Pumps As Turbines. Carravetta Et Al., (2018)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 226

Springer Tracts in Mechanical Engineering

Armando Carravetta
Shahram Derakhshan Houreh
Helena M. Ramos

Pumps as
Turbines
Fundamentals and Applications
Springer Tracts in Mechanical Engineering

Board of editors
Seung-Bok Choi, Inha University, Incheon, South Korea
Haibin Duan, Beijing University of Aeronautics and Astronautics, Beijing
P.R. China
Yili Fu, Harbin Institute of Technology, Harbin, P.R. China
Carlos Guardiola, Universitat Politècnica de València, València, Spain
Jian-Qiao Sun, University of California, Merced, USA
About this Series

Springer Tracts in Mechanical Engineering (STME) publishes the latest develop-


ments in Mechanical Engineering - quickly, informally and with high quality. The
intent is to cover all the main branches of mechanical engineering, both theoretical
and applied, including:
• Engineering Design
• Machinery and Machine Elements
• Mechanical structures and stress analysis
• Automotive Engineering
• Engine Technology
• Aerospace Technology and Astronautics
• Nanotechnology and Microengineering
• Control, Robotics, Mechatronics
• MEMS
• Theoretical and Applied Mechanics
• Dynamical Systems, Control
• Fluids mechanics
• Engineering Thermodynamics, Heat and Mass Transfer
• Manufacturing
• Precision engineering, Instrumentation, Measurement
• Materials Engineering
• Tribology and surface technology
Within the scopes of the series are monographs, professional books or graduate
textbooks, edited volumes as well as outstanding Ph.D. theses and books purposely
devoted to support education in mechanical engineering at graduate and postgraduate
levels.

More information about this series at http://www.springer.com/series/11693


Armando Carravetta Shahram Derakhshan Houreh

Helena M. Ramos

Pumps as Turbines
Fundamentals and Applications

123
Armando Carravetta Helena M. Ramos
Department of Civil, Architectural and Department of Civil, Architecture and
Environmental Engineering Georesources Engineering, CERIS
University of Naples Federico II Instituto Superior Técnico, University of
Napoli Lisbon
Italy Lisbon
Portugal
Shahram Derakhshan Houreh
School of Mechanical Engineering
Iran University of Science & Technology
Tehran
Iran

ISSN 2195-9862 ISSN 2195-9870 (electronic)


Springer Tracts in Mechanical Engineering
ISBN 978-3-319-67506-0 ISBN 978-3-319-67507-7 (eBook)
https://doi.org/10.1007/978-3-319-67507-7
Library of Congress Control Number: 2017953792

© Springer International Publishing AG 2018


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

The sustainable use of water and energy represents one of the major challenges
of the majority of the nations worldwide. In the number of requirements for Water
Supply Networks (WSN), a rational use of energy is becoming a key topic, espe-
cially considering the relationship between water pressure in the pipelines and
leakage: an optimal pressure level in all branches of the network grants, both, the
users’ satisfaction and the water savings. On the contrary, pressure excesses are
connected to an uncontrolled use of energy for water transfer and to the dispersion
of large water volumes.
Pressure management in WSN is not only a primary research topic in civil
engineering, but it is also a daily exigency of the network technical management.
Appropriate strategies have been developed for the optimal location of Pressure
Reducing Valves (PRVs) used for dissipating excess energy. In the meantime, in all
states the maximum effort is present to reduce the energy required in pumping
stations by increasing the pumps performances and by a better integration of the
pumping system in the network.
The environmental benefit coming from the use of renewable sources of energies
led in the last decades to a simplification of the connection of energy sources to the
electrical grid. Nowadays, also very small power plants, presenting peak power of
few kilowatts, are considered a valuable resource. Often the production by
renewables is promoted via national feed in tariffs. The exploitation of small energy
sources is even more convenient when the production is controlled by a energy
user. This is the case of the water transportation and the dissipation points along the
network start to be attractive for the transformation of hydraulic energy in
electricity.
For the small available powers in the PRV location, in the design of the
hydropower plant traditional turbines fail. The miniaturization of the turbine itself
and of the turbine control increases the cost of the electromechanical devices.
Therefore, an alternative was found in the use of pump working in inverse mode:

v
vi Preface

a Pump as Turbine (PAT). The main advantages of this technology are in the low
cost and in the reliability of the electromechanical devices. PAT-based power plants
are receiving an increasing interest in the international literature and the first pilot,
and real plants have been realized. The scope of the book is to give the state of the
art in the design of PAT technology, including, both, mechanical and hydraulic
aspects.

Napoli, Italy Armando Carravetta


Tehran, Iran Shahram Derakhshan Houreh
Lisbon, Portugal Helena M. Ramos
Acknowledgements

For the data on existing hydropower plant based on PAT technology, the authors
express gratitude to:
– Aquatec, Proyectos para el Sector del Agua, SAU
– Aguas Andinas
– Caprari S.p.A.
– Enco Engineering Consultants
– Empresa Municipal de Aguas de Murcia. SA. EMUASA
– Empresa Municipal de Aguas de Granada, SA, EMASAGRA.
The authors wish to thank to the project REDAWN (Reducing Energy Dependency
in Atlantic Area Water Networks) EAPA_198/2016 from INTERREG ATLANTIC
AREA PROGRAMME 2014–2020, CERIS (CEHIDRO-IST), and Federico II
DICEA Laboratory, for the support in the conceptual work and in the experiments on
PATs.
For the work of editing on the book chapters, a particular acknowledgment is for
the Ph.D. student, Mohammad Rezaee.

vii
Contents

Part I Theoretical Aspects


1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1 Micro/Pico Hydroelectric Power Stations . . . . . . . . . . . . . . . . . . . . 4
1.2 Micro/Pico Hydraulic Turbines . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.1 The Pelton/Turgo Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.2 The Cross Flow Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2.3 The Propeller Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.2.4 The Pump as Turbine (PAT) . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Field of PAT Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3.2 Natural Streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.3.3 Water Supply . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.4 Other Users . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2 Reverse Pump Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.1 PAT Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.1.1 The Volute . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.1.2 The Impeller . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.3 The Draft Tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2 PAT Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.3 Flow Pattern in a Pump and in a PAT . . . . . . . . . . . . . . . . . . . . . . 33
2.4 Application of Computational Fluid Dynamics (CFD) to a PAT . . . 36
2.4.1 Importance of CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.4.2 Mathematical Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.4.3 PAT Geometry Modeling and Mesh Generation . . . . . . . . . 38
2.4.4 CFD Solutions for a PAT . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.5 Cavitation in a PAT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.6 Pressure Surge Analysis in a PAT . . . . . . . . . . . . . . . . . . . . . . . . . 48
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

ix
x Contents

3 Industrial Aspects of PAT Design Improvement . . . . . . . . ......... 59


3.1 Impeller Refinement . . . . . . . . . . . . . . . . . . . . . . . . . . . ......... 60
3.1.1 Rounding of the Blade’s Leading Edge and the
Hub/Shroud Inlet Edges of a PAT . . . . . . . . . . . . . . . . . . . 60
3.1.2 Impeller Diameter (Trimming Effect) . . . . . . . . . . . . . . . . . 62
3.2 Inlet/Outlet Part Refinement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.3 PAT Reliability. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.4 PAT Life Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

Part II PAT Based Micro/Pico Hydropower Stations


4 PAT Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.1 Theoretical and Practical Methods . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.2 Estimation of Characteristic Curves . . . . . . . . . . . . . . . . . . . . . . . . 83
4.3 Rotational Speed Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
4.4 Impeller Diameter Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.5 PATs in Parallel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.6 PAT-Pump Direct Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
5 PAT Control Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.1 Operating Conditions in a WSN . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
5.2 PAT Working Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.2.1 The HR Mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.2.2 The ER Mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.2.3 The HER Mode. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.2.4 The SSP Mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.3 Plant Effectiveness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.3.1 System Capability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
5.3.2 System Reliability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
5.3.3 System Flexibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.3.4 System Sustainability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.4 PAT Selection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
6 Civil Engineering Design, Electromechanics, Installation and
Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
6.1 Civil Engineering Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
6.1.1 Intake and Turbine Inlet . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
6.1.2 The Powerhouse . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
6.1.3 Turbine Outlet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.2 Mechanical Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6.2.1 The Shaft . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
6.2.2 The Shaft Seal. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
Contents xi

6.2.3 The Bearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126


6.2.4 The Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.3 Electrical components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.4 Installation Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.5 Operation and Maintenance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

Part III Applications and Impacts


7 Location of a PAT in a Water Transmission and Distribution
System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
7.1 Selecting the Site . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
7.2 Flow Rate and Head Measurement . . . . . . . . . . . . . . . . . . . . . . . . . 142
7.3 Pressure Control in Water Transmission Systems . . . . . . . . . . . . . . 147
7.4 Identification of Dissipation Points . . . . . . . . . . . . . . . . . . . . . . . . . 148
7.5 Dynamic Control of Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
7.6 Numerical Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
7.6.1 Hydraulic Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
7.6.2 Control Valves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
7.6.3 Runaway Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
7.7 Pressure Surge Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
7.7.1 Basic Considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
7.7.2 PAT Shutdowns with Control Valve Closure . . . . . . . . . . . 167
7.7.3 PAT Start-Up and Control Valve Opening . . . . . . . . . . . . . 168
7.7.4 Overspeed Effect in a PAT . . . . . . . . . . . . . . . . . . . . . . . . . 168
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
8 PAT System Economic Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
8.1 Fundaments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
8.2 Time Value of Money . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
8.3 Methods of Economic Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . 178
8.3.1 Net Present Value (NPV) . . . . . . . . . . . . . . . . . . . . . . . . . . 179
8.3.2 Benefit/Cost Ratio (B/C) . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
8.3.3 Internal Rate of Return (IRR) . . . . . . . . . . . . . . . . . . . . . . . 179
8.3.4 Payback Period (T) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
8.3.5 Economic Feasibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
8.4 Effects of the Lifetime Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
xii Contents

9 Application of PAT Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189


9.1 Working Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
9.1.1 Malecòn, Spain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
9.1.2 San Vito di Cadore, Italy . . . . . . . . . . . . . . . . . . . . . . . . . . 194
9.1.3 Capodacqua, Italy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
9.1.4 Conejeras and Cartuja, Spain . . . . . . . . . . . . . . . . . . . . . . . 199
9.1.5 San Antonio, Chile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
9.1.6 Beliche, Portugal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202
9.2 Pilot Plants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
9.2.1 Naples, Italy, Pilot Plant . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
9.2.2 Lisbon, Portugal, Pilot Plant . . . . . . . . . . . . . . . . . . . . . . . . 208
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Acronyms

B/C Benefit–Cost Ratio


BEP Best Efficiency Point
BP Back Pressure
CFD Computational Fluid Dynamics
EE Electricity Energy
ELC Electricity Load Controller
ER Electric Regulation
GHG Greenhouse Gas
GWP Global Warming Potential
HER Hydraulic and Electric Regulation
HPP Hydropower Plant
HPRT Hydraulic Power Recovery Turbine
HR Hydraulic Regulation
IRR Internal Rate of Return
LCA Life Cycle Assessment
LCOE Levelled Cost of Energy
MHP Micro-Hydropower
MHS Micro-Hydro-Solution
MOC Method of Characteristics
MTTF Mean Time to Failure
NPSH Net Positive Suction Head
NPV Net Present Value
PAT Pump as Turbine
PLC Programmable Logical Controller
PRV Pressure Reducing Valve
RAE Relaxed Affinity Model
RNG Re-Normalization Group
SHP Small Hydropower
SSP Single-Serial-Parallel Regulation
SST Shear Stress Transport

xiii
xiv Acronyms

TTF Time to Failure


UPS Uninterruptible Power Supplies
WD Water Distribution
WDS Water Distribution System
WSN Water Supply Network
WSS Water Supply System
WT Water Transmission
WWTPS Waste Water Treatment Plant

Symbols
A Cross-sectional area
c Wave celerity
Cp Water demand peak coefficient
cos / Power factor
D Impeller diameter
d Pipe diameter
e Centrifugal PAT efficiency ratio
E Plant effectiveness
Eh Adsorbed or produced energy
E Mean annual production
f Dimensionless friction factor
F Geometrical size of pump stage
F Probability
Fpx Fmx Pressure and momentum forces
g Acceleration of gravity
h Centrifugal PAT head ratio
H Head drop
Nin Number of inhabitants
I Rotating mass inertia
I Electrical current
K Fluid bulk modulus of elasticity
Kf Leakage coefficient
Kv Valve coefficient
L Mechanical load
n Impeller rotational speed (rps)
N Impeller rotational speed (rpm)
Ns Specific speed
Nin Number of inhabitants
p Water pressure
p Centrifugal PAT power ratio
P Mechanical power
q Centrifugal PAT discharge ratio
Q Discharge
Acronyms xv

QRW Turbine discharge at runaway speed


QR Rated turbine discharge
r Friction factor
R Mechanical reliability
R Mean annual benefit
Re Reynolds number
t Time coordinate
T Mechanical torque
T Time observation period
TE Pipeline elastic time constant
u, v, w Components of local flow velocity
U Electrical potential
V Average cross-sectional velocity
x, y, z Space coordinates
a Shaft internal angle
b Exponent of the leakage equation
η PAT efficiency
ηP Plant capability
h Axial angle
nt Slip factor for turbine operation
k Failure rate
lQ Daily mean water demand
np Slip factor for pump operation
l Dynamic viscosity
lP Plant reliability
p Power number
q Volumetric mass
s Pipeline period
/ Discharge number
u Geometrical size of pump stage
uP Plant flexibility
vP Plant sustainability
w Head number
W Parameter that depends upon the pipe structural constraints
Part I
Theoretical Aspects
Chapter 1
Introduction

Abstract The topic of the book is introduced, based on existing power plants and
machinery, and the main definition of hydropower technology is given. Power
plants are differentiated in terms of plant size, stream characteristics and field of
application, and the design constraints in energy production within pressurized
networks are described. In this chapter, specifically, conventional hydro turbines
applied in micro/pico-hydropower stations are presented. The pump, as an effective
alternative device operating as a turbine, is introduced and compared with more
expensive conventional turbines. Finally, the field of application of the pump as a
turbine (PAT) is presented and discussed. The reader will be introduced to the
particular characteristics of small hydropower plants and will become acquainted
with the terminology used by professionals. In particular, he/she will be made
aware of the specific problem of energy recovery in water distribution networks for
different industrial sectors.

Nowadays water and energy systems are interdependent: the water energy nexus.
Many European and international agencies are addressing the question of opti-
mizing the energy efficiency of water management, treatment, distribution and end
use systems to consider the potential role, contributions and limitations of low
carbon options related to renewable resources. Clean energy is going to be
increasingly important, particularly taking into account recent environmental
limitations relating to problems such as greenhouse gas emissions and global
warming. Off-grid forms of renewable energy, such as solar power, wind power,
hydro power and biomass energy, can play a major role to address these problems.
Unfortunately, renewable energy is not always economically expedient in com-
parison with conventional forms of energy. Therefore, designing low-cost machines
with a higher efficiency is a topic of great importance for researchers and engineers.
Large hydroelectric power plants, intensively realized in Europe and North
America in the first half of the last century, are still now under design and con-
struction in many regions of Asia and Africa. The main advantage of such plants is
that they are characterized by the low cost of the energy produced (cost per-kW).
Moreover, their high flexibility in the fast and accurate regulation of power has

© Springer International Publishing AG 2018 3


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_1
4 1 Introduction

resulted in their becoming the main means of facing the peaks in energy demand. In
addition, the high capacity energy saving by pumped storage power plants makes
them an attractive method for adoption in many countries. The first use of a reverse
pump in hydropower is connected with the design of pumped storage power plants:
the same machine, which is used for pumping during the night, is also used as a
turbine in the peak energy consumption period.
Hydro-energy became even more attractive after the oil price crisis of the 1970s
and has become so again in recent years [1]. Unfortunately, most natural sources of
falling water have already been exploited in industrialized countries and environ-
mental concerns act as a limitation on the design of large storage capacity in
mountain areas. Nevertheless, the expectations of exploiting hydropower have
recently been growing again, due to the possibility of producing energy from small
water streams, as in the case of the environmental flows released from large dams,
or at the outlet of waste water treatment plants. However, the cost per-kW of the
energy produced by these micro hydropower stations can be higher than that of the
large hydroelectric power stations [1, 2].
Water supply systems are another potential application for hydro-energy. The
water industry is the fourth most energy intensive sector in Europe, contributing
heavily to climate change emissions. The distribution of water in pipe networks
consumes 45% of this energy [3]. As highlighted by EIP1 Water, the potential
energy savings available in transporting water through a network infrastructure are
very significant. A considerable proportion of these savings could come from the
installation along water supply systems of micro and pico hydropower stations,
either replacing existing pressure reducing valves or being placed in parallel or in
series with them, depending on regulation requirements. Due to the variable
operating conditions, in terms of flow rate and pressure drop, the cost per-kW of the
energy produced is a crucial aspect for the exploitation of this technology [4].

1.1 Micro/Pico Hydroelectric Power Stations

Recently, several publications have emphasized the importance of using simple


turbines in order to reduce the cost of the electrical energy produced. Accordingly,
there is the need for the installation of small hydroelectric power stations in many
developing countries.
Several methods have been proposed by different organizations for the classi-
fication of hydropower stations according to their power, head and flow rate. In
Table 1.1, a classification formulated by the Latin American Energy Organization
(OLADE) is presented [5]. In some countries, micro, mini and small hydropower
plants are generally referred to by the common name “small hydro power plants”.

1
European Innovation Partnerships.
1.1 Micro/Pico Hydroelectric Power Stations 5

Table 1.1 Small hydropower classification [5]


Type Capacity (kW)
Small hydropower station <5000
Mini hydropower station <500
Micro hydropower station <50
Pico hydropower station <5

Fig. 1.1 A
high-head/low-discharge
micro/pico hydropower
station

Micro/pico hydroelectric power stations are usually installed in remote areas as


off-grid or stand-alone power stations. The arrangement of a micro/pico hydro-
electric power station depends on the hydraulic characteristics of the site. Typically,
there are two site arrangements: high-head/low-discharge (Fig. 1.1) and
low-head/high-discharge (Fig. 1.2). Generally, a typical micro/pico hydroelectric
power unit consists of three main components: the turbine, generator and control
system. The turbine is the heart of the micro/pico hydroelectric power system,
converting the hydraulic energy of water to mechanical energy. The generator takes
the mechanical energy from the turbine by means of a connection with the shaft and
converts it into electricity. To control the frequency and voltage of the electricity
consumption, a control system should monitor the system, which can be of a
hydraulic, pneumatic or electric type. In addition, either a governor, as for large
hydroelectric power plants, can be used to control the inlet flow rate to regulate the
output of the generator, or an electronic load controller (IEC) can be applied to
control the consumption of electricity. The details of the equipment of a typical
micro/pico hydroelectric system are described in Chap. 5.
6 1 Introduction

Fig. 1.2 A
low-head/high-discharge
micro/pico hydropower
station (www.inforse.org)

1.2 Micro/Pico Hydraulic Turbines

There are many turbine types available to convert hydraulic energy to mechanical
energy. However, only a few can be applied in micro/pico hydroelectric power
stations. Pelton/Turgo, cross flow and propeller turbines and pumps as turbines
(PAT) are the types that have been widely used to produce energy in micro/pico
hydroelectric power units.

1.2.1 The Pelton/Turgo Turbine

The Pelton turbine is an impulse turbine, invented by Lester Allan Pelton (1829–
1908), and is perfect for high head hydro plants. The high pressure flow enters the
runner by means of a single/multi jet nozzle system. The turbine discharge can be
regulated by a needle valve to adjust the output power. The runner is of a cylindrical
shape consisting of several buckets, which have a special design to capture the
energy from the jets of water. The jets strike the centers of the buckets and exit from
the sides. The design and layout of a series of buckets is realized in order to avoid
the possibility that the flow into one bucket influences the others. The recom-
mended number of buckets is between 18 and 24 [6].
1.2 Micro/Pico Hydraulic Turbines 7

Fig. 1.3 The Pelton turbine


(www.ricklyhydro.com)

Fig. 1.4 The Turgo turbine


(www.renewablesfirst.co.uk)

A deflector is installed between the runner and the nozzle deviating the water jets
to prevent them from entering the runner in emergency situations, such as the
increase in the rotational speed in no-load conditions (Fig. 1.3 shows a pico Pelton
turbine).
The Turgo turbine is very similar to the Pelton turbine. However, the jets of
water enter the buckets at an angle of approximately 20°. Therefore, the Turgo
turbine can be applied in power units with a lower head and higher flow capacity.
Figure 1.4 shows a Turgo turbine and Fig. 1.5 compares the inlet flow in a Turgo
and a Pelton turbine.
8 1 Introduction

Fig. 1.5 Comparing the inlet


flow between a Turgo and a
Pelton turbine (www.tanasui.
co.jp)
Turgo

Pelton

1.2.2 The Cross Flow Turbine

The cross flow turbine is a hydraulic turbine developed by an Australian, Anthony


Michell, a Hungarian, Donát Bánki and a German, Fritz Ossberger. This turbine,
one of the impulse turbine types, has been widely used in small hydropower sta-
tions. Heads ranging from 2 to 400 m and discharges from 20 to 2000 l/s can be
covered by cross flow turbines. This turbine has a flat curve of efficiency covering
various flow rates. In fact, by acting on a simple governor, it is possible to send a
low flow rate to a smaller part of the runner to obtain a better efficiency (Fig. 1.6).
Water passes from the inlet channel and the guide vane and enters the runner.

Fig. 1.6 The cross flow turbine (www.ossberger.de)


1.2 Micro/Pico Hydraulic Turbines 9

Crossing the water from the runner twice (for which reason it is called “cross
flow”), it exits into the atmosphere by the outlet channel. Moreover, water can enter
the runner horizontally or vertically. The recommended number of blades is
between 10 and 34.

1.2.3 The Propeller Turbine

The Propeller turbine is of a reaction turbine type and is useful for low heads and
high flow rates. The turbine is installed in a pipe and the generator can be installed
inside or outside. The draft tube can be conic or cylindrical (Fig. 1.7). In the
absence of a spiral case the inlet can even be provided with a set of fixed/movable
guide vanes. The manufacture of this turbine is easier than the Kaplan or Francis
turbines.

1.2.4 The Pump as Turbine (PAT)

In recent years, small and micro hydro power plants have become a possible new
application area for pumps as turbines (PAT), which are aimed at replacing the
more expensive conventional turbines. One way of overcoming the high-cost
capital price of micro/pico hydropower stations is to use simple machines instead of
more complex conventional turbines. Industrial pumps can be effectively operated

Fig. 1.7 The propeller turbine (www.hydroturbine.cn)


10 1 Introduction

as turbines. Pumps are relatively simple machines, easy to maintain and readily
available in most developing countries. From the economic point of view, it is often
stated that pumps working as turbines, in the range of 1–500 kW, allow capital
payback periods of two years or less, a considerably shorter time than that of a
conventional turbine [5].
Several researchers have become interested in pumps operating as turbines for
many types of small applications. Pump manufacturers do not normally provide
turbine mode performance for their products. Therefore, in recent years, many
prediction techniques have been developed e.g. Sharma [7], Williams [8],
Alatorre-Frenk [9], and more recently Ramos and Borga [10, 11], Valadas and
Ramos [12], Singh [13, 14], Derakhshan and Nourbakhsh [1] and Yang et al. [15].
Due to the huge number of pumps on the market of all possible sizes, appropriate
machines, cheap and reliable, are readily available. Regarding maintenance, there
are many advantages compared with custom-made turbines.
As a result of the fixed geometric conditions within the casing and impeller,
pumps as turbines have a low part-load performance. This characteristic addresses
one of the most important challenges for a micro hydro system based on a PAT.
A pump operating as a turbine is very sensitive towards changing boundary
parameters, namely head and discharge. Hence a wrong pump selection, or rota-
tional speed, will result in a shift of the operating point (Fig. 1.8), delivering a
non-desired output, and ultimately even causing a possible failure of the project.
In Fig. 1.9, the main parameters under analysis have been converted into
non-dimensional parameters, such as discharge number (u = Q/ND3 with Q in
m3/s, N in rps, and D in m) and head number (w = gH/N2D2 with H in m). In this

Fig. 1.8 Typical performance curves for a PAT for various rotational speeds
1.2 Micro/Pico Hydraulic Turbines 11

Fig. 1.9 Performance curves of different PATs [16]

figure there is a large representation of head number (w) and discharge number (u)
curves, ranging from lower to higher specific speed values (Ns = NQ0.5/H0.75). The
lower specific speed w–u curves also have steeper slopes when compared to the
upper specific speed values. Each PAT has different values of maximum operating
efficiency (see BEP trend line), which are also related to the size of each machine
[16].
Centrifugal pumps can be easily rotated as turbines in heads ranging from 15 to
100 m and with flow rates ranging from 5 to 50 l/s (Fig. 1.10). Figure 1.11 shows
the application of different pumps working as turbines. Cross flow turbines can be
replaced by centrifugal, mixed flow and double suction PATs. The ranges for mixed
flow PATs (Fig. 1.12) are 5–15 m and 50–150 l/s for head and flow rates,
respectively. For heads ranging from 15 to 100 m and flow rates ranging from 50 to
1000 l/s, double suction pumps or multi centrifugal pumps in a parallel arrange-
ment, are suitable. Pump manufacturers usually produce centrifugal and mixed flow
pumps on a large scale, so these pumps are readily available on the market at a low
price [5].
In particular, it may be possible to use multistage pumps replacing Pelton/Turgo
turbines [5]. The best example is the HPRT (Hydraulic Power Recovery Turbine)
for recovery from outflow from a higher pressure to a lower pressure in such a way
that throttling occurs.
For heads ranging between 1 and 5 m and flow rates between 50 and 1000 l/s,
axial pumps can be used as propeller turbines (Fig. 1.12). However, these types of
pumps cannot usually be found on the market and therefore, since pump manu-
facturers supply them in a customized way, their price may not be so reasonable.
12 1 Introduction

Fig. 1.10 A centrifugal PAT based pico hydropower station, installed in the west of Iran

Fig. 1.11 Application of various pumps as turbines [5]


1.3 Field of PAT Application 13

Fig. 1.12 An axial PAT based micro hydropower station, installed in the north of Iran

1.3 Field of PAT Application

1.3.1 Introduction

The use of the most refined methods in the design of hydro power plants is
necessary to maximize plant efficiency and energy production. For many applica-
tions a difference of a few percentage points in PAT efficiency is crucial for the
economic sustainability of the project. The limited installed power of micro and
pico installations reduces annual incomes. Therefore, the difference between the net
present value (NPV) of cash inflows and the present value of cash outflows is
generally small.
The choice of the pump is strictly connected with the power plant operating
conditions, which are differentiated on the basis of the nature of the water resource,
of the flow conditions, and of the use of the water. In natural streams, depending on
the size of the body of water, the flow rate and water level variability can be
attributed to seasonal spring outflows, levels of rainfall and water uptakes. In large
hydro power plants, specific waterworks are created to stabilize the turbine flow and
head, as in the case of dams. In artificial streams, as in channels and in pressure pipe
14 1 Introduction

Fig. 1.13 Daily pattern of water demand for different end users [17]

transmission, the operating conditions can be more variable. In general, when the
power plant is installed close to the network intake, the working conditions are
appreciably stable. On the contrary, the operating conditions are extremely variable
when the micro and pico hydro power plants are installed close to the end user. In
Fig. 1.13, the daily trend of water demand for different end users is plotted,
showing the influence of the water use on the pattern of water demand [17]. In
pressurized water distribution systems, friction losses are strongly dependent on the
flow rate and a high pressure drop variability is also observed.
The identification of potential energy production points, the choice of the main
design parameters and, finally, the selection of the best design solution are crucial
aspects in micro hydro power (MHP) exploitation.
These elements will be described separately for natural streams, including all
solutions for free surface flow, for water supply, considering only pressurized water
flow conditions, and for all other uses, considering pressure and channel flow
conditions in irrigation and industrial environments.

1.3.2 Natural Streams

Using natural water sources to produce energy is not an innovative method.


Hundreds of years ago, our ancestors occasionally converted this natural hydraulic
potential to mechanical power by the use of simple water wheels. Water mills, for
instance, were widely used for grinding grain into flour in ancient civilizations (e.g.,
in Iran, Greece, Egypt and China). With the invention of generators, these natural
sources were easily converted into electricity energy. Today, these sources can be
found in the landscapes of every country. Apart from natural rivers and waterfalls,
stream flows are the main means of supplying power to the agriculture and aqua-
culture industries. Therefore, it is advantageous for farmers to produce their own
energy from these natural streams by micro/pico hydropower stations, a process
known as “Distributed Generation (DG)”.
1.3 Field of PAT Application 15

Fig. 1.14 PAT-based pico hydropower stations in Iran

Natural streams are generally available for heads from 1 to 100 m and for flow
rates from 1 to 1000 l/s up to 100 kW. However, the absence of exact hydrology
data leads to uncertainty in the selection of DG equipment. Due to geographical
distribution and economy issues, gathering hydraulic data as a package, as for large
hydropower plants, is generally impossible.
The most important factor is the accuracy of flow rate and head measurement
data in order to select the right machine. Some simple methods are available to
measure hydraulic data and these will be discussed in detail in further chapters.
Conventional turbines, such as Francis turbines, are extremely expensive in terms
of their application in these natural sources to produce energy from small water
flows. On the contrary, PATs have represented an economical and easy-to-use
alternative means that could be made available even in remote villages and rural
areas. For this reason, PATs have been widely used as turbines to produce off-grid
electricity from natural streams and to meet the basic requirements of the popula-
tion. Figure 1.14 shows two PAT-based pico hydropower stations in natural
streams.
The power generated by natural streams can be approximately estimated by:

P ¼ 9:81 QHg  5QH ð1:1Þ

where H (m), Q (l/s) η are head, flow rate and total efficiency (consisting of PAT
and generator), respectively.
Figure 1.15 shows a schematic view of a micro hydropower station established
on a natural stream site. The water controlled by an upper reservoir is released to a
penstock to enter the PAT by an inlet valve. Next, the water flows out to a tail water
reservoir by a suction pipe to return to its main path (i.e., the river). The mechanical
energy of the PAT, produced by the hydraulic potential, can be easily converted to
16 1 Introduction

Fig. 1.15 A schematic view of a pico/micro hydropower station established on a natural stream
site (www.sigma.cz)

electricity by an alternator. Finally, to ensure clean electricity, an ELC (Electronic


Load Controller) system is definitely recommended to monitor the system
automatically.

1.3.3 Water Supply

A typical scheme of water supply for an urban network is shown in Fig. 1.4. From
the water intake, represented by a spring or well withdrawal, long transmission
pipelines are used to convey the water close to the urban centers. Water tanks are
used to store the water and/or to stabilize the pressure head. Urban water networks
are used to supply the water to the end users.
The supply network is therefore split into two separate systems: water trans-
mission (WT) and water distribution (WD). Generally, the first is an open network
and the second a closed loop network. Flow variability in WT is limited and it is
determined only by seasonal variations, as an effect of differences in the budget
available and water demand. In WD the variability is large and it is determined by
the instant variations of user demand.
A typical daily pattern of water demand flow and pressure head (hm) in WD is
plotted in Fig. 1.17. The water demand is made dimensionless by means of the
ratio:
1.3 Field of PAT Application 17

Q ðt Þ
C p ðt Þ ¼ ð1:2Þ
lQ

where lQ is the daily mean water demand, and Q(t) is the flow demand during the
day. The span of the flow variability is determined by several factors, such as the
number of inhabitants, the presence of industrial and commercial use, and the
amount of water leakage. The maximum flow rate is connected to the aggregate
size.
Setting at Nin the number of inhabitants, the maximum value of Cp can be
expressed by the following formulation [18]:

17:12
Cpmax ¼ pffiffiffiffiffiffiffi þ 1:185 ð1:3Þ
Nin

The lower limit of the water flow span can be determined only experimentally,
with observed minimum night values of Cp, ranging between 0.2, for new or
maintained networks, and 0.5 in the presence of old networks presenting huge
leakage. High night values of flow are also observed in the presence of a large
spectrum of user types: residential, industrial and commercial.
Excess energy can be presented along any pipe branch of a water supply system.
In Fig. 1.16 a typical situation is shown, including a WT system and a very sim-
plified water supply system. In WT, the excess of energy is generally connected to
the difference in ground elevation between the spring and the final tank, larger than
the difference in the grade line. In WD, a second reason for the energy excess could
be connected to the necessity of granting optimal pressure values to the end users.
Excess energy was in the past dissipated by manual or automatic valves, the
former being used only in the presence of stable operating conditions (Fig. 1.18).
Along the main pipelines of water transmission systems, the opportunity of

Fig. 1.16 Water supply system for residential use [19]


18 1 Introduction

converting hydraulic energy has already been exploited by mini hydro power
plants, like the one represented in Fig. 1.19.
In WD, due to the marked variability of the flow rate and the pressure head, a
dynamic valve control is necessary to grant optimal pressure value to the network.
The pressure head in the internal nodes of the network has to be contained within a
strict range of values, the lower limit being represented by the minimum pressure
required by the end users, and the higher limit being fixed to grant a proper network
functionality due to safety reasons. In general, pressure control reduces leakage and
pipeline stress. This situation is shown in Fig. 1.20: a control valve is used to
dissipate the excess energy Dh. An automatic pressure reducing valve (PRV) can
adjust dynamically the entity of dissipation, in order to grant optimal pressure
values in the most critical node of the network (Fig. 1.21).
Micro and pico hydro power plants can replace the PRV granting the same
dissipation but with energy production. By the use of appropriate technologies,
excess energy (Dh) will be partially dissipated in the control valve and partially
used as the PAT head drop (d).

Fig. 1.17 Daily pattern of water demand flow and pressure head [4]

Fig. 1.18 Excess energy in a water supply system


1.3 Field of PAT Application 19

Fig. 1.19 A mini hydro power plant in a WT system

Fig. 1.20 Geometric relationship between head drops and geodesic elevations [4]

1.3.4 Other Users

WT is also common to different industrial sectors: in irrigation, waste water


treatment and water drainage. Different potentialities for energy production emerge
from each of these applications, even if the cost of the technology is, in some cases,
still prohibitive.
20 1 Introduction

Fig. 1.21 Localization of maximum and minimum pressure and PAT nodes in a real network [4]

Water transfer in irrigation is performed as in Fig. 1.22. From the intake,


pressure pipes or artificial channels supply water to a number of disconnecting
tanks. Next, the water is distributed to farmers by open loop networks. The main
difference, compared to residential supply, is that the use of irrigation networks is
limited to the arid season. This seasonal use of water represents an obstacle to the
introduction of hydro power plant in WD due to the lower economic benefits in
energy production.
In many cases this limitation is not present in WT. In fact, during the wet season,
water could be available at the intake and a power plant installed along the WT
route could work continuously by discharging water directly into the drainage
network.
As examples of hydro power application in WT, two situations are illustrated,
taken from the WT of the irrigation water utility of Consorzio di Bonifica della

Fig. 1.22 A water supply


system for irrigation
1.3 Field of PAT Application 21

Control room
Room gates

Squanderer

Fig. 1.23 Inlet of the water transmission system of Consorzio di Bonifica della Capitanata (Italy)
at the Occhito dam

Tavoliere tank

Tower P4 Celone tank

Tower P1 Tower P2

Tower P3

Region 6/A Region 6/B

Region 12

The tank S3
The tank S1 The tank S2

Fig. 1.24 Scheme of the water transmission system of Consorzio di Bonifica della Capitanata
(Italy)

Capitanata in southern Italy. The first case refers to the network intake at the feet of
the Occhito dam. The water head, determined by the static level in the reservoir, is
reduced in the dissipation chamber before the free surface channel of the WT.
A project has been funded for a 1.0 MW power plant. A bypass of the existing gate
will be realized, and energy will be produced by a turbine in the gate chamber.
Water will then be discharged in the dissipation chamber from the top (Fig. 1.23).
The second scheme refers to the dissipation node located at the outlet of WT
(Fig. 1.24). A constant pressure head for WD is granted by the presence of three
large elevated tanks (see Fig. 1.25). Four large pressure reducing valves (400–
1400 mm) are located at the feet of the tanks to dissipate the excess energy
(Fig. 1.26). The substitution of the PRV is planned by means of a micro hydro
power plant of approximately 600 kW in each node.
22 1 Introduction

Fig. 1.25 Elevated tank

The mining industry also represents a potential application of PATs. The process
of underground mining frequently requires large volumes of water and consumes a
considerable amount of energy. In these cases, an extensive network of water
supply pipes is present, which distribute water for the mining process across large
distances and deep underground. The drop in elevation from ground level to the
lowest point in the mine could be of several hundred meters, and the extent of the
mine could be of tens of kilometers in multiple underground directions. Due to the
large drops in elevation, PRVs are required at regular intervals, to maintain the
pressure at a optimal pressure value. Feasibility studies have demonstrated that each
PRV could have the capacity to recover about 10 kW of electricity.
The food industry is another potential application of PATs, as in the case of an
alcohol brewery or paper manufacturing plant. For example, the brewing process
consumes 3.5 l of water for every 1.0 l of beer produced. Water for production is
generally obtained from the local public water distribution network (WDN) and
also distributed within the processing plant within an internal WDN. Small head
drops are generally present along the WDN and are favorable for energy production
1.3 Field of PAT Application 23

Fig. 1.26 Pressure reducing valves at the inlet of the first elevated tank (P1)

due to the fact that they are usually associated with large flow rates. The same case
is observed at the outlet of a waste water treatment plant. It may happen that a
geodesic drop is present between the level of the treatment plant and the body of
water, a small stream or a lake. In these cases, thanks to the large flow rate, the
installation of a micro hydro power plant is feasible. The use of PATs can be
recommended to reduce the impact of the waterworks in any area of great envi-
ronmental value.
HPRTs are usually used instead of throttling valves to recover water power in
thermal power plants, and in oil and gas sector refineries. Reverse running pumps
(single and multi-stage centrifugal and axial types) can convert the return pipe line
differential pressure to mechanical energy helping to drive the centrifugal single or
double suction pump to pressurize the water of the main process pipe line. Since, in
the recycling system, the recovered energy is smaller than the required energy to
drive the main pump, an electromotor, a steam turbine or a diesel generator, on the
other side of the main pump shaft, is used to supply the extra energy required.
Figure 1.27 shows a typical multistage pump used as an HPRT.
Last, but not least importantly, it should be noted that the main challenge of
renewable energy is related to the sources. Since these sources (e.g. solar, wind and
tidal) are not permanent, a high-capacity energy storage system is required to
manage the energy production and consumption. Hence, pumped-storage is a
perfect solution for high-capacity energy saving, together with a battery. To reduce
the net cost in a pumped-storage system, a pump-as-turbine (PAT) can be used to
save energy (Fig. 1.28) [20].
24 1 Introduction

Fig. 1.27 A typical multistage pump used as an HPRT (www.tefkuwait.com)

Fig. 1.28 A PAT saving strategy in a hybrid solar-wind system with a battery [20]
References 25

References

1. S. Derakhshan, A. Nourbakhsh, Experimental study of characteristic curves of centrifugal


pumps working as turbines in different specific speeds. Exp. Therm. Fluid Sci. 32(3), 800–807
(2008)
2. H. Ramos (Editor), Guidelines for design of small hydropower plants. WREAN (Western
Regional Energy Agency and Network) and DED (Department of Economic Development—
Energy Division), Belfast, North Ireland, ISBN 972-96346-4-5 (2000)
3. A. McNabola, P. Coughlan, A.P. Williams, Energy recovery in the water industry: an
assessment of the potential of micro-hydropower. Water Environ. J. 28(2), 294–304 (2014)
4. A. Carravetta, O. Fecarotta, M. Sinagra, T. Tucciarelli, Cost-benefit analysis for hydropower
production in water distribution networks by a pump as turbine. J. Water Resour. Plan.
Manage 140(6), 4014002 (2014)
5. A. Nourbakhsh, S. Derakhshan, E. Javidpour, A. Riasi, Centrifugal & axial pumps used as
turbines in small hydropower stations. Hidroenergia 2010 International Congress on Small
Hydropower International Conference and Exhibition on Small Hydropower, pp. 16–19
(2010)
6. R. Hothersall, G. Anestis, C. Gürkök, Hydrodynamic Design Guide for Small Francis and
Propeller Turbines (UNIDO, 2004)
7. K.R. Sharma, Small Hydroelectric Project-Use of Centrifugal Pumps as Turbines. Technical
Report, India (Kirloskar Electr. Co, Bangalore, 1985)
8. A.A. Williams, Pumps as turbines used with induction generators for stand-alone
micro-hydroelectric power plants, Doctoral dissertation, Nottingham Trent University
(1992) URL: http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.262127
9. C. Alatorre-Frenk, Cost minimisation in micro-hydro systems using pumps-as-turbines,
Doctoral dissertation, University of Warwick (1994) URL: http://wrap.warwick.ac.uk/36099/
1/WRAP_THESIS_Alatorre-Frenk_1994.pdf
10. H. Ramos, A. Borga, Pumps as turbines: an unconventional solution to energy production.
Urban Water 1(3), 261–263 (1999)
11. H. Ramos, A. Borga, Pumps yielding power. Dam Eng. 10(4), 197–217 (2000)
12. M. Valadas, H. Ramos, Use of pumps as turbines to profit the available energy in irrigation
system. Water Resour. J. APRH. 24(3), 63–76 (2003)
13. P. Singh, Establishment of a test-rig for turbine for micro hydro and detailed testing of a pump
as turbine, Centre for Energy Studies, Indian Institute of Technology, New Delhi, India
(2000)
14. P. Singh, Optimization of the internal hydraulic and of system design in pumps as turbines
with field implementation and evaluation, Doctoral dissertation, University of Fridericiana,
Karlsruhe, Germany (2005)
15. S.-S. Yang, S. Derakhshan, F.-Y. Kong, Theoretical, numerical and experimental prediction
of pump as turbine performance. Renew. Energy 48, 507–513 (2012)
16. H. Ramos, A. Borga, M. Simão, New design solutions for low-power energy production in
water pipe systems. Water Sci. Eng. 2(4), 69–84 (2009)
17. A. Funk, W. DeOreo, Embedded energy in water studies, study 3: end-use water demand
profiles. Prepared by Aquacraft, Inc. for the California Public Utilities Commission Energy
Division, Managed by California Institute for Energy and Environment, CALMAC Study ID
CPU0052 (2011)
26 1 Introduction

18. J. Martínez-Solano, P.L. Iglesias-Rey, R. Pérez-García, P.A. López-Jiménez, Hydraulic


analysis of peak demand in looped water distribution networks. J. Water Resour. Plan.
Manage 134(6), 504–510 (2008)
19. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Energy production in water
distribution networks: a PAT design strategy. Water Resour. Manage 26(13), 3947–3959
(2012)
20. S. Sadeghpour, “Design and Analysis of the performance of a hybrid system of renewable
(hydro, wind, solar)”, MSc Dissertation, Iran University of Science and Technology, (Sep,
2015)
Chapter 2
Reverse Pump Theory

Abstract Users, researchers and engineers require a fundamental knowledge of


reverse pump theory. In this chapter first, a complete description of all PAT
components is given. Then, the theory of small reverse pumps is described, based
on the results of theoretical analyses, experiments and numerical simulations. PAT
characteristic and efficiency curves are introduced and, the influence of pump size
and rotational speed on such curves is considered. The possibility of using
numerical simulations instead of experimental results is discussed. Finally, addi-
tional aspects like cavitation and the incidence of pressure surges are presented and
different tools, commonly used for PAT sizing and hydropower plant design, are
described. The reader will be prepared, by means of the theoretical basis provided,
for the comprehension of the hydraulic and mechanical processes connected with a
PAT operation.

Around 1930, Thoma started his research on PATs by chance during tests to obtain
the characteristic curve of a pump. However, the first complete characteristic curves
of a pump working in turbine mode were presented by Knap in 1937 [1]. The
hydraulic behavior of a pump, when it rotates as a turbine, has changed significantly.
Although, the rotation speed and flow are reverse in turbine mode, the behavior of a
PAT is completely different compared with its pump mode. In general, a pump
works in turbine mode with higher heads and discharges for the same rotational
speed. In the past, the electromechanical characterization of a pump working in
turbine mode was limited to pump and storage hydro power plants, with the installed
power as large as hundreds of megawatts, as in the case of the Presenzano power
plant in Italy, which is equipped with two pump/turbines with 250 MW each. In
recent years the interest in the use of small pumps in mini and micro hydro power
plants has increased. The behavior of these small PATs is extremely different when
compared with the huge pumps of the past, because industrial PATs are generally
used without any optimization of the geometry for reverse mode application.

© Springer International Publishing AG 2018 27


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_2
28 2 Reverse Pump Theory

2.1 PAT Components

Depending on the PAT type, the components of each machine are different.
However, in general, PAT components are classified in two groups: rotary and
stationary. The rotary group consists of the impeller and shaft and the stationary
group includes the inlet parts (i.e., the inlet pipe and volute), outlet parts (i.e., the
suction pipe and draft tube), and the casing and mechanical parts (i.e., rings,
mechanical seals and bearings). However, the main hydraulic parts of a PAT are the
impeller, volute and draft tube (Fig. 2.1).

2.1.1 The Volute

Although the volute of a PAT is designed to collect the water from the impeller of a
pump, the design method for a pump is quite similar to a Francis turbine volute (i.e.,

A B C D E

Seal Cover Shaft Shaft Sleeve Channel

F G H J I K

Volute Outlet Impeller Eye Impeller Wear ring Inlet

Fig. 2.1 PAT components (www.processindustryforum.com)


2.1 PAT Components 29

Fig. 2.2 Plan view of a PAT


[18]

a constant velocity method). Therefore, when the flow is reverse, the volute of a
PAT will work properly as a distributor of the flow surrounding the impeller.
However, the most important role of a volute is the balancing of radial pressure and
forces. Figure 2.2 shows a typical volute of a PAT.

2.1.2 The Impeller

The impeller, the main component of a PAT, converts the hydraulic energy of the
inlet flow to the mechanical shaft. Although the impeller has been designed to
pressurize the liquid using the centrifugal force, it can also absorb the flow energy
with a reasonable efficiency. The impeller usually has 5–8 blades, with a hydrofoil
shape type. An impeller and the details of a blade are presented in Fig. 2.3.

Fig. 2.3 Impeller and details of a blade [19]


30 2 Reverse Pump Theory

Fig. 2.4 Draft tube of a PAT


[20]

2.1.3 The Draft Tube

A small kinetic energy is required at the PAT outflow to minimize energy losses.
Using a draft tube, the kinetic energy can be converted into potential energy.
Therefore, a draft tube is designed based on the maximum conversion of velocity to
pressure (according to the Bernoulli equation). Figure 2.4 shows a typical draft tube
of a PAT.

2.2 PAT Theory

When a pump runs as a turbine, the high pressure fluid enters the machine and then
leaves after the conversion of the energy has occurred within the impeller. Since the
flow direction and the rotational speed are reverse, the velocity triangles of both
modes change.
The pump and outlet velocity triangles in direct and reverse modes are shown in
Fig. 2.5. Although only the shaft rotation and flow direction are opposite in pump
and turbine modes, their velocity triangles are different. Considerations show that
the inlet flow angle to the impeller in a PAT is not exactly the same as the outlet
flow angle of the impeller in the pump mode. This angle (av) is approximately equal
to the volute angle. In fact, the volute operates as a guide channel in turbine mode.
Assuming the outlet flow angle from the impeller in reverse operation is equal to the
impeller inlet flow angle in direct operation (i.e., with no whirl at the impeller exit),
the same Euler heads for both modes can be considered [2]:

HpE ¼ HtE ð2:1Þ


2.2 PAT Theory 31

Fig. 2.5 Velocity triangles: a pump impeller outlet velocity triangle b turbine impeller inlet and
outlet velocity triangles [2]
32 2 Reverse Pump Theory

Due to the slip factor related to the finite blade number, the pump and turbine
theoretical heads can be written as:

Hp00 ¼ np HpE ð2:2Þ

HtE
Ht00 ¼ ð2:3Þ
nt
 
where np is the slip factor for the pump operation np \1 and nt is the slip factor
for the turbine operation, that is approximately equal to 1.0 [2].
Considering the hydraulic efficiency for each mode, Eqs. (2.2) and (2.3) can be
rewritten as:

Hp ¼ ghp Hp00 ¼ wghp HpE ð2:4Þ

Ht00 HtE
Ht ¼ ¼ ð2:5Þ
ght wght

resulting in:
 
1
Ht w2
h¼ ¼ ð2:6Þ
Hp ghp ght

The hydraulic efficiency of a pump at its best efficiency point (BEP) can be
computed using its overall efficiency by [2]:

g  g0:5
h ð2:7Þ

Finally, the turbine and pump head ratio can be summarized as:

Ht b
h¼ ¼ ; ða [ 1:0; b [ 1:0Þ ð2:8Þ
Hp gap

Assuming gp  gt . Therefore, a pump will work in higher heads in turbine mode


(h > 1.0).
The leakage ratio for reverse and direct operations is [3]:
sffiffiffiffiffiffi
Qlt Ht b0:5
¼ ¼ 0:5a ð2:9Þ
Qlp Hp g p
2.2 PAT Theory 33

In addition, the theoretical flow rate ratio for turbine and pump modes is [3]:
sffiffiffiffiffiffi
Q00t Ht b0:5
¼ ¼ 0:5a ð2:10Þ
Q00p H p gp

Thus, the actual flow rate ratio is:


!
Qt Q00 þ Qlt Q00p þ Qlp b0:5 b0:5
q¼ ¼ 00t  ¼ c ð2:11Þ
Qp Qp  Qlp Q00p  Qlp g0:5a
p g0:5a
p

where c > 1.0. Therefore, a pump will work in higher flow rates in turbine mode
(q > 1.0).
The values of a, b and c may change for various pumps depending on the pump
specific speed and size, mechanical tolerance and manufacturing accuracy. To
summarize, a PAT will work in higher heads and flow rates in turbine mode.

2.3 Flow Pattern in a Pump and in a PAT

Based on expert literature, it has now become clear that in reverse mode:
– the mechanical operation is very smooth and without noise;
– the maximum efficiency of the reverse mode is approximately equal to the
maximum efficiency of the direct mode;
– the output power at its BEP of a PAT is greater than the absorbed power of a
pump;
– the head and the flow rate at its BEP in reverse mode are higher than those in
direct mode.

Head ψ, π η
Power 20 1
Efficiency 0.9
18
PAT Pump
16 0.8
14 0.7
12 0.6
10 0.5
8 0.4
6 0.3
4 0.2
2 0.1
0 0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3
φ
Fig. 2.6 Characteristic curves of a pump in direct and reverse modes [1]
34 2 Reverse Pump Theory

Fig. 2.7 Shock and leakage


flow at the entry of an
impeller in a PAT [5]

On the basis of these important results, the use of a PAT is an acceptable


solution from the technical point of view. Dimensionless characteristic curves of a
pump in direct and reverse modes are compared in Fig. 2.6.
Compared to the direct mode, a pump in reverse mode works in higher heads,
flow rates and powers, but it still shows at its BEP a good value of efficiency [1].
Therefore, the most challenging question is: why are the maximum efficiencies so
high? The pump was not originally designed for reverse working; however, it works
at the same efficiency in turbine mode! To answer this question, losses in both
modes should be investigated. For instance, how great are the shock losses in an
inlet of an impeller in turbine mode compared to those in the pump mode? In fact,
the hydraulic losses are seemingly equal in both modes. Shock losses in the pump
mode usually occur at low velocities. But, due to the absence of a guide vane in a
PAT, these losses are not under control (Fig. 2.7 shows the shock and the leakage
flow at the entrance of the impeller in a PAT). In addition, when passing from an
increased flow rate in a convergence channel (i.e., in reverse mode) to a divergence
channel (i.e., in direct mode), the friction losses increase. Finally, due to a higher
pressure of the flow in a PAT, the leakages are also greater. It seems that there is a
balancing between the various lead losses at the same maximum hydraulic effi-
ciency (at BEP) in both modes (Fig. 2.8).
2.3 Flow Pattern in a Pump and in a PAT 35

Fig. 2.8 Comparison of losses in direct and reverse modes [9]

Therefore, it is believed that at the BEP of turbine and pump modes:


– leakages in reverse mode are higher than in direct mode;
– losses in the volute of a PAT (the convergence path) are less than those in the
pump mode (the divergence path);
– friction losses in the impeller are similar in both modes: on the one hand, they
increase due to the higher flow velocities in PATs and, on the other, they
decrease due to the convergence path of the flow;
– if we consider the BEP as the shock less point, head–flow rate curves can be
divided into two sections: flow rates less than the BEP flow rate and flow rates
greater than the BEP flow rate. Shock losses are considerably higher in the right
hand side of the curve of Fig. 2.8. Therefore, it is strongly recommended that a
PAT must be selected to work in the left hand side of its curve in Fig. 2.8. Low
rotational speeds of a PAT are not recommended.
Given these issues, shock losses have an important role in PAT behavior.
Therefore, for a certain head and flow rate, selecting a PAT with a higher specific
speed is preferred. Stepanoff argued that every good pump can work as a reasonable
turbine; however, each good turbine may not be a suitable pump. It is true even for
multistage pumps [4].
PATs, although suitable for energy production, are very sensitive. Due to the
higher working pressure, the leakages are greater. Therefore, it may be necessary to
improve the seals of the machine. In addition, due to the higher radial and axial
forces, the life of its bearings is shorter. Table 2.1 shows some of the advantages
and disadvantages of PATs.
36 2 Reverse Pump Theory

Table 2.1 Advantages and disadvantages of a PAT operation


Operation Advantages Disadvantages
Hydraulic Reasonable peak efficiency at the BEP Narrow efficiency curve in
off-design points
Adjusting head-flow rate by impeller trimming Higher PAT costs for impeller
or using a specific rotational speed trimming or velocity variation
Mechanical Operates without noise Mechanical seal required, due
to higher leakage
Vibrations are acceptable Short life cycle of bearings due
to higher power
Capability of using a belt instead of direct Strengthening of the shaft and
coupling to a generator to change the PAT coupling required
speed

Last but not least, their lower efficiency, in comparison with conventional tur-
bines, is the major negative point of PATs. Due to the lack of any guide vane, a
PAT has a lower efficiency in off-design points in comparison with conventional
turbines. Therefore, it is important that the pump geometry is modified for use as a
PAT in higher power situations.

2.4 Application of Computational Fluid Dynamics


(CFD) to a PAT

2.4.1 Importance of CFD

There are some hydrodynamic and structural fundamental requirements that all
hydraulic circuit designs have to respect. The variables that have an influence on the
hydraulics, energy efficiency and design are dependent on the local inlet discharge,
as well as on the configuration of the circuit elements (i.e., the trash rack, water
intakes and all the particular features of the hydraulic circuit, turbines and tail race).
All these components, included in the hydraulic circuit, will affect the available
head and the turbine discharge. Characteristics which are normally associated with
inefficiency and unsafe conditions include separate flow zones, vorticity, macro
turbulence intensity, induced vibrations, resonance effects, leakages, ruptures or
collapses, pressure fluctuations, cavitation and water column separation, significant
friction losses due to shear stress development, huge local pressure gradients,
longitudinal swirl vortex and regions of reversed recirculating flow.
Three decades ago, the absence of CFD models led to the use of theoretical analysis
to understand the behavior of a PAT. However, due to the one-dimensional (1D)
nature of such analysis, presenting a complete flow pattern in a PAT was impossible.
In fact, there are many losses, such as secondary flows, that cannot be computed by
one-dimensional models. On the other hand, using experimental methods to capture
the flow behavior in a PAT is really expensive and laboratory tests, measuring
external data such as inlet pressure, outlet pressure, flow rate, rotational speed and
power, are limited in the identification of a PAT’s characteristic curves.
2.4 Application of Computational Fluid Dynamics (CFD) to a PAT 37

Nowadays, CFD models are powerful tools for researchers and pump manu-
facturers for the following purposes:
– simulating the PAT working conditions for a complete set of boundary condi-
tions, whenever only the pump internal geometry is known and experimental
tests are incomplete or too expensive;
– optimizing the internal geometry of the PAT to increase its efficiency;
– verifying the influence of the PAT in the network for water-hammer analyses.
When the flow direction completely changes during the turbine mode, the flow
enters the volute and then the impeller, leaving the machine through the suction
pipe, which is now called the draft tube. Although the geometry of a pump is fixed
when rotating as a turbine, the flow pattern is quite different from its normal
rotation. Fortunately, CFD analyses can clearly show the flow field and the flow
pattern. Some researchers have tested the use of CFD to evaluate a pump working
as a turbine [2, 5–7]
In this section, CFD codes, from commercial solvers, are used to demonstrate
how a better understanding of the behavior of a PAT can be achieved.

2.4.2 Mathematical Approach

Navier-Stokes equations have a limited number of known analytical solutions, but


they are adequate for flow computational models, by means of numerical approa-
ches. CFD models solve Navier-Stokes equations, which are formulations of
momentum plus mass and energy conservation laws for fluid flows. The equations
are supplemented by fluid state equations defining the nature of the fluid, and by
empirical dependencies of fluid density, viscosity and thermal conductivity on
temperature.
Equations (2.12) show Navier-Stokes equations for incompressible flows. These
equations are based on differential equations of linear momentum for a Newtonian
fluid with density and viscosity as constants.
 2 
dp @ u @2u @2u du
qgx  þl þ þ ¼q
dx @x2 @y2 @z2 dt
 2 
dp @ v @ v @ v
2 2
dv
qgy  þl þ þ 2 ¼q ð2:12Þ
dy @x2 @y2 @z dt
 2 
dp @ w @ w @ w
2 2
dw
qgz  þl þ 2 þ 2 ¼q
dz @x2 @y @z dt

where: q: volumetric mass (kg/m3); g: acceleration of gravity (m/s2); p: pressure


(Pa); l: dynamic viscosity (kg/ms); u, v and w: velocity components for each
moving fluid particle, depending on x, y and z coordinates for a given t instant
(m/s).
38 2 Reverse Pump Theory

The incompressible fluid flow behavior is determined by the velocity and


pressure variables and their variations in time and space. In Eq. (2.12), that allow
the attainment of pressure and velocity fields, the velocity components at each point
x, y and z are vector fields and the pressure at each point is a scalar quantity.
Most of the flows that occur at the hydraulic circuits are turbulent, and CFD
models allow the numerical modeling of both laminar and turbulent flow regimes.
The turbulent flow occurs for high values of the Reynolds number, given by the
Eq. (2.13):

qUd
Re ¼ ð2:13Þ
l

where: d: is the pipe diameter (m) and U: is the flow mean velocity (m/s). When the
flow is turbulent, the variables present at each instant random fluctuations.
A fluid under low pressures can reach the vapor pressure at the local temperature
leading to the formation of vaporous cavities. The fluid undergoes a phase change
and cavities filled with vapor fluid and other dissolved gases are formed. When
analyzing areas subject to flow conditions that lead to the occurrence of cavitation,
the CFD model uses a homogeneous equilibrium model of cavitation in water.
CFD allows many calculations both in steady conditions (i.e., the performance
curves of the machine) and in transient flow regimes (i.e., machine stoppage and
start-up, valve closure and load-rejection). This kind of calculation can be very
complex and the computational complexity increases when dynamic simulations
are performed. Thus, great attention is necessary in order to find the minimum
number of elements in the 1D or 3D fluid dynamic mesh necessary to obtain
reliable results for the numerical PAT simulations.

2.4.3 PAT Geometry Modeling and Mesh Generation

To simulate adequately a PAT, the generation of a reliable geometry is very


important. The use of CADCAM software allows a reasonable model of a real
pump. The geometry should be derived according to the exact dimensions of the
prototype. The flow field is split into five component parts: the volute, impeller,
front and back chambers and outlet pipe, as shown in Fig. 2.9. This separation
allows each mesh to be generated individually and tailored to the flow requirements
in each particular component. To obtain a relatively stable inlet and outlet flow, the
pipe diameter should be extended four or five times in the PAT inlet and outlet
sections.
As the motion of the impeller blades relative to the stationary volute is crucial,
the analysis must involve multiple frames of reference. The volute and the outlet
pipe are set in the stationary frame and the impeller is set in the rotary frame. The
interfaces between the two stationary components, the rotary and stationary com-
ponents, are set to general grid and rotor-stator interface, respectively.
2.4 Application of Computational Fluid Dynamics (CFD) to a PAT 39

Fig. 2.9 CADCAM geometry of a PAT [2, 21]

In the process of numerical simulation, all domains except the leakage through
the balancing holes can be included. Therefore, the simulated efficiency includes
volumetric loss through the wear ring and flow loss caused by disc friction. The
volumetric loss caused by the balancing holes are usually neglected.
Different turbulence models are available for the simulation a PAT. However, in
the authors’ experience, k-e RNG and SST k-x are the best models to evaluate
Reynolds Stresses in Navier-Stokes equations. The fluid properties must be con-
sidered in accordance with the site information. For instance, ideal water at 25 °C
can be assumed. All the wall surface roughness within the control volume should be
selected based on the real pump.
Some boundary condition strategies are available. However, one recommended
strategy is to use the static pressure for the inlet and the mass flow rate for the outlet
[2]. By changing the mass flow rate, the performance curves of the PAT can be
achieved.
All numerical simulations face the necessity of making a compromise between
results accuracy and CPU time. In the CFD of a PAT, the primary factors increasing
calculation times are the small dimensions of the wear rings compared to the
dimensions of the computational domain and the time variability of the process for
the rotary frame of the impeller.
Several mesh models can be used to generate structured or unstructured grids for
each component. The possibility of using a variable grid dimension in the com-
putational domain is exploited to reduce the total number of elements without
increasing the mesh size in the wear ring zone. Due to the complexity of generating
a structured mesh based on the geometry, an unstructured grid can be generated for
the volute. In all the other regions, the use of a structured mesh is generally
preferred, especially for the impeller and for the inlet and the outlet pipes. In some
cases, due to the small size of the elements in the wear ring zone, an unstructured
40 2 Reverse Pump Theory

Fig. 2.10 A sample generated mesh for a PAT [2]: a Impeller, b Cut by the draft tube and impeller

Fig. 2.11 Mesh independency curve with a structured grid [2]

grid is used in that section to reduce the time required to set up the model.
Figure 2.10 gives a general view of a structured sample mesh.
A grid independent test of the turbine mode’s performance should be performed
with both kinds of grid. Figures 2.11 and 2.12 show a sample of this strategy for
structured and unstructured grid, respectively. As shown in Fig. 2.11, the effect of
the refinement of a structured grid is an increase in the apparent machine efficiency,
approaching the experimental value. Reducing the mesh size, the narrowest pas-
sages of the pump are better modeled, as the internal leakages are lower, and,
additionally, the computed efficiency is increased. When the mesh numbers are
around one million, the variation of efficiency is about 0.5%, as indicated. The same
2.4 Application of Computational Fluid Dynamics (CFD) to a PAT 41

Fig. 2.12 Mesh independency curve with an unstructured grid [6]

behavior is observed in Fig. 2.12, where the results of many simulations with an
SST k-x turbulence model on 24 meshes of different minimum element sizes are
plotted. The head increases with the decrease in the number of elements, and the
maximum error in relation to the value obtained with the 5 million elements mesh is
greater than 7.4%. The efficiency shows the same trend, with a maximum error of
18.8%. The spreading of the CFD results is determined by the uncontrolled mesh
geometry of the unstructured grid.
Special care should be taken in the machine inlet pipe mesh generation, in order
to ensure that the mesh between the interface of the inlet pipe and the front chamber
leakage is almost the same. In the case of multistage pumps, a minimum number of
three stages has to be simulated. Thus, the domain of the middle stage is far from
the boundaries and its fluid dynamic field is reliable.
y+ takes into account the effect of the wall on the shear stress. Reynolds models
aim to identify laminar sub-layers that can be used to calculate the height of the first
near-wall cell based on a specific y+ value in order to determine the effect of wall
friction more accurately. An appropriate value of y+ has to be selected based on the
turbulence model and wall-treatment method for a robust model simulation.
42 2 Reverse Pump Theory

The second aspect that could have an influence on the CFD results is the kind of
flow field variability considered in the numerical problem. Both steady and transient
flow simulations can be performed in order to obtain the performance curves of a
turbo-machine. The first type of simulation solves the steady state problem for a
fixed position of the impeller, while the second considers different positions of the
impeller for different time instants.
In the transient case, the solution is obtained as an average of the different
transient solutions. Obviously, steady state calculations significantly reduce the
computation time, but appreciable differences have been observed between steady
state solutions and the experimental results.

2.4.4 CFD Solutions for a PAT

Based on CFD models, a simulation of the hydrodynamic flow in a PAT can be


performed [8]. The computational analysis of each test aims to:
– obtain the distribution of descriptive physical parameters of the hydrodynamic
flow in planes that intersect the representative geometric model of a PAT;
– make a comparison between the velocity diagrams obtained experimentally, in
the field, and computationally.
In accordance with Sect. 2.4.2, in order to make a computer simulation of the
flow in the PAT and its installation, it is necessary to build a representative geo-
metric model by using a CAD system or a solid work which is then imported into
the CFD model (Fig. 2.13). The geometric model constructed results from the
meeting of several independent solid components, of which the most important are
the rotor and the volute.
The rotational behavior induced in the flow by the impeller rotational speed and
the shape of the respective vanes is observed in Fig. 2.14a. This variation results
from the centrifugal force induced by the rotor rotation.
The flow enters the volute and focuses radially on the rotor, which imposes a
speed of rotation. In turn, when the rotation of the rotor continuously varies the

Fig. 2.13 PAT model: a Complete geometric model, b the PAT rotor, and c the volute [8]
2.4 Application of Computational Fluid Dynamics (CFD) to a PAT 43

Fig. 2.14 Flow condition: a Stream lines (m/s) along the geometric model, and b the turbulence
intensity distribution (%) in a longitudinal plane of the rotor [8]

Fig. 2.15 Static pressure distribution: a a longitudinal plane at the ball valves, b a longitudinal
plane of the rotor, and c in a plane transverse to the rotor [8]

direction of the flow along the passage through the rotor, consequently, the outlet
flow direction is changed. The continuous variation of the direction of the flow
gives rise to a centrifugal force, which has the effect of spacing the flow rotor
rotation axis concentrating it on the periphery. The presence of this hydrodynamic
behavior justifies the higher flow velocity values which occur at the rotor periphery.
The pair with the highest tangential velocity values also occurs on the periphery of
the rotor values with the greatest turbulence intensity, as seen in Fig. 2.14b. The
rotation of the rotor also induces the increase in the flow turbulence intensity.
Since the PAT converts mechanical to electrical energy, the pressure that exists
in the PAT downstream of the rotor is lower than the pressure upstream
(Fig. 2.15a), creating the available head. The differential pressure reflects the net
head for each turbine’s operating conditions. Figure 2.15b shows a reduction in the
pressure from upstream to downstream and along the flow passage in the rotor. It
shows the extraction of energy from flow pressure, promoted by the rotor.
When a vortex appears in the rotor’s outlet and extends downstream, associated
with the turbulence and hydrodynamic instability, pressure fluctuations and loss of
efficiency occur. Figure 2.15b, c show lower values of static pressure, respectively
along the rotor axis and at the axis of the diffuser duct, which is the core of the
vortex that is formed at the output of the downstream rotor. The pressure reduction
characteristic of the cores of the vortices that occurs downstream of the rotors can
lead to cavitation. Under certain operating conditions, the lower values of pressure
44 2 Reverse Pump Theory

Fig. 2.16 Characteristic curves of a PAT obtained by CFD and experimentation [2]

occurring along the rotor axis (Fig. 2.15b) and along the axis of the draft tube
(Fig. 2.15c) are significantly higher than the saturation pressure water vapor.
Therefore, no vapor bubbles or cavitation occurs in this case.
Figure 2.16 shows a comparison of the characteristic curves of a PAT obtained
by CFD and experiments with a single stage pump.
In Fig. 2.17 a similar comparison is performed for a multistage pump, including
the difference between the characteristic curves of a PAT obtained by CFD, in
steady and unsteady state simulations, and experiments. The better performance of
the unsteady state simulations is evident. The most important ability of CFD models
is the capability of fluid flow analyses. Flow parameters such as velocity and
pressure can be easily studied with many available methods of presenting the
results. 2D and 3D color contours can be used to show the parameter variety from
inlet to outlet. Figure 2.18 shows a sample velocity contour in a PAT. Other
contours can also be depicted, if necessary. In addition, some extra stream flows can
be used to detect the appearance of secondary flows.
Additionally, knowledge of the pressure distribution in a PAT can be useful to
locate the point of minimum pressure. In the presence of low pressure at the PAT
outflow this point is a potential source of cavitation. Figure 2.19 shows the pressure
distribution in a multistage pump, for two different mesh resolutions.
Figure 2.20 shows the hydraulic head drop distribution in the different zones of a
PAT: the volute, gap between volute and impeller, impeller and outlet pipe. It is
clear that the impeller is the main component of a PAT for the efficiency.
Additionally, losses in the gap between the impeller and the case can have an
important role. Therefore, it is recommended to make efforts possible to improve
the efficiency of a PAT by refining the impeller and gap head drops.
2.5 Cavitation in a PAT 45

Fig. 2.17 Comparison of


steady and unsteady state
conditions [6]

2.5 Cavitation in a PAT

A major limitation in any hydraulic machinery operation is the phenomenon of


cavitation. Whenever, in the suction pipes of pumps and turbines, the local pressure
maybe drops to the values of the vapor pressure of the liquid, the liquid phase can
rapidly convert to the vapor phase. Pump manufacturers normally determine the
cavitation limitation of their products by NPSH1 (Fig. 2.21), directly looped by the
location of the pump installation above the tail water (Fig. 2.4):

1
Net Positive Suction Head.
46 2 Reverse Pump Theory

Fig. 2.18 Velocity contour in a PAT [21]

Fig. 2.19 Comparison between pressure fields: a a higher mesh resolution; and b a coarse mesh
[6]
2.5 Cavitation in a PAT 47

Fig. 2.20 Loss distribution in different zones of a PAT [21]

Fig. 2.21 Typical NPSH


curve of a pump

 
Pa Pv
hs;max ¼  þ NPSHp þ hf ð2:14Þ
qg qg

where hs;max (m) is the maximum suction head, Pa (Pa) is the atmosphere pressure,
Pv (Pa) is the water vapor pressure and hf (m) is the hydraulic loss in the suction
pipe. NPSHp (m) is the Net Positive Suction Head, available from a pump catalogue
(Fig. 2.21).
48 2 Reverse Pump Theory

Considering Pv for a normal water temperature (3 kPa for 25 °C), and Pa for an
average height of the site location (101 kPa sea level) and assuming 0.5 m for
losses in the suction pipe, hs,max can be estimated by:

hs;maxp  9:0  NPSHp ð2:15Þ

Providentially, hs,max for turbines is less than that for pumps. As a PAT works at
a higher pressure than a pump, hs,maxPAT < hs,max p for each corresponding rated
point in direct and reverse modes [9]. A PAT may work at a higher rotational speed
than its direct mode. Using affinity (or similarity) laws in turbo-machinery, the
Eq. (2.15) will be modified to:
 
NPAT 2
hs;maxPAT ¼ 9:0  NPSHp ð2:16Þ
Np

where, NPAT and NP are the rotational speeds of reverse and direct modes,
respectively.

2.6 Pressure Surge Analysis in a PAT

One of the main reasons for the still limited use of PATs in water supply networks
(WSN) is due to doubts about the behavior of this electromechanical device during
the lifetime of the WSN. The WSN technical management is responsible for pre-
venting any worsening in the running of the water service. This attitude is quite
comprehensible because water distribution is the main target of water companies
and a suspension of the service could lead to penalties and to customer
dissatisfaction.
Three aspects that have to be considered in micro hydropower design are rep-
resented by: (i) the effect on the WSN of the power plant start and stop procedures;
(ii) the effects on the WSN of power plant load rejection or power failures; and
(iii) the behavior of the PAT in the presence of pressure surges reaching the power
plant node from different points of the network.
A power plant failure is possible on account of either a PAT failure or an electric
grid failure. Any PAT failure could be of an electrical or mechanical type. The most
serious situation is represented by an abrupt PAT stop or by the occurrence of a
runaway or overspeeding of the machine due to a sudden black-out, namely a total
load-rejection.
It is important to understand that experimental results show differences in the
dynamic response of a PAT and a pressure relief valve (PRV) in the presence of
similar flow conditions (Fig. 2.22). In addition, the wave period is also different and
there is a fast damping of pressure oscillations in the presence of a PRV, with
higher extreme pressures than with a PAT.
2.6 Pressure Surge Analysis in a PAT 49

Fig. 2.22 Dynamic response of a a PAT and b a PRV [22]

Table 2.2 Maximum H0 (m) Allowable maximum DH/H0


allowable DH/H0 as a
function of the design head >100 0.15–0.20
100–20 0.25–0.35
<20 0.50

The hydraulic transient effects will strongly depend on the overall hydro-system
characteristics, including the number of units and the type of PAT. In all cases the
main objective is to prevent any serious damage to the penstocks or other pressurized
pipes, as well as to any other components of the hydro-system. Regarding structural
penstock protection both the maximum and minimum transient pressures need to be
controlled in order to avoid the possibility that: (1) the allowable maximum pressure
is exceeded, thereby preventing any breakdown or pipe burst; and (2) the
sub-atmospheric pressure is increased, thereby preventing any cavitation phenomena
and water column separation effects, as well as a potential pipe wall buckling event.
Typically, the allowable maximum relative transient head variations DH/H0 will
depend on the design head, the range and dependency, as shown in Table 2.2.
In all cases, the downsurge, or minimum transient pressure, should also be
properly considered in order that the pipe system is protected against atmospheric
pressures and vacuum formation (e.g., water column separation).
Any disturbance induced in the flow will propagate as a wave with a finite
relative velocity, known as the elastic wave celerity, whose value will depend on
the elastic properties of both the fluid and the pipe according to [10]
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K
c¼ ð2:17Þ
q½1 þ ðK=EÞW

where E—Young’s modulus of elasticity of the pipe wall; K—fluid bulk modulus
of elasticity (2  103 MPa for water); W—parameter that depends upon the pipe
structural constraints; q—fluid specific mass (1000 kg/m3 for water).
In small hydro schemes, when equipped with small inertia turbines, the
parameterization of water hammer effects might allow a better characterization of
50 2 Reverse Pump Theory

the dynamic behavior of these turbo-machines, assisting in the choice of the most
appropriate solutions.
Several numerical methods are available for the prediction of maximum and
minimum pressures in each network point in the presence of unsteady state con-
ditions [11]. Numerical simulations are commonly based on the one-dimensional
momentum and continuity equations:
   
@H @H f @V @V @V
g þ þ V jV j þ k þ sgn V c ¼0 ð2:18Þ
@x @t 2D @t @x @x

@H c2 @V
þ ¼0 ð2:19Þ
@t g @x

where V is the pipe cross average velocity, H is the piezometric head and c is the
pressure wave celerity. The last term in the momentum equation is used for highly
deformable pipe materials, producing a variability of wave celerity and a reduction
of the pressure peaks during the transient [12, 13]. The WSN interior points could
be solved using the method of characteristics (MOC). MOC transforms these
equations into a pair of ordinary differential equations valid along the characteristic
lines in the (x, t) plane. These equations can be replaced by algebraic equations that
are solved together with the other basic equations of the computational model.
MOC is the most common numerical scheme used to solve the transient problem in
a discrete number of pipelines nodes, based on the solution of Eqs. (2.18) and
(2.19), in the WSN initial flow conditions and in the WSN boundary conditions.
In the presence of a PAT in the hydropower node, generating a pressure surge or
reacting to a pressure wave generated in a different node of the network, a few
alternatives are available for transient calculations. In particular, the main alterna-
tives concern the schematization of the electro mechanical device.
During normal steady-state operating conditions the speed of a turbine is constant,
but it rises quickly when the load of the turbine is rejected. During the closure of a
flow control valve, the speed of the turbine rises to a maximum value, under runaway
conditions, that cannot surpass the allowable value fixed by the PAT manufacturer.
With reaction turbines, as PATs are, the runner speed changes the turbine dis-
charge and the overspeed conditions also influence the hydraulic circuit [14]. The
transient overspeed can induce an additional source, a water-hammer wave, espe-
cially in relation to PATs with small N values. A complete computer analysis should
be performed based on the PAT characteristic equations, the complete hydraulic
Eqs. (2.18) and (2.19) and the equation of the rotating masses of each unit:

dx
TH  TR ¼ I ð2:20Þ
dt

in which TH is the transient hydraulic torque (N m), TR is the resistant or electrical


torque (N m), I is the rotating mass inertia (kg m2) and x is the angular rotating
speed (rad/s).
2.6 Pressure Surge Analysis in a PAT 51

The moment of inertia of a pump reflects its resistance to variations of the wheel
while rotating on its shaft. Pumps with high rotating masses have higher values of
inertia: they take longer to reach the rated speed during start-up and to slow down
and lose power in the case of a planned or accidental shut off [10]. The inertia of the
PAT is given by the sum of the inertia of the pump and motor and the following
expressions (2.21) and (2.22) are proposed [10, 11]:
 1:435
P
Ipump ¼ 912 ð2:21Þ
N
 1:48
P
Imotor ¼ 118 ð2:22Þ
N

in which P is the power (kW); N is the rotational speed (rpm); and I is the inertia of
the rotating mass (kg m2).
Due to the need to test all the different failure scenarios, experimental or numerical
PAT characteristic curves could be used in the PAT node as internal conditions
between two network branches. This method can be considered fully one-dimensional,
treating the fluid mass in the PAT as a single node. In a more complex procedure [6, 7,
15] a new computational scheme is proposed coupling CFD and MOC. In Fig. 2.23
the compatibility conditions between the two computational schemes are reported: the
coupling was performed by modeling two pipe branches in CFD calculations, at the
inlet and outlet of the PAT, for flow regulation, and by imposing compatibility
equations between CFD and MOC at the extremes of the branches.
Differences are observed when running a CFD incompressible (rigid) or com-
pressible fluid (elastic) model [7]. In Figs. 2.24 and 2.25 time-head plots are shown

Fig. 2.23 Compatibility conditions between CFD and MOC


52 2 Reverse Pump Theory

Fig. 2.24 Head variation


after a sudden discharge
variation (rigid model) [4]

Fig. 2.25 Head variation


after a sudden discharge
variation (elastic model) [7]

after a sudden finite variation of discharge, from 35 to 30 l/s, with a rigid and elastic
model, respectively. According to the first model, after a sudden initial drop of the
head, the machine gradually tends to new steady conditions. These conditions are
achieved after 1/3 rounds of the impeller. The results of the elastic model are
significantly different. After the variation of the discharge, a pressure wave is
originated and reflected, causing severe head fluctuations. These fluctuations are
much wider than the maximum upsurge calculated with the rigid model simulation
and the effects of the discharge variation take a longer time to be dissipated.
The coupled CFD + MOC model, which was validated on experimental data at
the Instituto Superior, Tecnico University of Lisbon (Portugal), was acquired on the
simple hydraulic scheme of Fig. 2.26 [16]. Tests were performed for a sudden
closure of the downstream valve. With a dumping factor k = 0.34 in Eq. (2.18)
accounting for the viscoelastic behavior of the pipe system, a good agreement
between the experimental and numerical piezometric head immediately upstream
the regulation valve was observed (Fig. 2.27).
On the same hydraulic scheme, but with a different PAT and pipeline size, some
comparisons have been made between the transient results obtained by using a
one-dimensional model (MOC + one dimensional characteristic curve; CrCv in the

Fig. 2.26 Hydraulic scheme for water hammer analysis


2.6 Pressure Surge Analysis in a PAT 53

EXP
25
20 k=0
k = 0.34
15
10
H [m]

5
0
-5
-10
-15
0 2 4 6 8
t [s]

Fig. 2.27 Model and experimental heads at the downstream valve

following figures), the steady PAT characteristic curves obtained by a Suter


parameters calculation, or the coupled CFD + MOC model (CFD). The calculations
were performed for three different closure times of the stroking valves, Tc = 0 s
(Case I), Tc = 0.15 s = 0.75s (Case II) and Tc = 0.45 s = 1.125s (Case III). In
Figs. 2.28, 2.29 and 2.30, for Case I, Case II and Case III, respectively, the head

Fig. 2.28 Head and discharge in Case I [6]


54 2 Reverse Pump Theory

Fig. 2.29 Head and discharge in Case II [6]

and the discharge downstream the PAT, divided by the initial values, are plotted.
Some appreciable differences appear in the calculated values of discharge and head,
especially when the flow reverses. These differences are greater when the valve
closure is faster and in Fig. 2.28 some spikes in both the head and discharge
diagrams are visible. These spikes are generated by the response of the rotating
machine to the water hammer waves and they are greater when the wave is steeper.
Similar results were found by Ismaier and Schlücker in 2009 [17], in experi-
ments on the interaction between water hammer waves and pumps. They showed
that the rotating machine generates pulsations which interact with the water hammer
phenomenon and that the pulsation effect increases with the rotating speed. While
the usual 1-D models that use the steady characteristic curves do not allow a
simulation of the impulse response of the machine, the CFD + MOC model is a
reliable method to predict this phenomenon.
2.6 Pressure Surge Analysis in a PAT 55

Fig. 2.30 Head and discharge in Case III [6]

Finally, in Fig. 2.31 the net head-discharge curves of a PAT are plotted as
calculated by the coupled CFD + MOC model in unsteady conditions, and the
difference discrepancies between these curves and the steady characteristic curve
are clearly noticeable.
The control of transients and the dynamic behavior of hydropower plants are
fundamental for the design and exploitation, in order to guarantee a safe and reliable
solution. The effectiveness of new design criteria based upon computational tech-
niques should allow an analysis of the global behavior in order to identify possible
operational constraints, from the beginning of the design and exploitation process,
in particular situations that are vital for the total WSN safety.
56 2 Reverse Pump Theory

Fig. 2.31 Net head-discharge curves of a PAT under unsteady conditions in Cases I, II, and III,
respectively

References

1. S. Derakhshan, A. Nourbakhsh, Experimental study of characteristic curves of centrifugal


pumps working as turbines in different specific speeds. Exp. Therm. Fluid Sci. 32(3), 800–807
(2008)
2. S.-S. Yang, S. Derakhshan, F.-Y. Kong, Theoretical, numerical and experimental prediction
of pump as turbine performance. Renew. Energy 48, 507–513 (2012)
References 57

3. A. Williams, Pumps as Turbines used with Induction Generators for Stand-alone


Micro-hydroelectric Power Plants (Nottingham Trent University, UK, 1992)
4. A. Nourbakhsh, S. Derakhshan, E. Javidpour, A. Riasi, Centrifugal & axial pumps used as
turbines in small hydropower stations, in Hidroenergia 2010: International Congress on
Small Hydropower International Conference and Exhibition on Small Hydropower (2010),
pp. 16–19
5. S. Derakhshan, A. Nourbakhsh, Theoretical, numerical and experimental investigation of
centrifugal pumps in reverse operation. Exp. Therm. Fluid Sci. 32(8), 1620–1627 (2008)
6. A. Carravetta, O. Fecarotta, H. Ramos, Numerical simulation on pump as turbine: mesh
reliability and performance concerns, in 2011 International Conference on Clean Electrical
Power (ICCEP) (2011), pp. 169–174
7. O. Fecarotta, A. Carravetta, H. Ramos, CFD and comparisons for a pump as turbine: mesh
reliability and performance concerns. Int J Energy Env. 2(1), 39–48 (2011)
8. A. Pereira, H. Ramos, CFD for hydrodynamic efficiency and design optimization of key
elements of SHP. Int. J. Energy 1(6), 937–952 (2010)
9. J.-M. Chapallaz, P. Eichenberger, G. Fischer, Mini Hydro Power Group., and German
Appropriate Technology Exchange., in Manual on Pumps Used as Turbines (Vieweg, 1992)
10. H. Ramos (Editor), Guidelines for design of small hydropower plants. WREAN (Western
Regional Energy Agency and Network) and DED (Department of Economic Development—
Energy Division), Belfast, North Ireland, ISBN 972-96346-4-5 (2000)
11. E. Wylie, V. Streeter, in Fluid Transient in Systems (Englewood Cliffs, NJ 7632: Prentice
Hall, 1993)
12. B. Brunone, U.M. Golia, M. Greco, Effects of two-dimensionality on pipe transients
modeling. J. Hydraul. Eng. 121(12), 906–912 (1995)
13. G. Pezzinga, Evaluation of unsteady flow resistances by Quasi-2D or 1D models. J. Hydraul.
Eng. 126(10), 778–785 (2000)
14. H. Ramos, Simulation and Control of Hydrotransients at Small Hydroelectric Power Plants
(Technical University of Lisbon, Portugal, 1995)
15. X.X. Zhang, Y.G. Cheng, L.S. Xia, J.D. Yang, Dynamic characteristics of a pump-turbine
during hydraulic transients of a model pumped-storage system: 3D CFD simulation. IOP
Conf. Ser. Earth Environ. Sci. 22(3), 32030 (2014)
16. O. Fecarotta, Impiego Di Pompe Come Turbine: calcolo fluidodinamico per l’analisi
prestazionale e l’inserimento in rete (2011)
17. A. Ismaier, E. Schlücker, Fluid dynamic interaction between water hammer and centrifugal
pumps. Nucl. Eng. Des. 239(12), 3151–3154 (2009)
18. A. Nourbakhsh, A. Jaumotte, C. Hirsch, H. Parizi, in Turbopumps and Pumping Systems
(Springer Science & Business Media, 2007)
19. J.F. Gülich, Centrifugal Pumps (Springer, Berlin, 2008)
20. P. Singh, Optimization of the Internal Hydraulic and of System Design in Pumps as Turbines
with Field Implementation and Evaluation (2005)
21. S.-S. Yang, F.-Y. Kong, H. Chen, X.-H. Su, Effects of blade wrap angle influencing a pump
as turbine. J. Fluids Eng. 134(6), 61102 (2012)
22. H. Ramos, A.B. Almeida, Dynamic orifice model on waterhammer analysis of high or
medium heads of small hydropower schemes. J. Hydraul. Res. 39(4), 429–436 (2001)
Chapter 3
Industrial Aspects of PAT Design
Improvement

Abstract A PAT normally has a lower efficiency than a conventional turbine


applied at the same site. In the long term, a higher efficiency is important for users
in that more energy is produced. A pump’s geometry can be modified to improve its
efficiency in turbine mode. Several methods of PAT geometry modification are
presented for the improvement of PAT efficiency. Optimization of the internal
pump geometry for reverse use is fundamental also for the reduction of the cost of
PAT maintenance. Therefore, the industrial aspects of PAT reliability in terms of
different design methods, solutions and recommendations are described. Practical
technical skills relating to the PAT inlet/outlet parts and impeller geometry modi-
fication in order to improve the PAT efficiency are presented. Basic concepts in
relation to mechanical reliability are given and the analytical methods to assess the
reliability of an industrial PAT are provided.

Although pumps are much cheaper than conventional turbines, their efficiency is
usually lower than hydraulic turbines. Figure 3.1 shows a comparison of the effi-
ciency curves of different machines in respect of dimensionless flow rate.
Obviously, a PAT has a low efficiency even at its Best Efficiency Point (BEP). In
addition, the efficiency curve of a PAT is a narrow curve, especially in partial load.
On the other hand, hydro sites often have daily, monthly or seasonal fluctuation
flow rates. This means that a PAT maybe works in its partial load several times
during its daily operation. Therefore, to increase the range of PAT application, its
performance in part load and at BEP should be improved.
Considering PAT components, there are three geometric parts of a PAT that can
be modified to improve machine performance: the impeller, inlet part and outlet
part.

© Springer International Publishing AG 2018 59


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_3
60 3 Industrial Aspects of PAT Design Improvement

Fig. 3.1 Efficiency curves of


conventional turbines and a
PAT [1]

3.1 Impeller Refinement

It is crystal clear that the impeller, as the major component of a PAT, has the most
important effect on its performance. While the impeller was originally designed to
work in pump mode, its refinement can improve the machine’s performance in
turbine mode. In fact, the impeller diameter and width and the blade geometry
profile and its leading edge can be changed according to the target curve of
efficiency.

3.1.1 Rounding of the Blade’s Leading Edge


and the Hub/Shroud Inlet Edges of a PAT

Figure 3.2 illustrates modifications of the edges, which involve a rounding of the
blade shape and the sharp edges of the hub/shroud in the inlet of the PAT impeller.
On the blade edge profile, a rounding is made at the radius so as to be equal to half
of the blade thickness but great care is taken to ensure that the overall diameter is
not altered. Therefore, there is a certain trade-off in selecting the rounding radius.
The hub and shroud are subject to rounding both at their outer and inner edges.
These modifications could reduce the net flow separation losses component in the
blade and the impeller. However, they may also cause a rearrangement of the inlet
velocity triangle that could change the shock losses component. A loss reduction
associated with the shock wave coming into contact with smoother geometric
profiles and a complex interaction in the clearance between the volute and impeller
(with a rounded hub/shroud) can be expected.
3.1 Impeller Refinement 61

Fig. 3.2 a Rounding of the blade profile at the impeller inlet in turbine mode. b Rounding of the
hub/shroud inlet edges [2]

Figure 3.3 shows the experimental results of edge rounding. In these figures, w,
/, and p are defined as [2]:

gH Q P P
w¼ ;/ ¼ ;p ¼ 3 5 ;g ¼ ð3:1Þ
2
nD 2 nD 3 qn D qgHQ

where, g (m2/s), q (kg/m3), H (m), Q (m3/s), P (W) and η are the gravity accel-
eration, water density, head, flow rate, power and efficiency, respectively. n (rps) is
the rotational speed and D (m) is the impeller diameter.
As proven by experience, an efficiency improvement occurs in all flow rates
of the part load and overload zones. Rounding the impeller can improve the PAT’s
maximum efficiency by up to 5%. This modification can reduce the net flow sepa-
ration losses component at the blades and impeller and also cause a rearrangement in
the inlet velocity triangle and a change in the shock losses component.
62 3 Industrial Aspects of PAT Design Improvement

30 80
28 Rounded
26 Optimized 70
24 Initial
22 60
20
50
18
16
ψ 40 η(%)
14
12
30
10
8 20
6
4 10
2
0 0
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14
φ

Fig. 3.3 Experimental results for head number and efficiency [2]

3.1.2 Impeller Diameter (Trimming Effect)

Impeller trimming is a conventional method by which pump manufacturers aim to


adjust their pump following a client inquiry. Heads and flow rates decrease with a
trimming of the impeller. However, the pump’s efficiency also reduces. Reports
have shown that the same phenomena occurs for a PAT after impeller trimming.
Impeller trimming, therefore, although useful for hydraulic data adjustment, leads to
a lower efficiency. Figure 3.4 shows a sample of a PAT curve with various impeller
diameters (255, 235 and 215 mm) [3].
A PAT’s efficiency decreases with the trimming of its impeller diameter.
According to Fig. 3.4, as the impeller is trimmed from 255 mm to 215 mm, the
PAT’s efficiency at its BEP drops by 4.11%, The Q–H curve becomes increasingly
steeper and the Q–P curve moves downwards in an almost parallel way. The
decrease of the diameter shifts the flow rate at the BEP towards the left of its
performance curve.
Therefore, it is recommended that impeller trimming of a PAT should be
avoided due to the considerable decrease in efficiency.

3.2 Inlet/Outlet Part Refinement

PAT performance can be improved by a simple modification of the inlet and outlet
parts. Figure 3.5 shows a schematic flow pattern in both a volute outlet and impeller
inlet in a PAT. A lack of guide vanes leads to a second flow in the gap area between
3.2 Inlet/Outlet Part Refinement 63

Fig. 3.4 Head, power and efficiency curves of a PAT with various impeller diameters [3]

the impeller and the volute. In conventional turbines, the flow is guided reasonably
to the impeller with a minimum vortex.
Using a rib as a fixed guide vane can slightly improve the PAT performance.
Figure 3.6 shows a guide rib for a better guiding of the flow into the impeller.
Singh [5] reported an improvement of 1% in the BEP efficiency of a PAT with
Ns = 79.1.
In addition to the inlet modification of a PAT, a rearrangement of the outlet
elements can help to improve performance. According to the conventional reaction
turbine arrangement, a draft tube can be reasonably effective in the outlet energy
recovery of a turbine. The draft tube can convert the kinematic energy into potential
energy, and therefore, the flow losses decrease. Another reason for the draft tube
application is to increase the impeller outlet pressure to prevent cavitation arising.
However, the PAT is in a safe mode, which, considering the pump Net Positive
Suction Head (NPSH), means that there is no cavitation in a PAT. However, a
simple elbow or straight pipe can help the PAT performance improve. On the other
hand, economic considerations have the principal role in terms of PAT application.
Therefore, any decision on the use of a draft tube in a PAT should be made taking
into account both technical and economic issues (Fig. 3.7).
64 3 Industrial Aspects of PAT Design Improvement

Fig. 3.5 Flow pattern of a


volute outlet B and impeller
inlet A in a PAT [4]

Fig. 3.6 Flow pattern in a


volute outlet and impeller
inlet A in a PAT after the
insertion of a guide rib [4]
3.3 PAT Reliability 65

Fig. 3.7 A PAT arrangement


with a simple draft tube [1]

3.3 PAT Reliability

Reliability represents the probability that a component, system or process will work
without presenting any failure during a specified period of time, in under the correct
operating conditions, as specified by the manufacturer [6]. Any failure of an
engineering system occurs when the load, L (the external forces or demand),
becomes larger than the resistance, r, of the system for a specified duration, Dt.
Therefore, the reliability, R, of a system can be defined as the probability, p, that
during the time interval Dt, the load will not exceed the resistance:

R ¼ pðL\r ÞjDt ð3:2Þ

If the load, L, can be considered constant for specified conditions during Dt, the
reliability can be considered also as the probability that the time to failure, TTF, is
larger than Dt:
66 3 Industrial Aspects of PAT Design Improvement

R ¼ pðTTF [ DtÞjL ð3:3Þ

The reliability of an engineering system is often represented by an exponential


probability distribution [7]:

RðtÞ ¼ ekt ð3:4Þ

where k is the failure rate, equal to 1/MTTF, introducing the mean time to failure
MTTF [8].
Barringer [9, 10] performed accurate studies on the reliability of pumps and
demonstrated that the pump flow rate has a marked impact on pump reliability when
the pump is operating away from its BEP: the reliability is maximum at the BEP
and decreases the more the flow rate differs from the BEP flow rate. This behavior is
easily explained by the fact that the pump design is performed around the BEP
conditions, while the different kinds of load can increase moving far from the BEP
and exceeding the resistance of the pump components.
The reliability or the probability that the machine works without failure for the
whole period T, equals the probability that the machine does not fail in any of its M
components in the same time period. The reliability can be computed by the
equation:

 Y
M 
RðT Þ ¼ R1 ðT ÞjL1 . . .Rj ðT ÞLj . . .RM ðT ÞjLM ¼ R1j ðT ÞLj ð3:5Þ
j¼1

where each component j is subject to a different external load Lj.


Expressing the reliability by an exponential probability distribution, we have:

Y
M  Y
M PM
 kT
RðT Þ ¼ Rj ðT ÞLj ¼ ekj T ¼ e j¼1 j ð3:6Þ
j¼1 j¼1

and a failure rate:

X
n
k¼ kj Dti ð3:7Þ
i¼1

The reliability of the electro-machinery is expressed by:

RðT Þ ¼ ekT ð3:8Þ

And the average mean time to failure of the system under variable working
conditions, MTTF, can be calculated as 1 = 1/k.
The inherent life of the pump components at BEP can be described by using
Weibull statistical formats. For three well known industrial pump fabrication
3.3 PAT Reliability 67

standards (ANSI, ANSI enhanced and API), the Weibull parameters are given in
Table 3.1 BEP Weibull parameters.
Weibull characteristic values will be different, due to the reduction of the reli-
ability of each single component when the pump is working far from its BEP. The
working conditions, in terms of the scatter of the working flow rate with respect to
the BEP flow rate, for best practice (−10 to +5%), better practice (−20 to +10%)
and (−30 to +15%) are defined and the corresponding Weibull characteristic values
are given in Tables 3.2, 3.3 and 3.4. Any installation errors—such as a lack of
rotational balance, or suction pipeline alignment, or an incorrect foundation and
grouting—strongly influence the number of failures and specific reductions of
component reliability have to be considered [9].
The value of the pump reliability at the BEP, as calculated by 3.5–3.8 for the
three standards, is plotted in Fig. 3.8 as a function of the time period. The reliability

Table 3.1 BEP Weibull parameters


ANSI Enhanced ANSI API
Beta Eta Beta Eta Beta Eta
Impeller 2.5 300,000 2.5 300,000 2.5 400,000
Housing 1.3 300,000 1.3 300,000 1.3 400,000
Pump bearings 1.3 100,000 1.3 200,000 1.3 400,000
Seals 1.4 100,000 1.4 200,000 1.4 400,000
Shafts 1.2 300,000 1.2 300,000 1.2 400,000
Coupling 2 100,000 2 100,000 2 300,000
Motor bearings 1.3 150,000 1.3 150,000 1.3 150,000
Motor windings 1 150,000 1 150,000 1 150,000
Motor rotor 1 300,000 1 300,000 1 300,000
Motor starter 1.2 300,000 1.2 300,000 1.2 300,000

Table 3.2 Best practice Weibull parameters


ANSI Enhanced ANSI API
Beta Eta Beta Eta Beta Eta
Impeller 2.5 291,780 2.5 291,780 2.5 389,040
Housing 1.3 256,410 1.3 256,410 1.3 341,880
Pump bearings 1.3 97,120 1.3 194,240 1.3 388,480
Seals 1.4 96,770 1.4 193,540 1.4 387,080
Shafts 1.2 291,360 1.2 291,360 1.2 388,480
Coupling 2 98,010 2 98,010 2 294,030
Motor bearings 1.3 150,000 1.3 150,000 1.3 150,000
Motor windings 1 150,000 1 150,000 1 150,000
Motor rotor 1 300,000 1 300,000 1 300,000
Motor starter 1.2 300,000 1.2 300,000 1.2 300,000
68 3 Industrial Aspects of PAT Design Improvement

Table 3.3 Better practice Weibull parameters


ANSI Enhanced ANSI API
Beta Eta Beta Eta Beta Eta
Impeller 2.5 264,000 2.5 264,000 2.5 352,000
Housing 1.3 219,000 1.3 219,000 1.3 292,000
Pump bearings 1.3 79,000 1.3 158,000 1.3 316,000
Seals 1.4 88,000 1.4 176,000 1.4 352,000
Shafts 1.2 237,000 1.2 237,000 1.2 316,000
Coupling 2 92,000 2 92,000 2 276,000
Motor bearings 1.3 141,000 1.3 141,000 1,3 141,000
Motor windings 1 150,000 1 150,000 1 150,000
Motor rotor 1 282,000 1 282,000 1 282,000
Motor starter 1.2 300,000 1.2 300,000 1.2 300,000

Table 3.4 Commercial practice Weibull parameters


ANSI Enhanced ANSI API
Beta Eta Beta Eta Beta Eta
Impeller 2.5 204,000 2.5 204,000 2.5 272,000
Housing 1.3 210,000 1.3 210,000 1.3 280,000
Pump bearings 1.3 65,000 1.3 130,000 1.3 260,000
Seals 1.4 51,000 1.4 102,000 1.4 204,000
Shafts 1.2 195,000 1.2 195,000 1.2 260,000
Coupling 2 76,000 2 76,000 2 228,000
Motor bearings 1.3 117,000 1.3 117,000 1.3 117,000
Motor windings 1 145,500 1 145,500 1 145,500
Motor rotor 1 234,000 1 234,000 1 234,000
Motor starter 1.2 300,000 1.2 300,000 1.2 300,000

express the reduction of the machine MTTF for different working flow rates
compared to the BEP flow rate:

MTTFQ
lp ¼ ð3:9Þ
MTTFQBEP

The plot of Fig. 3.9 is obtained. In the same plot the pump characteristic curves
are reported together with the main loads affecting the reliability of the single pump
components.
It is clear now that the inherent reliability can be obtained only from the man-
ufacturer of the equipment, through handbooks or websites, or can be scaled from
operating experience. PAT applications are still too limited to allow a similar
analysis and, in particular, we are not completely aware of the load acting on the
individual pump component when working in inverse mode. Therefore, some
3.3 PAT Reliability 69

Fig. 3.8 Load, L, variation during the period, T

Fig. 3.9 Pump reliability and characteristic curves

degree of speculation will be necessary and the machine behavior will be deter-
mined for the whole PAT and not for the individual elements. In addition, the effect
of the installation grade on the reliability is at the moment completely unexplored.
To understand the shape of the reliability curve of a PAT, it is necessary to
compare the features of a pump working in direct or in inverse mode (Fig. 2.8). In
both cases, the BEP flow rate represents the best hydrodynamic condition for the
pump geometry and the shock losses at the BEP can be considered negligible. As
already mentioned in Chap. 2, in a machine working in direct mode the entity of
shock losses is higher for flow rates higher than the BEP than for flow rates lower
70 3 Industrial Aspects of PAT Design Improvement

Fig. 3.10 Comparison of


pump reliability in direct and
inverse mode

than the BEP. In a pump working in reverse mode there is the opposite result,
namely that the entity of shock losses is higher for flow rates lower than the BEP
than for flow rates higher than the BEP.
Shock losses induce machine vibrations and flow recirculations, reducing or
increasing the load on the bearings and seals, and reducing the PAT reliability for
flow rates lower than the BEP. In addition, when the flow rate approaches the
runaway condition the machine reaches an escape velocity with a large load on the
pump bearings and a further reduction in the reliability. Due to the presence of a
higher pressure of flow in inverse mode than in direct mode a larger load on the
machine seal will be present at high flow rates.
All these phenomena, according to the manufacturers, lead to a modification of
the pump reliability curve, as reported in Fig. 3.10.

3.4 PAT Life Cycle

The technical design of the machine to be used in a micro or pico hydro power
system is only the final, but nonetheless important, aspect in the decision process.
The installation of a PAT is in general determined on the basis of a comprehensive
analysis of the benefits of the system according to the energy recovery policy. The
reason for such a complex decisional approach for such a small investment
probably results from a number of different positive and negative considerations:
• The small direct income of a single plant in terms of the small installed power;
• The increase of the technical responsibility in the presence of a relatively
unexploited technology;
• The difficulty in the quantification of the additional benefit of the energy
recovery determined by the feed-in tariff;
3.4 PAT Life Cycle 71

• The complexity of the administrative procedure for the permission to undertake


the energy recovery;
• The benefit to the company image coming from the use of renewables;
• The possibility of accessing funding and credit for power plant design and
realization;
• The primary objective of water management consisting in water distribution.
To cover all these aspects, the decision process in power plant design involves a
far greater number of issues than in the case of pumping system design, and the
complete life cycle of the PAT is considered.
Bearing in mind that turbine technology is long established, the choice of tra-
ditional machines should always be preferred. Unfortunately, the smaller is the
power plant size and the larger are the variations of the turbine operating conditions,
the less is the economic viability of power plant based on traditional Francis,
Kaplan or Pelton turbines. When PAT becomes the only viable technology it is
obvious that some of the issues addressed become academic, as in the case of the
question of PAT and power plant reliability due to the lack of a large number of
on-site experiences. Moreover, the number of existing applications shows that, in
the presence of an appropriate technical design, power plants based on the PAT
technology are extremely reliable.
On the other hand, the use of commercial pumps simplifies the analysis in terms
of the management costs, because each individual repair operation is equal for a
pump or for a PAT and can be easily required of the pump manufacturer. The same
consideration can be made in terms of the environmental impact of the machine, at
the level of a single plant or at a global scale: these life cycle indicators can be
easily found in literature [11].
The effect of the feed-in tariff on the economic viability of the hydro power plant
varies considerably from one country to another. This is a subject in continuous
evolution but an accurate and recent analysis can be found in [12].
A number of recent papers have provided an accurate assessment of the life cycle
environmental balance of micro-hydropower energy recovery in the water industry
[13, 14]. These balances can be of great interest for policy makers and demonstrate
that micro and pico hydro power installation in the water industry has a decidedly
positive environmental impact. As an example of these analyses, in Fig. 3.11, the
normalized life cycle environmental burdens of micro hydropower are compared
with the marginal grid electricity generation impact coming from 300 MW gas
combined cycle plant, for the impact categories reported in Table 3.5.
72 3 Industrial Aspects of PAT Design Improvement

Fig. 3.11 Normalized impact category contributions for micro hydro power installations [13]

Table 3.5 Impact categories AP Acidification potential


for environmental impact
ARDP Abiotic resource depletion
assessment
FRDP Fossil resource depletion potential
GWP Global warming potential
HTP Human toxicity potential

References

1. S.V. Jain, R.N. Patel, Investigations on pump running in turbine mode: a review of the
state-of-the-art. Renew. Sustain. Energy Rev. 30, 841–868 (2014)
2. S. Derakhshan, B. Mohammadi, A. Nourbakhsh, Efficiency improvement of centrifugal
reverse pumps. J. Fluids Eng. 131(021103), 1–9 (2009)
3. S.S. Yang, F.Y. Kong, M.W. Jiang, X.Y. Qu, Effects of impeller trimming influencing pump
asturbine. Comput. Fluids, 67, 72–78 (2012)
4. S. Derakhshan, Efficiency improvement of centrifugal reverse pump (Ph.D. thesis). University
of Tehran (2008)
5. P Singh, Optimization of internal hydraulics and of system design for pumps as turbines with
field implementation and evaluation (Ph.D. thesis). University of Karlsruhe (2005)
6. Y. Tung, B. Yen, C. Melching, Hydrosystems Engineering Reliability Assessment and Risk
Analysis (McGraw-Hill Professional, New York, NY, USA, 2006)
7. K. Kapur, L. Lamberson, Reliability in Engineering Design, vol. 1 (Wiley, New York, NY,
USA, 1977)
8. K. Trivedi, Probability and Statistics with Reliability, Queuing, and Computer Science
Applications (Wiley, New York, NY, USA, 2002)
9. H.P. Barringer, Life cycle cost and good practices, in Proceedings of the NPRA Maintenance
Conference, San Antonio, TX, USA, 19–22 May 1998 (Barringer & Associates, Inc.:
Humble, TX, USA 1998)
References 73

10. H.P. Barringer, A life cycle cost summary, in Proceedings of the International Conference of
Maintenance Societies, Perth, Australia, 20–23 May 2003
11. Work on Preparatory studies for implementing measures of the Ecodesign Directive
2009/125/EC, ENER Lot 29 Working document European Commission, DG ENER 22
January 2014
12. O. Fecarotta, C. Aricò, A. Carravetta, R. Martino, H.M. Ramos, Hydropower potential in
water distribution networks: pressure control by PATs. Water Resour. Manag. 29(3) 699–714
(Feb, 2014)
13. J. Gallagher D. Styles, A.P. Williams, A. McNabola, Quantifying the environmental impact of
micro-hydropower in the water industry using life cycle assessment, avnIR conference, life
cycle in practice, Lille, France (2014)
14. A. McNabola, P. Coughlan, L. Corcoran, C. Power, A.P. Williams, I. Harris, J. Gallagher, D.
Styles, Energy recovery in the water industry using micro-hydropower: an opportunity to
improve sustainability. Water Policy 16(1) 168–183 (2014)
Part II
PAT Based Micro/Pico Hydropower
Stations
Chapter 4
PAT Selection

Abstract A simple methodology for PAT selection starting from the performance
of the pump in direct mode is given, based on a number of studies on PAT behavior
present in expert literature. The expected errors in terms of head and PAT efficiency
between the results of this methodology and the experimental data are also dis-
cussed, and the possibility of introducing correction factors based on the internal
pump geometry is considered. The reader will be introduced to the design proce-
dure of the PAT in small hydropower plants. In particular, simple methods for the
selection of the pump to be used in inverse mode and on the modification of the
PAT impeller geometry will be presented, together with the speed necessary to
obtain the match between the required design parameters and the PAT working
conditions.

Pump characteristic curves in reverse operation are not usually available on the
market or in manufacturers’ catalogues. Therefore, establishing a correlation
enabling the transition from the “pump” characteristics to the “turbine” character-
istics is the main challenge in using a pump as a turbine. Many researchers have
presented certain theoretical and empirical relations for predicting the PAT char-
acteristics at the best efficiency point (BEP). However, the results predicted by these
methods are not reliable for all pumps with different specific speeds and capacities.
In this chapter, the available methods of selecting a proper PAT to produce energy
will be presented.

4.1 Theoretical and Practical Methods

Various researchers have theoretically and practically presented relations to select a


proper PAT for known hydraulic data. Figures 4.1 and 4.2 show some data reported
for the centrifugal PAT head ratio (h) and discharge ratio (q) based on the pump
head and discharge respectively [1, 2]:

© Springer International Publishing AG 2018 77


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_4
78 4 PAT Selection

Fig. 4.1 Head ratios of the


tested PATs with various
pump specific speeds [1]

Fig. 4.2 Discharge ratios of


the tested PATs with various
pump specific speeds [1]

HT QT
h¼ ; q¼ ð4:1Þ
Hp Qp

In these figures, the h and q of the PATs are shown in relation to the pump
specific speed and maximum efficiency. The definition of the specific speed of the
pump is:

Q0:5
p
Nsp ¼ Np ð4:2Þ
Hp0:75

It can be observed that two pumps with the same specific speeds could have a
different h and q.
For Nsp < 50, a higher specific speed leads to a higher efficiency for the fixed
head drop. However, for Nsp > 50, it is the reverse. From Figs. 4.1 and 4.2, it is
clear that the lowest h and q are related to Nsp > 50.
A recent study developed by Novara et al. [3, 4] shows a set of PATs where the
HBEP and QBEP have been plotted on the logarithmic H-Q space as in Fig. 4.3,
highlighting with different colors the maximum efficiency of each machine for the
radial PAT range presented by Chapallaz [5].
4.1 Theoretical and Practical Methods 79

Fig. 4.3 PAT maximum efficiency ranges against the H-Q space [3, 4]

References [3, 4] have developed analyses based on several PATs and proposed
a three dimensional representation of PAT efficiency (where Ns corresponds to Nsp).
These selected machine PAT plus generator units are shown in Fig. 4.4, while a
two dimensional contour chart similar to the one proposed for pumps by [5] is
presented in Fig. 4.5.
According to Fig. 4.5, PATs with higher capacity and medium specific speeds of
50 have a better efficiency: to obtain the maximum efficiency in a pump working
in reverse mode, devices with lower rotational speeds and higher capacities are
recommended for specific speeds less than 50.
As mentioned in Chap. 3, based on PAT theory [6–11], pumps with a higher
efficiency work in lower h and q in turbine modes. From Figs. 4.6 and 4.7, the
following relations can be found:

1:2
h¼ ð4:3Þ
g1:1
p

1:2
q¼ ð4:4Þ
g0:55
p

From Fig. 4.8, the pump specific speed can be selected using the turbine specific
speed with the following estimated relation:

Nsp ¼ 1:125Nst þ 1:73 ð4:5Þ


80 4 PAT Selection

Fig. 4.4 Proposed 3D representation for PAT efficiency estimation against the real efficiency of
selected machines [3, 4]

where:

Q0:5
Nst ¼ Nt t
ð4:6Þ
Ht0:75

It is useful to mention that a pump in inverse mode operates at its BEP always
with a flow rate and head drop larger than in direct mode. Therefore, energy
4.1 Theoretical and Practical Methods 81

Fig. 4.5 2D contours of the proposed function for PAT efficiency plotted against the nominal
flow rate and specific speed of selected machines [3, 4]

Fig. 4.6 Head ratios of tested


PATs with various pump
maximum efficiencies [1]

Fig. 4.7 Discharge ratios of


tested PATs with various
pump maximum
efficiencies [1]
82 4 PAT Selection

Fig. 4.8 Turbine specific speed versus pump specific speed [1]

Table 4.1 Review of methods to determine non-dimensional head and flow parameters (adapted
from [16])
Method h q
Stepanoff 1 1
gt;BEP gp;BEP gp;BEP

Gopalakrishnan 1 1
ðgp;BEP Þ
2 gp;BEP

Childs 1 1
ðgp;BEP Þ ðgp;BEP Þ
2 2

Sharrna 1 1
ðgp;BEP Þ ðgp;BEP Þ
1:2 0:8

Alatorre-Frenk 1 0:85g5p;BEP þ 0:385


0:85g5p;BEP þ 0:385
2g9;5
p;BEP þ 0:205

Nautiyal [41] gp;BEP 0:212 g 0:212


41:667 ln N  5:042 30:303 lnp;BEP  3:424
ð s;q;pump Þ ðNs;q;pump Þ
Grover 2:693  0:229Ns;q;turbine 2:379  0:0264Ns;q;turbine

dissipations connected with the flux of water in the machinery are also greater in
reverse mode.
There are various methods available in literature to select h and q, based on
theoretical and practical issues presented by certain researchers. However, none of
the methods presented can predict the PAT behavior perfectly. The most commonly
used methods to obtain values of h and q are summarized in Table 4.1.
In Figs. 4.9 and 4.10, the results of the design methods of Eqs. (4.3) and (4.4)
are compared with the results of the methods presented by Stepanoff and Sharma.
4.2 Estimation of Characteristic Curves 83

Fig. 4.9 h of the tested PATs with various pump maximum efficiencies [1]

Fig. 4.10 q of the tested PATs with various pump maximum efficiencies [1]

4.2 Estimation of Characteristic Curves

Although a PAT may work at off-design conditions, most prediction methods have
only predicted the BEP of the PAT. Therefore, estimating the complete charac-
teristic curve of a PAT based on its BEP would be very expedient. Experimental
data have shown that the dimensionless characteristic curves of centrifugal PATs
based on their BEP are approximately the same. The above mentioned dimen-
sionless head and power curves of a PAT can be estimated as below, using second
and third order polynomials, respectively [12–15]:
 2  
Ht Qt Qt
¼ 1:03 0:55 þ 0:53 ð4:7Þ
Htb Qtb Qtb
 3  2  
Pt Qt Qt Qt
¼ 0:31 þ 2:15 0:89 þ 0:05 ð4:8Þ
Ptb Qtb Qtb Qtb
84 4 PAT Selection

The efficiency curve can be obtained for each point by:

Pt
gt ¼ ð4:9Þ
ðq:g:Qt :Ht Þ

However, it must be noted that this method can only provide an approximate
view of the characteristic curves of a PAT.
In a recent study [2] a large data set relative to the behavior of 17 different
pumps operating as turbines has been analyzed, as shown in Table 4.2. The fol-
lowing considerations were possible by means of the comparison of the classic
affinity law with experimental data:
– the agreement between the experimental and the calculated curves is worse
when the difference between the rotational speed of the prototype and that of the
simulated machine increases;
– the entity of the discrepancies is not dependent on the machine type;
– when the characteristic curves are calculated by means of the affinity law and
Suter parameters with a 20% difference in rotational speed compared to the
prototype, the error in the evaluation of the head drop is lower than 3%;
– out of the above mentioned range of differences in rotational speed compared to
the prototype, the error in the evaluation of the head drop could be as large as
12%, and the mean error in the whole range of rotational speed is 4.8%;
– the error in the evaluation of η is less than 15%, if the difference in rotational
velocity with the prototype ranges between −40% and + 50%;

Table 4.2 Machine data set PAT Pump type No. of stages Speed range (rpm)
[4]
a HCM 2 1550–3050
b HCM 2 1550–3050
c HCS 1 1550–3050
d HCS 1 1550–3050
e HCS 1 1550–3050
f SSS 1 780–1550
g SSS 1 1050–1550
h SSS 1 780–1550
i SSS 1 780–1860
j SSS 1 780–1860
k SSS 1 780–1550
1 SSS 1 780–1550
m SSS 1 750–1550
n HCM 3 1550–3050
o HCM 4 1550–3050
(HCM Horizontal Centrifugal Multi-stage, HCS Horizontal
Centrifugal Single-stage, SSS Submersible Semiaxial
Single-stage)
4.2 Estimation of Characteristic Curves 85

– out of the above mentioned range of differences in rotational speed compared to


the prototype, the error in the evaluation of η could be even larger than 40%, and
the mean error in the whole range of rotational speed is 7.1%.
To overcome these problems, a new model (Relaxation of the Affinity Equations—
RAE) for the estimation of the performance of semi-axial PATs has recently been
proposed, based on the following experimental evidence:
– the efficiency at the BEP attains its maximum value gMAX
B for a specific rota-
tional speed Nmax, and Nmax depends on some geometrical parameters of the
machine;
– the position of the BEP at a certain value of speed (N) depends on the ratio
(N/Nmax);
– generalized dimensionless performance curves h = h(q), p = p(q) and e = e(q)
can be defined for a given pump type, based on experimental results (instead of
Suter parameters), where:

Q H P g
q¼ h¼ p¼ e¼ ð4:10Þ
QB HB PB gB

In Fig. 4.11 the geometrical parameters of a semi-axial PAT useful for the
identification of Nmax are shown: D is the diameter of the runner, u the diameter of
the PAT body and F the length of the PAT stage. The following relation was found
to be representative of the experimental results:

Nmax ¼ aDb uc F d ð4:11Þ

The new generalized characteristic and efficiency curves determined by a cali-


bration with the experimental data set of the Submersible Semi-axial Single-stage
pumps of Table 4.2 are plotted in Fig. 4.11. By the use of RAE, the error in the
evaluation of the head drop has been reduced to 3.9%, and the error in the

Fig. 4.11 Geometrical


parameters of a semi-axial
PAT [17]
86 4 PAT Selection

Fig. 4.12 Generalized characteristic and efficiency curves by RAE [17]

evaluation of η has been reduced to 2.5%, compared with a 4.8 and 7.1% error,
respectively, relating to the classic affinity law (Fig. 4.12).
Even if no such an in-depth analysis has been performed for other pump types,
some recommendation can be derived in the use of the affinity law for all pump
types and in the design of mini and micro hydro power plants using industrial
PATs. The potential discrepancies in terms of pressure drop and efficiency between
the real PAT behavior and the theoretical prediction are small for small differences
in rotational speed compared to the test conditions (± 20%). Even in the presence
of large differences in the rotational speed the error in terms of pressure drop is
contained, but the difference in terms of efficiency could become huge. As a
consequence, in the absence of any experimental data coming from industrial tests,
the main problem in power plant design could be an overestimation of the PAT
production.
4.3 Rotational Speed Selection 87

4.3 Rotational Speed Selection

Centrifugal pumps normally operate at fixed rotational speeds, approximately at


1450 rpm or 2900 rpm, driven by one or two pairs of pole induction electro motors,
respectively. In reverse, a PAT’s rotational speed completely depends on the
available hydraulic potential of the power plant. Therefore, the rotational speed is
not input data for the selection of a proper PAT, especially in off-grid conditions.
Due to price restrictions, only 2 and 4 pole generators are available on the market.
Therefore, rotating the PAT at 1500 rpm or 3000 rpm for a frequency of 50 Hz and
at 1800 rpm or 3600 rpm for a frequency of 60 Hz is a priority for the user by
direct coupling to the generator. However, to reach these speeds, a sufficient head
and flow rate of PAT are required. It has been proven by experience that
medium-head PATs normally rotate at speeds lower than 1500 rpm. In this situa-
tion, transferring the power from the PAT to the generator can be achieved by
indirect coupling using a belt and pulleys. In this condition, the PAT arbitrary speed
can be converted to a generator nominal speed by selecting a suitable pulley ratio.
Figure 4.13 compares the direct and indirect power transmission in a PAT.
On the other hand, pump manufacturers normally present characteristic curves at
the two mentioned speeds. Therefore, a correlation is required to estimate a pump
characteristic curve, or at least its BEP at a desired rotational speed. Using the
affinity law in turbomachinery, Eqs. (4.12) and (4.13) can be used in the following
forms:
 
1:2 NPAT 2
h¼ ð4:12Þ
g1:1
p NP
 
1:2 NPAT
q¼ ð4:13Þ
g0:55
p NP

Therefore, the pulley ratio will be the ratio of the PAT rotational speed to the
generator rotational speed: NNPAT
G
.

Fig. 4.13 Direct and indirect power transmission methods for a PAT
88 4 PAT Selection

4.4 Impeller Diameter Selection

One of the other methods to adjust a pump working as a proper PAT according to
the site data, is trimming of the impeller diameter. Pump manufacturers usually use
this method to adjust their pumps’ characteristic curves and rated point according to
client data: these curves are normally provided by experimental measurements in
standard test rigs. For a PAT, establishing the same characteristic curves is costly
and in general only a small number of manufacturers have these design data already
available, and, even then, not for all pump families and types.
Yang et al. [16] have carried out experimental and numerical research on a
centrifugal pump used as a turbine, in order to study the effect of impeller trimming
and inlet angle on its performance. The results suggested that, as the impeller
diameter reduces from 255 mm to 215 mm, the PAT efficiency decreases by 4.11%.
As a result, due to the decrease in efficiency by diameter trimming, it is not
recommended that a PAT’s head and flow rate be adjusted by diameter trimming
rather than by rotational speed changing, whenever the PAT works in constant
hydraulic conditions.
A benefit of reducing the impeller diameter can be found in some types of
machine where impeller trimming changes significantly the shape of the Q-H curve,
as in the case of submersible semiaxial pumps. The steeper Q-H curve for the
trimmed impeller can be used to achieve working conditions characterized by
higher available head drops than those achieved by using a full size impeller pump.

4.5 PATs in Parallel

Parallel systems are usually used to optimize the management of the system. Pump
stations consisting of parallel pumps are typical examples in this respect. By par-
alleling multiple pumps, we can manage our flow rate and pressure requirement for
an optimal energy consumption. Hydraulic turbines are usually arranged in a par-
allel position to optimize electricity production depending on annual and seasonal
hydro-potential (flow rates and heads). Although a parallel arrangement is the most
popular in large and medium hydro power plants, sometimes small hydro potentials,
including mini and micro plants, are divided between parallel turbines for economic
or technical issues. Figure 4.14 shows a typical parallel arrangement of a small
hydropower plant.
A PAT, though cheaper than a conventional turbine, has a narrow efficiency
curve. Therefore, it is suitable for a constant flow rate and pressure. However, most
available hydro sites, including natural water and industrial water sectors, are not
sufficiently permanent for the replacement of a conventional turbine with a PAT. In
this situation, using multiple PATs in a parallel arrangement is very effective.
Figure 4.15 shows the typical arrangement of three PATs in parallel controlled by
valves. To select the proper PAT for a hydro site, one, two, three or more PATs
4.5 PATs in Parallel 89

Fig. 4.14 Parallel arrangement of a PAT mini hydropower plants

Fig. 4.15 The typical arrangement of three PATs in parallel controlled by valves
90 4 PAT Selection

1900 150
1800 140
1700 Site Head 130
1600
1500 Power- 1 PAT 120
1400 110
1300 100
Power- 3 PATs
1200

Efficiency (%)- H (m)


Efficiency- 3 PATs 90
P (kW)

1100
1000 80
900 70
800
60
700
Efficiency- 1 PAT 50
600
500 40
400 30
300
20
200
100 10
0 0
0 100 200 300 400 500 600 700 800 900 1000110012001300140015001600

Q (lit/sec)

Fig. 4.16 Output of a hydro site using three small PATs in parallel and one large PAT

should be considered, according to a feasibility study report. Figure 4.16 compares


the output of hydro sites consisting of one large PAT and three small PATs,
respectively. It is clear that using a parallel system, a site’s efficiency increases from
the minimum flow to the maximum flow rate.

4.6 PAT-Pump Direct Coupling

In the sixties Worthington created a turbo-pump to be used on ships, a CT model of


which is shown in Fig. 4.17. The idea was to use this device in submerged
applications, as in the case of bilge water, or for ballast water pumping. By the
direct coupling of a turbine runner and a pump impeller, this turbo-pump worked
using the hydraulic energy of the low flow-high pressure jet of the anti-fire boosters,
to transfer large water volumes with a smaller pumping head. Nowadays, this
principle can be used also in a WSN.
In Chap. 9, a real application using a PAT-pump direct coupling (P&P) is fully
described. The greatest advantage of such a system is the very low cost, limited to
the mechanical equipment, and the great efficiency, due to the absence of any motor
4.6 PAT-Pump Direct Coupling 91

Fig. 4.17 The Worthington


pump, CT model

or generator. Despite the simplicity of the idea, the design of a PAT-pump system
with direct coupling is not straightforward.
In the design of a P&P turbocharger the group is free to achieve any rotational
speed. The PAT provides the power for the pump and the rotational speed is set by
the combination of the performance curves of the two devices with the network
characteristics.
A simplified scheme of a WSN working in combination with a P&P plant is
shown in Fig. 4.18. The whole water supply system can be considered as two
separated network districts, which are connected by the P&P plant.
The residual head at the end of district 1 that can be turbined within the PAT is
represented by the difference HT = Hu1−Hd1−DHr1, where Hu1 is the head measured at
the end point of district 1, DHr1 is the head loss in the pipeline approaching the PAT
and Hd1 is the head downstream of the PAT. The lengths of the pipelines are not
representative of the real system and, if the values of Hu1 and Hd2 are measured near
the P&P system, the head losses DHr1 and DHr2 can be neglected. As a general case,
the two values Hd1 and Hu2 are considered different, but the outlet tank of the PAT
and the inlet tank of the pump are often the same. The presence of four tanks,
however, could even be unnecessary if the P&P system is inserted in a fully
pressurized network, and the values of the four variables, Hu1, Hd1, Hu2 and Hd2 depend
on time according to the network behaviour. QT is the flow rate available at the
PAT inlet.
The power QT the PAT is cQTHTηT, ηT being the PAT efficiency. Such power is
transmitted to the pump by the shaft which connects the two machines. QP is the
pumped flow rate while the total head required at the pump outlet is
HP = Hd2−Hu2 + DHr2, the meaning of the symbols being shown in Fig. 4.18. The
efficiency of the plant can be calculated as the ratio between the output hydraulic
power at pump side and the input hydraulic power at PAT side. Thus, the plant
efficiency η can be calculated as:
92 4 PAT Selection

Fig. 4.18 Use of a P&P in a water supply network [15]

g ¼ gT  gP ð4:14Þ

where gP is the pump efficiency and gT is the PAT efficiency.


The equations regulating the functioning of the P&P, written in dimensionless
form, are the following [14]:

HP ¼ H2u  H2d  DH2r ð4:15Þ


 2   !
HT QT QT
¼ aT þ bT þ c T nt ð4:16Þ
N2 N N

 3  2 !
PT QT QT QT
¼ aT þ bT þT þ dT nt ð4:17Þ
N3 N N N

 2   !
HP QP QP
¼ aP þ bP þ c P np ð4:18Þ
N2 N N

 3  2 !
PP QP QP QP
¼ aP þ bP þP þ dP np ð4:19Þ
N3 N N N

PT ¼ PP ð4:20Þ

In the seven variables: HT ; HP ; QT ; QP ; PT ; PP and N. The variables represent


respectively the head, the flow rate and the power of the PAT and pump. In
Eqs. 4.16–4.19 a cubic regression has been considered to fit the single stage
4.6 PAT-Pump Direct Coupling 93

experimental pump and PAT characteristic curves nt and np are the number of
stages of the PAT and pump.
Equations 4.16–4.18 can be rewritten in dimensionless form by assuming
hT = HT/N2 and hP = HP/N2:
 
hT ¼ aT q2T þ bT qT þ cT nT ð4:21Þ
 
hP ¼ aP q2P þ bP qP þ cP nP ð4:22Þ

In accordance with the conditions of Eq. (4.20), expressing that the axial power
is the same for the pump and the PAT, by assuming qT = QT/N and qP = QP/N,
Eqs. (4.17) and (4.19) can be rewritten as:
   
aT q3T þ bT q2T þ cT qT þ dT nT ¼ aP q3P þ bP q2P þ cP qP þ dP nP ð4:23Þ

Equations (4.21) to (4.23) link the four variables hT, qT, hP and qP. Considering
that qP/qT = QP/QT and that hP/hT = HP/HT, the relationship between the ratio of
the delivered discharge and the ratio of the delivered head is independent of the
rotational speed. Obviously, this relationship depends on the characteristic curves
and on the number of stages of the two hydraulic machines used in direct and
inverse mode in the P&P plant.
Some simulations have been performed on two Caprari machines, an HMU
model used as the pump and an NC80 model used as the PAT. The characteristic
curves were determined experimentally in the University of Naples Federico II and
by the manufacturer [14]. In Figs. 4.19 and 4.20 the head ratio and the efficiency of
the P&P plant for different numbers of pump stages and with nt = 1 are plotted,
respectively. It is evident from the figures that an appreciable P&P efficiency can be
obtained with a pumping flow rate varying between 10 and 30% of the turbine flow
rate, corresponding to a pumping head as high as three times the turbine head drop.
For the discharge pattern of a large urban district, presenting a peak coefficient of
about 1.2 and a night coefficient of about 0.8, and with reference to a number of
WSN scenarios of Table 4.2, the energy recovery of the P&P has been evaluated
and reported in Table 4.3. Case a, and Case b differ in terms of the turbine flow rate.
Supply conditions 1 and 2 differ in terms of the pumping conditions, presenting a
constant pumping head and a constant pumping flow rate, respectively. The P&P
efficiency is obviously low in all cases, ranging between 35 and 44%, but the annual
energy saving is so high as to justify the substitution of an existing pumping system
with this kind of unconventional power plant.
94 4 PAT Selection

Fig. 4.19 Head ratio of the


P&P plant for different
numbers of pump stages [15]

Fig. 4.20 Efficiency of the


P&P plant for different
numbers of pump stages [15]
References 95

Table 4.3 WSN scenarios for a P&P design [15]


Case a Case b
Supply condition 1  T ¼ 40L=sÞ
Variable QT ðQ  T ¼ 25 L=sÞ
Variable QT ðQ
Variable HT Variable HT
Variable QP Variable QP
HP ¼ 50 m HP ¼ 15 m
nP ¼ 4 nP ¼ 2
Supply condition 2  T ¼ 40L=sÞ
Variable QT ðQ  T ¼ 25L=sÞ
Variable QT ðQ
Variable HT Variable HT
QP ¼ 4; 4L=s QP ¼ 6; 4L=s
Variable HP Variable HP
nP ¼ 4 nP ¼ 2
Scenario Turbined Turbined Pumped Pumped Average Annual
average maximum average maximum efficiency energy
power power power power saving
(kW) (kW) (kW) (kW) (MWh)
1a 8.95 20.21 3.52 6.93 0.39 48.1–
77.0
1b 2.27 4.87 0.93 1.76 0.41 12.7–
20.3
2a 8.8 19.70 3.10 5.88 0.35 42.5–
68.0
2b 2.28 5.26 1.01 2.16 0.44 13.8–
22.0

References

1. S.-S. Yang, S. Derakhshan, F.-Y. Kong, Theoretical, numerical and experimental prediction
of pump as turbine performance. Renew. Energy 48, 507–513 (2012)
2. D. Novara, S. Derakhshan, A. McNabola, H. Ramos, Estimation of unit cost and maximum
efficiency for Pumps as Turbines, in Young Water Professionals (2017)
3. D. Novara, A. Carravetta, S. Derakhshan, A. McNabola, H. Ramos, A cost model for pumps
as turbines and a comparison of design strategies for their use as energy recovery devices in
Water Supply Systems, in EEMODS’17, Rome (eemods17.org), Energy Efficiency in Motor
Driven Systems (2017)
4. D. Novara, A. Carravetta, S. Derakhshan, A.M. Nabola, H. Ramos, Centrifugal pumps as
turbines cost determination and feasibility study for pressure reducing valve substitution in a
water supply system, Submitt. to Renew. Energy (2017)
5. J.M. Chapallaz, P. Eichenberger, G. Fischer, in A Manual on Pumps Used as Turbines
(Vieweg, 1992)
6. S. Derakhshan, A. Nourbakhsh, Experimental study of characteristic curves of centrifugal
pumps working as turbines in different specific speeds. Exp. Therm. Fluid Sci. 32(3), 800–807
(2008)
7. H. Ramos, A. Borga, Pumps as turbines: an unconventional solution to energy production.
Urban Water 1(3), 261–263 (1999)
8. H. Ramos, A. Borga, M. Simão, New design solutions for low-power energy production in
water pipe systems. Water Sci. Eng. 2(4), 69–84 (2009)
9. H. Ramos, D. Covas, L. Araujo, M. Mello, Available energy assessment in water supply
systems, in XXXI IAHR Congress (2005), pp. 11–16
96 4 PAT Selection

10. H. Ramos, M. Mello, P.K. De, Clean power in water supply systems as a sustainable solution:
from planning to practical implementation. Water Sci. Technol. Water Supply 10(1), 39–49
(2010)
11. P. Singh, Optimization of internal hydraulics and of system design for pumps as turbines with
field implementation and evaluation, Universität Karlsruhe (TH) Diss.
InstitutfürWasserwirtschaft und Kulturtechnik (2005)
12. A. Carravetta, G. del Giudice, O. Fecarotta, H. Ramos, PAT design strategy for energy
recovery in water distribution networks by electrical regulation. Energies 6(1), 411–424
(2013)
13. A. Carravetta, O. Fecarotta, R. Martino, L. Antipodi, PAT efficiency variation with design
parameters. Procedia Eng. 70, 285–291 (2014)
14. H. Nautiyal, V. Varun, A. Kumar, Experimental investigation of centrifugal pump working as
turbine for small hydropower systems. Energy Sci. Technol. 1(1), 79–86 (2011)
15. A. Carravetta, L. Antipodi, U. Golia, O. Fecarotta, Energy saving in a water supply network
by coupling a pump and a pump as turbine (PAT) in a turbopump. Water 9(1), 62 (2017)
16. S.-S. Yang, F.-Y. Kong, W.-M. Jiang, X.-Y. Qu, Effects of impeller trimming influencing
pump as turbine. Comput. Fluids 67, 72–78 (2012)
17. O. Fecarotta, A. Carravetta, H. Ramos, R. Martino, An improved affinity model to enhance
variable operating strategy for pumps used as turbines. J. Hydraul. Res. 54(3), 332–341
(2016)
Chapter 5
PAT Control Systems

Abstract The main rules for the selection of the regulation mode of a PAT based
on the operating conditions are presented, as measured at the point of the network
where the power plant will be located. The variability of operating conditions in
water supply networks is discussed and the influence of the daily demand variation
of the flow rate and available head drop is established, with the main regulation
modes introduced. Finally, the concept of power plant effectiveness is described
and the Variable Operating Strategy (VOS) is proposed as the best design proce-
dure. The reader will be introduced to the importance of PAT regulation in a small
hydropower plant, and to the main schemes of PAT installation. In particular,
he/she will be instructed on to how to perform an advanced power plant design
based on the local operating conditions in order to maximize the power plant
effectiveness.

The research into small PATs is much more recent than that into large pumps
working in reverse mode. The latter were installed in the last century in big pump
and storage power plants used to face peaks in energy use. In the last ten years
PATs have been recognized as the only viable solution for energy recovery in
WSNs. This is primarily due to two particular characteristics of small PATs: the
low cost of these devices, and their great reliability. Despite the lower efficiency of
PATs [1] when compared with classic turbines the former are preferred in WSNs in
consideration of their economic convenience and of the great importance of
reducing the costs of WSN maintenance.
In a classic turbine the high level of efficiency is determined by two factors: the
limited variation of the working point, and the presence of a mechanical regulation.
As the turbine size decreases the cost of the mechanical regulation increases. At the
same time in a WSN the flow rate and available head might vary in a wide range of
operating conditions, out of the range of the turbine working conditions. Therefore,
whenever a turbine becomes a not viable solution, a PAT could be a valid alter-
native. Other energy production devices exist or have been designed for pico and
micro hydropower plants, but the costs are still high because the production is not at
an industrial scale and the reliability is not proven for the limited application.

© Springer International Publishing AG 2018 97


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_5
98 5 PAT Control Systems

In this chapter, the operating conditions of PATs in a WSN are analyzed. Then,
the possible PAT regulation modes and working conditions for each mode are
described. Finally, a methodology for power plant design is given and the opti-
mization parameters are described.

5.1 Operating Conditions in a WSN

A WSN is a pipeline or channel system used to transfer water from the water source
to the end users. Pressure pipes are used more frequently than channels in a WSN
for potable water, the latter employed only in part of the network for water trans-
mission, Fig. 5.1. Pressure pipes are also preferred for irrigation in arid or semi-arid
regions.
Storage tanks, regulation valves, and pumping stations, are all parts of the
network. All these components are necessary to grant the flow of water from the
spring to the end users, while maintaining the pressure in the pipelines within an
optimal range. An excess of pressure increases the stress on the pipeline material
and has to be avoided in an urban network. An excess of pressure reduction is also
negative in view of the potential contamination of water due to fluid external to the
pipe.
Unfortunately, due to pipeline friction and ground level variations, in some
points of the network energy has to be supplied to the water by pumping stations,
leading to a higher cost of water transfer. In other points, equally expensively, the
excess of energy has to be dissipated in pressure reducing valves (PRV).
Hydropower plants can be installed to substitute PRVs and a large part of the
energy dissipated by the PRV can be recovered leading to a reduction in the costs
connected to water transfer.

Fig. 5.1 A water supply network (WSN) [2]


5.1 Operating Conditions in a WSN 99

A main difference between water transmission and water distribution is repre-


sented by the daily variation of the flow rate. Water transmission is used to transfer
water from the water source to storage tanks. The flow rate is usually constant
during the day and only a seasonal variability is present. This seasonal variability is
generally limited for drinking water, with the exception of that in touristic areas. On
the contrary, a large variation of water demand is present in irrigation, the use of the
network being limited to the summer season.
In water distribution the daily variability of flow rate is large, with a large water
demand in daylight hours, and a reduction of water use during the night. The
smaller is the dimension of the urban area, the larger is the extent of this variation.
For irrigation and industrial use the daily variation is more limited or indeed
completely absent depending on local factors, like the presence of automation in
irrigation or a continuous 24 h production in industry.
As an effect of friction losses, any variability of flow rate reflects on a variability
of pressure head along the pipelines, and on a variability of available head in the
energy production node. Therefore, in water transmission, due to the presence of a
constant flow and drop, and to the larger available power, traditional turbines are
preferred. Instead, in water distribution, the variability of the discharge and drop is
large and the available power small.
As an example, data from the monitoring station (Amendola, Pompeii, Southern
Italy) of a 10,000 inhabitant WSN in southern Italy are analyzed. In Fig. 5.2 the
measured values of flow rate and pressure head are plotted. As expected, the flow
rate is larger during the day and smaller during the night, but a large base flow is
always observed, due to the presence of different water uses and of appreciable
leakage. On the contrary, the pressure head is large during the night and decreases
during the day as an effect of flow resistance. Another feature of the plots is the

Fig. 5.2 Flow rate and pressure head in a real network


100 5 PAT Control Systems

Fig. 5.3 Daily pattern of


hourly averaged
dimensionless discharge and
head [2]

random fluctuation of flow rate and pressure head. The effects of this random
component can be reduced by performing an average over several days of the flow
rate and pressure head values measured at the same instant of the day. Another
classic manipulation is given
 in Fig. 5.3, which
 shows the values of cQ, i.e. the ratio
between the hourly lQ and the daily l Q averaged discharge, and cH, i.e. the
ratio between the hourly ðHQ Þ and the daily ðH  Q Þ averaged upstream head, in
relation to the daily time.
The maximum value of cQ is called the peak demand coefficient, Cp. It is well
known in water network research that the peak demand coefficient reduces with an
increase in the number of inhabitants, Nin. Therefore, the cQ behavior is not
invariant with lQ and the following relation can be used to scale the daily pattern:

lQ ðtÞ ¼ lQ  ½1 þ cm ðcQ ðtÞ  1Þ ð5:1Þ

where cm is a coefficient which value is fitted to obtain the observed variation


coefficient of the daily pattern [3].
We can also introduce a minimum night demand coefficient, CN, representing the
minimum value of cQ. The entity of the night demand coefficient depends on
several factors, such as the age of the pipelines and the state of maintenance. In
small residential areas this value is in general smaller due to the better control of
water leakage.
The difference between the instantaneous discharge and the time averaged dis-
charge can be relevant and a correct hydropower design should take into account
such variation. The instantaneous discharge can be calculated, based on the daily
pattern of the hourly averaged discharge, by means of a recent probabilistic model
(Mixed Distribution) by Gargano et al. [4]. This distribution is obtained by merging
two cumulative distribution functions, taking into account the probability of both
null and not null discharge. The following formulations can be used for a Monte
Carlo generation:
5.1 Operating Conditions in a WSN 101

8  pffiffi 
>
> Qd
¼ l ð t Þ  1  3
 ln 1FðtÞ
 CV ð t Þ
>
> Q p FðtÞF0 ðtÞ
>
<  
F0 ðtÞ ¼ exp 5  1000 Nin
lQ ðtÞ ð5:2Þ
>
>
>
>
>
: CV ðtÞ ¼ 0:1 þ 6
3=4 5=4
ð0:25Nin Þ lQ ð t Þ

where F0(t) is the probability of a null discharge, Nin the number of inhabitants
served by the urban supply system, and F(t) the probability of the not null dis-
charge, randomly generated by the Monte Carlo method. According to [4], the
instantaneous discharge of the Amendola monitoring station has been calculated for
a 60 s sampling rate and the results are plotted in Fig. 5.4, where they are compared
with the experimental data sampled each 15 min.
The relation between the measured flow rate and pressure is determined by a
large number of factors, such as: the shape and extension of the upstream water
network, the pipe material and age, and the presence and entity of water demand in
the upstream part of the network, etc. In Fig. 5.5 measured pressure head values are
plotted versus measured flow rates. The inverse proportionality between the pres-
sure head and flow rate is clear, but there is a large dispersion of data even in the
presence of a small data rate. When dealing with instantaneous data, the kind of the
relation between the random flow rate fluctuation and random pressure head fluc-
tuation is still unexplored. Carravetta et al. [5] showed that the fluctuations of the
Amendola pressure head data were normally distributed around the hourly HQ value
of the pressure head.

Fig. 5.4 Comparison between the generated discharge with a 1 min sampling rate and the (exp)
measured discharge with a 15 min sampling rate
102 5 PAT Control Systems

Fig. 5.5 Measured pressure head versus discharge

The pressure drop in the energy production node, ΔH, is given by the difference
between the upstream pressure head and the required downstream pressure head.
Due to the large variation of flow rate values and the completely different form of
the Q(H) distribution when compared with a PAT characteristic curve, it is evident
that any constraint in terms of back pressure value could be obtained only by
introducing electric or hydraulic procedures for PAT regulation. Therefore, a pre-
liminary question to address in the design of a hydropower plant in a WSN relates
to the possible flexibility of these pressure constraints.
It is possible to discuss separately three cases:
1. PRV strict replacement
2. PRV weak replacement
3. Maximum energy production.
In case 1 the hydropower plant is required to act strictly as an existing PRV. In
general, this case corresponds to the introduction of an energy recovery project in a
network where a pressure reduction strategy is in action to reduce water leakage. In
this condition Carravetta et al. [5] demonstrated that the financial benefit of the
energy recovery is large in presence of feed-in tariff but it is small when compared
with the financial benefit coming from the water savings.
In case 2 the back pressure constraints are relaxed to improve the economic
convenience of the hydropower station. This case is observed in small residential
areas or in the peripheral branches of a WSN. Due to the low available power, energy
recovery could produce a benefit only if the cost of equipment is reduced, compared
to that in case 1. This economy can be obtained by simplifying the PAT regulation.
In case 3 the only design requirement consists in the economic convenience of the
hydropower plant. This is frequently the case in water transmission, but it can be observed
also in new pipeline systems where any excess in back pressure can be tolerated.
5.2 PAT Working Conditions 103

5.2 PAT Working Conditions

The case of a PRV replacement is probably the most complete in terms of a


description of a PAT design strategy. In the framework of a pressure reduction
strategy the PRVs can be regulated automatically by imposing a constant back
pressure value, or dynamically by operating on the PRV to obtain a time variable
back pressure value. This difference will be reflected in the necessity of giving a
constant or time variable back pressure constraint in the design of the hydropower
plant, but it will not change the possible regulation modes.
The regulation of the PAT in a hydropower plant can be obtained hydraulically
or electrically (Fig. 5.6). The goal of PAT regulation is to match the WSN operating
conditions with the PAT working conditions. Depending on the reciprocal position
of the (Q, ΔH) WSN operating point and of the PAT ΔH(Q) characteristic curve, the
WSN operating conditions are forced to match the possible PAT working condi-
tions in three ways: (i) by bypassing part of the flow rate, so reducing the flow in the
PAT; (ii) by activating a PRV in series with the PAT, so increasing the pressure
drop of the PAT; or (iii) by varying the PAT rotational speed, so matching the WSN
operating point with the Q(ΔH) characteristic curve.
In Fig. 5.6 a comprehensive sketch of the PAT installation in the different
regulation modes is given. In the hydraulic regulation mode (HR) the PAT rota-
tional speed is taken as constant and the inverter is not present. In the electrical
regulation mode (ER) a single speed drive is present without any bypass or series
valve. Finally, the HR and ER modes can be coupled in the hydraulic electrical
regulation mode (HER) and all the equipment of Fig. 5.6 will be in use. In any case,
it is intended that the hydropower plant will be supplied with an external bypass
pipeline, in order to exclude the possibility of the use of plant with an on/off valve
during maintenance or anomalies. The PAT working conditions in the HR, ER and
HER modes are shown in Fig. 5.7a–c, respectively.

Fig. 5.6 Installation scheme


of a PAT in a hydropower
plant [6]
104 5 PAT Control Systems

Fig. 5.7 PAT working conditions in the different regulation modes

5.2.1 The HR Mode

In the HR mode, for a net-head, DHd, higher than the head-drop deliverable by the
machine, DHT (at the left of the PAT characteristic curve in Fig. 5.7a), the valve in
the series dissipates the excess pressure. Instead, when the discharge, Qd, is larger
(at the right of the PAT characteristic curve in Fig. 5.7a), the PAT produces a
head-drop higher than the available net-head: in this case, the bypass valve is
opened to reduce the discharge flowing in the PAT from Qd to QT.
The advantage of the HR mode is the simplicity in the power plant operation.
The two valves, the bypass regulating valve and series regulating valve, could be
diaphragm-operated control valves, like the one shown in Fig. 5.8. These valves
operate automatically by reducing the downstream pressure to a constant set-up
value, independently of the flow variations or upstream pressure. A disk, whose
position is forced by the fluid pressure on a diaphragm, restricts the cross section of
the pipe, determining the dissipation of excess energy in the series regulating valve,
or the opening of the bypass in the bypass regulating valve. In principle there is no
need in a power plant with the HR regulation mode of any control unit, electronic
pressure or flow measurement, or electric actuator.
The disadvantage of the HR mode consists in the presence of working conditions
existing far from the PAT’s best efficiency point (BEP), in any case in which the
flow rate varies within a wide range of values.

5.2.2 The ER Mode

In the ER mode, Fig. 5.7b, the operating speed (N) of the generator is changed to
match the load conditions determined by the instant flow discharge and head drop
values; namely, the PAT characteristic curve is modified to match the available
head. In the case of a pump asynchronous motor acting as a generator in reverse
5.2 PAT Working Conditions 105

Fig. 5.8 Diaphragm-operated control valve

mode, the variation of the rotational speed is obtained by varying with an inverter
the frequency of the stator excitation circuit.
The advantage of this regulation mode lies in the possibility of using the whole
flow rate in any WSN operating conditions to produce energy. This advantage is
often only a presumed advantage. In fact, it frequently happens that the PAT
characteristic curves in the whole range of rotational speeds do not cover the
complete variability of the WSN operating points in the (Q, ΔH) plane. In this case
the ER mode cannot be used for a strict PRV replacement.
The disadvantage of the ER mode consists in the higher cost connected with the
electronic regulation of the rotational speed, based on pressure or flow measure-
ments, and the use of an inverter to vary the frequency of the generator, of a rectifier
to flatten the current generated by the PAT, and of a second inverter to produce
energy at the grid frequency.

5.2.3 The HER Mode

In the HER mode, the valve stroking and operating speed can be selected to obtain
the derivable high drop. In general, for a given flow rate, the required back pressure
can be obtained for different rotational speeds, with a different regulation of the two
106 5 PAT Control Systems

regulation valves. Therefore, based on the characteristic, efficiency and reliability


curves of the PAT, an optimization algorithm can be set up to obtain a specific
power plant target, in terms of the production or reliability of the PAT.
Even if the HER regulation mode is the most complete and operatively sound
power plant set-up, the cost of the equipment and the complexity of the set-up
phases are the highest when compared with the other regulation modes.

5.2.4 The SSP Mode

The single-serial-parallel regulation (SSP) mode has been proposed as a means of


reducing hydropower plant costs in all cases where a PRV weak replacement is
possible [6]. This situation is usually present in small residential areas where the
range of flow values is very wide, due to the high daytime peaks in user demand and
the low nightly flow rate. Therefore, the new regulation mode is based on the use of
two PATs which could act in parallel for the larger flow rates. The scheme of PAT
installation is given in Fig. 5.9. Three on/off valves and two PATs are used to
obtain three different HPP working conditions:
(a) Valve I on, PAT A on, valve II and valve III Off, PAT B off—single PAT;
(b) Valve II and valve III on, PAT A and PAT B on, valve I off—series of PATs;
(c) Valve I and valve III on, PAT A and PAT B on, valve II off—parallel PATs.
In working condition (a) a single PAT is producing energy for discharge close to
average; in working condition (b) two PATs are producing energy in series small;
finally, in working condition (c) two PATs are producing energy in parallel for the
highest discharge.
The working conditions in an SSP regulation are shown in Fig. 5.10. For the
smallest flow rates working condition (b) is activated; next, for increasing flow rates
the power plant moves to working condition (a), finally for the largest flow rates
working condition (c) is in operation. It is clear that the exact ΔH value is rarely

Fig. 5.9 SSP installation mode


5.2 PAT Working Conditions 107

Fig. 5.10 Working and operating conditions in SSP

obtained in SSP, but a pressure in the downstream part of the network will reduce
below the optimal value for some period. In the calculations of Fig. 5.10 a 5 m
excess drop in the power plant is accepted, and the transition between the two
following conditions is determined when the dissipation connected to the previous
working condition exceeds this limit.
An initial advantage of SSP is that there is no need in a power plant with the HR
regulation mode for any control unit, electronic pressure or flow measurement, or
electric actuator. On/Off valves, the on/off command of a PAT, can be designed
pneumatically on the basis of the upstream flow pressure. A second fundamental
advantage lies in the possibility of reducing the cost of the equipment, a hydro-
power plant working in SSP costing approximately one third of the corresponding
cost of a plant working in HR. In fact, On/Off valves, Fig. 5.11, are much less
expensive than pressure reducing valves.

Fig. 5.11 Diaphragm-operated on/off valve (http://www.bermad.com/Data/Uploads/FP-400-


Engineering.pdf)
108 5 PAT Control Systems

The disadvantage of SSP lies in the rough regulation of the back pressure.
Therefore, this regulation mode is recommended only in peripheral areas of the
network which have an excess of pressure able to recover a small amount of energy
with a low cost device.

5.3 Plant Effectiveness

Within the framework of the most complete method for PAT selection, including
not only electro-mechanical efficiency, but also PAT reliability, flexibility and
sustainability [5], the effectiveness equation can be taken into account [7–10] as the
objective function in power plant optimization. Effectiveness responds to the
question of how well the product/process satisfies the end user’s demands. The
effectiveness equation can be written as follows:

E ¼ A1 . . .A1 . . .An ð5:3Þ

where A1 . . .A1 . . .An are the n performance indicators of the system, influencing its
overall effectiveness, ranging between zero and one.
For PAT design, the following four indicators have been suggested in terms of
evaluating power plant effectiveness E.

5.3.1 System Capability

This indicator shows how well the system performs according to expectations.
Considering the operating and working conditions of Fig. 5.7, the capability is
defined as:
Pn
DH Q g ðN ÞDti
Pn i i i
gp ¼ i¼1 ð5:4Þ
i¼1 DHi Qi Dti

Δti being the time interval with constant working conditions (Qi; ΔHi).
It is interesting to stress that only in the ER mode does the system capability
correspond to the average machine efficiency. In the other modes the system
capability gives the fraction of the available power that is transformed by the
hydropower plant into energy. In any case in which the target of the hydropower
plant design is to maximize energy production, the system capability will be the
only relevant performance indicator in the effectiveness equation.
5.3 Plant Effectiveness 109

5.3.2 System Reliability

Due to the variability in PAT working conditions, the load on the


electro-mechanical components of the PAT is not constant but variable in time. Let
us consider a cycle of operations whose period is T and with constant conditions for
each Δti time interval, as shown in Fig. 5.12. The reliability, or the probability that
the machine will work without failure for the whole period T, R(T), equals the
probability that the machine will not fail in any of the Δti intervals. If the reliability
is time-independent, i.e., the effect of the aging of the machine is not considered or
is considered as constant, and a correct maintenance program is carried out, the
reliability can be expressed by the equation:

Y
n
RðT Þ ¼ RðDt1 ÞjL1 . . .RðDti ÞjLi . . .RðDti ÞjLn ¼ RðDti ÞjLi ð5:5Þ
i¼1

Expressing the reliability by an exponential probability distribution:

Y
n Y
n Pn Pn
RðT Þ ¼ RðDti ÞjLi ¼ e i¼1 ki Dti ¼ e i¼1 ki Dti ð5:6Þ
i¼1 i¼1

For an average failure rate:


Pn
ki Dti
kav ¼ i¼1
ð5:7Þ
T

the reliability is expressed by:

RðT Þ ¼ ekav T ð5:8Þ

Fig. 5.12 Load L variation


during the period, T
110 5 PAT Control Systems

and the average mean time to failure of the system under variable working con-
ditions, MTTFav, can be calculated as 1/kav.
Finally, the system reliability, expressing the reduction of the mean time before
any failure of the system, compared to the mean time before a failure of the system
at BEP, is calculated as:

MTTFav
lp ¼ ð5:9Þ
MTTFB

The closer lp is to one, the less the system reliability is affected by the power
plant variable working conditions.

5.3.3 System Flexibility

For most of the design scenarios, the PAT design parameters are determined by the
numerical solution of WSN simulations. Despite the fact that pressure head and
discharge measurements are available in particular WSN nodes, not negligible
differences will be present between design and real life conditions. Even in the
presence of ad hoc measurements in the hydropower plant node, the new regulation
point will modify the WSN regime with scatters between the planned and realized
conditions. As a consequence, the system productivity could be lower than
expected. Assuming, during operations, real head drop Hd values different from the
design ones by ±10%, the system flexibility will be defined as the minimum of the
ratio between the plant efficiency at ±10% Hd, namely η±10%, and the design plant
efficiency will be:
!
gpþ 10% g10%
p
/p ¼ min ; ð5:10Þ
gp gp

The system flexibility is, therefore, an important design parameter: the closer /p
is to one, the less the global efficiency will vary around the design efficiency in the
presence of unexpected Hd variations around the design value.

5.3.4 System Sustainability

When the plant is provided with HR, each value of DHt and Qd can be matched by
the mutual regulation of the valves opening. Conversely, in ER, some operating
points could lie outside of the regulated region shown in Fig. 5.7: this results in a
BP value different from that assigned. Similarly, when a simplified regulation is
present (SSP), the BP value can be reasonably different from that assigned due to
5.3 Plant Effectiveness 111

the difference between the operating points and the characteristic curve, as shown in
Fig. 5.10. In both cases a penalty should be considered in the optimization process
and such a penalty can be included as a factor in the calculation of the effectiveness.
This new factor, namely plant sustainability, vp , ranges between 0 and 1 and has
been defined as:
 t !
DH  DHd  1
vp ¼ 1þa ð5:11Þ
BP

where a is a coefficient influencing the decrease in effectiveness when the produced


head drop is different from the design value.

5.4 PAT Selection

The starting point for PAT selection is a reliable set of data giving the daily pattern
of discharge and head. The PAT operating conditions can be determined both
experimentally, as measured by a monitoring station in the production node, or
numerically, by solving the governing equations of flow in the WSN. The reliability
of the operating conditions to be used in the design procedure and the presence of
smoothing in the flow and head fluctuations can influence the design results.
System flexibility could be used to account for all the uncertainties in PAT oper-
ating conditions.
The second step in the design procedure consists in the choice of PAT type (e.g.
centrifugal, semi-axial, etc.) and size range. This is not an easy task at the current
state of knowledge. In fact, it is very difficult to obtain characteristic and efficiency
curves from industry and, when these curves are available, they are referred to a
single rotational speed. Nevertheless, once the pump type is selected, for a given
ratio between the diameter of the runner and the diameter of the PAT body, it is
possible to extend the results obtained in a specific condition, e.g. for a specific
diameter of the runner and rotational speed, to the whole possible range of runner
diameters and rotational speeds. System flexibility could be used to account for all
the uncertainties in the definition of PAT working conditions.
The strategy for PAT design is the same for all regulation modes, and it is known
as the variable operating strategy (VOS) [2, 5]. The following steps are necessary:
1. A measured pattern of flow-rate and pressure-head conditions is assigned and
the available head is determined based on the required back pressure (BP);
2. The PAT type is considered (e.g. centrifugal, semi-axial, etc.);
3. A wide set of PAT characteristic curves is considered in the PAT operating
region;
4. For each curve the overall plant effectiveness is calculated by:
112 5 PAT Control Systems

E ¼ gp  lp  /p  vp ð5:12Þ

5. The PAT that maximizes the power plant effectiveness, i.e. the PAT having the
largest E value, is considered the optimal design solution;
6. The near-optimal machine is selected from the market and its turbine mode
curves are calculated to verify the actual effectiveness.
In order to perform step 3, the characteristic and efficiency curves for a whole set
of PATs at different rotational speeds, having their BEPs in the operating region,
are necessary. Such trail curves could be obtained from a single reference prototype
machine curve by using the turbomachine affinity law. Furthermore, since the Suter
parameters can be calculated from the data from a single PAT curve, the perfor-
mance curves of the similar PATs [11, 12] can be assessed. The possibility of using
a reference characteristic curve of a PAT based on Computational Fluid Dynamics
(CFD) rather than on an experimental procedure [13, 14] produces great advan-
tages. Indeed, shaft geometry and stage numbers (for multistage machines) can be
easily modified in numerical simulations.
The BP downstream the power plant could be considered in some cases a desired
value, instead of a strict condition, which is the case in respect of SSP regulation
modes. It is clear from Fig. 5.10 that the available head drop is dissipated exactly
only in a few moments of the day, when the power plant operating conditions match
exactly the network working conditions. In the rest of the day the BP will be larger
or even smaller than required. The same situation is possible in ER, in the presence
of a range of PAT operating conditions not covering the whole set of network
working conditions. In this case, the variation of the rotational speed in not capable
of obtaining the required cross matching between the characteristic curve and the
network working point. In both cases, step 1 of the VOS will be modified and the
BP will not be imposed as a fixed constraint but only the maximum value of
allowed BP in the network will be imposed.
Obviously, the results of the VOS in terms of the design solution, plant effec-
tiveness and system capability may vary significantly depending on the network
working conditions and on the plant operating conditions. In particular, the larger is
the range of flow rates and head drop the higher is the complexity of the PAT
regulation. Additionally, intermittent working conditions, as in the presence of a
dynamic control of pressure, will reduce the power plant effectiveness. In relation to
the PAT operating conditions, the coarser is the range of available design solutions
in terms of PAT runner diameters the more difficult will it be to get close, by means
of the near-optimal machine, to the best VOS solution.
As an example of power plant design, the results of a VOS application to the
data of Fig. 5.5, using a centrifugal single stage PAT (Ns = 44 [rpm kW1/2 m−4/3]),
are discussed for all the regulation modes of Table 5.1. The plant capabilities for
variable BP values and impeller diameters are plotted in Fig. 5.13 with reference to
the HR and ER regulation modes.
The field of application of a PAT family is larger in HR than in ER, as shown by
the feasible colored area of Fig. 5.13. Outside this area, a number of operating
5.4 PAT Selection 113

Table 5.1 Hydropower plant regulation modes


Equipment Case I Case II Case Case Case V
(Fig. 5.8) (HR) (ER) III IV (HER)
Inverter ✗ ✓ ✓ ✓ ✓
Valve A ✓ ✗ ✗ ✓ ✓
Valve B ✓ ✗ ✓ ✗ ✓

Fig. 5.13 Plant capability in HR and ER modes [15]

conditions could not be correctly managed by the power plant, and in this case the
BP will be larger or smaller than the optima: in HR this will happen when the head
drop is too small compared with the PAT characteristic curve or when the flow rate
is smaller than runaway flow rate; in ER it will be whenever the operating con-
ditions do not match the working conditions for any value of the rotational speed.
In Fig. 5.14 the zones of capability larger than 0.4 for all the possible modes of
Table 5.1 are shown, considering the combined use of valves and inverter. The HR
and ER modes are complementary design choices. HER (case V) is, obviously, a
very flexible plant choice, granted that a PAT family could work with a high
capability in the widest range of BPs, but case III, with an inverter and bypass,
seems a valid alternative. Obviously, cases III–V represent the most expensive
solutions, both in terms of the equipment cost and the power plant control.
The plant flexibility is shown in Fig. 5.15a, b for HR and ER, respectively. The
effect of the plant flexibility is to reduce further the effectiveness in part of the area
of low capability. On the contrary, a high flexibility corresponds to the high
capability values.
Contour values of plant reliability are plotted in Fig. 5.16 showing an appre-
ciable difference between the HR and ER modes. The use of an inverter grants a
high reliability to the design solutions with the best capability. Instead, the zones of
the highest reliability and capability are not coincident in HR, and some pump size
solutions presenting a high capability could represent a wrong solution in terms of
the length of the life of the electro mechanical components. Finally, the plant
114 5 PAT Control Systems

Fig. 5.14 Plant capability for all possible modes [6]

Fig. 5.15 Plant flexibility in HR and ER modes [15]

effectiveness is shown in Fig. 5.17. It is evident that the design solutions in the HR
and ER modes are completely different. In ER a larger pump size is necessary with
an increase in the equipment costs. In Table 5.2 optimal solutions in term of
effectiveness for different BP values are reported: the plant capability is also plotted.
For the analyzed scenario HR always achieves the best effectiveness with an
appreciable capability value. If the optimization is performed considering only plant
capability [15], the results will change and ER will give a larger energy production.
5.4 PAT Selection 115

Fig. 5.16 Plant reliability in HR and ER modes [15]

Fig. 5.17 Plant effectiveness in HR and ER modes [15]

Table 5.2 Optimal solutions in term of effectiveness for different BP values


BP (m) 10 20
D (mm) E (−) ηp (−) D (mm) E (−) ηp (−)
HR 179 0.529 0.567 178 0.531 0.549
ER 239 0.199 0.289 221 0.432 0.495
BP (m) 30 35
D (mm) E (−) ηp (−) D (mm) E (−) ηp (−)
HR 171 0.471 0.499 163 0.394 0.439
ER 231 0.422 0.495 250 0.306 0.410
116 5 PAT Control Systems

References

1. A. Carravetta, O. Fecarotta, R. Martino, L. Antipodi, PAT efficiency variation with design


parameters. Procedia Eng. 70, 285–291 (2014)
2. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Energy production in water
distribution networks: a PAT design strategy. Water Resour. Manage 26(13), 3947–3959
(2012)
3. R. Gargano, D. Pianese, Reliability as tool for hydraulic network planning. J. Hydraul. Eng.
126(5), 354–364 (2000)
4. R. Gargano, C. Tricarico, G. del Giudice, F. Granata, A stochastic model for daily residential
water demand. Water Sci. Technol. Water Supply 16(6)
5. O. Fecarotta, E. Al, Optimal regulation of a PAT in water supply systems for energy recovery.
Water Resour. Manage (2016)
6. A. Carravetta, O. Fecarotta, G. Del Giudice, H. Ramos, Energy recovery in water systems by
PATs: a comparisons among the different installation schemes. Procedia Eng. 70, 275–284
(2014)
7. D. Kececioglu, X. Tian, Reliability education: a historical perspective. IEEE Trans. Reliab. 47
(3), SP390–SP398 (1998)
8. M. Pecht, Product reliability, maintainability, and supportability handbook (CRC Press, Boca
Raton, FL, USA, 2009)
9. D.G. Raheja, M.L. Lindsley, Defect-free manufacturing in the nineties. in Proceedings of the
IEEE 1991 National Aerospace and Electronics Conference NAECON 1991, pp. 1019–1022
(1991)
10. R. Clements, Handbook of Statistical Methods in Manufacturing (Retroactive Coverage).
Upper Saddle River, NJ, USA (1991)
11. H. Ramos, A.B. Almeida, Dynamic orifice model on waterhammer analysis of high or
medium heads of small hydropower schemes. J. Hydraul. Res. 39(4), 429–436 (2001)
12. H. Ramos, A. Beta⋌mio de Almeida, Parametric analysis of water-hammer effects in small
hydro schemes. J. Hydraul. Eng. 128(7), 689–696 (2002)
13. O. Fecarotta, A. Carravetta, H. Ramos, CFD and comparisons for a pump as turbine: mesh
reliability and performance concerns. Int J Energy Environ. 2(1), 39–48 (2011)
14. A. Carravetta, O. Fecarotta, H. Ramos, Numerical simulation on pump as turbine: mesh
reliability and performance concerns. in 3rd International Conference on Clean Electrical
Power: Renewable Energy Resources Impact, ICCEP 2011, pp. 169–174 (2011)
15. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Pump as turbine (PAT) design in
water distribution network by system effectiveness. Water (Switzerland) 5(3), 1211–1225
(2013)
Chapter 6
Civil Engineering Design,
Electromechanics, Installation
and Operation

Abstract Information about the different components of a PAT and an overall


description of civil engineering design are presented. The electromechanical
transmission arrangements realized to avoid dissipative effects and to guarantee the
best operating conditions are defined and additional aspects associated with
installation procedures and operation and maintenance are described. A theoretical
and project-based design for the comprehension of a micro-hydropower scheme
connected with a PAT installation in water systems is presented. Different design
components and the rules to be followed by engineering designers are defined.

A PAT-based micro-hydropower plant includes civil engineering and electrome-


chanical aspects, the latter consisting in mechanical and electrical equipment. In
addition to civil engineering work, mechanical and electrical components are
required in a PAT, as well as a generator, control system and electricity transfer.
The PAT body belongs to the class of mechanical components but the generator and
control system are electromechanical systems. While a hydropower plant is a chain
operation from the water intake to the electricity consumer, an overview of its
components is very useful in terms of the design, selection and operation. The
installation and operation of micro-hydro solutions can improve energy efficiency
through electricity generation and simultaneously in the pressure control of water
systems, particularly in cities with considerable geodesic differences, or in irrigation
areas, drainage systems and pipelines, and in pipe sections where the pressure is
higher than an adequate or necessary level.

6.1 Civil Engineering Design

The civil works include several components: (i) the water intake and PAT inlet;
(ii) the conveyance system or hydraulic circuit; and (iii) the powerhouse and
machine outlet. For micro-hydro stations (MHS), the civil works are reduced and
relatively straightforward since the layout schemes are, basically, characterized by

© Springer International Publishing AG 2018 117


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_6
118 6 Civil Engineering Design, Electromechanics …

an existing supply infrastructure consisting in a pipe system with some valves or


control devices installed. The investment is generally limited and there is no
requirement to spend so much on hydraulic and structural design components.
These structures have a main advantage in the optimization of the use of local
equipment from existing infrastructures. Hence, an MHS design should essentially
be the responsibility of hydraulic and electrical experts, there being necessary only
a relatively reduced working team.

6.1.1 Intake and Turbine Inlet

In a typical hydropower station, the water intake can be by means of a frontal,


lateral or bottom drop, depending on the system characteristics [1–4]. Special
attention must be paid to the design of the inlet in order to avoid separation zones of
the flow and excessive head losses. It is also necessary in all system types to verify
the minimum submergence in order to avoid vortex formation and, consequently,
an entrance of air. The discharge control can be improved by special structures (e.g.
narrow sections or venture type weirs, gates or valves) depending on whether or not
the inlet approximation flow is pressurized or is of a free surface type, as presented
in Figs. 6.1 and 6.2.
A lower flow velocity at the turbine intake, as a uniform distribution of flow
velocity along the water outlet, allows a minimizing of the vorticity of the flow, and

(a) (b)

(c)

Fig. 6.1 Flow in a curve of 90 and 45°. a Longitudinal cut with flow separation zones in 90°.
b Velocity profiles and separation zones in 45°. c Pressure and flow distribution
6.1 Civil Engineering Design 119

Fig. 6.2 Typical lateral and frontal water intake: design and distribution of pressure and velocities

therefore the intensity of the turbulence and pressure losses induced. The increased
uniformity in the distribution of the flow velocity provides a reduction in the load
losses and an increase in the turbine efficiency and consequently can be regarded as
a significant improvement in the hydraulic efficiency of a water intake.
In irrigation systems the water intake can be of a frontal or lateral type (Fig. 6.2).
For drainage systems the water intake can be of a bottom drop type (Fig. 6.3).
The problem most often attributed to the formation of vortices in the water intake
is the loss of hydraulic efficiency, resulting from impairments (disorders) to the
flow, which
• gives rise to non-uniform flow conditions;
• promotes the entry of air in the flow, leading to the formation of adverse
operating conditions in hydraulic turbomachines, in particular vibrations, cavi-
tation and unbalanced loads. The release of entrapped air results in high over-
pressures that might lead to a collapse of the penstock/pipe;

Fig. 6.3 Bottom drop intake


120 6 Civil Engineering Design, Electromechanics …

• increases losses and reduces hydraulic efficiency; and


• promotes the dragging of solid waste into the water inlet, leading to a blockage
of the grids, so increasing the load losses (in the case of free open channel
intakes).
A possible vortex classification considers four main types [1]: Type 1—a
developed vortex with a deep nucleus and with drape air; Type 2—a superficial
depression, without drag bulb air but with a well-defined nucleus; Type 3—a quite
negligible depression with an unstable nucleus; and Type 4—a rotational movement
without a free-surface depression but with a superficial circulation.
The vortices are caused by a non-uniform acceleration of the flow. The per-
turbations that lead to a non-uniform speed can give vorticity. These disorders
include:
• conditions of asymmetric approach;
• irregularities in the surface geometry;
• an inadequate submergence;
• approach velocities exceeding 0.8 ms−1;
• a separation of flow and the formation of turbulence;
• abrupt changes in flow direction;
• obstructions to flow;
• and flow torrents.
The attack angle of the flow to the turbine runner requires an adequate
approximation structure to avoid turbulence formation (Fig. 6.4).
The knowledge learned through a proper modelling of hydraulic structures and
equipment has contributed to an improvement in their water and energy efficiencies
and to a reduction in the requirements regarding protection and control measures.
Another criterion, based on the maximum velocity flow in a pipe referred to as
the penstock, can give a first estimation for the adequate pipe diameter: 2–3 m/s for
low head plants; 3–4 m/s for medium head plants and 4–5 m/s for high head plants.

Fig. 6.4 The different types of flow inlet approximation angle and consequences on the velocity
distribution
6.1 Civil Engineering Design 121

These are typical values based on real cases. For penstocks different types of
materials can be used:
• steel and cast iron piping is widely used in hydropower design and is preferable
for very high head plants in that it allows a long service life. Expansion joints
should be provided at certain intervals, depending on local climatic conditions;
• pre-stressed concrete piping presents a reduced cost but a greater difficulty in
installation;
• plastic and glass fibre reinforced plastic piping induces a small friction loss, but
must be buried or wrapped to protect it from the damaging effect of sunlight;
and
• reinforced concrete piping, which has a long service life, little maintenance and
a low cost, but a high difficulty in installation and resistance problems.
The net forces due to pressure and momentum change can be calculated as
follows:

Fpx ¼ p A ðcos h1  cos h2 Þ Fpy ¼ p A ðsin h1  sin h2 Þ


ð6:1Þ
Fmx ¼ qQV ðcos h1  cos h2 Þ Fmy ¼ qQV ðsin h1  sin h2 Þ

where p = the pressure in the penstock; A = the cross-section area of the penstock;
Q = the discharge; and h 1 and h 2 = the angles, as shown in Fig. 6.5.
At each pipe change direction, the penstock and its supporting structures must be
designed to resist the forces resulting from changes in direction (Fig. 6.5).

6.1.2 The Powerhouse

The powerhouse can be included in a valve chamber if there is sufficient space or in


a chamber built for this purpose. The function of any powerhouses consists in
housing and protecting turbo-generator groups and any auxiliary equipment (e.g.
the safety and protection valves, electric boards, control equipment, remote con-
troller, switchgear panel and protection equipment (when applied) (Fig. 6.6)). The

Fig. 6.5 Resultant forces produced by weight, pressure and changes in momentum
122 6 Civil Engineering Design, Electromechanics …

Fig. 6.6 Powerhouse implementation in a valve chamber partially out of service (http://www.
hidropower.eu/)

powerhouse layouts need to allow an easy installation of the equipment as well as


access for inspection and maintenance of the turbines and all other equipment.
The design of a powerhouse needs to take into account the hydraulic constraints
of existing water systems (Fig. 6.7) for a good turbine inflow, the weight of the
equipment and the hydrodynamic forces that will be transmitted to the structure
including the massive concrete blocks (anchors) for the machine support.
Constructive procedures should be analysed in order to take advantage of the
local conditions, in order to avoid additional civil works, as presented in, which can
have a significant impact on the viability studies and on the final decision. Some
details of the last configuration with different design cuts are presented in Figs. 6.8
and 6.9.

Fig. 6.7 Implementations of 2 PATs for energy recovery at the entrance of a water treatment plant
(http://www.hidropower.eu/)
6.1 Civil Engineering Design 123

Fig. 6.8 Constructive process (http://www.hidropower.eu/)

6.1.3 Turbine Outlet

The use of CFD models provides considerable support to the design project of the
various components of the hydraulic circuits of hydropower plants, in particular the
exit flow conditions, in order to analyze the effect of the outlet configuration on the
flow stability (i.e. the velocity distribution, flow path and pressure variation). The
use of this tool also allows designers to anticipate critical issues and improve the
hydraulic/structural project to a level normally only achieved through studies on
physical models (Fig. 6.10).

6.2 Mechanical Components

When a pump is used as a turbine in a hydropower plant, its general arrangement


and components may be slightly changed [5–8]. However, as a pump is a more
straightforward machine than a conventional turbine, its arrangement is very easy.
124 6 Civil Engineering Design, Electromechanics …

Fig. 6.9 Adaptation of an existent valve chamber for the powerhouse (http://www.hidropower.eu/)

Velocity distribution Flow path Pressure variation

Fig. 6.10 Influence of the exit flow in the hydraulic-structural conditions

For example, depending on the PAT rotational speed, a direct coupling or


pulley-belt to transfer the PAT mechanical torque to the generator may be used. In
summary, the component of a PAT consists in wet or hydraulic parts and
mechanical parts. In Chap. 2 (Sect. 2.1), the wet parts (the volute, impeller and
draft tube) are briefly described. Here, the shaft, shaft seal, bearing and coupling are
introduced. Figure 6.11 shows the mechanical parts of a pump (or a PAT).
6.2 Mechanical Components 125

Fig. 6.11 Mechanical parts


of a pump (or a PAT) [9]

6.2.1 The Shaft

The shaft is the main part of a PAT, which transfers the PAT’s power to the
generator. Fortunately, the shaft in a pump is a standard item and its spare parts can
be easily supplied by the manufacturer. Its material is hard (i.e. ST 420 according to
ASTM) and has sufficient thickness to bear the PAT’s power. A pump is normally
designed to work at 1450 rpm (or 1750 for 60 Hz) and 2900 rpm (3500 for 60 Hz)
per shaft. Therefore, using a PAT for 1500 rpm (1800 rpm for 60 Hz) or less is safe
(the PAT power may be greater than the pump input power at the same rotational
speed). In other words, a pump shaft can definitely be used in turbine mode without
any change being made.

6.2.2 The Shaft Seal

The shaft seal provides sealing between the shaft and casing faces to achieve a
leak-proof segregation from inside the pump to the atmosphere. Shaft seals may be
divided into non-contact and contacting seals [9].
The simplest types of non-contact shaft seals are throttle bush seals which form a
small clearance between the stationary bush and the rotating sleeve. The clearances
are kept as narrow as possible to minimize leakage [9].
As for contacting shaft seals, the simplest type, long used in centrifugal pump
design, is the packed stuffing box. However, where these are no longer adequate, on
account of the excessive pressure difference, the circumferential speed or for other
reasons (e.g. the high leakage rate), rotating mechanical seals are employed [9].
Since a PAT’s working pressure is higher than a pump’s working pressure, it is
recommended that the mechanical seal be used.
Single mechanical seals are normally used if the pumped liquid is clean and does
not crystallize when exposed to the atmosphere and if the pumped fluid is
126 6 Civil Engineering Design, Electromechanics …

compatible with the environment. These seals are normally flushed with the
pumped liquid. Therefore, for a PAT, single mechanical seals are suitable
(Fig. 6.12).

6.2.3 The Bearing

Where the bearing load and rotational speed permit, as is generally the case with
standard pumps (and PATs), rolling contact bearings of the ball or roller type are
frequently used (Fig. 6.13). Such bearings provide both radial and axial guidance to
the shaft. To accommodate larger axial forces, self-aligning roller thrust bearings
are often used. Lubrication is with oil or grease, depending on the service condi-
tions [9].

Fig. 6.12 Single mechanical


seal [9]

Fig. 6.13 A ball or roller


type bearing, commonly used
in centrifugal pumps [9]
6.2 Mechanical Components 127

6.2.4 The Coupling

Considering the PAT rotational speed and the rated generator rotational speed,
direct coupling or indirect coupling (by mechanical gear or by belt and pulley) can
be selected for the mechanical power transmission.
The rated PAT rotational speed is to be fixed with regard to the condition of the
site head and flow rate. On the other hand, the generator rotational speed is to be
selected considering the frequency and the number of generator poles.
Therefore, if the rotational speeds of the PAT and generator are exactly the same,
they can be coupled directly by a simple coupling (Fig. 6.14a). Using direct cou-
pling is the best way to transmit the power.
However, such a direct coupling design is sometimes not applicable due to the
high generator cost, especially in the case of mini/micro/pico power plants when the
PAT rotational speed is not the synchronized speed. Therefore, the special facility
of speed increasing, named the belt-pulley system, is usually used in order to match
the speed of the PAT and that of the generator (Fig. 6.14b).
The PAT shaft and generator shaft are coupled with pulleys and a belt according
to the ratio of rotational speeds between the PAT and generator. In this way, the
user can select the desired rotational speed for the PAT.
The cost is relatively low but the belt lifetime is short and the power loss is
greater than that incurred with the use of the direct coupling method (approximately
2–5%).

6.3 Electrical components

The electromechanical converters that are usually employed in a PAT are asyn-
chronous machines, optimized for pumping operations. Such convertors, as pre-
viously mentioned, help to reduce acquisition costs, which, in turn, increases the
competitiveness of a PAT in a power generation scenario, especially considering
the fact that PATs allow for power values starting at 50 W.

Fig. 6.14 Direct and indirect PAT-generator shaft coupling methods


128 6 Civil Engineering Design, Electromechanics …

Another advantage in the use of electromechanical convertors in a PAT is that


the technology has been widely explored and with proven results. Asynchronous
machines are also called induction generators, being magnetized from the
three-phase winding in the stator and inducing slip currents on the rotor.
The rotor revolves at a speed slightly above synchronism and the currents induced
on it are of a frequency corresponding to the speed difference, with the slip defined as
the percentage difference between the synchronous speed and the rotation speed.
The torque results from the magnetic interaction between the stator and rotor
fluxes vary with the slip, as indicated on the torque/slip curve of Fig. 6.15.
When a PAT is installed in real conditions, it allows for head, flow and pressure
variation, and grid connected excitation [1]. The electrical machine or the motor
employed with a PAT is usually an asynchronous generator, whose characteristics
are presented with the machine (as in Table 6.1).
In order to test a PAT under different operating conditions an accurate com-
putational model through the MATLAB/Simulink software can be used [10]. This
is based on the no-load and blocked rotor rests, performed in order to determine the
resistive and inductive characteristics of the machine. A typical model is shown in
Fig. 6.16.
Asynchronous generators can only supply active power to the grid and must
import reactive power for their magnetization.
The procedure for grid connection requires a running up driven by the turbine.
The generator breaker is closed when the actual speed passes over the synchronous
speed, at a much reduced acceleration, in order to limit circulating currents when
closing. On the grid connection the generator absorbs a transient magnetizing
current of short duration. The turbine power is then increased at a suitable rate by
the wicket gate control, while the slip and stator current increase according to the
generator characteristics.

Fig. 6.15 Torque/slip curve

Operating point

Table 6.1 Characteristics of Operating frequency (Hz) 50


an asynchronous generator for
a typical PAT Operating voltage (V) 400
Motor winding (V) 400/230
Rated power (W) 550
cos/ at 4/4 load 0.81
Number of poles 6
Starting current ratio 3.9
6.3 Electrical components 129

Fig. 6.16 PAT MATLAB/Simulink model

The asynchronous generator control and handling towards the grid is simpler
than the synchronous but its application in small power plants (or mini-hydro)
schemes is normally limited to 2 MW machines due to the transient effects of grid
connection and the need to install capacitors for power factor correction.
Electrical loads existing on a given network may have a behaviour of three
types: purely resistive (e.g. a filament lamp), inductive (e.g. a coil, a fluorescent
lamp, or an induction motor) or capacitive (e.g. a condenser, or a cable under-
ground). Inductive and capacitive loads have a lag effect between the voltage wave
and current wave deriving from the fact that during a period of the voltage
waveform, this type of load stores a certain amount of energy, returning it to the
network subsequently, during a further period of time. This gap is the angle dif-
ference between the voltage and current waves, and is represented by /. For
resistive loads / = 0°, for inductive loads / = 90° and for capacitive loads
/ = −90°.
In fact, the amount of energy presented in a circuit having such loads is larger than
that which is used to perform useful work. This energy, known as apparent power (S),
comprises the power output i.e. that which is converted into work, called the active
power (P), and loads introduced by what is termed the reactive power (Q).
By feeding a given set of charges it is necessary to provide apparent power,
namely active power and reactive, and these can be calculated through the fol-
lowing Eqs. (6.1) and (6.2).

P ¼ U  I  cos/ ð6:2Þ

Q ¼ U  I  sin/ ð6:3Þ

where cos/ is called the power factor and allows a calculation of the value of the
apparent power which is converted into active power. The smaller the value of the
130 6 Civil Engineering Design, Electromechanics …

power factor, the greater the value of the reactive power and, consequently, the
greater the current needs of a given device, compared to a purely resistive load.
The power factor is a quantity of the utmost importance in a micro-water
environment, mainly because the larger the reactive power needs of a given
machine, the greater the charges imposed on the generators. These charges may
possibly not be maintained and/or may suffer overloads. Table 6.2 presents some
typical power factor values for certain items of electrical equipment.
One way to improve the power factor of a given installation is to use capacitor
banks.
An asynchronous or inducing machine has been massified as a motor, which has
contributed several factors, including strength, low maintenance, and efficiency and
flexibility driving variable loads. The induction machine as a generator entity
appears as a viable alternative to the synchronous machine in the context of mini-
and micro-production.
The flexibility of such an induction motor is also present in its use as a generator.
This flexibility makes it an ideal induction machine for use in variable speed drives.
The characteristic parameters of the equivalent circuit of an induction machine can
be calculated from the known pattern in no-load and testing loop circuits.
In relation to the use of electric energy (EE) in micro-hydro systems, it is a
requirement that the generation is performed either connected to the national grid or
independently (isolated). Whether performed in DC or AC, small-scale power
generation is subject to perturbations derived from rotational speed variations of the
turbine, which results in not inconsiderable voltage amplitude and frequency
fluctuations (this is not applicable to generation in DC). In isolated generation, the
use of EE is required so that these fluctuations do not result in damage to household
appliances, not prepared to handle large variations in voltage and current; in sys-
tems connected to the network, the use of EE is necessary to ensure that the
frequency and the amplitude of this voltage do not affect the values of the network.
In terms of power conversion, there are four major groups of inverters, namely:
• a DC/AC converter, also called an “inverter”. This is used to convert direct
current into alternating current. It is ideal for use in situations where the gen-
eration is made in DC.
• an AC/DC converter, also called a “rectifier”. this is used to convert alternating
current into direct current. It is ideal for use in situations where the generation is
made in AC but it is necessary to feed equipment working in DC.

Table 6.2 Typical factors for Equipment Power factor


certain items of electric
equipment Incandescent bulbs (filament) 1
Engines activating light loads 0.4 Inductive
Engines triggering heavy loads 0.8 Inductive
Fluorescent lamps 0.5–0.7 Inductive
6.3 Electrical components 131

• a DC/DC converter. This is used to change the amplitude of the DC voltage,


raising it or lowering it. It is ideal for situations in which a given device does not
have a power supply with the same amount as that generated.
• an AC/AC Converter. This is used to change the amplitude and/or frequency of
the alternating voltage. It is ideal for use in applications where a stable fre-
quency of the drive output is required (e.g. a connection to the network).
All these converters are applicable to small power water systems, either alone or
in conjunction with other methods.
One of the problems associated with micro generation systems is related to the
availability of the resource vs. energy needs. One of the solutions found to coun-
teract the problem of storage is the use of one or more batteries (called a “battery
bank”). These serve the purpose of storing energy for use in times of peak con-
sumption or sale to the network, and can also serve as an aid in stabilizing the
voltage waveform in unstable production times.
Another solution is to store energy, instead of (or in addition to) making the sale
to the network, or to dissipate energy loads, using the excess energy to power
pumps that pump the water to elevated tanks. This solution is particularly useful in
the case of water supply systems, since the company managers have to ensure a
minimum volume of water in tanks/reservoirs. The typical single line diagram of a
mini-hydro power station is shown in Fig. 6.17. The transformer is a static unit with
the purpose of stepping up the generated voltage to the grid connection voltage
level.
Mini-hydro power stations normally use the medium voltage switchgear,
switches housed in metal cubicles, these being standard market products satisfying
the European electrical regulations (CEI). The circuit breakers operate in SF6 or in a
vacuum also with standard ranges of breaking capacity and open/close times.
The automatic controls are based on a Programmable Logical Controller (PLC),
which receives on line information through transducers and digital input signals and
takes the necessary control actions for water utilization, flow or level control.
The PLC output is processed via suitable relays.

Fig. 6.17 Typical single line


diagram [1]
132 6 Civil Engineering Design, Electromechanics …

The PLC is also used for data processing and transmission via telephone
modems or radio signalling.
The generating unit behaviour shows the effects of the hydraulic circuit, the
turbine and the generator dynamics, with the regulation parameters very accessible.
Figure 6.18 shows the interaction of the various blocks.
For micro hydropower plants the transformers and switching gear can be located
inside the powerhouse.
As a practical application of a PAT with a generator group associated with a
machine of the asynchronous type, a horizontal axis, for direct coupling to the
turbine shaft, presents the following characteristics:
• Power (on the shaft) 57 kW
• Apparent power 60 kVA
• Rated voltage 0.4 kV
• Nominal power factor (provided as a generator) 0.85 (cap.)
• Synchronous speed 1000 rpm
• Protection class IP23
For the isolation of the group from the hydraulic circuit upstream and down-
stream, the application of two valves (e.g. the butterfly type) is applied, section-
alizing type DN 350. The electrical control of the opening of the downstream
isolating valve to move parallel with the network is performed according to the law
of opening established by automation.
In electrical fault situations the electric actuator of the downstream isolating
valve must proceed to its closing. Under these conditions the regulating valve of the
main circuit should start its opening in accordance with the indications from the
electrical or mechanical protection of the installation associated with the generator
and addressed in terms of its automation.
The command of the main circuit valve will, however, be borne by the existing
PLC, which communicates with the PLC group to receive directions to the
appropriate command of the flow control valve (existing).
The closure valve downstream the group must be secured by means of an
independent system of self-power, to avoid vibration problems when reaching the
runway speed by an overspeed of the group in voltage drop situations of the
receiving network. To achieve this objective, a UPS can be installed that should
have adequate power and autonomy for this purpose, and the power should be
increased compared to the values imposed for such quantities, by the electric
actuator that may be applied to the downstream valve group. Concerning the feed

Fig. 6.18 Power conversion scheme [1]


6.3 Electrical components 133

from the UPS, it is still recommended to install an AC/DC converter for 24 V to


feed the automation equipment and to protect the operation of the associated group.

6.4 Installation Procedure

After the various elements of the power station and the respective hydraulic circuit,
which constitute the hydropower project, have been identified, the design rules
specified by manufacturers should be followed. According to the project drawings,
the implementation of all the components of the power station in the adequate pipe
section should be applied. Usually the PAT is installed in a by-pass to the existent
pipe system, as shown in Fig. 6.19.
In the by-pass one or more PATs can be installed, depending on the system
characteristics (Fig. 6.20). This apparatus should follow the security procedures and
can be installed (when there is sufficient space) in an existing valve chamber, or
specifically in a chamber built for this purpose [1, 11].
For the isolation of the group from the hydraulic circuit upstream and down-
stream the application of two valves is applied (e.g. a butterfly or other type
depending on the pressures) sectionalizing the hydraulic circuit (Fig. 6.21). More
than one turbine can be installed in series or in parallel depending on the required
head or flow control (e.g. for a PAT in series—see Fig. 6.22).

Fig. 6.19 Installation in a by-pass to the existent hydraulic circuit (http://www.hidropower.eu/)


134 6 Civil Engineering Design, Electromechanics …

Fig. 6.20 Proposal for the installation in a by-pass to an existing PRV of a single PAT or two in
series

Fig. 6.21 Representation of the existing solution (left) and a single PAT installation (right)

The electrical control of the opening of the downstream isolating valve to move
parallel with the network is performed according to the law of opening established
by automation to guarantee safety and adequate operating conditions.
The schematic representation shows plan and profile views of the installed
elements associated with a PAT implementation when compared with the status
quo.
6.5 Operation and Maintenance 135

Fig. 6.22 Representation of the existing solution (left) and two PATs installed in series (right)

6.5 Operation and Maintenance

A micro hydro power plant is intended to be able to function without any operator
being present. There will, however, be a supervisor who will perform a weekly
monitoring of the operating conditions. According to the recommended setting for
the command and control system, in addition to the automatic control, the plant will
be equipped with all the devices necessary to operate in local manual control. In the
automatic operation of the group, it is expected that the flow regulation will be
realized by a suitable valve for this purpose and located downstream of the pump as
a turbine diffuser. It is commanded through respective automation, which will
reboot, and stop according to an appropriate time schedule relating to the needs in
the water system or with the flow regulated by a valve and sent to the robot group.
In remote situation it will be possible to carry out several commands for the
equipment, as well as the parameterization of different quantities associated with the
operation of the system through communication with the existing robot in the valve
chamber or powerhouse.
A maintenance schedule includes these types of inspections (according to the
PAT manual):
• Routine maintenance
• Routine inspections and
• Annual inspections
136 6 Civil Engineering Design, Electromechanics …

It is recommended to shorten the inspection intervals appropriately if the


pumped fluid is abrasive or corrosive or if the environment is classified as poten-
tially explosive.
Annual inspections are the most important and include checking the PAT
capacity, pressure and power.

References

1. H. Ramos (Editor), Guidelines for design of small hydropower plants. WREAN (Western
Regional Energy Agency and Network) and DED (Department of Economic Development—
Energy Division), Belfast, North Ireland, ISBN 972-96346-4-5 (2000)
2. H. Ramos, A. Borga, Pumps as turbines: an unconventional solution to energy production.
Urban Water 1(3), 261–263 (1999)
3. H. Ramos, A. Borga, Pumps yielding power. Dam Eng. 10(4), 197–217 (2000)
4. H. Ramos, A. Borga, M. Simão, New design solutions for low-power energy production in
water pipe systems. Water Sci. Eng. 2(4), 69–84 (2009)
5. A. Williams, Pumps as Turbines: A User’s Guide (Practical Action Publishing, London, UK,
2003)
6. G. Caxaria, D. Sousa, and H. Ramos, Small scale hydropower: generator analysis and
optimization for water supply systems, in World Renewable Energy Congress (2011)
7. H. Ramos, Simulation and hydraulic transients control in small hydropower plants. Modeling
and analysis of induced effects of turbogenerator overspeed, Instituto Superior Técnico (1995)
8. H. Ramos, A.B. Almeida, Dynamic orifice model on waterhammer analysis of high or
medium heads of small hydropower schemes. J. Hydraul. Res. 39(4), 429–436 (2001)
9. Centrifugal Pump Handbook, Third Edit (Sulzer Pumps Ltd Winterthur, Switzerland, 2010)
10. H. Ramos, A. Beta⁁mio de Almeida, Parametric analysis of water-hammer effects in small
hydro schemes, J. Hydraul. Eng. 128(7), 689–696 (2002)
11. H. Ramos, A.B. De Almeida, Control of dynamic effects in small hydro with long hydraulic
circuits. Int. J. Glob. Energy Issues 24(1/2), 47 (2005)
Part III
Applications and Impacts
Chapter 7
Location of a PAT in a Water
Transmission and Distribution System

Abstract The importance of pressure control in water transmission systems with


the identification of flow energy dissipation points is presented. The numerical
methods used to analyze the behavior of the hydraulic system, the control valve
maneuvers, and the over-speed effect associated with the dynamic behavior of the
PAT are discussed. The dynamic control of the pressure induced by the water
demand which varies during the day in quasi-steady state conditions and the
pressure surge control for unsteady regimes that can occur in PAT shutdown/start-
up and valve control closure/opening, as well as in a sudden load rejection with the
associated over-speed effect, are examined. An understanding of the system
behavior in terms of pressure control and in the presence of different dynamic
effects is provided, and, in addition, the different analyses that should be made for
each PAT type and water system requirement are described.

Small scale hydropower is emerging as a decentralized source to satisfy the local


demand for electricity. The installation and operation of micro hydro solutions can
improve energy efficiency through the electricity generation and simultaneously
pressure control in water systems, particularly in cities with considerable geodesic
differences, or in irrigation areas, drainage systems and pipelines, in sections where
the pressure is greater than that which is adequate or necessary. Moreover, in certain
cases, energy dissipation is imperative to minimize leakages, achieving important
monetary savings due to the reduction of water losses, system ruptures and inter-
ruptions of the service [1–4]. This necessary dissipation or control is usually
ensured by pressure reduction valves. The replacement of these valves with
micro-power solutions could allow the use of energy that is otherwise dissipated
[5]. The adaptation of water supply systems to produce energy has the advantage of
leveraging components that already exist (i.e., tanks, pipe lines and valves).
Furthermore, a continuous supply is guaranteed by the continuous consumption
throughout the day [5–10].
Different strategies can be found in expert literature for locations to install PATs
in water systems [11–15]. Some authors have proposed their placement close to
reservoirs/tanks and at the inlet nodes of network districts. The authors in [4]

© Springer International Publishing AG 2018 139


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_7
140 7 Location of a PAT in a Water Transmission …

suggested the identification of a number of nodes with energy potential within the
entire network, developing a complex methodology which combines GIS tech-
niques and sensitivity analyses. Thus, although some authors limit the placement to
a particular node, others imply long and time-consuming analyses.
The approach of using stochastic optimization algorithms has been investigated
in relation to WSN (Water Supply Network) design and the optimal location of
control valves with interesting results [4, 16, 17].
The development of novel techniques to find the location of micro-turbines
within water systems which maximizes the annual energy production is necessary.
The technique consists of detailed analyses with the use of mathematical models
which are based on different simulating strategies, addressing pressure constraints,
flow variability and the complexity of the networks. Pressure control is a subject of
concern for water managers and has a political visibility, mainly when periods of
scarcity of water resources occur or when the water supply is not sufficient in areas
of rapid expansion.

7.1 Selecting the Site

The location of a hydro generation site is mainly conditioned by head and flow
requirements. The power potential and energy output, the estimation of the need
and the economic feasibility must be studied. Hence, the location of energy con-
verters within networks needs to be optimized with respect to energy production
and economic value.
In any pipe of a network the head in each node of the hydraulic system will be
higher or equal to a minimum pressure pmin, usually imposed by regulations to
guarantee a quality service. In each pipe section, whenever the head is higher than
this minimum pressure, there is energy in excess. This excess of energy will vary in
time, as the demand in the network is not constant, affecting both the pipe flow and
the pressure in the nodes. Even if a potential for hydropower has been recognized,
the ideal location of turbines depends on the flow rates and respective velocity
restrictions (if any exist), which have daily variations, the head, which is dependent
not only on the minimum service pressures but also on the chosen machine, and the
geometry of the network
The hydraulic energy Eh (Wh) in any point of the pipe is defined as:

Eh ¼ cQHDt ð7:1Þ

where c is the water volumetric weight (N/m3), Q is the flow rate (m3/s), H is the
head (m), meaning the total energy subtracted from the elevation of the point, and
Dt (h) the time interval. This time interval can be considered as the duration of the
considered steady state.
7.1 Selecting the Site 141

To estimate the excess energy in the upstream and downstream nodes Eq. (7.2)
is applied:
 
Eexcess ¼ c Qt Hup  Hdn Dt ð7:2Þ

Considering a pipe system, (Fig. 7.1a), excess energy exists but none is avail-
able, since the most downstream point is at the minimum pressure and extra head
losses would cause this pressure to decrease.
In (Fig. 7.1b), the minimum pressure is not limited at the end point but the
excess downstream from this point is partly available. During the network design,
there is usually an effort to minimize this difference by using smaller diameters, as a
means to control the pressure. However, the installation of turbines can also be used
to dissipate this excess energy as well as simultaneously producing energy.
To quantify how much of the excess energy is available for hydropower pro-
duction, critical points must thus be identified. The critical point corresponds to
the position in a network where the difference between the total energy and the
minimum pressure is minimal but higher than zero. This difference is the head that
can be taken from the total energy line.
Depending upon the geographical characteristics of the implantation site, the
hydropower scheme can have a high, medium or low head. High head schemes can
be less expensive because the requirements are normally associated with small
discharge flows when compared with low head schemes that need more flow for the
same power.
The experience of the authors based on several studies resulting from laboratory
tests, analyses of field applications and hydraulic simulations and optimizations
allows the conclusion that the most favorable locations in networks are the
replacements of PRVs, if any exist, and the paths of the highest available heads and
discharge flows. Dissipation upstream from water tanks is often needed and typi-
cally the discharges are high. Within the network, however a larger energy pro-
duction does not necessarily imply better economic value since the costs related to
the construction and equipment of several micro power stations can justify a lower

Fig. 7.1 Excess energy assessment a system without available energy: the downstream node has
an imposed minimum pressure not allowing for any energy extraction b pipe system with available
energy: effect of the extraction at two critical points (adapted from [11, 16])
142 7 Location of a PAT in a Water Transmission …

revenue. Finally, the installation of turbines in a network should be analyzed with


small time steps, requiring sensitivity analyses, since the implementation of pres-
sure reductions in a network has an influence on the distribution of the discharge
flows, flow directions and pressures. The risk of inducing fragility in the network
should be analyzed with reference to different consumption scenarios.

7.2 Flow Rate and Head Measurement

Hydraulic simulator models allow an analysis of the behavior of hydraulic systems,


assessing various scenarios and strategies for the operation of the real system. One
advantage of hydraulic models is their ability to perform quasi-steady simulations,
often referred to as dynamic, i.e. to determine the evolution of the system over time
on an extended time scale. To undertake this analysis, it is necessary to set the
periodicity of simulation and construct a consumption pattern. Consumption pat-
terns are placed multiplicative factors that modify the consumption of nodes at
every moment depending on the standard time step (Fig. 7.2).
A simulation in an extended period can produce the following information:
– in nodes—the hydraulic load pressure;
– in pipe branches—the flow discharge, flow velocity and head losses.
These parameters are obtained by solving the mass conservation equations for
each node (7.3) and the energy conservation equation for each Sect. (7.4). The
non-linear equations are solved by the hydraulic solver commonly using the
“Gradient Method”:

Fig. 7.2 Options in an extended period simulation


7.2 Flow Rate and Head Measurement 143

X
NPJi
aij Qj þ DMi ¼ 0 ð7:3Þ
J¼1

 n1
Hi  Hk ¼ rj Qj Qj  ð7:4Þ

in which,
Qj The flow discharge in pipe j (m3/s);
NPJi The number of pipes connected to the node i;
DMi The consumption in node i (m3/s);
Hi The hydraulic load of node i (m);
rj The friction factor to the pipe j dependent on the head losses;
n The exponent of discharge flow for the head losses.

– Hazen-Williams

e ¼ 10:7 C 1:852 d 4:87 L; n ¼ 1:852 ð7:5aÞ

– Darcy-Weisbach;

e ¼ 0:083 fðe; d; QÞ d5 L, M ¼ 2 ð7:5bÞ

being:
C The Hazen-Williams coefficient;
e The absolute roughness (mm);
f The Darcy-Weisbach factor;
d The pipe diameter (m);
L The length of the pipe (m);
Q The discharge flow (m3/s).
A knowledge of the topography levels (Fig. 7.3) is crucial to define the critical
points of total energy available and, consequently, to evaluate the potential energy
available for production.
Whenever it is the necessity to control excessive pressures in network systems,
the influence area of a pressure reducing valve, as presented in Fig. 7.4, shows the
path flow where the turbine installation can operate.
Hence, the calibration proceeds in order to verify the flows and pressures in the
entire system (Fig. 7.5). By using a trial and error methodology or by applying an
optimization algorithm, the calibration of the network can be obtained in order to
make possible the optimization of the potential energy available, the identification
of the main characteristic variables and the development of viability studies and
performance indicators.
144 7 Location of a PAT in a Water Transmission …

Fig. 7.3 Topography where the hydraulic system is implanted

Fig. 7.4 Influence area of the installed PRV and recorded discharge flow pattern
7.2 Flow Rate and Head Measurement 145

Fig. 7.5 Calibration process: flow variation and pressure at the PRV (turbine site)

Table 7.1 BEP of the selected PAT


Q (l/s) H0 (m) PH (kW) PE (kW) g N (rpm) ns (rpm) (m, kW)
(−)
4.44 58 2.5 1.56 0.631 1520 11.86

After the selection of the most suitable PAT for each place previously identified,
the rated conditions for the best efficiency point (BEP) can be defined (Table 7.1).
The behaviour of the PAT depends on its specific speed (for low values the more
radial runner shape). The characteristic curves are characterized by higher values of
the head with the increasing speed (Fig. 7.6) (the contrary for high specific speed
PATs).
Based on the specific speed, ns, and the engine torque, M, it is possible to
estimate the electrical power, PE. Through the interception between the hydraulic
system curve and the PAT curve the operating point is identified (preferentially as
close as possible to the BEP) (Fig. 7.7).
Several analyses have to be developed for different consumption hours.
A comparison between the PRV and PAT performance is presented (Fig. 7.8) to
analyze the behavior of the entire system when the PRV is replaced by a PAT.
146 7 Location of a PAT in a Water Transmission …

Fig. 7.6 Characteristic curves of the selected PAT for different rotational speeds (N) and torque
values (M)

Fig. 7.7 Operation point: interception of the hydraulic system and PAT rotation characteristic
curves

Fig. 7.8 Pressure variation for the highest demand hour of the hydraulic system
7.3 Pressure Control in Water Transmission Systems 147

7.3 Pressure Control in Water Transmission Systems

In water supply systems, micro-hydropower can be used for the energy recovery
associated with excessive pressure control. One of the main concerns of water
system managers, throughout the world, has been the minimisation of water losses,
that frequently reach values of 30% or even 60% from all the water that supplies
drinking systems. Nowadays, the problem of water losses and their control in water
distribution systems (WDSs) assumes an increasing importance in light of the
current trend of privileging the sustainability of consumption and the protection of
the environment.
In order to guarantee an adequate technical performance there needs to be a
global evaluation of the system, which includes different scenarios in terms of
operational conditions for different restrictions of each component. Hence, only by
means of an integrated analysis of the behavior of the system based on support
instruments, it will be possible to address the necessary requirements to attain the
maximum efficiency in terms of pressure control.
The need to increase the efficiency in pressurized water networks has allowed the
development of new water management strategies [1, 5, 6, 15, 17, 18]. These
strategies have been focused in two different directions according to the water
pressurized network type (i.e. pumped or gravity system). In pump solutions, the
increase in efficiency in the network is directly correlated with the reduction of the
pressure, the operation rules correction and the design of the facilities (e.g. the
pump efficiency, leakage control, and establishment of optimum schedules). In
gravity systems, the efficiency improvement has been related to the reduction of
leakage levels through the installation of pressure reduction valves [3, 4, 21–24].
References [7, 8, 17, 25–29] have proposed the replacement of PRVs by
hydraulic machines, which could also generate energy. These systems provide two
benefits in terms of increasing the efficiency: on the one hand, the operation of the
PRVs reduces the pressure in the system and therefore also the leakages; on the
other hand, the generated energy contributes to improving the global energy effi-
ciency in water networks.
With the purpose of obtaining a solution that achieves simultaneously an opti-
mization of the number and placing of valves, as well as a refinement in the
adjustment of each valve opening, for a simulation in an extended period, a
methodology has been developed which is based on the number and the location of
the valves and which allows for the adjustment of the valve opening degree [3, 4].
In water network analysis, leakage is often modelled based on the equation of
flow through an orifice:

qj = Kf pbj ð7:6Þ
148 7 Location of a PAT in a Water Transmission …

with

X
M
Kf ¼ c  0:5  Lij ð7:7Þ
J¼1

where qj is the leakage flow at node j, pj the service pressure at node j and Kf is a
fixed leakage coefficient for the node, estimated as a function of the pipe and soil
characteristics, (for the exponent ß of the pressure in the leak equation, 1.18 being
adopted), c is the discharge coefficient of the orifice, which depends on the shape
and the diameter; Lij is the pipe length between nodes i and j, and M is the number
of pipe reaches connected to the node j.
The simulation is needed for a period of 24-h, with intervals of 1 h, with the
objective of minimizing the pressure, but without letting it drop lower than the
minimum pressure admissible (e.g. 30 m) in any node of the system. The opti-
mization model results can be used by EPANET (free software) in order to provide
each user with the possibility of defining how many valves should be used and
where they should be placed. It is interesting to observe a reduction in the maxi-
mum level of the pressures, changing from 46 to 38 m., this behavior depending on
the number and the localization of the valves.
The number of valves installed is less important than their localization. An
efficient strategy for the control and minimization of pressure is a good operational
tool for leakage reduction in the networks of water supply systems, without com-
promising the system operation. When methodologies for the location and opti-
mization of the opening and adjustment of valves are associated with a control
strategy, the profits are indubitably more significant (Fig. 7.9).

7.4 Identification of Dissipation Points

Energy recovery in water systems has been studied by different authors in terms of
replacing the pressure reduction valves (PRVs) with a mini-hydropower station [1,
30–32] selecting the pump as turbine (PAT) type, and maximizing the theoretical
energy available [16, 33].
The selection of the hydraulic machine is very important in the energy recovery
process, taking into account that water managers have to guarantee the quality of
service to the users, ensuring the minimum pressure established (normally 30 m at
each consumption node).
Hence, numerous authors have analyzed the use of PATs in non-conventional
heads, where the flow and head are not constant. These have been proposed at
remote areas [34–44] and for small applications in pressurized systems of distri-
bution, which are mainly allocated for drinking systems [1–17] or irrigation [25–
28]. The use of these machines is due to the wide range of pumps existing on the
7.4 Identification of Dissipation Points 149

Press
30.00
32.00
34.00
38.00
m

Press„
30.00
32.00
Val. 37A
34.00
38.00
m

Va. 01A
Press„
30.00
32.00
Val. 37A
Val. 28A
34.00
38.00
m
Val. 31A

Fig. 7.9 Distribution of pressure for the hour of lesser and greater consumption with no pressure
control, with one valve and with four valves, respectively

current market, and the low price and high availability of PATs compared with
conventional turbines [41].
Different authors [42–47] have analyzed the characteristic curve of the head
(Q-H) and the performance as a function of the non-dimensional parameters of
hydraulic machines.
In water systems, PRVs are utilized to guarantee the standardization and control
of the pressure, dividing the water grid into pressure areas according to the
topography. The maneuvering of PRVs creates a local head loss, with hydraulic
energy dissipation, through the decrease in the outlet pressure. PRVs create areas
with a controlled pressure and flow, where it is possible to control water losses
more efficiently, allowing a faster location and response. There are three behavior
types of PRV: (1) if the pressure outlet is higher than the set value, the valve closes
creating a head loss—PRV active (Fig. 7.10a); (2) if the pressure inlet is lower than
the set value, the valve opens, decreasing the head loss—PRV passive open
(Fig. 7.10b); and (3) if the outlet pressure is higher than the inlet pressure, the valve
closes (work as a retention valve)—PRV passive closed (Fig. 7.10c).
150 7 Location of a PAT in a Water Transmission …

Fig. 7.10 PRV behavior: a Active, b Passive open, c Passive closed

In water pipe systems, in particular in the transmission pipelines, hydraulic


power is abundant and relatively constant. This characteristic identifies transmission
pipelines as potential energy sources. However, in dissipation nodes, the flows and
head have significant variations.
The knowledge of the availability of power is a critical factor necessary to
predict and define the economic benefits of converting energy dissipation into
energy production [1].
A growing interest has been focused on exploiting water for drinking or irri-
gation purposes flowing in pipes or channels to produce electricity. As reviewed by
[26], micro-hydropower schemes are an effective solution for energy recovery from
WSNs and considerable research has been carried out in the past highlighting a
series of recurrent situations within water networks from which energy can
potentially be recovered:
• excessive pressure in correspondence with inlet ducts leading into storage
reservoirs;
• excessive pressure within gravity-fed water conveyance pipes;
• the replacement of PRVs with PATs;
• dissipated potential within irrigation networks; and
• sites at the inlet or outlet of Wastewater Treatment Plants (WWTPs) charac-
terized by low available net heads, albeit showing significant and constant flow
rates throughout the day.
In any pipe of a network at a certain instant with a steady flow regime, the total
energy line can be approximated by a straight line, between the available head in
each boundary node (upstream and downstream). The head in each node will be
higher or equal to a minimum pressure per min, usually imposed to guarantee a
quality service. In each point of a pipe, whenever the head is higher than this
minimum pressure, there is an excess of energy which will vary in time, as the
demand in the network is not constant, affecting both the pipe flow and the pressure
in the nodes [11, 17]. The available energy at a point in a water system can hence be
defined as the excess energy that can be extracted from the flow without causing the
pressure to drop below an imposed minimum value in any point of the system. To
7.4 Identification of Dissipation Points 151

quantify how much of the excess energy is available for hydropower production,
the excess energy points must first be identified. The excess energy point corre-
sponds to the position in a pipe system where the difference between the total
energy and the minimum pressure is minimal but higher than zero. This difference
is the head that can be taken from the total energy line (see Fig. 7.11).
Including the electric regulation by a VOS (variable operation speed) allows the
water system manager to optimize the energy recovery changing the operating point
or the pressure drop and flow values (Fig. 7.12).
In local systems with a higher pressure the installation of PRVs is necessary,
which allows the perception of introducing a local head loss into the system,
without constraints to the service or damage to the network. In Chap. 5 the possible
power plant regulation modes have been discussed, allowing the replacement of a
PRV with a PAT. If the PRV installation is successful, it can be replaced with a
PAT, which will reduce the excess of pressure and simultaneously can produce
electricity (Fig. 7.13).

Fig. 7.11 Pipe reach with


available energy: the
minimum pressure settings for
the upstream energy used and
the excess energy point
downstream show the
available energy to be
extracted

Fig. 7.12 Hydraulic and electric regulation mode


152 7 Location of a PAT in a Water Transmission …

Fig. 7.13 Scenario 1: a PRV at 2:00 AM; b PAT at 2:00 AM; c PRV at 7:00 AM; d PAT at 7:00 AM

In scenario 1, the night pressures are very high and there is a possibility of
reducing them with the installation of a bypass PRV that will work, only during the
periods of the lowest consumption and higher pressure (scenario 2—Fig. 7.14).
In scenario 2, the PRV already installed is replaced with a PAT (as in scenario 1)
and another PRV is installed that only works during the night period to control the
excess pressure (Fig. 7.14).

7.5 Dynamic Control of Pressure

Commonly, a hydraulic machine, when proposed to replace a PRV, has been a


pump working as a turbine (PAT). Numerous researchers have analyzed the
behavior of these machines in steady flow conditions. A review of the available
technology was developed by [1, 10, 11, 14, 16, 25, 28, 45–47]. An analysis of
performance and modeling in PATs was made while an examination of these
machines in water distribution networks including the design of an innovative
strategy to maximize the recovered energy when the flows vary during the day
[1–10, 12–17]. These strategies have been applied to determine and to maximize the
7.5 Dynamic Control of Pressure 153

Fig. 7.14 Scenario 2: a PRV at 2:00 AM; b PAT at 2:00 AM; c PRV at 7:00 AM; d PAT at 7:00 AM

theoretical recovered energy in both water systems for drinking water and irrigation
[11, 19, 20, 25].
However, the study of unsteady flow in these systems has been poorly analyzed
and the installation of PATs in pilot plants [5] means that it is necessary to analyze
such unsteady flow conditions in order to better estimate the overpressures that can
put the facilities at risk.
For a PAT with a low specific speed (i.e. a radial turbine) the characteristic
curves for different rotational speeds (N) and torques can be observed in Fig. 7.15.

Fig. 7.15 Characteristic curves of Multitec 32–2.1 (KSB)


154 7 Location of a PAT in a Water Transmission …

Fig. 7.16 Calibration process of a water distribution network in a hydraulic simulation (e.g.
EPANET). The topographic levels of the pipe system and the results of the pressure distribution
when a PRV or a PAT is installed

After the calibration of the flow at the entrance to the network and with a
knowledge of the topographic elevation of the pipe system, the pressure distribution
for the same daily hour with a PRV installed and a PAT replacement can be
obtained (Fig. 7.16).

7.6 Numerical Methods

7.6.1 Hydraulic Modelling

For steady state conditions the hydraulic simulation models of water transport and
distribution networks are used to help in the planning, preparation, and operating
diagnosis. The EPANET model was developed by the U.S. Environmental Agency
(EPA), which allows static and quasi-dynamic simulations of hydraulic behavior
and pressure distribution in the system. This simulator is widely accepted due to its
trustworthy results and free license, making it one of the most used water simulation
software products. EPANET uses the “Gradient Method” to obtain the equations of
continuity and energy conservation and the relationship between flow and head loss
that characterize the hydraulic equilibrium in the network. The relation between
flow-head loss in the pipe from node i to j is given by Eq. (7.8):
7.6 Numerical Methods 155

Hi  Hj ¼ hij ¼ rQnij þ mQ2ij ð7:8Þ

where H is the nodal head (m), h is the head loss (m), r is the resistance coefficient,
Q is the flow rate (1/s), n is the flow exponent, and m is the minor loss coefficient.
The flow continuity in the node is given by Eq. (7.9):
X
Qij  Di ¼ 0 for i ¼ 1; . . .; N ð7:9Þ
j

where Di is the consumption in the node (by convention, the flow that arrives at the
node is positive) (1/s). Thereby, knowing the head of the fixed nodes, it is possible
to obtain the heads, Hi and flows, Qij , of the network that satisfy the Eqs. (7.8) and
(7.9).
The unsteady flow in pressurized pipe systems with a greater length than
diameter can be analyzed, considering the one-dimensional (1D) model type,
through the mass and momentum conservation equations which derive from the
Reynolds transport theorem [48]. Water hammer equations have already been
introduced in Chap. 2. The assumptions applied in the classic one dimensional
water hammer models are [44]:
• the flow is homogenous and compressible;
• the changes of density and temperature in the fluid are considered negligible
when these are compared with the pressure and flow variations;
• the velocity profile is considered pseudo-uniform in each section, assuming the
values of momentum and the Coriolis coefficients to be constant and equal to
one;
• the behavior of the pipe material is considered linear elastic; and
• the head-losses are calculated by the uniform flow friction formula, which is
used in steady flow.
Water hammer equations can be simplified into a hyperbolic system of equations
[41, 49]. These equations can be presented as a matrix (7.10):

@U @FðUÞ
þ ¼ DðUÞ ð7:10Þ
@t @x

being:
   2   
H c
Q 0
U¼ ; FðUÞ ¼ gA ; DðUÞ ¼ JgA QjQj ð7:11Þ
Q gA H Q2

where J is the hydraulic gradient.


The solution of these equations is obtained through a discretized time interval for
each time step ‘Dt’ at a specific point of the pipe for each ‘Dx’, fulfilling the
Courant condition (Cr = 1) (7.12):
156 7 Location of a PAT in a Water Transmission …

Dx
¼c ð7:12Þ
Dt

The differential Eq. (7.10) can be transformed into linear algebraic equations,
obtaining the Eqs. (7.13) and (7.14). The application of these equations is
denominated the “Method of Characteristics” (MOC).

A  nþ1  fi1
n
Dx n  n 
C þ : Hin þ 1  Hi1
n
þ Vi  Vi1
n
þ Vi1 Vi1 ¼ 0 ð7:13Þ
c D
A  nþ1  fi1
n
Dx n  n 
C  : Hin þ 1  Hi1
n
 Vi  Vi1
n
 Vi1 Vi1 ¼ 0 ð7:14Þ
c D

where Hin þ 1 is the piezometric head in m w.c. at the pipe section i and time instant
n + 1; Vin þ 1 is the velocity in m/s at the pipe section i and time instant n + 1; Hi1n
n
is the piezometric head in mw.c. at the pipe section i−1 and time instant n; Vi1 is
n
the velocity in m/s at the pipe section i−1 and time instant n; and fi1 is the friction
factor in the section i−1 at time instant n.

7.6.2 Control Valves

The valves are system components, which are responsible for changing the flow
when its opening degree changes. Any operation in a valve modifies the opening
degree and varies the loss coefficient of the valve causing a flow variation in the
system, being one of the sources of the hydraulic transients. The closure time as
well as the valve type influence the type of water hammer (i.e. fast or slow
maneuver) for a system characterized by its diameter, length and pipe material.
For any maneuver, the loss coefficient of a valve is function of the opening
degree ðKv ðhÞÞ [50] and the behavior of the valve can be defined by Eq. (7.15):
pffiffiffiffiffiffiffi
Q ¼ Kv ðhÞ DH ð7:15Þ

where Q is the flow rate in m3/s; and Kv is the flow loss coefficient as a function
of the opening degree ðhÞ in m5/2/s. This coefficient can be obtained from the
manufacturer and DH is the head loss in the valve in m.
The ratio between the flow loss coefficient for a determined opening degree and
for the nominal flow loss coefficient (i.e. Kv0 flow coefficient for a completely
opened valve), can be defined by Eq. (7.16):

Kv
uðhÞ ¼ ð7:16Þ
Kv0
7.6 Numerical Methods 157

This ratio can also be defined as a function of the time, using the parameter s
between the flow coefficient for the determined opening degree Kv ðtÞ and the one
corresponding to the initial state Kv oð0Þ. This parameter is defined by Eq. (7.17)
[50]:

Kv ðtÞ uðtÞ
sð t Þ ¼ ¼ ð7:17Þ
Kv ð0Þ uð0Þ

When the closing or opening maneuver is total, the parameter s is one when the
steady flow is established (at the initial time) and zero when the time (t) is higher
than the closing (or opening) time (Tc).
The variation of parameters uðhÞ and sðtÞ along the maneuver time depends on
the type of valve, with consequences in the pressure values of the transient regime
due to the closure law. This parameter can be estimated by Eq. (7.18):

uðhÞ ¼ hb ð7:18Þ

where b is an integer number greater than zero. In a closure law of a valve, h is 1 or


0 when the valve is opened and closed respectively.
So, Eq. (7.18) can be rewritten:
 b
t
uðtÞ ¼ 1 ð7:18aÞ
Tc

Figure 7.17 shows different closures as functions of the b exponent. If the


exponent is one, the closure law is linear and the variation of the flow loss coef-
ficient is continuous. When the exponent is less than one, the variation of the flow
loss coefficient is higher at the end of the closure time (e.g. a diaphragm valve—
Fig. 7.17. left (f)). If the exponent is greater than one, the closure is higher at the
beginning of the maneuver (e.g. a butterfly valve), causing higher overpressures due
to the main closure occurring at the beginning when the velocity of the fluid is

1.00 1.00
(a) Gate Valve
(b) Butterfly Valve f
(c) Seat Valve
0.80 (d) Needle Valve 0.80
(e) Ball Valve
(f) Diaphragm Valve c f
d a
Av (t)/Av (0)

a
Kv (t)/Kv(0)

0.60 0.60

c e
0.40 b 0.40
b
(a) Gate Valve
(b) Butterfly Valve
0.20 e 0.20 (c) Seat Valve
d
(d) Needle Valve
(e) Ball Valve
(f) Guillotine Valve
0 0
0 0.20 0.40 0.60 0.80 1.00 0 0.20 0.40 0.60 0.80 1.00
Opening Degree Opening Degree

Fig. 7.17 Closures as a function of the valve type of flow coefficient (left) and the ratio between
area as a function of the opening degree (right) [25, 26]
158 7 Location of a PAT in a Water Transmission …

6 T EF 1.30
1.20
5 1.10
1.00
4 0.90
0.80
H/H0

Q/Q 0
0.70
3
0.60
0.50
2 0.40
0.30
1 0.20
0.10
0 0.00
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Opening Degree
H/Ho upstream H/Ho dowstream Q/Qo

1.2 T EF 1.10
1.00
1.0 0.90
0.80
0.8
0.70
H/H0

Q/Q 0
0.60
0.6
0.50
0.4 0.40
0.30
0.2 0.20
0.10
0.0 0.00
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Opening Degree
H/Ho upstream H/Ho dowstream Q/Qo

Fig. 7.18 H/H0 (PAT upstream and downstream) variation and Q/Q0 in a ball valve for turbulent
flow (up; Re = 100,000) and laminar flow (down; Re = 1000). Comparison between effective
closure % and total closure [27]

greater. Taking into account Eq. (7.18), the maneuver can be an instantaneous
closure ðb ! 1Þ, a convex closure (0 < b < 1), linear (b = 1), or a concave clo-
sure (b > 1) [51]. Although the closure law is known, depending on the opening
degree, this is not enough to establish the boundary condition, even knowing the
ratio of the free area as a function of the opening degree, which differs according to
the type of the valve. Therefore, the valve installed in a pipe system has a great
significance in terms of the generated transient.
The duration of the valve maneuver, the diameter, the type of closure law (linear
or non-linear) and the actuator type will influence the shape and values of the
piezometric line envelopes.
The effective time closure (Tef) is the real time of the valve closure (less than
the total time (TC)), which can induce a high discharge reduction, responsible for
the extreme water hammer phenomenon (see the effective closure % in Fig. 7.18).
7.6 Numerical Methods 159

This time is, mathematically, defined in Eq. (7.19) by the tangent to the point of the
curve where dq/dt is highest:

DQ
Tef ¼ dq ð7:19Þ
dt max

where q is the ratio Q/Qo (relative discharge value), DQ is the discharge variation in
the hydraulic system, and Qo is the discharge for total opening.

7.6.3 Runaway Conditions

The specific rotational speed (ns) is given by Eq. (7.20):

P1=2
ns ¼ n ð7:20Þ
H 5=4

where: ns is the specific speed of the machine in (m, kW); n is the rotational speed of
the machine in rpm; P is the power in the shaft, which is measured in (kW); and H is
the recovered head in (m). Therefore, ns is a characteristic parameter concerning the
runner shape and its associated dynamic behavior. In reaction turbines with a low
specific speed the flow drops with the transient overspeed. Conversely, for turbines
with a high specific speed the transient discharge tends to increase [37, 38, 41].
The flow across a runner is characterized by three types of velocities: the
absolute velocity of the water (V) with the direction imposed by the guide vane
blade for conventional turbines, the relative velocity (W) through the runner and the
tangential velocity (C) of the runner (Fig. 7.19).

nS (m,kW)
d1
v1 c
R 80 1
d1
a' w1 d1m
M'
c
v1
130 c1
b
w1
v' w' v1
210 c1
w1
d1m
v1
v2
250 w1 c1
v' d1m
c v1
a α 800 c
b2 w1 1
c1
a' d1m
M
β b w2 w v1 d
a2
300-700 c1
w1

Fig. 7.19 Velocity components across a reaction turbine runner (adapted from [6, 27])
160 7 Location of a PAT in a Water Transmission …

For a uniform velocity distribution assumed at the inlet and outlet of a runner,
the application of Euler’s theorem enables you to obtain the relation between the
motor binary and the momentum moment between these two sections:

BH ¼ qQðr1 V1 cos a1  r2 V2 cos a2 Þ ð7:21Þ

a and r being the angle and radius, respectively (see Fig. 7.19).
The output power in the turbine shaft is a result of the multiplication of the
binary by the angular speed x.

P ¼ BHx ð7:22Þ

where BH is the hydraulic torque in Nm and P the output power in W.


This results in the following Eq. (7.23), after some transformations to obtain the
discharge variation [37, 38, 41]:
g
Q¼A þ Bn ð7:23Þ
n

where η is the turbine efficiency, and coefficients A and B are defined by


Eqs. (7.24) and (7.25), respectively:
60 gH0
2pr2
A¼ ð7:24Þ
1
2pb0 r2 tan a0 þ 1
A2 tan b2

2pr2
B¼ 60
ð7:25Þ
1
2pb0 r2 tan a0 þ 1
A2 tan b2

depending upon the rotational speed value and the characteristic of each runner.
The subscript “o” denotes the outlet from the wicket gate; the subscripts “1” and
“2” are at the inlet to and outlet from the runner; bo is the runner height (or
free-area) in m; r is the radial distance in m; a is the angle that the velocity vector
(V) makes with the rotational velocity (C) in degrees; b is the angle that the runner
blades makes with the C direction in degrees; and A2 is the exit flow cross-section
area in square meters.
In accordance with this equation, the discharge regulation can be obtained by the
variation of bo, ao or b2. The bo variation is a fixed characteristic of a runner (related to
the height of the runner). According to [41], the ratio between the flow discharge in
runaway conditions (QRW) and the discharge for initial conditions of total opening (Q0)
QRW/Q0 tends to increase linearly with the increase of the specific speed (Fig. 7.20).
Furthermore, [37, 38, 41] determined the variations of the ratio Q/QBEP as a
function of N/NBEP for constant values of h (H/HBEP) in radial and axial conven-
tional turbine machines based on Suter parameters (Fig. 7.21) which are in
accordance with the dynamic behavior associated with the runner shape.
7.7 Pressure Surge Control 161

Fig. 7.20 Overspeed effect on the discharge variation of reaction turbines (adapted from [6, 37,
38, 41]

Pump-Turbine n s = 61 (m, kW) - Suter parameters Pump-Turbine n s = 397 (m, kW) - Suter parameters
-1.5 -2

-1.4 -1.8

-1.3 -1.6

-1.2
-1.4
Q/Q BEP
Q/Q BEP

-1.1
-1.2
-1
-1
-0.9
-0.8
-0.8

-0.7 -0.6

-0.6 -0.4
0 -0.1 -0.2 -0.3 -0.4 -0.5 -0.6 -0.7 -0.8 -0.9 -1 -1.1 -1.2 -1.3 -1.4 0 -0.2 -0.4 -0.6 -0.8 -1 -1.2 -1.4 -1.6 -1.8 -2
N/N BEP N/N BEP
h=1.4 h=1.3 h=1.2 h=1.1 h=1.4 h=1.3 h=1.2 h=1.1
h=1.0 h=0.9 h=0.8 h=0.7 h=1.0 h=0.9 h=0.8 h=0.7

Fig. 7.21 Q/QBEP as a function of N/NBEP and h for radial and axial machines (adapted
from [37])

7.7 Pressure Surge Control

7.7.1 Basic Considerations

The effective % of closure in valve maneuvers, and the start-up and shutdown of
radial and axial PATs of a smsall size (i.e. with a low inertia) will be analyzed. The
runaway conditions induced by the overspeed effect are common scenarios to be
taken into account. The overpressure and the flow cut effects are emphasized in the
162 7 Location of a PAT in a Water Transmission …

overspeed of radial machines, as well as the flow increase in axial machines. These
effects were also noted in [41] when developed for conventional turbines but in this
case the small inertia has a significant influence.
An approximate formula for upsurge estimation (DHM/Ho) has been developed
taking into account the start-up (or inertia) time of the water mass, TW (s), and the
guide vane, nozzle or valve closing time, TC (s) (Eq. 7.26):

DHM TW
¼ KC ð7:26Þ
Ho TC

with TW ¼ gH LV in which DH is the maximum head variation (m), Ho is the gross


M
o
head (m), KC is a factor that depends upon the turbine specific speed Ns, L is the
pipe length (m), V is the flow velocity (m/s) and g is the gravitational acceleration
(m/s2).
According to Lein’s formulation, the KC of Eq. (7.26) varies between 1.3 (for
higher N values) and 1.5 (for lower N values). On the other hand, Michaud’s
formula can be obtained from the same Eq. (7.26), assuming KC = 2. Due to the
runner overspeed effect in centrifugal PATs (as reaction turbines), DHM will also be
a function of the inertia of the rotating mass [37, 49], which is characterized through
the unit starting time, Tm (s), according to the Eq. (7.27):

WD2 N2R 3
Tm ¼ 10 ð7:27Þ
3575PR

in which NR stands for the rated wheel speed (r.p.m.) and PR is the rated turbine
power (kW), where the rated conditions for the turbines correspond to the maxi-
mum turbine discharge and the head of the best efficiency, and WD2 is the inertia
machine characteristic parameter (WD2 = 4gI (N m2)), with I the polar inertia
moment of the rotating mass (kg m2).
The turbine behavior characterization, as well as the dynamic response of the
hydraulic system, depend on the specific speed, Ns (m, kW), defined by Eq. (7.28):
pffiffiffiffiffiffi
PR
Ns ¼ NR ð7:28Þ
HR1:25

where HR is the rated head (m) or the head of the best efficiency.
The dynamic behavior of a PAT can be estimated based on two fundamental
dimensionless parameters (Eq. 7.29):

QRW NRW
aR ¼ and bR ¼ ; ð7:29Þ
QR NR

with QRW = the turbine discharge at the runaway speed, QR = the rated turbine
discharge.
7.7 Pressure Surge Control 163

NRW = the turbine runaway speed and NR = the rated turbine speed, all for the
rated turbine head.
These parameters depend on the turbine type [37, 38]: a low specific speed
reaction turbine has aR < or = 1 and a high specific speed has aR > 1.
Based on published information [6, 37] and, after an adequate adjustment of
available manufacturer turbine data, average values of aR and bR can be estimated
for small turbines, according to Eqs. (7.30) and (7.31):

aR ¼ 0:3 þ 0:0024 Ns ð7:30Þ

bR ¼ 1:6 þ 0:002 Ns ð7:31Þ

In accordance with case studies of micro and small hydropower plants equipped
with low inertia reaction turbines, as is the case of most PATs with low N values, it
is commonly assumed that bR is equal to 2  20%.
It has been shown that reaction turbine runaway conditions, in particular when
the specific speed is low, can induce dangerous overpressures during load rejection.
As can be seen in Fig. 7.22, for Turbines with low specific speeds the transient
discharge tends to decrease, while for higher specific speeds, the flow maintains or
increases with the transient overspeed [8].
In Fig. 7.23 the main parameters under analysis can be converted into
non-dimensional parameters, such as the discharge number (u = Q/ND3 with Q in
m3/s, the rotational speed in rpm, and D in m) and head number (w = H/N2D2 with
gH in m2/s2, N in rps, and D in m). In this figure there is a large representation of
head number (w) and discharge number (u) curves, from lower specific speed
values to higher specific speeds. The lower specific speed w–u curves also have
higher slopes when compared to the upper specific speed values. Each PAT has
different values of maximum operating efficiencies (see the BEP trend line), which
are also related to the size of each machine and the scaled hydraulics within it.

Fig. 7.22 Runner shape variation as a function of the specific speed and overspeed effect on
discharge variation [6, 37, 38]
164 7 Location of a PAT in a Water Transmission …

Fig. 7.23 Performance curves of PATs (adapted from [46, 47])

Based on the results obtained through systematic computer simulations,


Fig. 7.24 shows the dimensionless maximum upsurge or overpressure values
induced by full load rejection at the downstream end of a single uniform penstock,
DHM/H0, for bR = 2.0, as a function of aR. The symbols are defined as H0 = the
gross head (m), TC = the valve full closure time (s), TE = the pipeline elastic time
constant, TE ¼ 2Lc (s), Tw = the pipeline hydraulic inertia time constant Tw ¼ g H0
L U0

2
3
2
(s) and Tm = the unit starting time Tm ¼ WD
3575P  10
N
(s) with WD2 ¼ 4gI.

Fig. 7.24 Maximum upsurge induced by a full-load rejection based on the dynamic orifice model
technique [36]
7.7 Pressure Surge Control 165

When the runaway speed is attained in a very short time interval of order TE, the
overpressure due to the overspeed can be evaluated by the following modified
Joukowsky formula:

DHM ð1  aR Þ bR TE
¼ 2hw þ ð7:32Þ
Ho 20 Tm

with

cV TW
hw ¼ or hw ¼ ; when TC  TE ; ð7:33Þ
2gHo TC

hw being the Allievi parameter (typically hw < 1 for high-head systems).


The higher is the degree of simplification of the PAT the faster and simpler will
be the numerical resolution of the transient. Ramos and Almeida [37, 38]
demonstrated that in the presence of overspeed, a dynamic orifice equation could be
written for the turbine node, based on the turbine head-discharge equation.
Assuming a linear variation of the discharge with a rotating speed under runaway
conditions, a simplified solution of the transient equation can be found to be in good
agreement with the experimental results. Examples of the transient variations are
presented in Fig. 7.25, respectively for low and high NS runners.
For reaction turbines the simultaneous transient pressure and overspeed control
is much more difficult [36]. In order to evaluate, in a preliminary analysis, what the
turbine maximum relative overspeed is after a full load rejection, some approximate
formulas based on Tm can be used, such as
• Lein formula
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

Dn kTC DHC
¼ 1þ 1þ 1 ð7:34Þ
no Tm Ho

valid for Dn < 0.5. For PATs, k = 0.8 and TC is the control flow valve closing
time.

Fig. 7.25 Comparison between experimental and turbine model results of overpressure, discharge
and runner speed, for a low specific speed turbine (aR = 0.5; TE/Tm = 1.40; TW/TC = 0.12) and
for a high specific speed turbine (aR = 0.88; TE/Tm = 1.50; TW/TC = 0.12) [41]
166 7 Location of a PAT in a Water Transmission …

• Hadley formula
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

Dn kTC DHC 3=2
¼ 1þ 1þ 1 ð7:35Þ
no Tm Ho

where k is now related to the PAT characteristics as a function of the specific


speed Ns, as indicated in Table 7.2, and TC is the flow control valve time of
closure.
Typically, Dn/no should not exceed 0.6. To govern the normal runner speed,
Hadley estimated that the following relationship between the minimum Tm and TW
should obey Tm = 3.0 TW.
Thus, the pressure variation can be obtained, as represented in Fig. 7.26.
The overspeed control and speed regulation stability in micro hydroplants,
during normal operation with load demand changes, can be improved by increasing
the Tm value through the WD2 value (the flywheel effect) or by decreasing the TW

Table 7.2 Parameter for Ns (r.p.m.) (kW, m) k


Lein and Hadley formula
76 0.98
134 0.96
286 0.84
381 0.77
573 0.66
672 0.61
763 0.57

Fig. 7.26 Load rejection and control valve closure: pressure variation (m); flow variation (l/s);
and air/vapour volume variation (no cavitation problems)
7.7 Pressure Surge Control 167

Table 7.3 Runaway conditions (adapted from [41])


PAT Type Ns (m, Normal speed NR Runaway speed Runaway discharge
kW) (r.p.m.) NRW/NR QRW/QR
Radial PAT 10–100 500–1500 1.8–2.2 0.45–1.00
Axial PAT >100 75–150 2.0–30 0.80–1.90

value (e.g. by inserting an air-vessel or by selecting a larger pipe diameter). The


advanced electric speed-load regulators can also guarantee a better regulation sta-
bility criterion.
The WD2 and TC values should be guaranteed by the PAT manufacturers and are
agreed upon at design stage. Another effective way to control the transient over-
speed and pressure in hydrosystems equipped with PATs is to equip the spiral case
with special pressure regulators or relief valves which automatically open and
discharge water when there is a pressure peak (i.e. synchronous relief valves).
In micro or small hydropower plants connected to a large national electric grid,
the stability of speed regulation does not present any problem, because the grid has
the capacity to stabilise any unstable behaviour.
In exceptional conditions the PAT runner can be forced to accelerate under the
hydraulic power until a maximum limit or runaway speed NRW that can be attained
in 3 or 5 s (as presented in Figs. 7.25 and 7.26—the first pressure peak).
For PATs with a low specific speed, the flow drops with the transient overspeed
and flow control valve position. Conversely, for PATs with a high specific speed (as
for Axial turbines) the transient discharge tends to increase [41].
In Table 7.3 the PAT discharge reduction for low specific speeds is shown.

7.7.2 PAT Shutdowns with Control Valve Closure

This section shows the flow and the rotational speed variation when a sudden
shutdown happens. Figure 7.27 shows some real test results in a radial and an axial
PAT. The values of flow and rotational speed vary rapidly from the nominal values

Fig. 7.27 Experimental pressure values over time in a PAT shutdown with a fast valve closure
maneuver: radial (left) and axial (right) (adapted from [26, 27])
168 7 Location of a PAT in a Water Transmission …

Fig. 7.28 Experimental data and simulations for a PAT start-up and control valve opening in a
radial (left) and axial (right) machine (adapted from [26, 27])

to zero. These results show the dynamic behavior of the radial and axial PAT when
a transient induced downstream reaches the runners. It is interesting to observe that
the pressure wave passing through the runners and the pressure variation upstream
and downstream are both in phase (Fig. 7.27).

7.7.3 PAT Start-Up and Control Valve Opening

The PAT start-up and the fast opening of the downstream control valve for each
PAT type induce some variations in the rotational speed of the machine, passing
from a peak in runaway conditions until it reaches the synchronous rotational speed
[27]. This effect on the flow and rotational speed induces a pressure variation
upstream and downstream of a PAT (Fig. 7.28).

7.7.4 Overspeed Effect in a PAT

Interesting conclusions can be drawn for both types of runners. Figure 7.29 presents
the obtained values of pressure, flow and rotational speed for overspeed conditions.

Fig. 7.29 H/HR, Q/QR and


N/NR over time for a radial
PAT
7.7 Pressure Surge Control 169

Fig. 7.30 Q/QR as a function


of N/NR and H/HR for an
axial machine

The flow value decreases over time in all tests induced by the runner shape asso-
ciated with the low specific speed value. This decrease in the flow is related to an
increase in the rotational speed, being the minimum flow attained for the runaway
conditions.
The experimental data has been correlated with the values of rating conditions
(QR, HR, NR). The results obtained are contrasted with those relating to the radial
machine. Given a constant value of H/HR, the flow increases when the rotational
speed increases. The cases analyzed for a constant H/HR value are shown in
Fig. 7.30.

References

1. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Energy production in water distribution


networks: a PAT design strategy. Water Resour. Manag. 26(13), 3947–3959 (2012)
2. A. Carravetta, G. del Giudice, O. Fecarotta, H. Ramos, PAT design strategy for energy recovery
in water distribution networks by electrical regulation. Energies 6(1), 411–424 (2013)
3. L. Araujo, S. Coelho, H. Ramos, Estimation of distributed pressure-dependent leakage and
consumer demand in water supply networks, in The International Conference on Advances in
Water Supply Management, pp. 119–128 (2003)
4. L. Araujo, H. Ramos, S. Coelho, Pressure control for leakage minimisation in water
distribution systems management. Water Resour. Manag. 20(1), 133–149 (2006)
5. A. McNabola, P. Coughlan, L. Corcoran, C. Power, A. Prysor Williams, I. Harris,
J. Gallagher, D. Styles, Energy recovery in the water industry using micro-hydropower: an
opportunity to improve sustainability. Water Policy 16(1), 168–183 (2014)
6. H. Ramos (Editor), Guidelines for design of small hydropower plants. WREAN (Western
Regional Energy Agency and Network) and DED (Department of Economic Development—
Energy Division), Belfast, North Ireland, ISBN 972-96346-4-5 (2000)
7. H. Ramos, A. Borga, Pumps as turbines: an unconventional solution to energy production.
Urban Water 1(3), 261–263 (1999)
8. H. Ramos, A. Borga, Pumps yielding power. Dam Eng. 10(4), 197–217 (2000)
9. F. Vieira, H. Ramos, Hybrid solution and pump-storage optimization in water supply system
efficiency: a case study. Energy Policy 36(11), 4142–4148 (2008)
10. A. Williams, Pumps as Turbines: A User’s Guide (Practical Action Publishing, London, UK,
2003)
170 7 Location of a PAT in a Water Transmission …

11. I. Samora, P. Manso, M. Franca, A. Schleiss, H. Ramos, Energy recovery using


micro-hydropower technology in water supply systems: the case study of the city of
Fribourg. Water 8(8), 344 (2016)
12. A. Carravetta, O. Fecarotta, G. Del Giudice, H. Ramos, Energy recovery in water systems by
PATs: a comparisons among the different installation schemes. Procedia Eng. 70, 275–284
(2014)
13. G. Caxaria, D. Sousa, H. Ramos, Small scale hydropower: generator analysis and
optimization for water supply systems, in World Renewable Energy Congress (2011)
14. O. Fecarotta, C. Aricò, A. Carravetta, R. Martino, H. Ramos, Hydropower potential in water
distribution networks: pressure control by PATs. Water Resour. Manag. 29(3), 699–714
(2015)
15. O. Fecarotta, A. Carravetta, H. Ramos, R. Martino, An improved affinity model to enhance
variable operating strategy for pumps used as turbines. J. Hydraul. Res. 54(3), 332–341
(2016)
16. I. Samora, M. Franca, A. Schleiss, Simulated annealing in optimization of energy production
in a water supply network. Water Resour. 30(4), 1533 (2016)
17. I. Samora, P. Manso, M.J. Franca, A.J. Schleiss, H.M. Ramos, Opportunity and economic
feasibility of inline microhydropower units in water supply networks. J. Water Resour. Plan.
Manag. 142(11), 4016052 (2016)
18. I. Kougias, T. Patsialis, A. Zafirakou, N. Theodossiou, Exploring the potential of energy
recovery using micro hydropower systems in water supply systems. Water Util. J 7, 25–33
(2014)
19. M.A. Pardo, J. Manzano, E. Cabrera, J. García-Serra, Energy audit of irrigation networks.
Biosyst. Eng. 115(1), 89–101 (2013)
20. E. Cabrera, R. Cobacho, J. Soriano, Towards an energy labelling of pressurized water
networks. Procedia Eng. 70, 209–217 (2014)
21. A. Dannier, A. Del Pizzo, M. Giugni, N. Fontana, G. Marini, D. Proto, Efficiency evaluation
of a micro-generation system for energy recovery in water distribution networks, 2015 Int.
Conf. Clean Electr. Power, 689–694 (2015)
22. M. Giugni, N. Fontana, A. Ranucci, Optimal location of PRVs and turbines in water
distribution systems. J. Water Resour. Plan. Manag. 140(9), 6014004 (2014)
23. M. Abbott, B. Cohen, Productivity and efficiency in the water industry. Util. Policy 17(3–4),
233–244 (2009)
24. L.S. Araujo, H. Ramos, S.T. Coelho, Pressure control for leakage minimisation in water
distribution systems management. Water Resour. Manag. 20(1), 133–149 (2006)
25. M. Pérez-Sánchez, F. Sánchez-Romero, H. Ramos, P. López-Jiménez, Modeling irrigation
networks for the quantification of potential energy recovering: a case study. Water 8(6), 234
(2016)
26. M. Pérez-Sánchez, F. Sánchez-Romero, H. Ramos, P. López-Jiménez, Energy recovery in
existing water networks: towards greater sustainability. Water 9(2), 97 (2017)
27. M. Pérez-Sánchez, Methodology for energy efficiency improvement analysis in pressurized
irrigation networks. Practical application, Universidad Politècnica de València (2017)
28. M. Valadas, H. Ramos, Use of pumps as turbines to profit the available energy in irrigation
system, Water Resour. J.–APRH Novemb. 24(3), 63–76 (2003)
29. F. Vieira, H.M. Ramos, Optimization of operational planning for wind/hydro hybrid water
supply systems. Renew. Energy 34(3), 928–936 (2009)
30. F. Vieira, H. Ramos, Optimization of the energy management in water supply systems. Water
Sci. Technol. 9(1), 59–65 (2009)
31. H. Ramos, A. Almeida, Small hydropower schemes as an important renewable energy source,
Publicação com Ref. Int. Hidroenergia 99 (1999)
32. H. Ramos, A. Borga, Application of pumps in water supply systems for energy production,
WIT Trans. Ecol. Environ. 40 (2000)
References 171

33. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Pump as turbine (PAT) design in
water distribution network by system effectiveness. Water (Switzerland) 5(3), 1211–1225
(2013)
34. H. Ramos, M. Mello, P.K. De, Clean power in water supply systems as a sustainable solution:
from planning to practical implementation. Water Sci. Technol. Water Supply 10(1), 39–49
(2010)
35. M. Chaudhry, Applied Hydraulic Transients, 3rd edn. (New York, 2013)
36. C. Mataix, Turbomáquinas Hidráuilcas: Turbinas hidráulicas, bombas y ventiladores,
(Madrid, 2009)
37. H. Ramos, A.B. Almeida, Parametric analysis of water-hammer effects in small hydro
schemes, J. Hydraul. Eng. 128(7), 689–696 (2002)
38. H. Ramos, A.B. Almeida, Dynamic orifice model on waterhammer analysis of high or
medium heads of small hydropower schemes. J. Hydraul. Res. 39(4), 429–436 (2001)
39. H. Ramos, D. Covas, A. Borga, D. Loureiro, Surge damping analysis in pipe systems:
modelling and experiments. J. Hydraul. Res. 42(4), 413–425 (2004)
40. H. Ramos, M. Simão, A. Borga, Experiments and CFD analyses for a new reaction
microhydro propeller with five blades. J. Energy Eng. 139(2), 109–117 (2013)
41. H. Ramos, Simulation and control of hydrotransients at small hydroelectric power plants
(Ph.D. Thesis IST). Tech. Univ. Lisbon, Portugal (1995)
42. S. Derakhshan, A. Nourbakhsh, Experimental study of characteristic curves of centrifugal
pumps working as turbines in different specific speeds. Exp. Therm. Fluid Sci. 32(3), 800–807
(2008)
43. A. Nourbakhsh, S. Derakhshan, E. Javidpour, A. Riasi, Centrifugal & axial pumps used as
turbines in small hydropower stations, in Hidroenergia 2010: International Congress on
Small Hydropower International Conference and Exhibition on Small Hydropower (2010),
pp. 16–19
44. A.B. Almeida, E. Koelle, in Fluid Transients in Pipe Networks (Amsterdam, Elsevier ed,
Computational Mechanics Publications)
45. M. Arriaga, Pump as turbine—a pico-hydro alternative in Lao People’s Democratic Republic.
Renew. Energy 35(5), 1109–1115 (2010)
46. P. Singh, Optimization of the internal hydraulic and of system design in pumps as turbines
with field implementation and evaluation (2005)
47. S. Rawal, J. Kshirsagar, Numerical simulation on a pump operating in a turbine mode, in
International Pump Users Symposium (2007)
48. F.M. White, Fluid Mechanics (McGraw-Hill, Sixth edit, 2008)
49. M. Chaudhry, in Applied Hydraulic Transients, 2nd edn. (1987)
50. J. Abreu, R. Guarga, J. Izquierdo, in Transitorios y oscilaciones en sistemas hidráulicos a
presión. Valencia: U.D. Mecánica de Fluidos. Universidad Politécnica de Valencia (1995)
51. N. Subani, N. Amin, Analysis of water hammer with different closing valve laws on transient
flow of hydrogen-natural gas mixture. Abstr. Appl. Anal. 2015, 12 (2015)
Chapter 8
PAT System Economic Analysis

Abstract PATs are a viable and low-cost solution suitable for implementation in
water networks. The basic economic parameters, the time value of money, methods
of economic evaluation and the effects of the lifetime cycle are presented. The
determination of analytical formulas to predict the PAT purchase price is shown.
The installation of a PAT as a replacement of or complement to a pressure reducing
valve is presented in order to assess the producible energy and minimize the plant
economic payback time. The reader will be supplied with the theoretical basis of the
economic parameters used to estimate the cost of a PAT, together with methods for
economic evaluation and a definition of the best economic solution.

The final decision on whether or not a Pump as Turbine (PAT) scheme should be
constructed or the selection of the best design solution is generally based on a
comparison of the expected costs and benefits in relation to the useful life of the
project, assessed by means of economic analysis criteria. These analyses should be
performed together with feasibility studies as there is no guarantee that a project
suitable from a technical point of view is also advantageous from an economic point
of view. The effectiveness of the economic analysis as a decision tool for an
investor depends on the accuracy of the project cost and benefits estimates.
PATs are an unconventional solution for hydro power generation, suitable in
many scenarios when a conventional turbine unit would not be economically viable.
The use of PATs has proven to have several advantages over other turbine types,
namely compact dimensions, mass manufacturing, availability and short delivery
time, easy installation, operation and maintenance, reduced investment cost and a
longer lifespan compared to traditional turbines.

© Springer International Publishing AG 2018 173


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_8
174 8 PAT System Economic Analysis

8.1 Fundaments

In order to assess the feasibility of any small hydropower project an economic


analysis is needed. This analysis is usually based on an estimation of the costs and
revenues over an expected period of operation. The profitability is then evaluated
considering economic indexes that allow a comparison between different solutions
[1, 2].
The costs of a PAT scheme can be grouped in the three following categories:
• the capital costs, which may be defined as the sum of all expenditures required
to realize the micro hydro project;
• the annual operation costs resulting from the exploitation and maintenance of
the scheme during its useful life; and
• the repositioning costs related to the substitution of the equipment.
Within the capital costs it is important to consider studies and design, supervi-
sion during the construction, civil works, equipment (including hydromechanical,
electromechanical and electrical equipment) and the interconnection to the electrical
grid in the case of any sale of the energy.
An item for contingencies or unforeseen costs should also be included, which is
intended to cover the uncertainties associated with civil works from underestimated
excavation or land preparation works. The study, design and supervision costs
during the construction can be expressed as a percentage of the civil work and
equipment costs.
The civil work costs are small and can be deduced from the design. To evaluate
the equipment costs, budget prices from the suppliers should be obtained. This kind
of scheme can be explored in an abandoned mode, with additional equipment
required (e.g. smart control systems). The maintenance costs include two entries,
one related to the civil works and the other to the equipment. The former is
generally the smaller, representing from 0.25 to 0.5% of the capital costs of the civil
works, while the latter can reach about 2.5% of the respective capital costs.
From the investor’s point of view, the benefit in a PAT scheme is the annual
income from the energy production, which depends on the amount of energy
produced during the lifetime and on the energy sale contract conditions and tariff
policy, which are specific in each country.
The following variables can be defined [1]: n—the number of the project lifetime
periods (generally measured in years and equal to the duration of the legal permits);
r—the discount rate (constant throughout the project’s useful life); Ci—the capital
costs in year i; Oj—the annual operation cost for year j; Rj—the revenue in year j;
Pm—the repositioning costs foreseen for year m; and E—the  mean annual
production.
If the capital costs occur in the first k years, the energy production takes place
from year (k + 1) until year n and the equipment repositioning is foreseen for
year m, the following relations are obtained for the beginning of the n years period:
8.1 Fundaments 175

X
k
Ci
C¼ ð8:1Þ
i¼1 ð1 þ rÞi
Pn Oj
j¼k þ 1 ð1 þ rÞj
O¼ k
ð8:2Þ
ð1 þ rÞ
Pn Rj
j¼k þ 1 ð1 þ rÞj
R¼ k
ð8:3Þ
ð1 þ r Þ

Pm
P¼ ð8:4Þ
ð1 þ r Þm

The numerator of expressions (8.2) and (8.3) provides the present value (for the
beginning of year k + 1) of an annuity that will occur during n − k years, from
year k + 1 until year n. The denominator indicates the transference of the previous
value from year k to the beginning of year 1 (Fig. 8.1) [1].
Equation (8.2) considers that the portion of the annual operation costs related to
the legal permits will occur only when the scheme starts to operate, which is
generally a sufficiently realistic scenario. As the annual benefits in a small hydro-
power scheme are generally evaluated on a constant annual production basis (the
 formula (8.3) becomes:
mean annual production, E),
ðnkÞ
 ð1 þ rÞ nk1
R ð1 þ rÞ r
R¼ ð8:5Þ
ð1 þ rÞk

where R  is the mean annual benefit obtained by multiplying the E  by the unitary
sale price of the energy in the first year of the period under analysis (year 1).

Fig. 8.1 Compound of present value operations [1]


176 8 PAT System Economic Analysis

8.2 Time Value of Money

Based on the costs from manufacturers the PAT price per kW is calculated by
dividing the purchase cost of the pump and generator unit by the rated power at the
Best Efficiency Point (BEP), as shown in Fig. 8.2 [3–6]. Different colors have been
used to distinguish PATs coupled to an asynchronous generator having one, two or
three pairs of magnetic poles (pp), respectively.
As could be expected, PATs linked to a generator with a higher number of
magnetic poles also have a higher specific purchase cost.
There is a relation between the parameter Q√H and the price of centrifugal
pumps [4–6]. By plotting graphically such a parameter, a significant linear corre-
lation is presented, relating the Q (m3/s) and H (m) nominal values of a PAT to its
purchase price. For every PAT (and generator with a number of pairs of poles
pp = 1, 2 or 3) a best linear fit of cost function is derived, and maximum and mean
differentials are obtained, as seen in Figs. 8.3, 8.4, and 8.5 [5].
Another important parameter to assess the convenience of energy systems is the
Levelized Cost of Electricity (LCOE), defined as the ratio between the sum of all

€/kW pp = 1 pp = 2 pp = 3
10,000

1,000

100

P (kW)

Fig. 8.2 PAT unit cost versus power at BEP (0 < P < 200 kW)

y = 17,640.008x + 488.937
€ R² = 0.964
25000
20000
15000
10000
5000
0
0 0.2 0.4 0.6 0.8 1

Fig. 8.3 Cost versus Q∙RADQ(H) for PATs with pp = 1 (N = 3020 rpm in turbine mode,
Confidence range: ±35%; Mean variance: ±10.1%
8.2 Time Value of Money 177

€ y = 14.767,476x + 942,082
R² = 0,974
40000

30000

20000

10000

0
0 0.5 1 1.5 2

Fig. 8.4 Cost versus Q∙RADQ(H) for PATs with pp = 2 (N = 1520 rpm in turbine mode)
Confidence range: ±25%; Mean variance: ±9.3%

€ y = 17,939.572x + 1,235.877
R² = 0.976
25000
20000
15000
10000
5000
0
0 0.2 0.4 0.6 0.8 1

Fig. 8.5 Cost versus Q∙RADQ(H) for PATs with pp = 3 (N = 1005 rpm in turbine mode,
Confidence range: ±22%; Mean variance: ±7.9%)

costs over the lifetime of a determined production technology and the total
generated electrical energy. For the proposed installation, considering the PAT
lifetime to be 20 years and supposing a discount rate equal to between 4 and 8%, a
LCOE value of 0.025 €/kWh is a typical value for a PAT installation, which is
significantly lower when compared to other energy generation technologies, as shown
in Fig. 8.6. In general, the cost of the electromechanical devices (PAT + generator)
depends on the power plant size (as shown in Fig. 8.2). This cost, in accordance with
the present cost of electro-pumps, is much lower than the cost of a traditional turbine
group. At the present stage of PAT technology exploitation, the cost of the elec-
tromechanical devices in existing plants (see Chap. 9) has been found to be higher,
but still much lower than the cost of the traditional turbine groups.
The production of energy in the water distribution network can reduce signifi-
cantly the costs of the exploitation of and demand for external energy, becoming
auto-sufficient and eco-friendly. The energy production depends on the usable flow
of the daily demand pattern. The energy produced can be achieved by Eq. (8.6):
178 8 PAT System Economic Analysis

Fig. 8.6 LCOE of different


electricity generation
technologies

X
Energy ¼ Pu Dt ð8:6Þ

where, Energy is the energy (kWh), Pu is the power (kW) and Dt is the time interval (h).
The discount rate, , allows the assignment of a value to a monetary flux that
occurs at a different moment. The value of a monetary unit in year will be changed
in the present to monetary units [1].
The discount rate, r, allows the assignment of a value to a monetary flux that
occurs at a different moment. The value of a monetary unit in year n will be
changed in the present to 1=ð1 þ rÞn monetary units [1].
The present value (PV) gives the sequence of monetary fluxes referred to the first
year of the analysis period (Eq. 8.7):

X
n
1 ð1 þ rÞn 1
PV ¼ C ¼C ð8:7Þ
i¼1 ð1 þ r Þi ð1 þ r Þn r

8.3 Methods of Economic Evaluation

The evaluation of the profitability of a PAT scheme can be based on economic


indicators or parameters. Among these parameters we can identify four important
ones: the net present value (NPV), the benefit/cost ratio (B/C), the internal rate of
return (IRR) and the payback period (T).
As the results of the economic analysis are so sensitive to the value considered
for the discount rate, special attention should be paid in the establishment of this
value. However, in most situations, there is no well-defined and unique discount
8.3 Methods of Economic Evaluation 179

rate but a range of possible discount rates, making it advisable to perform a sen-
sitivity analysis to different discount rates [1].

8.3.1 Net Present Value (NPV)

The net present value, NPV, represents the cumulative sum of all expected benefits
during the lifetime of the project minus the sum of all its costs during the same
period, both expressed in terms of present values:

NPV ¼ R  C  O  P ð8:8Þ

If the NPV is negative, the project should be rejected as it is expected that the
benefits during its lifetime will be insufficient to cover the project costs. Assuming
that there are no restrictions with respect to the required initial capital availability,
among projects or alternative design solutions with a positive NPV, the best
investments will be those with a greater NPV.

8.3.2 Benefit/Cost Ratio (B/C)

The benefit/cost ratio, B/C, compares the present values of the benefits and costs of
the hydropower scheme on a ratio basis. This index can be defined as the ratio
between the present values of the net annual benefits and capital and repositioning
costs or as the ratio between the present values of the benefits and of the total costs,
that is to say

RO
B=C ¼ ð8:9Þ
CþP

The B/C parameter has much popular appeal since it gives an immediate indi-
cation of the “degree” of desirability of a project. If the benefit/cost is less than one,
the project is evidently undesirable. If it is exactly one the project has a marginal
interest and if it is greater than one, its implementation would seem justified, by as
much as the B/C is higher. A unitary B/C ratio implies a NPV equal to zero.

8.3.3 Internal Rate of Return (IRR)

The internal rate of return, IRR, is defined as the discount rate that makes the net
present value, NPV, equal to zero, and can be obtained by
180 8 PAT System Economic Analysis

Pn 1
j¼k þ 1 1 þ IRR j ðRj  OjÞ X
k
ð Þ 1 Pm
NPV ¼  Ci  ¼0
ð1 þ IRRÞ k
i¼1 ð1 þ IRRÞ i ð1 þ IRRÞm
ð8:10Þ

A discount rate equal to the IRR implies a unitary B/C ratio and a null NPV.
A process of trial and error can provide the solution of Eq. (8.10).
Among alternative design solutions with different internal rates of return the best
will be the one with the greatest IRR. If the achieved rates are greater than the
opportunity costs of the capital, the solutions are advantageous from an economic
point of view.

8.3.4 Payback Period (T)

The payback period or recovery period, T, represents the number of years it takes
before the cumulative forecasted cash flows equal the initial investment. Its value is
provided by the year when the cumulative cash flow (B-C) changes from a negative
value to a positive value.

8.3.5 Economic Feasibility

Once the characteristic curves of PATs to be installed have been determined, the
definition of the hydraulic power and the energy produced in the pipe system will
be addressed.
In Fig. 8.7 the total energy production per hour of a PAT operation is presented,
in the different regulation modes (no regulation (NR), hydraulic regulation
(HR) and through flow control and hydraulic and electric regulation (HER) with
flow and rotational speed control).
The economic analysis is performed for a period of 40 years and envisages the
PAT and valves (e.g. PRV) i.e. with electromechanical equipment replacement at
the year 20. Let us consider the energy selling price to be 0.09 €kW. In the
maintenance costs the following percentages are applied to the investment costs:
0.5% to civil construction, 2.5% to electric equipment and 1.5% to mechanical
equipment. The discount rates applied for the economic analysis are 6, 8, and 10%.
These discount rates are considered appropriate to estimate the time value of the
money, the risk of future cash flows and the cost of opportunity. In Fig. 8.8 the cash
flows for the different discounted rates are presented. In Table 8.1 an example of an
economic evaluation is shown.
8.4 Effects of the Lifetime Cycle 181

Fig. 8.7 Energy production

Fig. 8.8 Economic analyses

8.4 Effects of the Lifetime Cycle

The water industry is the 4th largest energy intensive sector in several EU countries
[7–14]. Most of the electricity used to treat and supply water is sourced from fossil
fuels, with an average carbon footprint of 483 g CO2 equivalent per kWh (g CO2
eq./kWh) consumed [11–14].
Overall, the UK water industry is responsible for 5 million tonnes of CO2
emissions annually [13], and a reduction in the demand for fossil-based electricity is
182 8 PAT System Economic Analysis

Table 8.1 Main parameters of the economic analysis


NR HR HER
Total investment (€) 19,255 25,120 47,202
Total energy production (kWh day) 457 1329 1402
Total energy production (kWh year) 166,925 485,084 511,575
IRR (%) 30 79 57
Discount rate (%) 8 8 8
Discounted cash flows (€) 119,968 453,554 447,257
B/C (–) 3.2 8.2 5.8
T (years) 5 2 2

a sustainability key objective in terms of economics, resource efficiency and


environmental responsibility. Water companies often have to respond to govern-
ment regulations that state that utility suppliers must monitor and reduce green-
house gas (GHG) emissions [7].
Renewable energy can provide an important solution to help water companies to
meet their GHG emission targets and provide long-term sources of energy for water
treatment and supply. In Europe, hydropower is considered the most suitable
technology for the water sector to adopt for the generation of electricity.
Micro-hydropower (MHP) installations have recently been identified as an area of
growing interest for water companies as they involve energy recovery within the
water infrastructure [8, 9]. These sites are located throughout the water infras-
tructure where excess pressure exists and can generate between 5 and 300 kW. In
addition to generating electricity, MHP installations can help to optimize a network
by acting as a mechanism for flow control, pressure management and subsequently
reducing water losses through leakage [8].
Higher efficiencies and longer payback periods lead to higher investment costs.
From Fig. 8.9, it can be found that most of the points obtained are above the water
turbine curves, except for some points with power values below roughly 10 kW,
when the payback period (S) is six years (i.e., non-cost-effective situations).
Due to the large existing market for water pumps, they tend to be less expensive
than conventional water turbines (Fig. 8.9a). In this analysis, the operational and
maintenance costs have not been taken into account since these costs have less
significance than the investment costs. Figure 8.9b shows that the cost effectiveness
of the system decreases with shorter payback periods. For a payback period of six
years, when compared with a conventional water turbine, the system is only eco-
nomically cost effective for a power output above 10 kW. With increasing power
outputs, higher initial investments are permitted, since the amount of energy pro-
duced and sold to the grid is higher, increasing the project’s profits. An increasing
trend of cost effectiveness with the increase of payback periods with varying H, Q,
and efficiency is shown in Fig. 8.10, in which the efficiencies 30, 40, 60, and 80%
correspond to the flow intervals Q1, Q2, Q3, and Q4, respectively.
8.4 Effects of the Lifetime Cycle 183

(a) Initial investment cost

Fig. 8.9 Investment costs in low-power machines for different expected payback periods [2]

Life cycle assessment (LCA) has previously been used to assess the environ-
mental impacts of renewable energy systems [10–14]. Carbon and other environ-
mental burdens of MHP installations in the water infrastructure are not reported
since there are no areas of land flooded.
The LCA of MHP installations needs to be carried out based on a number of
existing cases in practice, so facilitating the comparison of carbon and economic
payback periods. The carbon payback period is significantly shorter than the eco-
nomic payback period (typically less than 1 year). MHP presents new opportunities
to generate electricity from within the existing water infrastructure. A quantification
184 8 PAT System Economic Analysis

Fig. 8.10 Investment cost versus turbine power with varying H and Q for different payback
periods [2]

of the environmental impacts of electricity generation from three MHP case studies
(1–100 kW) in the water industry, using an LCA approach, is presented in 4.
Environmental burdens can be calculated per kWh of electricity generated over
nominal turbine operational lifespans. Compared with marginal UK grid electricity
generation in combined cycle turbine natural gas power plants, the normalized life
cycle environmental burdens for MHP electricity were reduced by: >99% for global
warming potential (GWP); >98% for fossil resource depletion potential; >93% for
acidification potential; and 50–62% for human toxicity potential. However, the
burden for abiotic resource depletion potential was 251–353% higher for MHP than
for marginal grid electricity. Different quantities of raw materials and different
installation practices led to a range in GWP burdens from 2.0 to 4.5 g CO2 eq./
kWh. When a very low site preparation requirement is compared with other sites
requiring substantial excavation works and material quantities the carbon payback
times ranged from 0.1 to 0.3 years, extending to 0.19–0.40 years for worst-case
scenarios based on a sensitivity analysis developed by [4]. The carbon payback
period for future MHP installations was estimated to increase by 1% annually, as
the carbon intensity of marginal grid electricity is predicted to decline. The study
[14] demonstrates that MHP installations in the water industry have a decidedly
positive environmental balance. LCA software, Open LCA version 1.6.1, can be
used for all calculations (Fig. 8.11), in combination with the International Life
Cycle Data System (ILCD) Handbook [14] midpoint indicators.
Studies examining the LCA of run-of-river hydro installations identified meth-
ods to reduce the carbon footprint and resource demands of these projects through
applying sustainable eco-design measures. These findings can potentially be applied
to water infrastructure sites to reduce the environmental impacts of MHP. Globally,
8.4 Effects of the Lifetime Cycle 185

Fig. 8.11 Process map and LCA system boundary for an MHP scheme [14]

the hydropower (HP) sector has significant potential to increase its capacity by
2050. A study developed by [5] quantifies the energy and resource demands of
MHP and small hydropower (SHP) projects and presents methods to reduce the
associated environmental impacts based on the potential growth in the sector. The
environmental burdens of three (50–650 kW) run-of-river HP projects were cal-
culated using LCA. The GWP for the projects to generate electricity ranged from
5.5 to 8.9 g CO2 eq./kWh, compared with 403 g CO2 eq./kWh for UK marginal
grid electricity. A sensitivity analysis took into account alternative manufacturing
processes, transportation, eco-design considerations and extended project lifespan.
These findings were extrapolated for technically viable HP sites in Europe, with the
potential to generate 7.35 TWh and offset over 2.96 Mt of CO2 from grid electricity
per annum. The incorporation of ecodesign could provide resource savings for these
HP projects: avoiding 800,000 tonnes of concrete, 10,000 tonnes of steel, and
65 million vehicle miles. Small additional material and energy contributions can
double a HP system lifespan, providing 39–47% reductions for all environmental
impact categories. In a world of finite resources, this research has highlighted the
importance of HP as a resource-efficient, renewable energy system.
The LCA results for a typical 100 kW scheme are presented in Fig. 8.12 and are
expressed as the contributions of the environmental impacts associated with activity
and material consumed in the HP installation process. While plastics account for the
186 8 PAT System Economic Analysis

Fig. 8.12 Material breakdown of different impact categories of a 100 kW micro hydropower
plant [14]

Table 8.2 GHG emissions Component MHP Lifespan (years)


of component (kg CO2 eq.) as
a fraction of the total kW Turbine and generator 4.3E−3 20–30
generated over the project Pipework 5E−4 100
lifespan) Concrete 1E−3 100
Reinforcing steel 6E−4 100
Power cable 4E−5 100
Contraction/expansion joint 3E−6 30–50

highest contribution to GWP, metals are predominantly responsible for the deple-
tion of mineral, fossil, and renewable resources.
In Table 8.2 the contribution of components to an MHP installation can be
observed (GHG emissions of component, expressed as kg CO2 eq., as a fraction of
the total kW generated over the project lifespan).
Micro-hydropower is a growing area of interest to water companies as potential
energy recovery sites can capture the excess energy within water infrastructure and
can generate between 1 and 200 kW. These data prove that the use of PATs in the
water industry has a decidedly positive environmental balance.

References

1. H. Ramos (Editor), Guidelines for design of small hydropower plants. WREAN (Western
Regional Energy Agency and Network) and DED (Department of Economic Development—
Energy Division), Belfast, North Ireland, ISBN 972-96346-4-5 (2000)
References 187

2. H. Ramos, A. Borga, M. Simão, New design solutions for low-power energy production in
water pipe systems. Water Sci. Eng. 2(4), 69–84 (2009)
3. D. Novara, H.M. Ramos, “Energy harvesting from municipal water management systems:
from storage and distribution to wastewater treatment,” IST, 2016
4. D. Novara, S. Derakhshan, A. McNabola, H. Ramos, “Estimation of unit cost and maximum
efficiency for pumps as turbines,” in Young Water Professionals, 2017
5. D. Novara, A. Carravetta, S. Derakhshan, A. McNabola, H. Ramos, “A cost model for pumps
as turbines and a comparison of design strategies for their use as energy recovery devices in
water supply systems,” in EEMODS’17, Rome (eemods17.org), Energy Efficiency in Motor
Driven Systems, 2017
6. D. Novara, A. Carravetta, S. Derakhshan, A.M. Nabola, H. Ramos, “Centrifugal pumps as
turbines cost determination and feasibility study for pressure reducing valve substitution in a
water supply system,” Submitt. to Renew. Energy, 2017
7. S. Rothausen, D. Conway, Greenhouse-gas emissions from energy use in the water sector.
Nat. Clim. Chang. 1(4), 210–219 (2011)
8. A. McNabola, P. Coughlan, L. Corcoran, C. Power, A. Prysor Williams, I. Harris,
J. Gallagher, D. Styles, Energy recovery in the water industry using micro-hydropower: an
opportunity to improve sustainability. Water Policy 16(1), 168–183 (2014)
9. A. McNabola, P. Coughlan, A.P. Williams, Energy recovery in the water industry: an
assessment of the potential of micro-hydropower. Water Environ. J. 28(2), 294–304 (2014)
10. B. Guezuraga, R. Zauner, W. Pölz, Life cycle assessment of two different 2 MW class wind
turbines. Renew. Energy 37(1), 37–44 (2012)
11. A. Pascale, T. Urmee, A. Moore, Life cycle assessment of a community hydroelectric power
system in rural Thailand. Renew. Energy 36(11), 2799–2808 (2011)
12. H.L. Raadal, L. Gagnon, I.S. Modahl, O.J. Hanssen, Life cycle greenhouse gas
(GHG) emissions from the generation of wind and hydro power. Renew. Sustain. Energy
Rev. 15(7), 3417–3422 (2011)
13. J. Gallagher, D. Styles, A. McNabola, A.P. Williams, Current and future environmental
balance of small-scale run-of-river hydropower. Environ. Sci. Technol. 49(10), 6344–6351
(2015)
14. B.M. Rule, Z.J. Worth, C.A. Boyle, Comparison of life cycle carbon dioxide emissions and
embodied energy in four renewable electricity generation technologies in New Zealand.
Environ. Sci. Technol. 43(16), 6406–6413 (2009)
Chapter 9
Application of PAT Technology

Abstract A number of existing hydropower plants are described, which have a


power generation capacity ranging between a few kW and a hundred kW, thus
proving the extreme flexibility of PAT technology. Several different designs are
shown, with the best solution depending on the power plant data and on the
problem constraints. In addition, the layout of some pilot research systems and the
main results obtained during testing are described. The reader will be able to discuss
the technical solutions with reference to power plants already realized or under
construction in water supply systems, comparing the expected and measured power
plant efficiency. Suggestions about the best layout solution for the different cases,
based on the results obtained in pilot plants, are provided.

A number of hydropower plants in water supply systems (WSS) are already using
PATs for energy production [1–11]. In many cases, these power plants were
designed to address specific site conditions [12, 13]. In other cases, the design
constraints correspond to more diffuse conditions. In all situations the choice of the
PAT working conditions is critical for the machine selection [6, 7]. The power plant
layout is fundamental in order to obtain a good plant performance. Small differences
between the real working conditions and the PAT best operating conditions could
produce a serious reduction of the plant efficiency [5, 8, 14, 13]. Examples of past
experiences are provided in order to increase the effectiveness of future plants.
In addition, a description is given of some of the research PAT test rigs
frequently cited in literature and of the main results obtained in these facilities
[9, 10, 12, 13].

© Springer International Publishing AG 2018 189


A. Carravetta et al., Pumps as Turbines, Springer Tracts in Mechanical Engineering,
https://doi.org/10.1007/978-3-319-67507-7_9
190 9 Application of PAT Technology

9.1 Working Plants

9.1.1 Malecòn, Spain

The Malecón site, in the Murcia region of Spain, was realized under the design of
Aquatec, Proyectos para el Sector del Agua. The power plant is located immedi-
ately outside the city of Murcia, along one of the three transmission pipelines
renovating the distribution network coming from the water tanks of the
Mancomunidad de Canales del Taibilla de Espinardo.
The power plant has to face a large variability of flow conditions, with a
maximum of 1000 m3/h, but a limited variability of pressure drop, between 20 and
40 m. These design conditions were addressed with a system having two PATs on
separate lines and a bypass working in parallel. A PRV was installed in only one
bypass line. The two PATs have different sizes and are centered to a different flow
rate, having a maximum power of 45 and 75 kW, respectively. The layout of the
Malecón central site is shown in Fig. 9.1 and a picture is given in Fig. 9.2.
It is interesting to observe that no dissipation valve is placed in series with the
PAT. Therefore, the pressure value at the plant outlet is determined uniquely by the
PAT characteristic curves.
A list of costs for the power plant realization is given in Table 9.1. Considering
an average energy production of 600 MWh per year, a 5–6 year payback period
was assumed.
From a technical point of view the presence of two PATs and a bypass increases
the problems connected with the power plant operation. In response to any variation
of the pipeline regime, a different configuration of the power plant has to be
automatically selected, from among the feasible operation modes. In particular, the
full incoming flow could pass along a single PAT line, or could be deviated to the
bypass line. Alternatively, the stream could also be split between the two PAT lines
and the bypass.

Fig. 9.1 Scheme of the Malecón hydro power plant


9.1 Working Plants 191

Fig. 9.2 The Malecón hydro power plant

Table 9.1 List of costs for Concept Cost


the power plant realization
PAT €61,253
Control box €54,325
Assistance €6724
Planning costs €16,852
Civil construction €19,350
Hydraulic equipment €31,056
Electrical grid (20 m) €6176
Administrative issues €6751
Others €16,751
Total €219,250

The plant is electronically regulated, has a remote connection with the main
control station in Murcia and a PLC for local work. The logic is based on fixed rules
about the flow and head at the station, normally with a fixed pressure downstream
the power plant. The PATs were tested on the site and, afterwards, the flow and
head rules were chosen to maximize the energy production.
The Malecón power plant is currently under operation and data from the plant
control system are available. In Fig. 9.3 the typical daily plant working conditions
are depicted, as measured on 4th September 2014. In the three plots the flow rate in
192 9 Application of PAT Technology

Drop

Fig. 9.3 Working conditions in the Malecón hydro power plant on 4/9/2014

each branch, the pressure head at the plant inflow and outflow, and the output
electric power are reported. The flow rate distribution was extremely variable with a
high peak at midday and a long period of null flow between midnight and the early
morning. The PATs work together, producing electricity on a total flow rate of
approximately 1000 m3/h, starting at 9:45 until 16:45. The bypass operates with an
excess flow rate of 500 m3/h. At 16:45, as a consequence of a reduction of the
incoming flow rate the 45 kW PAT is shut off and the production is limited to the
larger PAT.
In other days the pattern is much more complicated, because the logic of the
transmission line working conditions depends on the control of the Murcia city water
balance, as can be observed in the daily distribution of discharges, shown in Fig. 9.4
as measured on 20th October 2014. The logic of the PAT command seems mainly
based on the measured incoming flow, with different set-up levels for the two PAT
on/off switches. The 75 kW PAT works if the inflow is greater than 800 m3/h, while
9.1 Working Plants 193

Fig. 9.4 Working conditions in the Malecón hydro power plant on 20/10/2014

the 45 kW PAT works for inflow rates greater than approximately 900 m3/h, either
in combination with the 75 kW PAT or in a stand-alone configuration for inflows
smaller than 800 m3/h.
In the last plot in Fig. 9.5 a comparison is shown between the producer PAT
characteristic curves of each single PAT (45 and 75 kW) and the combined
characteristic curve (45 + 75 kW) and the measured working conditions on 20th
October 2014. The scatter between the measured head drop and PAT operating
head drop has to be attributed to head losses in the pipeline system of the power
plant, such as T-junctions, curves or diameter changes. The head losses seem more
relevant on the 45 kW line than on the 75 kW line. Considering that these parasite
dissipations amount to approximately 20% of the head drop recovered by the PAT,
it is evident that every effort is necessary to optimize the power plant pipe layout.
Unfortunately this is not always possible due to the size constraints.

Fig. 9.5 Measured working conditions versus PAT (Q, H) curves


194 9 Application of PAT Technology

9.1.2 San Vito di Cadore, Italy

The hydro generation site is located in San Vito di Cadore, a fashionable Italian
village in the Dolomite Alps. A map of the small San Vito WSS is plotted in
Fig. 9.6. The data on the plant were furnished by Enco Engineering consultants srl,
in charge of the technical design.
The water source is situated in the locality of Fontanes. The water stream reaches
the intermediate tank of Parié, at 1155 m of elevation, granting an optimal pressure
to the highest part of San Vito di Cadore. Next, a pipeline 600 m in length and
200 mm in diameter, supplies a second reservoir in the locality of Donarie, dom-
inating the main part of the village. An excess of hydraulic energy is present along
the pipeline at Donarie and a power plant was designed at that site to substitute an
existing pressure reducing valve, using the service building of the reservoir.
The main plant data are the following:
Qmax 17.00 l/s
Parié water level 1155.00 m.s.l.m.
Donariè water level 1078.50 m.s.l.m.
Head difference Parié-Donariè 76.50 m
Maximum head losses in water transmission 2.00 m
Net head drop 74.50 m
Maximum power 9.00 kW
Available energy per year 75,000 kWh/year
Only small variations of the flow rates are measured in the spring. Therefore, the
only design constraint for the power plant was to grant an average efficiency of

Fig. 9.6 Hydro power plant location


9.1 Working Plants 195

Table 9.2 Average Q/Qmax [–] f [%]


efficiency based on the simple
frequency distribution 0.4 10
0.6 10
0.8 30
1.0 50

0.7%, based on the simple frequency distribution of Table 9.2, accounting for the
risk of an occasional reduction in the spring capacity during dry years.
The hydraulic scheme of the hydro power plant is shown in Fig. 9.7. A vertical
multi-stage pump was used as the PAT. A plant productivity of 75,000 KWh/year
was assumed, with a total plant cost of approximately 110 K€, including 19% for
civil waterworks, 8% for electromechanical equipment, 6% for grid connection,
16% for technical costs and unforeseen eventualities, and the remaining amount for
VAT.
It is interesting to observe that the pressure reducing valves 3 and 4 are oil
operated and work in synchronism to exclude the possibility that the PAT, without
any overpressures in the pipeline in the case of a failure of the electromechanical
device or of the grid, shuts off. The purpose of the pressure reducing valve 4 is to
dissipate the excess energy.
The final layout of the power plant is shown in Fig. 9.8. In this final layout,
probably in the presence of dimensional constraints, the pressure reducing valve
was positioned immediately upstream the PAT. It is a general recommendation to
operate pumps and PATs with a suction straight line of a minimum length equal to
ten times the pipe diameter. The best practice in PAT design—the rotational shaft

Fig. 9.7 Hydraulic scheme


196 9 Application of PAT Technology

Fig. 9.8 San Vito di Cadore power plant

alignment, rotational balance, and grouting—and in PAT installation—the piping


alignment, length of suction pipe and foundation design—is a basic element to
obtain the best plant efficiency and reliability [1].
Based on tests performed on the power plant after construction, we have
obtained proof of the importance of observing such practices. In Fig. 9.9 design and
measured plant efficiencies are compared. The measured plant efficiencies were

Fig. 9.9 Design and measured efficiencies of the San Vito di Cadore power plant
9.1 Working Plants 197

found to be lower than expected in the whole range of flow rates with the exception
of the maximum WSS discharge when the pressure reducing valve was completely
open. For the smallest flow rates the discrepancy is less because the disturbance of
the PRV at the PAT inlet is also reduced.
Despite this unforeseen reduction in efficiency the power plant matched the
client requirements because the frequency distribution of Table 9.2 was centered on
the highest discharge values where the difference in the power plant performance
was smaller.

9.1.3 Capodacqua, Italy

The WSS of Pescara, a coastal city in central Italy, is supplied by a tank at 884 m
elevation. The water is partially provided by pumping at the Capodacqua tank at
850.42 m elevation. A sketch of the system is shown in Fig. 9.10. To reduce the
pumping costs a small hydro power plant was realized using the water of the Forca
Canapine spring at 1014.9 m of elevation. The hydro power plant works with a
flow rate in the range 12–35 l/s and the head drop existing between the Forca
Canapine spring and the Pescara Tank ranges between 114 and 140 m.
By a direct coupling of the hydraulic components of one pump and another used
in inverse mode, a flow rate ranging between 20 and 52 l/s can be sent to the
Pescara Tank with a positive drop of 40 m to 49 m, for the smallest and largest flow
rates, respectively.
In Fig. 9.11 the hydraulic scheme of the plant is shown. The data were provided
by Caprari spa, the company that realized the turbo-compressor used in the power
plant. It is interesting to note that, in the absence of a complete theory of a PAT
used as a component of a turbo-machinery system, the power plant was conceived

Fig. 9.10 Scheme of Pescara WSS


198 9 Application of PAT Technology

Fig. 9.11 Hydraulic scheme

assuming a range of pumping power corresponding to the available hydroelectric


power and selecting the pump on the basis of a few PAT test points. The
mechanical group used in the Capodacqua power plant is shown in Fig. 9.12.

Fig. 9.12 The Capodacqua turbo-compressor


9.1 Working Plants 199

9.1.4 Conejeras and Cartuja, Spain

These two twin power plants are placed at the entrance of the water supply network
to Granada, in Spain. All the energy generated will be used in a pumping station.
The waterworks have already been finished and only the administrative aspects of
the procedure remain to be completed.
The main plant data are the following:
– Generator power: 91 kW
– PAT efficiency: 74% for Conejeras and 84% for Cartuja
– Medium Flow: 1000 m3/h
– Medium Head drop: 36 m for Conejeras and 39 m for Cartuja
– Working hours per year: 8700 h for Conejeras and 5000 h for Cartuja
– Energy generated per year: 615 MWh at Conejeras and 450 MWh at
Cartuja
A list of costs is given in Table 9.3 for the power plants of Conejeras and Cartuja
and for a third plant realized by the Aquatec, Proyectos para el Sector del Agua,
company in Chile.
The layouts of the power plants of Conejeras and Cartuja are reported in
Figs. 9.13 and 9.14. The layout of the power plant refers to an HR regulation mode
during operation.

Table 9.3 List of cost for the power plants of Conejeras and Cartuja
Concept Conejeras plant Cartuja plant San Antonio plant
PaT €43,061 €40,973 €41,253
Control box €53,179 €51,062 €38,365
Assistance €6822 €6724 €6724
Planning costs €16,955 €16,852 €16,852
Civil construction €18,000 €18,000 €14,150
Hydraulic equipment €30,554 €29,315 €52,365
Electrical grid (150 m) €23,897 €23,823 €12,300
Administrative issues €6765 €6751 €6751
Others €17,328 €16,751 €16,751
Total €216,500 €210,200 €205,500
200 9 Application of PAT Technology

Fig. 9.13 Conejeras power plant

Fig. 9.14 Cartuja power plant

9.1.5 San Antonio, Chile

The San Antonio mini hydraulic installation has the same dimensions as those of
Conejeras and Cartuja. The plant is installed on one of the two lines supplying
water to the San Enrique network, Fig. 9.15, at the exit of the San Antonio Tank,
and it has been established as an alternative to the previous pressure reducing valve.
The main difference from the previous plants lies in the need to sustain the pressure
at a value greater than 5 bars to provide the functionality of the downstream part of
the network. The construction is close to the deadline.
9.1 Working Plants 201

Fig. 9.15 Scheme of the San Enrique WSS

The design conditions of the central installation are the following:


– Generator power: 90 kW
– PAT efficiency: 85%
– Medium Flow: 865 m3/h
– Medium Head: 43 mca
– Working hours per year: 7800 h
– Energy generated per year: 600 MWh
The list of costs is given in Table 9.3. The power plant scheme and layout are
reported in Figs. 9.16 and 9.17, respectively. The piping consists in the original
line, normally closed and used in an emergency, a second line with the PAT and a
third bypass line with a PRV to manage the highest flows at the PAT.

C - flow meter
VM - butterfly
valve
VSE - actual PRV
VS - new PRV
BFT -PAT
VC - gate valve
VMM=butterfly
valve with engine

Fig. 9.16 Scheme of the San Antonio power plant


202 9 Application of PAT Technology

Fig. 9.17 Layout of the San Antonio power plant

9.1.6 Beliche, Portugal

The Beliche dam in Portugal was built within the Águas do Algarve S.A.
multi-municipal WSS for irrigation and water supply purposes. Downstream from
this dam is the Beliche water treatment plant in which a micro-hydropower station
was installed in 2008 and connected to the grid in 2011 [2]. The micro-hydropower
plant has two PATs installed in parallel within an existing valve chamber, although
there is space for a future third PAT to be installed (Fig. 9.18). This system acts in
bypass to the main pipeline that supplies the water treatment plant.

Fig. 9.18 Location of the Beliche Dam


9.1 Working Plants 203

The design head in the Beliche hydropower station is 16.3 m and the corre-
sponding flow rate is 96.2 l/s, which is approximately the average inflow discharge.
The installed machines are horizontal single stage pumps, with a runner of 160 mm
[2]. To obtain the characteristic curves of this model, working in reverse operation,
a set of experimental tests was carried out in another water treatment plant in the
region, during the course of a master’s project in 2013 [3]. These tests were per-
formed under physical and hydraulic conditions similar to those verified in Beliche.
The first tests were performed with atmospheric downstream pressure and
considering different rotation frequencies. The electric efficiency and power
obtained in this situation are presented in Figs. 9.19 and 9.20, respectively. Since
the frequency that led to the best results (50 Hz) is also the frequency of the
national grid, new tests were carried out with 3 m of downstream pressure in order
to simulate the height difference between the turbine and the outlet connected to the
treatment plant. The results obtained are presented in Fig. 9.21.

Fig. 9.19 Efficiency curves for atmospheric downstream pressure

Fig. 9.20 Electric power for atmospheric downstream pressure


204 9 Application of PAT Technology

(a) Characteris c curve (b) Efficiency curve

(c) Electric power (d) Mechanical torque

Fig. 9.21 Test results for 1500 rpm (50 Hz) and 3 m of downstream pressure

9.2 Pilot Plants

9.2.1 Naples, Italy, Pilot Plant

The pilot plant is located in Naples, Italy, at the Department of Civil, Architectural
and Environmental Engineering (DICEA) of the University of Federico II. The
plant was realized within the ambit of the projects “WaterGRID” and
“BE&SAVE-AQUASYSTEM-SIGLOD”, with national and European funding.
The plant is constituted by a four loop laboratory network, reproducing the
operation of a small WSN. A scheme and an aerial view of the network are shown
in Figs. 9.22 and 9.23, respectively.
The laboratory water network is supplied by a pump which delivers a flow
discharge of up to 40–50 l/s at a maximum pressure head of 100 m. An air chamber
downstream of the pump operates as a small compensation tank, granting a constant
pressure head at the network inlet, and avoiding continuous pressure and flow
fluctuations. The network is made of cast iron, with a nominal diameter of 150 mm.
9.2 Pilot Plants 205

Fig. 9.22 Schematic representation of the laboratory network

Fig. 9.23 WDN laboratory network at DICEA Hydraulic Laboratory

A total of 19 motorized gate valves are installed at the network nodes. The valves
are remotely controlled by means of electric actuators, allowing the closing or
opening of the pipes for flow control.
A motorized needle valve is installed at the network inlet to regulate the pres-
sure, whereas manual butterfly valves are installed at the outlets, to regulate the
outflow. The network allows a simultaneously management of two PATs, installed
at the SS and MS nodes in Figs. 9.23 and 9.24a, b; an inverter is also installed, both
to vary the rotational speed of the impeller within the range 5–50 rps and to
206 9 Application of PAT Technology

Fig. 9.24 Detail views of the a SS and b MS nodes

maximize the hydropower generation. The data acquisition and settings are realized
by remote control, through the running of an in-house SCADA system (Fig. 9.25),
connected to the Programming Logic Controllers (PLCs) by means of a wired
connection with the Ethernet/IP protocol. PLCs are installed at the most significant
nodes of the network.
For a comprehensive measurement of the flows and pressures during the network
operation, 11 pressure transducers and 7 flow meters are deployed within the net-
work. The pressure is measured using piezo-resistive transducers with a pressure
range of 0–10 bars and an accuracy of 0.25%. The flow discharge is measured
using electromagnetic flow meters with a nominal pressure of 16 bars.
Algorithms have been developed to couple the pressure management and energy
production [15]. The set-point pressure is guaranteed at the critical node (i.e. the
node where the minimum pressure occurs), regardless of the inflow discharge and
the head losses within the network. At the same time, the maximum hydropower
consistent with the inflow and the available head is generated, in the cases both of
the inverter modulation and fixed velocity of the impeller.
Furthermore, the experimental characterization of several centrifugal pumps
operating as turbines has been performed, having different performances, geometric

Fig. 9.25 Main page of the SCADA system


9.2 Pilot Plants 207

configurations and numbers of stages. Specifically, by using the inverter, the


experimental characteristic curves have been estimated by varying the rotational
speed and the analytic formulations have been derived, as a function of the PAT
geometric characteristics [4].
The results of the laboratory experiments have been considered in both the
analysis and design of the installation of the PATs at the Naples WDN, in col-
laboration with the ABC “AcquaBeneComuneAziendaSpeciale” Water Agency of
Naples Municipality, co-promotor of the above mentioned projects. In greater
detail, their intervention has involved the Soccavo District Metered Area
(DMA) which supplies 5000 users (about 20,000 inhabitants).
The Soccavo WSN has an overall length of about 16 km, composed of cast iron
and steel pipes with diameters from 40 to 600 mm. Domestic electronic water
meters have been installed at the DMA, aimed at both monitoring and recording the
individual user’s consumption.
The water district is supplied by the Scudillo and Santo Stefano Nuovo water
tanks, with a specific head of 183 and 130 m a.s.l, respectively. The installation of
two PATs has been planned at the Via Epomeo and Via Nerva nodes, located at
89 m and 75 m a.s.l., respectively. Due to the high excessive pressure in the DMA,
the PAT installation has been considered both to reduce the pressure (and the
related water losses) and to test the reliability of small-scale hydropower generation
in urban areas (Fig. 9.26).

Fig. 9.26 Soccavo district of Naples


208 9 Application of PAT Technology

Fig. 9.27 Via Epomeo PAT manhole

The PAT installation at the Via Epomeo node (Fig. 9.27) has been completed
and its activation is currently under development.

9.2.2 Lisbon, Portugal, Pilot Plant

The pilot plant is located in Lisbon, Portugal, at the Instituto Superior Tecnico
(IST) [1]. The plant is constituted by a single pipeline equipped with a PAT,
reproducing the operation of a transmission network or of a branch of a WSN,
Fig. 9.28. A scheme of the system is shown in Fig. 9.29.
Water circulation is granted by a 4 kW pump, shown in Fig. 9.9, supplying an
air vessel, Fig. 9.10, responsible for providing a constant pressure (approximately
3 bars) to the pipeline, which has a length of approximately 110 m, and is com-
posed of two different materials.
A PAT installed at the Lab of IST aims to produce energy while reducing the
pressure between both up and downstream the PAT. PATs in serial or in parallel
can be studied too (Fig. 9.30).
9.2 Pilot Plants 209

Fig. 9.28 IST pilot plant

An induction machine has 0.55 kW of nominal power. It has 3 wires repre-


senting the three-phase system and it has no neutral wire. It can be set as star or
delta connection inside the metal case on it, following the instructions depicted on a
small paper inside the same. Its nominal speed is 910 rpm.
In order to verify different types of combinations (open/close or partial
open/close) of the emitters (such as irrigation hydrants or different DMA—district
210 9 Application of PAT Technology

Fig. 9.29 Schematic representation of the PAT laboratory network

Fig. 9.30 PAT—Etanorm 32-125 KSB and generator by an induction machine

meter areas of an urban zone), it is used five valves that would simulate the various
connections. These valves would be in some cases, fully open, fully close or
partially open/close to verify the PAT performance along different hydraulic con-
ditions. The PAT system operates in a loop pipe using a pump for the recirculating
(Fig. 9.31).
The air vessel is responsible for providing constant pressure within the network
(*3 bar). It has 1 m3 of capacity and its maximum of pressure is 4 bar (Fig. 9.32).
The pipe system is divided into two different pipe branches, being one in PVC and
another in PEAD material, as described in Fig. 9.29. Both pipes have 45 mm of
internal diameter and the PVC has a total length of 13.8 m and the PEAD 100 m.
The PVC pipe has 10 elbows (Fig. 9.32).
9.2 Pilot Plants 211

Fig. 9.31 Flow emitters and the recirculation pump

Fig. 9.32 Air vessel and pipe system


212 9 Application of PAT Technology

The plant is equipped with electronic measurement devices allowing the control
of hydraulic, mechanical and electrical parameters, in particular:
• an electromagnetic flowmeter
• pressure transducers
• a multimeter and
• a tachometer
The pilot plant is suitable both for the PAT characterization, and for the analysis
of the effect of the PAT on the network. In particular, for the length of the pipeline,
approximately 114 m, this is probably the only experimental plant allowing the
analysis of fast transients.
Wires enable the connection between the different electrical components and
sensors. Always ensure the good connection between all the equipment in order to
validate your data collection. If you want to test a connection, use the multimeter
with the sound option, to test the continuity of the wire.
The set of resistors try to simulate the load into the system. For each phase it is
used the same value of resistance. For this case is used a resistive load of 175 X for
each phase. The ideal way is to use a bank of variable resistors, which can provide a
huge range of resistance values and simplifies a lot the load regulation (Fig. 9.33).
The capacitors are responsible for different combinations of capacitance values
applied to the generator, allowing to simulate diverse levels of machine excitation.
This capacitance can be regulated through two switches that are located in each
capacitor, represented by number from 1 to 4 and 1 to 5. For higher numbers
combinations, the capacitance is higher. The capacitors can be connected in series
or in parallel. For a higher values of capacitance, the parallel connection is
advisable (Fig. 9.33).
With this switch it is possible to connect/disconnect the generator and capacitors
to the load. With the switch at “0”, the generator is only connected to the capacitors.
When you switch to the right “B-C” it connects the load to the system. With the
PAT and generator rotating, to produce electrical power, you need to change from
“B-C” to “0” (generator excitation) and after 2–3 s switch back to “B-C” (generator
is already excited). For lower rotational speeds (due to low flows) you will need

Fig. 9.33 Connection wires and resistors


9.2 Pilot Plants 213

Fig. 9.34 Capacitors and load switch for off-grid and switch for on-grid testing

higher values of capacitance applied to the generator to achieve the production of


power. For higher rotational speeds, you will need less capacitance applied to the
generator. There is a minimum of capacitance that is required to excite such gen-
erator (Fig. 9.34-top).
For an on-grid testing a specific switch (Fig. 9.34-bottom) to impose the grid
frequency to the generator and achieve the synchronous speed is necessary, being
the machine working as a motor. When this synchronous speed is exceeded by the
rotor, the machine starts to work as a generator and it is able to produce electrical
power. Before switching on directly the 3 cables of the generator and switch, the
direction that the PAT and generator rotates should be known. This is due to this
grid connection can impose the two directions of rotation, and the desired direction
214 9 Application of PAT Technology

is the one that follows the same when the system is working normally with an
imposed water flow. The direction of rotation is changed when you change one wire
connection of the 3 shown in the left Fig. 9.30 (whose connects with the 3 gen-
erator’s wires).
Responsible for measuring the electrical parameters of the generator such
parameters as Active Power (kW), Voltage (V) and Current (A), are the output of
the sensor in Fig. 9.35a. It reads three-phase system and has two different cables
with four small connection wires each one (neutral is not needed). One cable has
forceps/clips to measure the generator’s phase current and the other has connection
wires to measure phase voltage. The sensor (b) is very useful to measure and
analyse the rotor frequency increase during runaway conditions, such as discon-
nection of the capacitors from the machine (electrical transients). The sinusoidal
wave can be recorded, pressing the Run and Stop button. The data can be saved in
“.CSV” format and can be open in excel. To measure electrical data of the system,
such as resistance (ohm) and voltage (V) a multimeter is used (Fig. 9.35c).
The picoscope software allows readings of the pressure in different points of the
pipe system, such as up and downstream the PAT (Fig. 9.36a). After connecting

Fig. 9.35 a Fluke multimeter; b Oscilloscope; c Multimeter

Fig. 9.36 a Computer with Picoscope software; b Pressure transducer; c Electromagnetic flow
meter; d Photo/contact tachometer
9.2 Pilot Plants 215

the two pressure transducers (e.g. up and downstream the PAT in the respective
pipe pressure intakes) the software can be activated for pressure variation recording.
The pressure transducer sensor (b) is the hardware that allows the pressure data
collection through the Picoscope software. An electromagnetic flowmeter allows
flow measures in the pipe system (c). To measure the rotational speed of the
interconnection shaft between the PAT and generator the sensor photo/contact
tachometer (d) is used.
With the collected data, the dimensionless coefficients of flow and head, pre-
sented in Eqs. (9.1, 9.2), were obtained. Also, the rated values for flow, head,
efficiency and speed were calculated:

Q
u¼ ð9:1Þ
n  D2
gH
w¼ ð9:2Þ
n2  D 2

where u is the specific capacity, Q the flow, n the rotational speed, D the impeller
diameter, w the specific head, g the gravity acceleration and H the head.
The results obtained in Fig. 9.37 show the characteristic curve for PAT opera-
tion. While the correlation between flow and head showed good agreement
(Fig. 9.37a), the efficiency presented a higher variation. This can be explained by
the precision of power meter used to calculate the generator efficiency. Although,
both curves presented the expected behavior, with an increase of head with the
flow, and a maximum efficiency just before PAT maximum capacity.

Fig. 9.37 Dimensionless PAT characteristic curves: a Head and flow; b Efficiency and flow
216 9 Application of PAT Technology

Head (m)
4

2
N = 1100 rpm N = 720 rpm
0
1.500 2.000 2.500 3.000 3.500 4.000 4.500
Flow (l/s)

Fig. 9.38 PAT characteristic curves for different rotational speed

The PAT characteristic curve for different rotational speed are presented in
Fig. 9.38.
The capacitance applied to the machine and its output rotational speed can be
observed in Fig. 9.39. This speed is independent of the flow in the PAT and it only
depends on the chosen capacitance values. The increase of the capacitance causes a
decrease in the rotational speed. This is due to the different magnetic fields inter-
actions between the rotor and the stator, dependent on the excitation currents.
The hill chart is presented in Fig. 9.40 for PAT and global efficiencies. When
only the PAT is considered, the maximum efficiency is obtained for H/HR = 0.79,
Q/QR = 0.93 and N/NR = 0.72, while for the global operation the BEP is
H/HR = 1.53, Q/QR = 1.21 and N/NR = 1.07. It is noticeable the drop of generator
efficiency for low rotational speeds, while the PAT remains with similar values.

1200
1100
1000
900
N (rpm)

800
700
600
500
400
20 70 120 170
Capacitance ( μF)

Fig. 9.39 Capacitance versus N


References 217

Fig. 9.40 PAT hill chart: a PAT efficiency; b PAT + generator efficiency [1]

References

1. H.M. Ramos, P. Branco, R. Andrade, M. Zilhão, G. Meirelles, “PAT Pilot Station at IST -
Manual”, IST, 2017
2. I. Samora, H.M. Ramos, D. Covas, A.J. Schleiss, “Micro-generation in the multi-municipal
water supply system of Algarve. Beliche hydropower station (in Portuguese),” in SEREA - XII
Simposio Iberoamericano sobre planificación de sistemas de abastecimiento y drenaje
3. J.M. Livramento, “Micro-hydro in water conveyance system,” M.Sc. thesis, Universidade do
Algarve, 2013
218 9 Application of PAT Technology

4. F. Pugliese, F. De Paola, N. Fontana, M. Giugni, G. Marini, Experimental characterization of


two pumps as turbines for hydropower generation. Renew. Energy 99, 180–187 (2016)
5. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Energy production in water
distribution networks: a PAT design strategy. Water Resour. Manag. 26(13), 3947–3959
(2012)
6. A. Carravetta, G. del Giudice, O. Fecarotta, H. Ramos, PAT design strategy for energy
recovery in water distribution networks by electrical regulation. Energies 6(1), 411–424
(2013)
7. A. Carravetta, G. Del Giudice, O. Fecarotta, H. Ramos, Pump as turbine (PAT) design in
water distribution network by system effectiveness. Water (Switzerland) 5(3), 1211–1225
(2013)
8. A. Carravetta, L. Antipodi, U. Golia, O. Fecarotta, Energy saving in a water supply network
by coupling a pump and a pump as turbine (PAT) in a turbopump. Water 9(1), 62 (2017)
9. J. Du, H. Yang, Z. Shen, J. Chen, Micro hydro power generation from water supply system in
high rise buildings using pump as turbines. Energy (2017)
10. T. Lydon, P. Coughlan, A. McNabola, Pump-as-turbine: characterization as an energy
recovery device for the water distribution network. J. Hydraul. Eng. 143(8), 4017020 (2017)
11. M. De Marchis, G. Freni, Pump as turbine implementation in a dynamic numerical model:
cost analysis for energy recovery in water distribution network. J. Hydroinformatics 17(3)
(2015)
12. G.M. Lima, E. Luvizotto, B.M. Brentan, Selection and location of pumps as turbines
substituting pressure reducing valves. Renew. Energy 109, 392–405 (2017)
13. M.R.N. Vilanova, J.A.P. Balestieri, Hydropower recovery in water supply systems: models
and case study. Energy Convers. Manag. 84, 414–426 (2014)
14. H. Ramos, A. Borga, M. Simão, New design solutions for low-power energy production in
water pipe systems. Water Sci. Eng. 2(4), 69–84 (2009)
15. N. Fontana, M. Giugni, L. Glielmo, G. Marini, Real time control of a prototype for pressure
regulation and energy production in water distribution networks. J. Water Resour. Plan.
Manag. 142(7), 4016015 (2016)

You might also like