
and downstream ends, and at the 20, 40, 60 and 80% of the length of the pipe, and,
therefore, in this case at 160, 320, 480 and 640 m from the edges of the pipe.
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 51

Figure 4.7: Users interface of the program.

Very simple hand calculation show that the spatial discretization step and the
time step are:
PipeLength 800
ds = = = 0.8 m
NSections 1000

ds 0.8
dt = = = 0.001 s
celerity 1000
The velocity in the pipe, in steady flow (initial) conditions is equal to 0.566 m/s.
As the valve suddenly closes, the surcharge can be computed with the Allievi
Joukowskys formula (2.4), which yields:

c 1000
h = V0 = 0.566
= 57.7 m
g 9.81

As the wave celerity is equal to 1000 m/s the phase time of the system, i.e., the
time, the wave requires to reach the reservoir and to be back to the valve is equal
to 2 L/c = 1.6 s.
The results of the simulation carried out with the computer program are shown
in Figure 4.8; as can be seen, these results are perfectly consistent with the hand
calculations.
52 WATER HAMMER SIMULATIONS

180

160

140

120
Head (m)

100

80

60

40

20

0
0.0 1.5 3.0 4.5 6.0 7.5 9.0 10.5 12.0 13.5 15.0
Time (s)

Figure 4.8: Results of the first simulation: sudden closure and negligible
headlosses.

When a slower operation is performed at the downstream valve, results are


expected to change. In particular, the second simulation is performed considering
the valve closes in 3 s. In this case, the Michauds formula (2.16) has to be used,
which yields:

2 L V0 2 800 0.566
h = = = 30.8 m
g Tc 9.81 3

As described in Chapter 2, the maximum head is reached at time 2 L/c = 1.6 s;


results are shown in Figure 4.9. These results can be also compared with those
carried out with the graphical method in Chapter 3.
In order to check the effect of the headlosses, a third simulation has been
performed, letting again the valve closure time equal to zero and the roughness
n = 0.0125 m1/3 s. Results are shown in Figure 4.10 and are interesting because
of two aspects. The former is that the maximum value is not reached immediately.
The reason of this phenomenon can be found in Section 2.4, where it is explained
that at the first timestep only the first element is stopped, but the surcharge reached
is not enough to stop the whole water column, and therefore it stops in a larger
time. The latter aspect is due to the reduction of the amplitude of the oscillations:
again this is due to the headlosses, that induce a loss of energy (head) in the water
column, so that the surcharge fades with time.
The last simulation is carried out with the same parameter of the first, i.e., valve
closure time equal to zero and negligible headlosses, but with a smaller number of
section, let equal to 10; moreover, the time of simulation is equal to 30 s. Results are
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 53

140

120

100
Head (m)

80

60

40

20

0
0.0 1.5 3.0 4.5 6.0 7.5 9.0 10.5 12.0 13.5 15.0
Time (s)

Figure 4.9: Results of the second simulation: slow closure and negligible
headlosses.

180

160

140

120
Head (m)

100

80

60

40

20

0
0.0 1.5 3.0 4.5 6.0 7.5 9.0 10.5 12.0 13.5 15.0
Time (s)

Figure 4.10: Results of the third simulation: sudden closure and not negligible
headlosses.
54 WATER HAMMER SIMULATIONS

180

160

140

120
Head (m)

100

80

60

40

20

0
0 5 10 15 20 25 30
Time (s)

Figure 4.11: Results of the fourth simulation: sudden closure and negligible head-
losses; the number of computational sections is quite low and the
effects of numerical diffusion are significant.

reported in Figure 4.11. As can be seen, the angles of the square wave obtained with
the first simulation (in Figure 4.7) become smoother and the amplitude decreases;
this phenomenon is called diffusion. In this case, the phenomenon is not physical,
and it is due only by numerical errors (for this reason it is called numerical diffusion),
which will be discussed in detail in Chapter 5, being introduced here only because
the reader could find it by playing with the enclosed computer code, or with other
codes available.
Chapter 5
Numerical solution of the governing equations:
finite difference methods

More flexible methods are those based on finite differences [31]. The idea beyond
these methods is very simple, being based on the substitution of the differential
ratios with finite differences.
Obviously, while analytical solutions are defined continuously on the whole
domain, numerical solutions are defined only in a finite number of points. The
values computed on this finite number of points can be different from those which
would be carried out with an analytical solution, but they should converge when
the spatial integration step x decreases.
The unknown parameters are computed starting from those already known, i.e.,
from the values the parameters have at a former time.
To this end, these methods can be divided into explicit and implicit.
Explicit methods solve the differential equations at time t + 1 using data related
only to previous times (Figure 5.1).
In Figure 5.1, the derivatives /x and /t are computed as function of the
unknown variable values at time t + 1 and position i and of the known variable
values at time t and positions i 1, i and i + 1. The solution of the equations,
therefore, advances one point at a time for each timestep, and consequently each
equation contains only one unknown, which can be easily computed.
On the other hand, implicit methods solve the differential equations at time t + 1
using data related both to previous times and to the same time t + 1 (Figure 5.2).
As the latter data are unknown, the equations have to be solved simultaneously, and
this requires the solution of a system which is usually quite large; for this reason
implicit methods are computationally more demanding.
The advantage of the implicit methods is that they are unconditionally stable,
and that means the timestep may be larger. Explicit methods, instead, have to satisfy
the (necessary, but not sufficient) stability criterion: to this end, small timesteps have
to be chosen and, therefore, the time of simulation increases. However, in having a
too large timestep other problems may arise, because the truncation error increases
and the precision of the solution diminishes.
The approximation of a partial derivative with an algebraic ratio is possible as
it can be shown starting from the Taylor series, which allows the approximation of
any function with a polynomial.
56 WATER HAMMER SIMULATIONS
t

1 2 3 4 5 6 7

1 2 3 4 5 6 7

Figure 5.1: Explicit method example.

1 2 3 4 5 6 7
t
x

1 2 3 4 5 6 7

Figure 5.2: Implicit method example.

Let f (x), a real numerical function, defined in [a, b] and let x0 [a, b]. If f can
be derived, the Taylors series is:

 
f  2 f  (x x0 )2
f (x) = f (x0 ) + (x x ) + + ...
x x0 x2 x0
0
2!

n f  (x x0 )n
+ n +H (5.1)
x x0 n!

where H is the truncation error. The first derivative in a point is the slope of the
tangent to the curve that can be approximated by the straight line passing for x0
and x0+1 : in this case, the approximation is called forward difference; when the
points are x01 and x0 , the approximation is called backward difference. Different
combinations are possible, for instance when the derivative is approximated using
points x01 and x0+1 , the approximation is called central difference.
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 57

Expressing these concepts with formulae, the result is:



f  f (xi+1 ) f (xi )
= forward difference
x xi x

f  f (xi ) f (xi1 )
 = backward difference
x xi x

f  f (xi+1 ) f (xi1 )
= central difference
x xi 2 x
For the second-order derivative, the following expression can be used:

2 f  f (xi+1 ) 2 f (xi ) + f (xi1 )
2  =
x xi (x)2

The terms which have been neglected are called truncation error, which is a
measure of the accuracy of the solution and which diminishes when the distance
between points decreases.

5.1 The CourantFriedrichsLevy stability condition


Very often in the context of hyperbolic systems, the solutions are required for a
long duration; in these cases, a strong stability is required, in order to guarantee the
solution is limited for any value at any time. In order to have a stable solution, it is
necessary the errors due to the discretization are not amplified from each temporal
line to the following.
The stability criterion, which, as mentioned, is necessary but not sufficient, is
the following [32]:
t
= (5.2)
x
This condition is known as CourantFriedrichsLevy (CFL) and it is based on
the concept of dependency domain, which is defined as the set of points that
influence the value of the solution in a generic point (xj , tn ).
Referring to Figure 5.3, the area inside the triangle with vertices are (xj , tn ),
(xjn , t0 ) and (xj+n , t0 ) is the dependency domain. The lines drawn in the same figure
are the characteristics of the differential equations, as described in Chapter 4.
The stability condition requires the physical dependency domain to be inside
the numerical dependency domain. If this condition is not satisfied, the numerical
solution cannot converge towards the physical solution, because the change of a
condition within the physical dependency domain but not within the numerical
dependency domain would not have effect on the numerical solution.
From the mathematical point of view, thus, it is required that the spatial and
temporal steps respect the relation (5.2).
58 WATER HAMMER SIMULATIONS
(xj , tn)
tn

t0
(xjn, t0) xj (xjn, t )
0

Figure 5.3: Dependency domain.

t t
xj, tn xj, tn

0 0
x x

Figure 5.4: Physical and numerical dependency domain. On the left, the CFL con-
dition is respected, as the numerical domain contains the physical one;
on the right, the CFL condition is not satisfied.

As the characteristic lines are the following (cfr. Chapter 4):

dx
= V c c (5.3)
dt
the limitations to the timestep is given by:
 
 c t 
 
 x  1 (5.4)

This condition is shown in Figure 5.4. As mentioned, this condition may be


not sufficient to assure stability, especially when the boundary conditions are
improperly-posed.

5.2 The LaxWendroff method


This method [33] is explicit and it is widely used for the solution of linear hyper-
bolic differential equations with central differences. It is second order accurate both
in space and time [34] and it is obtained taking into account the first three terms
of the Taylors series. Being an explicit method, the CFL condition has to be
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 59

respected: in this case, it can be demonstrated that the method converges towards
the exact solution.
A linear system of differential equations can be written as:


u u
+A =0 (5.5)
t x

or, in compact form:


ut + A u x = 0 (5.6)
As mentioned, at the base of the LaxWendroff method is the Taylors series:

1
u(x, tn+1 ) = u(x, tn ) + t ut (x, tn ) + (t)2 utt (x, tn ) + (5.7)
2

Equation (5.6) yields:


ut = A ux (5.8)
and deriving again with respect to time:

utt = A uxt = A2 uxx (5.9)

Equation (5.6) implies:

uxt = utx = (A ux )x = A uxx (5.10)

Substituting these expressions of ut and utt into eqn (5.7) it yields:

1
u(x, tn+1 ) = u(x, tn ) t A ux (x, tn ) + (t)2 A2 uxx (x, tn ) + O(x3 , t 3 )
2
(5.11)
Neglecting the truncation error O(x3 , t 3 ) and approximating ux and uxx with
central differences, the LaxWendroff method is obtained, which can be written as:

2
ujn+1 = ujn A (uj+1
n
uj1
n
)+ A2 (uj+1
n
2ujn + uj1
n
) (5.12)
2 2

where j and n indicate the spatial and temporal positions, respectively, and =
t/x.
This method is defined as three-point support scheme: examining the points of
the grid which have been used in the computation of the new value, the stencil of
the method can be drawn (Figure 5.5) and, as can be seen, to compute the values in
the point ( j, n + 1) it is necessary the knowledge of the values in the nodes ( j, n),
( j + 1, n) and ( j 1, n).
60 WATER HAMMER SIMULATIONS
Time

n1

j1 j j1 Space

Figure 5.5: Stencil for LaxWendroff method.

Unfortunately, waterhammer is not a linear problem, and the system to be solved


is not precisely defined with eqn (5.5) but instead with:


u u
+ A (u) =E (5.13)
t x

However, LaxWendroff method can be extended to quasi-linear problems [35]


with the expression:

2
ujn+1 = ujn Anj (uj+1
n
uj1
n
)+ (Anj )2 (uj+1
n
2ujn + uj1
n
) (5.14)
2 2

5.3 Solving the governing equations


The method has to be applied to the equations governing the waterhammer
phenomenon, written in conservative form as follows:

p p V
+V + c2 =0 (5.15)
t x x

1 p V V J z
+ + V + + =0 (5.16)
x t x x
Substituting in eqns (5.15) and (5.16) the following expression:

p h
= (5.17)
t t


p h z
= (5.18)
x x x
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 61

Neglecting the term V z/x as it is very small, the equations become:

h h c2 V
+V + =0 (5.19)
t x g x

V h V
+g +V +J g =0 (5.20)
t x x
or, in matrix form:
       
 h   V c2 h  0 
    
 + g
  =  (5.21)
t  V   g V  x  V   g J 

The system to be solved can be finally written as:



u u
+ A(u) =E (5.22)
t x
where:
     
h V 
c2  0 
     
u=  A= g
 E=  (5.23)
V  g V  g J 
Again, the headlosses can be computed with the Chzys formula:

V |V |
J = (5.24)
2gD

The application of the LaxWendroff s formula (5.14) yields:

2
ujn+1 = ujn Anj (uj+1
n
uj1
n
)+ (Anj )2 (uj+1
n
2ujn + uj1
n
) + t Ejn
2 2
(5.25)
which can be written as:
       n 
 hn+1   hn  V c2 n hj+1 hnj1
 j   j  g
uj =  n+1  =  n 
n+1

 Vj   Vj  2 g V n
Vj+1 Vj1n
  2  n
j
  
2 V g
c2 n hj+1 2hnj + hnj1  0 n
 
+ n + t  
2 g V Vj+1 2Vj + Vj1
n n  J g 
j j

(5.26)
where:
   
 2 2V c2 
 (V + c 2
) 

A =
2 g  (5.27)

 2 V g (V 2 + c2 ) 
62 WATER HAMMER SIMULATIONS

Separating the two unknown parameters V and h, the following solving


equations are obtained:


Vjn+1 = Vjn (g (hnj+1 hnj1 ) + Vjn (Vj+1
n
Vj1
n
))
2
2
+ ((2 Vjn g) (hnj+1 2 hnj + hnj1 )
2
+ (c2 + (Vjn )2 ) (Vj+1
n
2 Vjn + Vj1
n
)) g t Jjn (5.28)


c2
hn+1
j = hnj Vjn (hnj+1 hnj1 ) + (Vj+1
n
Vj1 n
)
2 g
2 

+ ((Vjn )2 + c2 ) (hnj+1 2 hnj + hnj1 )
2
  
2 Vjn c2
+ (Vj+1 2 Vj + Vj1 )
n n n
(5.29)
g

Equations (5.28) and (5.29) are those implemented in the computer program.

5.4 Boundary conditions


As can be easily seen from the stencil of the LaxWendroff method shown in
Figure 5.5, at boundaries this method cannot be applied. In fact, upstream, to
compute the values at time n + 1 in the point j = 0, the solver should use the values
at time n in the point j = 1, j = 0 and j = 1; the point j = 1 does not exist; the
same applies for the downstream end of the pipe.
There are two main ways to overcome this problem: (1) the use of a different
scheme and (2) the definition of virtual points (called ghost cells) outside the
integration space, the characteristics of which will be discussed in section 5.4.2.

5.4.1 Asymmetrical schemes

A number of schemes useful to solve the problem at boundaries can be found


in the literature; not to be forgotten is the possibility to use the same method of
characteristics which has been presented in Chapter 4. However, as the goal of this
chapter is the use of finite differences, the very simple method called Upwind is
presented and implemented in the computer code. The stencils for the method are
shown in Figure 5.6.
According to the stencils shown in Figure 5.6, the expressions for the Upwind
method upstream and downstream are, respectively:

1
ux (xj , tn ) = [u(xj+1 , tn ) u(xj , tn )] + O(x; t) (5.30)
2 x
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 63
n-time n-time
1 1

0 0

1 1

2 1 0 1 2 2 1 0 1 2
i-space i-space

Figure 5.6: Stencil for the Upwind method. Upstream (left) and downstream
(right).

  1
ux xj , tn = [u(xx , tn ) u(xj1 , tn )] + O(x; t) (5.31)
2 x

Substituting in eqn (5.25) it yields:

A
ujn+1 = ujn (uj+1
n
ujn ) (5.32)
2

A
ujn+1 = ujn (ujn uj1
n
) (5.33)
2

Substituting in eqns (5.32) and (5.33) the expressions to be inserted in the


computer program can be obtained, as follows:

c2
hn+1
0 = hn0 [V0n (hn1 hn0 ) + (V1n V0n )] (5.34)
2 g


V0n+1 = V0n [g (hn1 hn0 ) + V0n (V1n V0n )] g t J0n (5.35)
2

c2
hn+1
n = hnn [Vnn (hnn hnn1 ) + (Vnn Vn1
n
)] (5.36)
2 g


Vnn+1 = Vnn [g (hnn hnn1 ) + Vnn (Vnn Vn1
n
)] g t Jnn (5.37)
2

Equations (5.34) and (5.35) are valid for the upstream conditions, while eqns
(5.36) and (5.37) are valid for the downstream conditions.
Obviously, more complex schemes can be developed. As example, in Figure 5.7
the BeamWarming scheme is presented.
64 WATER HAMMER SIMULATIONS
n-time n-time
1 1

0 0

1 1

2 1 0 1 2 2 1 0 1 2
i-space i-space

Figure 5.7: Stencil for the BeamWarming method. Upstream (left) and down-
stream (right).

For the BeamWarming method, the derivatives, according with the stencil, for
the upstream condition are:
1
ux (xj , tn ) = [3u(xx , tn ) 4u(xj+1 , tn ) + u(xj+2 , tn )] + O(x2 ) (5.38)
2 x

1
uxx (xj , tn ) = [u(xx , tn ) 2u(xj+1 , tn ) + u(xj+2 , tn )] + O(x) (5.39)
(x)2
while for the downstream condition are:
1
ux (xj , tn ) = [3u(xx , tn ) 4u(xj1 , tn ) + u(xj2 , tn )] + O(x2 ) (5.40)
2 x

1
uxx (xj , tn ) = [u(xx , tn ) 2u(xj1 , tn ) + u(xj2 , tn )] + O(x) (5.41)
(x)2
Again the insertion of these expressions into the eqn (5.25) yields:

2
ujn+1 = ujn + A (3ujn 4uj+1
n
+ uj+2
n
)+ A2 (ujn 2uj+1
n
+ uj+2
n
) (5.42)
2 2

2
ujn+1 = ujn A (3ujn 4uj1
n
+ uj2
n
)+ A2 (ujn 2uj1
n
+ uj2
n
) (5.43)
2 2
The equations to be inserted in the computer program can be obtained as follows:
 
 n  c2  n 
hn+1
0 = hn
0 + V n
0 3h 0 4h n
1 + h n
2 + 3V 0 4V 1
n
+ V n
2
2 g
 
2 2 V0n c2
+ ((V0 ) + c ) (h0 2h1 + h2 ) +
n 2 2 n n n
(V0 2V1 + V2 )
n n n
2 g
(5.44)
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 65

V0n+1 = V0n + [g (3hn0 4hn1 + hn2 ) + V0n (3V0n 4V1n + V2n )]
2
2
+ [2 g V0n (hn0 2hn1 + hn2 ) + ((V0n )2 + c2 ) (V0n 2V1n + V2n )]
2
g t J0n (5.45)
 
c2
hn+1
n = hnn Vnn (3hnn 4hnn1 + hnn2 ) + (3Vnn 4Vn1 n
+ Vn2
n
)
2 g

2 2 V0n c2
+ ((Vnn )2 + c2 ) (hnn 2hnn1 + hnn2 ) +
2 g

(Vnn 2Vn1
n
+ Vn2
n
) (5.46)


Vnn+1 = Vnn [g (3hnn 4hnn1 + hnn2 ) + Vnn (3Vnn 4Vn1
n
+ Vn2
n
)]
2
2
+ [2 g Vnn (hnn 2hnn1 + hnn2 ) + ((Vnn )2 + c2 )
2
(Vnn 2Vn1
n
+ Vn2
n
)] g t Jnn (5.47)

Unfortunately, this method, probably because of the presence of the second-order


derivatives, seems to be very unstable for the case of waterhammer and, therefore,
will not be implemented in the computer code.

5.4.2 Ghost cells

The idea of the method [36] is the addition of a further cell beyond the physical
domain of the problem, i.e., the numerical domain is defined for the points in the
set [1 . . . n + 1], while in the real domain the points belong to the set [0 . . . n].
The values of the parameters in the ghost cells have to be defined accordingly,
and a number of methods exists for this purpose.
A widely used method is based on internal solutions, i.e., the definition of a cell
n
Qj+1 through the values computed inside the domain, as Qjn or Qj1 n
.
The simplest approach is defined zero-order extrapolation which defines the
external parameters as:
n
Qj+1 = Qjn (5.48)

The insertion of this condition inside the computer program is very easy, as the
LaxWendroff equations can be used and simplified. For the upstream condition
66 WATER HAMMER SIMULATIONS

these become:

V0n+1 = V0n (g (hn1 hn0 ) + Vjn (V1n V0n ))
2
2
+ ((2 V0n g) (hn1 hn0 ) + (c2 + (V0n )2 ) (V1n V0n )) g t Jjn
2
(5.49)

  c 2  
hn+1
0 = hn0 V0n hn1 hn0 + V1n V0n
2 g

2 2 V0n c2
+ ((V0 ) + c ) (h1 h0 ) +
n 2 2 n n
(V1 V0 )
n n
(5.50)
2 g
while for downstream conditions the equations become:

Vnn+1 = Vnn (g (hnn hnn1 ) + Vnn (Vnn Vn1
n
))
2
2  2
+ ((2 Vnn g) (hnn + hnn1 ) + (c2 + Vnn ) (Vnn + Vn1
n
))
2
g t Jjn (5.51)
c2
hn+1
n = hnn (Vnn (hnn hnn1 ) (Vnn Vn1
n
))
2 g

2 2 Vnn c2
+ ((Vn ) + c ) (hn + hn1 ) +
n 2 2 n n
(Vn + Vn1 )
n n
2 g
(5.52)

It is obviously possible the use of the so-called first-order extrapolation based on


the filtering of the internal solutions through a linear function:
n
Qj+1 = 2 Qjn Qj1
n
(5.53)

Many schemes are available, and the best scheme, which can be success-
fully used for any conditions unfortunately does not exist. However, the methods
presented in this chapter are normally valid for most real cases.

5.5 The computer code


The code for using the LaxWendroff scheme is as simple as that shown in Chapter
4 to implement the characteristic method. The core procedure, to solve internal
points (which is the actual LaxWendroff scheme), is shown in Figure 5.8. Again,
it can be seen that this procedure uses the exact equations (5.28) and (5.29).
In order to compare the carried out results with those obtained in Chapter 4,
the same simple plant is simulated: with an upstream reservoir and a downstream
valve which closes with assigned law. The UpWind and the Ghost Cell Zero Order
to solve the boundary conditions are implemented, and the related codes are shown
in Figures 5.9 and 5.10.
NUMERICAL SOLUTION OF THE GOVERNING EQUATIONS 67

procedure ComputeInternalPoints;
var i:integer;

begin
for i:=1 to n-1 do
begin
vp[i]:=v[i]-0.5*(dt/ds)*((h[i+1]-h[i-1])*g+v[i]*(v[i+1]-v[i-1]))+
0.5*(dt/ds)*(dt/ds)*(2*g*v[i]*(h[i+1]-2*h[i]+h[i-1])+
(cel*cel+v[i]*v[i])*(v[i+1]-2*v[i]+v[i-1]))-
Lambda*v[i]*abs(v[i])*dt/(2*diam);

hp[i]:=h[i]-0.5*(dt/ds)*(v[i]*(h[i+1]-h[i-1])+(Cel*Cel/g)*(v[i+1]-v[i-1]))+
0.5*(dt/ds)*(dt/ds)*((v[i]*v[i]+Cel*Cel)*(h[i+1]-2*h[i]+h[i-1])+
(2*Cel*Cel*v[i]/g)*(v[i+1]-2*v[i]+v[i-1]));
end;
end;

Figure 5.8: Core procedure of the solution program: calculation of the unknown
(head and velocity) for the points internal to the pipe LaxWendroff
scheme.
procedure UpWind;
begin
(* Reservoir positioned upstream *)
hp[0]:=h0;
vp[0]:=v[0]-0.5*(dt/ds)*(g*(h[1]-h[0])+v[0]*(v[1]-v[0]))-
Lambda*v[0]*abs(v[0])*dt/(2*diam);

(* Valve positioned downstream *)


if t<tc then
begin
tau:=power((1-t/tc),m);
vp[n]:=v0*tau;
hp[n]:=h[n]-0.5*(dt/ds)*(v[n]*(h[n]-h[n-1])+
(cel*cel/g)*(v[n]-v[n-1]));
end else
begin
tau:=0;
vp[n]:=0;
hp[n]:= h[n]-0.5*(dt/ds)*(v[n]*(h[n]-h[n-1])+
(cel*cel/g)*(v[n]-v[n-1]));
end;
end;

Figure 5.9: Core procedure of the solution program: calculation of the unknown
(head and velocity) for the points at the boundary of the pipe UpWind
scheme.
68 WATER HAMMER SIMULATIONS

procedure GhostZero;
begin
(* Reservoir positioned upstream *)
hp[0]:=h0;
vp[0]:=v[0]-0.5*(dt/ds)*((h[1]-h[0])*g+v[0]*(v[1]-v[0]))+
0.5*(dt/ds)*(dt/ds)*(2*g*v[0]*(h[1]-h[0])+
(cel*cel+v[0]*v[0])*(v[1]-v[0]))-Lambda*v[0]*abs(v[0])*dt/(2*diam);

(* Valve positioned downstream *)


if t<tc then
begin
tau:=power((1-t/tc),m);
vp[n]:=v0*tau;
hp[n]:=h[n]-0.5*(dt/ds)*(v[n]*(h[n]-h[n-1])+
(1/g)*Cel*Cel*(v[n]-v[n-1]))+
0.5*(dt/ds)*(dt/ds)*((v[n]*v[n]+Cel*Cel)*(-h[n]+h[n-1])+
(2*Cel*Cel*v[n]/g)*(-v[n]+v[n-1]));
end else
begin
tau:=0;
vp[n]:=0;
hp[n]:=h[n]-0.5*(dt/ds)*(v[n]*(h[n]-h[n-1])+
(1/g)*Cel*Cel*(v[n]-v[n-1]))+
0.5*(dt/ds)*(dt/ds)*((v[n]*v[n]+Cel*Cel)*(-h[n]+h[n-1])+
(2*Cel*Cel*v[n]/g)*(-v[n]+v[n-1]));
end;
end;

Figure 5.10: Core procedure of the solution program: calculation of the unknown
(head and velocity) for the points at the boundary of the pipe Ghost
Cell Zero Order scheme.

5.6 Again the simple application


The simulated plant is exactly that described in section 4.4 and therefore its charac-
teristics will not be repeated here. The users interface is very similar to the previous
program, with slight differences; it is reported in Figure 5.11. As mentioned, in this
case both the spatial interval and the time step can be decided by the user, and the
computer program computes the CFL condition.
As can be seen (and the reader is invited to try), when the CFL number is higher
than 1, the code crashes and it is impossible to see the results.
However, some interesting comparisons can be performed for CFL numbers
smaller than 1. In Figure 5.12, two simulations are presented: in both cases the
time step is very small (t = 0.0008 s) but in one case the number of computational
Figure 5.11: Users interface of the program.

200
CFL  1.0
180
CFL  0.1
160

140

120
Head (m)

100

80

60

40

20

0
0.0 1.5 3.0 4.5 6.0 7.5 9.0 10.5 12.0 13.5 15.0
Time (s)

Figure 5.12: Results of the simulation of a sudden closure and negligible head-
losses, with LaxWendroff scheme, UpWind scheme for boundaries
and different CFL numbers.
70 WATER HAMMER SIMULATIONS

section is let equal to 1000 (and this implies CFL = 1.0) while in the latter case, in
order to increase the velocity of computations the section number is equal to 100,
and therefore CFL = 0.1.
As can be seen, even if the code does not crash, when CFL = 0.1 noticeable
instabilities appear and the carried out solution can be said to be less accurate,
while with CFL = 1.0, the solution is identical to that of the theoretical.
The shown simulations have been performed with the UpWind scheme for the
boundary conditions, but results carried out with the Ghost Cells are very similar.
Chapter 6
Devices Boundary conditions

In a complex plant many devices can be inserted. In this chapter few of them
are presented, and in particular those purposely studied for reducing the pressures
generated during transients, because, as it has been shown in the past chapters,
these pressures can be very high and dangerous. Moreover, a section discussing the
possibility to use a simple model for pumps is also presented.
As waterhammer is known since many decades (or centuries), a large number
of devices have been studied and implemented, and the research is still going on.
The devices presented in this chapter are the most widespread, and their discussion
and implementation in the computer programs attached to the book should allow
the reader to generalize the method and to apply it to any technical case.
These devices are modelled as boundary conditions and positioned upstream or
downstream the pipes; pipes are, therefore, solved with the methods developed for
internal points (Chapters 4 and 5) and boundary conditions are solved in a different
computer procedure, as described in the following.

6.1 Surge tanks


6.1.1 Simple surge tanks

Hydroelectric plants are normally composed by an upstream reservoir with the


intake, a tunnel where water flows with relatively low velocity (to minimize the
headlosses) and a penstock, where water flows at high velocity and the differences
in elevations are maximized, in order to retrieve the maximum power (Figure 6.1).
Between the tunnel and the penstock is normally positioned a surge tank, i.e.,
a reservoir with the aim to receive the inflowing water when the valve situated
downstream the penstock is closed and the turbines are turned off. This has already
been mentioned in Chapter 3.
When the flow in the penstock is abruptly stopped, the water flowing in the
tunnel does not stop immediately but, due to its inertia, flows into the surge tank,
increasing its level up to a maximum, that occurs when all the kinetic energy of the
water in the tunnel has been transformed, except for the headlosses, into potential
energy. After that, the flow reverses and the level in the surge tank decreases. The
process continues with a series of oscillations, damped by the headlosses of the
system.
72 WATER HAMMER SIMULATIONS
45 m 45 m

Upper expansion
A
chamber
Cross
section AA
Cross
section CC
0

1,75
2,5

0
2,5
3,44
29 m Cross
100 section DD

B
Cross
section BB
2,10 2,10
B

4
3.4 Lower expansion
C
chanber D

Tunnel C D
70 m

Figure 6.1: Hydroelectric plant the Plima-Lasa plant, North-East of Italy.

The model implemented in a computer code using the method of characteristics


is very simple, and it uses the following equations:

1. positive characteristic line (inside the tunnel);


2. negative characteristic line (inside the penstock);
3. continuity equation for the three pipes: tunnel, surge tank and penstock;
4. equality of the heads for tunnel and surge tank;
5. equality of the heads for surge tank and penstock;
6. equality of the head at the base of the surge tank and the water elevation in the
tank;
7. variation of the elevation z within the surge tank, computed as mass con-
servation, evaluating the inflow discharge Qtank = Atank z/t where Atank is
obviously the cross section of the surge tank.

Equations are, respectively:

htunnel,j htunnel1,j1 c Vtunnel,j Vtunnel1,j1


+
t g t
Vtunnel1,j1 |Vtunnel1,j1 |
+c (6.1)
2gD
DEVICES BOUNDARY CONDITIONS 73

Downstream tunnel section


Q surge tank Upstream penstock section
Surge tank base section
Tunnel section 1

Penstock section 1

Q tunnel

Q penstock

Figure 6.2: Sketch of the notation used for developing the computer code.

hpenstock,j hpenstock+1,j1 c Vpenstock,j Vpenstock+1,j1



t g t
Vpenstock+1,j1 |Vpenstock+1,j1 |
c (6.2)
2gD

Qtunnel = Qtank + Qpenstock (6.3)


htunnel = htank (6.4)
htunnel = hpenstock (6.5)
Ztank = htank (6.6)
Qtank
Z = t (6.7)
Atank

In the above reported description the subscripts tunnel, penstock and tank
have been used to indicate the same point positioned at the junction with the
three elements but related to each of the element (Figure 6.2). With the subscripts
tunnel 1 and penstock + 1 the sections in the tunnel immediately upstream and in
the penstock immediately downstream the surge tank are designated, respectively.
The above equations act as downstream boundary condition for the tunnel and
as upstream boundary condition for the penstock. These equations, together with
those for the internal points and the initial conditions, allow the solution to the
problem.
74 WATER HAMMER SIMULATIONS

It is to be observed that no characteristic equation is written for the surge tank,


for which eqn (6.6) is written instead: this implies the static pressure distribution
and the non-elastic behaviour of the water in the surge tank.
As can be seen in the example reported in Figure 6.1 and in other examples
that can be found in the literature, these devices are very large and, therefore,
are very expensive: any reduction of these dimensions is appreciated. To this end,
it is possible to shape the base of the surge tank in order to generate a further
(concentrated) headloss, which is active only during transients and it has the purpose
to reduce the oscillation amplitudes (restricted orifice).
The effect of this orifice can be modelled with a slight change in eqn (6.4), with
the more general:
htank = htunnel + k Vtank |Vtank | (6.4 )
The coefficient k is chosen to represent the effect of the restricted orifice,
noting that during steady flow (Vtank = 0) and when the orifice is not present
(k = 0) htank = htunnel [37].
At the beginning of transients, the headlosses due to the orifice produce an high
value of pressure at the base of the surge tank, which fosters the deceleration of the
velocity in the tunnel and, therefore, damps the oscillations in the tank.
If the value of k is very low, the effect of the restricted orifice is negligible; on
the other hand, if it is too high, the pressures at the beginning of the transient are
very high: this reduces the oscillations inside the surge tank and also its effect in
damping the pressure peaks in the tunnel.
It is, therefore, possible to devise an optimal dimension of the restricted orifice,
so that the pressures reach immediately their maximum value, and they stabilize
at that level for the whole duration of the transient. In other words, the tunnel is
subjected for the whole transient duration to the maximum value of pressure, which
is slightly smaller than the pressure it would reach without this added headloss and,
therefore, it reduces the oscillations inside the surge tank.
To design a simple surge tank, the Thomas stability condition [3840] has to be
satisfied, that, for small oscillations, the minimum cross sectional area of the surge
tank is:
Atunnel Ltunnel
Amin tank = (6.8)
2 g (Htunnel /Vtunnel
2
) Hnet
where Htunnel is the headloss in the tunnel, Hnet is the difference in elevation (so
neglecting the headlosses) between the water elevation in the surge tank in static
conditions and the elevation of the downstream turbine. Normally, this minimum
value is multiplied for a safety factor that can vary between 1.2 and 1.5.

6.1.2 Different types of surge tanks

As mentioned above, dimensions of surge tanks are very large. Moreover, in the
practical engineering other targets are required, as the reduction of the oscillation
period and the duration of the transient; furthermore, if a surge tank is connected to
DEVICES BOUNDARY CONDITIONS 75

Riser
Upper expansion chamber

Lower expansion chamber

Tunnel Penstock

Figure 6.3: Sketch of a differential surge tank.

a reservoir which water level is not constant but shows wide variations, this surge
tank would be oversized for most of the period of use when the water level in the
reservoir is low.
To overcome these drawbacks various types of surge tanks have been studied
[41, 42]. In this section the basic types are described, but the creativity of designers
has suggested several others, according to the needs of the specific projects.
One of the widespread types is built with expansion chambers, as sketched in
Figure 6.3.
In general, the higher expansion is kept above the highest level of water elevation
in the reservoir, while the lower expansion is built below the level of water obtained
with maximum discharge and lowest water elevation inside the upstream reservoir.
Under these conditions, during steady flow, the water surface is always contained
within the riser of the surge tank.
If, during transients, water is contained in the riser, the surge tank behaves
as in the simple case (Section 6.1.1). When, instead, the expansion chambers are
also affected by the motion, the level in the tank rises (or lower) quickly until it is
contained in the riser, to rise (or lower) much more slowly when water reaches one
of the expansions. With these surge tanks, then, an high (or low) pressure quickly
establish at the base of the riser while the expansions limit the amplitude of the
oscillations. This system, surely advantageous, may introduce complications to the
analysis, as waves can be produced in the galleries during emptying and filling [43].
In order to slightly increase the pressures at the base of the riser, sometime it is
continued beyond the bottom of the chamber, an overflow is created, as sketched in
Figure 6.4.
A similar device has been studied by Kammuller [44] to take advantage of the
effect of rapid lowering of the level in the riser; the device is sketched in Figure 6.5.
76 WATER HAMMER SIMULATIONS
Upper expansion chamber

Riser

Overflow

Figure 6.4: Overflow in the upper expansion chamber of the surge tank.

Riser

Lower expansion chamber c

o o

Tunnel Penstock

Figure 6.5: Lower expansion chamber with Kammulers device.

Because of this device, the lower expansion chamber starts to empty only when the
water level is below the section 00, i.e., when from the conduit c the air can enter
the chamber. The pipe a, much smaller than pipe c, serves only to expel the air
when the chamber fills.
A different idea to reach the same target is the use of a differential surge tank,
or Johnson type [45, 46], equipped with an optimal orifice for the inflowing water,
but also with an internal riser that overflows in the outer tank when the maximum
water level is reached. This type of surge tank is sketched in Figure 6.6.
When the valve positioned downstream the penstock opens, and, therefore,
water starts flowing in the penstock, the decrease of the pressure under the orifice
is equal to the maximum lowering of the water level in the riser, as the orifice is
optimal for this operation.
DEVICES BOUNDARY CONDITIONS 77

Orifice

Q1

Q2
Q Penstock
Q Tunnel

Figure 6.6: Differential surge tank, or Johnsons type.

When the valve closes, water stops in the penstock and flows into the surge
tank; the orifice is small and, without the internal riser, pressures at the base of the
surge tank would be too high. The velocity inside the riser (related to discharge
Q1 ) is high and the water surface reaches the maximum level when the level in the
outside tank (given by discharge Q2 ) is still very low; when the water level reaches
the top of the riser, it overflows in the outside tank thus stabilizing the value of the
pressure.
The surge tank is, therefore, optimized for both the operations of opening and
closing of the plant.

6.2 Air chambers


These devices are the most common in pumping stations, where an abrupt stoppage
of the flow induces in the downstream pipe low pressure values that can have
significant impact on the system. These very low pressures, in the worst cases,
produce cavitation, an undesirable phenomenon that will be described later in the
book.
Air chambers are closed reservoirs, positioned downstream the pump, filled
by both water and pressurized air [47, 48]. When the pump stops, due to this
device the flow in the downstream pipe continues, fed by the air chamber, and its
velocity slowly decreases, therefore limiting the waterhammer effects. During the
transient, the volume of air increases and its pressure decreases until it reaches a
minimum; afterwards the flow inside the pipe reverses, and in this second phase,
the air volume decreases and the pressure increases up to its maximum value.
The process progresses with a series of oscillations, damped by headlosses.
78 WATER HAMMER SIMULATIONS

Air chamber

Valve Section 0

Figure 6.7: Sketch of an air chamber.

Variables that govern the boundary conditions are the head in the air chamber
h (or the pressure p), the velocity at the inlet of the chamber Vs and the volume of
air in the chamber itself W .
Equations that solve the system (in addition to the usual characteristic equation)
are the mass conservation for the system air chamber-pipe and the polytropic law
that describes the behaviour of the air inside the chamber.
Considering the air chamber positioned downstream, the pumps or the valve
which suddently stops the incoming discharge, and therefore upstream the subse-
quent pipe, i.e., in the section 0 of the computational domain (see Figure 6.7) the
negative characteristic equation is:

c V1,t |V1,t |
h0,t+1 h1,t (V0,t+1 V1,t ) c t = 0 (6.9)
g 2gD

The continuity equation for the air inside the chamber is written as:

Wt+1 = Wt + A Vo,t t (6.10)

The polytropic equation that describe the behaviour of air is:

ho,t W0,t
m
= c.te (6.11)

where m depends on the polytropic process and can vary between 1.0 and 1.4 for
the isothermal and adiabatic (considering dry air at 20 C) processes, respectively
[49]; this parameter is normally let equal to 1.4, the differences in the carried out
results being very limited and this value being on the safe side.
Without the additional headloss, as those described for surge tanks in Section
6.1, the assumption is that the head in the first section of the downstream pipe is
equal to that inside the chamber. Actually, even if it is possible to design orifices
for these devices as well as for the surge tank, they are not normally used, as
the economic saving is negligible and an error in design can worse the normal
DEVICES BOUNDARY CONDITIONS 79

conditions. However, as will be seen in Section 6.4.3, the implementation of an


headloss model is very easy.
In the thirties of the twentieth century, Evangelisti [50, 51] devised a graphical
method to dimension air chambers, but only under the hypotheses of no headlosses
and non-elastic behaviour of the system. With numerical models as those presented
in this book, the dimensioning of these devices is carried out by trial and errors,
selecting a volume for the air chamber and numerically testing its effect in the
system.
With regard to the installation of air chambers, it is to be observed that they
are normally equipped with a compressor in order to re-establish the mass of air
within the chamber, as, because of the high pressures, air tends to dissolve in
water and then to be convoyed away. For some installations there is the possibility
of having the aeriform contained in a pouch which isolates it from the liquid, so
that no dispersions occurs and, therefore, no compressor is required. However, the
possibility to recharge that pouch should always be granted; a possible drawback
of this system is given by the life expectancy of the pouch.

6.3 Relief valves and rupture disks


Let us consider again the simple plant of Figure 2.1. As already mentioned, when the
downstream valve closes, the pressure increases up to a value than can be computed
with eqn (2.1) or (2.9) if the closure is abrupt or slow, respectively.
Relief valves work with the principle sketched in Figure 6.8. The valve is closed
by a spring, which is calibrated to keep the valve closed under static and steady flow
conditions, with a tolerance of approximately 5%. When, during a transient, the
pressure increases and exceeds the calibrated value of the spring, the valve opens
and water outflows.
The discharge outflowing from the valve can be computed with:
"
Q = (x) A 2gH (6.12)

where x is the distance between the disk and the outflow of the valve. Obviously,
if that distance is zero, the disk prevents the discharge to outflow and, therefore,
(x) = 0; for small distances, Cozzo [52] experimentally carried out the values of
the efflux coefficient . This coefficient is affected from the disk because of the
turbulence and its value cannot be determined theoretically.
The use of these typology of valves requires attention because of the possibility
of their resonant behaviour. Tanda and Zampaglione [53] presented a real case where
the installation of these valves rose instabilities in the system: they oscillated with a
frequency close to the natural frequency of the plant, inducing resonance. In other
words, not only the pressures did not decrease, but they actually increased, with the
risk of driving the system to the collapse. The plant the authors studied is sketched in
Figure 6.9: at the end of each of the four branches, a regulation valve is positioned;
immediately upstream of the latter, in order to reduce the pressures generated during
80 WATER HAMMER SIMULATIONS

Relief valve

Valve

Figure 6.8: Sketch of a relief valve.

Regulation valves

1 12 8

2 13 9

3 14 10

5 6 7 4 15 11

Relief valves

Figure 6.9: Plant studied by Tanda and Zapaglione [53] which showed instabilities
when the anti-waterhammer valves were installed.

transients, not only one, but two relief valves were positioned. This decision was
taken because of an error in the dimensioning of the relief valves, which were not
sufficient to reduce the pressures as it was desired. However, the positioning of two
valves for each branch was probably the cause of the generated instabilities.
Some authors [19] point out that the discharge capacity of the valve should not
be too large, because if that discharge is insufficient to keep the valve fully open,
then hammering and oscillation may occur. Manufacturers develop valves that use
damping mechanisms which allow a quick opening, as requested in order to control
the phenomenon, but a slow closure, to eliminate any risk of resonance.
Bianchi and Mambretti [54] showed that the use of different valve typology
which adopts the same principle (they open when a given pressure is exceeded)
may solve the problem. The valve studied in the mentioned paper (which will be
presented in Section 6.5.5), in fact, opens quite quickly, being its velocity similar
to the relief valves; the difference with the latter is that once it is completely open,
it remains in this position until the transient has ended.
The use of rupture disks is now quite widespread, which work using the same
principle: they open (break) when the threshold pressure is reached, to allow the
DEVICES BOUNDARY CONDITIONS 81

water outflow. Also in this case instabilities are not possible, as obviously the disk
does not close again at the end of the transient; it is also obvious that the installation
of these devices should be designed only for emergencies, as they require their
manual substitution and that means the plant has to be stopped, at least partially.
Moreover, when the rupture threshold is close to the working pressure, fatigue has
been detected and after some time from the installation the disks can leak.

6.4 Centrifugal pumps


Pumps have always been part of complex plants and therefore a large number
of models have been implemented to describe their behaviour. In this book only
centrifugal pumps are covered, as they are the most commonly installed pumps; they
produce transients only when they change the pumped discharge. Other pumps,
as reciprocating pumps or even the hydraulic ram (see Chapter 1) can produce
waterhammer as by-product of their operations.
The first and simplest idea was that the pump inertia is so low that the time
of stopping of the pumps is negligible, and therefore no models are needed. For
instance, all the charts prepared in the first mid of twentieth century to design air
chambers adopted this hypothesis, which is surely on the safe side, but that may be
too demanding.
Therefore, during the years many methods for pump modelling have been pre-
sented in the scientific literature. In this paragraph, according with the aim of the
book, not all these models are covered, but only a simple model, quite widespread,
is shown, discussed and implemented in a computer code.
When a pump is stopped, the rotor slows down with a trend that can be computed
equalling the kinetic energy variation of the rotor and the power of the hydraulic
flow [10, 19]:

d I 2 Q H
= (6.13)
dt 2

where I is the rotor inertia, that can be computed by:

G D2
I= (6.14)
4g

G being the weight of the rotor and D their inertial diameter. The value of G D2
is normally provided by the companies producing the pumps, but there have been
studies about its computation [55]. In (6.13), is the angular velocity of the pump,
is the efficiency and is the specific weight of the fluid. As usual, Q is the
discharge and H is the head of the pump.
Equation (6.13) and the characteristic equation of the system allow the solution
of the problem.
82 WATER HAMMER SIMULATIONS

Actually, eqn (6.13) can be generalized to model the starting phase of the
pump [56]:
d I 2 Q H
= P (6.13 )
dt 2
where P is the power provided to the pump. When the pump is stopping, P = 0 and
eqn (6.13 ) is obviously equal to (6.13). In steady flow,

Q H
=P

that yields to
d I 2
=0
dt 2
that means the angular velocity is constant. The problem in modelling the starting
phase of the pump is the determination of the variation of the power P when the
pump starts, which is normally not known and that makes difficult the construction
of a reliable model. However, differently from a pump trip that cannot be controlled,
the start-up of the pump can always be planned, and it should be to avoid undesiered
surges. It is quite common nowadays to provide the pumps with electrical devices
able to variate their speed in steady flow conditions and that are also able to soft
start them at the beginning of the operations. Cheaper devices are able to control
only the start up of the pumps (the so-called soft-starter) but not to change their
velocity in normal operations. In any case, the manufacturers are able to produce
a large variety of methods and devices to ramp up the pumps and the value of P
in eqn (6.13 ) should be modelled accordingly.
A further difficulty is given by the incomplete knowledge of the characteristic
curves of the pumps when they are not in steady flow conditions. Therefore, when
these curves are not provided by the producers of the pumps, the modeller is forced
to accept the similitude hypotheses [57]:

Q1 n1
= (6.15)
Q2 n2
2
H1 n1
= (6.16)
H2 n2

where n is the rotations per minute of the pump, tied to the angular velocity
by the:
2 n
= (6.17)
60
Equations (6.15) and (6.16) are acceptable only when the velocity of the pump is
not too far from that of the steady flow, and therefore their use is suggested only
when no better information are available.
DEVICES BOUNDARY CONDITIONS 83

From (6.13 ), as the inertia I is constant, the variation of the rotation with time
dn/dt is equal to:
I 2 dn Q H
n = P (6.18)
900 dt
and, therefore, the pump rotation at time t + 1 can be computed as:

Q H 900
nt+1 = nt + P dt (6.19)
nt I 2

The curve of the pump is normally given as second-order polynomial, i.e., in the
form:
h = a V2 + b V + c (6.20)
where h is the head of the fluid and V its velocity. The parameters a, b and c charac-
terize the behaviour of the pump. The problem is how to scale these parameters to
produce new curves to be used in unsteady flow in accordance with the similitude
laws (6.15) and (6.16).
The following assumptions are needed to produce the required rules [56, 58].

Assumption 1. The curve is scaled so that when V = 0 it results ht+1 = ct+1 =



nt+1 2
ht ; as a consequence:
nt
2
nt+1
ct+1 = ct (6.21)
nt

Assumption 2. The derivatives of the curves must be equal when V = 0; as a


consequence:

nt+1
bt+1 = bt (6.22)
nt

 
nt+1
Assumption 3. When vt+1 = vt nt
it must be ht+1 = 0, and therefore:

at+1 = a = c te (6.23)

An example of the curve scaling is shown in Figure 6.10.

This model can be simply inserted in a computer code that solves the internal
points with one of the methods presented in Chapters 4 and 5; the pump is considered
as a boundary condition that can be positioned upstream or downstream, even if
the former is the normal position for a pump.
84 WATER HAMMER SIMULATIONS

n  100%
1.0 n  80%
n  50%
0.8
H/Hmax

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0
V/Vmax

Figure 6.10: Curves of the pump for different rotations.

Assuming the pump is positioned upstream, the numerical solution is carried


out coupling the negative characteristics and the equation of the pump:

c Vt,1 |Vt,1 |
ht+1,0 ht,1 (Vt+1,0 Vt,1 ) c t = 0 (6.24)
g 2gD
at+1 Vt+1,0 |Vt+1,0 | + bt+1 Vt+1,0 + ct+1 = ht+1,0 (6.25)

where at+1 , bt+1 and ct+1 are computed by eqns (6.21)(6.23).


Equation (6.19) shows that a larger pump inertia implies a slower stoppage of
the pump itself. This suggests the idea to artificially increase the pump inertia in
order to lengthen the time of pump stopping and therefore reducing (according to
Michauds formula) the pressures related to transients [59]. This target is normally
reached designing flywheels, i.e., metal disks rotating with the mobile part of
the pump, at the same velocity. This solution is surely feasible, especially when
the pumps are continuously working. Limitations are given by the capability of the
motor and of the equipments and, for large installations, by the increased loads
and vibrations. Moreover, when pumps are often turned on and off, the drive shaft
might result over-stressed by the increased mass of the rotor. However, the actual
effectiveness of this solution will be discussed in Section 6.5.6.

6.5 Other methods for controlling the pressures


The reader can find in the literature more methods to try to reduce the pressures
generated during the transients, but, as the parameters that govern the phenomenon
are the difference in the initial and final velocities and the celerity of the wave, these
are also the ways in which actions can be taken.
DEVICES BOUNDARY CONDITIONS 85

To reduce the wave celerity, it is possible to reduce the density of the fluid
or to increase the deformability of the pipe. The reduction of the density of the
fluid can be obtained bleeding air in the liquid: but probably this method produces
more disadvantages than advantages and it is better to avoid it. Ramenieras [60]
developed a device which consisted a small flexible hose in the pipe where the air
is trapped, so that the effective bulk elasticity modulus of the system is reducd.
Ghilardi and Paoletti [61] studied the possibility to add a flexible viscoelastic
pipe to be used as pressure surge suppressor; this pipe deforms during transients
so that the celerity and the consequent pressures decrease. The exact behaviour
of this added pipe is still under investigation [62, 63]. Moreover, it is still to be
investigated how the repeated pressure fluctuations influence the life expectancies
of these pipes.

6.6 The computer codes

Again the computer codes are based on the method of characteristics, and only
the boundary conditions are changed in order to model the devices presented in the
chapter. The routine for solving internal points is presented in Chapter 4 together
with those for solving the boundary conditions related to the reservoir and the valve.

6.6.1 The simple surge tank

Boundary conditions for the simple surge tanks are those described in Section 6.1.1
and in the computer program these are reported as shown in Figure 6.11. As can be
seen with very simple calculations, the equations have been solved algebraically

SA:=f1*cel1*v[n1-1]*abs(v[n1-1])/(2*g*d1)*dt;
SB:=f2*cel2*v[n1+1]*abs(v[n1+1])/(2*g*d2)*dt;
SC:=h[n1-1]+cel1*TankArea*ZTank/(g*Area1*dt)+
+cel1*Area2*h[n1+1]/(Area1*cel2)+cel1*Area2*SB/(Area1*cel2)+
-cel1*Area2*V[n1+1]/(g*Area1)+cel1*V[n1-1]/g-SA;
SD:=1+cel1*TankArea/(g*Area1*dt)+Area2/Area1;

ZpTank:=SC/SD;
hpd:=ZpTank;
hps:=ZpTank;

VpTank:=(ZpTank-ZTank)/dt;
vpd:=(ZpTank-h[n1+1]-SB)*g/cel2+V[n1+1];
vps:=((ZpTank-ZTank)*TankArea/dt+Vpd*Area2)/Area1;

Figure 6.11: Boundary conditions for the solution of a simple surge tank.
86 WATER HAMMER SIMULATIONS

and then implemented. Four service variables (SA, SB, SC and SD) have been used
to ease the formulas.
A simple example can be studied, still referring to a plant with an upstream
reservoir, a valve downstream and a simple surge tank to divide the tunnel and the
penstock.
The characteristics of the plant are as follows. The discharge in steady flow
conditions is Q0 = 15 m3 /s and the head of the upstream reservoir is 100 m.
The tunnel has a diameter of Dtunnel = 2.80 m, length Ltunnel = 4200 m,
Mannings roughness n = 0.0143 s m1/3 , celerity ctunnel = 1000 m/s.
The characteristics of the penstock are: diameter equal to Dpenstock = 1.90 m,
length Lpenstock = 1400 m, Mannings roughness n = 0.0111 s m1/3 , celerity
cpenstock = 1000 m/s.
The surge tank has no orifice and it is simple in type; its diameter is
Dtank = 11.50 m and therefore the area of the surge tank is Atank
= 104 m2 .
The downstream valve closes instantaneously.
The steady flow water velocity in the penstock is equal to 5.29 m/s, the phase
time is 2.8 s and therefore to compute the maximum pressure, the Michauds formula
(Chapter 2) has to be used, which yields:

2 L V0 2 1400 5.29
h = = = 302 m
g Tc 9.806 5

Results from the computer programs are saved in the files heads.txt and velocities.txt
in a simple text format that can be imported in any spreadsheet.
The plot of the heads computed at the downstream section, where the valve
is positioned, are shown in Figure 6.12, and a very good agreement can be
observed with the Michauds result. The oscillations in the surge tank are reported
in Figure 6.12. Results are reported in two different figures because of the large
differences between the two phenomena, both in terms of pressure values and in
terms of time.
As can be seen, the high pressures generated inside the penstock are strongly
damped by the surge tank, and in the tunnel only the pressures given by the
oscillations within the surge tank propagate.
The results presented in Figure 6.13 are those already shown in Chapter 3, when
this solution, based on elastic hypothesis, was compared with the solution carried
out with the simplified hypothesis of anelastic fluid.

6.6.2 The simple air chamber

Boundary conditions reported in Section 6.2 are quite simple, but in this section
a number of models are implemented, with increasing complexity. The simplest
model is mentioned in section 6.2 and its implementation is shown in Figure 6.14;
to call this routine in the computer program attached to the book the option Simple
Air Chamber has to be selected. With the same program is also possible to simulate
DEVICES BOUNDARY CONDITIONS 87
400

350

300

250
Head (m)

200

150

100

50

0
0 10 20 30 40 50 60
Time (s)

Figure 6.12: Results at the downstream valve.

110
108
106
Water elevation (m)

104
102
100
98
96
94
92
90
0 100 200 300 400 500 600
Time (s)

Figure 6.13: Oscillations of the water in the surge tank.

Wtp:=Wt+Area1*V[0]*dt;
hp[0]:=h[0]*power(Wt/Wtp,m);
Vp[0]:=V[1]+g*(hp[0]-h[1])+V[1]*abs(V[1])*dt/(2*d1);

Figure 6.14: Boundary conditions for the solution of an air chamber.


88 WATER HAMMER SIMULATIONS
180

160

140

120
Head (m)

100

80

60

40

20

0
0 15 30 45 60 75 90 105 120 135 150
Time (s)

Figure 6.15: Results of an instantaneous valve closure in a plant without the air
chamber.

the closure of an upstream valve without air chamber, and therefore comparisons
are possible.
For instance, let us consider the case of a pipe with diameter of 0.3 m, length
1000 m, Mannings roughness coefficient 0.0125 s m1/3 and celerity 1000 m/s.
The steady flow discharge is equal to 0.050 m3 /s and downstream is positioned a
reservoir and the fixed head is equal to 100 m. Upstream there is a valve which
closes completely and instantaneously.
The results computed in the upstream section without the air chamber
(Figure 6.15) and with an air chamber with 1 m3 volume (Figure 6.16) are shown,
with the same scale on the xy axes to allow a better comparison.
As can be seen the air chamber has the double effect of reducing the pressures
generated during the transient and of increasing the time of the oscillations.

6.6.3 Air chamber with headlosses

As mentioned above, this very simple boundary condition can be complicated by


inserting an headloss in the passage from the chamber to the conduit. In Section
6.2 it has also been mentioned that, in the real world, it is not usual to insert a
valve in the air chambers which has the same aim of an orifice in the surge tank;
however, it has been implemented and the results are commented, as it is a quite
simple improvement, it is useful for understanding the behaviour of such a device
and allows an interesting discussion.
DEVICES BOUNDARY CONDITIONS 89
180

160

140

120
Head (m)

100

80

60

40

20

0
0 15 30 45 60 75 90 105 120 135 150
Time (s)

Figure 6.16: Results of an instantaneous valve closure in a plant with an air chamber
of 1 m3 volume.

Wtp:=Wt+Area1*V[0]*dt;
hpChamber:=hChamber*power(Wt/Wtp,m);
hp[0]:=hpChamber-KOrifice*V[0]*abs(V[0]);
Vp[0]:=V[1]+g*(hp[0]-h[1])/cel1-f1*V[1]*abs(V[1])*dt/(2*d1);

Figure 6.17: Boundary conditions for the solution of an air chamber with head-
losses (orifice or valve).

According to the description of an orifice, the difference between the heads


inside the chamber (hChamber ) and in the pipe (hPipe ) are given by the headlosses,
so that:

hChamber = hPipe + k V |V |

The constant k has to be appropriately selected and depends on the geometry of


the orifice (or valve). The code is shown in Figure 6.17 and the results for different
values of the parameter k are shown in Figure 6.18, for the section where the valve
is positioned (section 0).
As it can be seen (Figure 6.18), the increase of the parameter k produces a sudden
diminution of the pressure in the section 0, followed by the usual oscillations which
have a reduced amplitude.
90 WATER HAMMER SIMULATIONS
120
K=0
115 K=5
K = 10
K = 20
110

105
Head (m)

100

95

90

85

80
0 5 10 15 20 25 30
Time (s)

Figure 6.18: Results of an instantaneous valve closure in a plant with an air cham-
ber of 1 m3 volume for different values of the parameter K for the
headlosses.

However, the very simple code here presented shows to be unstable, as can be
seen when looking at the data related to k = 20. This result is very interesting and
deserves a deeper analysis.
At the beginning of the transient, because of the concentrated headloss, a low
value of pressure at the base of the air chamber is generated; this consequently
reduces the flow velocity and, as the headlosses depend on this latter parameter,
the pressure at the base of the air chamber increases again. This rapid variation of
the pressure values propagates downstream, producing instabilities. The method
of characteristics, in fact, is very stable by itself, but it is strongly affected by
instabilities coming from the boundary conditions and this is precisely the case.
Actually, the instability is surely numerical, as it is not known to the author
any literature case reporting instabilities in the use of air chambers. However, the
reduction of the time step do not produce any improvements. The solution, in this
case, would be the change of the integration method and the insertion of limiters,
as will be discussed in Chapter 7, where the problem will be taken up again and
solved.

6.6.4 Air chamber and valve

Finally, it is possible to study the case of a valve that closes in finite time and it is
positioned together with an air chamber. Equations in this case are the following:

1. Negative characteristic equation (6.9);


2. Head in the pipe at section 0 is equal to head at the valve;
DEVICES BOUNDARY CONDITIONS 91

if t<tc then
begin
tau:=power((1-t/tc),m);
vpChamber:=v01*tau;
end else
begin
tau:=0;
vpChamber:=0;
end;

SA:=-h[1]+cel1*v[1]/g-cel1*f1*v[1]*abs(v[1])*dt/(2*g*d1);
SB:=Wt+QChamber*dt/2;
SC:=-cel1*vpChamber/g+SA;
SD:=cel1*dt/(2*g*Area1);
SE:=cel1*SB/(g*Area1)-SC*dt/2;
SF:=-h[0]*Wt-SC*SB;

QpChamber:=0;
Wtp:=Wt+(QpChamber+QChamber)*dt/2;
vp[0]:=vpChamber+QpChamber/Area1;
hp[0]:=h[1]+cel1*vp[0]/g-cel1*v[1]/g+cel1*f1*v[1]*abs(v[1])*dt/(2*g*d1);

Figure 6.19: Boundary conditions for the solution of an air chamber when the
upstream valve is not instantaneously closed.

3. Head in the pipe at section 0 is equal to head in the chamber;


4. Continuity equation for the node;
5. Continuity equation for the chamber (6.10);
6. Polytropic equation (6.11).

As can be seen with very simple calculations, the equations can be solved
algebraically and they have been implemented in the computer code. Four service
variables (SA, SB, SC and SD) have been used to ease the formulas. The computer
program is shown in Figure 6.19; the results carried out for different time of closure
of the valve are reported in Figure 6.20.

6.6.5 Valve modelling: an example

The description of all the valves available on the market is obviously impossible,
thus in the following only one analysis is reported in order to show how this problem
may be faced. The complete analysis is reported in [54].
92 WATER HAMMER SIMULATIONS
120

100

80
Head (m)

60

40

20
Tc  0 s
Tc  10 s
0
0 15 30 45 60 75 90 105 120 135 150
Time (s)

Figure 6.20: Effects of time closures of a valve in a plant with an air chamber with
volume equal to 1 m3 .

The valve is sketched in Figure 6.21 and it is set to work as relief valve. It has
two compartments: the lower (I) is positioned in direct contact with the pipe where
the discharge flows and the upper (II) has the function of controlling the opening.
As relief valve, it should be installed on a derivation of the main pipe in order to
put in communication, when it opens, the water in the pipe with the atmosphere,
thus reducing the pressures.
The lower compartment is divided by a breechblock (1) in two sections: down-
stream (a) and upstream (b). The former is directly connected to the atmosphere so
that, during the efflux, here the pressure is atmospheric but for negligible headlosses.
The upper compartment is also divided in two sections: the former (c) above
the diaphragm and the latter (d) between the diaphragm and the body of the valve.
The diaphragm of the upper compartment and the breechblock of the lower
compartment are connected by a rigid rod (2); this connection is not sealed and,
therefore, the same pressure can be found in the sections (a) and (d).
Actually, the valve is for multipurpose and its use depends on the device called
pilot (4) that controls its functioning. To use the valve to reduce waterham-
mer effects, the pilot has to put in communication, either the main pipe or the
atmosphere.
During steady flow, the upper section is in contact with the main pipe. The pres-
sure in the section is therefore equal to that of the main pipe; the same value of
pressure is also in the lower part of the diaphragm (3) and on the upper surface
of the breechblock (1) because of the mentioned hydraulic continuity between the
DEVICES BOUNDARY CONDITIONS 93

(4)
II

c d
(3) II
a

(2) (1)

Diversion to the
relief valve

Main pipe

Figure 6.21: Sketch of the valve modelled.

sections (a) and (d), while on the lower surface of the breechblock is the atmo-
spheric pressure. The valve is closed due to the force directed downwards which is
applied on the breechblock.
When in the pipe the pressure increases, for instance because a downstream
valve closure that starts a transient, the pilot, which has minimum inertia, impose
the almost instantaneous interruption of the communication between the section (c)
and the main pipe, connecting the section (c) with the atmosphere. In this condition,
the force on the upper surface of the diaphragm is very small as it is the force applied
on the lower surface of the breechblock (1); on the contrary, the force acting on the
lower surface of the diaphragm and directed upwards is larger than the force acting
on the upper part of the breechblock (1) and directed downwards because of the
larger area of the diaphragm (3). Therefore the valve opens.
The time required to the valve to be completely open is not negligible because
the upper section of the valve is in communication with the atmosphere through a
94 WATER HAMMER SIMULATIONS

Hd
Qd

X1 P3A2 Finertia

Diaphragm

P2A2

P2A1
Breechblock

Figure 6.22: Installation of the valve and symbols for modelling.

very small pipe, which creates a resistance to the emptying of this volume. As a
consequence, the opening of this valve is not as fast as those purposely designed
to this aim.
On the other end, its closure is remotely controlled and therefore oscillations
(and the possible instabilities of the system, mentioned in Section 6.3) are avoided.
The regulation of the opening pressure threshold is obtained by acting on a
spring of the pilot.
The unknown parameters are the head hc and the discharge outflowing Qc in the
upper section of the upper compartment, and the displacement x of the connection,
i.e., the entity made by the diaphragm, the breechblock and the rod which connects
the two.
To solve the problem, equations are (i) the dynamic equilibrium of the connec-
tion, (ii) the continuity equation for the upper section of the upper compartment
and (iii) the equation that ties the outflowing discharge Qc and the internal head hc .
With the symbols used in Figure 6.22, the equations above-mentioned are:

2 (p1 + p) A1 + p2 A2 = p2 A1 + p3 A2 + FInertia


"
Qc = oref Aoref 2 g hc
Qc
x = x + dt
A2
DEVICES BOUNDARY CONDITIONS 95

In the equations:

p2 A2 is the force acting under the diaphragm of area A2 and tends to open the
valve;
p2 A1 is the force acting on the upper surface of the diaphragm and tends to close
the valve;
p3 A2 is the force due to the pressure p3 , which corresponds the head hc , acting
on the upper surface of the diaphragm and tends to close the valve;
2 (p1 + p) is the force due to the pressure p1 + p acting on the lower surface
of the breechblock during transients (the meaning and values of the parameter
will be discussed in the following);
FInertia is the inertia force, obtained multiplying the mass of the entity made by
the diaphragm, the breechblock and the rod which connects the two by its
acceleration;
oref , Aoref are the efflux coefficients and the area of the very small pipe (diameter
equal to 5 mm) that connects the upper section and the atmosphere.

It is to be considered that p2 = 0 as it is the atmosphere.


These equations allow to compute the value of the head in the section (c)
and consequently the displacement of the diaphragm and the breechblock, i.e.,
of the opening of the valve. Once this opening is known, it is possible to compute
the effluent discharge by means of the efflux law.
Coefficients and oref have to be determined through laboratory tests. Coef-
ficient is used to reduce the force on the diaphragm in dependence of the
distance between the diaphragm itself and the outlet. It has been estimated in
steady flow conditions measuring the force exerted by the jet on the diaphragm
at different degrees of opening of the valve. Similar tests have been carried out for
the determination of the coefficient oref , and finding it is also a function of the
distance x.
Numerical tests have been performed on the three plants shown in Figure 6.23,
with different configurations and characteristics, for details refer [54].
An example of the carried out results is reported in Figure 6.24. Generally
speaking, the efficacy of these valves when they are installed in the plants has been
demonstrated. However, their effect is limited when a very fast response is required;
for instance, when the downstream valve closes very fast or when the phase of the
plant is very short (low lengths, high celerity, etc.) or even when the required
reduction is great.
The analysis performed to characterize the valve allowed the authors to build
a numerical model of the valve itself, which has been inserted in a number of
plants that could not be studied on physical models and that can be used for
other simulations. Obviously, all results have to be intended as general orientation
for understanding the plant and, moreover, many features cannot be theoretically
derived, and experimental tests are very often required, especially when the systems
are very complex.
96 WATER HAMMER SIMULATIONS

Relief valve Regulation valve

Regulation valves

1 12 8

2 13 9

3 14 10

5 6 7 4 15 11

Relief valves

Figure 6.23: Plants numerically analysed for determining the characteristics of the
valve.

6.6.6 Pumps

In the computer code just turning off the pump has been implemented and, therefore,
eqn (6.19) where P = 0.
The first step for the solution of the problem is the evaluation of the new number
of rotation of the pump, which is computed as shown in Figure 6.25; then the
working point of the system is carried out using the same procedure adopted for
the steady flow, but with the new values of the pump parameters.
If the velocity becomes negative, in the computer code it is equal to zero, in
order to avoid reverse flow; this position is quite close to the reality, as the reverse
flow should be avoided also in the real world and appropriate valves (checkvalves)
are installed downstream the pumps in order to prevent it. However, the behaviour
of the system since the flow velocity is zero as the section of the pump strongly
DEVICES BOUNDARY CONDITIONS 97

160 Upstream valve


Downstream valve

120
Heat (m)
80

0 20 40 60 80 100
Time (s)

Figure 6.24: Numerical results carried out for the plant A shown in Figure 6.23.

np:=nn-900*9806*v[0]*Area1*h00/(Efficiency*Inertia*Pi*Pi*nn)*dt;

PumpATrans:=PumpA;
PumpBTrans:=PumpB*(np/NumRot);
PumpCTrans:=PumpC*(np/NumRot)*(np/NumRot);

Figure 6.25: Boundary conditions for the pump.

depends on the rest of the plant; therefore, the code that has been developed is to
be considered valid up to this point.
The plant to be simulated is with a pump positioned upstream; then there is a
single pipe whose length is equal to 3000 m, diameter equal to 0.45 m and Mannings
roughness n = 0.0125 m1/3 s. Downstream a reservoir whose water elevation is
equal to 10 m is positioned.
The coefficients of the pump are:

a = 12.72 m1 s2
b = 1.32 s
c = 53 m

Inertia of the pump is equal to G D2 = 12 N m2 , speed is equal to 2980 rpm,


efficiency = 0.88. The value of inertia is computed with the pump filled with
water and also adding the inertia of the motor.
The curve for the pump is shown in Figure 6.26, together with the curve of the
headlosses of the plant and the working point of the system.
98 WATER HAMMER SIMULATIONS
60

50

40
Head (m)

Working point
30

20

10 Curve of the pump


Curve of the plant
0
0 50 100 150 200 250 300
Discharge (l/s)

Figure 6.26: Head balance curve for the pump, coupled with the plant headloss
curve. The working point is highlighted.

1.6 3500
Flow velocity GD2  12
1.4 Flow velocity GD2  24 3000
Pump rotation GD2  12
Pump rotation GD2  24
1.2
2500
Rotation (rpm)
Velocity (m/s)

1
2000
0.8
1500
0.6
1000
0.4
500
0.2

0 0
0 0.2 0.4 0.6 0.8 1
Temp (s)

Figure 6.27: Flow velocity and pump rotation for the simulated plant and two
different values of inertia.

As can be easily seen, in the pipe a discharge of 221 l/s is flowing, with a velocity
equal to 1.39 m/s.
The simulation is related to the sudden shut off of the energy of the pump, while
the pump stops in a time dependent on the inertia and the system. Simulations have
been compared with those obtained doubling the inertia.
DEVICES BOUNDARY CONDITIONS 99

As can be seen in Figure 6.27, the increase of the value of inertia has a visible
effect on both the flow velocity and the rotation of the pump; however, the flow
velocity reaches zero in a very short time, while the pump continue to rotate, but
with no effect. In the figure, continuous lines are related to computed velocities,
while dashed lines to the estimated values. Actually, larger pumps can continue their
rotation for minutes, but this does not solve the problems due to the waterhammer
which effects are created by the velocity of the water inside the conduit. Therefore,
it is difficult to say whether the increase of the inertia of the pumps has appreciable
effects and can be suggested instead of the most common devices as the air chambers
or others, especially because of the problems that they can create; probably no
general answer exists.
This page intentionally left blank
Chapter 7
Instabilities

As mentioned in the previous chapters, changes from the steady flow conditions
can be considered as disturbances that can cause serious problems; with time, these
disturbances dampen and the flow reaches a new steady regime. Theoretically,
this new regime is reached asymptotically, in practice after a finite time. However,
in some cases, not only the oscillations does not attenuate, but also do they increase
and the system collapse. In the previous chapters such instabilities of the system
have been encountered: in Chapter 6 they were related to surge tanks, which require
a minimum dimension in order not to give rise to undamped oscillations; then, a
real case of resonance was presented, due to valves inappropriately installed. Other
cases are presented in the scientific literature, as those found in leaking valves [64]
or in the behaviour of check valves, if not correctly installed [43]. All these cases
depend on both the system and the external acting force characteristics.
The instabilities may not be in the system, but in the computer code, which
crashes for numerical problems not related to the actual plant. Again, this case has
been encountered in Chapter 6.
In this chapter the two cases are addressed.

7.1 Vibrations
7.1.1 General remarks

The problems related to vibrations are very complex and textbooks are available
to deal with them [65, 66], and therefore in the following only few reminders of
theory, which will be used later in the book, are presented.
For a pipe with length L made with only one material and with constant diameter
and thickness, the fundamental period Tt is [67]:
4L
Tt = (7.1)
c
where, as usual, c is the celerity of the wave. This is also the highest harmonic; for
instance, Tt /3 is the period of the third harmonic, Tt /7 of the seventh harmonic
and so on.
The frequency that corresponds to the fundamental period, or highest harmonic,
is called resonance frequency, which refers to an oscillation that leads to a pressure
102 WATER HAMMER SIMULATIONS

k1 k2 k3

m1 m2 m3

x1 x2 x3

Figure 7.1: System with three degrees of freedom.

amplification. If T is the period, the corresponding frequency (rad/s) is equal to:

2
= (7.2)
T

while the circular frequency f (cycle/s) is equal to:

1
f = (7.3)
T

The number of coordinates required to define a system is said degrees of freedom.


Figure 7.1 shows a system made of three masses connected in series with three
springs. As can be seen, the status of this system can be described with the
three coordinates x1 , x2 and x3 and therefore it has three degrees of freedom.
A system made of a mass and a spring has a natural frequency n equal to:
"
n = k/m (7.4)

where, with the notation used in Figure 7.1, k is the elastic constant of the spring
and m is the mass. A system with n degrees of freedom has n natural frequencies,
and each of them defines a normal mode of vibration.
When an oscillatory force with frequency f is applied at this system, it oscil-
lates with the frequency of the forcing momentum and an amplitude that depends
on the ratio f /n . If the two frequencies have the same value and no friction is
applied, the amplitude of the oscillations tends to infinitive, because at each cycle
the system increases its energy, which cannot be dissipated. If friction is considered,
the amplitude of the oscillations increases until the energy acquired in each cycle
INSTABILITIES 103

vp[n]:=v0*cos(t*Period);
hp[n]:=h[n-1]-cel*(vp[n]-v[n-1])/g-
cel*Lambda*v[n-1]*abs(v[n-1])*dt/(2*g*diam);

Figure 7.2: Code implemented in the computer program to simulate an oscillating


downstream velocity.

is equal to that dissipated by the friction: therefore the system does not increase its
energy and the oscillations have finite amplitude.
In the description of hydraulic systems, the fluid is described as made of an
infinite number of masses and springs in series, with the springs representing the
fluid elasticity. As a consequence, an hydraulic system has infinitive degrees of
freedom.
This assumption can be noted observing the equations that describe the move-
ment: if a system is governed by an ordinary differential equation, then in this
system the number of degree of freedom is equal to the equation order; while, if
a system is governed by partial differential equations, the number of degree of
freedom is infinitive.

7.1.2 Computer program for oscillating velocity

To check and to understand the theory shown in Section 7.1.1, a simple computer
program is developed, very similar to that presented in Chapter 4, but having as
downstream boundary condition the following:

2 c
V [n] = V0 cos t N (7.5)
4L

where t is time, 2c/4L is the fundamental frequency of the system and N is a


parameter to adjust the period of the oscillating velocity. The negative characteristic
line is used to compute the head.
Obviously, the condition does not represent an actual system or valve, but its
function is purely educational.
The code is shown in Figure 7.2.
As application, let us consider the case of a plant with an upstream reservoir
(the head is constant and equal to 100 m) a pipe 10,000 m length and 0.3 m diam-
eter, celerity equal to 1,000 m/s discharge equal to 70 l/s; downstream a valve is
positioned. Headlosses are negligible.
Two simulations have been carried out. In the former, the frequency of the oscil-
lations is let equal to the natural frequency of the system, i.e.: = 2c
4L
= 5 .
In the latter simulation, the frequency of the downstream valve is equal to the
half of the above computed, i.e.: = /10.
104 WATER HAMMER SIMULATIONS
3000

Force in resonance
2000 Force in non resonance

1000
Head (m)

0
0 50 100 150 200 250 300

1000

2000
Heads below 10.33 m are not physically possible
because of column rupture. They are shown only
to demonstrate the effect of the resonance.
3000
Time (s)

Figure 7.3: Results of simulations with oscillating downstream head.

Results of simulations are shown in Figure 7.3; obviously, when the pressure
drops below 10.33 m the water column breaks and the actual behaviour is differ-
ent from that seen in the figure (see Chapter 8), but the simulation has not been
interrupted because the target was the visualization of the behaviour of a system in
resonance.
As can be seen, when the forcing momentum has a frequency of = /10
(this is obtained inserting in the user interface the parameter N Periods = 0.5),
oscillations are limited and have a period equal to 40 s; when the forcing momentum
has a frequency equal to the natural frequency of the system, i.e., = /5 (and
this is obtained inserting in the user interface the parameter N Periods = 1.0),
the pressures increase more and more. In the simulations, the oscillations are not
damped having accepted the hypothesis of negligible headlosses. However, if the
forces that can contrast the increase of the oscillations are limited, heads increase
and the system eventually breaks.

7.2 Transfer matrix method


The case presented in the Section 7.1.2 is very simple and the result of the simulation
was known in advance from vibration theory. Obviously, and unfortunately, real
cases are much more complex.
Different methods have been developed to study this kind of problems. Numer-
ical models as revised in this book may not be able to clarify the problem, because
INSTABILITIES 105

it is not always clear whether unexpected oscillations that are noted during simu-
lations have physical or numerical origin (in other words, if it is a problem of the
computer program, or the plant can actually develop unwanted oscillations).
When the origin of the oscillations is not clear, a different method is used to
evaluate the system stability, which develops overall matrices for a fluid system
[68]; this method is widely used also in mechanics, structural and electrical engi-
neering. This method works in the frequency domain, linearizing the equations
that govern the phenomenon and considering sinusoidal both the motion and the
pressure fluctuations.

7.2.1 General remarks

Generally speaking, a linear system which input variables x1 , x2 , , xn are tied to


the output variables y1 , y2 , . . ., yn by the equations:

u11 x1 + u12 x2 + + u1n xn = y1


u21 x1 + u22 x2 + + u2n xn = y2
(7.5)



un1 x1 + un2 x2 + + unn xn = yn

in matrix notation can be written as:


 x
y = U (7.6)

where x and y are the vectors that contain the variables, and they are called state
vectors; the matrix U  contains the parameters of the system and it is called transfer
matrix.
In the hydraulic systems the variables to be transferred are, as mentioned, veloci-
ties (or discharges) and heads (or pressures). As a consequence, in this case the input
data are the upstream velocity and head, while the output data are the downstream
velocity and head, i.e.:
   
V  V 
zinput =  upstream  zoutput =  downstream  (7.7)
hupstream hdownstream

and, therefore, the system is governed by the equation:

 z
zoutput = U (7.8)
input

Obviously, systems can be composed by combinations of series or parallel pipes,


valves, junctions, pumps and all the devices are only partially described in the book.
For each of these subsystems, an equation has to be written based on the transfer
matrix; then, the description of the whole system is carried out joining all these
subsystems in a larger matrix.
106 WATER HAMMER SIMULATIONS

7.2.2 Application to hydraulic transients

The governing equations have been shown in Chapter 2; in the following they are
written again for the sake of simplicity and in the form which will be used in this
chapter [11].

Continuity equation:
gA
Ht + Q x = 0 (7.9)
c2
Momentum equation:
1 Qn
Hx + Qt + f =0 (7.10)
gA 2 g D An
In eqn (7.10) the exponent n depends on the headlosses. As can be seen, in
these equations the unknown variables are the heads H (instead of the pressures p)
and the discharges Q (instead of the velocities V ); moreover, with a more concise
notation, the generic partial derivative has been indicated:
A
Ab = (7.11)
b
The hypothesis that the transient is developed around average values of head H and
discharge Q is introduced. Therefore the following equations can be written:

H = H +h (7.12)
Q = Q+q (7.13)

where h and q are the oscillation of the head and discharge values around the steady
flow values. Deriving with respect to the space and the time, and remembering that
in steady flow the discharge is constant in the space and time, while the head is
constant only with respect to the time since it changes in the space (because of the
headlosses), it yields:
Qn
H x = Hx + h x = f + hx (7.14)
2 g D An
Ht = Ht + ht = ht (7.15)
Qx = Qx + qx = qx (7.16)
Qt = Qt + qt = qt (7.17)

The term related to the headlosses can be written as follows:


n n1
Qn (Q + q)n Q Q
f =f =f +f n q
2gDA n 2 g D An 2 g D An 2 g D An
(7.18)
INSTABILITIES 107

Substituting these equations in eqns (7.9) and (7.10) yields:


gA
ht + qx = 0 (7.19)
c2
n1
1 Q
hx + qt + f n q = 0 (7.20)
gA 2 g D An
Introducing the terms:
1
L= inertance (7.21)
gA
gA
C= 2 capacitance (7.22)
c
n1
Q
R = f n resistance (7.23)
2 g D An
The system finally becomes:

qx + C ht = 0 (7.24)
hx + L qt + R q = 0 (7.25)

This system can be solved with the method of separation of the variables that
consists in deriving the first equation with respect to the time t and the second with
respect to the space x, and then to subtract them from each other; afterwards, the
process is repeated inverting the order, i.e., deriving the first equation with respect
to x and the second with respect to t. At the end of the whole process the following
equations are obtained:

qxx = C L qtt + C R qt (7.26)


hxx = C L htt + C R ht (7.27)

A new hypothesis is to be introduced: that the variable h can be computed as the


product of two independent functions X = X (x) and T = T (t); if this hypothesis
is verified, the integral of the system can be computed. Substituting, the following
equations can be carried out:

1 d2 X
= 2 (7.28)
X dx2

1 d2 T dT
CL 2 +CR = 2 (7.29)
T dt dt

where 2 is a complex value independent from x and t. Integrating it yields:

X = A1 e x + A2 e x (7.30)
T = A3 est (7.31)
108 WATER HAMMER SIMULATIONS

where s is again a complex value independent from x and t, while A1 , A2 and A3


are the integration constants. Now it is possible to tie the variables and s:

2 = C s (L s + R) (7.32)

Considering that h(x, t) = X (x) T (t), it yields:

h = est (C1 e x + C2 e x ) = H (x) est (7.33)

and finally it is possible to define the expression:

C s st
q= e (C1 e x C2 e x ) = Q(x) est (7.34)

The variable is called propagation constant and represents the complex


frequency, i.e.:
s= +i (7.35)
where is referred to the damping of the system and is the oscillation circular
frequency. Finally, the characteristic impedance can be defined as:


Zc = (7.36)
Cs

This impedance is a complex function independent of space and time; this function
and the propagation constant are functions of fluid and pipe characteristics and the
complex frequency.
The functions H (x) and Q(x) are the complex head and the complex discharge,
respectively, and depend on the spatial variable only. With these equations it is
possible to define the values of the integration constants function of the boundary
conditions. Once these constants are known, transfer matrix can be built for each
element of the system.

7.2.3 Description of simple system: pipes

In the simple case of a single pipe, represented in Figure 7.4, the hydraulic variables
have to be transferred from upstream to downstream.
As when x = 0 it must be H = Hupstream and Q = Qupstream , from the definition
of H (x) and Q(x), given in eqns (7.33) and (7.34), the values of the integrations
constants C1 and C2 are carried out:

1
C1 = (Hupstream Zc Qupstream ) (7.37)
2
1
C2 = (Hupstream + Zc Qupstream ) (7.38)
2
INSTABILITIES 109
H upstream H downstream

Q upstream Q downstream

Figure 7.4: Sketch of a simple pipe and related symbols.

Finally, substitution yields:

H (x) = Hupstream cosh( x) Zc Qupstream sinh( x) (7.39)


Hupstream
Q(x) = sinh( x) + Qupstream cosh( x) (7.40)
Zc

To compute the values of the complex head and discharge downstream the pipe,
the simple substitution of x = l allows are to obtain:

Hdownstream = Hupstream cosh( l) Zc Qupstream sinh( l) (7.41)


Hupstream
Qdownstream = sinh( l) + Qupstream cosh( l) (7.42)
Zc

These relations allow to write the transfer matrix for the element pipe, accord-
ing to the notations in Figure 7.4, indicating as positive the discharges outflowing
from each element and introducing the functions:

1
T = (7.43)
Zc tanh( l)
1
S= (7.44)
Zc sinh( l)

The final compact transferring relation is:


     
 T S   Hupstream   Qupstream 
  =  (7.45)
 S T   Hdownstream   Qdownstream 

7.2.4 Description of simple system: valves and effluxes

Transfer matrixes in these cases are obtained linearizing the efflux equations, which
does not imply a severe error if the pressure differences induced by the valve are
small compared with the static head.
110 WATER HAMMER SIMULATIONS

If valves have a dynamic behaviour, this is assumed oscillatory sinusoidal; this


method is obviously possible to study non-periodical behaviours too, decomposing
the motion in a series of harmonics through the Fourier analysis, which, however,
will not be described in this book [69].
In the case of a free efflux (i.e., in the atmosphere), say H the head upstream
the efflux and Q the outflow discharge, the well-known formula can be used:
"
Q = Aefflux 2gH (7.46)

where is the efflux coefficient. Recalling eqns (7.12) and (7.13) and letting
= A, it yields:
 1/2
Q Q+q + H +h
= = (7.47)
Q Q H

that becomes:

q  h 1/2
1+ = 1+ 1+ (7.48)
Q H
Explaining h, the formula becomes:
   
2 2q q2
h= 1+ + 2 1 H (7.49)
2 + 2 + 2 Q Q

2
Neglecting the terms 2 and q2 /Q and rearranging, eqn (7.49) becomes:

2 q 2
h= H H (7.50)
+2 Q +2

When << , eqn (7.50) can be simplified as:

H 2
h=2 q H (7.51)
Q

In Section 7.2.2, eqns (7.33) and (7.34) have been developed; considering only their
real parts the following are derived:

h = H (x) sin( t) (7.52)


q = Q(x) sin( t) (7.53)

The hypothesis that the valve has an oscillatory sinusoidal behaviour means that:

= k sin( t) (7.54)
INSTABILITIES 111

where k is the amplitude of the valve motion. With due simplifications the following
is obtained:
H 2k
Hupstream = 2 Qupstream H (7.55)
Q
Because of the hypothesis of free efflux (i.e., directly in the atmosphere),
Hdownstream = 0, and therefore:
H 2k
Hdownstream = Hupstream 2 Qupstream + H (7.56)
Q
Finally, from the continuity equation:

Qupstream = Qdownstream (7.57)

In conclusion, denoting with:


H
U = 2 (7.58)
Q

H k
V = 2 (7.59)

the transfer matrix is:
     
 Qdownstream   1 0 0   Qupstream 
     
 Hdownstream  =  U 1 V   Hupstream  (7.60)
     
 1  0 0 1  1 

7.2.5 Global matrix of a system

Let us consider the usual very simple system, as in Figure 7.5, composed by three
elements in series:

an upstream reservoir;
a pipe;
a valve with free efflux.

For the pipe (Section 7.2.3):


     
 T S   H A   QA 
 
 S T   H B  =  QB  (7.61)
pipe pipe

and for the valve (Section 7.2.4):


 valve     valve 
Q   0   Qupstream 
 downstream   1 0 
 valve     valve 
H   1 V   Hupstream 
 downstream  = U (7.62)
     
 1  0 0 1  1 
112 WATER HAMMER SIMULATIONS

Upstream reservoir

Downstream valve

Figure 7.5: Sketch of a simple plant.

It has to be noted that, in the case of free efflux, obviously is:


valve
Hdownstream = Hatm = 0 (7.63)

Moreover, the upstream head can be denoted as:

H B = Hupstream
valve
= Hpipe
B
(7.64)

Finally, from the continuity equation:


B
Qpipe = Qupstream
valve
(7.65)

Developing the matrixes contained in eqn (7.63) and (7.64), keeping in account the
(7.65) and collecting the common terms, the following equation is carried out:

V U
HB = S HA + (7.66)
U T U +1
With the matrix notation, eqn (7.66) can be expressed as:
 B   SU V   A 
H     
   T U +1 T U +1   H 
 =   (7.67)
 1   0 1   1 

In order to study the free vibrations, then, being negligible the external causes, it
is necessary the analysis of eqn (7.67). The represented system allows non-trivial
solutions if the determinant of the matrix is zero; therefore, it is to be computed the
complex variable s = + i such that:
S U
=0 (7.68)
T U +1
The purpose of the analysis is to identify if there are values of the variable function
of with positive values: in this case, the system is unstable.
INSTABILITIES 113

7.2.6 A simple application

The plant to be studied is that described and tested in Section 7.1.2, i.e., a pipe
10,000 m length and 0.3 m diameter, celerity equal to 1000 m/s where flows a dis-
charge equal to 70 l/s. Headlosses are negligible and the upstream head is equal to
100 m.
The solutions of eqn (7.68) have to be studied.
First, the variables inertance L (eqn (7.21)), capacitance C (eqn (7.22)) and resis-
tance R (eqn (7.23)) are computed (the latter term is zero because of the hypothesis
of negligible headlosses):
1
L= = 1.442
gA

gA
C= = 6.934 107
c2
To compute the characteristic impedance of the system, eqns (7.32) and (7.36) have
to be used:
"
= C s (L s + R) = 9.978 104 s


Zc = = 1438.975
Cs
It is to be reminded that in the above equations, s is the free variable. To study the
behaviour of the pipe, eqns (7.43) and (7.44) are needed:

1 1
S= =
Zc sinh( l) 1438.975 sinh(9.978 s)

1 1
T = =
Zc tanh( l) 1438.975 tanh(9.978 s)
Because of the valve, eqn (7.58) is to be introduced:

H
U = 2 = 2857.143
Q

Eqn (7.68) in the complex variable s = + i can finally be written as:

1.986
1.986 + tanh(9.978 s) =0
cosh(9.978 s)

The study of this function is possible with any spreadsheet and little effort. It can
be shown that no positive values of exists to set the equation equal to zero: this
result means that the system is stable. As mentioned, this is true when no external
114 WATER HAMMER SIMULATIONS

force are considered; on the contrary, the presence of an external force oscillating
with the natural frequency of the system would bring it in resonance.
In conclusion, to decide whether a system is stable, different methods should
be applied. However, a good practice is always the installation of slow-response
devices, which reduce the risk of undesired instabilities.

7.3 Numerical instabilities


In the case of waterhammer, it may happen that the computer program crashes even
when the real plant under analysis is not problematic; this is often to be ascribed to
numerical instabilities.
Numerical schemes based on the series of Taylor, as the LaxWendroff s, can
be analysed developing the series. As it has been shown in Chapter 5, the solution
for a linear system like the following:

u u
+A =0 (7.69)
t x

or, in compact form:


ut + A u x = 0 (7.70)
can be expressed as:

1
u(x, tn+1 ) = u(x, tn ) + t ut (x, tn ) + (t)2 utt (x, tn ) (7.71)
2

Equation (7.71) shows that the series of Taylor has been stopped at the second
order and this implies that the truncation error is given by the following neglected
term, which is:
uttt = A3 uxxx (7.72)
To study the behaviour of the solution, it is possible to look closely to the
original equation (7.70) with a slight modification in order to correlate it with the
truncation error. Hedstrom [70] shows that the modified equation for the Lax
Wendroff method is:

ut + A ux = uxxx uxxxx (7.73)

where:

(x)2 (t)2
= A A2 1 (7.74)
6 (x)2

(x)3 (t)2
= A 1 A 2
(7.75)
6 (x)2
INSTABILITIES 115
3
2,5
2
1,5
1
0,5
0
0,5
1
1,5
2
2.5 2 1.5 1 0.5 0 0.5 1 1.5 2 2.5

Figure 7.6: Example of the instabilities.

In eqn (7.73), the two addends have different physical meaning. Generally speak-
ing, even spatial derivatives as uxxxx are related to dissipation (or amplification)
terms, while odd terms are related to dispersion. The term uxxxx modifies the
trend of the solution when |c | < 1 and that confirms the need to let CFL < 1 and
at the same time close to one, in order to have stable and non-dissipative solutions.
If the spatial term can be neglected, eqn (7.73) is reduced to:

ut + A ux = uxxx (7.76)

which is called dispersive equation and allows to explain the formation of oscilla-
tions close to the points where the presence of large gradients of the dependable
variables is noted, as shown in the Figure 7.6 where a square wave is integrated
using LaxWendroff method. The effects of diffusion have been encountered and
highlighted in Chapter 4 (see Figure 4.11), but without an appropriate analysis,
which will be performed in the following.
In this book the analysis of the reasons why the instabilities occur is not per-
formed, as more specialized books are available in the scientific literature [31];
however, as this problem is found quite frequently, some methods to solve it, or at
least reduce it, are shown. In particular, the following will be analysed:

the change of the CFL parameter;


the use of a first order code;
the use of flux limiters;
the use of artificial viscosity.
116 WATER HAMMER SIMULATIONS
Diffusion errors LaxWendroff
1.2
1
0.8 CFL  0.25
0.6 CFL  0.50
0.4 CFL  0.75
0.2 CFL  1.00
0
0 20 40 60 80 100 120 140 160 180

Dispersion errors LaxWendroff

0.5

0 20 40 60 80 100 120 140 160 180


k x

Figure 7.7: Diffusion and dispersion errors, function of CFN and angle of phase
parameters (after Ref. [71]).

7.3.1 Changing CFL number

As discussed in Chapter 5 and mentioned above, the easiest way to try to reduce
the numerical oscillations is to keep the CFL number close but lower than 1;
the consequences of the variations of this number have already been shown in
Figure 5.12.
In the paragraph above (Section 7.3) a brief analytical description of the oscil-
lation problem has been given through the modified equation (7.73), introducing
a dissipative (or amplifying) and a dispersive term. Changing the CFL number
implies a change in the numerical value of these parameters with obvious impli-
cations on the stability and convergence of the solution. Figure 7.7 shows [71] the
relative values of the parameters and when the CFL number changes, for the
LaxWendroff method; on the x-axis the phase angle of the solution1 is reported. As
can be seen, in the case of the LaxWendroff method, the only value that guarantees
no dispersion along the whole axes of the pipe is CFL = 1, as already mentioned.
When the phase angle increases, for CFL values lower than 1, the dissipative and
dispersive terms become more important.

1
The phase angle of the kth harmonic is a real number expressed in radiants and equal to
k x.
INSTABILITIES 117

150

125

100

(m)
Head[s]
75

Head
4000
50

3000 25

0
2000 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
Time (s)

1000
Head (m)

0
0 5 10 15 20 25 30 35 40 45 50
1000

2000

3000

4000
Time (s)

Figure 7.8: Results with boundary conditions solved with the BeamWarming
method and internal points with LaxWendroff (CFL = 1).

It has also been already mentioned that the condition on the CFL number is
not sufficient to guarantee stability: Figure 7.8 allows to focus on a particular
case. Results have been carried out with the usual simple plant where the valve
is instantaneously closed; the pipe length is equal to 10 000 m and the celerity is
1000 m/s. Therefore, the perturbation starts at time 0 s and it is back every 20 s;
as can be seen from Figure 7.8, when the perturbation is back to the valve, the
instabilities increase until the computer program crashes. The problem in this case
is obviously given at the boundary conditions, which have been solved with the
BeamWarming scheme and cannot be solved operating on the CFL number.

7.3.2 First order methods

As already described, the numerical instabilities are due to the truncation errors
and are part of the intrinsic diffusive properties of the selected numerical method
of solution. Usually, if the main terms of the error are odds, as it happens in the
second order schemes, these dispersive errors generate the oscillations. A viable
118 WATER HAMMER SIMULATIONS
1.5

0.5

0.5
0 0.2 0.4 0.6 0.8 1

Figure 7.9: Numerical diffusion in the integration through first-order schemes.

alternative is therefore the reduction of the order of the numerical scheme, using a
first order method, for instance the mentioned LaxFriedrichs.
LeVeque [36] illustrates the results carried out with first order schemes although
they do not show non-physical oscillations in the presence of discontinuities, and
they are less accurate with respect to second order methods. As can be seen in
Figure 7.9, in fact, where a generic case of first-order integration is shown, the
solution is approximated, but the numerical solution tends to be smoother than the
analytical because of the numerical dispersion.
In this case, the equivalent equation at the third order is [71]:

ut + A ux = D uxx uxxx (7.77)

where:

(h)2 (t)2
D= 1 A2 (7.78)
2 t (h)2


(h)3 (t)2
= 1 A2 (7.79)
3 (h)2
The term D uxxx is dissipative; in other words, increasing the positive parame-
ter D, the solution tends to flatten, becoming increasingly smoothed where the first
derivative is discontinuous. The term uxxx is normally not negligible, although
the dispersive effect is usually not really noticeable, because hidden by the dissi-
pative, much larger. As the numerical dispersion is quite small, no oscillations are
recorded; however, an increase of the celerity when > 0 and a diminution when
< 0 can be noted. As a consequence, observing the position of the wave crests of
Figure 7.9 compared with those of the exact solution, it can be seen that the Lax
Friedrichs method has, generally speaking, a positive dispersive error: as > 0
the celerity changes and the carried out solution is faster than the exact solution.
INSTABILITIES 119
Diffusion errors LaxFriedrichs
1.2
1
0.8
0.6
0.4
0.2
0
0 20 40 60 80 100 120 140 160 180

Dispersion errors LaxFriedrichs

5 CFL  0.25
4 CFL  0.50
3 CFL  0.75
CFL  1.00
2
1

0 20 40 60 80 100 120 140 160 180

k x

Figure 7.10: Amplification and dispersion errors, function of CFN and angle of
phase parameters (after Ref. [71]).

In Figure 7.10 for the method of LaxFriedrichs the values of the coefficients
D and are reported, when changing the phase angle and the CFL number.
Letting the CFL number equal to 1, the centered LaxFriedrichs method can
be considered a good choice; also in this case, the CFL condition is not sufficient
to guarantee a solution free of errors of amplification or dispersion, and therefore,
the use of a first-order scheme is to be deeply evaluated. The absence of spurious
oscillations in fact, in general could be compensated by the presence of numerical
errors of different nature such as dispersion.

7.3.3 Flux-limiters

A very interesting option to be used to deal with these problems is called Flux-
Limiter; this expression joins a number of methods that generally guarantee good
results. The common ground of all these methods is the insertion in the numerical
code of a new term which allows to obtain a second-order accuracy when it is
possible, but that smooths it when the numerical solution becomes unstable.
The general idea is the construction of an hybrid numerical flux, carried out
combining a second-order scheme with a monotonic first-order scheme so that the
former or the latter prevails when the solution is regular or irregular, respectively.
120 WATER HAMMER SIMULATIONS

One of these methods is the so-called total variation diminishing (TVD) [72];
the discretized total variation of a solution u for a scalar conservation law is:
#
TV (u) = |ui+1 ui | (7.80)
i

and a numerical scheme is called TVD if for any time and any u the following
inequality is satisfied:
TV (un+1 ) TV (un ) (7.81)
When the inequality (7.81) is met, the scheme respects the monotonic trend of
a function, or, in other words:

in the computational domain no new extremes can be created;


the value of a local minimum is not decreasing, while the value of a local
maximum is not increasing.

Obviously, within a scheme with such properties, starting from a correct


solution, non-physical oscillations cannot be created.
Godunov [73] demonstrated that linear first-order TVD schemes do not exist,
and therefore in a second order scheme it is necessary to introduct non-linear
correction terms.
These terms are known as limiters, as they limit the flux between the cells.
Generally speaking, a flux-limiter scheme can be expressed as:

H = HL + (HH HL ) (7.82)

where HL is a first order monotonic flux (Lower), HH is a second order flux


(Higher),  is a flux limiter. It is desirable that the function  tends to 1 when the
solution is regular and that tends to 0 when the solution is discontinuous; moreover,
it must guarantee that the scheme converge to the right solution. After having
arbitrarily defined HL and HH , the shape of the flux limiter  is defined accordingly.
As example, HL can be defined as the Upwind first-order numeric flux:

ujn+1 = ujn Anj (ujn uj1


n
) (7.83)

while as HH , the second-order LaxWendroff scheme, can be used:

1 n   1  
ujn+1 = ujn Aj uj+1
n
uj1
n
+ A2 2 uj+1
n
2 ujn + uj1
n
2 2
(7.84)

The parameters have the usual meaning and in particular = t/x. The hybrid
flux is, therefore:

1 n
Q(u, j) = Anj ujn + A (1 ) (uj+1
n
uj1
n
) (jn ) (7.85)
2 j
INSTABILITIES 121
 2

2
2

Figure 7.11: Admissibility region for a flux-limiter method.

where jn is an index related to the regularity of the solution and it is defined as:

ujn uj1
n
jn = (7.86)
n
uj+1 ujn

Therefore, when jn 1 data are presumably regular, while when jn 0 is


likely the presence of discontinuities. LeVeque [36] and Hirsch [31] show that, for
the function (jn ) to be consistent and to allow the convergence towards the real
solution it is necessary the function is confined in an admissibility region, defined
by eqn (7.87) and shown in Figure 7.11.
 
 ( n ) 
 j 
 ( j1  2
n
) jn , j1
n
(7.87)
 jn 

The inequality (7.87) can be also expressed as:

0 (jn ) 2 jn 0 (jn ) 2 (7.88)

According with Ref. [31, 36], for the method to be second-order accurate the
following conditions have to be satisfied as well:

the gradient of the function (jn ) is limited;


(1) = 1;
 is continuous in jn = 1.

Combining LaxWendroff method with the boundary conditions given by


BeamWarming, for which () = , it is possible to definite a new, second-order,
admissibility region, shown in Figure 7.12.
122 WATER HAMMER SIMULATIONS

 2 

2 2

1 1

Figure 7.12: Second-order region.

Any curve chosen for the function () which falls in this region guarantees
that the carried out second-order method is converging towards the exact solution.
Solutions carried out, even when discontinuities of the first derivative are present,
are oscillations free and give a good approximation of the real value.
The most common limiters are the following, represented also in Figure 7.13 to
show their position inside the above defined second-order region.

Superbee limiter:

() = max(0, min(1; 2), min(; 2)) (7.89)

Van Leer limiter:


|| +
() = (7.90)
|| + 1
MinMod limiter:
() = max(0, min(1; )) (7.91)

The described limiters, only a sample from those available in literature, allows to
modify the solution depending on the problem and on the implemented integration
scheme. Therefore, it is not possible an a priori suggestion of the best limiter
that guarantee proper solutions for each analyzed case, but it is necessary to make
that choice as a result of an analysis of the problem.

7.3.4 Artificial dissipation

To improve their solution, LaxWendroff [33] developed this method, which con-
sists in the insertion in the numerical code of terms that model an artificial viscosity,
INSTABILITIES 123

 2 

2 2

1 1

Figure 7.13: Flux-limiters: Superbee (dash dot), Van Leer (continuous) and
MinMod (dashed).

or dissipation, directly derived from the discretization of the equations. These terms
have the effect to increase the viscosity around the critical points, although they
are negligible in the regions where no oscillations are produced and the solution is
regular.
Reminding the base scheme of the LaxWendroff method:

n 2
ujn+1 = ujn A (uj+1 uj1
n
)+ A2 (uj+1
n
2ujn + uj1
n
) (7.92)
2 2

which can be written in the general form as:

ujn+1 = ujn (Hj+1/2


n
Hj1/2
n
) (7.93)

where
n
Hj+1/2 = h(ujn ; uj+1
n
) (7.94)

n
Hj1/2 = h(ujn ; uj1
n
) (7.95)

are the numerical fluxes, and in this case they can be expressed as:

1  
n
Hj+1/2 = A (uj+1
n
uj ) (A2 (uj+1
n
uj )) (7.96)
2
1  
n
Hj1/2 = A (uj1
n
uj ) (A2 (uj1
n
uj )) (7.97)
2
124 WATER HAMMER SIMULATIONS

vp[i]:=v[i]-0.5*(dt/ds)*((h[i+1]-h[i-1])*g+v[i]*(v[i+1]-v[i-1]))+
0.5*(dt/ds)*(dt/ds)*(2*g*v[i]*(h[i+1]-2*h[i]+h[i-1])+
(cel*cel+v[i]*v[i])*(v[i+1]-2*v[i]+v[i-1]))-
Lambda*v[i]*abs(v[i])*dt/(2*diam);
if abs(h[i+1]+h[i]+h[i-1])>0.01 then
begin
CorrV:=eps*(dt/(2*ds))*(abs(v[i])+cel)*abs(v[i+1]-2*v[i]+v[i-1])*
(v[i+1]-v[i-1])/(h[i+1]+h[i]+h[i-1]);
vp[i]:=vp[i]-CorrV;
end;
hp[i]:=h[i]-0.5*(dt/ds)*(v[i]*(h[i+1]-h[i-1])+(Cel*Cel/g)*(v[i+1]-v[i-1]))+
0.5*(dt/ds)*(dt/ds)*((v[i]*v[i]+Cel*Cel)*(h[i+1]-2*h[i]+h[i-1])+
(2*Cel*Cel*v[i]/g)*(v[i+1]-2*v[i]+v[i-1]));
if abs(h[i+1]+h[i]+h[i-1])>0.01 then
begin
CorrH:=eps*(dt/(2*ds))*(abs(v[i])+cel)*abs(h[i+1]-2*h[i]+h[i-1])*
(h[i+1]-h[i-1])/(h[i+1]+h[i]+h[i-1]);
hp[i]:=hp[i]-CorrH;
end;

Figure 7.14: Computer code for LaxWendroff method with artificial viscosity.

The addition of the viscosity coefficient changes the eqns (7.96) and (7.97) as
follows:

1
n
Hj+1/2 = [A (uj+1
n
uj ) (A2 (uj+1
n
uj ))] D(ujn ; uj+1
n
) (uj+1
n
uj )
2
(7.98)

1
n
Hj1/2 = [A (uj1
n
uj ) (A2 (uj1
n
uj ))] D(ujn ; uj1
n
) (ujn uj1 )
2
(7.99)

The function D represents the artificial viscosity. Inserting that term in the
general LaxWendroff scheme, the expression is:

2
ujn+1 = ujn A (uj+1
n
uj1
n
)+ A2 (uj+1
n
2 ujn + uj1
n
)
2 2

+ [D(ujn ; uj+1
n
) (uj+1
n
uj ) D(ujn ; uj1
n
) (ujn uj1 )] (7.100)
2

In order to have a stabilizing effect, function D has to be positive. In practice,


n
such term is expressed through a polynomial function of (uj+1 ujn ) and (ujn uj1
n
).
INSTABILITIES 125
120

100

80
Head (m)

60

40

Air chamber without orifice


20
Air chamber with orifice

0
0 10 20 30 40 50 60
Time (s)

Figure 7.15: Effects of artificial viscosity for the case of air chamber with orifice.

Many alternatives are possible [7476]. In particular, MacCormack and


Baldwin [76] proposed the following definition:
 
|V | + c  2 u 
D = x2  2 (7.101)
p x

Finally, the LaxWendroff scheme becomes:

2
ujn+1 = ujn A (uj+1
n
uj1
n
)+ A2 (uj+1
n
2 ujn + uj1
n
)
2 2
|uj+1
n
2 ujn + uj1
n
|
(|V | + c) (uj+1
n
uj1
n
) (7.102)
2 hj+1 + hj + hj1
n n n

Again there is the problem of a correct determination of the coefficient , which if


too low the artificial viscosity has no effect and if too high the computer program
crashes; therefore, a fine tuning is required and depends on the case to be solved.
Modelling transients is never a trivial problem, but when oscillations are present it
becomes almost an art.
However, the correction has been inserted in the computer program, as shown
in Figure 7.14, and the results for the case of air chamber with and without orifice
are shown in Figure 7.15. As can be reminded, in Section 6.6.3 (see Figure 6.18),
instabilities were found using the method of characteristics to try to simulate the
presence of an air chamber with orifice. With the use of the LaxWendroff s method
and limiters, the problem is completely solved.
This page intentionally left blank
Chapter 8
Effects of air and cavitation

During transients it may happen that pressures fall below atmospheric, especially
when steady flow pressures are low; sometimes pressures can fall even down to
the vapour pressure related to the local temperature of the fluid with consequent
formation of cavities1 . The pressure inside these cavities is not larger than the
vapour pressure mentioned above, if no air is artificially inserted.
The formation of those cavities can bring two types of danger to the plant. In
the former, the danger of crushing for excess of external pressure and the rupture
of parts particularly delicate (for instance, in some treatment water plants, some
kinds of membranes). In the latter, in the subsequent phases of the phenomenon,
the cavities can abruptly collapse, which can produce very high pressures.
Moreover, the effect of the presence of air is noted in the transients for two
different cases: the change of the value of the celerity of the propagation wave and
an increased flow resistance.
Finally, the cavitation can become so large to produce the so-called column
rupture, which effects are significant and very difficult to be reproduced with a
numerical model.
All these effects are briefly discussed in this chapter. However, this topic is still
subject to research and many aspects need to be disclosed.

8.1 Cavities
8.1.1 Formation of the cavities

As mentioned, when the pressures goes below the value of vapour pressure, first we
have the formation of bubbles and then the phenomenon of water column rupture.

1
Theoretically, the absolute pressure of the fluid can fall to zero and only then the rupture
of the column should happen; in fact, absolute pressure equal to zero means that normal
stresses are zero; positive absolute pressure means that normal stress are of compression and
negative pressures means that the stresses are of traction. As water (and all fluids, generally
speaking) breaks under traction, negative absolute pressures are not possible. However, in
the real world, before reaching this theoretical limit, the dissolved air in the water frees and,
therefore, cavities can form much earlier.
128 WATER HAMMER SIMULATIONS

In case of equilibrium, the Henrys law [77] applies:

C/p = constant (8.1)

where C is the gas concentration and p, is the pressure. Obviously, during transients
the equilibrium is never reached: gas tends to develop preferentially from free sur-
faces (in this case, for instance, from existing bubbles) and a diffusive phenomenon
starts, where concentration is function of position and time.
The governing equation is, therefore, of the type2 :


C
D 2 C + v grad C = (8.2)
t

where D is the diffusivity constant of the liquid. Boundary conditions are quite
difficult to be set and the equation itself is normally not analytically integrable.
Usually two types of conditions are placed where the water column breaks. If
no air is inserted, the simplest condition is to consider the cavity as a point where
the pressure is constant and equal to vapour pressure.
A more sophisticated way to deal with the bubbles is the hypothesis that the gas
inside them follows the polytropic law:

p W m = constant (8.3)

where W is the volume of gas and m is the exponent of the law, which falls in the
range from 1.0 (for isotermic transformations) to 1.4 (for adiabatic). This way it is
possible to study the bubbles stability in function of their dimensions.
Because of the surface tension and imposing the same pressure inside and out-
side a spherical bubble, the equation that provides the pressure p of the liquid
outside the bubble with radius r is:
 r 3 m 2A
0
p = p0 + pv (8.4)
r r

where pv is the partial pressure of the vapour inside the bubble, A is the constant of
surface tension and the subscript 0 refers to the steady conditions. Equation (8.4)
is represented in Figure 8.1.
As can be seen in Figure 8.1, pressure decreases to a minimum given by the
condition:
dp p0  r0 3m 2 A
= 0 3m + 2 =0 (8.5)
dr r r r

 
2C 2C 2C
2
It is to be reminded that 2 C = x2
+ y2
+ z 2
and grad C = C C C
, ,
x y z
.
EFFECTS OF AIR AND CAVITATION 129
P

r 2A
P  Po ( ro )3m r Pv

rmax
r

Pmin

Figure 8.1: Equilibrium between external pressure and bubble radius.

which allows to compute the maximum dimension of the bubble as:

( 13m
1
)
2A
rmax = (8.6)
3m p0 ro3m

To this maximum value of the bubble radius rmax corresponds the minimum
pressure:

( 13m
1
)
1 3m
pmin = 2 A 1 p0 r03m + pv (8.7)
3m 2A

As m > 1, pmin < pv .


Looking again at Figure 8.1, the stable part of the function (r < rmax ) is related
to the gaseous cavitation. If the pressure of the environment p becomes lower than
pmin , the bubble becomes instable and tends to expand; as this expansion is quite
fast, at least with respect to the diffusion time, the liquid becomes vapour, which
remains inside the bubble: this phase is called vaporous cavitation.
This exchange between vaporous and liquid phases is much faster than the
emission and reabsorption of air. The former phenomenon is quite frequent and
cavities can be considered almost always filled with vapour; the latter phenomenon
might be neglected.
As a consequence, the current practice to verify the presence of cavitation, which
consists in comparing the absolute pressure of the liquid with the vapor pressure
is fully justified; moreover, if no external air is introduced, the cavities are filled
only with vapour which is produced and absorbed during the different phases of
the transient.
130 WATER HAMMER SIMULATIONS

Finally, if air is introduced into the cavities, obviously their absolute pressure
will not be constant, but function of both the volume of the cavity and the mass of
air that is introduced (and therefore of the device which allows the entry).

8.1.2 Collapse of the cavities

When the pressure in the pipe increases, the two stubs of the water column,
upstream and downstream, are quite close, being separated by a cavity which volume
tends to zero.
Indicating with vupstream and vdownstream the velocities of the front of the upstream
and downstream stubs, respectively, in the moments immediately before the col-
lapse is vm > vv and the piezometric head is the same for the two parts and the
cavity as well. During the moments immediately following the collapse, there is an
unique value not only of the piezometric head but also of the velocity vc . It can be
demonstrated that:
vm + vv
vc = (8.8)
2
and, therefore, the pressure due to the transient can be computed as:
c c c v m vv
H = (vc vv ) = (vm vc ) = (8.9)
g g g 2

8.1.3 Description of the motion in the presence of cavities

A problem that might be faced is the evaluation of the position of the cavity: if
its position can be assumed known with a good approximation (because it can be
supposed positioned close to the regulatory devices, which originate the waterham-
mer transients), the equations to be written have to allow the determination of the
upstream and downstream fronts of the two parts in which the water column is
divided.
To this end, it is necessary to write an equation that governs the volume of the
cavity, function of its pressure. Therefore, it is possible to:

1. suppose that the pressure inside the cavity is equal to the vapour pressure and
therefore constant;
2. write the characteristic equation for the stub of interest;
3. compute the volume of the cavity W evaluating the differences between the
velocities of the upstream and downstream stubs and considering that W =
A (vm vv ) t (being A the section of the pipe).

It is to be highlighted that in the past the description of the water column separa-
tion was performed under the non-elastic hypothesis. However, for this phenomenon
more studies, both numerical and experimental, are necessary as the descriptions
developed so far are very approximate and there is a need to a better understanding
of the physics of the system.
EFFECTS OF AIR AND CAVITATION 131

8.2 Changing of celerity


As mentioned in the introduction to this chapter, the dimensions of gas bubbles
have two important consequences related to the celerity of the pressure waves and
to the headlosses.
With regards to the celerity, qualitatively the significant decrease of the density
and the correspondent increase of the fluid elastic bulk modulus can be considered:
in other words, a fluid composed of water and vapour bubbles is less dense and
more compressible than a fluid composed of water alone. As a consequence, it can
be verified that the value of celerity presents impressive decreases even for very
low values of vapour concentration.
A fast computation can be performed as follows. Let us have, within a unit
volume of liquid, a number Nr of bubbles all with the same radius r. The so-called
vacuum fraction is therefore equal to:

4
= r 3 Nr (8.10)
3
The density of the mixture m is equal to:

m = v + (1 ) l (8.11)

where v and l are the densities of the pure components vaporous and liquid,
respectively. The volumetric compressibility of the mixture m is therefore equal to:

dp dp
m = dV = 1
(8.12)
V
d
dr
dr
dp
dp +
dp

If, for the sake of simplicity, the hypothesis of isothermal transformation is


introduced (i.e. m = 1), eqn (8.12) becomes:

1
m = 1 3 r3
(8.13)

+ 3 p0 r03 2 A r 2

It can be easily demonstrated that m < when r < rmax and has no meaning when
r > rmax . If, as usual, the mixture is inside a deformable pipe with diameter d,
thickness e and elastic modulus E, the value of celerity to be used for transient
analysis is given by:
$
%
% 1
cm = & 1 m d

(8.14)
m
+ Ee

Therefore, cm can be significantly lower than the usual value computed with the
formula (2.10) carried out in Chapter 2, even for low values of . During transients,
therefore, significant reductions of the celerity have to be expected when values of
 = 0 are expected.
132 WATER HAMMER SIMULATIONS

Valve

Flow Pressure
meter transducer

Figure 8.2: Experimental set-up.

8.3 Water column separation


As described in Section 8.1, when the pressure in the pipe goes below that of satu-
rated vapour, cavities start to form. When the stress is high, the cavities enlarge to
a point which is called water column separation, when the cavity actually separates
the two sections of the liquid pipe.
For each of the two sections, which are separated by the cavity, the cavity itself
is a condition of constant head, equal to the vapour pressure if no air is inserted,
artificially or through devices.
Supposing a plant as that sketched in Figure 8.2 (which has been installed
in laboratory and where tests have been performed, as will be mentioned in the
following), where the valve is positioned upstream and the reservoir downstream,
a sudden closure of the upstream valve produces an immediate pressure drop; in
certain conditions, the pressure drops down to 10.33 m when the water column
separates3 .
The upstream valve is normally modelled imposing the velocity and retrieving
the head from the equation of the system. When cavitation occurs, the boundary
conditions change: in the position of the valve a constant pressure is imposed, equal
to the mentioned vapour pressure, and the velocity has to be computed from the
equation of the system.

3
As known, the value of 10.33 m corresponds to the absolute pressure equal to zero and
therefore the threshold under which the stresses would be of traction; as the liquid do not
stand the traction stresses, the water column separates.
EFFECTS OF AIR AND CAVITATION 133

Direction of the flow in steady conditions


Valve

Air

Water

Figure 8.3: Cavitation recorded during laboratory tests.

Concluding, until the water column is separated, the condition to impose


upstream is that of a constant reservoir, with head h0 = 10.33 m; when the two
sections reattach, the upstream condition again becomes on the velocity.
If in the cavity (upstream section) air is artificially inserted, the absolute pressure
is not zero but results to be function of (i) the volume of the cavity and (ii) the amount
of mass of air inserted.
A picture of the phenomenon is shown in Figure 8.3, taken through a transparent
pipe positioned downstream the valve. The whole cavitation phenomenon is shown
in a short movie enclosed to the book.

8.4 Additional resistance terms

8.4.1 Models

It has already been stated that the presence of the bubbles induces additional energy
dissipations compared to ordinary hydraulic resistances because of (i) the conver-
sion of mechanical work into heat in the cycles of compression and decompression
of the bubbles and (ii) the energy dissipated by viscosity in the liquid surrounding
the bubbles. Moreover, some authors (e.g. [78]) believe that the headlosses during
transients are higher than those computed in steady flow not only because of the
presence of the bubbles but also because of the differences in the velocity profile.
These effects are very difficult to quantify.
On the other hand, referring to a constant celerity (and equal to that calculated
in the absence of bubbles) and to conventional headlosses can lead to results too
much on the safe side.
134 WATER HAMMER SIMULATIONS

However, it is clear that new resistance terms have to be inserted; to evaluate


these new terms, many models are available in the literature, which have been
classified into six categories [79]. This new term may depend on:

the instantaneous variations of the average flow velocity V ;


the variations of both the average flow velocity V and the local acceleration
V /t;
the variations of the average flow velocity V , the local acceleration V /t
and the convective acceleration V /x;
the variations of both the average flow velocity V and the diffusive term 2 V /x2 ;
the variations of the average flow velocity V and of the flow history, i.e., the
record of the past accelerations;
the variations of the instantaneous velocities along the longitudinal and transver-
sal directions.

As it is not in the aim of the book to give a complete exposition of all the methods,
but only to show how the problem can be dealt, the most common models will be
shown and implemented, presenting their critical points also. The models presented
in the following are grouped under the name of instantaneous acceleration based
(IAB) in which the headlosses are evaluated adding a new term to that related to
steady flow, i.e.:
J = JS + JU (8.15)
where JS is the usual steady flow term and JU is the new unsteady formulation.
Brunone and Greco [80] evaluated this new term as:

Q
JU = k1 (8.16)
t

where k1 is an appropriate damping parameter, with the condition



Q
if sign(Q)  = sign then k1 = 0 (8.17)
t

This model has been changed [81, 82] as the need for keeping in account also the
convective acceleration has emerged; therefore, the new formulation is:

k2 c V
JU = 1 (8.18)
g V t

where:
 V 
V =  V
t
 (8.19)
x

and k2 is again an appropriate damping parameter, normally different from k1 .


EFFECTS OF AIR AND CAVITATION 135

Pezzinga [83] commented that the sign minus in eqn (8.18) is not justified and
in certain cases it can bring to an unrealistic increase of oscillation, and therefore
proposed the following:

k3 V V k3 c V
JU = + sgn V (8.20)
g t x g x
Again with k3 a damping parameter is indicated.
Starting with this model, Ramos [84] modified again the expression in order
to improve the model performances and to match experimental results. All these
models can be decomposed in order to separate the two accelerations (local and
convective), as follows:

kp V V ka c V
JU = + sgn V (8.21)
g t x g x

According to Ramos, the term g1 V t


does not produce any additional damp
to the signal, but it is responsible only of a change in the value of the celerity,
which implies a variation of the frequency of the signal. On the other hand, the
term gc V
x
is responsible only for the signal damping during the transient. There-
fore, as the physical effects associated to the two terms are different, it is possible to
separate them and to use two different parameters, which in eqn (8.21) are indicated
as kp and ka and that are generally different.
However, in the numerical tests we carried out, the two effects cannot be com-
pletely divided, as, even when changing the value of one parameter and keeping
constant the other, the carried out solution changes in both the dissipative effects
and the variation of velocity; therefore, the performances of this last model were
found not to differ too much from those of Pezzinga.

8.4.2 Parameters

The main problem in the use of these models is the choice of the appropriate value
for the damping parameter(s). To this end, in the scientific literature many sets of
values that reflects the different conditions are reported.
For instance, Brunone et al. [81, 82] suggest a single value k1 = 0.085; Bergant
and Simpson [79] found values within the range [0.085, 0.330]; Bughazem and
Anderson [85] within the range [0.065, 0.150] and Wylie [86] within the range
[0.02, 0.03].
The usual suggestion, however, is to calibrate this parameter using experimental
data, as no reliable analytical or graphical procedures to estimate a parameter are
available yet. Comparing the results of the above-mentioned models it has been
demonstrated that:
k1 100 k2 100 k3 (8.22)
Expression (8.22) tries to tie the parameters of the mentioned formulations; in the
following some empirical estimations of the required parameters is presented, with
136 WATER HAMMER SIMULATIONS

the note that the values that can be carried out with these procedure are strongly
approximated and should be verified.
Pezzinga [87] tried to build an abacus tying the parameter k to the plant char-
acteristics. As Vardy and Brown [88] have shown, the value of the parameter k
decreases when the Reynolds number Re increases; however, a number of different
parameters have to be introduced to properly describe the behaviour of k, which
are /D (the relative roughness of the pipe, where is the roughness and D is the
pipe diameter) and especially:

g J0 L
y0 = (8.23)
c V0

where J0 and V0 are the head grade line and the velocity in steady flow, respectively.
Varying these non-dimensional parameters, a number of abacus have been
drawn, similar to that presented in Figure 8.4, shown as an example.
With only one experimental test performed, Carravetta et al. [89] claims there
is the possibility to give an appropriate estimation of the required parameter. The
base of this method is the consideration that [90] the term related to steady flow
headlosses JS is negligible when compared with JU , at least in the cases of complete
closure that ends in few phases. Therefore, Carravetta et al. [89] tied the damping
parameter k1 to the peak values recorded in two consecutive periods ht and ht+1
as follows:
2
ht+1 1
= (8.24)
ht 1 + k1

k2

/D
0.05
0.03
0.015
0.008
0.003
0.001
0.0003
0.00005
0.00001

Re0
102 103 104 105 106 107 108

Figure 8.4: Abacus to estimate the damping parameter k (after Ref. [87]).
EFFECTS OF AIR AND CAVITATION 137

8.5 Laboratory experiments


8.5.1 Experimental set-up

In order to verify the effects of the sudden closure of a valve that creates depression
in a closed pipe, a very simple experimental plant has been built in the Hydraulic
Laboratory of the Politecnico di Milano, as sketched in Figure 8.2: it is a single
pipe that links two reservoirs, the elevation of the upstream reservoir being equal
to 17.7 m, while the elevation of the downstream reservoir is equal to 5 m.
A pipe which diameter is 300 mm is connected to the upstream reservoir; the
diameter of this pipe becomes 150 mm and the final part of the plant is constituted
of an iron pipe with diameter D = 52 mm and thickness e = 5.0 mm; this latter
pipe is wound in a roll of 26 coils for a total length of 90 m, and then connected to
the downstream reservoir.
The transient is recorded only in this latter pipe: the flow is interrupted by
means of a very fast electro-valve positioned on this pipe and pressures are recorded
through a transducer positioned immediately downstream the valve. The discharge
is measured with an electromagnetic device.
Tests carried out in steady flow showed a Strickler ks resistance parameter equal
to 110 m1/3 /s, i.e., Manning equal to 1/110 = 0.009091 m1/3 s. The value of
the celerity can be computed with the usual formula (2.10) given the pipe charac-
teristics and known that the iron elasticity modulus is E = 2.0 1011 N/m2 , water
compressibility = 2.14 109 N/m2 and water density = 1000 kg/m3 . It resulted
c = 1388 m/s.

8.5.2 Experimental tests

During the years many tests have been performed, and it is obviously not possible
to present all of them here. Among the many tests, five have been selected and dis-
cussed in this chapter, which are useful for a further discussion. The characteristics
of the five tests are shown in Table 8.1. Moreover, in Figures 8.58.9 the recorded
pressures are shown for each test.

Table 8.1: Characteristics of the experimental


tests discussed in the book.

Test no. Discharge Velocity Head U/S


() (l/s) (m/s) (m)
1 0.112 0.0527 5
2 0.228 0.1074 5
3 2.5 1.18 8.1
4 3.0 1.41 9.7
5 3.3 1.55 10.5
138 WATER HAMMER SIMULATIONS
20

15

10
Head (m)

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

5

10
Time (s)

Figure 8.5: Test #1.

30

25

20

15
Head (m)

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
5

10
Time (s)

Figure 8.6: Test #2.

As can be seen, tests 1 and 2 do not reach the value of zero absolute pressure
(10.33 m), but the pressure drops under the atmospheric, and therefore bubbles
are expected. Tests 35, instead, are characterized by water column separation
(cavitation): in this case the increase of the velocity cannot force an increase of
the depressions as obviously absolute pressure cannot fall below zero; however,
there is an increment of the time during which the water column separation occurs.
EFFECTS OF AIR AND CAVITATION 139
80

70

60

50

40
Head (m)

30

20

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
10

20
Time (s)

Figure 8.7: Test #3.

100

80

60
Head (m)

40

20

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

20
Time (s)

Figure 8.8: Test #4.

When the water column reattaches, a large turbulence is recorded and then the
oscillations around the mean value start.
In test #1, the temporal distance between the first two pressure peaks is equal
to 0.15 s. It is therefore possible to compute the celerity:

2L 2L 2 90
= c= = = 1200 m/s (8.25)
c 0.15
140 WATER HAMMER SIMULATIONS
100

80

60
Head (m)

40

20

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

20
Time (s)

Figure 8.9: Test #5.

As can be seen, the experimental value of the celerity recorded in test #1 is much
lower than that theoretically computed above, as expected because of the presence
of air bubbles.

8.6 Computer code


A computer code has been developed with different options in order to check
whether it is possible to simulate these cases with the methods described so far.
The computer program has been developed starting from the code that uses the
LaxWendroff method for the internal points and the Ghost-Cell Zero-Order for
the boundary conditions. It implements the Pezzingas method [83] as presented
in Section 8.4.2 and in the following better specified. Moreover, the possibility of
evaluating the cavitation is given.
For the sake of brevity, in this book not all the passage are reported to implement
the method of Pezzinga with the LaxWendroff s integration scheme, but only the
final formulae:

g k3 c
Vjn+1 = Vjn (hnj+1 hnj1 ) (Vj+1
n
Vj1
n
)
2 1 + k3 1 + k3
2 V n g
j k3 c g
+ (hnj+1 2hnj + hnj1 )
2 1 + k3 (1 + k3 )2
  
c2 k3 c 2 g
+ + (Vj+1 2Vj + Vj1 )
n n n
t Jjn
1 + k3 1 + k3 1 + k3
(8.26)
EFFECTS OF AIR AND CAVITATION 141

Figure 8.10: Users interface for the program to simulate unsteady flow headlosses
and cavitation.


c2
hn+1
j = Vj (hj+1 hj1 ) +
hnj n n n
(Vj+1 Vj1 )
n n
2 g

2 c2
+ (Vj ) +
n 2
(hnj+1 2hnj + hnj1 )
2 1 + k3
  
Vjn c2 k3 c 3
+ (Vj+1 2Vj + Vj1 )
n n n
(8.27)
g g (1 + k3 )

As can be seen in Figure 8.10, which reports the usersinterface, in the computer
program there is the possibility to insert two coefficients k: the former, named
Pezzingas coefficient, is used when cavitation does not occur; if there is cavitation,
then a different coefficient is used. In the latter case, however, it is not possible
to estimate this parameter through relations as those of Carravetta and Vardy and
Brown (see Section 8.4) as the flow characteristics are quite different.
To reproduce test #1, the first method which has been applied is the standard
one, with no rupture of the water column and no added resistance terms. Then,
the method of Pezzinga is used, letting k = 1.5. The results from the application of
these two methods are compared with the results of test #1 and shown in Figure 8.11.
As can be seen, in the classical model, as presented in Chapters 4 and 5, the oscil-
lations are not appropriately damped by the steady flow resistance term; moreover,
there is a slight difference in the celerity of the waves. With the added resistance
142 WATER HAMMER SIMULATIONS
20
Experimental data
Classical model
15
Resistance terms, k  0.1

10
Head (m)

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0

5

10
Time (s)

Figure 8.11: Test #1 compared with classical model (no added resistance terms)
and with the model where the Pezzingas resistance term is added.

term, the improvement is highly visible, and the simulated pressures are very close
to those recorded. However, it can be observed as the presence of the resistance
term is also influencing the celerity of the wave, as it has been already discussed.
It should also be noted that the maxima values recorded from the experiments
are higher than the theoretical values. This phenomenon is probably due to the
turbulence, as will be discussed in Chapter 9.
Obviously, the simulation of the experiments where cavitation occurs is more
difficult; to do that with the computer program, it is necessary to tick the option
Water Column rupture in the usersinterface and to choose an appropriate parameter
for the model to be used in case of cavitation.
Different parameters have been used for the simulations as reported in
Figure 8.12.
The first simulation has been performed using the same values for the two
parameters and equal to that used for the above reported simulation of test #1
(kPezzinga = kCavitation = 0.1). As can be seen, the time when the column reattaches
is perfectly estimated, but unfortunately the peak is highly overestimated. Moreover,
while in the real data the headlosses due to the turbulence that follows the water
column reattachment are enough to avoid subsequent separations, the computer
model does not dissipate sufficient energy and therefore behaves as if it was a series
of ruptures.
As mentioned above, the increase of the value of k brings to a larger damping of
the oscillations but also influences the celerity. In the second simulation the values
kPezzinga = 0.1 and kCavitation = 1.5 have been used. Here, the time when the water
column reattaches is delayed; however, the value of the peak is decreased and,
EFFECTS OF AIR AND CAVITATION 143
140 Experimental data
kPezzinga  kcavitation = 0.1
120
kPezzinga  0.1; kcavitation  1.5
kPezzinga  kcavitation  1.5
100

80
Head (m)

60

40

20

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
20
Time (s)

Figure 8.12: Test #3 compared with the model where the Pezzingas resistance term
is added, for different parameters.

therefore, more similar to that recorded. In any case, however, the energy dissipa-
tions that happens when the water column reattaches are much higher than those
simulated by the model.
The need to differentiate the damping parameter k in case of cavitation is clear
observing the third simulations in the same figure: the results carried out leaving
kPezzinga = kCavitation = 1.5 are too damped towards the end of the simulation, i.e.,
when the cavitation phenomena are concluded.
However, even if playing with the parameters the results may be improved
(and the reader is invited in trying), it is quite clear that this topic is still very
uncertain and researchers should investigate it more deeply.
This page intentionally left blank
Chapter 9
Advanced models

As it has been seen in Chapter 8, classic models fail when applied to complex cases.
To this end, more complex models have been developed, even if the research in this
field is still open and the efforts of the researchers working all around the world are
yet to be considered concluded.
In this chapter two-dimensional (2D) models are shown, and the need to keep in
due account the effects of turbulence are also shown, which demonstrates in some
cases there might be the need to move towards three-dimensional (3D) models.
Obviously, these models have drawbacks which will be described briefly.

9.1 2D models
The evaluation of the headlosses during transients can be more adequately per-
formed through 2D or quasi-2D models, even if the computational times increase.
In these models it is possible to keep in due account the non-uniformity of the
velocity distribution in the transversal section.
Bratland [91] proposed the so-called cylinder model for non-laminar flow, in
which the continuity equation is solved in one dimension and the momentum equa-
tion in two dimension. This formulation has been extended to the turbulence flow
by Modica and Pezzinga [92] and Pezzinga [93], who allowed to clarify the nature
of the simplifications.
The reference system in two dimensions is shown in Figure 9.1; the equations
that describe the phenomenon have to be rewritten using this system.

Figure 9.1: Two-dimensional reference system.


146 WATER HAMMER SIMULATIONS

9.1.1 Continuity equation

The axially symmetric two-dimensional field in an elastic circular pipe can be


described by means of a quasi-two-dimensional model based on the continuity
equation written as [94]:

(u) 1 (rv)
+ + =0 (9.1)
t x r x
where x is the distance along the pipe, r is the distance from the central axis and u
and v are the velocities along x and r, respectively.
Analysing the numerical results carried out by Vardy and Hwang [94], it resulted
that the velocity v along the radius and its derivative take very low values both in
laminar and turbulent flow, and that justifies the choice of Bratland [91] to write the
continuity equation neglecting the term 1r (rv)
x
; eqn (9.1), therefore, becomes:

u
+ =0 (9.2)
t x
This equation can be integrated along the transversal section, neglecting the
derivative of with respect to x and, therefore, obtaining:

H c2 Q
+ =0 (9.3)
t gA x

9.1.2 Momentum equation

The 2D momentum equations can be written as [93]:

u u u H 1 x 1 (r )
+u +v = g (9.4a)
t x r x x r r
v v u H 1 1 (rr )
+u +v = g (9.4b)
t x r r x r r

where x and r are the longitudinal and radial components, respectively, of the
normal stresses (pressures), and is the shear stress.
Also in this case it is possible to introduce some simplifications in the system.
In particular, Pezzinga [93] posed the following hypotheses:

1. As described in Section 9.1.1, the transversal velocity v can be neglected together


with all its derivatives;
2. In longitudinal direction the derivative of the normal stress x is also neglected,
together with the residual convective term;
3. For each time step a single value of the piezometric head is assumed in each
section, which means the variations of the stresses in the radial direction is
negligible, i.e.: H
r
= 0.
ADVANCED MODELS 147

The momentum equation can, therefore, be written as:

u H 1 (r )
+g + =0 (9.5)
t x r r

Finally, considering that dA = 2 r dr, eqn (9.6) becomes:

u H 2 (r )
+g + =0 (9.6)
t x A

The model is called quasi 2D because the dependant variable u is function of x,


r and t, while H is a function of x and t.

9.2 Headlossess
There is the need to develop a model for evaluating the shear stress inside the
pipe. In this section two families of model are presented.
The former is called two-layer turbulence model: in the generic cross section
two layers are identified, with different hydrodynamic characteristics, and for each
of them a flux law is developed. Closest to the pipe wall there is a delimited field
where viscous stresses are higher than those turbulents and, therefore, the flow
is laminar. Outside that zone, where turbulent stresses are prevailing, the flow is
obviously turbulent. In the scientific literature, a number of methods have been
developed, which divide this area into three or even five subareas, where different
flux laws are defined, in order to get as close as possible to the actual conditions.
The latter family of model presented is called low Reynolds number k model
that allows the further generalization of the shear stress evaluation but requires
higher computational times.
9.2.1 Pezzingas model

Referring to the works of Bratland [91], Vardy and Hwang [94] and then of Pezzinga
[93] who considered turbulent flow, the shear stress can be evaluated through a
model based on the Newtons law inside the laminar layer; outside this layer, the
hypothesis of the Prandtls mixing length is used:
 
u  u  u
= l 2   (9.7)
r r r

where l is the mixing length and is the kinematic viscosity. To evaluate the length
l the following expression has been used [49, 95]:

l y
= e(y/R) (9.8)
R R
where y is the distance from the pipe wall, R is the radius of the circular pipe and
is a parameter dependent on the steady flow Reynolds number Re0 .
148 WATER HAMMER SIMULATIONS
k
0.42
R/  507
0.41 R/  252
R/  126
0.40 R/  60
R/  30.6
0.39 R/  15
smooth
0.38 Eq. (9.9)

0.37

0.36 R
103 104 105 106 107 108

Figure 9.2: Parameter function of Re0 (after [93]).

Pezzinga [93] computes the parameter interpolating the experimental results


carried out by Nikuradse [57, 96, 97], as shown in Figure 9.2 and, therefore,
obtaining the equation:

83,100
= 0.374 + 0.0132 ln 1 + (9.9)
Re0

In order to compute the thickness of the laminar sublayer , it is convenient to make


the hypothesis that inside this layer the velocity is linear and outside it the velocity
profile follows a logarithmic law. Let the speed friction u tied to the shear stress
at the pipe walls w from the:

p
u = (9.10)

Inside the laminar sublayer the linear equation to describe the velocity is:

u u y
= (9.11)

The logarithmic profile of velocity outside the sublayer is defined by:



u
= 2.5 ln +B (9.12)

being the roughness and B is a parameter dependent on the parameter:

u
Re = (9.13)

ADVANCED MODELS 149

The thickness of the sublayer is computed requiring that the two profiles match, i.e.:

u
= 2.5 ln +B (9.14)

The parmeter B is dependent on Re by the expression:



3.32
B = 8.5 2.5 ln 1 + (9.15)
Re

9.2.2 k model

The model presented in the Section 9.2.1 can be generalized with the so-called low
Reynolds number k model. These families of models have been developed [98]
to describe the turbulence and integrate the NavierStokes equations simplified to
the Reynolds equations. Eichinger and Lein [99] applied this method to describe
the pressure fluctuations recorded during transients; this procedure has been lately
used, with modifications, by Pezzinga and Brunone [100] and Zhao and Ghidaoui
[101]. This method can be applied in a quite wide range of Reynolds numbers, from
laminar to purely turbulent flow. The shear stress is evaluated as:

u
= ( + t ) (9.16)
r

where t is the value of turbulent viscosity, computed with:

k2
t = fu cu (9.17)

k being the kinetic energy for mass unit and is a parameter for the dissipation
effects.
These variables are governed by the following transport relations [102]:
  2
k 1 t k u
r + r + =0 (9.18)
t r r k r r
  2
1 t u 2
r + ce1 f1 r + c1 f2 = 0 (9.19)
t r r k r r k

Pezzinga and Brunone [100] showed that in case of transients, the formulation that
produce best results is that of Lam and Bremhorst [103]. The parameters to be
applied with this method have been evaluated through semi-empirical formulations
and they are:

20.5
fu = [1 exp(0.0165 y )] 1 + (9.20)
Rt
150 WATER HAMMER SIMULATIONS
3
0.05
f1 = 1 + (9.21)
fu

f2 = 1 exp(R2t ) (9.22)

where: Rt = and y = ky . Boundary conditions for k and are: k = 0 and
K2
2
= rk2 at the pipe walls; k
r
= 0 and
r
= 0 on the central axis. Numerical
parameters have the following values [103]: cu = 0.09; k = 1; = 1; 1 = 1.44;
2 = 1.92. Finally, the initial values of k and are obtained imposing the steady
flow conditions.

9.3 Cavitation

In Chapter 8 the possibility to have vaporous or gaseous cavitation has been men-
tioned, reminding that, as water is normally saturated with air, mainly the latter
case happens in the real plants; moreover, Wylie [104] showed that gaseous cavita-
tion models are able to reproduce quite well the vaporous cavitation phenomenon.
As a consequence, in the following only that kind of model is described. The
presence of air, as shown in Chapter 8, increase the headlosses, already larger
than those expected using the steady flow formulas. In the scientific literature two
families of models can be found: those that analyse the effects of the free gas
in the fluid (release gaseous cavitation model) and those that analyse the ther-
mic exchanges between the free gaseous and the liquid phases (second viscosity
cavitation model).

9.3.1 Release gaseous cavitation model

Under this type fall those models that consider constant the mass of gas [11] and
keep in account the whole gas release process [105].
The density of the mixture can be evaluated as:

mRT mRT
= 1 W + g (9.23)
p p

where m is the mass for specific volume; R is the constant of the gases; T is the
temperature and p is the pressure. Considering the density of the mixture as function
of the pressure and of the mass of the gas free, the continuity equation in an elastic
pipe can be written as [105]:

p c2 mRT m c2 RT c2 Q
1+ 0 2 + c02 0 + 0 =0 (9.24)
t p t p A x

where c0 is the celerity for an homogeneous liquid. To apply a consistent numerical


scheme, the equation is to be written in conservative form. Introducing the variable
ADVANCED MODELS 151

, defined as:
p c2 mRT c2 m
= 0 + 0 (9.25)
g gp g
The continuity equation becomes:

 c02 Q
+ =0 (9.26)
t A x

Neglecting the convective term, the variation in time of the air mass can be computed
as constant, as mentioned, but the carried out results show that this model is not
able to correctly reproduce the experimental data. On the other hand, Zielke et al.
[106] introduced a model where the air mass is variable, proving the hypothesis
that there is an initial quantity of mass m0 and the water is air saturated; then the
gas is released following the law:

m 1
= (ps p) (9.27)
t RT

where is the solubility constant of the gas, is the relaxation time, ps the saturated
pressure and p the pressure of the gas.
Generally speaking, the models that consider the air mass as constant are able to
represent the main characteristics of the phenomenon and in particular the effects
of propagation of the increased deformability of the mixture; the models that try to
use a release law are potentially able to describe the energy dissipations not due to
the friction. Obtaining the value of the free mass from eqn (9.27) and inserting it
in eqn (9.25) it is possible to get the value of the pressure p:
'
+  + 4c02 mRT
p= (9.28)
2

Finally, the piezometric head is given by:

p + p v pa
H =z+ (9.29)
g

where pv is the vapour pressure and pa is the atmospheric pressure. It is to be


highlighted that to the gas pressure is added also the vapour pressure, as both are
present inside the bubbles, even if the vapour mass might be considered negligible
[107].
The problem is the determination of the initial mass m0 and of the relaxation
time . The former parameter influences the oscillation phases, while the latter
their dampening. The calibration of these parameters is very difficult and analytical
solutions that allow an empirical evaluation do not exist yet. So far, only estimations
based on experimental data are provided [105, 108].
152 WATER HAMMER SIMULATIONS

Finally, the momentum equation can be written under the hypothesis of a two-
dimensional field with axial symmetry in a circular pipe [93]:

u H 1 (r )
+g + =0 (9.30)
t x r r

Again, it is necessary to introduce a model that allows the definition of the shear
stress ; the models described in Section 9.2 can be used.

9.3.2 Second viscosity cavitation model

Some authors ascribed non-negligible effects to dissipations due to the thermic


exchanges between the free gaseous and liquid phases [109]. In order to evaluate
these effects, in the momentum equation a new term is added, called pseudo-
velocity or volumetric velocity . This term allows to merge the mechanical and
thermodynamic effects in only one expression, which can be easily inserted in the
momentum equation. In the following model [110], the free gas for unit volume
has been considered as constant [11] for the sake of simplicity.
The continuity and momentum equations, written under the usual hypotheses
of a velocity profile in axial symmetry are:

p c2 Q
+ =0 (9.31)
t A0 x

u H 1 (r ) 1 x
+g + =0 (9.32)
t x r r x
In this case only the term dependent on the volumetric viscosity is added, thus
obtaining [110]:

u H 1 (r ) 1 u
+g + =0 (9.33)
t x r r x x

Referring to [110] for details, the involved parameters can be computed as:
2
k 1 m R T c4 2 m R T c4
= + 2
= (9.34)
k p p2

p
c = c0 ' (9.35)
p2 + c02 m R T

where c0 is the celerity evaluated when m = 0, is the density of the mixture


water-vapour, which can be computed with eqn (9.23), is the relaxation time,
k is a coefficient dependent on the gas transformation, m is the mass of the gas
ADVANCED MODELS 153

for unitary volume, R is the constant of the gases, T is the temperature, p is the
pressure and  is the relaxation time of the equivalent gas.
The governing equations are therefore the continuity equation, as expressed in
Section 9.1.1, and the momentum equation which finally is:

u H 1 (r ) (c0 )2 m R T p
+g + + =0
t x r r x p2 + (c02 )2 m R T t
(9.36)

With a procedure similar to that used for the continuity equation, the variable 
can be defined as:
 
' p
 = (c0 )2 m R T arctg " (9.37)
(c0 )2 m R T

Moreover, imposing the pipe is circular and therefore dA = 2 r dr and the final
equation is:
u H 2 (r )  2 
+g + + =0 (9.38)
t x A x t
Again, the parameters needed to be used in the model should be calibrated on
experimental data, as no empirical relations are available yet.
It is to be reminded that in this expression the hypothesis also of constant
gas/vapour mass is made, while Cannizzaro and Pezzinga [111] extended the model
to be used also when this mass is variable, and therefore improving the results.

9.4 Numerical schemes


The integration of the equations is carried out using a cylindrical grid where width is
constant in the longitudinal direction, while the area is constant in the radial direc-
tion, as shown in Figure 9.1. Velocities are defined on the median circumference,
while the shear stresses are defined on the internal and external circumferences.
The celerity of the mixture in a deformable pipe is computed by:
c0
c= ' (9.39)
c02 mRT
1+ p2

The integration of the equations can be carried out by means of the method of
characteristics [94] although the dependence of the celerity on the pressure would
suggest more appropriate and different methods. As in the momentum equation for
a given grid the velocities in the neighbouring grids are present, for each transversal
section a non-linear equation system has to be solved. This implies the methods like
the LaxWendroff cannot be used, as they are valid for quasi-linear systems. Hirsch
[31] shows that it is still possible to extend the LaxWendroff scheme to non-linear
systems too, if the final equation is considerably different from that presented in
154 WATER HAMMER SIMULATIONS

the preceding chapters. Alternatively, to solve the problem the McCormack method
[76] has been successfully used; this method is second order accurate both in space
and time; the matrix to be solved is tri-diagonal and dependent on the unknowns
vector given by the velocities. The solution is obtained through an iterative method,
coefficients are computed at each iteration and the first trial values are let equal to
those computed at the preceding temporal step [112].
Generally speaking, the results carried out are significantly better than those
obtained with one dimensional models; unfortunately, if a one-dimensional model
takes few seconds to produce a solution, two-dimensional ones may take, for the
same problem, several hours. The time required to give a solution can even increase
if complex turbulence models, like k, are applied. An idea for further research
could be the development of one-dimensional code able to shift to a two-dimensional
when and where it is needed. This involves the determination of a threshold of
validity of one-dimensional models, or acceptability of its results, which so far is
not yet quantified in the scientific literature.

9.5 Further problems


A simple experimental plant has been set up in the Hydraulic Laboratory at Politec-
nico di Milano being very similar to that presented in Chapter 8, the valve to
interrupt the flow is positioned downstream the plant and the pressures are recorded
immediately upstream the valve with a very sensitive pressure transducer, able to
sample at 1000 Hz. The recorded pressures are shown in Figure 9.3: as can be seen,

4.0
Recorded data
3.5 Theoretical value

3.0

2.5
Pressure (bar)

2.0

1.5

1.0
0
0.5

0.0
4500 4600 4700 4800 4900 5000 5100 5200 5300 5400 5500
Time (ms)

Figure 9.3: Experimental data carried out with the closing valve positioned at
0.10 m from the pressure transducer.
ADVANCED MODELS 155
3.5
Recorded data
3.0 Theoretical value

2.5
Pressure (bar)

2.0

1.5

1.0

0.5

0.0
17.1 17.2 17.3 17.4 17.5 17.6 17.7 17.8 17.9 18.0 18.1
Time (s)

Figure 9.4: Experimental data carried out with the closing valve positioned at
1.50 m from the pressure transducer.

the peaks reach higher values than the expected, theoretical, computed with the for-
mulas presented in Chapter 2. This is unexpected and unjustifiable considering even
different profile of velocities that could be reproduced using the models described
in this chapter.
The high turbulence recorded is attributed to its proximity to the valve, which
in this configuration was equal to 0.10 m; therefore the distance between the valve
and the transducer has been increased to 1.5 m and the test repeated. Values are
slightly different because the flowing discharge (which has been measured with
an electromagnetic device) was not perfectly the same in the two tests. However,
when the valve is moved, the recorded data show an almost perfect agreement
with the expected theoretical value and the turbulence practically disappeared
(Figure 9.4).
That means, on one hand, that to reproduce the effects of turbulence, the models
that have been shown in this book, although complex, are still not appropriate, and
probably the interested user should try the most complex fully three-dimensional
codes, accepting to face all the related problem for instance, the choice of the
appropriate turbulence model, the complexity in building the model, the time
required for simulation that would practically prevent its use for complex networks.
On the other hand, the user should be aware of these effects and these peaks, but
also consider their limited extension in time and space, and therefore whether they
are significant for the plant to be designed or verified.
This page intentionally left blank
Chapter 10
Case studies

All the theories and formulas presented in this book can be applied to real cases;
in the following, studies of existing plants are presented in order to give an
idea of the approach that can be used. Three cases are presented with increasing
complexity.

10.1 Simple pressure pipe for petroleum


products in Djibouti
In this case the object of the study was the evaluation of the unsteady flow pressures
that can be found in a terminal pipe conveying petroleum products in a plant located
in Oriental Africa. The study has been performed with an elastic model integrated
using the method of characteristics.

10.1.1 Plant characteristics

The plant to be modelled consisted in a simple pipe, with the following boundary
conditions: upstream one or two pumps operating in parallel and downstream a
motorised valve.
System characteristics are as follow. Pipe diameter is D = 16 = 400 mm (but
in the last 20 m where D = 12 = 300 mm); total length is equal to L = 2200 m;
pipe thickness is equal to e = 7.92 mm. Pipe is built in carbon steel, with elasticity
modulus equal to E = 2.0 1011 N/m2 and its junctures are welded.
Transported fluid is hydro carbonic combustible derived by petroleum, with
density equal to = 850 kg/m3 and compressibility coefficient equal to = 1.5
109 N/m2 .
In steady flow, discharge Q and velocity V in the pipe are, respectively, equal to
Q = 750 m3 /h V = 1.66 m/s (where D = 16 ) when only one pump is working.
When two pumps are working these values can be simply doubled1 , obtaining
Q = 1500 m3 /h V = 3.32 m/s.

1
Note that this simplification is not strictly correct if the pumps are working with fixed
velocity; however, it was discussed with the client, considered on the safe side and finally
accepted, especially because the headlosses were not known.
158 WATER HAMMER SIMULATIONS

As in the last end of the pipe the section is smaller, outfall velocity is equal to
V = 2.95 m/s (when only one pump is working) and V = 5.89 m/s (when the two
pumps are working together).
Pressure downstream the pump is p = 1.1 106 Pa.

10.1.2 Expected scenarios

Expected movements of the devices in the plant that have been simulated are:

1. the downstream motorised valve gets closed and upstream pumps are still
working (they will be switched off at the end of the transient);
2. pumps are switched off and then the motorised valve is closed (case of pumps
failure);
3. pumps are switched on and then the motorised valve is open.

Due to the lack of more precise information on the valve closure, the linear clo-
sure hypothesis is adopted, using the law mentioned in Chapter 4. This hypothesis,
however, is quite cautious and thus can be adopted.

10.1.3 Case 1

The first scenario is with the upstream pumps working and the downstream valve
getting close. This is the normal procedure adopted to switch off the pumps. The
static head (around which the unsteady flow pressures oscillate) is equal to the head
of the pump when the discharge is zero. From pump curves it is deduced that this
value is equal to 12 atm.
Headlosses in the pipe have been computed so that at the outfall the pressure
is equal to 1 atm, i.e., it is assumed that the upstream pressure (of the pumps) is
necessary and sufficient to convey the discharge Q = 1500 m3 /h.
In Figure 10.1 the results of this simulation have been reported; it can be seen
that the unsteady flow pressures oscillate, as expected, around the static head equal
to 12 atm. On the other hand, the maximum pressure reached is lower than that
computed with the Michauds formula in Chapter 2, which brought 15.45 atm;
instead, the maximum pressure value reached with the computer program is equal
to 14.34 atm. This result is ascribed to the headlosses in the downstream end of the
plant, where the pipe has smaller diametre.

10.1.4 Case 2

The second scenario described in the Section 10.1.2 is when the upstream pumps are
turned off while the downstream valve is still open. The computer program is built
modelling the upstream pumps as if they were a valve that closes in fixed times,
selected in Tc = 15 and 30 s, which is the range of time in which the discharge
is expected to stop. The model was simplified because no precise information
CASE STUDIES 159
16

14

12
Pressure (atm)

10

0
0 10 20 30 40 50 60
Time (s)

Figure 10.1: Results of simulations for case 1.

were provided. Downstream the liquid outfalls in the atmosphere with concentrated
headloss equal to V 2 /2 g; the coefficient has been calibrated in order to
have atmospheric pressure in the downstream section of the plant, in steady flow
conditions. The value of the coefficient carried out in steady flow conditions has
also been used in unsteady flow simulations.
Simulations show that in both cases (Tc = 15 and 30 s) pressures fall below
absolute zero. As largely discussed these are non-physical values: after reaching
the points shown in Figure 10.2 indicated as Vacuum, the numerical model does
not produce realistic results.
It has to be noted that this case should be avoided and that the correct sequence
for turning off the pumps is that described in case 1. This second scenario happens
only in case of pumps failure, and therefore not very frequently. The curves in
Figure 10.2 are related to the case of both pumps failure (power failure); however,
simulations also showed that when only one pump is working, the pump failure
brings the system to cavitation, even if of less important magnitude.

10.1.5 Case 3

The third scenario described in Section 10.1.2 is the opening of the downstream
valve while the upstream pumps are already working. This is the correct starting
procedure. The downstream valve is modelled with linear opening law, considering
it is completely open in 36 s.
In Figure 10.3 the results are shown; it can simply be observed that the pressures
decrease, never reaching vacuum, and then there are oscillations around the steady
flow value.
160 WATER HAMMER SIMULATIONS
14
12 Closure: 15 s

10 Closure: 30 s

8
Vacuum
6
Pressure (atm)

4
2
0
0 10 20 30 40 50 60
2
4
Pressures below 1 atm are not
6
possible. Left as model output.
8
Time (s)

Figure 10.2: Results of simulations for case 2.

14

12

10
Pressure (atm)

0
0 10 20 30 40 50 60
2
Time (s)

Figure 10.3: Results of simulations for case 3.


CASE STUDIES 161

10.1.6 Conclusions

The problem of transients in a very simple plant has been studied by means of the
elastic model equations integrated with the method of characteristics.
Different scenarios have been drawn and then evaluations have been carried
out. Results are quite consistent with those obtained with empirical methods (slight
differences are noted); carried out results show pressures are within the range of
acceptability for the construction plant standard, if the computational hypothe-
sis assumed can be considered valid and if the correct sequences of starting and
stopping of the plant are applied.
A problem of cavitation might arise in case of pumps failure, which is obviously
not the correct sequence of stopping the system and therefore that happens quite
rarely. However, according to the client, this outcome can be also accepted.
In normal operations, carried out results show pressures are within the
acceptable limits and therefore the plant could be considered safe.

10.2 A more complex example for seawater


treatment plant in Tanzania
For the design of a seawater treatment plant in Tanzania, we have been asked
to verify the steady and unsteady flow conditions of the pumping station both in
existing and design conditions. Moreover, we have been asked to determine whether
any remedial is required for the pumping station and, in this case, to compute its
type and dimensions.
In this paragraph, after describing the sketch of the two systems as they have
been implemented in the computing code, some results of unsteady flow simulations
are presented.
The mathematical model used for unsteady flow simulations is based on the
method of characteristics.

10.2.1 Plant characteristics

The two plants (existing and to be designed) are shown in Figures 10.4 and 10.5
with their computational model.
As can be seen, some pipes and junctions had to be inserted into the model in
order to keep in account particular devices as the air chambers.
Pumps characteristics are reported in Table 10.1, while pipes are described in
Table 10.2 and nodes in Table 10.3.
In Table 10.2 celerity has been computed with the usual formula (2.10) where
the parameter is equal to 1.
Finally, note that the existing air chamber volume is equal to 6.37 m3 : but this
is the total volume of the chamber, while the volume of the air inside the chamber
162 WATER HAMMER SIMULATIONS

P1

P2 J1
AC1
P3
01 02

LEGEND:

Tank Junction

Pump Air chamber

1 4 10
12 18

15 20
9
19
17 21
3 6 7 11
14

Figure 10.4: General sktetch and numerical implementation existing plant.

is smaller. As we do not know that volume, we suppose it is about he half, so equal


to 3.2 m3 .

10.2.2 Unsteady flow simulations: existing plant

In the case of the existing plant, it can be easily seen that the times in which
the pumps stop (approximately equal to 20 s) is much higher than the phase time
of the system, so that unsteady flow pressures are quite low.
In Figure 10.6 is shown the static head (equal to the elevation in the down-
stream tank) and the pressures in the upstream end of pipe 15 (see Figure 10.4,
lowerpart). At the beginning (time t = 0) the head is higher than the static value;
then as the pumps stop, there is a sudden drop in the pressure; afterwards, the
head oscillates around the static value, as expected. As can be seen, oscillations
are very small. Other pipes have similar values. During the simulations the clapets
CASE STUDIES 163

P4

J2
P5
AC2
P6

P4 Line J1(a)J3(a) J3

P5 J1 03
AC1
P6 Line J1(b)J3(b)

LEGEND:

Tank Junction
Pump Air chamber
29

22 25 36
30
33
37
38
35
24 27 28
32

8
39 48

18 40
43
44 46 49
1 4 9 42
12 15 19
45 47
7 41

Figure 10.5: General sktetch and numerical implementation plant to be designed.

positioned downstream the pumps close in about 6 s, even if the pumps continue
their rotating motion.
Finally, we checked these results against the Michauds formula. As well known,
this formula is rigorously valid only for a single pipe; however, in this case we let
164 WATER HAMMER SIMULATIONS

Table 10.1: Pumps characteristics.

nreg Qreg Hreg


Pump (rpm) (m3 /h) (m)
P1 1450 1500 78
P2 1450 1500 78
P3 1450 400 85
P4 1450 1500 85
P5 1450 1500 85
P6 1450 1500 85

Table 10.2: Pipes characteristics.

From Diameter Length n Thick. Cel.


node To node (m) (m) (m1/3 /s) (mm) (m/s)
P1 AC1 0.40 3 80 6 939
P2 AC1 0.40 3 80 6 939
P3 AC1 0.35 3 80 6 976
AC1 Air cham. 0.55 12 80 6 850
AC1 J1 0.60 35 80 6 826
J1 O1 0.60 525 80 6 826
J1 O2 0.35 525 80 6 976
P4 AC2 0.60 3.5 80 6 826
P5 AC2 0.60 3.5 80 6 826
P6 AC2 0.60 3.5 80 6 826
AC2 Air cham. 0.60 10 80 6 826
AC2 J2 0.80 15 80 6 745
J2 J1 0.60 14 80 6 826
J2 J3 0.60 540 80 6 826
J1 (a) J3 (a) 0.60 525 80 6 826
J1 (b) J3 (b) 0.35 525 80 6 976
J3 O3 0.90 30 80 6 713

L = 550 m, V = 1.6 m/s and obviously Tc = 6 s, obtaining:

2 L V0 2 550 m 1.6 m/s


h = = = 30 m
g Tc 9.806 m/s2 6

As expected, the carried out value does not perfectly agree with those obtained
through the simulation code.
CASE STUDIES 165

Table 10.3: Nodes characteristics.

Node Elevation (m) Notes


P1 1137
P2 1137
P3 1137
AC1 1133.75 Air chamber: 6.37 m3
J1 1136.62
O1 1205.147
O2 1205.147
P4 1136
P5 1136
P6 1136
AC2 1135.69
J2 1135.69
J3 1135.69
O3 1204.78

1230

1225

1220

1215
Head (m asl)

1210

1205

1200

1195

1190
Pipe 15upstream
1185 Static head (downstream tank)
1180
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 10.6: Existing plant: pressures.

10.2.3 Plant to be designed

In the following, results of simulations for the plant to be designed are shown having
considered three cases: no air chamber, an air chamber of 3 m3 and an air chamber
of 5 m3 .
166 WATER HAMMER SIMULATIONS
1230

1225

1220

1215
Head (m asl)

1210

1205

1200

1195

1190
Pipe 44upstream
1185 Static head (downstream tank)
1180
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 10.7: Pressures in the existing pipes, having eliminated the air chambers.

At first, results presented are related to pipes 44 and 45 because they are already
existing and the new upgrade must not worsen the present situation (in the figures
only the values for pipe 44 appear as pipe 45 values overlap). Then results for the
pipes 38, 39 and 48 are also presented, as they are the main pipes in the new plant.
Note that the referred air chamber volume is related to steady flow conditions,
while it obviously changes during transients.
Also in this case simulations have been carried out considering the presence of
a clapet positioned downstream the pumps, which closes when water velocity gets
negative.
It has to be noted that in some points the piezometric line intersects the pipe:
this means that sometime, during transients, pressures can fall below atmospheric.
In these cases, vent valves open and let air entry the pipe; this air can change the
systems behaviour, as discussed in Chapter 8.

10.2.4 No air chambers

As expected, increasing the discharge, and so the velocity, brings to higher pres-
sures to be recorded during transients. However, changes are not dramatic, and
probably the key factors are the short length of the pipe (less than 600 m) and the
time of the valve closure (more than 15 s). For the new pipes, similar considera-
tions may be drawn (Figure 10.8). When the velocity becomes negative, the clapet
positioned downstream the pump closes and waterhammer pressure oscillations
become regular.
CASE STUDIES 167
1230

1225

1220

1215
Head (m asl)

1210

1205

1200

1195

1190 Pipe 38upstream


Pipe 39downstream
1185 Pipe 48downstream
Static head (downstream tank)
1180
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 10.8: Pressures in the new pipes, having eliminated the air chambers.

1230

1225

1220

1215
Head (m asl)

1210

1205

1200

1195

1190
Pipe 44upstream
1185 Static head (downstream tank)
1180
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 10.9: Pressures in the existing pipe with 3 m3 air chamber.

10.2.5 Air chamber 3 m3 volume

As can be seen in Figure 10.9, due to the distance between the new air chamber
and the old pipe, the effect of the former is small. Anyhow, the presence of an
air chamber helps in slowing down the system response and in cutting the higher
frequency waves.
168 WATER HAMMER SIMULATIONS
1230
1225
1220
1215
Head (m asl)

1210
1205
1200
1195
1190
1185
1180
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 10.10: Pressure values in the existing pipe with 5 m3 air chamber.

1230

1225

1220

1215
Head (m asl)

1210

1205

1200

1195

1190 Pipe 38upstream


Pipe 39downstream
1185 Pipe 48downstream
Static head (downstream tank)
1180
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 10.11: Pressure values in the new pipes with 5 m3 air chamber.

10.2.6 Air chamber 5 m3 volume

For the reason mentioned in Section 10.2.4 (the distance between the new air cham-
ber and the old pipe) an increase in the air chamber volume has little effect in
reducing the overpressure peaks; that can be seen in Figure 10.10 for the existing
pipe and in Figure 10.11 for the new pipes.
CASE STUDIES 169

The reduction of the pressures is not dramatic, probably because waterhammer


effects are already quite small.
The main effect in installing an air chamber in this position is probably the
increase of the response time of the system, which become less sensitive to higher
frequency perturbations.

10.2.7 Conclusions

In this Section unsteady flow simulations are performed with different hypoth-
esis, related to the possible scenarios for a seawater plant to be potentiated in
Tanzania.
Results show pressure values are within the acceptable range and no cavitation
occurs; anyway, in some points of the pipe pressures drop under atmospheric during
transients. In these points the positioning of an air valve was designed. Note that
the presence of air inside the pipelines changes remarkably the system behaviour
in a way that cannot be properly computed with current knowledge.
The implementation of the new plant obviously worsen the existing pipe
conditions, but it seems this variation is not dramatic.
Moreover, it seems that the size of the new air chamber does not affect much
the pressure values in the existing pipe, probably because of the distance between
these two elements.
Anyway, a new air chamber installation is suggested (volume equal to 3 m3 ),
in order to smooth the higher frequency perturbations. It has also been shown
that the analysis of this plant, performed quite easily with the computer pro-
grams developed in this book, could not have been performed with empirical
methods.

10.3 A very complex example for seawater


treatment plant in Algeria
10.3.1 The plant to be modelled

The seawater plant to be modelled is located in Algeria and it is intended to supply


potable water to 25% of the Algerian capital citys population. The project is the
first private reverse osmosis potable water project in Algeria and the largest mem-
brane desalination plant in Africa. The plant is fed with open intake Mediterranean
seawater.
As can be seen in Figure 10.13, the plant comprises nine first-pass trains,
each capable of continuously producing 25,100 m3 /day of permeate. In this figure,
the intake is depicted by the Storage tanks (S), positioned downstream of which
are the main pumps, which guarantee the upstream minimum pressure (NPSH)
required for the functioning of the following pumps.
WHS
170 WATER HAMMER SIMULATIONS

CH010.tex
Figure 10.12: Sketch of the seawater treatment plant.

S J J J J J J J J J
I P
FILTER

24/8/2013
J J J J J J J J J J
I P
ERI ERI ERI ERI ERI ERI ERI ERI ERI

B B B B B B B B B
RO RO RO RO RO RO RO RO RO

12: 11
H H H H H H H H H
I P O O O O O O O O O
FILTER
J J J J J J J J J J
I
REMINERALIZER
P
VESSEL
J J J J J J J J J O

Page 170
J Junction
P Main Pump
B Booster
H High Pressure Pump
O Outfall
I Intake
Generic Valve
WHS
CH010.tex
Figure 10.13: Numerical implementation of the seawater treatment plant.

60 101 144 187 230 273 316 359 371


103 146 189 232 275 318 361 373
1 5 J J J J J J J J O
11
I P 15
9 19 21 62 64 104 107 147 150 190 193 233 236 276 279 319 322 362
23 25 66 109 152 195 238 281 324

24/8/2013
J J J J J J J J J
2 6 36 77 120 163 206 249 292 335
12 76
16 45 35 44 34 86 85 129 119 128 172 162 171 215 205 214 258 248 257 301 291 300 344 334 343
I P 27 J J 68 J J 75 J J 118 J J J J 204 J J 247 283 J J 290 J J 333
111 154 161 197 240 326

30 32 71 73 114 116 157 159 200 202 243 245 286 288 329 331
40 31 41 81 72 82 124 115 125 167 158 168 210 201 211 253 244 254 296 287 297 339 330 340
J J J J J J J J J J J J J J J J
52 57 93 98 136 141 179 184 222 227 265 270 308 313 351 356
49 90 133 176 219 262 305 348
46 B 87 B 130 B 173 B 216 B 259 B 302 B 345 B
53 56 94 97 137 140 180 183 223 226 266 269 309 312 352 355

12: 11
47 394 91 134 177 220 260 263 306 349
J 398
399 395 88 J 99
390 387 131 J 142
407 403 174 J 185
416 412 217 J
228
425 421 J 271
434 430 303 J 314
443 439 J 357
452 448
55 396 J 96 388
J 139 405
J 182 414
J 225 423
J 268 432
J 311 441 354 450
48 401 400 89 392 391 132 410 408 175 419 417 218 428 426 261 437 435 304 446 444 347 455 453
H 402
H 393
H 411
H 420
H 429
H 438
H 447
H 456
3 7 51 383 404 413 422 431 440 449
13
54 95 138 181 224 267 310 353
I P 17 O O O O O O O O
10 20 397 63 389 386 406 409 415 418 424 427 433 436 442 445 451 454
24 26
J J 67 J 110 J 153 J 196
J 239
J 282 J 325
J
4 8 22 65 108 151 194 237 280 323
14 18

Page 171
I P
102 145 188 231 274 317 360 366 372
105 148 191 234 277 320 363 369 374
61 J J J J J J J J O

J Junction

P Main Pump

B Booster

H High Pressure Pump

CASE STUDIES
O Outfall

I Intake
Generic Valve (Concentrated Headloss)

171
172 WATER HAMMER SIMULATIONS

The membrane elements (RO) of each train are fed by a dedicated high pressure
(HP) pump, a booster (B) pump, and an array of 32 ERI2 model PX-220 Pressure
Exchanger energy recovery devices. All nine pumps are fed from a common supply
manifold, which is maintained at a pressure greater than 3.5 bar in order to meet
the minimum NPSH requirement of the HP pump.
The PX device arrays are fed from a second supply manifold, which is
maintained at a pressure of greater than 1.3 bar in order to meet the minimum
discharge-pressure-requirement of the PX devices. Because the PX device arrays
and HP pumps are supplied separately, the PX device arrays are isolated from flow
and pressure variations that may occur as individual HP pumps go on- and off-line.
The HP pumps selected for the plant operate at 1084 m3 /h and approximately
88% efficiency; the booster pumps operate at 1351 m3 /h and approximately 89%
efficiency.
The RO system divides the flow into two parts, one of which (permeate) is
collected to the remineralizer vessel and then to an outfall storage, while the other
(brine) is conveyed through the ERI device to recover energy and then disposed.

10.3.2 A peculiar device: energy recovery PX

In a reverse osmosis (RO) water desalination system, a high percentage of water


(brine) has to be discarded, consuming a great deal of energy used to obtain the
high pressure necessary in these plants. As a consequence, many systems have been
developed in order to recover this energy.
One of the most advanced devices is implemented in the ERI PX developed for
specific use in sea water reverse osmosis plants and transfers pressure from a high
pressure line to a low pressure one [113].
The Algerian desalination plant, designed by General Electric Water section,
was at that time the second biggest seawater reverse osmosis desalination plant in
the world, and it used one array of 32 PX-220 devices on each of the nine first-pass
trains for a total of 288 PX units, the purpose being to minimize the foreseen high
costs of processing the expected 200,000 m3 /day of water. The first pass train size
is to be 25,100 m3 /day, so that the plant is able to operate at full capacity with just
eight trains running [114].
In order to build a complete model and to simulate waterhammer processes, a
laboratory investigation of the PX behaviour during flow transients was performed,
as its behaviour in unsteady flow conditions was never been tested before, and these
results were used for the subsequent simulations.

10.3.3 Laboratory plant

In order to verify the behaviour of PX during unsteady flow, several experimental


tests were performed in the Hydraulic Laboratory of the Politecnico di Milano [115].

2
PX, PX Pressure Exchanger, ERI and the ERI logo are registered trademarks of Energy
Recovery, Inc.
CASE STUDIES 173

Low pressure line

Opening / Tank
closing valve Flowmeter

Pressure
transducer

ERI-PX

Flowmeter

Pressure
transducer Pumps

Outlet High pressure line

Figure 10.14: Sketch of a configuration of the experimental plant (valve on the


high pressure line).

A system of pumps brought water to the upstream tank positioned about 9 m


above the underground floor. A pipe with diameter D1 = 300 mm was derived from
this tank; a reduction brought the diameter to D2 =150 mm and a third reduction
brought the diameter to the final one D3 = 2 . This was the low pressure circuit.
A similar high pressure circuit was regulated by two high pressure pumps (see
Figures 10.14 and 10.15).
The final parts of the plant consisted of two iron pipes with diameter D3 = 2
of spiral shape with 26 coils for a total length L of about 90 m. These pipes
were installed in order to allow the development of the waterhammer waves in
all reflection phases. Celerity of the wave was evaluated as equal to 1388 m/s.
There followed two electromagnetic flowmeters, an electro-valve able to close
and open almost instantaneously, and the pressure transducers connected to the
data collector and the computer. Tests were performed with the valve situated
in different positions on the two lines, in order to fully investigate the devices
behaviour.

10.3.4 Laboratory tests

Numerous tests were performed by opening and closing the valve in different posi-
tions: on the high pressure or low pressure lines, and upstream or downstream of
the coils (in these cases always downstream of the ERI device). Moreover, the valve
174 WATER HAMMER SIMULATIONS

High pressure pumps

Opening/closing valve

Pressure
Flowmeters ERIPX 220
transducer

Figure 10.15: Laboratory plant. The valve is located on the low pressure line
upstream of the coil and downstream of the EXI PX.

was also positioned upstream of the PX. In what follows, only the tests judged most
significant are described.
During the tests, discharges were set the same for the two lines and equal to
3 l/s, which gave a velocity equal to 1.41 m/s.
Figure 10.16 shows the results of a test with the closure valve positioned down-
stream of the spiral pipe of the high pressure line. As expected, a large development
of pressures on the high pressure line can be observed. On the other hand, overpres-
sures on the low pressure line are also noticeable, but they are almost negligible.
The transient ends in a few seconds before the pumps reach the static head.
The reliability can be verified by comparing the peak recorded (Figure 10.16)
with the expected value computed with the AllieviJoukowski formula:

c 1388
h = V = 1.41 = 200 m 20 bar
g 9.806

Similar tests were performed with the closure valve positioned on the low pres-
sure stream. Moreover, the valve was opened on the two lines so as to have initial
depression values. The results show that in this case too the disconnection was
almost complete. When the valve on the high pressure line was opened, a depres-
sion occurred in the same line, but the effect on the low pressure line was practically
negligible. When the valve on the low pressure line was opened, the effect on both
lines was insignificant.
Repeatability tests were performed in order to verify the coherence between
results. In fact, in Figure 10.16 two tests are reported, performed with the same
conditions and which prove that the results are equivalent.
CASE STUDIES 175
25

HP line - Test 1
20 LP line - Test 1
HP line - Test 2
LP line - Test 2
Pressure (bar)

15

10

0
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 10.16: Pressure values recorded in the high and low pressure lines.
Instantaneous closure on the high pressure line.

10.3.5 Model of the seawater plant in Algeria

Analysis of the laboratory test results showed that, during transients, high pressure
and low pressure lines could be considered completely disconnected. This simpli-
fied the implementation of the mathematical model to simulate the actual plant. In
this case, the plant schematically represented in Figure 10.12 can be divided into
two independent parts where the ERI PX is located [115]. The resulting comput-
ing scheme is shown in Figure 10.13. Twenty tests, as listed in Table 10.4, were
performed with this mathematical model.
The results obtained for all tests showed that the plant was at no risk of dangerous
pressures or depressions even in the worst situations which could be foreseen.
For instance, simulation 5 modelled the stoppage of both the main pumps
positioned in the lower part of the plant (see Figure 10.13). As can be seen in
Figure 10.17, the pressure dropped and then started to oscillate in typical fashion,
but no problematic values were recorded.
A similar test (simulation 3) was performed by modelling the stoppage of the
two main pumps in the upper part of the plant (see again Figure 10.13). Owing to
the complete separation of the two parts of the plant produced by the ERI device,
there was no need to perform crossed stoppage (stoppage of one pump in the upper
part and one in the lower part; stoppage of both main pumps in the upper part and
both in the lower, and so on).
Simulation 10 modelled the whole train stoppage which means that the high
pressure and booster pumps stopped and the valve positioned on the RO system
is closed. A sudden fall in the pressure downstream of the RO could be observed,
176 WATER HAMMER SIMULATIONS

Table 10.4: List of all tests performed on the Hamma plant with the mathematical
model.

Device Nodes of the


whose trip mathematical
Test # was modelled Details about the models rules model involved
1 Remineralizer Closure valve Vrem : Tc = 5 s Node 366
vessel
2 Upstream Stopping of pump P1,sup Node 5
3 pumps Stopping of pumps P1,sup and P1,inf Nodes 5 and 6
4 Stopping of pump P2,sup Node 7
5 Stopping of pumps P2,sup and P2,inf Nodes 7 and 8
6 Stopping of pumps P1,sup P1,inf Nodes 5 6
P2,sup P2,inf 78
7 Train 1 Stopping of pump PHP Node 48
The following operations were Node 394
simultaneously performed:
a. Stopping of pump PHP
b. Closure of valve VRO,PERM : Tc = 5 s
8 Stopping of pump PBOOST Node 46
9 ROs breakage, VRO,PERM : Tc = 5 s Node 394
10 Whole train stopping
The following operations were
simultaneously performed:
a. Stopping of pump PHP Node 48
b. Stopping of pump PBOOST Node 46
c. Closure of valve VRO,PERM : Tc = 5 s Node 394
11 Train 4 Stopping of pump PHP
The following operations were
simultaneously performed:
a. Stopping of pump PHP Node 175
b. Closure of valve VRO,PERM : Tc = 5 s Node 177
12 Stopping of pump PBOOST Node 173
13 ROs breakage VRO,PERM : Tc = 5 s Node 177
14 Whole train stopping
The following operations were
simultaneously performed:
a. Stopping of pump PHP Node 175
b. Stopping of pump PBOOST Node 173
c. Closure of valve VRO,PERM : Tc = 5 s Node 177
(Continued)
CASE STUDIES 177

Table 10.4: Continued.

Device Nodes of the


whose trip mathematical
Test # was modelled Details about the models rules model involved
15 Train 8 Stopping of pump PHP
The following operations were
simultaneously performed:
a. Stopping of pump PHP Node 347
b. Closure of valve VRO,PERM : Tc = 5 s Node 349
16 Stopping of pump PBOOST Node 345
17 ROs breakage RO VRO,PERM : Tc = 5 s Node 349
18 Whole train stopping
The following operations were
simultaneously performed:
a. Stopping of pump PHP Node 347
b. Stopping of pump PBOOST Node 345
c. Closure of valve VRO,PERM : Tc = 5 s Node 349
19 All trains Stop of all trains at the same time
The following operations were
simultaneously performed:
a. Stopping of pump PHP Node 48
b. Stopping of pump PBOOST Node 46
c. Closure of valve VRO,PERM : Tc = 5 s Node 394
d. Stopping of pump PHP Node 89
e. Stopping of pump PBOOST Node 87
f. Closure of valve VRO,PERM : Tc = 5 s Node 91
g. Stopping of pump PHP Node 132
h. Stopping of pump PBOOST Node 130
i. Closure of valve VRO,PERM : Tc = 5 s Node 134
j. Stopping of pump PHP Node 175
k. Stopping of pump PBOOST Node 173
l. Closure of valve VRO,PERM : Tc = 5 s Node 177
m. Stopping of pump PHP Node 218
n. Stopping of pump PBOOST Node 216
o. Closure of valve VRO,PERM : Tc = 5 s Node 220
p. Stopping of pump PHP Node 261
q. Stopping of pump PBOOST Node 259
r. Closure of valve VRO,PERM : Tc = 5 s Node 263
s. Stopping of pump PHP Node 304
t. Stopping of pump PBOOST Node 302
u. Closure of valve VRO,PERM : Tc = 5 s Node 306
(Continued)
178 WATER HAMMER SIMULATIONS

Table 10.4: Continued.

Device Nodes of the


whose trip mathematical
Test # was modelled Details about the models rules model involved
v. Stopping of pump PHP Node 347
w. Stopping of pump PBOOST Node 345
x. Closure of valve VRO,PERM : Node 349
Tc = 5 s
20 Electricity failure stopping of
all pumps and VRO,PERM closure.
The following operations were
simultaneously performed:
a. Stopping of pumps P1,sup Nodes 5 6
P1,inf P2,sup P2,inf 78
b. Stopping of pumps PHP Nodes 48 89
132 175 218
261 304 347
c. Stopping of pumps PBOOST Nodes 46 87
130 173 216
259 302 345
d. Closure of valves VRO,PERM : Nodes 394 91
Tc = 5 s 134 177 220
263 306 349

70

60

50
Pressure (m)

40

30

20

10

0
0 5 10 15 20
Time (s)

Figure 10.17: Pressure values inside the pipe 17 (pipe 18 values overlap)
simulation 5.
CASE STUDIES 179
700

600

500
Pressure (m)

400
Pipe 452
300

200 Pipe 353

100

0
0 2 4 6 8 10 12 14 16 18 20
Time (s)

Figure 10.18: Pressure values inside the plant, simulation 19.

but even upstream of that system, where overpressures were expected, values were
within the acceptable range.
Figure 10.18 reports the results of simulation 19 when all trains were closed
(and hence was probably one of the most dangerous situations that may occur).

10.3.6 Conclusions

The Algerian desalination plant designed by General Electric Water section, the
largest seawater reverse osmosis desalination plant in Africa and the second largest
in the world, has been simulated using the methods described in this book. Such a
large plant requires careful design, and a hydraulic model for steady and unsteady
flow simulations has been developed accordingly.
A laboratory investigation was performed in order to investigate the ERI PXs
behaviour during transients, and its results have been reported in this book.
The laboratory results clearly show that the PX produced an almost perfect
hydraulic disconnection between high and low pressure lines (although from a
quality point of view the two lines were obviously not disconnected), and this
simplified the mathematical model implementation.
A sketch of the Algerian plant and its mathematical representations have also
been provided, with some of the results obtained from thorough investigation of the
problems most likely to occur.
The results show that the plant can be considered safe, from the hydraulic point
of view, in case of both steady and unsteady flows.
180 WATER HAMMER SIMULATIONS

10.4 Final remarks


This paragraph presented the applications of the methods to analyze complex plants
during transients, as described in this book.
Simplest plants, as the first presented in this paragraph for petroleum product
in Djibouti, show that the simplest models can be applied and the differences with
the results carried out with most complex models are negligible. However, when
the complexity of the plant increases, the oldest and simplest models are not yet
sufficient to describe the behavior of the plant, and the carried out results are not
appropriate to establish the measure to contrast the effects of waterhammer. As it
has been shown for the plant in Tanzania, the design of some devices, in that case
it was an air chamber, has to be performed through more complex models: with
them, different solutions must be tried in order to see the effect of the introduced
modifications and to select the most appropriate.
Some very complex plants are very difficult to be modeled. The Algerian
plant, shown as third example, included devices which behavior was not known in
unsteady flow conditions; this required the performance of laboratory tests. After
the tests were carried out, the numerical model has been built and the carried out
results allowed the evaluation of the plant performances.
However, field tests should always be scheduled and performed before starting
the normal activities of the plant because, as we tried to highlight in the whole
book, the topic is quite complex and not still resolved, and therefore a deep analysis
should be performed in most practical cases.
References

[1] Menabrea L.F., 1858, Note sur les effects de choc de leau dans les conduits.
C. R. Acad. Sci. Paris, 47, 221224.
[2] Michaud J., 1878, Coup de blier dans les conduits Etudedes moyens employs
pour en attnuer les effets. Bulletin de la Socit Vaudoise des Ingnieurs et des
Architects, 4(3), 5664; 4(4), 6577.
[3] Joukowsky N., 1898, Waterhammer. Mem. Imp. Acad. Soc. St. Petersburg
translated by O. Simin, 1904, Proc. Am. Waterworks Assoc., 24, 341424.
[4] Allievi L., 1903, Teoria generale del moto perturbato dellacqua nei tubi in
pressione. Annali della Societ degli Ingegneri ed Architetti italiani, Milano.
[5] Allievi L., 1913, Teoria del colpo dariete Atti del Collegio degli Ingegneri e degli
Architetti italiani.
[6] Evangelisti G., 1965, Teoria generale del colpo dariete col metodo delle caratter-
istiche LEnergia Elettrica, Vol. XLII, Milano.
[7] Evangelisti G., 1969, Waterhammer Analysis by the Method of Characteristics,
LEnergia Elettrica Vol. XLVI, Milano.
[8] Angus R.W., 1935, Simple Graphical Solution for Pressure Rise in Pipes and Pump
Discharge Lines, J. Eng. Inst. Canada, 7281, February.
[9] Bergeron L., 1935, Etude des variations de rgime dans les conduits deau: Solution
graphique gnrale, Rev. Gen. Hydraul., Paris, 1, 1225.
[10] Streeter V.L., Wylie E.B., 1967, Hydraulic Transients, McGraw Hill, New York.
[11] Streeter V.L., Wylie E.B., 1993, Fluid Transients in Systems, Prentice Hall,
New Jersey.
[12] Chaudry M.H., 1979, Applied Hydraulic Transients, Van Nostrand Reinhold,
New York.
[13] Brunone B., Ferrante M., 2001, Detecting leaks in pressurised pipes by means of
transients. Journal of Hydraulics Research, IAHR, 39(5).
[14] Mpesha W., Chaudhry M.H., Gassman S.L., 2002, Leak detection in pipes by
frequency response method using a step excitation. Journal of Hydraulic Research,
40, 11.1 (5562).
[15] Wang X.W., Lambert M.F., Simpson A.R., Liggett J.A., Vitkovsky J.P., 2002, Leak
Detection in Pipelines using the Damping of Fluid Transients. Journal of Hydraulic
Engineering. ASCE, 128(7), 697711.
[16] Kapelan Z.S., Dragati S.A., Godfrey W.A., 2003, A hybrid inverse transient
model for leakage detection and roughness calibration in pipe network. Journal
of Hydraulic Research, 41(5), 481492.
[17] American Society of Civil Engineers ASCE, 1975, Pipeline design for water and
wastewater, Report of the Task Committee on the Engineering Practice in the Design
of Pipelines, Philadelphia, PA.
182 REFERENCES

[18] American Society of Civil Engineers ASCE, 1992, Pressure Pipe Design for Water
and Wastewater, Committee on Pipeline Planning of the Pipeline Division of ASCE,
New York, NY.
[19] Ellis J., 2008, Pressure transients in water engineering, Thomas Telford Publishing,
London.
[20] Cant M., 2001, Mastering Delphi 6, Sybex Inc., Alameda, CA.
[21] Stephenson D., 1989, Pipeline design for water engineers, Elsevier Science
Amsterdam.
[22] Mays L.W., 2000, Water distribution system handbook, The McGraw-Hill Compa-
nies, New York.
[23] Todini E., 2006, On the convergence properties of the different pipe network
algorithms, 8th Annual Water Distribution Systems Analysis Symposium, Cincinnati,
OH, August 2730.
[24] Mariotte E., 1686, Trait du mouvement des eaux et autres corps fluids, Paris, chz
Estienne Michallet. Edme Mariotte died in 1684, and the book has been edited by
M. de La hire.
[25] DAlembert J.R., 1747, Recherches sur les Cordes Vibrantes, Paris.
[26] Evans L., 1998, Partial differential equations. American Mathematical Society,
Providence, 740 pp.
[27] Bergeron L., 1961, Water hammer in hydraulics and wave surges in electricity
(translated under the sponsorship of the ASME), Wiley, New York.
[28] Parmakian J., 1963, Water-hammer analysis, Dover, New York.
[29] Lister M., 1960, The numerical solution of hyperbolic partial differential equations
by the method of characteristics, in A. Ralston and H.S. Wilf (Eds.) Numerical
methods for digital computers, Wiley, New York.
[30] Abbott M.B., 1966, An introduction to the method of characteristics, Thames and
Hudson, London.
[31] Hirsch C., 2007, Numerical computation of internal and external flows, 2nd Edition,
Elsevier, Oxford, UK.
[32] Godunov S.K., Ryabenkii V.S., 1987, Difference schemes, Elsevier, Amsterdam.
[33] Lax P.D., Wendroff B., 1960, System of conservation laws, Comm. Pure Appl.
Math., 13, 217237.
[34] Thomas J., 1995, Numerical partial different equations: finite difference methods,
Springer-Verlag, New York.
[35] Roe P.L., 1981, Approximate Riemann solvers, parameter vectors and difference
schemes, J. Comput. Phys, 43, 357372.
[36] LeVeque R.J., 2002, Finite volume methods for hyperbolic problems, Cambridge
University Press.
[37] Maione U., 1961, Perdite di carico nella strozzatura di un pozzo piezometrico,
LEnergia Elettrica, 38, 330338.
[38] Thoma D., 1910, Zur Theorie des Wasserschlosses bei Selbsttaetig Geregelten
Turbinenanlagen, Oldenburg, Munchen, Germany.
[39] Jaeger C., 1958 Contribution to the stability theory of systems of surge tanks,
Trans. ASME, 80, 15741584.
[40] Jaeger C., 1960, A review of surge-tank stability criteria, J. Basic Eng., Trans.
ASME, December, 765783.
[41] Pickford J., 1969, Analysis of water surge, Gordon and Breach Publisher,
New York.
REFERENCES 183

[42] Pearsall I.S., 1963, Comparative experiments on surge tank performance, Proc.
National Engineering Laboratory, 177, 951970.
[43] Ellis J., Khairulla L.M., 1974, Oscillations in a surge tank with upper and lower
expansion galleries, Water Power, November, 359364.
[44] Noseda G., sine data, Problemi di moto vario, Ed. Istituto di Idraulica e Costruzioni
Idrauliche, Politecnico di Milano.
[45] Johnson R.D., 1915, The differential surge tank, Trans. Am. Soc. Civ. Eng., 78,
760784.
[46] Gibson W.L., Shelson W., 1956, An experimental and analytical investiga-
tion of a differential surge-tank installation, Trans. A. Soc. Mech. Eng., 78,
925938.
[47] Angus R.W., 1937, Air-chambers and valves in relation to water hammer, Trans.
Am. Soc. Mech. Eng., 59, 661668.
[48] Paoletti A., 1972, Il transitorio negli impianti elevatori muniti di casse daria,
LEnergia Elettrica, 6.
[49] White F.M., 2003, Fluid mechanics, 4th edition, McGraw Hill Series in Mechanical
Engineering, New York.
[50] Evangelisti G., 1935, Sul calcolo delle oscillazioni di carico nelle condotte degli
impianti di sollevamento LElettrotecnica, June.
[51] Evangelisti G., 1938, Il colpo dariete nelle condotte elevatorie munite di camere
daria LEnergia Elettrica, 9.
[52] Cozzo G., 1973, Azione di una piastra piana sullefflusso da un boccaglio, Memorie
e studi dellIstituto di Idraulica e Costruzioni Idrauliche del Politecnico di Milano,
1973.
[53] Tanda M.G., Zampaglione D., 1991, Analisi di fenomeni di risonanza in un sis-
tema distributore al termine di unadduttrice a gravit Dipartimento di Ingegneria
Idraulica, Ambientale e del Rilevamento, Politecnico di Milano.
[54] Bianchi A., Mambretti S., 2000, Le idrovalvole di scarico rapido per lattenuazione
delle sovrapressioni di colpo dariete, LAcqua, 6.
[55] De Martino G., 1973, Sul calcolo del GD2 neli impianti di sollevamento, LEnergia
Elettrica, 8.
[56] Mambretti S., Orsi E., 2007, Un modello di simulazione dei transitori di un sistema
di pompaggio. Proc. of the Third Seminar La ricerca delle perdite e la gestione
delle reti di acquedotto, 185192, Morlacchi editore, Perugia.
[57] Streeter V.L., Wylie E.B., 1985, Fluid mechanics, 8th Edition, McGraw-Hill,
New York.
[58] Mambretti S., 2004, Fenomeni di moto vario nelle correnti in pressione, pp. 128,
Aracne editrice, Roma.
[59] Miller D.S., 1978, Internal flow systems, Volume 5 in the BHRA Fluid Engineering
Series, Cranfield.
[60] Remenieras G., 1952, Dispositif simple pour rduire la clrit des ondes
lastiques dans le conduits en charge, La Houille Blanche, Special A,
172196.
[61] Ghilardi P., Paoletti A., 1986, Additional viscoelastic pipes as pressure surge
suppressors, Paper E2, Proc. 5th Int. Conf. Pressure Surges, BHRA, Hannover,
Germany.
[62] Pezzinga G., 2002, Unsteady Flow in Hydraulic Networks with Polymeric
Additional Pipe, J. Hydraulic Eng., 128(2), 238244.
184 REFERENCES

[63] Meniconi S., Brunone B., Ferrante M., Massari C., 2012, Transient hydrodynamics
of in-line valves in viscoelastic pressurized pipes: long-period analysis, Exp. Fluids,
52, 265275.
[64] Kolkman P.A., 1984, Gate vibrations in P. Novak (Ed.) Developments in hydraulic
engineering 2, Elsevier Applied Science Publishers, Barking, UK.
[65] Tongue B.H., 2002, Principles of vibration, 2nd edition, Oxford University Press.
[66] Inman D.J., 2008, Engineering vibration, 3rd Edition, Pearson Prentice Hall,
New Jersey.
[67] Jaeger C., 1963, The theory of resonance in Hydropower systems. Discussion
of incidents and accidents occurring in pressure systems. Trans. ASME, Ser. D,
85, 631.
[68] Chaudry M.H., 1979, Resonance in pressurized piping systems, J. Hydraulic Div.
ASCE, 96, HY 9, 18191839.
[69] Howell K.B., 2001, Principles of Fourier analysis, CRC Press, USA.
[70] Hedstrom G., 1975, Models of difference schemes for ut + ux = 0 by partial
differential equations, Math. Comp., 29, 969977.
[71] Quarteroni A., 2008, Modellistica numerica per problemi differenziali, Springer-
Verlag, Milano.
[72] Harten A., Lax P.D., 1984, On a class of high resolution total variation stable
finite-difference schemes, SIAM J. Numer. Anal., 21(1), 123.
[73] Godunov. S.K., 1959, A finite difference method for the computation of discon-
tinuous solutions of the equations of fluid dynamics, Matematicheskii Sbornik, 47,
357393.
[74] Jameson A., 1982, Transient aerofoil calculations using the Euler equation,
in P.L. Roe (ed.), Numerical methods in aeronautical fluid dynamics, New York.
[75] Pulliam T.H., 1984, Artificial dissipation models for the Euler equation, AIAA
Paper 85-0438, AIAA 23rd Aerospace Sciences Meeting, Reno, Nevada.
[76] McCormack R. W., Baldwin B.S., 1975, A numerical method for solving the
Navier-Stokes equation with application to shock-boundary layer interaction,
AIAA Paper, 175.
[77] Narayanan K.V., 2006, A Textbook of Chemical Engineering Thermodynamics,
8th Printing, Prentice-Hall, New Delhi.
[78] Van De Sande E., Belde A.P., Hamer B.J.G., Hiemstra W., 1980, Velocity profiles in
Accelerating pipe Flows started from the Rest, Proceeding of the 3rd International
Conference on Pressure Surges, BHRA Fluid Engineering, Canterbury, England, A1,
114.
[79] Bergant A., Ross Simpson A., Vitkovsky J., 2001, Developments in unsteady pipe
friction modelling, J. of Hydraulic Res., 39(3), 249257.
[80] Brunone B., Greco M., 1990, Un modello per la ricostruzione di fenomeni di colpo
dariete anche in presenza di cavitazione. Riscontro sperimentale, Proceedings of
22nd Convegno di Nazionale di Idraulica e Costruzioni Idrauliche, Cosenza, Italy,
4, 114160.
[81] Brunone B., Golia U.M., Greco M., 1991, Some remarks on the momentum equation
for the fast transient, Proceedings of International Meeting on Hydraulic Transient
and Water Column Separation, Valencia, IAHR, E. Cabrera, M.A. Fanelli (eds),
201205.
[82] Brunone B., Golia U.M., Greco M., 1991, Modelling of fast transient by numerical
methods, Proceedings of International Meeting on Hydraulic Transient and Water
Column Separation, Valencia, IAHR, E. Cabrera, M.A. Fanelli (eds), 273282.
REFERENCES 185

[83] Pezzinga G., 2000, Affidabilit di modelli semi-empirici di turbolenza per la val-
utazione delle resistenze di attrito in condizioni di moto vario, LAcqua, 78(1),
2938.
[84] Ramos H., Covas D., Borga A., Loureiro D., 2004, Surge damping anal-
ysis in pipe systems: modelling and experiments, J. Hydraul. Res., 42(4),
413425.
[85] Bughazem M. B., Anderson A., 1996, Problems with simple models for damping in
unsteady flow, Proceedings of the 7th International Conference on Pressure Surges
and Fluid Transients in Pipelines and Open Channels, BHR Group, Harrogate, UK,
537548.
[86] Wylie E. B., 1996, Frictional effects in unsteady turbulent pipe flow, Applied
mechanics in the Americas, M. Rysz, L.A. Godoy, L.E. Suarez (eds), Vol. 5, The
University of Iowa Press, Iowa City, Iowa, 5, 2934.
[87] Pezzinga G., 2000, Evaluation of Unsteady Flow Resistances by Quasi-2D or 1D
Models, J. Hydraul. Eng., ASCE, 126(10), 778785.
[88] Vardy A.E., Brown J.M.B., 1996, On turbulent, un steady, smooth-pipe flow,
International Conference on Pressure Surges and Fluid Transients, BHR Group,
Harrogate, England, 289311.
[89] Carravetta A., Golia U.M., Greco M., 1992, Sullattenuazione spontanea delle flut-
tuazioni di pressione durante i transitori di colpo dariete, Atti del XXIII Convegno
di Idraulica e Costruzioni Idrauliche, Firenze, E67E79.
[90] Golia U.M., 1992, Sulla valutazione delle Forze Resistenti nel Colpo dAriete,
Pubblicazione n .639 del Dipartimento di Idraulica, Universit degli Studi di Napoli
Federico II.
[91] Bratland O., 1986, Frequency-dependent friction and radial kinematic energy vari-
ation in transient pipe flow, Proceedings of the 5th International Conference On
Pressure Surge, BHRA, 95101.
[92] Modica C., Pezzinga G., 1992, Un modello quasi bidimensionale per il moto vario
elastico in regime turbolento, Atti del XXIII Convegno di Idraulica e Costruzioni
Idrauliche, vol. 4, E191E205, Firenze.
[93] Pezzinga G., 1999, Quasi-2D model for Unsteady Flow in Pipe Networks, Journal
of Hydraulic Engineering, 125(7), 676685.
[94] Vardy A.E., Hwang K.L., 1991, A characteristics model of transient friction in
pipes, Journal of Hydraulic Engineering, 29(5), 669684.
[95] Prandtl, L., 1925, Z. angew. Math. Mech., 5(1), 136139.
[96] Nikuradse J., 1932, Laws of turbulent flow in smooth pipes, NASA TT F-10, 359
(Translation of VDI-Forsch).
[97] Nikuradse J., 1933, Laws of flow in rough pipes, NASA TM 1292, 359 (Translation
of VDI-Forsch).
[98] Jones W.P., Launder B.E., 1972, The prediction of laminarization with a two-
equation model of turbulence, International Journal of Heat and Mass Transfer,
15, 301314.
[99] Eichinger P., Lein G., 1992, The influence of friction on un steady pipe flow,
Unsteady flow and fluid transients, R. Bettes, J. Watts (eds), Balkema, Rotterdam,
The Netherlands, 4150.
[100] Pezzinga G., Brunone B., 2006, Turbulence, fiction and Energy dissipation in
transient pipe flow, Vorticity and turbulence effects in fluid structures interac-
tions: an application to hydraulic structure design, M. Brocchini, F. Trivellato (eds),
WIT Press, UK, 213236.
186 REFERENCES

[101] Zhao M., Ghidaui M.S., 2006, Investigation of turbulence behaviour in pipe
transients using a k model, Journal of Hydraulic Research, 44(5), 682692.
[102] Fan S., Lakshminarayana B., Barnett M., 1993, Low-Reynolds number k model
for unsteady turbulent boundary-layer flows, AIAA Journal, 31(10), 17771784.
[103] Lam, C.K.G., Bremhorst, K.A., 1981, Modified form of the k model for predicting
wall turbulence. Journal of Fluids Engineering, 103(3), 456460.
[104] Wylie E.B., 1984, Simulation of Vaporous and Gaseous Cavitation, Journal of
Fluids Engineering, ASME, 106, 307311.
[105] Cannizzaro D., Pezzinga G., 2002, Influenza del rilascio di gas sulle dissipazioni
in transitorio con cavitazione, Proc. XXVIII Convegno di Idraulica e Costruzioni
Idrauliche, Potenza, 1, 619626.
[106] Zielke W., Perko H.D., Keller A., 1990, Gas Release in Transient Pipe Flow,
Proceedings of the 6th International Conference on Pressure Surge, BHRA, C1,
Cambridge, 313.
[107] Wylie E.B., 1992, Low void fraction two-component two-phase flow, in R. Bettes,
J. Watts (eds), Unsteady flow and Fluid Transients, Balkema, Rotterdam, 39.
[108] Scaccia M., Cannizzaro, D., Pezzinga, G., 2004, Calibrazione di modelli di cav-
itazione gassosa per mezzo di micro-algoritmi genetici, Proc. XXIX Convegno di
Idraulica e Costruzioni Idrauliche, Trento, 299305.
[109] Ewing D. J. F., 1980, Allowing for free air in water hammer analysis, Proceedings
of the 3rd International Conference on Pressure Surges, BHRA, Canterbury, UK,
127146.
[110] Pezzinga G., 2003, Second viscosity in transient cavitating pipe flows, Journal of
Hydraulic Research, 41(6), 656665.
[111] Cannizzaro, D., Pezzinga, G., 2005, Energy dissipations in transient gaseous
cavitation, Journal of Hydraulic Engineering, ASCE, 131, 724732.
[112] Ming Z., Ghidaoui M.S., 2003, Efficient quasi-two-dimensional model for water
hammer problems, Journal of Hydraulic Engineering, 10071013.
[113] Stover R.L., 2004, Development of a fourth generation energy recovery device.
A CTOs Notebook, Desalination, 165, 313321.
[114] Stover R.L., Martin J., Nelson M., 2007, The 200,000 m3 /day Hamma seawater
desalination plant largest single-train SWRO capacity in the world and alternative to
pressure center design, Proceedings IDA World Congress, Gran Canaria, Spain.
[115] Mambretti S., Orsi E., Gagliardi S., Stover R., 2009, The behaviour of energy
recovery devices in unsteady flow conditions and application in the modelling of the
Hamma desalination plant, Desalination, 238, 233245.
...for scientists by scientists

Stochastic Methods in Engineering


I. DOLTSINIS, University of Stuttgart, Germany

The increasing industrial demand for reliable quantification and management


of uncertainty in product performance forces engineers to employ probabilistic
models in analysis and design, a fact that has occasioned considerable research
and development activities in the field. Notes on Stochastics eventually
address the topic of computational stochastic mechanics. The single volume
uniquely presents tutorials on essential probabilistics and statistics, recent
finite element methods for stochastic analysis by Taylor series expansion as
well as Monte Carlo simulation techniques. Design improvement and robust
optimisation represent key issues as does reliability assessment. The subject is
developed for solids and structures of elastic and elasto-plastic material, large
displacements and material deformation processes; principles are transferable
to various disciplines. A chapter is devoted to the statistical comparison of
systems exhibiting random scatter. Where appropriate examples illustrate
the theory, problems to solve appear instructive; applications are presented
with relevance to engineering practice.
The book, emanating from a university course, includes research and
development in the field of computational stochastic analysis and
optimization. It is intended for advanced students in engineering and for
professionals who wish to extend their knowledge and skills in computational
mechanics to the domain of stochastics.
Contents: Introduction, Randomness, Structural analysis by Taylor series
expansion, Design optimization, Robustness, Monte Carlo techniques
for system response and design improvement, Reliability, Time variant
phenomena, Material deformation processes, Analysis and comparison of
data sets, Probability distribution of test functions.
ISBN: 978-1-84564-626-4 eISBN: 978-1-84564-627-1
Published 2012 / 378pp / 158.00

WIT Press is a major publisher of engineering research. The company prides itself on producing books by
leading researchers and scientists at the cutting edge of their specialities, thus enabling readers to remain at the
forefront of scientific developments. Our list presently includes monographs, edited volumes, books on disk,
and software in areas such as: Acoustics, Advanced Computing, Architecture and Structures, Biomedicine,
Boundary Elements, Earthquake Engineering, Environmental Engineering, Fluid Mechanics, Fracture
Mechanics, Heat Transfer, Marine and Offshore Engineering and Transport Engineering.
This page intentionally left blank

S-ar putea să vă placă și