You are on page 1of 310

Numerical Modeling of Active Hydraulic Devices and Their

Significance for System Performance and Transient Protection

by

Qinfen (Katherine) Zhang

A thesis submitted in conformity with the requirements

for the degree of Doctor of Philosophy

Graduate Department of Civil Engineering

University of Toronto

© Copyright by Qinfen (Katherine) Zhang 2009


ABSTRACT

Numerical Modeling of Active Hydraulic Devices and


Their Significance to System Performance and Transient Protection
by
Qinfen (Katherine) Zhang
Graduate Department of Civil Engineering
University of Toronto
2009

The thesis numerically explores the use and behavior of Active Hydraulic Devices

(AHDs), creating a new capability to simulate and control a pipe system’s transient

performance.

Automatic control valves are the first type of AHDs studied in this research. Due

to the challenges inherent in the design of a pressure relief valve (PRV), the general

principles of PRV use and selection are studied along with the system’s response to the

PRV parameters. A new application of PID (proportional, integral and derivative) control

valve is envisioned that combines a remote sensor at the upstream end of a pipeline to

create a non- or semi- reflective boundary at the downstream end. Case studies show that,

with such a boundary, the reflection and resonance of pressure waves within the pipeline

are sometimes eliminated and invariably limited.

The second type of AHDs studied in this research is the governed hydro turbine,

the most complicated hydraulic component in terms of transient analysis and


waterhammer control. A complete numerical model is developed for the turbine

installations in either urban water networks or conventional hydropower generation

systems. Using the model, transient simulations for several realistic hydro projects are

presented along with various transient control measures.

ii
Acknowledgements

My recognition and heartfelt thanks to Professor Bryan W. Karney, for his continuous

and enthusiastic supervision and financial support throughout my studies at the

University of Toronto, his influence on me not only in academic aspect but also in other

aspects of my life.

I appreciate Prof. Stanislav Pejovic for his knowledge, guidance, insight,

comments and the opportunity that he brought for turbine model application in MeS

Hydro project. I am also grateful to the team Professors of my Ph.D. committee,

including Barry Adams, Brent Sleep, Heather MacLean and Christopher Kennedy, for

their encouraging and invaluable comments on my thesis and presentations. I also thank

the external reviewer of my thesis, Dr. Anton Bergant of Litostroj Power, for the time he

allocated to reading my dissertation, his enthusiasm for this work and the invaluable

suggestions he offered for strengthening it. That a researcher and scholar of his high

calibre was involved with the evaluation of my work is an honor.

During my study at University of Toronto, I enjoyed the company, friendship and

help from Yimin Zhang, Lihong Sheng, Julia Koycheva, Andrew Colombo, Bong Seog

Jung, Bahman Nasar, Yves Filion et al. My special thanks to Andrew Colombo for his

help on my written English. It was a great and valuable experience for the transition from

a new immigrant to a professional in Canada.

iv
Most importantly, accomplishing a PhD during my middle life is only impossible

due to the continuous support of my family, the direct support from my dear husband Hui

(Howard) Guo and our older son Tingxi (Tim) Guo, and the understanding of my

extended families in China.

I experienced huge personal changes and emotional waves during the years that I

was perusing my PhD. In 2007, I was heartbroken for losing and finally lost my dear

mother. While on April 23rd, 2009, just one week after my final defense, I gave birth to

my second son Aaron Guo. These are the real pain and joy of our real lives in this world;

thank God for his incredible love and blessing through this journey.

v
Contents

Abstract
Acknowledgments
Contents
List of Tables
List of Figures
Notation

1 Introduction……………………………………………………………………….....1
1.1 Active hydraulic devices and system transient control………………………...1
1.2 Thesis objectives……………………………………………………………….4
1.3 Thesis overview and layout……………………………………………….……4
1.4 Publications related to thesis research………………………………………….7

2 Pressure relief valve Selections and Transient Control…………………………...9


2.1 Introduction…………………………………………………………………...10
2.2 Types of PRVs and their applications…………………………………………11
2.3 PRV operation in pump and turbine systems………………………………….14
2.4 Case studies…………………………………………………………………...18
2.4.1 Brief description of the system………………………………………18
2.4.2 Case study 1: upstream control………………………………………18
2.4.3 Case study 2: downstream control…………………………………...22
2.5 Summary and conclusions…………………………………………………….24

3 Transient Simulation of Active Hydraulic Devices with


an Introduction to Automatic Control…………………………………………...36
3.1 Numerical method of pipe transient flow…………………………….………36
3.1.1 Governing equations and method of characteristics (MOC)………...37

vi
3.1.2 Extended MOC equations in pipe networks…………………………40
3.2 Controllers and active hydraulic devices……………………………………..43
3.2.1 Open-loop and closed-loop control………………………………….43
3.2.2 Feedback and feedforward control…………………………………..44
3.2.3 Evolution of controllers……………………………………………...46
3.2.4 Elements in a control system………………………………………...53
3.3 Generalized mathematical model for active hydraulic devices……………….55
3.3.1 Extended MOC equations……………………………………………55
3.3.2 AHD’s dynamic characteristics………………………………………56
3.3.3 PID controller equation………………………………………………57
3.4 Summary……………………………………………………………………...58

4 “Non-Reflective” Boundary Design via Remote Sensing and PID Control


Valve………………………………………………………………………………...60
4.1 Introduction to “non-reflective” boundaries…………………………………..61
4.2 Transient performances with dead-ends and valve control…………………...64
4.3 Mathematical model for conventional/local PID control valve………………69
4.3.1 Extended MOC equations……………………………………………70
4.3.2 Valve discharge equation…………………………………………….70
4.3.3 PID controller equations……………………………………………..71
4.4 Consideration and mathematical model of “non-reflective” valve opening.…74
4.5 Simulations and case studies………………………………………………….79
4.5.1 Case studies with non-zero initial flow………………………………79
4.5.2 Case studies without initial flow (static initial state)………………...86
4.6 Frequency analysis and “non-reflective” boundary verification……………...88
4.6.1 System responses to pressure oscillations with various frequencies...89
4.6.2 “Non-reflective” boundary verification using hydraulic impedance
method……………………………………………………………………97
4.7 Tuning PID controller………………………………………………………....99

vii
4.8 Summary and conclusions…………………………………………………...103

5 Fundamentals of Transients in Hydro Turbine Systems……………………….105


5.1 Hydro turbine classification…………………………………………………106
5.1.1 General classification of hydro turbines……………………………106
5.1.2 Hydro turbine classification in context of numerical modeling……108
5.2 Hydro turbine parameters……………………………………………………111
5.3 Hydro turbine characteristics………………………………………………..116
5.4 Model hill diagrams and conversion………………………………………...118
5.4.1 Francis turbines……………………………………………………..118
5.4.2 Kaplan turbines……………………………………………………..124
5.4.3 Impulse turbines…………………………………………………….125
5.4.4 Pump-turbines...…………………………………………………….127
5.5 Layout of water conveyance system in hydro systems……………………...130
5.5.1 Conventional hydropower system………………………………….130
5.5.2 Energy recovery turbines in municipal water supply system………133
5.6 Governor and control system………………………………………………..133
5.6.1 Three levels of turbine flow control………………………………..135
5.6.2 Mechanism of a speed-control-governor…………………………...137
5.7 Fundamental knowledge of synchronous turbine-generator units…………..139
5.7.1 Load distribution and unit operation with a governor……………...139
5.7.2 Startup and load acceptance………………………………………..142
5.7.3 Shutdown and load rejection……………………………………….143
5.7.4 Speed-no-load (SNL) and runaway condition…...…………………144
5.8 Features of transients caused by governed turbines…………………………145
5.8.1 Key features of turbine transient analysis…………………………..145
5.8.2 Speed variation……………………………………………………...146
5.9 Summary.……………………………………………………………………148

viii
6 Hydro Turbine Modeling and Model Verification……………………………..149
6.1 Mathematical model of turbine boundary condition………………………...149
6.1.1 Head balance equation without turbine-attached valves……………150
6.1.2 Head balance equation with turbine-attached valves……………….153
6.1.3 Speed change equation (torque balance equation)………………….157
6.1.4 Governor equation………………………………………………….159
6.2 Combination of basic equations for different scenarios……………………..161
6.3 Computer implementation…………………………………………………...164
6.3.1 Steady-state simulation……………………………………………..164
6.3.2 Transient simulation and flow chart of programming……………...164
6.3.3 Data input and output……………………………………………….166
6.4 Model and program verification……………………………………………..166
6.4.1 Comparison with Wylie’s simulation……………………………….166
6.4.2 Comparison with field measurement……………………………….172
6.5 Summary………………………………………………………………….....177

7 Case Study 1: Energy Recovery Hydro Turbines in Las Vegas Water Supply
System……………………………………………………………………………..178
7.1 Project background and objectives…………………………………………..179
7.2 System and hydraulic characteristics………………………………………..181
7.3 Transient scenarios and worst case identification…………………………...183
7.4 Simulation and field tests on turbine load rejections………………………..185
7.4.1 Field tests…………………………………………………………...185
7.4.2 Simulation vs. field test……………………………………………..186
7.5 Effect of hydrant flow on turbine unit operation…………………………….196
7.5.1 Upstream hydrants………………………………………………….196
7.5.2 Downstream hydrants………………………………………………204
7.6 Summary and conclusions…………………………………………………...210

ix
8 Case Study 2: Evaluation of Transient Performance and Control Measures for
Bone Creek Hydropower Project.…………………………………………..…...212
8.1 Introduction to Bone Creek project………………………………………….213
8.1.1 System description and data………………………………………...213
8.1.2 Objectives of transient analysis.……………………………………216
8.1.3 Simplifications and assumptions made in transient analysis……….217
8.2 Conventional open surge tank scheme………………………………………218
8.2.1 Transient performance of original system design…………………..218
8.2.2 Sensitivity analysis of surge tank distance………………………….225
8.2.3 Sensitivity analysis of wicket-gate closure time……………………230
8.2.4 Sensitivity analysis for surge tank design parameters.……………..236
8.3 Air-cushioned surge chamber scheme……………………………………….240
8.3.1 Design parameters & input data of simple cylinder air surge tank…241
8.3.2 Transient performance with simple cylinder air surge tank………...243
8.3.3 Sensitivity analysis of air chamber height………………………….243
8.3.4 Transient performance with a pipe-like air surge tank……………...248
8.4 Summary and conclusions…………………………………………………...253

9 Thesis Summary and Suggestions for System Transient Performance


Evaluation…...…………………………………………………………………….256
9.1 Perspectives on system transient performance evaluation…………………..258

References………...…………………………………………………………………....262

Appendix Ⅰ: Turbine Characteristics Input Data Sample and Variable Index…269

Appendix Ⅱ: Turbine Characteristics Output Data Sample……………………...275

Appendix Ⅲ: Turbine Boundary Input Data Sample and Variable Index……….279

Appendix Ⅳ: Turbine Boundary Output Data Sample and Variable Index……..287

x
List of Tables

Table 5.1 Classification of hydro turbines

Table 6-1 System data of Wylie’s example

Table 6-2 System data of MeS hydro project

Table 7-1 Linden turbine-generator parameters

Table 8-1 System data of Bone Creek hydro project

Table 8-2 Sensitivity of surge tank distance to transient performance

Table 8-3 Sensitivity analysis of simple tank diameter

Table 8-4 Sensitivity analysis of feeder and connector pipe diameter

Table 8-5 Sensitivity of cylinder air surge tank height to transient performance

Table 8-6 Sensitivity of air chamber pipe length

xi
List of Figures

Figure 1.1 Complexity of hydraulic devices

Figure 2.1 Operation of main valve and PRV

Figure 2.2 Synchronous operation of turbine and PRV

Figure 2.3 PRV installations in a simple pipeline system

Figure 2.4 System transient performance varying with the design of upstream PRV

Figure 2.5 Sensitivity analysis of upstream PRV parameters

Figure 2.6 System transient performance varying with the design of downstream PRV

Figure 2.7 Sensitivity analysis of downstream PRV parameters

Figure 3.1 MOC grid for single pipe

Figure 3.2 Generalized node with an external flow

Figure 3.3 ON/OFF control

Figure 3.4 Proportional control

Figure 3.5 Integral control derived from the area under the control error curve

Figure 3.6 Illustration of the need for derivative control

Figure 3.7 Loop of PID-control-valve system

Figure 3.8 Active hydraulic device in a pipe network

Figure 4.1 Scheme of a branched system and initial pressure head

Figure 4.2 Comparison of transient responses to dead-ends and small orifices

Figure 4.3 Traveling pulse waves and “tailored” valve reflections

Figure 4.4 Case studies: System scheme and initial steady state

xii
Figure 4.5 Development of steady-oscillatory flow in frictionless pipeline with fixed

valve opening at downstream, due to the upstream pressure oscillations

Figure 4.6 Responsive PID control valve vs. fixed-opening-valve in frictionless pipeline

(f = 0)

Figure 4.7 Responsive PID control valve vs. fixed-opening-valve in frictional pipeline (f

= 0.012)

Figure 4.8 Responsive PID control valve vs. fixed-opening-valve for static initial states

with frictional pipeline (f = 0.012)

Figure 4.9 System responses to forced pressure oscillation with various frequencies in a

frictionless pipeline (f = 0)

Figure 4.10 System responses to forced pressure oscillation with various frequencies in a

frictional pipeline (f = 0.012) (Fixed valve vs. responsive PID control valve)

Figure 4.11 Effect of controller parameters on system pressure control

Figure 5.1 Illustrative Hill Diagram of Francis turbine

Figure 5.2 Francis turbine characteristic curves in Suter parameters

Figure 5.3 Interpolation scheme of turbine characteristics

Figure 5.4 Illustrative characteristics of a Kaplan turbine

Figure 5.5 Illustrative characteristics of a Pelton turbine

Figure 5.6 Illustrative characteristics of a Turgo turbine with inclined nozzle

Figure 5.7 Illustrative characteristics of a pump turbine

Figure 5.8 Water-supply modes in conventional hydropower stations

xiii
Figure 6.1 Schematic of turbine boundary condition

Figure 6.2 Scheme of the turbine attached valves

Figure 6.3 Flow chart of turbine boundary condition programming

Figure 6.4 Comparisons with Wylie’s simulations

Figure 6.5 Comparison of Wylie’s turbine characteristics with other Francis turbine

characteristics

Figure 6.6 Schematic waterway of two parallel turbine units

Figure 6.7 Comparison of numerical modeling and field measurement

Figure 7.1 Energy recovery hydro turbine and system constituents

Figure 7.2 Turbine transient responses to 100% load rejection

Figure 7.3 Main pipeline pressure responses to 100% load rejection

Figure 7.4 Turbine transient responses to 50% load rejection

Figure 7.5 Main pipeline pressure responses to load rejection of Sloan Forebay turbine

unit

Figure 7.6 Sketch of South Valley Lateral System

Figure 7.7 Isolated turbine responses to 10 upstream hydrants flow

Figure 7.8 Online turbine responses to 10 upstream hydrants flow

Figure 7.9 Isolated turbine responses to single upstream hydrant flow

Figure 7.10 Online turbine responses to single upstream hydrant flow

Figure 7.11 Isolated turbine responses to 10 downstream hydrants flow

Figure 7.12 Online turbine responses to 10 downstream hydrants flow

xiv
Figure 7.13 Isolated turbine responses to single downstream hydrant flow

Figure 7.14 Online turbine responses to single downstream hydrant flow

Figure 8.1 Planned system layout with differential surge tank

Figure 8.2 Sketch of planned differential surge tank

Figure 8.3 Originally specified wicket-gate closure law

Figure 8.4 System responses during full load rejection with planned system layout and

design and original specified wicket-gate closure law (T1 = 6 s, T2 = 8 s)

Figure 8.5 Turbine speed rise during full load rejection with planned system layout and

design and wicket-gates refusal to close (runaway condition)

Figure 8.6 Envelope of max. and min. HGLs along main penstock during full load

rejection with 20 m diameter of surge tank and originally specified WG

closure law (T1 = 6 s, T2 = 8 s)

Figure 8.7 Sensitivity analysis of surge tank distance

Figure 8.8 System responses during full load rejection with a closer differential surge

tank and original specified wicket-gate closure law (T1 = 6 s, T2 = 8 s)

Figure 8.9 Turbine speed rise during full load rejection with a closer differential surge

tank and wicket-gate refusal to close (runaway condition)

Figure 8.10 Sensitivity analysis of wicket-gate closure time

Figure 8.11 System responses during full load rejection with planned differential surge

tank and extended wicket-gate closure law (T1 =11 s, T2 = 8 s)

xv
Figure 8.12 System responses during full load rejection with extended wicket-gate

closure law (T1 = 35 s, T2 = 15 s) and flywheel (adding inertia 28000 kg.m2)

for no surge tank scenario

Figure 8.13 Sensitivity of simple tank diameter D

Figure 8.14 Sensitivity of feeder and connector pipe diameter Dc

Figure 8.15 System layout with suggested underground air-cushioned surge tank

Figure 8.16 Sketch of simple cylinder air-cushioned surge tank

Figure 8.17 System responses during full load rejection with originally specified

wicket-gate closure (T1 = 6 s, T2 = 8 s) and simple cylinder air surge tank

(1000 m3 in volume)

Figure 8.18 Sensitivity analysis of the height of simple cylinder air surge tank

Figure 8.19 System responses to full load rejection with originally specified wicket-gate

closure (T1 = 6 s, T2 = 8 s) and a pipe-like air surge tank (500 m3 in volume)

Figure 8.20 System responses to full load rejection with extended wicket-gate closure (T1

= 18 s, T2 = 8 s) and pipe-like air surge tank (500 m3 in volume)

Figure 8.21 System responses to full load rejection with originally specified wicket-gate

closure (T1 = 6 s, T2 = 8 s) and a greater size of air surge chamber (840 m3 in

volume)

Figure 8.22 Sensitivity of pipe-air-chamber length to peak pressure during full load

rejection with originally specified wicket-gate closure (T1 = 6 s, T2 = 8 s) and a

pipe-like air surge tank (500 m3 volume)

xvi
Notation

a – wave speed

a0 – turbine wicket gate opening.

a
B – pipe constant in MOC equations, B =
gA

D – internal diameter of pipe cross section

D1 – diameter of turbine runner

d – diameter of Pressure Relief Valve and bypass line

E – valve energy dissipation coefficient, a valve size parameter determined by the energy

dissipation potential of the valve

EB – energy dissipation coefficient of turbine bypass valve

EC – energy dissipation coefficient of turbine control/shut-off valve

ES – energy dissipation coefficient of turbine surge valve

e – control error or deviation of the process variable u(t) from its set point u*;
u (t )
dimensionless error is defined as e = 1 − .
u*
FDR –Relative Friction Decay Rate

f – Darcy-Weisbach friction factor of a pipe

g – acceleration due to gravity

H or H(t) – instant pressure head difference across a hydraulic device

H0 – pressure head difference across a hydraulic device at initial steady state

H1 or H1(t) – instant pressure head at the inlet of a hydraulic device

xvii
H10 – initial pressure head at the inlet of a hydraulic device

H1*(t) – the dynamic set point of the pressure head at the inlet of a hydraulic device

H2 or H2(t) – instantaneous pressure head at the outlet of a hydraulic device

H20 – initial pressure head at the outlet of a hydraulic device

H2*(t) – the dynamic set point of the pressure head at the outlet of a hydraulic device

Have – average turbine head over years

HC – pressure head across the turbine control/shut-off valve

HR or HR(t) – instantaneous pressure head at reservoir

HR0 – initial pressure head at reservoir

HTail – water level in tailrace channel or downstream reservoir

Hmax – maximum turbine hydraulic head

Hmin – minimum turbine hydraulic head

HR – the design/rated turbine head

H (x, t) – piezometric head (HGL) at time t and position x.

H
h – relative/dimensionless head of a turbine, h =
HR

I – polar moment of inertia of rotating fluid and mechanical parts in the turbine-generator

unit, I =WRg2

KC – proportional gain of controller

L – pipe length

xviii
N1 – number of pipes at a junction, whose flow direction is assumed toward the node

N2 – number of pipes at a junction, whose flow direction is assumed away from the node

Ndy – number of wicket gate settings;

Nblade – number of runner blade angle settings.

n – rotational speed of turbine unit (rpm)

nE – synchronous speed of generator

ng – specific speed of a turbine or pump

ns – specific speed of a turbine

nR – rated speed or normal speed of turbine unit

nrun – runaway speed of turbine unit

nD1
n11 or n1' – unit speed of a turbine, n 1 =
'

P – rated energy produced from turbine or generator per second (kW, Hp)

PG – generator power output

PT – turbine power output

P
P11 or P1' – unit power of a turbine: P1 =
'

D H 3/ 2
2
1

Q or Q(t) – instantaneous flow in pipeline or passing through a hydraulic device

Q* – dynamic set point of the flow passing through a hydraulic device

Q0 – initial steady state flow in pipe system or passing through a hydraulic device

Q1 or Q1(t) – instantaneous flow at the inlet of a hydraulic device (also called external

flow)

xix
Q2 or Q2(t) – instantaneous flow at the outlet of a hydraulic device (also called external

flow)

QBV – discharge of turbine bypass valve

QCV – discharge of turbine control/shut-off valve

QSV – discharge of turbine surge valve

QR – design/rated turbine discharge

Q
Q11 or Q1′ – unit discharge of a turbine, Q1′ =
D12 H

Qext – external flow at a junction, governed by a hydraulic device or boundary condition

fΔx
R – pipe constant in MOC equations, R =
2gDA2

Rg – gyration radius of turbine-generator rotating parts

SET1 – high-pressure set point of PRV

SET2 – low-pressure set point of PRV

T – turbine shaft torque exerted by the water

T
T11 or T1' – unit turbine shaft torque, T1' =
D13 H

TR – rated turbine shaft torque exerted by the water

Tg – resistant shaft torque from the generator

TV1 – opening time period of PRV

TV2 – closure time period of PRV

xx
Td – derivative time constant of controller; dashpot time constant in turbine speed

governor

Tf – the shortest time period for closure of turbine wicket gate

Ti – integral time constant of controller

Tm – mechanical strating time of turbine-generator unit, Tm = I ωR2 TRωR = IωR TR

Tα – turbine governor’s promptitude time constant without temporary speed droop

Tα' – turbine governor’s promptitude time constant with temporary speed droop, that is,

Tα' = Tα + δ Td

t – elapse of time

u(t) – process variable to be controlled

u* – set point or desired value of process variable

V (x, t) – fluid velocity whose direction is defined as the same as the distance x

Q
v – relative/dimensionless flow of a turbine, v =
QR

W – weight of turbine-generator rotating parts and fluid contained


h
WH – Suter parameter of turbine head, WH ( x) = , in which the
α + v2
2

α
coordinate x = tan −1
v
β
WB – Suter parameter of turbine torque, WB ( x) = , in which the
α + v2
B

α
coordinate x = tan −1
v

x – distance along the centerline of a pipeline

y – dimensionless opening of turbine wicket-gates or a control valve

xxi
n
α – relative/dimensionless speed of a turbine, α =
nR

T
β – relative/dimensionless torque of a turbine, β =
TR

δ – turbine governor’s temporary speed droop

σ – turbine governor’s permanent speed droop

γ– the specific weight of water

τ – dimensionless valve opening

τ B – dimensionless opening of turbine bypass valve

τ C – dimensionless opening of turbine control/shut-off valve

τ S – dimensionless opening of turbine surge valve

η G – generator efficiency

ηT – turbine efficiency

η – efficiency of a turbine-generator unit.

ω – angular speed of a turbine, ω = 2πn 60

ω R – rated angular speed of a turbine, ω R = 2πn R 60

Δt – time step, or dimension of time in MOC grid

Δx – dimension of the pipe distance in MOC grid

ϕ – runner blade angle for Kaplan turbines

xxii
Chapter 1 Introduction

1.1 Active Hydraulic Devices and Transient Control

Unsteady or transient states are inevitable for any pipeline carrying water or any other

substance. Indeed, transients can be caused by routine switch on/off of a device, system

initialization (pipeline filling), accidents occurred during normal operation (e.g., power

failure, load rejection) or any other event that causes a flow change. When a pressured

pipe system experiences a transient event, both the physical components of the system

and the services provided by the system could be at risk due to the possible extreme high

pressure, unacceptable low or negative pressure, vibration of devices, or the overspeed of

any associated turbo machines. Therefore, the unsteady condition and transient flow

should be well managed and controlled to maintain the normal operation of the system.

Although transient flow is transmitted through pipelines, such flows are

almost universally initiated and modified by hydraulic devices in the system. In other

words, boundary conditions are usually the key to control system transients, and they

play crucial roles in determining both the nature and magnitude of system responses

during unsteady states. The term “Active Hydraulic Device” (AHD) is coined here to

describe a category of hydraulic component with the potential to create, manipulate and

control the system transients, typically by direct or active modification of their head-

discharge characteristic. Thus, as a class AHDs include all types of control valves,

1
hydraulic turbines, pumps and so on. Each device specifies a certain relationship between

pressure and flow, and both the flow and associated pressure are usually adjustable to

certain degree. Interestingly, a pump or turbine usually adjusts its working head and flow

through the operation of its control valve or wicket-gates, although the speed or load

variation of the hydraulic machine could also lead to certain changes of its flow-head

relationship. Thus, the control valves or wicket-gates are in this sense more active than

hydraulic machinery in terms of transient control. Moreover, even less active hydraulic

devices, such as reservoirs, surge tanks, and air chambers, are usually operated in

conjunction with any control valves used for waterhammer protection during the system

transients. Each hydraulic device has different level of complexity and activity that can

be conceptually visualized in Figure 1.1, although the placement of the device in the

diagram needs to be considered approximate only.

Hydro Turbines
Variable Speed Pumps
Activity Level

Fixed Speed Pumps


Modulating Valves
Air Chambers
Water Tanks
On-off Valves
Reservoirs

Modeling and Operational Complexity

Figure 1.1 Complexity of hydraulic devices

2
It is not surprising that most transient events in a pipe system are created by the

actions of these devices either on purpose or by accident, while different active devices

(or strategies) are implemented in different systems to control and limit system transients.

Specifically, a pump with Variable Frequency Drive (VFD), by dynamically altering

pump speed, allows not only a more controlled start/stop of the pump, but also a more

precise match of pressure and flow to system requirements. Similarly, a governed turbine,

by regulating the unit speed, is able to change the opening of wicket-gates in response to

the demand from the power grid system. The particular strategy of Misaligned Guide

Vanes (MGV), specially designed for a high-head reversible turbine, is a typical active

control strategy, which modifies the “S-shape” turbine character by allowing pre-opening

of some guide vanes before the movement of others and mediates the instability at zero-

load operating region and reduces the difficulty of synchronization (Liu, et al., 2007). In

addition, by automatically adjusting the valve opening, control valves are able to relieve

excessive transient pressures (e.g., through pressure relief valves and air valves), and or

to at least partly sustain the desired pressure or flow rate at some point of system

(modulating valves), or even to create “non-reflective” boundaries for

eliminating/limiting reflection and resonance of pressure waves in the pipe system.

Overall, no hydraulic device is a panacea for all pipeline surge and waterhammer

problems, yet a combination of different components often provides a cost effective and

holistic solution to unsteady flow challenges. The resulting control system must be

examined thoroughly and comprehensively utilizing appropriate computer software for

transient simulation.

3
1.2 Thesis Objectives

This thesis aims to develop new capacity of transient analysis modeling and control for

several AHDs and their associated protection strategies. A generalized mathematical

model is first adapted for the hydraulic devices with PID (proportional, integral and

derivative) type of controllers since automatic control is inherently relevant to the AHDs;

the governing equations in the general model of AHDs are quite readily created for an

active control valve. In order to demonstrate the capability of an automatic control valve

for manipulating the wave reflection and resonance in a pipeline system, an attempt is

undertaken to modify the model of a PID control valve by combing a remote sensor with

the controller to create a “non-reflective” boundary condition in the pipe system.

Meanwhile, based on the general mathematical model of AHDs, with the knowledge of

hydro systems and characteristics of turbine-generators, a numerical model of governed

turbine unit is established with the intent of analyzing and controlling the system

transient performance for realistic hydro projects. More broadly, this thesis endeavors to

place the issue of transient analysis of AHDs within the larger context of comprehensive

evaluation methodology of system hydraulic transient performance and protection.

1.3 Thesis Overview and Layout

The progress of this thesis was guided by the creation of conference and journal articles

and project reports; some of the core material has been disseminated in different venues.

This dissertation is organized around these papers and reports, but effort is made to link

4
them logically using the “AHD” concept as their essential core, though readers might not

find the traditional “linear” narrative throughout the thesis.

Automatic control valves are the first type of active hydraulic devices studied

in this research. Such a valve could be either a cause of transient flow or a measure used

to protect system from detrimental transients; it is often a key element in a water system

to control the flow rate, reverse flow, pipe pressure or water level in a tank. Due to the

challenges inherent in the design of a Pressure Relief Valve (PRV), the general principles

of their use and selection are studied along with the sensitivities of a system’s response to

the PRV parameters (Chapter 2). Chapter 2 elucidates what to pay attention when

designing or applying an AHD (i.e., PRV) to protect a system from excessive

waterhammer pressures.

Chapter 3 introduces the general numerical modeling challenges associated

with AHDs. In this chapter, the extended Method of Characteristic (MOC) for transient

analysis and theory of automatic control and controllers are reviewed and a generalized

mathematical model of AHDs is described. This chapter also briefly reviews the relevant

literature, providing a foundation for the overall thesis research.

Chapter 4 is one of the core contributions of this research; it creates a new

strategy of transient pressure control using an AHD (i.e., the PID-controlled valve).

Based on the numerical model of conventional PID control valves, a new application of

PID control valve is envisioned that combines a remote sensor at the upstream of a

pipeline to create a non- or semi- reflective boundary at the downstream end. Case study

shows that, at such a boundary, the valve is able to adjust its opening automatically in

5
response to the pressure changes remotely sensed to eliminate or mediate the reflection

and resonance of pressure waves within a pipeline system.

The second type of active hydraulic device studied in this thesis is the

governed hydro-turbine, the most complicated hydraulic component in terms of transient

analysis and waterhammer control. The motivation and interest in this hydro system

research arises partly from the author’s previous background, as well as renewed interest

worldwide in hydro-related businesses (especially small hydro) motivated by growing

environmental concerns and improved hydro technologies. Meanwhile, waterhammer

has been a serious issue in many hydro systems, and numerous accidents or troublesome

behavior is associated with the operation of turbine units (Pejovic, et al., 2007). Although

there are a number of numerical models developed for transient analysis of hydro systems,

few of the published ones are particularly designed for the applications in water networks.

Chapter 5 reviews the hydro turbine and its control system, including the

treatment of turbine model characteristics. Then, Chapter 6 develops a comprehensive

numerical model for turbine installations in either conventional hydro pipeline systems or

complex municipal water networks. Following that, as case studies of transient

performance evaluation and system transient control, two realistic hydro projects are

presented, including an energy recovery hydro-turbine project in urban water supply

system (Chapter 7) and a conventional hydroelectric power project (Chapter 8).

Chapter 9 summarizes the thesis research and addresses the perspective that

system performance and transient protections should be evaluated at different technical

levels and in a more broad way.

6
1.4 Publications Related to Thesis Research

The following list comprises the journal articles and conference papers that have either

been published, or are under work for publication that are related to the thesis. As

aforementioned, the core contributions of this thesis have been disseminated in

publication format.

1. Qinfen Zhang, Bryan W. Karney and David L. McPherson. (2008) “Pressure Relief

Valves Selection and Transient Pressure Control”. August 2008, J. of AWWA. Vol.

100, No. 8, p62-69 (closely linked to Chapter 2).

2. Bryan Karney, Qinfen Zhang, Feng Wang and Stanislav Pejovic (2008). “Non-

reflective Boundary Design through Remote Sensing and PID Control Valve”.

Published and orally presented at Pressure Surges 10th International Conference,

BHR Group, Edinburgh, Scotland, May14-16, 2008 (loosely linked to Chapter 3

and 4).

3. Katherine Qinfen Zhang, Bryan Karney (2008). “Energy Recovery Hydro Turbines in

water Supply system”. Published and orally presented at HydroVision 2008,

Sacramento, California USA, July 14-18, 08 (somewhat linked to Chapter 7).

4. Stanislav Pejovic, Bryan Karney, Qinfen Zhang. (2004) “Water Column Separation in

Long Tailrace Tunnel”, Proceedings of HYDROTURBO 2004 international

conference, Brno, Czech Republic, Oct. 2004 (loosely linked to Section 6.4.2,

Chapter 6).

7
5. Zhang Qinfen, Karney Bryan. (2003) “Pipe System with Micro-Turbines:

Waterhammer Considerations”, Proceedings of FEDSM’03 4th ASME-JSME Joint

Fluids Engineering Conference, Hawaii, USA, July 2003 (loosely linked to Chapter

5, 6 and 7).

Chapters 2 and 4 are based on the published journal paper entitled “Pressure

Relief Valves Selection and Transient Pressure Control” (No.1 in the Publication List)

and conference paper “Non-reflective Boundary Design through Remote Sensing and

PID Control Valve” (No. 2 in the Publication List). The materials of other published

papers are referenced in the thesis. For these thesis-related papers, I wrote the papers and

did the majority of research and analysis behind them. My Ph.D thesis supervisor, Prof.

Bryan Karney, was naturally a significant coauthor for the publications, providing ideas

and insights, and proofreading/editing the manuscripts. Other coauthors on more limited

aspects include David L. McPherson, Stanislav Pejovic and others, they all performed the

customary duties of coauthor by way of editorial suggestions and appropriate caveats. I

have received their written permission and endorsement to include the material derived

from, or based on, our joint publications.

8
Chapter 2 Pressure Relief Valve Selection and

Transient Pressure Control

Valves are the first type of active hydraulic devices studied in this thesis, specifically

automatic control valves which are treated in depth. While by no means simple, such

valves serve as an excellent starting point for any foray into active hydraulic devices

because they are considerably less complex than a turbine control system or variable

speed pump. The operation of automatic control valves could be either a cause of

transient flow or, conversely, a measure to protect a system from the detrimental effects

of such unsteady flow; valves are key elements in water system for regulating flow rate,

preventing reverse flows, limiting pressures, or moderating tank water levels. Due to the

challenges inherent in the design of a Pressure relief valve (PRV), perhaps better referred

to as a Surge Relief Valve (SRV), the general principles of PRV/SRV use and selection

are studied in this chapter along with an exploration of the sensitivity to system response.

This chapter highlights the appropriate design and application of an AHD (i.e., SRV) to

protect a system from excessive waterhammer pressures.

This work is now published in the Journal of the American Water Works

Association (August, 2008) and was co-authored with Bryan Karney and David L.

McPherson. As the first author, I wrote the full early draft of the paper, and did all the

numerical runs and analysis. The idea of this work originated with Prof. Karney and

various editorial suggestions and practical caveats were provided by the paper’s

coauthors.

9
Abstract: A pressure relief valve (PRV) is installed for surge protection and pressure

relief in a pipeline system. The valve, which is normally closed, is designed to open

rapidly once its pressure setting is met or exceeded, thus permitting fluid discharge to

relieve pressure. A PRV’s effectiveness depends on system properties, the surge

characteristics, and the way in which the valve’s attributes and settings are configured.

This chapter illustrates the design challenges pertaining to PRVs and shows that an

appropriately designed PRV can protect some systems from extreme pressures but that

inappropriate use can actually worsen a system’s transient response. The general

principles of PRV use and selection are presented along with a sensitivity analysis of

PRV parameters. Although this understanding is essential to effective system design, the

final selection of a PRV is made through numerical simulation that evaluation of PRV

viability and cost-effectiveness in specific systems.

Key Words: Pressure relief valve (PRV), Parameters selection, Numerical simulation,

Transient performance, Sensitivity analysis, Transient pressure control.

2.1 Introduction

Transient events, in particularly rapid changes of flow rate, can cause serious problems in

pipeline systems and networks. High waterhammer pressure can permanently deform or

rupture a pipeline and its components; low pressures can collapse a pipe, inducing leaks,

causing service disruption, and contaminating water in the pipeline.

Numerous transient control strategies have been developed, including changes

within the pipe system (pipe diameter, thickness, alignment, profile or other hydraulic

10
components), wave speed reduction, optimized operation, and installation of dedicated

devices such as automatic control valves, surge tanks, and air chambers (e.g., Wylie et al.,

1993; Karney and Simpson, 2007). Automatic control valves, including pressure relief

valves (PRVs), flow- or pressure-regulating valves, air valves and check valves, are

common and often cost-effective. Depending on the type, a valve is used to control

transient conditions either by reducing the rate of flow velocity change or by discharging

or admitting air into the pipeline. When triggered by a pressure that is beyond its preset

limit, a PRV opens to allow flow. The resultant outflow causes a pressure drop and

presumably reduces the maximum pressure. Conversely, inflow through the valve

compensates for reduced flow and can limit low pressures and/or avert cavitation. A

PRV must have a low physical inertia so that it can respond rapidly to the detected

pressure fluctuations and open before the set-point is greatly exceeded (Chaudhry, 1979).

As portrayed in the case studies presented in this chapter, delays in valve opening can

compromise system transient protection.

2.2 Types of PRVs and Their Applications

Depending on the application and industry, various types of PRVs are used. A safety

valve (also referred to as an overpressure pop-off valve), usually applied in steam or gas

conveyance systems, is a spring- or weight-loaded valve that opens once the pressure in

the pipeline exceeds its set point and closes immediately when the pressure drops below

such threshold. Thus, a safety valve is either fully open or fully closed. If not activated

too frequently, a rupture disk can be used as an alternative to a safety valve. This type of

valve is not a “true” valve but rather an opening in the pipe. The opening is covered by a

11
diaphragm that ruptures and relieves pressure when the pressure set point is exceeded.

One disadvantage of a rupture disk is that it continues to discharge until it is replaced.

In liquid applications, a PRV is usually mounted on the discharge side of a pump,

hydro turbine, or main cutoff valve. It is typically a pilot-controlled throttling valve,

opened or closed either hydraulically or by a servomotor, with opening and closing rates

that can be individually set. It is distinct from a pressure-regulating valve, although both

have pilot systems. In fact, a regulating or modulating valve typically uses a feedback

device (usually, a PID type controller) to accurately sustain a pressure set point in

response to the sensed pressure or pressure difference. By contrast, a PRV is usually

triggered by an event (On/Off controller) and, once triggered, follows a predefined

opening and closing motion. More precisely, a PRV opens to release water when the

pipeline pressure at the valve inlet exceeds a high set point (referred to as SET1) and the

normal discharge of the valve is to a zone of lower pressure or to an open discharge area

such as a pump station wet well or adjacent storm sewer, pond or stream. In addition, a

PRV can open to admit water through a bypass line when the valve’s downstream

pressure in the main pipeline falls below a low-pressure set point (referred to as SET2).

If the pressure of the water source linked with the PRV is higher than that at the pipeline

downstream of the control valve, the valve opens to supply fluid, thus compensating for

the reduced flow and limiting the magnitude of the downsurge and the subsequent

upsurge reflected from the downstream pipeline. A PRV set in this mode is commonly

referred to as a pump- or valve-station bypass assembly.

PRVs have been used successfully in a wide range of hydraulic conditions and

operating scenarios to reduce the adverse transient conditions within a pipeline system.

12
However, they should not be used without proper assessment of their ability to

adequately protect the system. Because a PRV is a reactionary device, many hydraulic

systems will not benefit from a PRV. For example, if the PRV is used to relieve high

pressure, the pressure relief will initially be local to the PRV itself, because its operation

obviously depends on a local discharge of fluid. Protection of the pipeline as a whole will

depend on a variety of factors that involve the complex interplay between the strength

and source of the original surge condition, the way in which the valve action modifies the

wave propagation, and the way in which these conditions interact with the system’s

strength. In some cases, a PRV may only provide quick local protection, making other

mitigation strategies more cost effective. One particular misapplication is to deploy a

PRV in the bypass mode to protect a pumped system with a rising main from an

uncontrolled shutdown (power failure). Under these conditions, the available pressure on

the suction side of the pump station may not allow sufficient flow to effectively limit the

resultant downsurge. The first step to develop a positive transient control system using a

PRV is to understand the valve’s operation and limitations. A preliminary numerical

evaluation provides the design engineer with insights into the suitability of a PRV for

controlling surge pressures.

Once a decision has been made to use a PRV, it must be carefully and

appropriately designed. There are four design considerations for a PRV. The first defines

the valve’s location. In a high-pressure relief mode of operation, the PRV should be

positioned so that the high pressure and flow can be diverted around the pressure-

sensitive areas and excess flow can be discharged to an appropriate place. The remaining

three design considerations relate to the PRV characteristics and valve parameters,

13
including the valve size (d), the high- or low-pressure set point (SET1 or SET2), and the

opening and closure time periods (TV1/TV2). Each valve parameter profoundly

influences system performance. For instance, an oversized PRV will not be cost-

effective (the system safety enjoys little improvement), while an undersized valve cannot

effectively alleviate excessive pressure surges. Using a fully dynamic transient model,

TransAM, a simplified pipeline system is analyzed in order to illustrate both the general

principals of PRV design and the sensitivity of the response to control valve parameters.

2.3 PRV Operation in Pump and Turbine Systems

To protect the system from unacceptable low pressure or excessive high pressure during

the sudden pump shutdown, a PRV can sometimes be installed on a bypass line around

the pump station. When a pump’s power supply is interrupted, the reduced flow at the

pump causes an imbalance in flow and a rapid reduction in pressure. The net effect is that

a rarefaction wave propagates into the discharge line. When the low pressure wave

reflects off a downstream boundary (e.g., water tank), the pressure will be normalized to

the free water surface level in the tank and thus reflected back into the system,

establishing pressures that are sometimes higher than those originally experienced before

the flow was reduced. After this reflected positive wave has reached the pump, and the

pump check valve has closed, the check valve effectively becomes a “dead-end”, creating

another reflection and magnification. In theory, the reflected wave doubles at a closed

valve, producing a higher transient pressure. In a pumped system, a PRV will open at the

initial low-pressure set point once power to the pump motor is cut off. The PRV, now

open in anticipation of the returning positive wave, will offer a means for releasing water

14
from the system and prevent the onset of a high transient pressure condition. If the

operating suction pressure is not positive in a pumped system, a partial or full vacuum

pressure will develop at the PRV once the valve is opened. Should a full vacuum develop,

the severity of the transient pressure may be exacerbated because of cavity collapse. In

the absence of positive pressure, particular measures (such as an air valve) might be

needed when contemplating whether a PRV is an appropriate control device.

Main Valve Movement


Relative Valve Opening

PRV Movement

TV1 TV2
Time (s)

Figure 2.1 Operation of main valve and PRV

Unlike with regular startup and shutdown, power failure is always sudden and

unpredictable. In Figure 2.1, the main valve closure is equivalent to the reduction of

pump flow due to the power failure. This kind of system (e.g., typical of certain types of

booster pumping stations) assumes a positive suction operating head. As a result, the

PRV in the bypass line opens rapidly to admit water into, or release water from, the

pipeline back to the suction side of the pump. The PRV is then gradually closed to

15
eventually shut down the system and bring it to rest; a slowly closing valve has little

effect on the transient condition. The check valve, which is typically installed at each

pump, will close upon flow reversal. However, when the check valve is closed, the

rotational moment of inertia that controls the rate of pump run-down during a power

failure will be effectively eliminated (Ruus and Karney, 1997; Chaudhry, 1979). As a

result, the energy dissipation required to slowly bring the system to rest and dampen the

adverse transient condition is left to the PRV bypass assembly.

The application of a PRV in a hydropower plant depends on many factors. It is

more commonly used in small hydro turbine installations, where large surge protection

measures such as a surge tank and air chamber are not essential. In large hydropower

plants, a PRV usually functions as a secondary surge protection measure and is not

typically a substitute for a surge tank. For a power plant isolated from the electric grid,

the most common configuration is to have the PRV in the bypass line kept partly open to

provide for the maximum anticipated rapid load changes. As shown in Figure 2.2(a),

when the turbine unit rapidly accepts a load, the PRV begins to close almost

simultaneously with the opening of the wicket gates in order to compensate the turbine

flow from the PRV bypass line. Alternatively, following a load reduction and as the

wicket gates are closing, the PRV opens to divert the turbine flow as illustrated in Figure

2.2(b); subsequently, the PRV closes slowly. Yet, when a full load rejection occurs, either

in the synchronous or isolated operation of the turbine unit, all the turbine flow is

switched to the PRV, as shown in Figure 2.2(c). In this way, a much more constant flow

velocity in the penstock is maintained during either load acceptance or load rejection.

16
Some water is wasted through the PRV operation; however, the amount of discharged

water is insignificant compared with the total cost of transient control (Chaudhry, 1979)

Figure 2.2 Synchronous operation of turbine and PRV

17
2.4 Case Studies

The design of a PRV to control adverse transient pressure is system specific and depends

on the physical and hydraulic conditions in the system and the nature of its response to

operation of the PRV during the transient event. Fully dynamic hydraulic transient

modeling can be used to illustrate the system’s reactions and estimate its transient

performance associated with the selection of various PRV parameters. To illustrate, a

simple system with a PRV installation at both upstream and downstream locations, is

discussed here.

2.4.1 Brief description of the system

As shown in Figure 2.3, two reservoirs are linked by a uniform pipeline with a length L =

500 m, diameter D = 1.0 m, friction factor f = 0.012, and wave speed a = 1200 m/s. The

water level is 15 m at upstream reservoir and 12 m at downstream reservoir. At each end

of the pipeline, there is a primary valve and a PRV; the PRV is situated on a short bypass

line that connects either end of pipeline to its associated reservoir. At initial steady state,

both main valves are fully opened, both PRVs are fully closed, and flow in the pipeline is

Q0 = 1.73 m3/s.

2.4.2 Case study 1: upstream control

Main-valve-1 at the upstream end is fully closed within 5 s, causing an incident

rarefaction pressure wave in the pipeline. When the pressure at the outlet of PRV-1 (i.e.,

the pressure at the outlet of main-valve-1) falls below its set point (SET2), the PRV opens

18
to admit water and avoid the downsurge at the outlet of main-valve-1, thus reducing the

subsequent upsurge reflected from the downstream reservoir.

15 m

12 m

Main-Valve-1 Q
Datum

PRV-1 PRV-2

Figure 2.3 PRV installations in a simple pipeline system

Comparison of different upstream PRV parameters. The transient pressures in

the pipeline are compared in Figure 2.4 for the following systems: one without PRV

protection, one with an appropriately designed PRV, and four with poorly designed PRVs.

Without PRV protection (Figure 2.4(a)), rapid closure of main valve would result in

approximately 260 m of maximum pressure head and -240 m of minimum pressure head

in the system. The maximum head 260 m is extremely high, and this may cause pipe

rupture or impose extra cost for a pipe of greater strength. The negative pressure head (-

240 m) is a numerical value from the simulation model and does not factor in

vaporization and cavitation. However, the modeled -240 m rarefaction wave is

completely unacceptable because of the occurrence of cavitation and subsequent cavity

collapse. Cavitation will lead to vapor cavity formation and subsequent cavity collapse

(the so-called column rejoinder effect), which can cause extremely high pressure. The

19
large negative pressures indicate that the response is unacceptable when the PRV is not

operated, and thus protection from transient pressure is necessary for this system. As

shown in Figure 2.4(b), operation of an appropriately designed PRV will limit maximum

and minimum transient pressure heads along the pipeline to within the range of 3.6 m and

22 m. However, inappropriate selection of PRV parameters, such as an undersized PRV,

a pressure set point for SET2 that is too law, a PRV opening that is too slow, and closure

that is too fast, would result in poor performance and unacceptable transient pressures, as

evident in Figure 2.4 (c), (d) (e) and (f), respectively. However, Figure 2.4 also shows

that even a poorly designed PRV provides some protection and is certainly preferable to

completely ignoring PRV deployment.

Sensitivity analysis of upstream PRV parameters. Proper selection of a PRV’s

control parameters is important, particularly because these parameters are inevitably

uncertain. For example, activation of a PRV depends on exactly how and where the

pressure is sensed and how this information is transformed by both the ongoing transient

and control valve action. Sensitivity analyses would reveal how the system’s transient

performance changes with variation of each PRV parameter, and could aid in identifying

rules for PRV design. Therefore, the sensitivities of each parameter of PRV-1 (d, SET2,

TV1, TV2) were analyzed and results are summarized in Figure 2.5.

Figure 2.5(a) indicates that an increase in valve size would improve the transient

performance, but only up to a certain limit; an increase in valve size beyond 0.5 m would

not effectively improve transient performance in this case study. Design and field

experience indicate that PRV diameters usually range from one twentieth to one third of

the main pipe diameter, with most values falling near the middle of this range. The logic

20
for achieving this type of range is evident in Figure 2.5(a), although not exactly followed.

The slight discrepancy that appears can be explained as follows: in this case study, only 3

m of pressure head existed along a 500-m length of pipe; thus the driving heads in this

system is very low. This implies that in order to discharge a certain amount of water, a

larger PRV is needed. The ultimate choice of PRV size is usually a compromise between

cost, which increases rapidly with diameter, and hydraulic control and protection, which

usually provide diminishing returns with PRV diameter.

Figure 2.5(b) illustrates that the variation in transient pressures diminishes with an

increase in the low-pressure set point (SET2). An issue is how quickly the bypass or

PRV will open immediately after a transient down-pressure wave is detected in

association with the motion of the primary valve. Because of the system configuration

presented in case study, there is no improvement in response when SET2 is beyond 15 m,

because the valve stays open once the set point is higher than this upstream reservoir

pressure. In general, the lower or more sensitive the threshold pressure for PRV

activation, the better the transient protection, because the valve will begin its response

earlier. However, there is an obvious trade-off, because a valve that activates quickly is

more prone to activate not only for important design events but also for normal or routine

transients that pose no threat to the system. A valve that opens prematurely not only

requires more frequent routine maintenance but, depending on system configuration, can

waste both energy and fluid from the system.

Figure 2.5(c) shows that the variation of opening time period (TV1) does not bear

on the maximum transient pressure in this case study; in general, the shorter the opening

time (TV1), the less severe the resulting negative pressure. When TV1 is less than 5 s, a

21
reduction in TV1 no longer changes the minimum transient pressure. When TV1 is longer

than 10 s, the negative pressure would occur in the pipeline and rapidly becomes more

severe with increasing TV1. In this case study, the appropriate opening time for the PRV

should be within 5-10 s when both ease of PRV operation and the prevention of negative

pressures in the pipeline are considered. As mentioned here, and discussed by Chaudhry

(1987), low valve inertia or highly responsive control system is required for a PRV to

achieve a response and activation. These components can increase either, or both, valve

cost and maintenance requirements. However, the timing of PRV operation and the

transient response are unique to each hydraulic system and essentially govern the needs

and costs of the control valve, the configuration of the system, and the nature of the

disturbance that actually generates the transient response.

As shown in Figure 2.5(d), slowing PRV closure (i.e., increasing TV2) results in

smaller variations in transient pressures, which benefits hydraulic control but diverts

more water during PRV operation. In addition, when TV2 is shorter than 60 s, the range

of maximum and minimum pressures becomes more sensitive to the variation of TV2. In

this case study, 60- 80 s is suggested for the PRV closure time period.

2.4.3 Case study 2: downstream control

In this case study, main-valve-2 at the downstream end is fully closed in 5 s, causing an

incident upsurge in the pipeline. When the pressure at the inlet of PRV-2 (i.e., the

pressure at the inlet of main-valve-2) exceeds its set point (SET1), the PRV opens to

discharge water into the downstream reservoir to mediate the pressure rise in the

upstream pipeline.

22
Comparison for different downstream PRV parameters. As shown in Figure

2.6, if there is no PRV protection, the rapid closure of main-valve-2 results in 232 m of

the maximum pressure and -200 m of the minimum pressure (see previous text regarding

pressure in excess of full vacuum condition) in the system. Operation of an appropriately

designed PRV will reduce the envelope of extreme pressures to between 4.2 and 25.8 m.

An undersized PRV, or the inappropriate setting of the PRV set point or valve timing,

would result in unacceptable transient conditions, shown in Figure 2.6 (c-f).

Sensitivity analysis of downstream PRV parameters. The sensitivities of each

PRV-2 parameter of (d, SET1, TV1, TV2) were analyzed and are shown in Figure 2.7;

the results are similar to those from the previous case study. Figure 2.7(a) illustrates that

the most economically efficient and hydraulically effective diameter of the PRV is

around 0.2-0.3 m, which is well within the range from one fifth to one third of the main

pipe diameter.

Figure 2.7(b) shows that the range of transient pressures becomes narrower as

SET1 is reduced. However, there is little improvement in transient control when SET1 is

set lower than 15 m, which is the water level in the upstream reservoir. In addition, a

negative or partial vacuum pressure resulting from the boundary reflection will become

more severe if SET1 is set higher than 32 m. In this case study, a proper SET1 value of

15-32 m should be selected.

Figure 2.7(c) demonstrates how a reduction in PRV opening time (TV1) will

reduce the maximum transient pressure in the system, though only slightly affecting the

minimum transient pressure in the pipeline. Curtailing TV1 to less than 20 s does not

23
change the transient response any further. In this case study, results indicate that the

system is rather insensitive to the selection of TV1 as long as TV1 is not extremely long

(e.g., longer than 60 s).

Figure 2.7(d) shows that a slow closure of PRV-2 would improve transient

pressure performance. In this case study, 60 s or longer is required for TV2 to prevent the

negative pressure from occurring in the pipeline. The trade-off is that a slower closing

time will result in more water being diverted out of the system during the PRV operation.

2.5 Summary and Conclusions

The objective of PRV design is to transform an existing hydraulic system to a new one

capable of a more acceptable transient response. In evaluating a PRV, the first step is to

qualitatively assess the PRV’s viability as a protection device. Doing this requires a

comprehensive understanding of how a PRV contributes to transient mitigation,

especially appreciation that a PRV is a reactionary device that establishes a pathway to

either introduce or relieve energy from the hydraulic system. Many designs have failed

because a PRV was not designed or deployed correctly. However, if such a device is

found to be a viable surge-control strategy, a series of interdependent design parameters

must be considered in its design. The basic principles of PRV designing and its

associated parameter selection have been described in this chapter and are summarized as

follows:

It is obvious that an undersized PRV would be insufficient to protect a system

from extreme transient pressures. There is a misunderstanding in practice,

24
that is, the bigger the valve diameter is the safer the system will be. However,

this intuition is not always valid, and there is no value in over-sizing a PRV

because, beyond a certain point, there is little further improvement in its

waterhammer performance but potential under pressure problems due to

larger flow discharge, even though its cost continues to increase.

If the pressure set point is exaggerated (i.e., too low for the low-pressure set

point or too high for the high-pressure set point), the PRV will not adequately

provide transient pressure control.

If the PRV opens too slowly or closes too quickly, it could cause dangerous

waterhammer incidents, although in general a PRV helps to alleviate extreme

waterhammer pressures.

The case studies presented in this chapter suggested common principles to follow

when selecting a PRV at preliminary design stage, but the detailed values of each

parameter will be different from case to case. For instance, the valve opening time period

TV1 is required for 5-10 s in the upstream PRV, while it could be as long as 20 s in the

downstream PRV. Therefore, to make wise and informed choices for these parameters at

final design stage, a fully dynamic hydraulic simulation is required to verify the

performance of each pipeline system.

25
Max. H

Steady State H

Pipeline

Min. H

(a) Simulation without PRV protection

Max. H
Min. H

(b) Simulation with an appropriately designed PRV


(d = 0.9 m, TV1 = 5 s, TV2 = 80 s)

Figure 2.4 System transient performance varying with upstream PRV design (Con’d)

26
Max. H
Steady State H
Pipeline

Min. H

(c) Simulation with an under-sized PRV


(d = 0.2 m, TV1 = 5 s, TV2 = 80 s)

Max. H

Min. H

(d) Simulation with a too-low-pressure set point PRV


(d = 0.9 m, SET2 = 2 m, TV1 = 5 s, TV2 = 80 s)

Figure 2.4 System transient performance varying with upstream PRV design (Con’d)

27
Max. H

Min. H

(e) Simulation with a too-slow-opening PRV


(d = 0.9 m, TV1 = 20 s, TV2 = 80 s)

Max. H

Min. H

(f) Simulation with a too-fast-closure PRV


(d = 0.9 m, TV1 = 5 s, TV2 = 40 s)

Figure 2.4 System transient performance varying with upstream PRV design
(Note: Default SET2 is equivalent to SET2>=15m, i.e., the PRV is activated instantly when Main-Valve-1 closes)

28
Max. / Min. Waterhammer Pressure
vs. PRV Diameter

Max. and Min. Pressure (m) 80


60
40
Max.
20
0
-20 Min.
-40
-60
-80
0 0.2 0.4 0.6 0.8
PRV Diameter (m)

(a) PRV size (d) (TV1= 5s, TV2=80s, SET2=0)

Max./ Min. Waterhammer Pressure


vs . PRV Low Pressure Set Point

40
Max. and Min. Pressure

30
Max.
20
10
(m)

0
Min.
-10
-20
-30
-40
0 5 10 15 20 25 30

PRV Low Pressure Set Point (m)

(b) PRV pressure set point (SET2) (TV1=5s, TV2=80s, d=0.9m)

Figure 2.5 Sensitivity analyses of upstream PRV parameters (Con’d)

29
Max. /Min. Waterhammer Pressure
vs. PRV Opening Time Period

30

Max./ Min. Pressure (m)


Max.
20

10

-10
Min.
-20

-30
0 5 10 15 20 25

PRV Opening Time Period (s)

(c) PRV opening time period (TV1) (TV2=80s, d=0.9m, SET2=0)

Max./ Min Waterhammer Pressure


vs. PRV Closure Time Period

50
Max. and Min. Pressure (m)

40

30
Max.
20

10

0
Min.
-10

-20
0 20 40 60 80 100 120 140

PRV Closure Time Period (s)

(d) PRV closure time period (TV2) (TV1=5s, d=0.9m. SET2=0)

Figure 2.5 Sensitivity analyses of upstream PRV parameters

30
Max. H

Steady State H
Pipeline

Min. H

(a) Simulation without PRV protection

Max. H

Min. H

(b) Simulation with an appropriately designed PRV


(d = 0.5 m, TV1 = 10 s, TV2 = 70 s)

Figure 2.6 System transient performance varying with downstream PRV design
(Con’d)

31
Max. H

Steady State H
Pipeline

Min. H

(c) Simulation with an under-sized PRV


(d = 0.1 m, TV1 = 10 s, TV2 = 70 s)

Max. H

Steady State H

Min. H

(d) Simulation with a too-high-pressure set point PRV


(d = 0.5 m, SET1 = 50 m, TV1 = 10 s, TV2 = 70 s)

Figure 2.6 System transient performance varying with downstream PRV design
(Con’d)

32
Max. H

Min. H

(e) Simulation with a too-slow-opening PRV


(d = 0.5 m, TV1 = 60 s, TV2 = 70 s)

Max. H

Steady State H
Pipeline

Min. H

(f) Simulation with a too-fast-closure PRV


(d = 0.5 m, TV1 = 10 s, TV2 = 40 s)

Figure 2.6 System transient performance varying with downstream PRV design
(Note: Default SET1 is equivalent to SET1<=15m, i.e., the PRV is activated instantly when Main-Valve-2 closes)

33
Max. / Min. Waterhammer Pressure
vs. PRV Diameter

180

Max. and Min. Pressure 140


100
60 Max.
(m)

20
-20 Min.
-60
-100
0 0.2 0.4 0.6 0.8
PRV Diameter (m)

(a) PRV size (d) (TV1=10s, TV2=70s. SET1=0)

Max./ Min. Waterhammer Pressure


vs . PRV High Pressure Set Point

100
Max. and Min. Pressure

80
Max.
60
(m)

40
20
0
Min.

-20
0 10 20 30 40 50 60
PRV High Pressure Set Point (m)

(b) PRV pressure-set point (SET1) (TV1=10s, TV2=70s, d=0.5m)

Figure 2.7 Sensitivity analysis of downstream PRV parameters (Con’d)

34
Max. /Min. Waterhammer Pressure
vs . PRV Opening Time Period

80

70
Max./ Min. Pressure (m)

60

50
Max.
40

30

20
Min.
10

0
0 10 20 30 40 50 60
PRV Opening Time Period (s)

(c) PRV opening time period (TV1) (TV2=70s. d=0.5m, SET1=0)

Max./ Min Waterhammer Pressure


vs. PRV Closure Time Period

120

100

80
Max. and Min. Pressure (m)

60

40 Max.
20

-20
Min.
-40

-60

-80
0 20 40 60 80 100
PRV Closure Time Period (s)

(d) PRV closure time period (TV2) (TV1=10s. d=0.5m, SET1=0)

Figure 2.7 Sensitivity analysis of downstream PRV parameters

35
Chapter 3 Transient Simulation of Active Hydraulic

Devices with an Introduction to Automatic Control

This chapter provides an introduction to the numerical modeling of AHDs. Since

automatic control is a defining element of AHDs, the key feature is a general

mathematical model for the transient analysis of AHDs with a PID (proportional, integral

and derivative) type controller. Familiarity with the material contained in this chapter is

important for understanding the overall thesis and is based on a literature review of

automatic control and hydraulic devices. Specifically, the Method of Characteristics

(MOC) and its implementation to pipe networks is first reviewed; this is followed by a

brief introduction of control theory as well as the evolution and variety of controllers.

Finally, a set of basic governing equations is presented for the transient simulation of a

generalized AHD in a pipe network.

3.1 Numerical Method of Pipe Transient Flow

The momentum and continuity equations governing transient flow in closed conduits are

classified as quasi-linear hyperbolic partial differential equations for which no analytical

solutions are available for a general pipe system. With respect to pipe systems and the

resolution of the governing equations, the superiority of the MOC over finite element or

finite difference methods in terms of both simulation accuracy and computational effort

has been demonstrated (e.g., Karney & McInnis, 1992).

36
3.1.1 Governing Equations and Method of Characteristics (MOC)

The momentum equation (3-1) and continuity equation (3-2) can be derived for unsteady

closed conduit flow based on certain reasonable assumptions and approximations (Wylie,

etc., 1993 and Chaudhry, 1987).

j+1
△t
j P
C+ C-
j-1
x
Δx Δx
A C B
i-1 i i+1
Figure 3.1 MOC grid for single pipe

f
L1 = gH x + Vt + V | V |= 0 (3 - 1)
2D
a2
L2 = H t + Vx = 0 (3 - 2)
g

As shown in Figure 3.1, x is the distance along the centerline of the conduit,

and t is the time. In the equations, H (x, t) is the piezometric head (HGL), V (x, t) is the

fluid velocity whose positive direction is defined as the same as the distance x. Both H

and V are dependent on x and t, and the subscripts represent partial derivatives. The other

constants include f = Darcy-Weisbach friction factor, a = wave speed and g = acceleration

due to gravity.

37
Introducing an unknown multiplier λ , equations (3-1) and (3-2) can be

combined into:

⎡ ⎤ ⎡ λa ⎤ fV | V |
2
g
L = L1 + λ L2 = λ ⎢ H x + H t ⎥ + ⎢V x + Vt ⎥ + =0 (3 - 3)
⎣ λ ⎦ ⎣ g ⎦ 2D

Comparing the terms in the brackets of equation (3-3) with the full derivatives of H and V:

dH dx dV dx g λa 2 dx
= Hx + H t and = Vx + V t , suggests letting = = .
dt dt dt dt λ g dt

g dx
This substitution implies that λ = ± , = ±a and that equation (3-3) can be
a dt

expressed as a system of two pairs of ordinary differential equations, that is, the

following two sets of compatibility equations (3-4) to (3-7). They are also referred to as

the characteristics equations because equation (3-4) is only valid along C+ characteristic

line (3-5) for the arriving pressure wave while equation (3-6) is only valid along C-

characteristic line (3-7) for the leaving wave:

⎧ g dH dV fV | V |
⎪ + + =0 compatibility equation (3 - 4)
+ ⎪ a dt dt 2D
C ⎨
⎪ dx = + a C + characteristic line (3 - 5)
⎪⎩ dt

⎧ g dH dV fV | V |
⎪− a dt + dt + 2 D = 0
−⎪
compatibility equation (3 - 6)
C ⎨
⎪ dx = − a C − characteristic line (3 - 7)
⎪⎩ dt

In Figure 3.1, to solve the unknown point P from the known points A and B,

the compatibility equations can be integrated along AP (C+) and BP (C-), respectively,

38
using the pipe discharge Q and cross-sectional area A instead of velocity V. Here, a linear

approximation is used for the friction term; that is,

a fΔ x ⎫
H j
− H i j−−11 + ( Q i j − Q ij−1− 1 ) + Q ij | Q ij−1− 1 |= 0 ⎪

2
i
gA 2 gDA

a fΔ x
H j
− H i j+−11 − ( Q i j − Q ij+−1 1 ) − Q i | Q i +1 |= 0 ⎪
j j −1
i
gA 2 gDA 2
⎪⎭

These equations are only valid if Δx = a ⋅ Δt , which has to be met for the pipe

a
discretization and time step choice. Introducing two pipe constants B = and
gA

fΔx
R= , the above integrating equations can be written as:
2gDA 2

H ij = H i−j1−1 − B(Qij − Qi−j1−1 ) − RQij | Qi−j1−1 | ⎫⎪



H ij = H ij+−1 1 + B(Qij − Qi+j1−1 ) + RQij | Qi+j1−1 | ⎪⎭

These are further re-written as:

C +: = C P − BPQi
j j
H i (3 - 8)
C −: = C M + BM Q j
j j
H i (3 - 9 )

⎧⎪ C P = H i- 1 j −1 + BQ i- 1 j −1
in which ⎨
⎪⎩ C M = H i +1 j −1 + BQ i +1 j −1
⎧⎪ B P = B + R|Q i- 1 j −1 |

⎪⎩ B M = B + R|Q i +1 j −1 |

Here, the subscript ‘P’ represents positive ‘+’ sign; and ‘M’ represents minus ‘-’ sign in

the characteristic equations. Therefore, starting from the initial flow state (t = 0),

equations (3-8) and (3-9) can be solved, so the transient pressure head H and flow Q at

39
any internal pipe node along the distance x (i =1, 2, 3, ….) and at any time step t = j ⋅ Δt

(j =1, 2, 3,….) can be obtained step by step:

C P BM + C M BP
=
j +1
Hi (3 - 10)
B P + BM
C − Hi H i − CM CP − CM
j +1

= P = =
j +1
Qi (3 - 11)
BP BM BP + BM

Yet, only one of characteristic equations [either (3-8) or (3-9)] is available for the node at

a pipe’s upstream and downstream ends, and one boundary condition at each end has to

be provided to define either the flow Q (t, x=0 or L), or the pressure H (t, x=0 or L), or the

relation between H and Q. Thus, one characteristic equation and one or more boundary

conditions are always combined to determine the transient conditions at the pipe end.

3.1.2 Extension of MOC in Pipe Network

The MOC has been extended to network applications and the resultant solution is

algebraically simple and computationally flexible (Karney and McInnis, 1992). The

extended MOC is implemented in this research for the numerical development of

different AHDs.

In a pipe network, a junction (i.e., several pipe ends meeting together as

depicted in Figure 3.2) is the most common boundary condition and can usually be

assumed frictionless. Let the set N1 be the pipes whose flow direction is assumed toward

the node and N2 the set of pipes whose flow direction is assumed away from the node.

Qext is external flow governed by some boundary condition, with a positive magnitude

referring to flow away from the junction. If there is no external flow at the junction, Qext

40
=0. HP is assumed as the common piezometric head at a junction and QPK the flow in

each pipe. Here the second subscript k indicates a particular pipe in the set. For any pipe

whose flow direction is assumed toward the junction, we use the C+ compatibility

equation (3-8), which gives:

CP H
QP = − P K
K ∈ N1 (3 - 12)
K
BP B P K K

For any pipe whose flow direction is assumed away from the junction, we use C-

compatibility equation (3-9), which gives:

CM H
QP = − + P K
K ∈ N2 (3 - 13)
K
BM BM K K

The continuity equation requires that the total flow is zero at the junction; that is:

N1 N2

∑Q − ∑Q − Q = 0
k =1
PK
K =1
PK ext
(3 - 14)

Substituting equations (3-12) and (3-13) into equation (3-14), the following extended

MOC equation can be obtained:

H P = C ′ − B ′Qext (3 - 15)

in which

−1
⎛ 1 1 ⎞⎟ ⎛ C C ⎞
B ′ = ⎜⎜ ∑ +∑ ⎟ ; C ′ = B ′⎜ ∑ P + ∑ M ⎟
⎜ k∈ N B ⎟
k k
(3 - 16)
⎝ k ∈N B P k ∈ N B M ⎠
1
k
2
⎝ k P
k ∈N B
M ⎠
1
k
2
k

41
Qext

Figure 3.2 Generalized node with an external flow

It is interesting that the equation (3-15) at a junction has the same form as the

C+ compatibility equation for the downstream end of a simple pipe. In this way, any

junction within a network (with or without external flow) can be treated as equivalent to a

downstream node. Note that if only one pipe is connected to a junction (the so-called one-

degree-node) we can deduce B ′ = BP or BM , and C ′ = C P or C M from equation (3-16).

This indicates that the extended MOC equation is a generalized solution for any junctions

in the system, implying a higher efficiency in numerical modeling. It should be noted

that the junction without any connecting pipe (the so-called zero-degree-node) is not

allowed in the extended MOC model because B Pj = BM j = 0 would occur in equation (3-

16). Meanwhile, a junction only permits attachment of one “device”. When more than

one device is associated with a junction, the extra devices have to be treated either by

combining them as one “device” or setting at least one characteristic section of the

connection pipe between them when the system is discretized and modeled.

42
3.2 Controllers and Active Hydraulic Devices

Active hydraulic devices are intrinsically associated with control theory and automatic

controllers. In mathematics and engineering, control theory deals with the behavior of

dynamical systems. The desired output of a system is called its reference. When one or

more output variables of a system need to follow a certain reference over time, a

controller attempts to manipulate the input to the system to obtain the desired output

effect.

3.2.1 Open-loop and Closed-loop control

There are two basic types of control, open-loop control and closed-loop control. For

example, in an automobile's cruise control system, the output variable is vehicle speed

and the input variable is the engine's throttle position which influences engine torque

output. A simple way to implement cruise control while driving is to lock the throttle

position. However, on hilly terrain, the vehicle will slow down going uphill and

accelerate going downhill. This type of control is called open-loop control because there

is no direct connection between the output of the system and its input. A drawback of

open-loop control is that it requires perfect system knowledge (i.e., it needs to know

exactly what inputs will produce the desired output). Open-loop control can be used in a

well-characterized system so that one can adequately predict which inputs will

necessarily achieve the desired output. It is usually a simple system, and the mathmtaical

relationship between inputs and outputs is clear and an analytical solution is possible.

For example, the rotational velocity of an electric motor may be sufficiently characterized

for the supplied voltage to make feedback unnecessary.

43
To avoid the shortcomings of open-loop control, feedback is introduced into a

controller to regulate the operational states of a dynamic system. The process output is

measured with sensors and acted upon by the controller; the result (i.e., control signal) is

input to the process, closing the loop. This control type is known as closed-loop control,

which offers certain key advantages over its open-loop counterpart, such as disturbance

rejection, a stabilized pocesses, reduced sensitivity to parameter variations, improved

reference tracking performance, and guaranteed performance even with model

uncertainties when the model structure does not match perfectly the real process and the

model parameters are at least partly inaccurate. For instance, using the closed-loop

control in the automobile's cruise system, the engine's throttle position can be adjusted to

reach the desired driving speeds no matter what the road condition is (Tan et al., 1999;

http://en.wikipedia.org/wiki/Controller_(control_theory)).

3.2.2 Feedback and Feedforward Control

Closed-loop control is also called feedback control since, in this form of system

regulation, the sensed process outputs are used as “feedback” to control subsequent

outputs or states of a dynamic system. The household thermostat is an example of

feedback control. The feedback controller relies on measuring the governing variable, in

this case the temperature of the house, and then adjusting the output, whether the furnace

or heater is on or off. However, feedback control usually results in intermediate periods

where the controlled variable is not at the desired set-point. With the thermostat example,

the furnace or heater would switch on only after the house temperature falls below such a

point.

44
In some systems, closed-loop and open-loop control are used simultaneously.

In such cases, the open-loop control is termed feedforward and serves to further improve

reference tracking performance. Feedforward control can avoid the slowness of feedback

control because, with feedforward control, the disturbances are measured and accounted

for before they have time to affect the system. In the thermostat example, a feedforward

system may measure the fact that a door is opened and automatically turn on the furnace

before the house can get too cold. The difficulty with feedforward control is that the

effect of the disturbances on the system must be accurately predicted and there should be

few surprise disturbances. For instance, if an unmonitored window were opened, the

feedforward-controlled thermostat might still let the house cool down.

To achieve the benefits of feedback control (controlling unknown disturbances

and not having to know exactly how a system will respond to disturbances) and the

benefits of feedforward control (responding to disturbances before they can affect the

system), there are combinations of feedback and feedforward that can be used. Dead-

time compensation and inverse response compensation are examples where feedback and

feedforward control are used together. Dead-time compensation is used to control devices

that take a long time before exhibiting any adjustment to an input change and uses an

element to determine how an instant change in the controller’s input will affect the

controlled variable (output) in the future. The controlled variable is also measured and

used in feedback control. Inverse response compensation involves controlling systems

where a change at first affects the measured variable one way, but then later affects it in

the opposite way; for example, when one eats candy, at first it provides an energy boost,

but subsequently, one may feel tired or even lethargic. In this way, it can be difficult to

45
control this type of system with feedback alone. Therefore, a predictive feedforward

element is necessary to inform of the reverse effect that a change in state might entail in

the future (Tan et al., 1999; http://en.wikipedia.org/wiki/Controller_(control_theory)).

3.2.3 Evolution of Controllers

Controller design evolves with the development of associated science and technology.

Most control systems in the past were implemented using mechanical systems or solid

state electronics. Pneumatics were often utilized to transmit information and control using

pressure. Nowadays, most systems and industries rely on computers for controller design,

which obviously renders implementation of complex control algorithms easier than with

a mechanical system.

On/Off Control. On/Off control is the simplest and most intuitively akin to

direct manual control, like the thermostat that turns the furnace on or off depending on

the ambient temperature’s departure from the set-point. It is undoubtedly the most widely

used control type for both industrial and domestic service. It has an output signal u,

which may be changed to either a maximum or a minimum value, depending on whether

the process variable is greater or less than the set-point. As shown in Figure 3.3, the

control law is described by (Tan, et al., 1999):

u = u max , e > 0,
u = u min , e < 0,

where e is the control error, e = r − y ; y is the controlled process variable and r is its set-

point (reference). The minimum value of the control output is usually zero (off), the

46
maximum value of the control output could be either a positive or negative constant. The

mechanism to generate On/Off control is usually a simple relay.

One well acknowledged and significant disadvantage of On/Off control is that

it oscillates around the constant set-point. This will directly affect the process variable

which also oscillates around the desired value. If, for example, On/Off control is applied

to regulating a tank’s water level with the aid of a valve which can only open or shut

completely, the On/Off controller therefore naturally opens and shuts the valve

alternatively. This can cause rapid wear and tear on the moving parts in the actuating

device and the “ringing” or oscillation phenomenon may be intolerable in certain cases.

The solution in most On/Off controllers is to establish a dead zone, or hysteresis, of about

0.5% to 2% of the full range. This dead zone straddles the set-point so that no control

action takes place when the process variable lies within the dead zone.

u
umax

umin
e

Figure 3.3 ON/OFF control

Propotional/Correspondance/Modulating Control. Except for the use of a

small dead zone to mitigate signal oscillations associated with On/Off control, one

alternative is to use a small gain for the controller when the error e is modest and use a

47
large gain when the error is large. This can be achieved with a proportional or P-

controller, a basic continuous control mode. The control rule in a P-controller is given by

(Tan, et al., 1999):

u = u max , e >e 0 ,
= u 0 + K c e, − e0 < e < e0 , (3-17)
= u min , e < e0 ,

where u0 is the level of control signal when there is no control error, and Kc is the

proportional gain (or sensitivity) of the controller. In fact, Kc indicates the changes in

control signal per unit change in the error signal; it is indeed an amplification and may be

adjusted by the operator. The proportional gain Kc will be positive if an increase in the

input variable requires an increase in the output variable (direct-acting control), and it

will be negative if an increase in the input variable requires a decrease in the output

variable (reverse-acting control). A typical example of a direct-acting system is

controlling the flow of cooling water. When the temperature increases, the flow must be

increased to maintain the desired temperature. Conversely, a typical example of a

reverse-acting system is controlling the flow of steam for heating. If the temperature

increases, the flow must be decreased to maintain the desired temperature. The P-

controller can also be described graphically as in Figure 3.4.

While the signal oscillations with On/Off control may be quenched with

proportional control, a new problem arises. For most systems, the P-controller will never

entirely remove errors at steady state. In other words, after the transients have died down,

there may remain a deviation between the set-point and the process variable. Since the

48
u − u0
control error is given by e = , at least one of the following conditions must be true
Kc

to meet the zero control error at the steady state (Tan, et al., 1999):

1) Kc is indefintely large, which effectively reverts P-control back to On/off control,

bringing with it the issue of instablity or requiring physical impossibilities like

infinitely large valves;

2) The steady state control signal, uss, equals, the level of the control signal when there is

no control error, u0, that is, uss = u0. If so, when the error is zero, the controller only

provides the steady state control action so the system will settle back to the original

state which is probably not the new set-point desired for the system. In other words,

uss = u0 cannot be generally satisfied by all set-points. Even if u0 can be adjusted

relative to the set-point, it is still necessary to know at least the process static gain

before the adjustment can be done.

u
umax

u0

umin

-e0 0 e0 e
Proportional Band

Figure 3.4 Proportional control

49
PI control. To eliminate the remaining steady state error with pure P-control,

the integral or reset action can be introduced. The I-part control is able to find the correct

value for u0 automatically in response to any set-point without needing to know the

process static gain. Integral control action usually is combined with proportional control

action, although it is possible but less common than using integral action by itself. The

combination (PI-control) is favorable because of the advantages of both control actions.

The control signal in a PI-controller is given by (Tan, et al., 1999):

⎛ 1 ⎞
u = K c ⎜⎜
⎝ Ti
∫ edt + e ⎟⎟

(3-18)

where Ti is the integral time of the controller. The constant level of u0 found in the P-

controller has thus been replaced by the integral:

Kc
Ti ∫
u0= edt (3-19)

The integral of the control error is effectively proportional to the area under the curve

between the process variable and the set-point (Figure 3.5).

To better understand the capacity of a PI-controller for the elimination of

steady state error, it is assumed that the closed-loop system is stable and that a steady

state control error existed despite having a PI-controller, which implies e ss ≠ 0 . At steady

state, the integrator will continuously accumulate the error signal at the input and thus the

control signal u will be either rising or falling, depending on whether the error is positive

or negative. The P-part has a constant value corresponding to Kcess and thus will not

50
affect the analysis. If the control signal rises or falls, the process variable will also rise or

fall. This in turn means that the error e = r - y cannot be constant at steady state and thus

contradicts the assumption that the error is stationary. Thus, it is impossible to have a

non-zero steady state error when the controller has an intergral part and the closed-loop is

stable.

Figure 3.5 Integral control derived from the area under the control error curve

A PI-controller thus resolves the problem of remaining stationary error and the

problem of oscillation associated with On/Off control. A PI-controller is therefore an

effcient controller without any significant faults and is often sufficient when the control

requirements are low to modest.

PID control. Both the P- and the I-componets of a PI-controller operate on

past control errors, and do not attempt to predict future control errors. This characteristic

limits the achieveable performance of the PI-controller. The problem is more clearly

illustrated in Figure 3.6. The two curves in the figure show the time graph of the control

51
error for two different processes. At time t, the P-part, being proportional to the control

error, is the same for both cases. Assuming that the I-parts, being proportinal to the area

under the two control error curves, are also equal in the two cases. This means that a PI-

controller gives exactly the same control signal at time t for the two processes. However,

it is clear that there is a great difference between the two cases if the rate of change in the

control error is considered. For Repsone I in Figure 3.6, the control error changes rapidly

and the controller should reduce its output to avoid an overshoot in the process variable

occuring in the future. For Response II, the control error changes more sluggishly and the

controller should react strongly in order ro reduce the error more rapidly. Derivative or

rate control indeed carries out this type of compensation.

e e

I II
P P
t t

Figure 3.6 Illustration of the need for derivative control

It is not even theoretically possible to have a control action based solely on the

rate of error signal change (de/dt), since the derivative controller output would be zero if

the error is large but unchangaing. Thus, derivative control is usually combined with at

least a proportional control. The D-part of PID controller is proportional to the predicted

52
error at time t+Td, where Td is the derivative time of the controller. The control law for

the PID controller is described as (Tan, et al., 1999):

⎛ 1 de ⎞
u = K c ⎜⎜ e + ∫ edt + Td ⎟⎟ (3-20)
⎝ Ti dt ⎠

Since set-point r is normally constant or has rather abrupt changes (to reflect

changes in circumatnces), it will thus normally not contribute to the derivative term. For

this reason, it is common parctice to apply the derivative action only to the process output

y, except when the process variable is required to track a continually changing set-point.

3.2.4 Elements in a Control System

As shown in Figure 3.7, a control loop usually consists of the following hardware

elements:

1) Sensors that monitor the ongoing process condition;

2) Data transmission system (wireless, wire, or cable transmitters) that carries the

measured data from sensors to controller and signals from controller to the final

control element;

3) Controller that compares the “process variable” received from the transmitter with the

set-point (i.e., the desired process), and then makes decisions and sends corrective

signals to the final control element;

4) Actuator is a pneumatic, hydraulic or electrically powered device that supplies force to

open or close the valve, or moves the other final control elements.

53
5) Final control element that manipulates the process. Control valves are the most

commonly used final control elements; others include wicket-gates in turbine units;

MGV (misaligned guide vanes) device at high-head pump-turbines, VFD (variable

frequency drive) on pumps; dampers; and devices that regulate (throttle) electric

energy such as silicon-controlled rectifiers.

PID
Controller
Control signal for
valve operation

Sensor

Pipeline Control Valve

Figure 3.7 Loop of PID-control-valve system

AHDs are typically combined with automatic controllers, and different types of

AHDs use different controllers and actuators. For instance, an On/Off controller or P-

controller is used in pressure relief valves (PRV), check valves, and air valves; and a PID

controller is used in some regulating control valves, speed-governed hydro-turbines, and

variable speed pumps.

54
3.3 Generalized Mathematical Model for AHD Transient Simulations

A general mathematical model for a PID controlled AHD is presented next; specific

features, design functions and case studies for individual devices are discussed in the

following chapters.

3.3.1 Extended MOC Equations

Figure 3.8 shows an AHD in a pipe network. Using the extended MOC equation (3-15) to

relate the nodal heads to the flow rate across the device, the compatibility equations at

AHD’s upstream and downstream node, respectively, are written as:

H1

H2
Pipe
Pipe

Q Q Pipe
Pipe Node1 Node2
AHD

Figure 3.8 Active Hydraulic Device in a pipe network

55
H 1 = C1′ − B1′ Q (3 - 21)
H 2 = C 2′ + B2′ Q (3 - 22)

in which B '1 , B ' 2 , C1′ and C2′ are calculated using equation (3-16). By combining the two

extended MOC equations, a head-flow relationship at the AHD is obtained:

H = H 1 − H 2 = (C1' + C 2' ) − ( B1' + B2' )Q (3 - 23)

in which, H = the head difference across the AHD. This equation links the device to the

reminder of whole system, since it will be resolved by a moment-by-moment simulation

in the connected pipe system starting from the initial steady state.

3.3.2 AHD’s Dynamic Characteristics

As mentioned in Chapter 1, a control system is essential to make a hydraulic device

“active”. The head-discharge relationship of an AHD can usually be dynamically

modified via its control system to change the opening of its control valve or wicket-gates.

The generalized dynamic characteristics of an active hydraulic device is expressed as:

dy
F (Q, H , y, , ...) = 0 (3 - 24)
dt

in which y = the opening of control valve or turbine wicket-gates and Q = the flow rate

passing through the AHD. For a control valve, equation (3-24) is the valve discharge

equation, representing the relationship between the valve resistance and valve opening;

more specifically, that is equation (4-3) in Chapter 4. For a hydro-turbine, equation (3-24)

represents a set of characteristic curves (reference to different turbine characteristics

shown in Section 5.3 Chapter 5).

56
3.3.3 PID Controller Equation

Instead of a predefined opening or closing motion for most conventional valves with

On/Off controllers, the opening of a PID control valve is automatically regulated in

response to the sensed pressure or pressure difference, which is desired to control.

Consequently, y (t) in equation (3-24) is unknown in a PID control valve and needs to be

dynamically evaluated by the characteristics of the PID controller. While for a speed-

governed hydro-turbine, the wicket-gate opening is adjusted instantaneously by tracking

the rotational speed of the turbine unit in response to the variation of power load on the

unit.

Depending on the actuator’s power source (pneumatic, electric or hydraulic)

and the internal design of a valve and controller, the equation of the PID controller could

be established corresponding to each process variable for which control is sought.

Usually, we need to establish a relationship between the PID control signal (amplified

control error) and the actuator action, which is expressed by:

1 de
∫ edt + T
t
u (t ) = − K C ( e + d )
Ti 0 dt
dy
= F ( y, ,...) (3 - 25)
dt

Here, KC, Ti and Td are three parameters of PID-controller, representing the

dy
characteristics of a controller. The function F ( y, ,...) represents the action of the
dt

actuator. More specifically, equations (4-5) and (4-6) are two different actions taken by

the PID control valve for different control variables. And the turbine governor equation

57
(6-19) is another example of PID controller equation, which is to adjust wicket gate

opening for the speed control of turbine unit.

In sum, the mathematical model of a PID controlled AHD basically consists of

three governing equations, as presented above. However, it is possible to have additional

governing equations for some AHDs. These additional equations might include a torque

balance equation for a pump or hydro turbine, reflecting the variable power load or the

speed change of the turbine unit, e.g., equation (6-17) in Chapter 6

3.4 Summary

In this chapter, the Method of Characteristics (MOC) and its implementation in pipe

networks is first reviewed and then a brief introduction of control theory, as well as the

evolution and variety of controllers, is provided. Finally, a mathematical model for the

transient analysis of a general AHD, combined transient response with control theory, is

presented. AHD models must capture three specific relationships:

1. the characteristics of the hydraulic component, that is, the heart of the component,

whether valve, turbo machine or storage element;

2. the control behavior, that is, what action the device takes as a result of the hydraulic

event it creates or experiences; and

3. the wave behavior of the reminder of the system, captured in this case by the Method

of Characteristics (MOC), that is a moment-by-moment simulation of arriving and

leaving waves.

58
With the fundamental understanding of AHDs, the next chapter develops a new

active device to create a “non-reflective” boundary condition in a pipeline through a PID

control valve combined with a remote pressure sensor.

59
Chapter 4 “Non-Reflective” Boundary Design via

Remote Sensing and PID Control Valve

This chapter represents one of the core contributions of thesis research. Using the

concept of AHD, a new strategy is created to actively control the reflection and

resonance of pressure waves in a pipeline system, using a so-called “non-reflective”

boundary. The system performance is evaluated and compared for both scenarios with

and without application of this strategy, and the role of “non-reflective” or “semi-

reflective” boundary design is demonstrated.

The core results of this chapter have been published and presented at

Pressure Surges 10th International Conference (BHR Group, Edinburgh, Scotland, May

14-16, 2008), co-authored with Bryan Karney, Feng Wang and Stanislav Pejovic. As the

second author, I wrote the full early draft of the published paper, developed the

numerical model and did all the numerical runs and analysis. The initial idea of this

work originated with Prof. Karney and Stanislav Pejovic; Feng Wang helped to

incorporate the model into the program TransAM and also was involved in discussion of

the details of model “conception” during the time when he visited the University of

Toronto in 2006.

Abstract: This chapter explores a new approach to limit the surge pressures by

creating so-called “non-reflective” (or “semi-reflective”) boundary in a pipe system. This

60
boundary condition is established using the combination of a remote sensor and a control

system to operate a relief valve. In essence, the idea is to sense the pressure change at a

remote location (say, at a tee to a dead-end link) and then to use the measured data to

adjust the opening of an active control valve at the end of the line so as to eliminate or

attenuate the wave reflections at the valve and thus to control system transient pressures.

This novel idea is explored here initially by the means of numerical simulation and

shows considerable potential for transient protection, as demonstrated in case studies.

Using this model, wave reflections and resonance can be effectively eliminated for

frictionless pipelines or initially no-flow conditions and better controlled in more

realistic pipelines for a range of transient disturbances. In addition, the feature of even

order of harmonics as well as “non-reflective” boundary conditions during steady

oscillation, obtained through time domain transient analysis, are verified by hydraulic

impedance analysis in the frequency domain.

4.1 Introduction to “Non-Reflective” Boundaries

One of the most interesting aspects of transient events is that the phenomenon is

primarily created and controlled by the action of boundary devices in the system. The

actions of, or changes in, these boundaries both initiate the transient event and control its

severity, creating at operating devices a sequence of velocity and pressure fluctuations

that then propagate through the pipe system. The action or response of other devices and

components, either by design or by accident, then reflects, refracts, amplifies or

attenuates the primary pressure and velocity waves.

61
There is an intriguing connection between water hammer control and the use

of “tailored” or active boundary conditions that control wave reflections. Certainly such

“non-reflective” boundary conditions have been partially explored before and used in the

past to represent certain network junctions and components (Almeida and Koelle, 1992;

Pejović-Milić A., Pejović S, and Karney B., 2003). Indeed in a conventional sense,

various “semi-reflective” boundaries are the basis of water hammer protection, and their

role as general energy dissipation device in systems with dead-ends is extended and

explained in the following sections; however, as some devices can be sources of

resonance, special consideration must be given to the frequency domain. Moreover, the

physical, mathematical and numerical aspects of such boundaries are considered and

developed along with possible applications that move these boundary conditions toward

use and application in systems and devices.

With the development of electronic and computer technologies, dynamic PID

(proportional, integral and derivative) controllers are now being used more frequently to

maintain a desired performance in a water system; the variable to be controlled may be

turbine speed, turbine power output, pump speed, pump torque, pump discharge flow,

water level in tanks, upstream or downstream pressure at valve, and so on. This

developed control capability also provides an alternative arrangement for suppression of

hydraulic transients in pipe systems. In fact, local/conventional PID control valves, even

when modulating, have implications for transient control by maintaining a desired

pressure or flow in the system. However, the combination of remote sensing and PID

control valve could provide a broad range of waterhammer protection through designing

of “non-reflective” boundaries into a pipe system. Yet, the relevant literature to date has

62
mainly focused on the responses of transient flow to the action of PID control valves by

coupling hydraulic transient analysis and control theory (Koelle 1992, Koelle & Poll

1992, Lauria & Koelle 1996, Bounce & Morelli 1999, and Poll 2002), it is an innovation

to study how to actively control a valve in response to the remotely sensed pressure for

system waterhammer protection.

The goal of this chapter is to explore how “non-reflective” boundaries can be

designed and applied for transient protection, particularly at problematic locations like

dead-ends or cul-de-sacs in a distribution system. Though having the same role to limit

the transient pressure in pipe systems, this approach is different from the operation of a

pressure relief valve (PRV) that opens a by-pass line to release excessive flow when the

pressure in pipeline exceeds the set point. Such as an arrangement is also quite distinct

from a local/conventional pressure sustaining valve (PSV) or backpressure valve

(holding the pressure at valve inlet or outlet) that modulates the valve opening to

maintain the set point corresponding to the locally sensed pressure (Hopkins, 1998). The

key issue in the remote control is that the transformation of transient pressure waves

between the remote sensor and active control valve.

In this chapter, an example is first presented to demonstrate the possibility of

dangerous waterhammer occurring in the system due to reflections at dead-ends. The role

of a control valve to dissipate the transient energy and thus protect the system from

excessive transient pressures is illustrated. Then, the mathematical model for the

local/conventional PID control valves is addressed as a prerequisite to the solution of

remote sensing and “non-reflective” valve opening, and the key novel features in the

remote control model are discussed. After that, case studies involving a successful

63
numerical application of the remote control model are presented. Theses case studies

show the ability of “non-reflective” boundary to control the reflection of pressure wave

and potential resonance within the pipeline. Moreover, the developed “non-reflective”

boundary condition during the steady oscillation is verified using hydraulic impedance

analysis in frequency domain. Finally, using the developed simulation tool, the selection

and tuning of PID controller parameters are discussed based on sensitivity analyses.

4.2 Transient Performances with Dead-ends and Valve Control

A dead-end is often a tricky arrangement in a pipe system. Such an arrangement includes

a closed valve, which carries no discharge and can cause unexpected high pressure in the

system. When a pressure wave is transmitted into the dead-end pipe, the flow or velocity

is stopped rapidly at the end and the wave transmission also terminates there, which

induces a doubled value in local pressure and creates an increased pressure wave that

returns into the system. This is the so-called dead-end reflection. By contrast, a pressure

wave would be fully reflected with reverse sign from a constant head reservoir. In other

words, neither a reservoir nor a dead-end is intrinsically dissipative; they both reflect

waves but conserve energy. However, a partially open valve at a reservoir dissipates

energy and acts somewhat between a dead-end and a reservoir. If the size of the valve

opening is systematically adjusted, a value can be found for a given system so that the

disturbance/excitation does not reflect at all. This setting thus produces a “non-

reflective” boundary with a maximum rate of transient/oscillation energy dissipation.

64
Consider a pipeline system shown in Figure 4.1. Initially, the terminal control

valves at the end of branches are fully closed. If the pipes do not leak and have been

open to the reservoirs for some time, no flow will occur in the system and the head will

be uniformly 100 m as found in the reservoirs. However, following Boulos et al.

(Boulos, Lansey and Karney 2006), suppose the second section of the main pipeline

(from B to C) is pressurized to a uniform value of 130 m. If this initial condition is

released, two pressure waves, both 15 m in amplitude, are created and propagate into the

system. The response to the traveling pulse waves are simulated and shown in Figure 4.2,

representing the envelope of maximum and minimum transient pressures along the pipe

length from A to C and then from C to F. More specifically, in Figure 4.2 (a), the

terminal valves keep fully closed, and we see that the wave is reflected and magnified by

the dead-end due to the overlapping of incident and reflected pressure waves. In Figure

4.2 (b), the terminal valves now are opened to 10% of the full size in 1 second when the

pulse waves start to travel, which reduces the maximum pressure but causes the

unacceptable negative pressure. Figure 4.2 (c) represents the condition that the terminal

valves are opened to 0.35% of full size in 1 second, and in this case the wave energy is

largely dissipated when it arrives at this small orifice. These differences in system

response can be exploited. Indeed, even when the valve opening is chosen somewhat

arbitrarily, the transient pressures are likely to be at least partly mitigated. So, what if the

valve opening is systematically refined to eliminate the wave reflection?

65
130 m

100 m 100 m

A B C D E
80 m

F G

Figure 4.1 Scheme of a branched system and initial pressure head

(Note: Pipe length AB = BC = CD = DE = CF = DG = 1000 m, friction factor f = 0.012, and the


diameter of main pipeline D = 1 m, and the diameter of branches d = 0.5 m.)

Max. H

Initial H

Pipeline

Min. H
A B C F

(a) Transient response in the system with dead-end branches

66
Max. H

Initial H
Pipeline

Min. H
A B C F

(b) Transient response in the system with 10% orifices at branch ends

Max. H

Initial H

Pipeline Min. H

A B C

(c) Transient response in the system with 0.35% orifices at branch ends

Figure 4.2 Comparison of transient responses to dead-ends and small orifices

67
(a) Sketch of system configuration

(b) Initial pressure head along the pipeline

Valve Opening=0% Valve Opening=10% Valve Opening=20%

Valve Opening=30% Valve Opening=60% Valve Opening=100%

(c) Transient pressure head along the pipeline after reflections from both ends

Figure 4.3 Traveling pulse waves and “tailored” valve reflections

68
Further insight is obtained by comparing the responses for different sizes of

valve opening in a single pipeline system, sketched in Figure 4.3 (a). In this system, a

uniform pipeline links two reservoirs with the same constant head; and a control valve is

installed at the right hand reservoir. A traveling pulse wave is initially created within the

middle section as described in the literature (Boulos, Lansey and Karney 2006), see in

Figure 4.3 (b). Figure 4.3 (c) shows the sign of the wave reflection shifts as the right

hand boundary shifts with the valve opening increasing from fully closed (i.e., dead-end)

to fully opened (i.e., constant-head reservoir). By systematic adjustment, a valve

opening of about 30% is found in this system to eliminate all wave reflections.

In practice, such “non-reflective” boundary would and could only be “tailored”

by an automatic control valve, which measures and dynamically adjusts in response to

the incoming pressure waves.

4.3 Mathematical Model for Local/Conventional PID Control Valve

PID control valves are usually installed at the connection of subsystems to maintain the

desired operating condition in hydraulic networks. Most probably, the subsystems were

originally designed for separate operation and then have been connected or expanded due

to the urbanization and development of water distribution networks (B. Bounce & L.

Morelli, 1999). For instance, to ensure the minimum flow demand in downstream

subsystem, an accurate and continuous control of the pressure or flow rate in the

connection pipeline is needed.

69
The mathematical model and numerical simulation of PID control valves

provides a tool to better understand the system hydraulics. Usually, there is a built-in

PID controller and sensor at a control valve, as shown in Figure 3.7. For different control

variables, the mathematical models are largely the same except for a slight difference in

the PID controller equation.

4.3.1 Extended MOC Equations

To simulate a PID control valve in a pipe network (reference to Figure 3.8), the extended

Method of Characteristics (MOC) is used to calculate the nodal heads at the valve inlet

and outlet, respectively (Karney and McInnis, 1992). The MOC equations are repeated

here:

H 1 = C1′ − B1′ Q (4 - 1)

H2 = C2′ + B2′ Q (4 - 2)

in which B '1 , B ' 2 , C1′ and C2′ are calculated as shown in Chapter 3.

4.3.2 Valve Discharge Equation

Valve discharge equation defines the relationship between the flow passing through a

valve and the head difference across the valve. The commonly used valve discharge

equation is as follows:

Q = τ ES H1 − H 2 ( 4 − 3)

70
in which ES is a valve conductance parameter determined by the energy dissipation

potential of the valve, and τ is the dimensionless valve opening. For the steady flow Q0

Q0
and the corresponding head loss of H0, τ = 1 and E S = ; and for no flow case
H 10 − H 20

with the valve closed, τ = 0 (Wylie, et al., 1993).

Instead of the predefined opening or closing motion of a conventional valve,

the opening of a PID control valve is adjustable in response to the sensed pressure or

pressure difference, which is desired to be controlled. That is, τ (t ) is unknown for a PID

control valve and needs to be dynamically determined by the characteristics of the PID

controller.

The valve relationship, equation (4-3), is suitable for quasi-steady flow only

and might not be valid for rapid transient flow, because during transient state it may

occur that the flow direction is inconsistent with the head difference across the valve

when valve opening is significantly small. This phenomenon has similar influence as the

backlash or dead time of the valve. Future research should consider this inconsistency in

the model, but the challenge is that there is a little data on transient behavior of valves

and other components of the hydraulic system.

4.3.3 PID Controller Equations

The output signal or response of a typical parallel-structured PID controller is given as

(Tan, et al., 1999, page 9):

1 t de
r (t ) = K C (e +
Ti ∫
0
edt + T d
dt
) (4 - 4)

71
where e is the controller error, that is the deviation of the process variable u(t) from its

set point u*. Usually, the desired set point u* is a given constant and the dimensionless

u (t )
error is defined as e = 1 − . For a control valve, u(t) could be either of the inlet
u*

pressure head H1(t), outlet pressure head H2(t), or the flow passing through the valve Q(t).

In a physical system, the values of these control variables are continuously measured by

sensors; while in numerical counterpart, the values of these variables are simulated step

by step.

Constants KC, Ti and Td are, respectively, the proportional gain, integral time

and derivative time constants of a PID-controller; they represent the characteristics of the

controller. The questions about how to select these parameters and how they affect the

speed and stability of the control process are discussed in the last section of this chapter.

The control law expressed by equation (4-4) is general for all types of PID-

controllers. It is straightforward to set the parameter Td to zero for a PI-controller;

furthermore, for a controller that has proportional part only (i.e., a P-controller), Ti is

given a large value. Besides, for a series-structured controller with given parameters, the

parameters for corresponding parallel type can always be obtained by the relationships

between these two controller structures. On the other hand, given the parameters for a

parallel type controller, it is not always possible to obtain the corresponding parameters

for the series type (Tan, et al., 1999, page 17). Therefore, equation (4-4) for parallel type

PID controllers will work for series type as well by transforming the given parameters to

those for the corresponding parallel type.

72
Depending on the power source of the actuator (pneumatic, electric or

hydraulic) and the internal design of the valve and controller, the operational equation of

the PID controller could be established corresponding for each specified process variable.

For the control of pressure at the valve inlet, H1, the output signal (the amplified error) of

the controller is set equal to the rate of valve flow reduction (Hopkins, 1998); that is,

1 t de dQ
r (t ) = K C ( e +
Ti ∫
0
edt + T d
dt
)=−
dt
(4 - 5)

H 1 (t )
here e = 1 − . The hydraulic implication of the above control action can be
H1 *

explained as follows: when pressure H1(t) at the inlet of the valve begins to exceed the

set point H1* ( e < 0 ), the valve would open slightly to discharge the excess water

dQ
volume ( > 0 ). By contrast, if the pressure H1(t) at the valve inlet begins to decay
dt

below the set point H1* ( e > 0 ), the valve would throttle and reduce the discharge

dQ
( < 0 ). With a PID controller, the automatic adjustment of the valve opening would
dt

be smooth and continuous.

Similarly, for the pressure control at the outlet of the valve H2, the output

dQ
signal of the controller would be set equal to the rate of valve flow increase ( + ).
dt

That is, the minus sign is removed from the right-hand side of equation (4-5) and one

H 2 (t )
would use e = 1 − instead.
H2 *

73
For the control of valve flow rate Q, the signal of control error is set equal to

the change of the head difference across the valve, H, and this change is based on the

initial steady state, that is,

1 t de
r (t ) = K C ( e +
Ti ∫ edt + T
0
d
dt
) = H 0 − H (t )

= [ H 10 − H 20 ] − [ H 1 (t ) − H 2 (t ) ] (4 - 6)

Q(t )
here e = 1 − .
Q*

In sum, for a local/conventional PID control valve, two MOC equations, the

valve discharge equation and the PID controller equation constitute the mathematical

model of the boundary condition in the pipe network. Therefore, the four unknown

variables at valve boundary ( τ , H1, H2, Q) can be numerically resolved using the finite

difference method.

4.4 Consideration and Model of “Non-Reflective” Valve Opening

Based on the developed mathematical model for a local PID control valve, the physical

consideration and mathematical model to search for a “non-reflective” boundary by

using remote sensing and PID control valve is elucidated in this section.

Theoretically, a remote sensor could be located either upstream or downstream

of the control valve wherever an incident pressure surge occurs, since the pressure surge

will usually propagate towards both directions. However, this research focuses on the

74
case of upstream incident surge, and the main idea is illustrated by a simple system

shown in Figure 4.4.

f=0

f = 0.012

H1
HR
H1 H2

Sensor Q Valve

Figure 4.4 Case studies: System scheme and initial steady state

In this system, it is assumed that a series of pressure surges, HR(t), (say, that

represent periodic waves following a sinusoidal law) occur at an upstream reservoir due

to a certain disturbance, where a sensor is installed and linked to a PID control valve at

downstream reservoir. The pressure waves would then propagate at speed a towards the

downstream end along the pipeline. The wave would arrive at the valve inlet within L/a

seconds. To limit the wave reflection from the valve, the PID controller adjusts the valve

opening continuously and accurately to "absorb" the upcoming incident wave when it

passing through the valve (i.e., to dissipate the wave energy at the valve). As a result, at

any instant time t, the pressure at the valve inlet H1(t) should maintain the same

magnitude as the upcoming wave when it reaches the valve. In other words, the set point

75
of valve inlet pressure, H1*, would be equivalent to the “successor” of the upstream

reservoir pressure at a L/a time ahead, i.e., HR(t-L/a). Since the upcoming pressure

changes with time, the set point, H1*, must change with time as well, which is so-called

time-variable or dynamic set point, H1*(t).

The pressure at the valve inlet H1(t) will be tracked in this “non-reflective”

boundary model, correspondingly, the mathematical model for the local PID control

valve with H1 control case, including the extended MOC equations (4-1) and (4-2), the

valve discharge equation (4-3) and the PID controller equation (4-5), is applicable.

However, there is a key difference in the controller equation; in particular, the control

error e in “non-reflective” boundary model is the deviation of instant pressure head H1(t)

H 1 (t )
from its dynamic set point H1*(t), that is, e = 1 − . Instead of a given constant for
H 1 * (t )

the set point in the local control valve, the set point H1*(t) in this case is an unknown

variable (transmission of HR), dependent of the remote pressure waves and pipe system

features, which could significantly complicate the discretization of the governing

equations and their numerical solution. Yet, if H1*(t) could be predicted at any instant

time t, then the PID controller will send the actuator a series of “commands” (the error

signals) to adjust the valve opening continuously to achieve H1(t) = H1*(t). Therefore,

the only remaining issue is how to determine the dynamic set point, H1*(t), according to

the remotely sensed pressure HR(t)?

For a frictionless pipeline without reflections from the downstream boundary,

the upstream pressure wave HR(t) would have no change when it propagates to the front

of valve in L/a seconds. Thus, at any instant, the set point for the pressure at the valve

76
inlet is equivalent to the pressure at the remote sensor that has been recorded at prior

time of L/a. Mathematically, we have:

H 1 * (t ) = H R (t − L / a ) (4 - 7)

However, for a more realistic pipeline having friction resistance, the

magnitude of a pressure wave decays somewhat as it propagates downstream. The key

and challenging question is how to transform transient pressures between the two ends of

the pipeline with friction. Since no analytic solutions are available, the set point must be

estimated based on the remotely measured pressure waves and pipe system features. In

this research, a relative Friction Decay Rate (FDR) is introduced based on the hydraulic

grade line at initial steady state, which is defined as the ratio of pressure heads at the

valve inlet (H10) and at the remote sensor (HR0):

H 10
FDR = (4 - 8)
H R0

Certainly, FDR is system specific but constant for a particular system with certain initial

hydraulic condition. Using the FDR defined in equation (4-8), the magnitude of pressure

wave when it arrives at the valve could be largely estimated:

H 1 * (t ) = F DR × H R (t − L / a ) (4 - 9)

Note that equations (4-8) and (4-9) are consistent with and valid for

frictionless pipelines as well. When the pipeline friction is negligible we have HR0 = H10

and FDR = 1, and thus equation (4-9) reduces to equation (4-7). Therefore, for both

frictional and frictionless pipelines, the set point in controller equation (4-5) could be

77
dynamically estimated using equations (4-8) and (4-9). Interestingly, for those cases

with an initially static state (i.e., the initial flow Q0 = 0 in the system), there is no initial

headloss along the pipeline no matter the pipeline is frictional or frictionless, so we also

have HR0 = H10 and FDR = 1, and equation (4-9) still reduces to equation (4-7).

Through this FDR compromise, we are actually using the steady state FDR to

estimate the frictional decay under transient states. Fortunately, the case studies using

this model have sufficiently shown the potential of PID control valve with remote

pressure sensor in limiting the pressure oscillations and resonance. As better pressure

control would arise from a better estimate for the dynamic set point, H1*(t); but this

value is challenging to find since there is no universally appropriate estimate for transient

pressure decay rate. These difficulties arise from the complexity of the friction term and

wave interference in the momentum equation of transient flow, which is not only

dependent of the roughness or other properties of pipe system and the hydraulic

conditions, but also relevant to the direction and frequency of the propagating waves.

Finally, it is understood that the pressure at the remote location for the L/a

second earlier, HR (t-L/a), could be always retrieved at any instant time t in both physical

and numerical systems. In the physical system, the value of HR (t-L/a) was measured and

recorded at the remote sensor, and then sent instantly to the controller by wireless or

cable transmission at any time t. Thus, at least two pressure sensors are required in the

system, one is to measure the control variable H1 (t) at the valve and another is to

measure remotely to obtain the dynamic set point H1*(t). While in the numerical

counterpart, the pressure at the remote sensor node for L/a second time ahead, HR(t-L/a),

was already simulated and stored, which could obviously be retrieved at any instant time.

78
4.5 Simulations and Case Studies

4.5.1 Case Studies with Non-Zero Initial Flow

The same system, as shown in Figure 4.4, is studied to illustrate the application of

current mathematical model. In this system, a long horizontal pipeline (with L = 5000 m,

diameter D = 1.0 m, and wave speed a = 1000 m/s) links two constant head reservoirs.

At the entrance of downstream reservoir, there is a valve with constant ES = 7 m2.5/s and

initial opening τ 0 = 0.1. The upstream reservoir has 30 m water head initially (HR0 = 30

m) and downstream reservoir has a constant head of 15 m. The water level at upstream

reservoir starts to fluctuate sinusoidally, which induces transient events in the system.

The amplitude of oscillatory wave is 2 m; the cyclic period is 10 s, which equals 2L/a, so

that resonance would be expected to take place. The instant pressure at the upstream

reservoir can be described as:

H R (t ) = H R 0 + 2 sin( 2π t / 10) (4-10)

In the first case study, it is assumed the pipeline is frictionless, i.e., the Darcy-

Weisbach friction factor f = 0. At initial steady state, the hydraulic head at the valve inlet

equals the reservoir head, that is, H10 = HR0 = 30 m, and the flow rate in the system and

passing through the valve is Q0 = 2.71 m3/s.

If the downstream valve remains at its initial opening τ 0 = 0.1 (i.e., fixed

orifice), the oscillations of upstream pressure (i.e., incidental pressure wave, or forced

vibration) will cause hydraulic resonance and pressure amplification in the middle part of

the pipeline. Figure 4.5 shows the development of the steady-oscillatory flow in the

79
pipeline; the solid line represents the pressure oscillations at the middle-point of the

pipeline and the dashed line represents the pressure oscillations at the valve inlet. There

is a L/2a time difference (i.e., 1/8 phase difference) between these two pressure waves,

and the amplitudes of both waves are initially small and grow gradually until they finally

(within 300 s) stabilize at a resonance condition. This mode shape of the pressure waves

can be understood and explained as follows: in this system, the upstream incidental wave

will reach the valve inlet in 5 s (L/a = 5 s), and it will take 7.5 s for the first wave peak to

arrive at the valve inlet (the oscillatory period of upstream incidental wave is 10 s).

During the time period 5 to 7.5 s, the pressure at the valve inlet, indeed at any internal

pipe node, is the superposition of the incident wave and the reflected wave from the

fixed valve at downstream reservoir. Since the reflection at downstream boundary in this

case is negative, the first wave peak is reduced when it arrives at downstream valve.

However, the pressure waves with different amplitudes would continuously proceed and

be reflected. The result of superposition increases until the maximum amplitude reached

and the steady-oscillatory flow condition developed in the system. This simulation result

with fixed valve opening is also represented by the dashed lines in Figure 4.6. The value

of the steady amplitude at each position of pipeline depends on the system frequency and

resistance characteristics. As matter of fact, this phenomenon resulting from the forced

vibration of upstream pressure (with fixed downstream valve opening) is equivalent to

the responses caused by periodic valve motion at downstream end (while keeping the

upstream reservoir head constant), as summarized in Figure 8.2 on page 204 of Chaudhry

(1979).

80
To eliminate or limit the reflection and superposition of pressure waves in the

pipeline, a sensor is installed at the upstream node, and a PID controller positioned at the

downstream valve. The constants of PID controller are taken as KC = 250, Ti = 0.002 s

and Td = 0.5 s. The total simulation time is run for 800 s, and the time step is 1/100 s that

is sufficiently small to look into the detailed valve responses.

The developed remote control model for frictionless pipeline is used in this

case study to simulate the system responses to the oscillations of pressure head at

upstream reservoir. Figure 4.6 compares the system responses for the responsive PID

control valve (black/solid lines) and the conventional valve with fixed orifice (red/dashed

lines). The solid/black lines in Figure 4.6(a) show that the maximum and minimum

pressures along the pipeline remain the same as those introduced at upstream reservoir,

demonstrating no wave reflection or resonance ever occurred in the system by using

remote sensing and PID control valve. By contrast, if the valve opening is fixed at the

initial size ( τ = 0.1 ), the pressure waves are reflected and superposed in the pipeline,

resulting in the increased envelope of maximum and minimum pressures along the

pipeline, see the red/dashed curves in Figure 4.6(a). In Figure 4.6(b), the black/solid

curves show the opening of PID control valve is adjusted periodically in response to the

oscillating incident pressure waves, as expected, which exactly eliminates the wave

reflections at the valve (by exciting a contrary wave) and thus remains the same mode

shape of pressure oscillation as of the upstream incident wave; while the red/dashed lines

show that the amplitude of pressure waves at the inlet of fixed orifice is reduced at the

beginning and grows until stabilized at certain value because of resistance of the valve.

81
In the second case study, the friction factor f = 0.012 is given for the pipeline.

At the initial steady state, the hydraulic grade line declines along the pipe, the pressure at

the valve inlet is H10 = 19.37 m, and the flow in the system and passing through the valve

Q0 = 1.46 m3/s. The same parameters of PID controller, total simulation time and time

step as in the first case study are used here.

To simulate the system responses to the oscillations of pressure head at

upstream reservoir, the frictional decay rate FDR defined in equation (4-8) and the

dynamic set point estimated by equation (4-9) are applied. The simulated PID control

results are compared with the corresponding conditions for fixed-opening-valve case. As

shown in Figure 4.7, the reflection is not completely eliminated because of the roughly

estimated FDR in transient state. However, the simulation results converge to a steady

oscillation flow within 400 s, and the comparison in Figure 4.7(a) shows that the

maximum and minimum pressure envelope using the PID control (solid/black curves) is

smaller than the case of fixed-opening-valve (dashed/red curves), which demonstrates

the reflections and resonance of pressure waves are constrained by using the remote

sensing and PID control valve. The smaller the initial flow in the system, the more

reduction in the magnitude of pressure amplitude envelopes, and this point can be

demonstrated by the following case study with zero initial flow. Figure 4.7(b) shows

that the opening of the PID control valve is adjusted periodically in response to the

proceeding waves, and the steady oscillation remains at the valve inlet with the same

decayed amplitude as initial steady state. Yet, for the case of fixed-opening-valve, the

amplitude of pressure waves at the valve inlet varies with time, reduced at the beginning

and increasing gradually until a steady-oscillatory flow developed over about 300 s.

82
60

50
Pressure head (m)
40

30

20

10

0
0 50 100 150 200 250 300
Time (s)

H(L) H(L/2)

Figure 4.5 Development of steady-oscillatory flow in frictionless pipeline with fixed-

valve-opening at downstream, due to the upstream pressure oscillations

83
60

Max./Min. pressure head (m)


50
Max. H
40

30 S. S.
20
Min. H
10

0
0 1000 2000 3000 4000 5000
Distance along pipeline (m)

PID V alve Fixed Valve Steady State

(a) Maximum and minimum pressure head envelopes

34 0.4
H1
32
0.3

30
H1(m)

Tau

0.2
28

0.1
26
Tau

24 0
0 20 40 60 80 100 120 140 160 180 200
Time (s)

PID V alve Fixed V alve

(b) Valve opening and inlet pressure wave varying with time

Figure 4.6 Responsive PID control valve vs. fixed-opening-valve in frictionless


pipeline (f=0)

84
60

Max./Min. pressure head (m)


50

40 Max. H

30
S. S.
20

10 Min. H
0
0 1000 2000 3000 4000 5000

Distance along pipeline (m)

PID Valve Fixed Valve Steady State

(a) Maximum and minimum pressure head envelopes

22 0.4
H1

20
0.3

18
H1(m)

0.2 Tau
16

0.1
14
Tau

12 0
0 20 40 60 80 100 120 140 160 180 200
Time (s)

PID Valve Fixed Valve

(b) Valve opening and inlet pressure wave varying with time

Figure 4.7 Responsive PID control valve vs. fixed-opening-valve

in frictional pipeline (f=0.012)

85
4.5.2 Case Studies without Initial Flow (Static Initial State)

In the system sketched in Figure 4.4, if the constant heads are 30 m at both upstream and

downstream reservoirs, there is no flow in the system and the initial valve opening is of

no consequence. For both cases with frictionless and frictional pipeline, the wave

reflection and resonance can be completely eliminated and the “non-reflective” boundary

achieved. In this system, the oscillatory head causes the flow, and the flow direction

shifts as the head oscillates around the original constant level. At any specific point of

the pipeline (e.g., at the valve), the magnitude of oscillatory flow is small and the

average value with time is zero. The f value by itself seems hardly influence the wave

reflection.

Figures 4.8 shows the simulation results for the case study with friction factor

f=0.012. Without the flow (and thus without resistance from flow), the resonance would

be stabilized in around 600 s for the fixed valve opening case and the range of the

max./min. pressures are much larger than that with non-zero flow case. While with

responsive PID control valve, there is no wave reflection and resonance at all in the

system, and the oscillations remain steady at any point and at any time, as the same as

the upstream incident pressure oscillations.

86
130
110 Max. H

Max./Min. Pressure Head (m)


90
70
50
30
10
-10
-30
-50 Min. H
-70
0 1000 2000 3000 4000 5000
Dis tance along Pipeline (m )

Fixed Valve (Tau=0.1) PID Control valve

(a) Maximum and minimum pressure head envelopes

36 0.8

32 0.7

28 0.6
H1
24 0.5

Q (m /s)
H1 (m)

20 0.4
3

16 0.3

12 0.2

8 0.1
Q
4 0

0 -0.1
0 10 20 30 40 50 60 70 80 90 100
Time (s)

(b) Inlet pressure wave and flow at the PID valve varying with time

Figure 4.8 Responsive PID control valve vs. fixed-opening-valve for static initial

states with frictional pipeline (f=0.012)

87
4.6 Frequency Analysis and “Non-Reflective” Boundary Verification

Based on the time domain transient analysis using method of characteristics (MOC), the

numerical model for “non-reflective” boundary design has been developed in previous

sections. However, for a periodic oscillation originating at a remote location, transient

analysis in the frequency domain is a more practical and efficient way to reveal the

oscillatory conditions in the fluid system.

In the frequency domain, there are two types of analyses: frequency response

analysis and free vibration analysis. They apply the basic unsteady flow equations, using

complex variables and a mathematical shorthand common in variation problems. The

former is based on fully developed steady-oscillatory forced vibrations (i.e., a persistent

excitation). The objective is to obtain the system responses to the explicit excitation

imposed on the system. The free vibration of a fluid system is caused by some temporary

excitation. The initial fluid motion is of little interest in the Free Vibration Analysis, it

aims only at the residual time-dependent oscillation. Thus, it doesn’t need to know the

excitation, but the natural frequencies of the entire system would be defined and the

vibration mode shape of the system be identified. The analysis is also useful in

evaluating system stability, in judging system performance, and identifying critical

resonance conditions.

In this section, oscillation is introduced through various harmonics that might

occur in the pipe system. Then, the system responses to the forced vibrations (upstream

pressure oscillations) with different frequencies and the applicability of developed “non-

reflective” model are checked. After that, the steady “non-reflective” boundary

88
conditions for the frictionless pipeline system, obtained by the traditional MOC, would

be verified using the method of hydraulic impedance in the frequency domain.

4.6.1. System Responses to Pressure Oscillations with Various Frequencies

Unexpected resonance could be destructive in practical hydraulic systems. The

consequences of resonance in fluid systems range from objectionable operating

conditions, such as instability, noise, and vibration, to fatal damage of system elements

overstressed during severe pressure oscillations. Thus, the phenomenon of hydraulic

resonance should be predicted and prevented.

In this case study, the incident pressure oscillation at upstream reservoir is one

type of forced excitation. The fundamental period of pipeline system T0 = 4L/a = 20 s

(i.e., natural frequency is 1/20 Hz) and the given period of forced excitation T = 10 s (the

forcing frequency is 1/10 Hz), and the system responses to this forced excitation are

shown in Figure 4.6(a) and Figure 4.7(a) for frictionless and frictional pipeline,

respectively. For a fixed orifice at downstream reservoir, the system responses

demonstrate the characters of second harmonics, and the maximum amplitude of

pressure oscillation, occurred at the middle point of the pipeline, is about 12 and 9 times

as large as the incident pressure oscillation, respectively, for frictionless and frictional

cases. Now, what if we change the frequency of the forced excitation? Could we adjust

downstream valve opening to eliminate the potential resonance in the system using the

developed “non-reflective” boundary design?

89
(a) No resonance (T0/T = 1)

(b) 2nd order of harmonics (T0/T = 2)

(c) No resonance (T0/T = 3)


Figure 4.9 System responses to forced pressure oscillation with various frequencies
in a frictionless pipeline (f=0) (Con’d)

90
(d) 4th order of harmonics (T0/T = 4)

(e) No resonance (T0/T = 5)

(f) 6th order of harmonics (T0/T = 6)


Figure 4.9 System responses to forced pressure oscillation with various frequencies
in a frictionless pipeline (f=0) (Con’d)

91
(g) 8th order of harmonics (T0/T = 8)

(h) 10th order of harmonics (T0/T =10)

(i) 20th order of harmonics (T0/T =20)


Figure 4.9 System responses to forced pressure oscillation with various frequencies
in a frictionless pipeline (f=0)

92
Frictionless pipeline system. In this system, for the fixed valve opening at downstream

reservoir, even order of harmonics exist, as shown in Figure 4.9 (b), (d) and (f)-(i). The

even harmonics indicates that the reflective characteristic at downstream orifice is

similar to a reservoir (negative reflection) (Wylie, et al., 1993 and Chaudhry, 1987). This

can be verified by comparing the hydraulic impedance of the fixed valve with the

characteristic impedance of the system. Hydraulic impedance in a fluid system is defined

as the ratio of the complex head to the complex discharge at a particular point in the

system (Wylie, et al., 1993. Pg301-302).

For the frictionless pipeline system as shown in Figure 4.4 (reservoir-pipeline-

orifice at reservoir), the characteristic impedance of fluid system is calculated as follows:

a 1000 m/s
ZC = = ≈ 130 s/m2 (4 - 11)
gA 9.81 m/s × 3.1416×1 /4 m
2 2 2

While for the fixed orifice or valve, the valve equation (4-3) can be written as:

Q = τ ES H0 (4 - 12)

in which H 0 is the head drop across the valve for the mean flow Q . From the initial

steady state, we have: τ = 0.1, E S = 7 m 2.5 / s, H 0 = 15 m, Q = 2.711 m 3 / s . So the

hydraulic impedance of this fixed valve:

H V 2H 0
ZV = = ≈ 11 s/m 2 (4 - 13)
QV Q

93
here HV and QV are the complex head and flow at the oscillatory valve, respectively.

Thus, we have Z V < Z C , this is the condition that even harmonics occurs in the system.

On the contrary, if Z V > Z C , the odd harmonics could occur, which indicates the orifice

would provide a response similar to a dead-end. If we adjust the valve opening to make

Z V = Z C , then the orifice becomes “non-reflective”.

When the cyclic period of the forced vibration is given as fractional a part of

the fundamental period (T0 = 4L/a =20 s), the different orders of harmonics and different

mode shapes of pressure waves would occur in the pipe system. However, the frequency

change doesn’t affect the amplitude of each harmonics if the amplitude of incidental

pressure oscillation remains the same. The order of harmonics equals the system

fundamental period T0 divided by the period of forced vibration T. If we give even

number of T0 /T (i.e., T = 10 s, 5 s, 3.33 s, 2.5 s, 2 s and 1 s, as shown in Figure 4.9),

excessive energy influx to the system during oscillatory flow leads to resonance. In an

ideal lossless system (for this frictionless pipeline system the only energy dissipation

occurs at the downstream valve), there is generally no energy transmission in steady-

oscillatory flow, although alternating energy conversion between kinetic energy and

pressure energy may occur. With terminal wave reflection, steady-oscillatory motion

shows a combination of forward and reflected waves that results in a standing wave.

Within the standing-wave pattern, energy is converted from pressure energy to kinetic

energy, then back to pressure energy, and so on. If we give odd number of T0 /T (as

shown in Figure 4.9 (a), (c) and (e) for T = 20 s, 6.67 s and 4 s), the mode shape of the

pressure waves is quite different from the even harmonics, the energy dissipated

gradually and resonance does not occur in the system.

94
It is not surprising to find the resonance with different orders of harmonics

and amplification of pressure head can be completely eliminated in frictionless pipeline

by designing a “non-reflective” boundary. The simulated results are all the same as the

black/dotted line in Figure 4.6 (a), no matter how the period of the forced vibration T

changes. However, it is noticed that for the higher frequency forced oscillations (smaller

T), the PID integral and derivative parameters (Ti and Td) require smaller values to obtain

the precise adjustment of “non-reflective” valve opening.

Frictional pipeline system with non-zero initial flow. For the frictional pipeline case,

we have a similar finding regarding even harmonics when we change the period of the

forced vibration. The application of remote sensor and PID control valve cannot

completely eliminate the reflections and resonance if the initial flow is not zero, but the

amplitude of the pressure waves is significantly reduced for each harmonics, as shown

Figure 4.10. In this Figure, the different responses to the forced oscillations at the

upstream end are compared for the system with a fixed-opening-valve and “non-

reflective” boundary condition at the downstream end of pipeline.

95
Figure 4.10 System responses to forced pressure oscillation with various
frequencies in a frictional pipeline (f=0.012) (Fixed valve vs. responsive PID control
valve)

96
4.6.2. “Non-Reflective” Boundary Verification using Hydraulic Impedance Method

From the viewpoint of frequency domain, the automatically adjustable PID valve creates

an artificial excitation and the consequence of this designed valve-oscillation would

exactly cancel out the effect of incidental pressure oscillation at the upstream reservoir.

For the frictionless pipeline system, we have verified the condition of a “non-reflective”

boundary, that is, Z V = Z C , for the developed steady-oscillatory flow (i.e., after 300 s of

PID valve adjustment when the amplitude of the pressure waves in the pipeline

stabilized). This steady valve-oscillating condition has been obtained from the numerical

simulation in the time domain, which uses the “non-reflective” boundary model

developed in the previous section.

For the frictionless pipeline system, the characteristic impedance has been

calculated as equation (4-11). For an oscillating valve, the hydraulic impedance at the

upstream side of the valve can be calculated by

H V 2 H 0 2 H 0 TV
ZV = = − (4 - 14)
QV Q τ QV

We already know τ = 0.1, H 0 = 15 m, Q = 2.711 m3 / s , and here HV, QV and T V are the

complex head, flow and opening at the oscillatory valve, respectively.

From numerical simulation in the time domain, we found the maximum valve

opening (τ max = 0.107 , amplitude = 0.007) corresponding to the minimum valve flow

(Qmin = 2.695 m3/s, amplitude = 0.016 m3/s) and minimum valve inlet head (H1min = 28.0

m, amplitude = 2.0 m). On the contrary, the minimum valve opening ( τ min = 0.094 ,

97
amplitude = 0.006) corresponds to the maximum flow (Qmax = 2.726 m3/s, amplitude =

0.015 m3/s) and maximum valve inlet head (H1max = 32.0 m, amplitude = 2.0 m). In other

words, we have complex hydraulic values:

⎡ 2π (t + L / a) ⎤ ⎡ 2π t ⎤
H V = 2.0 sin ⎢ ⎥ = 2.0 sin ⎢ +π⎥ (4-15)
⎣ 10 ⎦ ⎣ 10 ⎦

⎡ 2π (t + L / a) ⎤ ⎡ 2π t ⎤
QV = 0.0155 sin ⎢ ⎥ = 0.0155 sin ⎢ +π⎥ (4-16)
⎣ 10 ⎦ ⎣ 10 ⎦

⎡ 2π t ⎤
TV = 0.0065 sin ⎢ ⎥ (4-17)
⎣ 10 ⎦

It is noticed that there is a phase difference of (2πL / a) / 10 between the oscillation of

valve opening (TV) and the oscillations of valve flow (QV) and inlet head (HV). So, we

HV 2 H 0 2 H 0 TV
can also calculate Z V = ≈ 129 s/m2, or Z V = − ≈ 137 s/m2 (from
QV Q τ QV

equations (4-16) and (4-17), we know TV and QV have opposite sign; the numerical error

is acceptable due to only 3 digitals of the simulation results recorded). Therefore, the

“non-reflective” boundary condition ZV ≈ Z C =130 s/m2 has been verified.

In the case without initial flow but the pipeline with friction, as shown in

Figure 4.8, we obtained the same pressure and flow oscillations at the valve as described

HV
in equations (4-15) and (4-16), and thus Z V = ≈ ZC =130 s/m2 can be verified.
QV

Moreover, by changing the amplitude of the incident pressure waves at the upstream

reservoir, the amplitude of induced flow oscillations would change proportionally, so the

98
HV
value of Z V = remains near 130 and “non-reflective” boundary condition would be
QV

always achieved for static initial state cases.

To further verify this law of “non-reflective” boundary condition, the wave

speed of pipeline is reduced to 500 m/s, so the system characteristic impedance also

a
reduces to ZC = ≈ 65 s/m2 . The “non-reflective” boundary condition, Z V = Z C , could
gA

be verified as well by the corresponding numerical simulation results in the time domain.

4.7 Tuning PID Controller

The final tuning of the parameters (KC, Ti and Td) for a PID controller would be

important during the commissioning stage of the system. Similar to the trial and error

method used for a physical system, the numerical model could provide a tool for

preliminary selection of these parameters. To better understand how the variation of

each parameter affects on the system control results, a sort of sensitivity analysis for

three controller parameters is performed here, using the same system shown in Figure 4.4

with frictionless pipeline as aforementioned, and the comparative results are summarized

in Figure 4.11.

Proper selection of controller parameters means finding a compromise

between the requirement for fast control and the need of stable control. More

specifically, with increases in the proportional gain (KC), the speed of control increases

but the stability of control reduces. Figure 4.11 (a) shows that given the same simulation

time period (200 s), the reduction of KC (slower control) enlarges the maximum and

99
minimum pressure envelope. Clearly, KC must be greater than a certain value to

effectively control the system. In this case study, a KC of about 50, or greater, is needed.

The tendencies for the variation of integral time (Ti) are opposite to KC. With

Ti increases, the speed of control reduces while the stability of control increases. Figure

4.11 (b) shows that the maximum and minimum pressure envelope expands with the

increase of integral time Ti (slower control), and Ti ≤ 0.001 s is required in this system.

The derivative part produces both faster and more stable control when Td

increases. However, this is only true up to a certain limit and if the signal is sufficiently

free of noise (calculation error is a kind of noise in numerical system). If Td rises above

this limit it will result in reduced stability of control. As we know that the function of the

derivative part is to estimate the change in the control a time Td ahead. This estimate will

naturally be poor for large values of Td. Another consideration is the noise and other

disturbances. The noise is amplified to a greater extent when Td increases, and thus it is

often the noise that sets the upper limit for the magnitude of Td. The above theoretical

analysis could be verified by the numerical simulations in Figure 4.11 (c), which shows

that the control stability is better when Td = 0.5 s (dashed lines) than that when Td =

0.005 s (dot-dash lines). However, when Td = 5 s the stability of control is poor.

In addition, the simulations also show that for high frequency oscillatory flow

(e.g., 1 s of the period of incidental pressure wave), the smaller values for both the

integral time constant Ti and derivative time constant Td are required.

The stability and speed of control process are associated with the parameters

of controller. The mathematical model and numerical simulation are useful for selection

100
and tuning of the controller parameters, which could save time and cost in

commissioning of the physical system.

60
Max./Min. pressure head (m)

50
Max. H Kc=250
40
Kc=50
S. S. H
30
Kc=5
20 Min. H S.S.

10

0
0 1000 2000 3000 4000 5000
Distance along pipeline (m)

(a) Max./Min. pressure envelopes expanding with reduction of proportional gain KC


(Ti = 0.001 s, Td = 0.5 s)

Figure 4.11 Effect of controller parameters on system pressure control (Con’d)

101
40
38

Max./Min. pressure head (m)


36
Max. H Ti=0.001
34
32 Ti=0.01
30 Ti=0.02
S. S. H
28
S.S.
26
24 Min. H
22
20
0 1000 2000 3000 4000 5000
Distance along pipeline (m)

(b) Max./Min. pressure envelopes expanding with increase of integral time Ti


(KC = 250, Td = 0.5 s)

40
38
Max. H
Max./Min. pressure head (m)

36
34
Td=0.5
32
30
Td=0.005
S. S. H
28 Td=5
26 S.S.
24
22 Min. H

20
0 1000 2000 3000 4000 5000
Distance along pipeline (m)

(c) Max./Min. pressure envelopes varying with different derivative time Td


(KC = 250, Ti = 0.02 s)

Figure 4.11 Effect of controller parameters on system pressure control

102
4.8 Summary and Conclusions

Dead-end branches are a common component in pipe networks; of such locations, dead-

end reflection may cause unexpected high pressures when system experiences transients,

which should and can be avoided by using a suitable dissipative valve.

The creation of “non-reflective” (or “semi-reflective”) boundaries through

remote sensing and PID control valve is a new concept to more accurately limit dead-end

reflection and resonance in pipelines. This idea is explored here by the means of

numerical simulation, and a considerable potential for transient protection has been

demonstrated in the case studies. Using the model developed in this paper, the wave

reflection and resonance could be eliminated for frictionless pipelines or the pipelines

with static initial states; while for the pipeline with some friction, the pressure waves’

reflection at the valve and superposition within the pipeline can be effectively limited.

Moreover, the theory of hydraulic impedance in the frequency domain, regarding the

condition of even order of harmonics and “non-reflective” boundary conditions, has been

verified by the numerical simulations using the transient analysis model developed in the

time domain.

However, complications and challenges may arise when the model of “non-

reflective” (or “semi-reflective”) boundary is applied in real systems. First, the model is

developed based on a single pipeline system with the remote sensor at upstream end and

PID control valve at downstream end. It is feasible to apply the model for a branched

pipeline in a complex system with remote sensor at the junction and a terminal valve at

the branch-end, but further application to a pipe loop or an arbitrary pipeline with other

103
components between the two ends would be significantly challenging, since the primary

pressure wave would be reflected, refracted, or attenuated by those components and thus

the theoretical estimate of pressure set point become almost impossible. Besides, in real

systems, the uncertainties (or frequency-relevance) in the magnitudes of waves speed,

pipe friction factor and pressure decay rate at transient state may cause significant over-

or under-estimate of pressure set point and thus require for careful system calibrations in

advance. The upstream pressure disturbance may not be a steady oscillation as in the

case studies, and then the reflections at the upstream reservoir would further complicate

the “non-reflective” valve opening control. Thus, at present the design of “non-

reflective” boundary, though promising, calls for further research as well as cooperation

with the device manufacturers.

104
Chapter 5 Fundamentals of Transients in Hydro
Turbine Systems

This chapter supplies the required background for the second type of AHD studied in this

thesis, a governed hydro turbine. A hydro turbine is the most complicated hydraulic

component in terms of transient analysis and waterhammer control. Motivation and

interest in hydro system research arises partly from the author’s previous background, as

well as a global reawakening of interest in the hydro industry (especially small scale

facilities). Unavoidably, waterhammer remains a serious issue in many hydro systems

where accidents and disruptions are caused by imperfect operation of turbine units

(Pejovic et al, 2007). Although there are several numerical models developed for hydro

system transient analysis, none of those published is directly applicable to a water network.

To better understand the transient flow associated with turbine operation, the

fundamental knowledge of turbine and hydro system, including turbine classification,

turbine characteristics, governor and control equipment, and the common arrangements of

hydro systems, are reviewed in this chapter.

105
5.1 Hydro Turbine Classification

Although turbine classifications are similar throughout the world, the specific terminology

of different turbine types and models may vary from one region to another.

5.1.1 General Classification of Hydro Turbines

Hydraulic turbines are classified, with respect to their hydraulic action or energy

conversion, into two major categories: reaction turbines and impulse turbines, as shown in

Table 5.1. The reaction turbines include Francis and Propeller types; in these types, flow

takes place under pressure in a closed chamber and only a part of the available water

energy is converted into kinetic energy at the entrance to the runner, with a substantial part

remaining capable of doing work on the runner through pressure. Some propeller-type

turbines have fixed runner blades, i.e., the angle of runner blade ϕ is fixed at a near optimal

position for the design operating condition. Another family of propeller turbines, the

Kaplan turbines, has adjustable blades; the blade angle ϕ can be adjusted within a certain

range to achieve higher efficiencies over a wide spectrum of flow conditions. Both the

runner blade angle and wicket gate position are regulated through governor control system

by a cam mechanism (a designed relationship between wicket gate servo stroke and runner

blade servo stroke). Kaplan turbines are typically used on run-of-the-river hydroelectric

plants with relatively high flow rates but low heads (Tong, et al., 1997 and Liu, et al.,

1997).

106
Table 5.1 Classification of Hydro Turbines

Impulse Turbines Pelton


(small ns) Turgo
Turbines
Francis
Reaction Turbines Propeller (with fixed runner blades)
Kaplan (with adjustable runner blades)

In an impulse turbine, a jet of water issuing from a nozzle impinges on the

vanes/buckets of the turbine wheel/runner (usually only one or two vanes/buckets at a

time), which is exposed to atmospheric pressure. All of the available potential energy of

the water is first converted into kinetic energy with the help of a nozzle. The impulse

turbine is a logical choice for high-head installations and it is also suitable for many lower

head sites if the discharge is relatively small. Because of this, it is the turbine of choice for

many small-scale hydropower plants. The primary advantages of the impulse turbine are

its simplicity and ease of maintenance. One of its disadvantages is that the impeller must be

set above the highest level of the downstream pool. This requirement implies that

run-of-the-river plants would tend to waste available head in dry seasons when the river

discharge is low and the downstream pool level depressed.

According to the main flow direction within the runner passage, hydro turbines

can be classified as types of axial-flow, cross-flow, radial-flow (Thompson) etc.; they also

be classified as vertical, horizontal (Bulb, Pit) and inclined types according to the

107
arrangement of the main shaft.

A pump-turbine is a special type of turbo machine used in pump-storage hydro

plants. All pump-turbines are reversible; that is, when water enters the spiral case at the

periphery and flows inward to the machine, it acts as a turbine; when water enters at the

center (or eye) and flows outward, it acts as a pump. The direction of rotation is, of course,

opposite in the two cases. Depending on the design pressure heads, pump-turbines may be

of the Francis, mixed-flow, or axial-flow type. A pump-turbine is connected to a motor

generator which acts as either a motor or generator depending on the direction of rotation

(Franzini, 1997).

Some pump-turbines are speed-adjustable either in pumping or generating mode

because, with variable speed, it is possible to operate the unit at or near the optimum speed

for the particular head and output and thus obtain a higher efficiency than with a

single-speed unit.

5.1.2 Hydro Turbine Classification in the Context of Numerical Modeling

To describe a turbine in computer code, it is necessary to precisely define which turbine

parameters are varying and which are fixed. In the numerical model developed in this

thesis, turbines are classified as:

1) ‘Impulse’ (equivalent in most ways to an external discharge valve);

108
2) ‘Fixed WG’ (with fixed wicket gates and fixed runner blades);

3) ‘Francis’ or ‘WG’ (wicket gates adjustable but runner blades fixed);

4) ‘Kaplan’ (both wicket gates and runner blades are adjustable).

Impulse Turbines. The flow of an impulse turbine is a function of head and nozzle

opening only. In other words, the variation of turbine speed has no direct influence on the

turbine flow. Some impulse turbines (Pelton) may have dual flow control systems, i.e., a

deflecting board and a needle valve. To avoid excessive acceleration of the runner during a

load rejection, it is necessary to block its water supply; the deflector servomotor shifts the

deflector to the required deflecting position for a short period of time (2-3 seconds or more)

while the discharge through the nozzle has not been changed. Simultaneously, the nozzle

servomotor moves the needle to the required position. The full travel of the needle may

take 30-40 seconds or longer. During movement of the needle, the deflector is withdrawn

gradually from the jet and at the end of its motion has no effect on the flow. In other words,

under load-rejection, the deflecting board does not change the flow rate but makes the jet

alter its direction to temporarily relieve the force impacting on the Pelton wheel. Therefore,

there is no need to incorporate the deflecting board into a waterhammer analysis. In the

current version of TransAM, the impulse turbine is simplified and treated as a valve

discharging to the atmosphere. That is, the turbine governor is not included in the model

and the turbine characteristics are not required; however, the nozzle characteristics have to

109
be provided for numerical simulation and the setting of the nozzle or needle valve

operation determined by the servomotor control system.

‘Fixed WG’. In this thesis, the variable Ndy represents the number of wicket gate

settings; and Nblade is the number of runner blade angle settings. For turbines with fixed

wicket gates and fixed blades, Ndy=1 and Nblade=1. An inversely operating pump would

fit this description and there is only one characteristic curve for this type of turbine. Also

for this type of turbine, an upstream control valve is usually necessary to govern inflow

into the turbine; the valve can be considered as either a turbine-attached-valve (see in a

later section) or a separate component in the system. In both cases, no governor is involved

in the turbine simulation.

Francis or ‘WG’. This type of turbines includes Francis turbines and propeller

turbines with fixed runner blades. The position of wicket gates can be adjusted but the

runner blades are fixed (i.e., Nblade=1). There are Ndy sets of characteristic curves

corresponding to different wicket gate openings (Ndy>1). The range and intervals of

wicket gate opening are dependent on the available model hill diagram. During turbine

operation, wicket gate setting is typically controlled by a speed-control-governor and

adjusted by a servomotor system. In this case, a governor equation has to be included in the

turbine simulation.

110
Kaplan Turbine. For a Kaplan turbine, both the wicket gate opening and runner

blade angle are adjustable (Ndy>1 and Nblade>1). A governor controls both the runner

blade angle and wicket-gate opening through a designed cam mechanism. In the current

simulation (adapted to TransAM) version, the Kaplan model has been considered but not

fully implemented.

5.2 Hydro Turbine Parameters

The terminologies commonly used in hydro engineering are reviewed in this section. They

include both dimensional and dimensionless turbine parameters.

1) Turbine head (H): the hydraulic head differential between the high- and low- pressure

reference sections of the turbine unit (in feet, meters, etc.). Hydraulic head is defined

as the energy available for per unit weight of water at a section, including the elevation,

water pressure and kinetic energies.

Net head is the hydraulic head differential between points just in front of, and behind,

the turbine unit (say, from the inlet of the spiral case to the outlet of draft tube).

The maximum and minimum turbine hydraulic head (Hmax and Hmin) establishes the

allowable head range for turbine operating conditions.

The average hydraulic head (Have) is an average of available head over the years. The

111
Design/Rated head (HR) is determined based on the range of available net head, under

which the rated turbine speed and runner size are selected such that the best overall

efficiency would be achieved over the operating range of turbine flow.

2) Turbine Discharge (Q): the flow rate through the turbine (m3/s, ft3/s, etc.).

Design/Rated discharge (QR) is the turbine flow required to produce the rated power at

the design head.

3) Turbine Speed (n): the rotational speed of turbine unit, typically in rpm.

Rated speed or normal speed (nR) is the turbine rotation speed under normal operating

conditions; it is typically the speed of which the unit enjoys its best efficiency.

Synchronous speed (nE) is the speed of the generator when regulated by the frequency

of the power grid system to which the unit is connected, that is, nE = 60 f p . Here, f

is the frequency of the a.c. supply system in Hz (50-60 Hz), the permitted variation of

frequency +/- 0.2 Hz for a big system, and +/- 0.5 Hz for a small system; p is the

number of pairs of generator poles. For a unit having a common shaft connecting the

turbine and generator directly, the synchronous speed is the rated speed of the turbine

and unit.

Runaway speed (nrun) is the highest speed associated with a particular position of

needle valve(s), wicket gates and/or runner blades, after all transient waves have been

112
dissipated, when the generator is disconnected from its load and network under a

specific turbine head. At steady state, for a particular head and opening position, a

turbine’s runaway speed can be tested and measured on the model. However, the

runaway speed during transient states will often surpass its steady counterpart because

of waterhammer effects. Runaway speed is not the same as the speed rise during the

closure of wicket gates (particularly for fast closures) when load is rejected.

4) Power (P): the rated energy produced from turbine or generator per second (kW, Hp).

Turbine output is the waterpower transferred to the generator or motor through the

main shaft. The rated turbine output depends on the rated flow, rated head and the

corresponding efficiency η T ; that is,

PT = 9.81ηT Q R H R (5 - 1)

The rated output of a turbine-generator unit is obtained using the efficiency of

generator η G multiplying with the rated turbine output to obtain

PG = ηG PT = 9.81ηGηT Q R H R (5 - 2)

5) Efficiency (η): the ratio of output to input power.

PT
Turbine efficiency: ηT = (5 - 3)
γ QH

Here, γis the specific weight of water. Turbine output can be expressed as PT = Tω ; T

113
is the shaft torque exerted by the water; ω = πn 30 , the angular speed of the turbine.

The efficiency of a turbine-generator unit can be expressed as:

PG
η = η Tη G = (5 - 4)
γ QH

6) Specific Speed: Specific speed is defined differently for a turbine and a pump:

n P n Q
For turbines ns = H 5 / 4 or ng =
H 3/ 4
(5 - 5a)

n Q
For pumps ng = (5 - 5b)
H 3/ 4

Here, P is the turbine power at maximum efficiency; Q is the pump flow at maximum

efficiency. The specific speed ns or ng represents the speed of a series of geometrically

homologous turbines (or pumps) at H =1 m and P=1 kW (or flow Q=1 m3/s), which is

unrelated to the machine size. Specific speed is an important turbine parameter, largely

reflecting a turbine’s performance and typically indicative of turbine type or model.

7) Unit Parameters: Unit parameters are defined for standard-sized runner (i.e., throat

runner diameter D1=1 m) under unit net head (i.e., H =1 m), and used to describe

turbine model characteristics in hill diagrams. They are sometimes also referred as

specific parameters.

114
Q
Unit discharge: Q11 or Q1′ = (5 - 6)
D 1
2
H

nD1
Unit speed: n11 or n 1 =
'
(5 - 7)
H

P
Unit Power: P11 or P1 =
'
(5 - 8)
D H 3/ 2
2
1

T
Unit Torque: T11 or T1 =
'
(5 - 9)
D13 H

8) Suter Parameters: Based on the following four dimensionless turbine parameters:

H Q T n
h= ,v = ,β = ,α = (5 - 10)
HR QR TR nR

the independent variable x and two dependent variables WH and WB regarding turbine
B

head and torque (i.e., the Suter parameters) are expressed as follows:

α
x = tan −1 (5 - 11)
v

h
WH ( x) = (5 - 12)
α + v2
2

β
WB ( x) = (5 - 13)
α + v2
2

The above definition of x is consistent with pump characteristics used in the simulation

115
model TransAM. The relationships between WH ~x and WB ~x are the turbine’s head

and torque characteristics in Suter parameters.

5.3 Hydro Turbine Characteristics

To simulate transient flow through turbines, the relationships between the net head,

discharge and/or other turbine parameters have to be specified. The flow rate through a

turbine depends upon various parameters. For an impulse turbine, the discharge through

the turbine Q is related to the net head H and the opening of nozzle-needle-valve y only, so

the relationship of its flow and head is as simple as a valve. However, for a Francis turbine,

the discharge of the turbine Q is related to the net head H, rotation speed n and wicket gate

opening y (the variable y represents the dimensionless opening of turbine in this thesis); the

flow through a Kaplan turbine also depends upon the runner blade angle ϕ. The curves

representing the relationships between these parameters are called turbine characteristics,

with combined model characteristic curves (i.e., hill diagram in unit parameters) defining

the relationships among various turbine parameters for a series of homologous turbines.

The prototype turbine characteristics, also known as turbine performance curves, are all

converted from model hill diagram. They include curves of efficiency vs. output (i.e., η –

P), output or efficiency vs. discharge vs. (i.e., P – Q or η-Q ), output or efficiency vs. head

(i.e., P–H or η – H), and discharge, output or efficiency vs. speed (i.e., Q–n, P–n or η –n).

116
Except for large and important units, turbine designers and manufacturers usually

do not conduct load rejection tests for model turbines (Standard laboratory for model tests

cannot do transient tests); that is, the turbine characteristics are usually obtained from

steady-state cavitation free model tests, from which the prototype turbine’s performance

curves would be plotted. However, an actual turbine cannot avoid experiencing hydraulic

transients, and the effect of waterhammer does influence the head–discharge relationship

and thus, to some extent, all other characteristic curves. Particularly, in no-load conditions,

the runaway speed at steady state cannot exceed a certain value at a particular net head and

gate opening. During a transient, the prototype speed at no-load condition would exceed

the steady-state runaway speed for a short duration. To account for this, the curves are

extrapolated for speed values beyond the steady-state runaway value by assuming they

follow the same trend observed for inferior speeds (Chaudhry, 1987).

In addition, because of the assumption of equal efficiencies in the homologous

relationship between the model and prototype, the prototype efficiency should be modified

after conversion of turbine parameters from model to prototype. For medium and large

turbine units, the prototype efficiency is usually higher than that determined by model tests

at the same operating point because of scale effects and the limitations of manufacturing

technologies for small-size models. Various empirical formulas for the normal operating

zone have been proposed for efficiency correction (IEC 193, 1965).

117
5.4 Model Hill Diagrams and Conversion

There are four quantities associated with turbo-machine characteristics: the pressure head

H, the discharge Q, the rotational speed n, and the shaft torque T (or efficiency η ). Two of

these may be considered independent. For instance, if Q and n are given, H and T (or η )

can be determined from the turbine characteristics. The shaft torque T and efficiency η can

be converted from each other in the range of the normal operating zone based on the shaft

power balance equation; that is, Tω = γ QH η for a pump and Tω = γQHη for a turbine.

As aforementioned, the available turbine characteristics are usually expressed in

unit parameters and obtained from model tests. For simulation, the model relations have to

be converted for the prototype turbine. However, the prototype turbine’s relationships

between H, Q, n and T (or η ) are difficult to handle in numerical simulation because they

all may cross zero and change sign during a transient. Thus, the Suter parameters, WH and

WB, are used to describe a turbine’s characteristics in transient simulation; it is particularly


B

necessary for reversible pump-turbines and turbines operated in pump mode.

5.4.1 Francis Turbines

Figure 5.1 is an illustrative hill diagram for a typical Francis turbine, which is usually

furnished by the turbine designer or manufacturer. At each wicket gate setting (e.g., y = y1),

a series of n 1' , Q 1' values are read from the Figure; and then the following dimensionless

variables can be calculated with known rated turbine parameters:

118
H 1 ⎛ n 2 D12 ⎞ nR2 ⎛ D12 nR2 ⎞ α 2
h= = ⎜ ⎟ =⎜ ⎟ (5 - 14)
H R H R ⎜⎝ n1'2 ⎟⎠ nR2 ⎜⎝ H R ⎟⎠ n1'2

Q Q1' D12 H HR ⎛ D12 H R ⎞ '


v= = =⎜ ⎟Q1 h (5 - 15)
QR QR HR ⎜ QR ⎟
⎝ ⎠

T γQHη ⎛ γQR H R ⎞ hv vhη


β= = = ⎜⎜ ⎟⎟ η = (5 - 16)
TR TRω ⎝ TRω R ⎠ α αη R

According to the above equations (5-14) to (5-16), the dimensionless variables h, βand v2

are proportional to α2 (i.e., h ∝ α 2 , v ∝ α , β ∝ α 2 ). Therefore, no matter what the relative

speed α is, the ratios of h /(α 2 + v 2 ) and β /(α 2 + v 2 ) are independent of α. For

convenience, α =1 (i.e., n=nR) is assumed in the data conversion.

n1′ η

y2

y1

Q1'

Figure 5.1 Illustrative hill diagram of Francis turbine

119
The steps to convert a hill diagram (Figure 5.1) to Suter parameter curves (Figure

5.2) are summarized as follows:

(1) For a given opening (e.g., y = y1), read one set of n 1' , Q 1
'
from the model hill diagram;

y=0
WH = 2
h β y=0 .2 .4 .6 .8 1.0
.2 WB =
α + v2 α 2 + v2
.4
.6
.8
1.0

α α
x = tan − 1 x = tan −1
v v

Figure 5.2 Francis turbine characteristic curves in Suter parameters

(2) With α =1, calculate h, ν, β from equations (5-14) to (5-16), respectively;

(3) Calculate x, WH and WB from equations (5-11) to (5-13), respectively;

(4) Repeat the above steps (1)-(3), two curves of WH ~x and WB ~x at opening y1 can be

plotted;

(5) Repeat the above steps (1) – (4) for each of wicket gate opening to plot a set of curves

of WH ~x and WB ~x at each wicket gate opening.

120
If the model characteristic curves are in terms of unit power instead of unit flow,

equations (5-15) and (5-16) would be replaced by following (5-15a) and (5-16a),

respectively; and the same procedure as above described would be followed to plot curves

of WH ~x and WB ~x at each wicket gate opening.

Q P / γ H η D12 H 3 / 2 P1' ⎛⎜ H R D12 ⎞ P1' h



v= = = = (5 - 15a)
QR QR Q R γη H ⎜ Q R γ ⎟⎠ η

T Tω P D12 H 3 / 2 P1' ⎛ D12 H R3 / 2 ⎞ P1' h 3 / 2


β = = = = = ⎜⎜ ⎟⎟ (5 - 16a)
TR TR ω TR ω TRω ⎝ TRω R ⎠ α

In the case of prototype turbine characteristics, represented by the head H vs. Q

and turbine shaft torque T (or power P) vs. Q, the data conversion to Suter parameters can

be calculated directly from prototype variables, which is even more straightforward.

Interpolation and Extrapolation of Turbine Characteristics. Since the

characteristic curves in Suter parameters are usually built with intervals ∆x=3ºand ∆y

=0.1-0.2, it is necessary to interpolate the values of WH and WB at a specified instantaneous


B

operating point (x, y) for turbine simulation. Experience points to linear interpolation as the

most reliable method. In this study, the concept of shape function is implemented, which is

actually a two-step linear interpolation but more straightforward and conceptually clear.

121
WH
(WB)
y
y2
A3 A4

y1 P (x, y)
A1 A2

x1 x2 x

Figure 5.3 Interpolation scheme of turbine characteristics

Assuming that Suter parameters WH(x, y) and WB(x, y) are related to independent
B

variables x and y as:

h
W H ( x, y ) = = a 0 + a1 x + a 2 y + a 3 xy (5 - 17)
α + v2
2

β
W B ( x, y ) = = b0 + b1 x + b2 y + b3 xy (5 - 18)
α + v2
2

Here, coordinates (x, y) represent turbine instant operating point (P); a0, a1, a2, a3 and b0, b1,

b2, b3 are interpolation constants. To obtain the values of WH(x, y) and WB(x, y), the B

interpolation constants need to be solved first. In fact, the Gaussian Elimination Method is

employed to find these interpolation constants based on four nearby known operating

points A1, A2, A3 and A4 (see in Figure 5.3). Specifically, the interpolation procedures are

described as follows:

122
1) Determine the instantaneous operation point P (x, y). The coordinate x can be

calculated according to the instantaneous turbine speed n and discharge Q, and the

coordinate y is the gate opening;

2) Find four nearest known points around P (x, y). From data tables of Suter parameter

characteristics, the points A1(x1, y1), A2(x2, y1), A3(x1, y2) and A4(x2, y2) and their

corresponding known values of WH(1,1), WH(2,1), WH(1,2) and WH(2,2), and WB(1,1), B

WB(2,1), WB(1,2) and WB(2,2) can be determined;


B B B

3) Substitute coordinates (x1, y1), (x2, y1), (x1, y2) and (x2, y2) and values of WH(1,1),

WH(2,1), WH(1,2) and WH(2,2) into equation (5-17) to obtain the appropriate equation

(5-19) and solve it for constants a0, a1, a2 and a3; similarly, substitute the values of the

coordinates and WB at the four points into equation (5-18) to obtain the determinant
B

equation (5-20) and solve it for constants b0, b1, b2, and b3;

4) Finally, substitute the coordinates of instant operating point (x, y) and constants a0, a1,

a2 and a3 or b0, b1, b2 and b3 back into equation (5-17) or (5-18) to obtain the

instantaneous turbine head or torque.

123
⎧a 0 + a1 x1 + a 2 y1 + a 3 x1 y1 = W H (1,1)


⎪a 0 + a1 x 2 + a 2 y1 + a 3 x 2 y1 = W H ( 2,1)

⎨ (5 - 19)
⎪a + a1 x1 + a 2 y 2 + a 3 x1 y 2 = W H (1,2)
⎪ 0


⎩a 0 + a1 x 2 + a 2 y 2 + a 3 x 2 y 2 = W H ( 2,2)

⎧ b 0 + b1 x1 + b2 y1 + b3 x1 y1 = W B (1,1)


⎪b0 + b1 x 2 + b2 y1 + b3 x 2 y1 = W B ( 2,1)

⎨ (5 - 20)
⎪b + b x + b y + b x y = W (1, 2)
⎪ 0 1 1 2 2 3 1 2 B



⎩b0 + b1 x 2 + b2 y 2 + b3 x 2 y 2 = W B ( 2, 2)

When the operating point P(x, y) is beyond the range of the characteristic curves,

extrapolation is necessary to obtain the values of WH(x, y) and WB(x, y).


B

5.4.2 Kaplan Turbines

Figure 5.4 shows the hill diagram of a Kaplan turbine, depicting the optimum relationship

between the runner blade position (blade angle ϕ) and wicket gate opening (a0). The

design of a Kaplan turbine provides a stepless control of both wicket gate opening and

blade position to achieve the best turbine efficiencies for a wide discharge range.

For a Kaplan turbine, using the same procedure as that for a Francis turbine, two

sets of curves WH ~x and WB ~x can be obtained for the series of wicket gate openings (y1,

124
y2, y3 ….) and for the series of runner blade angles (ϕ1, ϕ2, ϕ3 …), respectively. The values of

WH and WB for an instant operating point (x, y,ϕ) have to be interpolated or extrapolated by

two sets of curves, that is, WH (x, y), WB (x, y) and WH (x, ϕ), WB (x, ϕ).

Figure 5.4 Illustrative characteristics of a Kaplan turbine


(Revised based on the original source: Small hydropower series no. 4, p40 Fig.20)

5.4.3 Impulse Turbines

Figures 5.5 and 5.6 show the features of hill diagrams of Pelton and Turgon turbines.

Firstly, with variation of n1' the efficiency increases or declines rapidly. Secondly, with

variation of the turbine discharge by means of the nozzle needle, the efficiency remains

high over a wide range of operating conditions. Thirdly, at constant nozzle opening, the

value Q1' remains constant irrespective of the value of n1' .

Since an Impluse turbine is simulated as a valve, there is no need to convert the hill

125
diagram, but the nozzle characteristics (the relationship between the flow and nozzle

opening) should be specified by the manufacturer. However, the turbine characteritics

must be used to simulate the speed rise and runaway condition at load rejections.

a0 = const. η = const.

n11

ηmax

Q11

Figure 5.5 Illustrative characteristics of a Pelton turbine

a0 = const. η = const.

n11

ηmax

Q11

Figure 5.6 Illustrative characteristics of a Turgo turbine with inclined nozzle

126
5.4.4 Pump Turbines

The discharge and torque characteristics of a typical pump turbine are shown in Figure 5.7

(a) and (b). There are different characteristic curves for various gate openings and they

overlap with each other. On these charts, it is possible to find more than one value of unit

flow or unit torque at a given gate opening and a given unit speed, as illustrated in Figure

5.7 (c). The multi-valued area (so-called S-characteristic in the runaway zones) results in a

difficulty in application of these curves to transient simulation. For reversible machines

with the low specific speed, such multi-valued nature is pronounced as the unit’s operation

passes through this region, alternately experiencing positive and negative discharges as the

wicket gates close after load rejection. To overcome the challenge, various representations

have been used in numerical modelling and each presents certain merits and limitations

(Chaudhry, 1987 and Pejovic, et al., 1983).

The head WH and torque WB characteristics are defined as single-valued in the


B

entire area covered by the polar angle x = tan −1 (α / v) (Suter, 1966; Thorley, et al., 1966):

h
WH ( x) = SIGN (h) (5 - 21)
α + v2
2

β
WB ( x) = SIGN ( β ) (5 - 22)
α + v2
2

When using the Suter curves, difficulties arise from an uneven distribution of

127
curves at the wicket gate opening. At small openings, the curves space themselves out

further, and for a definite angle x when the wicket gates are being closed, the values of WH

and WB become indefinitely great. Consequently, the selection of interpolation and


B

extrapolation methods of these characteristics has at least some influence on the accuracy

of transient calculation results.

0.8 TAU=0.00
TAU=1.50
0.6
TAU=3.00
TAU=6.00
Unit Flow Q11 (cms)

0.4
TAU=8.80
0.2 TAU=11.9
TAU=14.9
0 TAU=17.9
-80 -60 -40 -20 0 20 40 60 80
TAU=20.9
-0.2
TAU=23.8
-0.4 TAU=26.8
TAU=29.9
-0.6 TAU=32.9

-0.8
Unit Speed n11 (rpm)

(a) Unit Discharge

Figure 5.7 Characteristics of a pump turbine (Con’d)

128
2500

2000
TAU=0.00
1500 TAU=1.50
TAU=3.00
1000
Unit Torque T11 (N.m)

TAU=6.00
500 TAU=8.80
TAU=11.9
0 TAU=14.9
-80 -60 -40 -20 0 20 40 60 80 TAU=17.9
-500
TAU=20.9
-1000 TAU=23.8
TAU=26.8
-1500 TAU=29.9
TAU=32.9
-2000

-2500

Unit Speed n11 (rpm )

(b) Unit Torque


Unit Discharge Q11

Unit Speed n11

(c) Multivalued characteristics

Figure 5.7 Illustrative characteristics of a pump turbine

129
5.5 Water Conveyance System Layout in Hydro Systems

5.5.1. Conventional hydropower system layout

In Figure 5.8, several water supply modes in conventional hydropower stations are

sketched; the “+” sign represents the necessary valve or gate; “×” indicates that the valve

may or may not be there depending on the situation and design. Mode (a) with separated

penstocks is suitable for the cases with short distance between the reservoirs/surge tank

and powerhouse. Modes (b) and (c) with combined water-supply pipes are economical for

long penstock cases, but the risk of hydraulic transients in the common water supply

pipeline could be severe in the case where all units simultaneously reject their power loads

and it is necessary to install a shut-off valve upstream of each turbine unit.

In this study, each turbine unit (not a hydropower station) is simulated as an

individual boundary condition (coded as BCTYPE=’TURB’), although several turbine

units are often arranged in parallel in a hydro station (as shown in Figure 5.8). This

numerical treatment is adopted for three reasons: firstly, each turbine unit might be

operated under different conditions or regulated in different ways; secondly, the pipeline

from the upstream junction to the turbine entrance can be long and pipes’ properties often

change along their route, and thus they usually cannot be neglected; finally, it is more

convenient in numerical modeling for each individual turbine unit than for a group of units.

In many conventional hydropower projects, the downstream conduit system

130
comprises a short draft tube, followed by a free-surface flow tunnel or an open channel. In

this case, the downstream conduit is typically negligible for transient analysis, and the

turbine unit is called a “one-node” turbine boundary condition (i.e., coded as a nodal

dependence number NDN=1). However, the pressurized tunnel/pipeline, even with a surge

tank, is common for those underground powerhouses and for turbine installations within

multi-use systems where the discharge water from the turbines may be directed to urban

utilities, industry users or irrigation systems.

Theoretically, the developed turbine model is applicable to any water system, no

matter how complex the system is, because the numerical models of almost all other

hydraulic devices have been previously developed in the simulation program TransAM.

For instance, the forebay or reservoir is a Constant Head Reservoir; the surge tank may be

a Linear Reservoir or a Tank With Overflow or an Air Chamber; and the valve distant from

the turbine can be a valve-in-line.

131
A B
T

Reservoir T
/ Forebay T

(a) Separated Water-supply Pipes

A B
T

T
Surge
Tank T

(b) Grouped Water-supply Pipes

A B
T

T
Surge
Tank
T

(c) Combined Water-supply Pipes

Figure 5.8 Water-supply modes in conventional hydropower stations

132
5.5.2 Energy recovery turbines in municipal water supply system

Unlike in a conventional hydropower system, there are three major features governing the

layout of energy recovery turbine units in municipal water supply systems: 1) there is

usually a long pipeline and pipe loops/networks downstream of the turbine, so transient

analysis should be carried out both upstream and downstream; 2) topographic variation is

often more significant along the pipeline, so the highest point should be closely scrutinized

to prevent vacuum pressure and cavitation; and 3) many different hydraulic devices in the

system (such as pumps, hydrants, etc.) could interact with the hydro turbine units.

5.6 Governor and Control System

The power load of a turbine-generator unit varies constantly with demand fluctuations and

the turbine governor and control system is designed to equate the dynamic power produced

by the water to the power required by the electrical system in order to ensure a stable power

supply frequency (either a large or isolated network). Also, it governs the turbine when the

unit starts up, is brought on line (synchronized), and shuts down normally or in an

emergency.

The governor control system for hydro turbines is basically a feedback regulation

system which senses the speed and power of the generating unit, or the water level of the

forebay of the hydroelectric installation, and takes a particular action for operating the

133
discharge/load controlling devices in accordance with the deviation of the actual control

variable from the reference/set point. An isolated turbine unit should use a feedback

control system to adjust its speed and power output. For the small and grid-connected units,

water level controllers may be used.

The governing system comprises of the control section and the mechanical

hydraulic actuation section. The control section may be mechanical (PI type controller),

analogue electronic or digital (PID type controller). The present trend is toward use of

digital governing control systems in hydroelectric units. The advantages of a

microprocessor-based system over the earlier analogue governors (based on solid state

electronic circuitry) include higher reliability, self diagnostic features, modular design,

flexibility of changing control functions via software, stability of set parameters, reduced

wiring and easy remote control through optical fiber cables.

The turbine control actuator can be hydraulic, mechanical (motor) or a load

actuator. A load actuator is used in micro hydro range (through adjusting the shunt/bypass

of the load bank), while mechanical (motor operated) actuators may be used for units up to

1000 kW in size. Overall, hydraulically controlled actuators (pressure oil system with oil

servomotor) are most commonly used. The actuator system compares the desired actuator

position with its instant position. In most hydroelectric units it requires positioning of the

wicket gates in reaction turbines, runner blades in Kaplan turbines and needles in Pelton

turbines (2008 MNREG, India).

134
5.6.1 Three Levels of Turbine Flow Control

The output of turbine power is controlled by adjustment of turbine flow discharge; there

are three levels of turbine flow control.

1) Speed-control governor

A speed-control governor is the most popular control device. During normal operation, it

continually adjusts the gate opening (and/or runner blade angle) according to the variations

of power load exerting on the turbine shaft to maintain the unit speed at the

rated/synchronous speed. In addition, it operates to ensure a smooth shutdown or start-up

of the turbine unit.

Large hydroelectric installations are usually equipped with hydraulic-mechanical

or hydraulic-electric governors. Depending upon the criteria for the corrective action, the

hydraulic governors may be classified into three types: a) temporary-droop governors; b)

accelerometric governors; and c) proportional-integral-derivate (PID) governors. In a

temporary-droop governor, the corrective action of the governor is proportional to the

speed deviation (α =N/NR) and its integral with time. In an accelerometric governor, the

speed change is proportional to the derivative of speed deviation ( dα dt ). And PID

governor is the sum of actions proportional to the speed deviation (α), its derivative

( dα dt ), and the integral of speed deviation with time. The PID was introduced in the

early seventies and is now being used extensively.

135
2) Non-speed-control governor

Small hydroelectric installations generally have little effect on the frequency of the power

grid and thus can have governors without speed regulation for cost savings.

Non-speed-control governors may be either hydraulically or electrically-operated to

control the gate opening (and turbine discharge) based on measuring forebay water levels.

Their function is to bring the turbine to near synchronous speed for start-up, to regulate

load after synchronous speed has been achieved and to shut down the unit during both

normal and emergency conditions.

3) Inlet control valve

In cases where load regulation is not required (with small or micro capacity sizes), the

governor is not needed but a main control valve or shut-off valve must be installed and able

to shut down the unit under normal and emergency conditions.

In a large turbine system, the main control valve is also equipped with a governor

(dual control system). The control valve could be a butterfly, a deadweight butterfly, a gate,

ball, or rotary valve. It is installed immediately upstream of the turbine entrance to shut off

the flow when the turbine water passage is under inspection, overhaul or standing idle. In

addition, it is also used to shut off the water flow in the emergency of wicket gate failure

during load rejection in order to protect the turbine-generator unit from a protracted

runaway speed condition.

136
5.6.2 Mechanism of a Speed-Control-Governor

With respect to the mechanism of a speed-controller, two basic concepts require

clarification: (1) the turbine speed change is controlled only by the net torque acting on the

shaft of turbine-generator; and (2) the speed change, acting through the governor, controls

the main servomotor. The main servomotor is the governing system device that adjusts the

energy input the turbine. This mechanism might be a needle valve, wicket gates, runner

blades, or a combination of them and other elements.

In a turbine unit, if friction is negligible, we have the following torque balance

equation (speed change equation):


I = T − Tg (5 - 23)
dt

Here, I =WRg2, is the polar moment of inertia of rotating fluid and mechanical parts in the

turbine-generator unit, in which W is the weight of turbine-generator rotating parts and

fluid contained and Rg is the gyration radius; ω is the angular speed of the unit

( ω = 2πn / 60 where n is in rpm); T is the torque produced by water flowing through the

unit; Tg is the resistant torque from the generator.

With a rise in the electrical load, Tg increases and the unit slows down, the

governor would activate the gate opening to augment the flow discharge, allowing the

hydraulic torque T to increase until a new torque equilibrium is achieved. The reverse logic

137
holds when electric load (and Tg) decreases.

The dynamic characteristics of a governing system represent the variation of

adjustable parameter (e.g., the speed n) with time during the governing process. The

stability of a governing system is essential; besides, the good quality of governing system

also requires minimizing the amplitude and duration of the deviation of turbine speed from

the rated speed within the smallest number of oscillations.

The static characteristics of a governing system indicate the relationship between

the adjustable parameters, usually the speed n, and the output power Pg during steady state.

There are two types of static characteristics. One is the no-droop governing system, in

which the speed of turbine is unrelated to the load Pg; no matter the load value, the speed

keeps at the rated value nR. The second governing system relationship allows for a small

difference between the new steady state and the old one, i.e., there is a permanent speed

droop σ. Usually σ is between 0 to 0.08, which is a relative value and defined as

σ = (nmax − nmin ) / n R , nmax is the speed at no-load condition, nmin is the speed at rated load

condition, and nR is the rated speed. Most units in the network adopt a drooped governing

system because it facilitates a fixed load; those units that provide the network as a whole

with frequency control use a no-droop governing system.

138
5.7 Fundamental Knowledge of Synchronous Turbine-Generator Units

Synchronous units (on-line units) mean that the turbine-generators feed their output to a

large electricity grid. In the modern hydro industry, most turbine units are linked to

electricity networks. Therefore, it is necessary to understand the fundamentals and basic

features of synchronous units.

5.7.1 Load Distribution and Unit Operation with a Governor

In a large interconnected system, there are usually many tie lines between power stations,

and each power station would adjust its output to cover local load and tie line loads, which

can be explained as the load borrowed from somewhere else. The tie lines do not start and

end at generating stations, thus it is difficult to control the amount of load for each tie line.

The frequency of the power system must remain stable to ensure the quality of

power supply. The question is how to operate a unit at a plant to maintain this master

frequency. “Frequency Signaling” is used to operate the plant or the unit. A clock

reflecting the speed of the unit at the plant is compared with a master clock reflecting the

system frequency and the difference shows whether the unit’s speed is too high or too low.

The output of the plant/unit is then changed according to the given frequency error.

The governor usually enables straightforward synchronization, particularly for

the units that carry peak loads. The speed-drop characteristic is also desirable to keep

139
several turbines working in parallel; otherwise the load carried by each unit cannot be fixed

when the new balance is being reached while the system experiences load changes. In fact,

the permanent speed drop causes a slight speed alteration as the load changes; however,

this small speed-drop could be re-adjusted after the load variation, which inevitably

requires expensive electrical equipment. However, the final allowable master frequency

fluctuation is only 0.1-0.4%, and thus can be neglected for hydraulic transient analyses.

Therefore, the assumption of synchronous speed for on-line turbine units is reasonable and

the torque balance (or speed change) equation is invalid for the units connected to a large

power network.

For unit speed fluctuations, the governor would adjust the wicket gate opening

slightly and accordingly to maintain the constant master frequency, while, for a larger load

variation of on-line units, the adjustment involves both the governor function and load

re-distribution among units in the power system. Load re-distribution is planned by the

dispatcher who determines the proportion of total load to be carried by each individual

plant or unit. A massive unplanned load change (e.g., a large power generation plant is

accidentally shut down) may cause too low a frequency of the local power network and

thus automatically shut down other generating units connected to the network as a

protective measure. This action would cascade down to the neighboring and connected

power networks (because the local load demand is transferred to the neighboring power

140
networks) if there were no isolation measures taken among the interconnected networks, a

unfolding of events experienced during the August 2003 North American Blackout.

Close frequency control and proper load distribution among units and plants are

requirements which cannot be simultaneously met by a governor because they are

contradictory. It is therefore practical to maintain that feature of the governor which

assures the best distribution of the load and realizes as constant a frequency as possible via

auxiliary corrective means that can be applied gradually enough so as not to offset the

already established proper load distribution. This indicates that for a significant load

change of a particular unit, it is the only way that the turbine gate is opened/closed at a

prescribed rate or by a time function to the position at which turbine output will be equal to

the required final output. In other words, the gate opening cannot be adjusted by the

governor for a required significant load change (Chaudhry, 1987).

In this thesis, the above recommendations are followed when the unit is connected

to a large electricity network. However, further discussion from the literature (Zaruba,

1993) is summarized below:

a) During the time when the generator is connected to a power network, the moment of

inertia considered in the speed change equation is not constant and is defined by the

relation Ic = I + fI (t). Here, I is the constant inertia of the turbine unit, and the

function fI (t) represents the effect of the power network.

141
b) The resistant torque Tg of the generator may be dependent on the turbine speed, but

also on the connected power network. The effect of the network may change with time.

The resistant torque employed in the speed change equation is defined by the relation,

Tg = (n-nE) fn(t) + fc(t). The functions fn(t) and fc(t) represent the effect of the

connected power network (the effect of self-regulation and the moment of loading),

(n-nE) is the speed deviation from the synchronous speed.

c) With above two corrections, the control regime of the turbine blades angle and

wicket-gate opening is not set in advance but is defined by the function of turbine

speed governor, i.e. the governor equation.

5.7.2 Startup and Load Acceptance

To start up a unit, the wicket gates are first turned on to an opening at which the generated

power is sufficient to offset static friction. This gate opening is maintained until the unit

speed reaches about 60% of the rated speed; then the wicket gates are closed to the position

of speed-no-load and during a short period of time the unit is running at the synchronous

speed and ready for load acceptance. Since no significant flow change occurs during the

synchronizing process (unit start-up), pressure surges are only concerned during load

acceptance; that is, when the turbine opens from the speed-no-load condition

(corresponding to zero load at speed-no-load opening and synchronous speed) to the

142
normal operation condition (corresponding to the target load at final opening and

synchronous speed).

5.7.3 Shutdown and Load Rejection

There are two types of load relief responses. One is associated with the normal shutdown

of the unit and the other is load rejection. Load rejection is a sudden unexpected load

cut-off, which is a frequent occurrence in hydro units that results from the power failure,

device malfunction or some other accident; load rejection actually serves as a kind of

protection for the system and units.

In normal unit shutdown, the gates are first closed to a gate opening

corresponding to speed-no-load condition. In this nil efficiency operating condition, the

turbine is still connected to the line and running at its synchronous speed; from this

opening, the gates are slowly closed to zero and the unit is taken out of grid/service. The

speed-no-load opening can be found from turbine characteristic curves (zero-efficiency or

zero net torque point), and it varies from 2-3% at high head plants to up to 35-40% at low

head propeller wheels.

However, when load rejection occurs, the unit is instantly disconnected from the

system. Then, the unit has to be quickly closed to prevent excessive runaway condition,

which may induce severe draft tube pressure drops.

143
5.7.4 Speed-No-Load (SNL) and Runaway Condition

The speed-no-load (SNL) gate is the lowest gate opening at which the turbine maintains

synchronous speed and carries no (zero) load. In this condition, the turbine power is just

sufficient to overcome the turbogenerator windage and friction losses at synchronous

speed, so the net output is zero. The value of SNL opening varies with the net head, since

lower head needs greater flow (greater gate opening) to offset the losses. SNL gate opening

with corresponding flow and head (zero-efficiency or zero net torque point) are usually

provided by turbine performance curves.

To keep the unit running at synchronous speed when the gate opening is less than

the SNL opening, an outside power source must be supplied, a procedure known as

motoring the unit.

When a load rejection occurs, the wicket gates are supposed to be closed quickly

to prevent extended periods of high over speed. However, if the closure fails, such as when

there is a malfunction of the governor, the imbalanced torque induces an acceleration of

the unit shaft. With the turbine speed growing, friction loss escalates until finally the

available hydraulic power is fully dissipated by friction losses in the machinery; the net

output of the turbine is zero, and the acceleration is reduced to zero with the speed reaching

its maximum value. As aforementioned, this is defined as runaway condition and the

maximum rotational speed is called runaway speed.

144
The runaway speed range varies according to unit type. Typical runaway speeds

for Kaplan turbines are 2.5 to 3 times normal speed, whereas with Impulse and Francis

turbines they are rarely over twice normal. Runaway speed values are measured during the

steady-state model tests of the turbine. In operation, a unit cannot exceed the runaway

speed for a particular net head and gate opening. Turbo-generating machinery must be

designed to be strong enough to withstand runaway speed, but with a highly reduced safety

factor. However, as mentioned in Section 5.3, the prototype speed may exceed the

steady-state runaway speed for a short duration in transient states since the instantaneous

head is governed by waterhammer effects. Therefore, the turbine steady state characteristic

curves have to be extended to obtain the transient runaway speed. The same concepts and

observations about runaway condition hold for both isolated turbines and the turbines

originally interconnected.

5.8 Features of Transients Caused by Governed Turbines

5.8.1 Key Features of Turbine Transient Analysis

Compared to pumps, the modeling of a governed turbine includes two new features that

further complicate the problem: first, the governor produces a change in wicket-gate

position or runner blade angle in response to a speed change, therefore an equation is

needed to describe the dynamic response of the governor. Second, turbine characteristic

145
curves are different for every gate position and, thus, additional characteristic curve data

must be available for the numerical solution. For a pump, only one curve is needed for each

of the head and torque characteristics, respectively (a pump with variable characteristics –

other than pump speed – is highly exceptional); for a turbine, two families of curves

represent H and T characteristics with different gate positions have to be provided.

5.8.2 Speed Variation

A change in turbine operating condition results in transient flow in the turbine and in the

associated hydraulic system. This change may be due to a load adjustment, a unit start-up,

or a unit stoppage, either normally or by accident. Among other things, these adjustments

cause an instantaneous imbalance between the power absorbed by the generator and that

produced by the water; the resulting shaft force changes the rotation speed of a

disconnected unit. For reaction types of turbines, such as Kaplan and Francis turbines, the

turbine speed exerts a considerable influence on the turbine flow, and the speed changes

should therefore be taken into consideration in transient analysis. However, the speed

variation is small and can be neglected if a unit keeps on line and the power of this unit is

only a small portion of network power, since the speed has to correspond to the

synchronous speed whenever its load increases or reduces.

When we calculate the waterhammer caused by load rejection or load variation,

the speed variation must be taken into account for the following two cases:

146
(a) The unit of turbine-generator is operating stand-alone or in a small power system;

whether for the case of load rejection or load variation, the speed change has an influence

on the discharge and efficiency for most turbines types. The selection of criteria to

determine whether a power system is large or small is somewhat complicated. In general,

any unit in a system that would be supplying 40% or more of the total capacity should be

designed as an isolated unit. The relationship between system capacity and load change

should also be compared. Any load change of 10% or more of the system capacity should

be analyzed to determine its effect on frequency (Gordon, 1961).

(b) If the unit is initially on-line (connected to a large network), when the load is rejected

it would be disconnected from the power network immediately and thus become isolated.

The maximum allowable speed rise is approximately 45-65%. During load

rejection, reducing the duration of regulation (the closure time of wicket gates or control

valves) can decrease the speed rise but will bring about an excess waterhammer (pressure

rise), so that the reasonable closing time becomes a compromise between the pressure rise

and speed control, which is called a Guaranteed Regulation Calculation.

Is it necessary to calculate the speed change of an impulse turbine for the purpose

of waterhammer analysis? The answer is no, since the speed change has no influence on

the turbine discharge. But for the purpose of speed control, the torque equation is suitable

for any type of turbine, though naturally each turbine has its own characteristics curves.

147
5.9 Summary

In this chapter, the classification of turbines, their key parameters, and the characteristics

of some typical turbines are reviewed. Following this, hill diagram conversions for Francis

turbine are introduced. Then, the layouts of a water conveyance system for conventional

electricity dedicated hydropower and energy recovery turbine systems are briefly

discussed. And the hydraulic control devices and their working mechanisms are

summarized. Finally, the basics of synchronous units and different scenarios of transient

simulations, as well as the features of turbine transient analysis, are introduced.

The focus of this overall review is to establish the relevance of the system’s

characteristics to the transient numerical modeling, which is essential for an understanding

of the numerical model of hydro turbine and transient analysis in the next chapter.

The challenges for hydro turbine modeling arise from several factors: (1) turbine

operation interacts with hydraulic, mechanical and electrical systems; (2) turbine

characteristics have more interdependent parameters, such as the turbine head, flow, speed,

efficiency, wicket-gate opening and even the blade angle for Kaplan turbines; and (3) in

some parts of the water passage, such as, in the runner and draft tube, the flow is fully

three-dimensional and might be mixed with air. Despite above challenges, reasonable

accuracy in prediction of transient peak values is required to prevent waterhammer

problems and failures.

148
Chapter 6 Hydro Turbine Modeling and Model
Verification

This chapter is another core contribution of the thesis. It develops a numerical model and

computer program for transient analysis of hydro turbine units, which is incorporated

into TransAM, a transient analysis model specifically designed for complex networks. To

date, few commercially available transient analysis models for the turbine installation

allow for the existence of a complex network. In this model, the turbine units could be

either isolated or interconnected; and with or without attached valves. The simulation

results are compared a published example (Wylie, et al., 1993) and also verified by a set

of field data measured during the test of large hydro plant. The developed model has

become a useful tool to evaluate transient performance and to find effective measures to

transient mediation and control in the systems with installation of turbine units (as shown

in case studies of Chapters 7 and 8).

6.1 Mathematical Model of Turbine Boundary Condition

A general mathematical model for active hydraulic devices has been developed in

Chapter 3. However, hydro turbine units have their own features and variables and thus

their governing equations are developed in this section.

149
6.1.1 Head Balance Equation without Turbine-Attached Valves

For the configuration of turbine system as shown in Figure 6.1, the turbine upstream

node is often set at spiral case inlet and the downstream node at draft tube exit. There is

an energy balance (or head balance) equation between these two nodes:

2 2
Q1 Q
F1 = ( H 1 + ) − ( H 2+ 2 2 ) − H = 0 (6 - 1)
2 gA12
2 gA2

Here, H is the difference of hydraulic head (i.e., the energy per unit of weight of water)

between the turbine upstream and downstream nodes, the so-called turbine net head. H1

and H2 are instantaneous piezometric heads, including the nodal elevation and pressure

heads. To account for the kinetic energy head, the nodal external flow (Q1 or Q2) and

pipe cross-sectional area (A1 or A2) at upstream and downstream node, respectively, are

used to calculate the velocity head at the nodes.

Pipe
Pipe

U/S Node T D/S Node

Q1 Q Q2
Pipe Pipe

Figure 6.1 Schematic of turbine boundary condition

150
As described in Chapter 3, taking the turbine upstream and downstream nodes as

the junctions in a pipe network, and H1 and H2 can be calculated from the extended MOC

equation, that is,

H 1 = C1' − B1' Q1 (6 - 2)

H 2 = C 2' − B2' ( −Q2 ) (6 - 3)

Here, the constants B1′ and C1′ at the turbine upstream node and the constants

B2′ and C 2′ at the downstream node can be calculated by equation (3-16). It has been

noted that the sign of external flow Q1 is positive and of Q2 is negative, and their values

are equivalent if water losses are negligible at the turbine boundary condition. With the

turbine flow Q, we have the flow continuity equations at nodes A and B:

Q1 = Q = Q2 (6 - 4)

This relation explicitly excludes any small capacitance or water storage within the

turbine, which is usually an excellent approximation except when considerable air is

present. If air or other capacitance effects are desired to be included, minor medication of

the solution procedure can be undertaken.

The draft tube that connects the turbine to the downstream reservoir can

sometimes be very short (e.g., straight cone draft tube); indeed the tailrace could be

non-pressurized, so the transient effect in water passage of turbine downstream would be

151
negligible. In this case, the water level in tailrace channel (or downstream reservoir),

HTail, is directly used in head balance equation (6-1), which results in:

2 2
Q1 Q2
F1 = ( H 1 + 2
) − ( H Tail + 2
)−H =0 (6 - 1a)
2 gA1 2 gA2

This configuration of turbine connection to system is called one-node-turbine boundary

condition, since the turbine unit is associated with only one node at the end of system

pipeline, and the variable NDN=1 in TransAM program. A NDN value of 2, for the

two-node-turbine boundary conditions, is typical in multi-use hydro systems, energy

recovery urban water systems, or many conventional hydropower plants especially with

underground powerhouses.

Based on equations (6-1) to (6-4), it is straightforward to obtain a general form

of the head balance equation, with respect to the turbine flow Q and net head H:

F1 ( Q , H ) = 0 (6 - 5)

Furthermore, substituting the turbine flow Q = QR v and head characteristic

relationship H = H R (α 2 + v 2 )(a0 + a1 x + a2 y + a3 xy) into this equation, a general form of

head balance equation in dimensionless variables α, v and y can be obtained:

F1 (α , v , y ) = 0 (6 - 6)

The above equation is referred to the first basic/governing equation of the turbine

152
boundary condition. In this equation, both the turbine gate opening y and the speed α

become explicit unknown variables though they are implicit in equation (6-5). Moreover,

the turbine characteristics have been implied in this equation.

6.1.2 Head Balance Equation with Turbine-Attached Valves

In hydro turbine systems, a main control valve (i.e., a shut-off or control valve) is often

mounted upstream of the turbine, either inside the powerhouse or immediately outside it.

The pipe between the valve and spiral-case entrance is typically only 2-3 m long, which

can create numerical challenges. To avoid too short a pipe section, and thus too small a

time step, it might be more appropriate to combine the control valve within the turbine

boundary condition rather than treating it as an individual hydraulic device.

In some reactive turbine installations, a Surge/Pressure Relief Valve (SRV or

PRV) is used to mediate waterhammer pressures during load rejections, together with, or

instead of, a surge tank. Turbine SRVs are also regulated by the turbine governor; their

function and working mechanism is discussed in Section 2.3 of Chapter 2. This type of

SRV could be combined with turbine boundary condition in numerical modeling.

These turbine-combined valves are also referred to “attached valves” in this

thesis. They occur more often for the small/micro hydro turbines in urban water systems

than for the large conventional hydropower plants. One reason is that a dedicated surge

tank is often economically inefficient for small/micro turbines in urban water systems;

153
moreover, the connection pipes between the valve and turbine are often relatively short

for the urban long pipeline systems.

In the current code, a maximum of three attached valves can be combined within

each turbine unit, as shown in Figure 6.2, including an upstream main control valve

(TvalTYP=’TCV’), a bypass valve (TvalTYP=’Bypass’), and a surge relief valve

(TvalTYP =’Surge’). There is little difference between the ‘Bypass’ valve and ‘Surge’

valve in modeling, since they both have multiple settings for different parameters.

The “degree” of a node is defined as the number of pipes connected to the

node/junction. Each hydraulic device (i.e., a boundary condition in pipe system) is

considered to be associated with one or two nodes; a zero degree of node (an unattached

device) is not permitted. Therefore, depending on the distance from the turbine, the main

control valve can be taken as either a ‘valve-in-line’ or a turbine-attached valve. As a

separate b/c –‘valve-in-line’, there has to be a connection pipe between the valve and the

turbine, so its simulation has already been completed in TransAM. As for the solution

of “TCV”, the valve head loss has to be accounted for in the head balance equation. In

addition, the nodal continuity equations can be used in the flow relationship between Q,

QSV, QBV and QCV (see Figure 6.2) when the surge valve and bypass valve are present in

the system. The detailed description of turbine attached valves can be found in Appendix

Ⅲ.

154
If an attached turbine main control valve is included, the additional head loss HC

across the ‘TCV’ has to be taken into account in the head balance equations. Thus

equation (6-1) turns to (6-7), and (6-1a) turns to (6-7a):

2 2
Q1 Q2
F1 = ( H 1 + 2
) − ( H Tail + 2
) − H − HC = 0 (6 - 7a)
2 gA1 2 gA2
2 2
Q1 Q
F1 = ( H 1 + ) − ( H 2+ 2 2 ) − H − H C = 0 (6 - 7)
2 gA12
2 gA2

The continuity equation (with the usual assumptions) is applied in the scheme of Figure

6.2, and we have:

Q1 = QCV = Q + Q SV + Q BV = Q 2 (6 - 8)

Q CV = τ C E C HC (6 - 9)

Q SV = τ S E S H (6 - 10)

Q BV = τ B E B H (6 - 11)

As indicated in Figure 6.2, HC represents the head loss across the control valve ‘TCV’.

The subscripts C, S, B in equation from (6-8) to (6-11) indicate different valve types,

such as, QCV is the flow through the control valve, QSV is the flow through surge valve,

and QBV is the flow through bypass valve. As usual, τ represents the relative valve

opening, E represents valve resistance coefficient. For bypass valve and surge valve,

valve opening and coefficient can be set flexibly; for example, setting EB = 0 indicates

155
there is no bypass valve present; setting τ B = 0 if the bypass valve is present but fully

closed. However, a logical variable is necessary to identify whether ‘TCV’ is present or

not, and then to decide which head balance equation to use in the model.

H1

HC

HGL
H

QCV H2
Q

U/S Node Q1 Q2 D/S Node


‘TCV’
QSV

Surge Valve

QBV

Bypass Valve

Figure 6.2 Scheme of the turbine attached valves

Substituting equation (6-9) to (6-11) into equation (6-8), the headloss across the

control valve can be derived:

156
2
⎡ Q v + (τ S E S + τ B E B ) H ⎤
HC = ⎢ R ⎥ (6 - 12)
⎣ τ C EC ⎦

Further substituting equation (6-8) to (6-12) into the head balance equation (6-7) or

(6-7a), the general form of turbine Q and H relationship, F1 (Q, H) = 0, can be derived as

well for the turbine boundary condition with attached valves. Therefore, the head balance

equation, as the first governing equation, always implies a relationship between turbine

flow Q and head H no matter ‘TCV’ present or not, no matter with or without bypass

valves.

6.1.3 Speed Change Equation (Torque Balance Equation)

The torque balance equation has been described in Section 5.6.2 of Chapter 5.

Multiplying ω at both sides of equation (5-23), we have:


Tω − Pg = Iω (6 - 13)
dt

In above equation, Pg is the power absorbed by the generator, reflecting the variable

power demand. During normal operation, the variation of Pg with time is typically

planned or can be predicted in advance. Divided by the product of TR ω R at rated

operation condition, equation (6-13) can be further deduced to

T ω Pg ω d ωωR
− =I ⋅ωR (6 - 14)
TR ω R TRω R TRω R dt

157
Using dimensionless turbine performance parameters α , β as defined in Chapter 5, the

equation (6-14) can be further transformed to:

Pg dα
βα − = Tmα (6 - 15)
PR dt

or

dα Pg
Tm =β− (6 - 16)
dt PRα

Here, Tm = I ωR2 TRωR = IωR TR is the mechanical starting time. Integrating (6-16) over

∆t, the second basic equation for a turbine boundary condition can be derived:

1 Pg Tm
β− − (α − α 0 ) = 0 (6 - 17)
PR α Δt

where α 0 represents the value of dimensionless turbine speed at the previous time-step;

β = (α2 + v2) (b0 + b1 x+ b2 y + b3 x y) represents the torque characteristics of the turbine,

so equation (6-17) can be expressed in dimensionless variables using the general form of

F2 (α , v, y ) = 0 (6 - 18) .

The speed-change equation is invalid for interconnected units, since in this

case the inertia of associated power network has to be taken into account. Instead, the

constant synchronous speed can be assumed for the interconnected turbine units.

158
6.1.4 Governor Equation

As introduced in Chapter 5, the function of a governor is at all times to supply the turbine

with a discharge which, multiplied by instantaneous net head and the instantaneous

efficiencies of turbine and generator, would match the power demand on the generator

shaft. Specifically, the objectives of governing are to minimize speed variations caused

by load changes and to provide adequate stability and a quick return to normal speed.

When the load changes, the unit will tend to speed up or slow down. The governor has no

way of knowing of load change before hand, but it detects the speed changes and acts to

compensate for the speed change by opening or closing the turbine gates. The action of

the governor is always a correcting action; there must be a deviation from normal speed

(a speed error) detected by governor to produce correcting action to compensate for this

error.

The equation of a PID speed-control governor can be derived (Wylie, et al.,

1993):

d2y dy dα
Td Tα + Tα' + σ ( y − 1) + (α − 1) + Td =0 (6 - 19)
dt 2
dt dt

Where y = dimensionless turbine gate opening; α = dimensionless turbine speed as

defined in Chapter 5; Td = dashpot time constant; Tα = promptitude time constant without

temporary speed droop; Tα' = Tα + δ Td , promptitude time constant with temporary

159
speed droop; δ = temporary speed droop; σ = permanent speed droop.

Generally, Tα << Td , the above equation reduces to

dy dα
Td δ + σ ( y − 1) + (α − 1) + Td =0 (6 - 20)
dt dt

Let dy/dt=Z, the rate of the movement of turbine gate opening. To discrete over Δt, we

have:

y − y0
Z= (6 - 21)
Δt

d 2 y dZ Z − Z 0 ⎛ y − y0 ⎞
= = =⎜ − Z 0 ⎟ Δt (6 - 22)
dt 2
dt Δt ⎝ Δt ⎠

Substituting the above two equations into the governor equation (6-20) and integrating

over Δt, we have:

⎛ y − y0 σΔt
⎞ T
( y − 1) + Δt (α − 1) + 1 (α − α 0 ) = 0
'
⎜ − Z 0 ⎟ + α ( y − y0 ) + (6 - 23)
⎝ Δt ⎠ Td Tα Td Tα Td Tα Tα

Here, subscript ‘0’ represents the values at the previous time step; and the dimensionless

opening y represents the main servomotor position that may control a needle valve,

wicket gates, runner blades or a combination of these or other elements. For a Kaplan

turbine, both the wicket gate opening and the runner blade angle have been obtained once

y is solved if the cam relationship between runner blade servo stroke and wicket gate

160
servo stroke has been designed at the specific net head; otherwise, the runner blade angle

at the instant operating condition has to be found from the set of turbine characteristic

curves of WH (x, ϕ) and WB (x, ϕ).

Equation (6-23) applies for mechanical, electric or digital governing system, as

long as the coefficients in the equation are properly identified. For different type of

governor, the governor equation may changes. Yet, the third basic equation for a turbine

boundary condition regarding its governor can always be generalized as

F3 (α , y ) = 0 (6 - 24)

6.2 Combination of Basic Equations for Different Scenarios

The head balance equation, the speed change equation and the governor equation

discussed above are the three basic equations, which govern a turbine boundary condition.

They are applied in different combinations according to the specific problem and

operation conditions. Several common scenarios and the corresponding equation sets are

summarized as follows.

1) Isolated Units

Isolated unit refers to the unit operating stand-alone or supplying its power generation to

a small isolated power system. A speed control governor is essential for an isolated unit.

161
Head balance equation F1 (α, v, y)=0, speed change equation F2 (α, v, y)=0, and governor

equation F3 (α, y)=0 constitute a set of governing equations for transient analysis of an

isolated turbine unit. Therefore, three unknowns gate opening y, turbine speedα and

turbine flow v can be obtained.

For isolated units without governors (usually only run-of–river installations with

small capacity), the servomotor stroke of turbine gates are usually set at full opening

position, or set manually to match the site flow condition, that is, y = y (t) is known. If

there is a governor installed but it is out of function, then the gate position remains at the

opening of initial state during the transients, that is, y = y0 is known. Obviously, the

governor equation is no longer valid under these two scenarios. So, the head balance

equation F1 (α, v) = 0 and speed change equation F2 (α, v) = 0 are applied to solve two

unknowns of turbine speed α and turbine flow v during the transient states.

As aforementioned, any turbine units would be disconnected when load

rejection happens. So, for transient analysis of a load rejection, the units are always

treated as isolated ones.

2) Interconnected Units

Whenever the demand changes in a large electricity system, there is a load re-distribution

among tie line loads. In other words, the load variation is assumed not only by a specific

unit but by all connected units, and the load Pg applied on the specific unit remains

162
around the dynamical power produced by the turbine, i.e., Tωηg = γQHηηg = Pg during

adjustment of the gate opening. Therefore, the normal speed change equation no longer

exists for an interconnected unit.

If the permanent speed droop σ and the dynamic variation of unit speed during

the process of adjustment are negligible, the speed of connected turbine can be

considered as constant and it is the synchronous speed which depending on system

frequency (n = nE). α =1 is known; and the governor equation is reduced to F3 (y)=0

which means the gate opening y is actually set directly.

The speed regulation governor is not as necessary for those units carrying on

block load (also called as base load) in the daily load diagram, particularly for those units

with extremely long penstocks. In cases with extremely long penstocks, a longer time

period for both opening and closing of gate is necessary to limit the waterhammer when

the unit rejects or accepts any load, which would cause a greater over speed. Therefore, it

is appropriate to let the unit only produce a constant load within the grid to avoid

frequently adjusting flows, with consequent savings to the governing system. With

constant speed (n=nR), α =1 is known; without the governor, the turbine opening has to

be set automatically or manually, i.e., y=y (t) is prescribed, which is applied into the first

basic equation: F1 (v, y)=0, then the turbine flow v can be obtained step by step.

163
6.3 Computer Implementation

A computer program for the solution of a turbine boundary condition was developed in

FORTRAN90 based on the mathematical model developed above. Except for some minor

additional and modifications to a few of the related codes, only a few new subroutines

had to be associated with turbine boundary condition. These include:

Turbdat.f90 — To read the input data and se up the initial conditions.

Turbintss.f90 — To initialize the steady state of turbine boundary condition.

Turb.f90 — Core subroutine of turbine b/c, the solution to the set of basic equations.

ShapeWHH.f90 — To interpolate the turbine head characteristics.

ShapeWBB.f90 —To interpolate the turbine torque characteristics.

6.3.1 Steady-State Simulation

In modeling, the input initial data from the input file are often inconsistent with each

other, so it is necessary to start simulation with initial steady state to eliminate the “false

transient conditions”.

6.3.2 Transient Simulation and Flow Chart of Programming

The transient simulation follows the procedure charted in Figure 6.3.

164
No Inverse
With Generator?
Pump
Yes

No Synchronous
Isolated Unit? Speed (α=1)

Yes

Yes (No Governor, or Governor Malfunctioned)


Governed Unit?
Valve-Controlled Unit

The opening of wicket-gate is


prescribed, and/or the flow
Basic equations: controlled by shut-off valve or
F1 (α, v, y)=0 other valves. Basic Equations:
F2 (α, v, y)=0 y=y(t)
F3 (α, y)=0 F1(α, v, y)=0
(α=1 and remove F2 (α, v, y)=0 F2 (α, v, y)=0
for Interconnected Units) (α=1 and remove F2 (α, v, y)=0
for Interconnected Units)

Solve the Eqs. by


Newton’s Method

Figure 6.3 Flow chart of turbine boundary condition programming

165
6.3.3 Data Input and Output

A new interface was created to effectively input system data and to visualize output of

the simulation results. For further development of the turbine boundary condition, a

sample set of data for a turbine boundary condition and the definition and description of

each input variable is provided in Appendix Ⅲ. A general summary of turbine boundary

condition is includes in the output data table. A sample output data abstracted from the

data file of *.BND and the definition of each variable in the time series are listed in

Appendix Ⅳ. The output data can also be found and reflected in Debug and Graphic

files.

6.4 Model and Program Verification

6.4.1 Comparison with Wylie’s Simulation

An example from literature (Wylie, et al., 1993) is used to verify the model and computer

program of turbine boundary condition developed in this thesis. This is a

reservoir-pipe-turbine system, in which a turbine unit is located at the downstream end of

the penstock. The draft tube is short and thus can be ignored, so it is a one-node turbine

boundary condition, which means the turbine downstream pressure head is constant. The

turbine unit is assumed to be an isolated one, and a PID speed control governor is used to

adjust the wicket gate opening during load rejection and variation, but the maximum rate

166
of wicket gate closure is limited by the shortest closure time period Tf = 6.5 s. The

turbine and system data are listed in Table 6-1.

Table 6-1 System data of Wylie’s example

Pipeline Turbine Unit Governor

L = 411 ft HR = 269 ft Td = 3.7 s

A = 254 ft2 NR = 200 rpm Tα = 0.325 s

a = 4100 ft/s QR = 4025 ft3/s Tf = 6.5 s

TR = 3.03(106) ft-lb δ = 0.18

WR g 2 =35.3(106) lb-ft2 σ = 0.0

Three cases are simulated. The first case is that the unit initially generating 61.7

MW of power and then rejected within 0.1 s; the second case is that the same initial load

reduces to 44.8 MW in 1 s; and the third case is that the same initial load increases to 81

MW. The simulated responses of turbine dimensionless variables to above three load

variations are showed graphically in Figures 6.4, 6.5 and 6.6, respectively, in which the

solid lines represent simulations of this thesis; and the dashed lines represent the

simulations from Wylie’s classic book, of which the program and data files are directly

cited from the appended disk of the book. By the way, other case studies do not show this

obvious difference between these two interpolation methods.

167
The comparison in Figure 6.4 shows that two simulations are generally quite

close to each other. However, the difference is noticed in the case of load rejection, as

shown in Figure 6.4 (a), for the simulation of turbine head at the range of wicket-gate

opening from y = 0.4 to y = 0.0. The turbine head, which is actually the pressure head

at the end of penstock in this case, analyzed by this model, has obvious “waves” while

Wylie’s simulation is quite flat. This difference is believed to arise largely from the

slightly different interpolation methods used for turbine characteristics and the particular

turbine characteristics. Figure 6.5 (a) shows that the slope of this particular turbine

characteristic curves WH (x, y) change with coordinates x when wicket-gate opening y is

small (less than 0.5). This feature requires a special treatment of interpolation for the

large gaps of WH values between the wicket-gate openings y; in Wylie’s simulation,

linear interpolation is used twice each time, between the openings y first and then

between the coordinates x; and the values of WH at certain range of small openings and

small coordinates are replaced by the values of coordinates x when they are interpolated

between the openings y. While in this developed model, the concept of shape function is

implemented for interpolation of WH between the openings y and coordinates x, which

causes the numerical instability in the range of small openings for this particular example.

However, this feature of WH curves is not necessarily found, or very rarely found, in

many other Francis turbine characteristic curves. Figure 6.5 (b) shows a typical Francis

turbine’s WH curves generated by the program developed in Chapter 5, they keep a

steady slope with coordinates x though big gaps exist between different openings. So far,

168
this numerical instability has never been found in the simulations of actual hydro projects

using the model developed in this thesis. However, certainly the interpolation method

could be further improved, particularly for each individual hydro turbine characteristics

and especially for the range of small openings.

(a) Responses to Load Rejection (61.7- 0 MW)

1.4
y
1.2
Dimensionless Turbine

1 v
0.8
Variables

h
0.6

0.4 ALPH

0.2
BETA
0

-0.2
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 6.4 Comparisons with Wylie’s Simulations (con’d)

169
(b) Responses to Load Reduction (67.1- 44.8 MW)

1.2

Dimensionless Turbine Variables 1 y


v
0.8
h
0.6
ALPH

0.4 BETA

0.2

0
0 2 4 6 8 10 12 14 16 18 20 22 24 26
Time (s)

(c) Responses to Load Increase (61.7- 81.0 MW)

1.2
Dimensionless Turbine

1
y
0.8 v
Variables

0.6
h
ALPH
0.4 BETA

0.2

0
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 6.4 Comparisons with Wylie’s Simulations

170
(a) Wylie's_Francis_Turbine_Characteristic_Curve

10

7
y= 0.0 0.1 0.2 0.3
6
WH(x,y)

4
0.4
3
0.5
2

0
-9 -6 -3 0 3 6 9 12 15 18 21 24 27 30 33 36 39 42 45 48 51 54 57 60 63 66 69 72 75 78 81 84 87 90

x=arctg(v/a) [deg]

(b) MeS_Francis_Turbine_Characteristic_Curve

12
11
y=0.13
10
y=0.22
9 y=0.39
8 y=0.48
y=0.57
7
y=0.65
WH(x, y)

6 y=0.74
5 y=0.83
y=0.91
4
y=1.0
3
2
1
0
-1
-2
-9 -6 -3 0 3 6 9 12 15 18 21 24 27 30 33 36 39 42 45 48 51 54 57 60 63 66 69 72 75 78 81 84 87 90

X=arctg(v/a) [deg]

Figure 6.5 Comparison of Wylie’s turbine characteristics with other Francis turbine
characteristics

171
6.4.2. Comparison with Field Measurement

The MeS hydropower plant in Iran has eight 250 MW units, as shown in Figure 6.6,

every two-units sharing a common penstock (Pipe #1) and a common tailrace tunnel

(Pipe #6). The units are submerged 13 m below the tailwater to prevent cavitation, but

the submersion is still insufficient based on the research (Pejovic, et al. 2004). This plant

has been built and operated for several years. However, during the simultaneous load

rejection of two units, there is a risk of water column separation and subsequent rejoinder

in the draft tube, the event that could potentially cause serious damage on both the

machine and water conveyance system (Pejovic, et al. 2004). Therefore, field

investigation and numerical simulation were carried out in 2004 to find an effective way

to prevent the excessive low pressures occurring in the draft tube.

To further verify the numerical model developed in this chapter, one set of data

measured during the site test at this power plant is used to compare with the simulation

results. As shown in Figure 6.7, this is the case that one unit rejects its full load while

another unit is closed. The (blue) thinner lines in the figure represent measurement and

the (red) thicker lines are from simulation. The greater time period of measured pressure

fluctuation is caused by the air chamber effects of the void in the draft tube; the bigger

the void, the greater the wavelength. There are two pressure taps set at turbine

downstream, one is positioned at a cross-section of the draft tube cone (3.2 m below the

turbine spiral case); another is positioned at the sluice gate where is near the draft tube

172
exit. The calculated pressure surges at the inlet of spiral case and outlet of draft tube are

initially calibrated to agree with the measurement. The calculated pressure at the

measured section of draft tube cone is retrieved based on the head balance equation from

the simulated pressure at the draft tube outlet (sluice gate) by considering the vertical

variations of both elevation and diameter. As expected, the simulated minimum pressure

values at the draft tube cone are lower than the measured ones for both initial steady and

transient states, because the role of vortex core void cannot be mathematically simulated

by one-dimensional model and thus the simulated pressure represents the average value

on the cross-section; despite this, the simulation agrees quite well with measurement

values.

173
368-372 m

Unit #3
Pipe #2 222-235m
Pipe #1 Pipe #4

Pipe #6
Pipe # 3 Pipe #5
Unit #4

Figure 6.6 Schematic waterway of two parallel turbine units in MeS plant

Table 6-2 System data of MeS Hydro Project

Pipe #1 Pipe #2/3 Pipe #4/5 Pipe #6 Turbine Unit Governor

L = 255.4 m L = 73.6 m L = 76.6 m L = 393.9 m HR = 142 m Td = 3.7 s

D = 9.5 m D = 5.81 m D = 7.315 m D = 11.0 m NR=187.5 rpm Tα = 0.325 s

a = 1200 m/s a = 1200 m/s a = 600 m/s a = 600 m/s QR = 191.5 m3/s Tf = 20 s

f = 0.03348 f = 0.02485 f = 0.01454 f = 0.02264 PR = 250 MW δ = 0.18

WRg2=7.1*107 N.m2 σ = 0.0

174
(a) Turbine Speed

160 90

150 80
70
140
60

Opening (%)
Speed (%)

130 50
120 40

110 30
WG Opening 20
100
10
90 0
80 -10
50 60 70 80 90 100 110
Time (s)

(b) Penstock Pressure

184
182
Penstock Pressure (m)

180
178
176
174
172
170
168
166
164
162
50 60 70 80 90 100 110
Time (s)

Figure 6.7 Comparison of numerical modeling and field measurement (con’d)

(Note: measurement; simulation)

175
(c) Pressure at Sluice Gate

22

Flap Gate Pressure (m)


20
18
16
14
12
10
8
6
4
2
0
50 60 70 80 90 100 110
Time (s)

(d) Pressure at Draft Tube Cone

22
Draft Tube Cone Pressure

20
18
16
14
12
(m)

10
8
6
4
2
0
-2
-4
50 60 70 80 90 100 110
Time (s)

Figure 6.7 Comparison of numerical modeling and field measurement

(Note: measurement; simulation)

176
6.5 Summary

This chapter demonstrates an application of AHD model to a system containing a hydro

turbine. A transient analysis model, including the head balance equation derived from

extended MOC equations, the speed change equation (i.e., the torque balance equation of

turbine shaft) and the governor equation (i.e., the relationship of PID speed controller

and corrective actions), has been developed for hydro turbine units. The numerical

solution and computer program have been coded to facilitate model implementation in

either pipe networks or conventional hydropower plants.

The applications of three basic equations for both isolated and interconnected

units, and for different operating scenarios are discussed in detail, and the computer

program and its input and output data are introduced as well for advanced TransAM

users.

Finally, the model and program have been verified by both Wylie’s simulation

example and a field test. Next, the developed turbine model and program are applied in

the new case studies presented in the next two chapters.

177
Chapter 7 Case Study 1: Energy Recovery Hydro
Turbines in Las Vegas Water Supply System

This chapter presents one application of the developed turbine model for the transient

analysis of small hydro turbines installed in Las Vegas water supply system for the

purpose of energy recovery in the place of the more traditional pressure reducing valves.

The transient analysis was carried between 2002 and 2004 during the feasibility study

and design stages of the project. The project has now been put in operation since 2006,

and field tests of load rejections were performed during system commissioning; field

data were kindly provided by the design engineer, Kinetrics Inc. Given that the original

purpose of the analysis was to predict transient pressures conservatively, and that the

results presented here are essentially “as is” without subsequent calibration, and taking

into account the complexity of the network system, the agreement shown in this chapter

between the numerical simulation and field test measurement for turbine transient

responses on load rejections is considered acceptable. Overall, this study provides

some additional verification of the developed turbine model, and to provide an indication

of the accuracy that might be expected when predicting transients in topologically

complex systems with active hydraulic devices.

This chapter also explores a novel application, specifically that of the effect of

178
fire-fighting flow on the operation of turbine unit. Such a flow disturbance would

obviously not be considered normal in the context of a more typical turbine application.

7.1 Project Background and Objectives

Small or micro hydro turbines can be applied not only in dedicated hydroelectric plants

but also in multiple-use systems and for power recovery purposes. In fact, installation of

micro-turbines in many water conveyance and irrigation systems is gradually becoming a

popular alternative to the wasteful dissipation of energy in devices such as pressure

reducing valves and head loss chambers. The use of small or micro turbines in zones with

a significant topographic difference not only produces renewable energy, but also can

avoid the use of stronger (and thus more expensive) downstream pipes and may even

reduce downstream water leakage. Moreover, extensive (and sometimes oversized)

storage tanks in urban water supply systems may have a valuable hydraulic potential first

for energy storage and then for subsequent electricity production during periods of high

electrical demand, thus allowing the municipal system to sometimes function somewhat

as a pumped storage scheme.

An important issue in any turbine installation is that of system hydraulics. In

particular, waterhammer in systems with turbines can cause numerous problems,

especially in systems with long pressurized pipelines. These concerns originate largely

from the complexity of turbine operation whose performance is influenced by the


179
operation of both the governor and the generator, from the properties of the water

conveyance network, and from the properties of the power grid connected to the unit.

Yet, even allowing for this complexity, at least a conservative estimate is required in

prediction and calculation of transient performance to avoid waterhammer problems and

associated system failures.

To take advantage of the flow rates (to feed the urban water distribution system)

and pressure differences (mostly from topographical difference) traditionally dissipated

by pressure reducing valves, four vertical Francis turbines were proposed to be installed

individually at Horizon Ridge, Bermuda, New Sloan (Linden) and Sloan Forebay ROFC

(Rate of Flow Control) stations in the Las Vegas water supply system. The system

hydraulics associated with the proposed turbine installations becomes a new issue for the

water supply system. Because of the complexity of long pipeline profiles, multiple

hydraulic devices and pipe connections in the system, the most vulnerable points to

waterhammer are not necessarily at upstream and downstream nodes of turbine (thus not

at the points which have typically been a focus of waterhammer studies in conventional

hydropower stations).

The developed turbine model was applied to analyze the proposed turbines and

associated pipe systems. The main objectives were to identify and predict the most severe

transient pressures in the pipeline under all possible conditions related to turbine

operations; and to identify possible mitigation techniques that could be implemented to

180
alleviate transient pressures.

Based on the turbine model characteristics provided by the manufacturer and

associated water system data converted from a supplied EPANET data file, the full load

rejection was identified as the most severe transient event. The corresponding numerical

simulation results provided designers with conservative estimates of the required pipe

pressure ratings and the minimum pipe internal pressures. Finally the key points in the

system to which more attention should be paid during design and operation were

identified.

7.2 System and Hydraulic Characteristics

The scheme summarized in Figure 7.1 shows the main constituents in one water district

of the water supply system. The Zone L1 is a part of water distribution network. The peak

daily water demand of this zone (approximately 22,000 gpm of flow) is supplied through

Linden ROFC station (Rate of Flow Control) from Pump Station A (shortened as PS A) to

the Reservoir S and/or through the 42-inch mains from the Reservoir S to the PS B.

During the summer months, some pumps at the PS B also can feed water to the zone L1,

typically pumping 13,500 gpm while 16,000 gpm is being taken from the ROFC station.

Because of the topographic difference between PS A and the distribution zone L1 the

pressure head across the ROFC can range from 165 to 240 feet, which was originally

dissipated by pressure reducing valves. To take advantage of this pressure head and the

flows that feed the downstream network, a Francis turbine manufactured by CHC

181
(Canadian Hydro Components Ltd.) has been installed within one valve train of ROFC

vault. The parameters of the turbine unit are summarized in Table 7-1.

Table 7-1 Linden turbine-generator parameters

Dth N11 Q11 NR QR HR Turbine Turbine Generator

(mm) (rpm) (cms) (rpm) (cms) (m) Efficiency Output Nameplate

(%) (kW) (kW)

385 57.8 0.764 1200 0.92 65.5 91.4 558 522

(Lower-elevation Distribution Zone L1)


Reservoir S
42’’ Mains (41,200’ long)

Hydrant 30’’(68’ long)

36’’ (80’ long)


Feed to
2 Spherical Proposed
Zone L2
Surge Tanks
Turbine
in ROFC
96’’(80’ long)

PS A Discharge
Pipeline 78’’
PS A
PS B

(Higher-elevation Water Supplier)

Figure 7.1 Linden Energy Recovery Hydro Turbine and System Constituents

182
The total length of mains is approximately 41,200 ft, including the 11,000 ft

portion of the line from Reservoir S to the ROFC and the 30,200 ft line between the

ROFC and PS B. The inner diameter of most of these pipes is 42 inches, and some are 60

inches. The friction factor for the Hazen-William’s formula was taken as 135

(equivalent Darcy-Weisbach friction factor f = 0.012), and the wave speed is 3500 ft/s; a

sensitivity analysis was performed and demonstrated that the transient response was only

mildly dependent on these specific values.

7.3 Transient Scenarios and Worst Case Identification

There are three potentially unfavorable situations that require special attention when

turbine units are installed and operated in a water supply system.

1) The first and most severe situation is that any control valve or the wicket-gate of

the turbine closes essentially instantly because of an extreme accident, such as the

sudden failure of the control ring or some other event not necessarily related to

turbine load rejection. This situation is termed “Instantaneous Closure” since the

flow at the turbine would be cut-off to zero in a negligible time. Not surprisingly,

the numerical simulation of “Instantaneous Closure” shows it would cause severe

negative and positive pressures within the connected pipe system.

However, this event can be considered highly unlikely for this application; in fact,

considerable effort is made to prevent such an occurrence. Specifically, through

183
careful design of the wicket-gate control ring actuator, this event is routinely made

extremely improbable. Protecting against such an event would be expensive and

would almost certainly include surge protection equipment what may never

subsequently be used.

2) The second, and more realistic, unfavorable transient event is caused by turbine

load rejection. The load rejection of a low- or medium-specific-speed Francis

turbine causes a rapid flow reduction or choking as the turbine accelerates to

runaway speed, and thus causes the receiving and supplying distribution system to

go through a phase of transient re-adjustment, as conditions adjust to the new flow

through the turbine. When load is rejected from large turbine-generator units, the

wicket-gate is usually governed to close gradually after load rejection, thus

reducing the time period (and intensity) of runaway speed (though the wicket-gate

closure can cause additional pressure surges in the system). For large machines that

cannot tolerate excessive runaway speeds, the calculation of wicket gate responses

is indispensable, as is the prediction of both the associated waterhammer and the

unit speed-rise.

However, the small turbine-generators considered in this project can be designed to

tolerate full runaway speeds for relatively long time periods (e.g., 60 minutes), and

thus the pressure surges in the system associated with only the load rejection itself

become the main transient event of concern. The 600-second (10-minute) time

184
period for the wicket-gate closure planned for this project is effectively unbounded

relative to the time constants of the hydraulic system, and thus the results are quite

similar to the full runaway situation. In other words, a controlled load rejection

(with wicket gate closure) and a full runaway condition (without wicket gate

closure) are effectively equivalent in this system. Thus, in summary, analysis and

prediction of the transient response associated with a load rejection and full

runaway condition is the major objective.

3) If there are hydrants nearby the turbine unit, either upstream or downstream, could

the in emergency use have significant impact on operation of the turbine unit?

How should the turbine unit be operated properly as a protection if this happens?

In addition, water column separation may become a concern because of the long

pressured downstream pipelines in a water distribution system. These topics, as one

of special issues for energy recovery turbines in water supply system, are addressed

as well in this chapter.

7.4 Simulation and Field Tests on Turbine Load Rejections

7.4.1 Field Tests

The load rejection tests were carried out on site during the commissioning of the turbine

systems; these tests permit an evaluation of the turbine’s transient performance (e.g., the

185
max. speed rise, the transient pressures, the turbine unit stresses and etc.). However, the

tests were no originally intended to verify the transient analysis model, and indeed the

data records are somewhat lacking for model verification. At the Linden

turbine-generator unit, the load was rejected at 50% and 100% wicket gate positions.

During the field tests, almost full flow is in the parallel piping path (train #1). The data

were recorded in steps of 1 or 2 (or 5) seconds. Since the rate of wicket gate closure is

at 6% per minute (full closure in 15 minutes), there is a long period of turbine runaway

condition with slowly decreasing flow before the system becomes to a final steady state.

7.4.2 Simulation vs. Field Test

Figure 7.2 shows the transient response of turbine parameters when the load is rejected

initially from the rated operation condition (100% wicket gate opening) and with

wicket-gate closure in 900 s. As expected, the transient responses are mild. This mild

response is achieved in several ways: (1) a flow of 0.924 m3/s (almost half of total system

flow) is continuously passed through the paralleled pressure reducing valve and bypass

piping line down to the main pipelines; and (2) the wicket gate closes extremely slowly,

so the discharge reduction is basically from the speed rise of the turbine at the initial stage

rather than from wicket gate closure, and the speed rise restricts the turbine flow from

initially 0.97 m3/s to 0.75 m3/s in the first 5 seconds.

In this system, the main pipelines are at the turbine downstream, so the

186
simulation of associated downstream pressure drop and subsequently pressure rise from

wave reflection are the major interest of the project. The comparison of turbine flow –

Figure 7.2 (a), speed rise – Figure 7.2 (b), upstream pressure rise – Figure 7.2 (c), and

downstream pressure drop – Figure 7.2 (d), shows a reasonable agreement between the

field test measurement and numerical simulation. Obviously, though, the simulation

results are generally somewhat more conservative, as indeed they were intended to be.

For instance, the maximum speed rise in Figure 7.2 (b) is 2070 rpm from simulation

while it is 2060 rpm from field test (the only two peak speed data recorded); the

simulated pressure at the turbine downstream in Figure 7.2 (d) shows the lowest pressure

drop to 68.6 m (more conservative peak pressure drop) and a slower transient energy

dissipation process (showing two pressure drop waves), while it has only one pressure

wave (demonstrating a fast transient energy dissipation process) with the lowest drop to

70.5 m from field test.

The observed discrepancies almost certainly arise because, in a more realistic

pipeline system, there are many more energy dissipation mechanisms than could be

captured in the numerical model. For example, small branch pipes are neglected in the

model yet they enhance energy dissipation in the system; this also holds for leaks which

cause water and energy losses from the distribution system; additional friction headlosses

occur at the pipe junctions, and many other mechanisms. The efficient dissipation

associated with waterhammer in complex networks has been commonly observed in other

187
studies.

The upstream pressure seems to rise gradually; this is because Pump Station A is

continuously supplying the same amount flow to the system, which results in the

accumulation of water in the double spherical surge tank, a process that occurs as the

turbine flow is gradually reduced.

An alternate reason for the noted difference between the simulation and field

measurement might be the complexity of flow conditions inside the turbine water passage

(three-dimensional and two-phases) or from incorrect input values of system parameters.

This difference would include, but not necessarily be limited to, variations in wave

speeds, pipe friction factors, device settings and system demands. It is challenging to

obtain the accurate values of these parameters from a realistic system, because many

factors could influence these parameters. For instance, a small amount of air in the

pipeline could significantly affect the value of wave speed and thus change the pressure

wave patterns and dissipation rate. A careful calibration of system parameters would be

required to minimize the differences between numerical simulation and field

measurement. However, from the perspective of engineering design practice, it is most

important to conservatively predict the peak pressure heads in the system and the peak

speed rise of the turbine unit. For this purpose, the differences shown in Figure 7.2 are

acceptable.

188
Figure 7.3 shows the transient pressure responses in the associated downstream

mains when the turbine unit experiences the full load rejection (i.e., the 100% initial

wicket gate opening). The maximum HGL are well below the pipe ratings; and the

minimum HGLs are generally above the pipeline profiles. Thus, no additional measures

are required for transient mitigation and control.

Figure 7.4 shows a comparison of numerical simulation and test data for the

transient responses of turbine parameters to the 50% load rejection (i.e., at the 50% of

initial wicket gate opening). The same trends and agreement are demonstrated as in

Figure 7.2 (100% load rejection case), except the transient response is milder for this

case.

189
(a) Turbine Flow Reduction During 100% Load Rejection

1.0

0.9

0.8
Turbine Flow (m3/s)

0.7

0.6

0.5
From Field Test
0.4
From Simulation
0.3

0.2

0.1

0.0
0 5 10 15 20 25 30 35 40 45 50
Tim e (s)

(b) Turbine Speed Rise During 100% Load Rejection

2100

1900
Turbine Speed Rise (rpm)

1700

From Field Test


1500 From Simulation

1300

1100
0 5 10 15 20 25 30 35 40 45 50
Tim e (s)

Figure 7.2 Turbine Transient Responses to 100% Load Rejection (con’d)

190
(c) Turbine Upstream Pressure Rise During 100% Load Rejection

132
Turbine Upstream Pressure Head (m)

130

128

126
From Field Test
124
From Simulation

122

120
0 5 10 15 20 25 30 35 40 45 50

Tim e (s)

(d) Turbine Downstream Pressure Variation


During 100% Load Rejection

80
Turbine Downstream Pressure Head (m)

76

72

From Field Test


68
From Simulation

64

60
0 5 10 15 20 25 30 35 40 45 50
Tim e (s)

Figure 7.2 Turbine Transient Responses to 100% Load Rejection

191
Pipe Ratings

Max. H
S.S. H

Min. H

Pipe Profile

(a) Distance from ROFC (Turbine) to PS B

Pipe Ratings

Max. H

Min. H

Pipe Profile

(b) Distance from ROFC (Turbine) to Reservoir S

Figure 7.3 Main Pipeline Pressure Responses to 100% Load Rejection

192
(a) Turbine Flow Reduction During 50% Load Rejection

1.0

0.9
From Field Test
0.8
From Simulation
Turbine Flow (m3/s)

0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85
Time (s)

(b) Turbine Speed Rise During 50% Load Rejection

2000

1800
Turbine Speed Rise (rpm)

1600 From Field Test

From Simulation
1400

1200

1000
0 10 20 30 40 50 60 70 80 90 100
Tim e (s)

Figure 7.4 Turbine Transient Responses to 50% Load Rejection (con’d)

193
(c) Turbine Upstream Pressure Rise During 50% Load Rejection

138
Turbine Upstream Pressure Head (m)
136

134

132

130

128

126 From Field Test


From Simulation
124

122

120
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85
Tim e (s)

(d) Turbine Dow nstream Pressure Variation


During 50% Load Rejection

80
Turbine Downstream Pressure Head

76

72
(m)

From Field Test


68
From Simulation
64

60
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85
Tim e (s)

Figure 7.4 Turbine Transient Responses to 50% Load Rejection

194
Figure 7.5 shows a simulation result for one main pipeline of SNWA water

supply system due to the load rejection of hydro turbine unit installed at New Sloan

Forebay (the detailed description and data of this turbine system are not provided here).

The maximum pressure rise at the pipe sections around the lowest position of the pipeline

(Las Vegas Wash) did exceed the existing pipe rating. So, the topographic variations in

the water supply system may create transient concerns due to the installation of hydro

turbine units.

Pipe Ratings

Max. HGL

S.S. HGL

Min. HGL

Pipe Profile

Las Vegas Wash

Distance from RMR to SF Turbine along 78'' New EVL Pipeline (ft)

Figure 7.5 Main Pipeline Pressure Responses to Load Rejection of Sloan Forebay
Turbine Unit

195
7.5 Effect of Hydrant Flow on Turbine Unit Operation

Hydrants are indispensable devices in water supply and distribution systems. Depending

on the population, the size and layout of a water network, the required fire flow rate

ranges from 1000 gpm to 3500 gpm (i.e., 65-220 l/s), the design duration 4-10 hours.

The operation of hydrant and water demand for the fire fighting can cause significant

transient surges in the system. If the hydrants are located close to turbine units, they

could induce undesirable impacts on the turbine operation. In this section, the effect of

fire fighting flow on the operation of turbine unit is explored; and the behavioral

responses of the turbine to the hydrant demands are graphically shown in the following

case studies with the hypothetical hydrants nearby the turbine units.

7.5.1. Upstream Hydrants

In the system sketched in Figure 7.6, since the downstream end of the turbine at Horizon

ROFC station is directly linked to a water tank, there is no possibility of hydrants

existing in its downstream pipeline. One to ten (intentionally extreme) hydrants are

assumed to locate at 2000 ft upstream of the turbine, where is marked as a star in Figure

7.6. The total flow in that pipeline is around 190 mgd (7.4 m3/s), and the turbine rated

flow is 33.75 mgd (1.48 m3/s). Initially, the turbine unit is operating at the rated condition

either isolated or connected to a large electricity system, and the hydrant flow is suddenly

initiated. The fire fighting demand would initially induce a rapid flow increase in the

196
associated pipeline and turbine, and thus result in a subsequent pressure drop at the

turbine, as well as a series of turbine behavioral variations, including the changes of gate

opening (automatically regulated by the governor) and rotational speed for an isolated

turbine unit.

Figure 7.7 shows the responses of an isolated operating turbine unit to 50 mgd

(2.2 m3/s) of upstream fire flow, which is equivalent to the extreme case of 10 hydrants

operating instantaneously (opened within 1 s). In this and the following figures, the

dimensionless turbine parameters v, h, ALPH (α), BETA (β) are defined as in equation

(5-10), Chapter 5; and they are turbine’s relative flow, head, speed and torque,

respectively; and y is the dimensionless wicket gate opening. The downsurge arrives at

the turbine in 0.5 –1.5 s from the hydrants, and then the turbine upstream head drops 60

feet relative to the initial head. This reduction of available waterpower causes the turbine

to rotate at a lower speed (α). Meanwhile, to balance its power load (resistance on the

turbine shaft), the turbine governor tries to open the wicket gates wider to pass more flow.

However, in this case even at full gate opening the available waterpower is still not

sufficient to meet the requirement of power load, so the rotation of the unit has to slow

down and finally would switch off automatically.

Figure 7.8 shows the responses of an online operating turbine unit to 50 mgd

(2.2 m3/s) of upstream fire flow. Since the online unit would transfer/distribute its power

load to the interconnected generator units, the unit speed and gate opening would keep at

197
their initial values. However, its flow (v) and dynamic torque (β) would keep at a lower

level because of the reduction in net head. Therefore, the turbine unit would finally reach

a new balanced condition with a lower power output.

Figures 7.9 and 7.10 demonstrate the turbine responses to single hydrant flow

(0.22 m3/s), the variation patterns of turbine behavior are the same as previously but the

variation ranges are naturally much milder, e.g., the speed of isolated turbine goes down

to 90% of the rated value at 16 s. This speed drop is much milder compared with the case

of 10 hydrants, in which the speed is lowered to 22% of the rated speed at 16 s. However,

in power system applications, to avoid tripping of underfrequency relays, the allowable

speed deviation of isolated turbine unit (i.e., the acceptable system frequency excursion)

is around 2-3% only (Gordon, 1985); so even a single hydrant flow would lead to the

stoppage of turbine unit operation.

Finally, attention should be paid to the pressure drop at the turbine downstream.

The sudden reduction of flow at the first initiation of hydrant operation would result in a

pressure drop at turbine downstream as well, which may cause the local pressure to drop

below atmospheric pressure for a short time period.

198
West-end SVL East-end Point
Warm Springs R. Bermuda ROFC
WL=2422’
& Turbine

P
T Horizon Ridge
108’’
ROFC & Turbine Burkholder R. &
84’’ 90’’
River Mountains PS
Horizon Ridge
R. WL=2365’ T
Hydrant
Black Mountain R.
WL=2220’

To 2300 Zone
South Valley R.
WL=2530’
To 2420 Zone

Parkway R.
WL=2495’
Bermuda R.
WL=2422.1’

Figure 7.6 Sketch of South Valley Lateral System

199
Upstream hydrant flow 2.2 cms
Isolated turbine unit

1. 4
Turbine dimensionless variables

1. 2
y
1 v

0. 8 h

ALPH
0. 6
BETA
0. 4

0. 2

0
0 2 4 6 8 10 12 14 16
Time (s)

Upstream hydrant flow 2.2 cms


Isolated turbine unit

200
Turbine upstream and downstream head

Upstream head
180
Downstream head
160
140
120
100
(ft)

80
60
40
20
0
-20
0 1 2 3 4 5
Time (s)

Figure 7.7 Isolated turbine responses to 10 upstream hydrants flow

200
Upstream hydrant flow 2.2 cms
Online turbine unit

Turbine dimensionless variables 1. 4


1. 2
1 y
v
0. 8
h
0. 6 ALPH

0. 4 BETA

0. 2
0
0 1 2 3 4 5
Time (s)

Upstream hydrant flow 2.2 cms


Online turbine unit

200
Turbine upstream and downstream head (ft)

180 Upstream Head


160 Downstream Head
140
120
100
80
60
40
20
0
-20
0 1 2 3 4 5
Time (s)

Figure 7.8 Online turbine responses to 10 upstream hydrants flow

201
Upstream hydrant flow 0.22 cms
Isolated turbine unit

1. 2
Turbine dimensionless variables

y
1. 1
v
h
1. 0
ALPH

0. 9 BETA

0. 8

0. 7
0 2 4 6 8 10 12 14 16 18 20
Time (s)

Upstream hydrant flow 0.22 cms


Isolated turbine unit

200
180
Turbine upstream and

160
downstream head (ft)

140
120
Upstream Head
100
Downstream Head
80
60
40
20
0
-20
0 1 2 3 4 5
Time (s)

Figure 7.9 Isolated turbine responses to single upstream hydrant flow

202
Upstream hydrant flow 0.22 cms
Online turbine unit

1.2
1.2
Turbine dimensionless variables

y
1.1
1.1 v

1.0 h
1.0 ALPH
0.9
BETA
0.9
0.8
0.8
0.7
0 1 2 3 4 5
Time (s)

Upstream hydrant flow 0.22 cms


Online turbine unit

200
180
160
Turbine upstream and
downstream head (ft)

140
120 Upstream Head
100 Downstream Head
80
60
40
20
0
-20
0 1 2 3 4 5
Time (s)

Figure 7.10 Online turbine responses to single upstream hydrant flow

203
7.5.2. Downstream Hydrants

In the system schemed in Figure 7.1 (Linden hydro turbine system), it is assumed that

hydrants are installed on the 42-inch mains at the turbine downstream, the location is

about 2900 ft of distance away from the turbine (see the star in Figure 7.1). The turbine

rated flow is 0.92 m3/s, which is also the total initial flow in the system for this case

study.

Figure 7.11 shows the responses of an isolated operating turbine unit to 1.0 m3/s

of downstream fire demand, which is equivalent to the case of 5 hydrants operating

instantaneously (opened within 1 s). Such a demand is unavailable in this system; thus,

then what would happen with this over demand? Since the location of the hydrant is not

far away from the turbine unit, this sudden flow is similar to a sudden opening of a valve,

which will cause an immediate pressure drop and subsequently a series of pressure waves

at turbine downstream pipeline. Even in this case of unrealistically over demand at

turbine downstream pipeline, the turbine upstream pressure is only slightly disturbed, and

the fluctuations in the turbine head would cause the variations in the turbine flow, speed

and wicket gate opening, which are gradually attenuated over time. The maximum speed

deviation is within the range of 82.5% to 115.5% of normal speed, and the main concern

is the negative pressure from the first down surge at the turbine downstream.

Figure 7.12 shows the responses of an online operating turbine unit to 1.0 m3/s of

204
downstream fire flow. Since the online unit would transfer/distribute its load to the

interconnected power system, so the turbine speed and wicket gate opening remain at

their initial values. The simulation shows that the turbine flow fluctuates with the turbine

head.

Figures 7.13 and 7.14 demonstrate the turbine responses to a smaller hydrant

flow (0.22 m3/s, equivalent of single typical hydrant); the variation patterns of turbine

behavior are similar to the previous 10 hydrants’ flow case, but the variations are much

milder. However, the transient waves would not be attenuated with time in this case

study; this might be a coincidence of resonance that depends on the characteristics of

pipelines.

The above discussions regarding the interaction between the hydrant flow and

turbine unit operation are based on certain hypothetical situations, particularly about the

speed change of isolated unit; the purpose is to reveal the potential significance of the

hydrant flow to the turbine operation. In any case, it is only prudent to place turbines as

far as possible from actively operating hydraulic devices to reduce the potential impacts

on the turbine operation. Under certain circumstance even the demand from a single

hydrant could cause instability of turbine operation.

205
Downstream hydrant flow 1.0 cms
Isolated turbine unit
3.5
Turbine dimensionless 3.0
2.5 y
variables

2.0 v

1.5 h

1.0 ALPH

0.5 BETA

0.0
-0.5
0 5 10 15 20 25 30 35 40 45 50
Time (s)

Downstream hydrant flow 1.0 cms


Isolated turbine unit
600
500
Turbine upstream and
downstream head (ft)

400
300
200
100
0 Upstream H
-100 Downstream H

-200
0 5 10 15 20 25 30 35 40 45 50
Time (s)

Figure 7.11 Isolated turbine responses to 5 downstream hydrants flow

206
Downstream hydrant flow 1.0 cms
Online turbine unit

Turbine dimensionless variables 3.0

2.5 y

2.0 v

1.5 h

1.0 ALPH

BETA
0.5

0.0
0 5 10 15 20 25 30 35 40 45 50

Time (s)

Downstream hydrant flow 1.0 cms


Online turbine unit

600
Turbine upstream and downstream

500
Upstream H
400
Downstream H
head (ft)

300

200

100

0
0 5 10 15 20 25 30 35 40 45 50
Time (s)

Figure 7.12 Online turbine responses to 5 downstream hydrants flow

207
Downstream hydrant flow 0.22 cms
Isolated turbine unit
2.5
Turbine dimensionless

2.0 y

v
variables

1.5
h

ALPH
1.0
BETA
0.5

0.0
0 10 20 30 40 50
Time (s)

Downstream hydrant flow 0.22 cms


Isolated turbine unit
600
Turbine upstream and

500
downstream head (ft)

Downstream H
Upstream H
400

300

200

100

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

Time (s)

Figure 7.13 Isolated turbine responses to single downstream hydrant flow

208
Downstream hydrant flow 0.22 cms
Online turbine unit

1.8
Turbine dimensionless

1.6 y
variables

v
1.4
h
1.2
ALPH
1 BETA

0.8

0.6
0 5 10 15 20 25 30 35 40 45 50
Time (s)

Downstream hydrant flow 0.22 cms


Online turbine unit
600
Turbine upstream and downstream

500

400
head (ft)

300

200
Upstream H
100 Downstream H

0
0 5 10 15 20 25 30 35 40 45 50
Time (s)

Figure 7.14 Online turbine responses to single downstream hydrant flow

209
7.6 Summary and Conclusions

This chapter presents an application of energy recovery hydro turbines in an urban water

supply system. The main points obtained from this project are summarized below:

(a) During load rejection of a turbine unit, the net pressure head at the lowest point of

upstream pipeline could exceed the existing pipe rating, and thus could lead to pipe

rapture and flooding. Therefore, it may be necessary to raise the pipe rating for

installations using hydro turbines.

(b) Compared with conventional hydropower projects, there are often more topographical

variations along the pipeline profiles in water supply systems. The pressure drop

caused by load rejection could be a concern at the highest position in the downstream

pipeline of the turbine unit, which may cause water column separation and subsequent

rejoinder.

(c) The potential interaction with active flow adjustments may cause serious interruption

of turbine unit operation, so the location of turbine installation should be isolated as

much as is reasonable from any sudden system disturbance.

(d) However, water storage tanks in the system may function as surge tanks, there is

seldom any waterhammer concern for the turbine installation near a water tank.

Appropriate sites for energy recovery turbine installation must clearly take into

210
account transient conditions. Topological complexity, such as associated with small

branched pipes in water networks, may dissipate transient energy and can provide

additional waterhammer protection.

Overall, there is considerable potential for energy recovery in urban water

supply and distribution systems, and hydro installations are often efficient ways of

generating clean and sustainable energy. Simulation tools, such as the developed in this

thesis, allow potential transient problems caused by turbine operation to be predicted and

avoided.

211
Chapter 8 Case Study 2: Transient Performance

Evaluation and Control Measures for Bone Creek

Hydropower Project

This chapter presents a case study of waterhammer analysis and control for the Bone

Creek hydroelectric power project in Western Canada. This is a typical hydropower

system with a long penstock, and is currently at design stage. Since there is no ideal

topographically and geologically feasible location for construction of open surge tank in

the proximity of the powerhouse, the surge tank was provisionally designed to be over

1000 m upstream of the powerhouse, connected by a branch pipeline bifurcated from the

penstock. Transient analysis shows the simultaneous full load rejection of two turbine

units would cause excessive high and low (negative) transient pressures in the penstock.

Therefore, various hydraulic system schemes are re-considered and simulated to seek for

a design plan with acceptable system transient performance; meanwhile, sensitivity

analyses of several important design parameters are performed to find the most

economical solution.

This chapter originated from the submitted consulting work for Bone Creek

Hydro Project; the thesis author carried out the numerical analyses and drafted the report

under the supervision of Prof. Karney.

212
8.1 Introduction to Bone Creek Project

8.1.1 System description and data

As sketched in Figure 8.1, this hydroelectric facility is proposed to have a 5957 m long

buried steel penstock from the intake to the powerhouse (Pipe No. 1 and 2). A

bifurcation (Node No. 2) is located 559 m upstream from the powerhouse (at the ground

elevation of 781 m) and a surge tower is tentatively planned at ground elevation of 873.5

m and distance of 578 m upstream from the bifurcation along a buried branch pipeline

(so-called surge pipeline, Pipe No. 3).

As shown in Figure 8.2, the originally proposed surge tower is a differential

surge tank with an internal 1.6 m diameter riser and a 10 m diameter external tank. The

top of the tank is at elevation (short as “EL”) 889.46 m, the top of the internal riser is at

EL 887 m, and the tank bottom is at EL 871 m. The 1.6 m diameter of riser is directly

connected with the sloped surge pipeline at EL 867 m below the tank bottom.

The upstream reservoir water level (short as “WL”) is at EL 882 m and

tailwater level is at EL 695 m; the centerline EL at the intake is 873 m. The system design

flow is 14.5 m3/s (for two units operating at full power load), and the expected head loss

at design flow is 41.4 m during normal operation. So, the steady state WL at the location

of surge tank is around 845 m, in somewhere of the surge pipeline as shown in Figure 8.2,

which is lower than the bottom of the surge tank.

Two turbine-generator units are designed for this facility, both of a double-runner

Francis turbine type directly connected to a generator through a horizontal shaft. Each

unit has a rated flow of 7.25 m3/s, a rated speed of 900 rpm and a rated power of 9,000

213
kW. The runner throat diameter is 0.7 m and the inertia of Turbine-Generator rotating
T

assembly is 10,700 kg.m2 per unit. The originally specified wicket-gate closure law is

shown in Figure 8.3 and entails a closure with first 70% closure in 6 seconds (T1 = 6 s)

and the last 30% over 8 seconds (T2 = 8 s).

Other dimensions around the powerhouse include the branched pipe (Pipe No.

4 or 5) to each turbine unit is 1.4 m in diameter, the center line EL at spiral case inlet is

694.52 m, the EL at draft tube exit and tailrace bottom is 691.52 m.

Figure 8.1 Planned system layout with differential surge tank

(Note: The drawing is directly copied from TransAM interface; the red digitals represent the Node
No./Identifiers and the black digitals represent the Pipe No./Identifiers)

214
Table 8-1 System data of Bone Creek Hydro Project

Pipe #1 Pipe #2 Pipe #3 Pipe #4-7 Turbine Unit Governor

L = 5398 m L = 559 m L = 578 m L = 10 m HR = 145.6 m Td = 3.7 s

D = 1.98 m D = 1.98 m D = 1.6 m D4-5 = 1.4.m NR= 900 rpm Tα = 0.325 s


D6-7 = 2.0 m
a = 1000 m/s a = 1000 m/s a = 1000 m/s a = 1000 m/s QR = 7.25 m3/s Tf = 20 s

f = 0.012 f = 0.012 f = 0.012 f = 0.012 PR = 9.0 MW δ = 0.18

WRg2=1.07*105N.m2 σ = 0.0

Figure 8.2 Sketch of planned differential surge tank

215
1

0.8
WG Opening

0.6

0.4

0.2

0
0 2 4 6 8 10 12 14 16

WG Closure Time T1+T2 (second)

Figure 8.3 Originally specified wicket-gate closure law

8.1.2 Objectives of Transient Analysis

A number of requirements for system transient performance need to be met under all

potential transient scenarios during turbine unit operations. In particular, the maximum

peak pressure rise (at spiral case inlet) needs to be no greater than 50% of the static head

(i.e., 280.5 m net head or 975.5 m EL) and the minimum pressure heads anywhere in the

system should be at least 2 m to avoid the possibility of negative pressures. The peak

head is expected around 880 m EL at the bifurcation and 882 m EL at the surge tank. The

maximum speed of turbine unit is not to exceed 150% (later relaxed to 160%) of normal

speed (i.e., 1350 rpm or 1440 rpm). The allowable runaway speed, given by the turbine

manufacturer, is 1570 rpm. Runaway speed occurs when the power load is rejected but

the turbine gates do not close due to system malfunction; and the maximum runaway

216
speed is also called full runaway speed, associated with full load rejection. The event of

simultaneous full load rejection of both units has been recognized as the most demanding

or unfavorable transient scenario during the operation, so the discussions in this case

study refer almost exclusively to this event. In addition, the requirements are interlinked

in that it is easier to meet the pressure rise constraint if the speed requirement is relaxed

through a slower closure of the wicket gates.

Transient analysis shows that the tentatively planned surge tank and system layout

would not automatically achieve all of the specified constraints, in particular, after full

load rejection with the originally specified wicket-gate closure (also shown in Figure 8.4).

The objective of the case study is to suggest possible changes in the system layout, the

surge tank scheme, or wicket-gate closure law that will control the transient response and

to allow the system to perform well under emergency conditions.

8.1.3 Simplifications and Assumptions Made in Transient Analysis

The representation of devices, system characteristics and constraints was made on the

basis of data provided and the standard hydraulic assumptions in modelling turbine

systems. The standard assumptions include valid algebraic head loss relations for friction,

constant wave speed, and dominantly one-dimensional (1-D) transient flow.

Because the actual turbine characteristic data was unavailable at the time of the

analysis (for confidentiality issues), the double-runner Francis turbine unit is modeled as

a single regular Francis turbine and the corresponding model characteristics for two

different-sized turbines from other projects are tested during simulation; the most

conservative simulation results turned out to be for the larger-sized of these two units,

and it is these results that are presented in this chapter.

217
The above approach may naturally cause some slight inaccuracy in the simulation

results; especially for the turbine runaway speed (since the model efficiency is more

sensitive than the head characteristics). However, the approach is believed to reasonably

reflect the system hydraulic features and to be able to capture the peak transient pressures

and speed rise caused by the load rejection of turbine units. Moreover, sensitivity studies

help to provide a control on the nature of the assumptions and their influence on the

modelling exercise.

8.2 Conventional Open Surge Tank Scheme

8.2.1 Transient Performance for Original System Design

Based on the planned system and differential surge tank scheme (Figures 8.1 and 8.2),

numerical simulations are performed using the developed simulation code for two units

simultaneously rejecting their full loads, and the results are summarized in this section.

As shown in Figure 8.4 (c), the turbine head goes up to 270 m maximum with the

flow cut-off by the originally specified wicket-gate closure law. As shown in Figure 8.4

(a) and (b), the maximum pressure head occurs at the inlet of turbine spiral case, nearly

965 m in EL; and the minimum pressure head is negative (about -22 m in theoretical

value), existing at upstream sections of the main penstock and thus cavitations would be

caused in the main pipeline. In addition, the peak pressure at the pipe bifurcation node is

around 920 m as shown in Figure 8.4(d); and the peak water level in the differential surge

tank is 884 m as shown in Figure 8.4(e).

218
As shown in Figure 8.4(f), the maximum speed rise is about 1290 rpm with

the originally specified wicket-gate closure, which is below the allowable maximum

speed 1350 rpm (or 1440 rpm).

Yet, the simulated full runaway speed is 1780 rpm, as shown in Figure 8.5,

which is higher than the specified allowable value or 1570 rpm given by the manufacturer.

However, interesting, the speed rise is below the allowable value if the turbine

characteristics (Hill Diagram) of a smaller size is used in simulation. Because the

runaway speed is mainly associated with turbine characteristics, it is suggested to consult

with the turbine manufacturer for the allowable runaway speed. In author’s view, there is

justification to down play concerns relating to runaway speed, particularly since this

scenario is highly unlikely in modern hydropower plants, equipped with advanced

automatic control and alarm system. With the appropriate mechanical design, the main

control valve at the penstock (butterfly or spherical valve) can be automatically closed if

the wicket-gates were to malfunction. The modern alarm system also ensures the

emergency reactions in the shortest time period, including possible manually intervention

of system closure in the emergency. One option that was briefly considered was to,

increase in the inertia of rotating parts; but this approach won’t significantly reduce the

runaway speed. This can be understood as follows: when a unit’s load is rejected, the

major resistance torque would be removed from the turbine shaft, but the dynamic torque

remains due to the water that is continuously arriving as the wicket-gate or the main

control valve fail to act to cut off the inflow. This huge unbalanced torque would not only

accelerate the turbine unit, but the runaway speed that would be achieved would be

largely dictated by the incoming flow.

219
The negative pressure predicted to occur at the upstream pipe sections is the major

concern for the current system layout. This negative pressure essentially results from

wave reflections from the upstream reservoir: a rapid high-pressure rise originates from

the downstream gate closure at the turbine unit, and this is turned into a negative wave

when it reflects off the upstream reservoir. Thus, the real issue is the significant inertia of

the water in the pipe between the powerhouse and the surge tank location. The 1157 m

long distance from the surge tank to the powerhouse, and the associated inertia of water

in it, reduces the ability of the originally proposed tank to mitigate transient events.

Moreover, because of the long distance of the surge tank and the associated inertia

of water it represented, the transient performance cannot be significantly improved by

increasing the size of surge tank only. If the original location of surge tank is retained, the

simulation shows increasing the surge tank diameter even from 10 m to 20 m while

lowering its bottom from 871 m EL to 860 m EL results in nearly identical peak upsurge

pressures and as well negative pressures. Figure 8.6 shows the envelope of maximum

and minimum pressures for this larger-sized tank, along the main pipeline, during the full

load rejection with the originally specified wicket-gate closure law.

220
(a) Envelope of max. and min. HGLs along main penstock

(b) Envelope of max. and min. HGLs from upstream reservoir to surge tank

Figure 8.4 System Responses during full load rejection with planned system layout

and design and originally specified wicket-gate closure law (T1=6 s, T2= 8 s) (Con’d)

221
(c) Pressure variations at spiral case inlet

(d) Pressure variations at bifurcation node

Figure 8.4 System Responses during full load rejection with planned system layout

and design and originally specified wicket-gate closure law (T1=6 s, T2= 8 s) (Con’d)

222
(e) Water level and flow at surge tank

(f) Turbine speed rise

Figure 8.4 System Responses during full load rejection with planned system layout

and design and originally specified wicket-gate closure law (T1=6 s, T2= 8 s)

223
Figure 8.5 Turbine speed rise during full load rejection with planned system layout

and design and wicket-gates refusal to close (runaway condition)

Figure 8.6 Envelope of max. and min. HGLs along main penstock during full load

rejection with 20 m diameter of surge tank and originally specified WG closure law

(T1=6 s, T2= 8 s)

224
8.2.2 Sensitivity Analysis of Surge Tank Distance

In order to better understand the influence of the distance between the differential surge

tank and the powerhouse, a sensitivity study was briefly undertaken. It is realized that

this location of the tank may be constrained by geologic and topographic conditions on

the site, but understanding this variable is key not only to the design of the surge tank but

to its possible substitution with a suitably sized air chamber.

Thus, the possibility is considered to install the surge tank closer to the

transient source (i.e., turbine unit) in order to reduce the peak pressure rise and eliminate

the negative pressures in the system. The question is how the reduced distance affects the

transient performance caused by load rejection. A sensitivity study is performed for this

project and the result is shown in Table 8-2 and Figure 8.7.

In Table 8-2 and Figure 8.7, the distance refers to the pipeline length L from

the powerhouse to the surge tank, that is, L = L2 + L3; and L2 = L3 is assumed in this study.

The maximum and minimum pressures (Hmax and Hmin) represent for the peak pressure

heads of upsurge and downsurge within the whole system following full load rejection.

The maximum unit speed is the peak turbine overspeed under the same scenario.

As shown in Table 8-2 and Figure 8.7, the distance L significantly affects the

transient response. In order to reach the desired transient performance during the full load

rejection, the surge tank would need to be installed within 200 m of the powerhouse;

further shortening this distance does not provide additional protection for the given

wicket gate closure law.

225
Table 8-2 Sensitivity of surge tank distance to transient performance

L (m) Hmax (m) Hmin (m) Max unit speed (rpm)


1157 270 -21.9 1290
600 219 -8.4 1225
400 199 -1 1197
200 184 3 1168
100 184 3 1157

Hmax (m) Hmin (m)

290 5

Minimum pressure head in system


Maximum pressure head in system

270 0

250
-5
230

(m)
(m)

-10
210
-15
190

170 -20

150 -25
0 200 400 600 800 1000 1200

Distance from Powerhouse to Surge Tank (m)

Figure 8.7 Sensitivity analysis of surge tank distance

226
As shown in Figure 8.8, the hydraulic pressures would be within the desired

range if the surge tank moved to the location within 200 m range from the powerhouse.

Yet, the full runaway speed would remain at 1780 rpm when the wicket gates refuse to

close (yet the flow is chocked by over-speeding runner), as shown in Figure 8.9. As

aforementioned, the speed rise at full runaway condition is mainly affected by the turbine

characteristics and turbine head; it is hardly reduced by changing system layout for this

particular high head turbine unit.

(a) Envelope of max. and min. HGLs along main penstock

Figure 8.8 System responses during full load rejection with a closer differential

surge tank and originally specified wicket-gate closure law (T1=6 s, T2= 8 s) (Con’d)

227
(b) Pressure variations at spiral case inlet

(c) Pressure at bifurcation node

Figure 8.8 System responses during full load rejection with a closer differential

surge tank and originally specified wicket-gate closure law (T1=6 s, T2= 8 s) (Con’d)

228
(d) Water level and flow at surge tank

(e) Turbine speed rise

Figure 8.8 System responses during full load rejection with a closer differential

surge tank and originally specified wicket-gate closure law (T1=6 s, T2= 8 s)

229
Figure 8.9 Turbine speed rise during full load rejection with a closer differential

surge tank and wicket-gate refusal to close (runaway condition)

8.2.3 Sensitivity Analysis of Wicket-gate Closure Time

Slowing down the closure of wicket-gates will reduce the peak pressures of original

upsurge at the turbine and thus largely eliminate the negative pressures reflected from the

upstream reservoir; however, this will also increase the maximum speed rise. However,

the transient simulations show that the maximum turbine speeds under the given closure

law of wicket-gates are well below the allowable maximum speed 1440 rpm (the relaxed

speed constraint), so it is possible to control both the transient pressures and turbine

speeds within desirable range by extending the wicket-gate closure time.

Due to the insensitivity of the second time period (T2) of wicket-gate closure,

T2 = 8 s is kept constant while changing T1 from 5 s to 15 s. The sensitivity study is based

230
on the original system layout with (1) the planned differential surge tank, (2) simple surge

tank without differential features, or (3) no surge tank in the place (certainly neither the

bifurcated pipeline). The simulation results are compared in Figure 8.10 for these three

scenarios.

The effects are only slightly different when using a simple surge tank or a

differential surge tank in this project. For both scenarios, it is found T1 needs to be

increased to 11 s so that the negative pressure can be eliminated and the speed rises are

still tolerable. The transient performance is shown in Figure 8.11 for the planned

differential surge tank and extended wicket-gate closure time T1 =11 s.

However, without a surge tank in the system, it is impossible to meet both

requirements of pressure head and speed rise during the transient by only adjusting the

speed of wicket-gate closure. Additional simulations show that further increasing T1 up

to 35 s (and T2 = 15 s), the maximum pressure head could be reduced to the desirable

value (278 m) with no negative pressures predicated in the system, but the speed would

rise to 1846 rpm. To reduce the maximum speed to below 1440 rpm, a flywheel with

inertia of 28000 kg.m2 would need to be added, which is infeasible due to its excessive

size and cost. The system transient performance under this scenario is shown in Figure

8.12.

231
20

-20
Hmin (m)
-40

-60

-80 Differential Surge Tank

-100 Simple Surge Tank


No Surge Tank
-120
5 6 7 8 9 10 11 12 13 14 15
WG Closure Tim e T1 (s)

(a) Min. pressure head in main pipeline vs. wicket-gate closure time T1

550

500

450
Hmax (m)

400
Differential Surge Tank
350
Simple Surge Tank

300 No Surge Tank

250

200
5 6 7 8 9 10 11 12 13 14 15
WG Closure Tim e T1 (s)

(b) Max. pressure head in system vs. wicket-gate closure time T1

Figure 8.10 Sensitivity analysis of wicket-gate closure time (Con’d)

232
1800

1700

Differential Surge Tank


Max Speed (rpm) 1600
Simple Surge Tank

1500 No Surge Tank

1400

1300

1200
5 6 7 8 9 10 11 12 13 14 15
WG Closure Tim e T1 (s)

(c) Max. speed rise of turbine unit vs. wicket-gate closure time T1

Figure 8.10 Sensitivity analysis of wicket-gate closure time

(a) Envelope of max. and min. HGLs along main penstock

Figure 8.11 System responses during full load rejection with designed differential
surge tank and extended wicket-gate closure law (T1 = 11 s, T2 = 8 s) (Con’d)

233
(b) Pressure variations at spiral case inlet

(c) Turbine speed rise

Figure 8.11 System responses during full load rejection with designed differential
surge tank and extended wicket-gate closure law (T1 = 11 s, T2 = 8 s)

234
(a) Envelope of maximum and minimum HGLs

(b) Turbine speed rise

Figure 8.12 System responses during full load rejection with extended wicket-gate
closure law (T1 = 35 s, T2 = 15 s) and flywheel (adding inertia 28000 kg.m2) for no
surge tank scenario

235
8.2.4 Sensitivity Analysis for Surge Tank Design Parameters

Sensitivity of diameter D of surge tank: For simple surge tank, the diameter change will

not cause an observable difference for either the maximum waterhammer pressure or the

overspeed, but the water level in the tanks does change with tank D and thus the tank

height could be reduced by increasing the tank diameter. The sensitivity analysis results

are shown in Table 8.3 and Figure 8.13.

With a differential surge tank, the diameter change will not cause an

observable difference for either the maximum waterhammer pressure or the overspeed,

and the variation of water level in the surge tank is also very slight. This is because the

current differential model is actually "a simple tank plus an overflow riser", and the

overflow water won't enter back into the big tank. This simplification maybe a concern

for small diameter of tank (e.g., D = 8 m or 6 m), the overflow would stabilize the water

level oscillations in the tank.

236
Table 8-3 Sensitivity Analysis of Simple Tank Diameter

Tank Dia Tank Cross- Turbine Over- Max. HGL HGL Tank Min.
D Section Area Max H speed at Spiral Inlet @ Tee WL @ Tank Height

(m) (m2) (m) (rpm) (m) (m) (m) (m)


10 78.54 226.6 1439 921.11 904.4 890.5 5.5
12 113.10 226.6 1439 921.11 904.4 889.0 4.0
14 153.94 226.6 1439 921.11 904.4 888.1 3.1
16 201.06 226.6 1439 921.11 904.4 888.4 2.4
18 254.47 226.6 1439 921.11 904.4 886.9 1.9
20 314.16 226.6 1439 921.11 904.4 886.6 1.6
Notes:
1. WG closure time T1=13 s, T2=18.33 s;
2. Feeder and connector pipe diameter d = 5 ft = 1.524 m, A=1.824 m2;
3. The resistance coefficient at the tank bottom orifice is defined as R = h f / Q 2 , and the
coefficient Rin = Rout = 0.005 is assumed;
4. Tank height needs to add the safety margin.

6.0

5.0
Net Height of Tank (m)

4.0

3.0

2.0

1.0

0.0
10 12 14 16 18 20

Tank Diameter (m)

Figure 8.13 Sensitivity of simple tank diameter D

237
Sensitivity of the connector pipe diameter Dc: When the connector pipe

diameter is reduced, the associated magnitude of resistance coefficient is increased as

square rate of the pipe/orifice area. Moreover, the resulting maximum waterhammer

pressure and overspeed do increase with the decrease of connector pipe diameter Dc. For

differential surge tank, when Dc reduces to 4 ft from current 5.25 ft, the maximum

pressure head increases from 212 to 216 m; moreover, overspeed increases from 1437 to

1440 rpm. Interestingly, the magnitude of resistance coefficient itself is insensitive to the

system transient performance, while the magnitude of connector/orifice area does result

in a changed transient performance. Various simulation results are summarized in Table

8-4 and Figure 8.14.

Table 8-4 Sensitivity Analysis of Feeder and Connector Pipe Diameter

Pipe Dia Pipe Dia Pipe Cross- Resistance Turbine HGL WL


D D Section Area Coefficient R Max. H Over-Speed @ Tee @ Tank

(ft) (m) (m2) (m) (rpm) (m) (m)


3 0.914 0.657 0.042 291.7 1549 974.7 888.0
4 1.219 1.167 0.013 251.5 1484 930.9 889.7
5 1.524 1.824 0.005 226.6 1439 904.4 890.5
6 1.829 2.627 0.002 210.9 1408 892.2 890.8
Notes:
1. WG closure time T1=13 s, T2=18.33 s;
2. Simple Surge Tank with D = 10 m;
3. The resistance coefficient at the tank bottom orifice is defined as R = h f / Q 2 (Rin =
Rout), which is inversely proportional to the square of connector pipe area, plus 10%
of additional head loss in the connector pipe;
4. The connector pipe has the same diameter as the feeder pipeline

238
891.0 300

Max.pressure head at power house


890.5 290
280
Water Level in Tank (m) 890.0
270
889.5
260

(m)
889.0 250
240
888.5
230
888.0
220
887.5 210

887.0 200
3.0 3.5 4.0 4.5 5.0 5.5 6.0
Pipe Diameter (feet)

WL @ tank (m) Turbine Hmax (m)

(a) Tank water level/max. pressure head at PH vs. pipe diameter Dc

1000 1560
980 1540

Turbine unit speed rise (rpm)


960
Pressure head at Tee (m)

1520
940
920 1500

900 1480
880 1460
860
1440
840
820 1420

800 1400
3.0 3.5 4.0 4.5 5.0 5.5 6.0
Pipe Diameter (feet)

HGL @ Tee (m) Speed Rise (rpm)

(b) Pressure at Tee and turbine speed rise vs. pipe diameter Dc

Figure 8.14 Sensitivity of feeder and connector pipe diameter Dc

239
8.3 Air-Cushioned Surge Tank Scheme

With the advance of underground engineering technology, more and more hydro projects

with high-head and long-headrace-tunnel are being constructed (e.g., most pump-storage

hydro projects). For this type of hydro project, the topographic condition often doesn’t

permit easy excavation of a large conventional open surge tank. Therefore, an innovation

in recent hydropower plant design is to replace the conventional open surge tank by an

air-cushioned chamber, especially when the geological condition is favorable and thus

there is no concern with air leakage or specific water treatment issues. In many cases, it

saves significantly on the initial investment compared with the conventional surge tank,

though the air pressure in the chamber needs to be well monitored and controlled during

operation, and a long time period may be necessary to compensate for the compressed air

subsequent to the system drainage and overhaul. In addition, the required stable cross-

sectional area is sometimes larger than that of conventional surge tank though the total

volume of chamber is less, so a long-hall-shaped tunnel with arched-cover is often

constructed. Such an arched chamber also helps stabilize the rock structurally while still

providing the required cross-sectional area. In other words, the detailed design of an air

surge tank must be optimized to maximize the economical and technical benefits (Bergh

Christensen 1982; Broch E. 1984, Gu Zhaoqi 1986, Zhang Jian et al. 2000, Liu Deyou et

al. 2000; Fan Boqin et al. 2005)

In the Bone Creek project, due to the topographical constraint on the

conventional open surge tank, the use of an air-cushioned surge tank near the powerhouse

was explored. As shown in Figure 8.15, an air-cushioned surge tank can be designed and

240
installed at the main penstock line as close as possible to the powerhouse (L1 = 5882 m,

L2 = 75 m, D1 = D2 = 1.98 m are assumed in the simulations), and the surge pipeline is

certainly eliminated. A simple cylindrical tank as sketched in Figure 8.16 is first

simulated; then a more realistic pipe-like air chamber is explored. As is typical in this

kind of system, the upper part is filled with compressed air gas and the lower part stores

the water that is exchanged with the pipeline as the pressure changes. Taking advantage

of air cushion allows the peak pressure of hydraulic surge to be suppressed during

transient events.

Figure 8.15 System layout with suggested underground air-cushioned surge tank

(Note: The drawing is directly copied from TransAM interface; the red digitals represent
the Node No./Identifiers and the black digitals represent the Pipe No./Identifiers)

8.3.1 Design Parameters and Input Data for Simple Cylinder Air Surge Tank

Figure 8.16 helps define the key input parameters for simulation of an air-cushioned

surge tank include:

1) The diameter of the tank D=10 m and its height h=12 m, although the key issue is that

the tank has a total volume of about 1000 cubic meters (air and water combined);

241
2) The elevation of tank bottom is 700 m;

3) The net hydraulic head loss resistance (R, defined as h f = RQ 2 ) for inflow and

outflow are assumed around 0.0001 s2/m5, but this would vary depending on the

relative size of connection to the main penstock line;

4) The local free atmosphere pressure = 9.55 m WC based on local elevation of the

project;

5) The polytropic exponent of the air gas, n = 1.6. Researches in China and Norway show

that n = 1.2-1.4 may not be sufficiently safe for prediction of the maximum gas

pressure and maximum water level in the tank (Zhang Jian, et al, 2004).

6) The initial water depth in the chamber Z0 = 4 m, so the height of air-chamber l0 = 8 m;

7) There are two ways to input the initial absolute air pressure P0 in the chamber. One

way is to input the number of P0 directly; another way is to input the static hydraulic

head (same as the WL at upstream reservoir 882 m) and then the software program

will automatically calculate the initial pressure P0 in the air chamber.

Figure 8.16 Sketch of simple cylinder air-cushioned surge tank

242
8.3.2 Transient Performance with Simple Cylinder Air Surge Tank

Based on the system model (Figure 8.15) and the above design of air-cushioned surge

tank (other system data are the same in Table 8-1), the transient event of full load

rejection is simulated, and the results are shown in Figure 8.17. The maximum peak

pressure at the spiralcase inlet is 960 m (Figure 8.17 (b)), and the maximum speed is

1142 rpm with normal closure of wicket-gate (Figure 8.17 (f)). Due to its better location

(shorter wave reflection time and reduced inertia of the intervening water), the concerns

with negative pressures are eliminated. Compared with the conventional surge tank

under the same condition, the peak of the pressure waves is reduced, though it is more

persistent. It is found that the transient waves have not declined in 600 s but they have

declined in 200 s for the corresponding case with open surge tank (Figures 8.4 and 8.8).

Another feature of air-cushioned tank is that the slope of pressure waves is related to the

water level, the peak pressure waves are sharper than the downsurge waves (Fan Boqin,

2005).

8.3.3 Sensitivity Analysis of Air Chamber Height

Table 8-5 and Figure 8.18 show the sensitivity of the height of air-cushioned surge tank

to the system transient performance while the initial water depth in the chamber, Z0 ,

keeps at 4 m; both the maximum turbine speed and maximum pressure head at the inlet of

spiralcase increase rapidly with the decrease of tank height. To control the peak pressure

below EL 975.5 m (i.e., the maximum head below 280.5 m), the minimum required

height of air-cushioned surge tank for this project would be between 11 to 12 m with as

per design described above.

243
(a) Envelope of max. and min.. HGLs along main penstock

(b) Pressure variations at spiralcase inlet


Figure 8.17 System responses during full load rejection with originally specified
wicket-gate closure (T1=6 s, T2= 8 s) and simple cylinder air surge tank (1000 m3 in
volume) (Con’d)

244
(c) Water level and flow at air-cushioned surge tank

(d) Nodal pressure head and air gas pressure head at air-cushioned surge tank
Figure 8.17 System responses during full load rejection with originally specified
wicket-gate closure (T1=6 s, T2= 8 s) and simple cylinder air surge tank (1000 m3 in
volume) (Con’d)

245
(e) Air volume in air-cushioned surge tank

(f) Turbine speed rise

Figure 8.17 System responses during full load rejection with originally specified
wicket-gate closure (T1=6 s, T2= 8 s) and simple cylinder air surge tank (1000 m3 in
volume)

246
Table 8-5 Sensitivity of cylinder air surge tank height to transient performance

Tank Height (m) Max. Speed (rpm) Max. Pressure Head (m)

8 1156 319

10 1146 284

12 1142 265

14 1140 252

16 1139 243

18 1138 237

350 1160

Max. Pressure Head


330 1155

Maximum Turbine Speed


Maximum Pressure Head

Max. Turbine Speed

310 1150

(rpm)
(m)

290 1145

270 1140

250 1135

230 1130
8 10 12 14 16 18

Air Chamber Height


(m)

Figure 8.18 Sensitivity analysis of the height of simple cylinder air surge tank

247
8.3.4 Transient Performance with a Pipe-like Air Surge Tank

From a construction and cost perspective, a more feasible plan for this system is to install

a pipe-like chamber, parallel to the penstock, starting from the bottom of the last steep

section near the powerhouse (75 m upstream from powerhouse and the tank bottom

elevation is 700 m). The air chamber pipe would use the same material and size as the

adjacent penstock (i.e., 1.98 m in diameter), and would be tentatively designed for 160 m

long sloping upwards at 16.1 degree (assuming 40% of volume is initially water-

occupied), so the total volume is around 500 m3 (a value that could be later optimized

through simulation). The pipe-like air chamber has hemispherical heads at each end and

an outlet pipe in the bottom end, and the outlet pipe diameter is about 1200 mm (or 1980

mm) with a 45-degree bend connecting to the penstock. An access manhole of 600 mm

and make-up compressed air supply will be set up at near the top of the air chamber.

If the originally specified wicket-gate closure law is retained, the above

dimension of air chamber is insufficient to control the system transient response within

the desirable range. As shown in Figure 8.19, there is no negative pressure head in the

system but the peak waterhammer pressure (1020 m EL) at the spiralcase inlet is

considered too high. Moreover, this high-pressure surge is hardly to be mediated by

extending the wicket-gate closure time. As shown in Figure 8.20, even if the closure time

T1 extends to 18 s (and T2 = 8 s), the peak pressure is again 1014 m EL, though the

overspeed 1402 rpm is acceptable for the relaxed criterion.

248
(a) Envelope of maximum and minimum HGLs along main penstock

(b) Water level and flow at air-cushioned surge tank

Figure 8.19 System responses to full load rejection with originally specified wicket-
gate closure (T1=6 s, T2= 8 s) and a pipe-like air surge tank (500 m3 in volume)

249
(a) Envelope of maximum and minimum HGLs along main penstock

(b) Turbine speed rise

Figure 8.20 System responses to full load rejection with extended wicket-gate
closure (T1= 18 s, T2= 8 s) and pipe-like air surge tank (500 m3 in volume)

250
(a) Envelope of max. and min. HGLs along main penstock

(b) Turbine speed rise

Figure 8.21 System responses to full load rejection with originally specified wicket-
gate closure (T1=6 s, T2= 8 s) and a greater size of air surge chamber (840 m3 in
volume)

251
However, as Figure 8.21 shows, the peak pressure is reduced to 974 m EL

when the volume of air chamber increases to 840 m3 (say either by increasing the pipe

diameter from 1.98 m to 2.6 m with a constant length of 160 m; by increasing pipe length

from 160 m to 280 m keeping the 1.98 m in diameter; or by using a pipe with 250 m

length with 2.07 m diameter).

Table 8-6 and Figure 8.22 show the peak waterhammer pressure at the inlet of

spiralcase reduces with the increase of the total length of air-chamber-pipe when using

the same diameter 1.98 m. All these analysis results are based on the originally specified

wicket-gate closure law and 40% of initial water level in the pipe-like air chamber.

Simulation also shows that if the total chamber volume and initial water level

are constant, there is no difference in the results when the pipe-like air chamber is

simulated as an actual slopped pipe-like chamber (with shorter but greater cross-section

due to the slopped angle) or as an equivalent vertical air chamber.

Table 8-6 Sensitivity of air-chamber-pipe length

Air-chamber-pipe
length L (m) Max. HGL (m)
160 1020
200 998
240 984
250 981
280 974

252
Peak Pressure Head vs. length of Air-chamber-pipe

1020

Peak Pressure Head EL (m) 1010

1000

990

980

970
150 170 190 210 230 250 270 290

length of Air-chamber-pipe L (m)

Figure 8.22 Sensitivity of length of air-chamber-pipe to peak pressure during full


load rejection with originally specified wicket-gate closure (T1=6 s, T2= 8s) and a
pipe-like air surge tank (500 m3 volume)

8.4 Summary and Conclusions

This chapter documents the search for feasible solutions to the excessive transient

pressures and speed rise associated with two units experiencing simultaneous full load

rejection. More specifically, systematic simulations and sensitivity analysis of important

system parameters were performed. The main conclusions are summarized as follows:

1. For the open surge tank scheme to be a viable way meeting all the imposed

constraints of head change and speed rise, the surge tank should be constructed

within 200 m distance from the powerhouse if both the planned surge tank scheme

and originally specified wicket-gate closure law were maintained. Unfortunately,

such a location is infeasible due to the topographic constraints.

253
2. The simulations show that increasing the size of the surge tank could not improve

system transient performance effectively and sufficiently if both the planned surge

tank location and wicket-gate closure law were remained.

3. However, more positively, adjusting the speed of the wicket-gate closure does

significantly reduce the peak of original upsurge and reduces/eliminates the extent

and duration of the possible negative pressure in the system. Although there is a

slight increase in the speed rise of the turbine unit, this speed rise is still tolerable.

To be more specific, a staged closure curve is quite feasible: if the first 70% of the

wicket-gate opening is closed linearly in 11 s, and the final 30% closed in 8 s, then

both the transient pressure and turbine speed rise can be kept within the desired range

for the current system layout and surge tank design. Moreover, the transient

response, both in terms of head rise and speed rise, is insensitive to the duration of

the second stage closure; variations of up to 50% have no visible influence on surge

conditions.

4. There is no significant difference between the effects of differential surge tank and

simple surge tank; but the transient response would be much worse if there is no

surge tank in the system.

5. For the case with no surge tank, the peak pressure rise can be constrained by

extending wicket-gate closure time to 35 s for the closure of the first 70% of the

wicket-gate opening; but the peak speed would be too high. Moreover, increasing the

inertia of Turbine-Generator rotating parts by adding a flywheel is an ineffective way

to reduce the peak speed rise to the desired range.

254
6. When the overspeed constraint is relaxed to 160% of normal speed (1440 rpm), the

wicket gate close time can be extended to 14 s for the first 70% of the full gate

opening and 18.7 s for the remaining 30% of opening, and thus the currently planned

system layout and open surge tank scheme (either simple or differential surge tank)

would be sufficient to meet the constraints of the transient pressures. This probably is

the most efficient way to eliminate the negative pressure and excessive high

pressures in the system during full load rejection.

7. Another effective way to improve the system’s transient performance is to replace

the differential surge tank with an air-cushioned surge tank nearby the powerhouse.

Such an air-cushioned surge tank could be installed underground nearby the

powerhouse provided the geological conditions are favorable, or could be configured

as a suitable length of pipe. To be effective in meeting the hydraulic constraints, a

proposed pipe-like air chamber needs approximately 840 m3 in volume (the pipe with

250 m length and 2.1 m in diameter is suggested). Such a dimension would be

associated with the originally specified wicket-gate closure law. Sensitivity studies

have shown that the exact shape for configuring the air chamber volume is of less

consequence. Moreover, as long as the connecting pipe is of adequate size, the

nature of the connecting pipe to the air chamber is of strictly secondary importance

compared to the volume of air provided. Moreover, given the fact that many of the

air chamber’s requirements are symmetrical in that they involve both load acceptance

and load rejection (upsurge and downsurge), there is little advantage to arranging a

significant variation between inflow and outflow orifice characteristics.

255
Chapter 9 Thesis Summary and Suggestions for System

Transient Performance Evaluation

This thesis focuses on the transient analysis associated with a general class of hydraulic

components, herein called active hydraulic devices (AHDs), and on several specific

realizations of these devices. AHDs are of strategic importance in the context of a

hydraulic system, as they are either the source of transient disturbances, or a critical

strategy for surge control and migration.

In particular, a new strategy using an AHD was explored which seems to

eliminate/mitigate wave reflection and potential resonance in pipelines in order to ensure

adequate transient control and protection. The application of AHDs entails a profound

effect on system behavior and, in turn, response to such devices is influenced by the

details of system design and investment as well as its operation and management. In fact,

it is emphasized that there is no ‘safe’ territory for AHDs and no simple way to achieve

better transient performance; rather, one needs to understand the nature of a system’s

response with or without AHDs. Moreover, AHDs play an important role not only in

transient management but also relative to overall system control and behavior.

Automatic control valves are the first type of active hydraulic devices studied

in this research. Such a valve could be either a cause of transient flow or a measure to

protect a system from its detrimental effects; it is often a key element in a water system

used to control the flow rate, reverse flow, pipe pressure or water level in a tank. Due to

256
the challenges inherent in the design of a pressure relief valve (PRV), the general

principles of PRV use and selection are studied along with the sensitivities of a system’s

response to the PRV parameters. Then, a numerical model for conventional PID

(proportional, integral and derivative) control valves is studied. Based on the model, a

new application of PID control valves is envisioned that combines a remote sensor

upstream in a pipeline to create a non- or semi- reflective boundary at the downstream

end. The case study shows that, at such a boundary, the valve is able to adjust its opening

automatically in response to the remotely sensed pressure changes in order to eliminate or

mediate the reflection and resonance of pressure waves within the conduit.

The second type of active hydraulic device studied in this research is the

governed hydro-turbine, the most complicated hydraulic component in terms of transient

analysis and waterhammer control. Although there are a number of numerical models

developed for analyzing transient flows cuased by turbine operations, to the author’s

knowledge, previously published studies were not directly applicable to employment in a

water network. In this thesis, a complete numerical model is developed for turbine

installations in either urban water networks for energy recovery purposes or

conventional/pumped-storage hydropower generation systems. Two case studies are

undertaken and presented for realistic hydro projects.

The numerical models developed in this research represent an enhanced ability

to simulate and even control a pipe system’s transient performance. It is essential to

evaluate the transient performance and control efficacy for a pipe system design and to

identify and discard the inappropriate design schemes based on suitable design criteria

(e.g., the acceptable range of max. and min. pressures).

257
9.1 Perspectives on System Transient Performance Evaluation

Transient flow is unavoidable in any dynamic hydraulic system, and threats to system

integrity not only arise from pressure surges but from potential resonance and auto-

oscillations initiated by transient pressure or flow fluctuations in poorly designed and

tested systems. Furthermore, even though less visible, excessive pressure drop

experienced during transient variations is also crucial. While pressure extremes are well

appreciated as a source of trouble in systems, difficulties persist from system overload

and even pipe burst is not rare if careful prediction and system design are neglected.

However, decreased pressure, especially during vacuum conditions (as discussed in the

case study of Chapter 8) can cause a series of detrimental phenomena, such as air release

from solution, water vaporization, water column separation and cavitation. A pipe could

collapse because of negative pressure or burst due to high pressure from water column

rejoinder following separation; air and contaminants could be drawn into a pipeline

through any breach of system integrity (transient intrusion) and water quality risks being

compromised by possible pollutant intrusion for municipal water supply systems; while

the ingress of air could in turn markedly degrade transient pressures.

Overall, transient control and protection has proven to be a complex task in

water systems, and there are many helpful techniques for controlling or influencing

system response. These can be categorized according to three basic approaches: 1) to

enhance the strength of the system or change the pipeline profile; 2) to control the source

of transients (such as avoiding check valve slam, moderating rapid valve opening); and 3)

to suppress the transient using various hydraulic devices (such as PRVs, bypass lines, air

valves, surge tanks, and air chambers). Some devices may require lower initial

258
investment but higher maintenance cost (or pose an elevated risk of malfunction) during

the operation stage. The question is how to combine these techniques to form an

integrated strategy for system transient control and protection? What trade-offs and

benefits are expected to arise from the application of these devices? Which design

scheme or solution is the more cost-effective and energy-efficient for a specific system or

particular issue?

Measures of waterhammer mitigation should be ultimately refined and

scrutinized by technical evaluations (maximum/minimum transient pressure, probability

and safety performance), evaluation at the economic level (cost-effectiveness, including

risk cost due to transient failure), and considerations at the environmental level (energy-

efficiency and greenhouse gas emissions).

Ultimately, as part of future work, the assessment of system transient

performance should not only be based on the traditional deterministic transient analysis

(using the numerical models developed in the thesis) but also on a stochastic model and

risk analysis, because randomness is the nature of transient flow due to the variability of

initial conditions (hydrological variations) and the uncertainties in system parameters and

variables. In fact, the impetus for a shift from the traditional deterministic approach to a

transient stochastic analysis in hydropower facilities is three-fold:

Firstly, there is a need for adequate hydraulic performance assessment; transient

performance should be evaluated in a more informative and comprehensive way based on

the spectrum of pressures that systems experience. Beyond just knowing the extreme

values of transient pressures, an estimate of the likelihood of their occurrence would be

enormously useful. This requires statistical characterization and simulation of probability

259
distributions representing transient pressures, rather than the more limited knowledge of

max./min. pressure obtained from deterministic analysis under worst-case scenarios.

Secondly, there is a structural design motivation. In recent decades, there has

been an obvious discrepancy between the popular reliability-based design approach and

traditional deterministic waterhammer evaluation. Excessive pressure unleashed by

transient events is one of the random loads endured by a hydraulic structure, and its

stochastic information is essential to the reliability-based design of such constructions, in

which the normative loads, partial load factors and safety factors are derived from the

probability characteristics of each load. Furthermore, there is an obvious discrepancy

between the current structural design methodology and traditional deterministic transient

load analysis. Indeed, the advanced hydraulic structural design based on reliability theory

calls for a stochastic analysis of transient flow and an analysis of the statistical features of

transient pressures in pipelines and other hydraulic facilities; however, the pressure

surges as loads used in hydraulic structural design are still analyzed using a deterministic

approach by calculating the maximum values under the worst case scenarios. Therefore,

the design guidelines or regulations about transient loads in pressure pipe systems need to

reflect the observations of stochastic transient and risk analysis. For this reason, statistical

data associated with transient analysis in conventional hydro systems was collected

during my previous studies in China (Zhang et al. 1997, 1998, 2000s).

Thirdly, stochastic transient analysis can provide useful insights and contribute

to life cycle risk, economic and environmental assessment studies. Life cycle assessment

is a systematic methodology for evaluating the environmental impacts of different design

options. The life cycle monetary and environmental expenses are associated with the risk

260
of system failure, from which no system is completely immune. Even for a well-designed

system with appropriate waterhammer provision, the probability of failure due to

transient events is never absent because of potentially wide hydrological variations,

component failure, possible nefarious actions and other inherent uncertainties. Expected

damages due to transient failure (i.e., the product of an incident’s risk of occurrence and

its corresponding damage) must be taken into account in the life cycle monetary cost and

environmental burden of a system. In fact, only when the probabilistic characteristics of

extreme pressures are obtained from a stochastic transient model can risk analysis be

performed and integrated within the life cycle analysis (LCA) framework for holistic

system assessment.

At the economic and environmental levels, LCA can be implemented to assess

the monetary cost and energy consumption of competitive alternatives during the

protracted service period. Some design schemes may require lower initial investment but

result in high maintenance and reparation costs during the operation stage (Zhang et al.,

2007). Reconsidering the case study of Bone Creek Hydro Project presented in Chapter 8,

various technically feasible alternatives, such as the air-cushioned chamber deployed near

to the powerhouse, or the turbine-generator units manufactured according to an upgraded-

standard that are able to withstand higher overspeed during load rejection (with extended

wicket-gate closure time), could be assembled to eliminate the negative pressures in the

penstock upstream sections and reduce excessive high pressure to within acceptable

range, but which alternative is the best solution? To answer this question, the life cycle

cost including the risk cost, as well as the energy used and greenhouse gases generated

through the lifespan of each alternative solution should be evaluated and compared.

261
References

Almeida A. B. de, Koelle E., 1992. Fluid Transients in Pipe Networks. Pg.111-112.

Co-published by Computational Mechanics Publications, Southampton Boston and

Elsevier Science Publishers Ltd., London New York, 1992.

Boulos P., Lansey K., and Karney B., 2006. Comprehensive Water Distribution Systems

Analysis Handbook for Engineers and Planners, Chapter 9, pg 9-9 to 9-11. MW

Soft, Inc., 2nd edition, 582 pp., 2006.

Bounce B. & Morelli L., 1999. Automatic control valve-induced transients in operative

pipe system, J. of Hydraulic Engineering, ASCE. Vol.125, No.5, pp543-542, 1999.

Broch E., 1984. Development of Unlined Pressure Shafts and Tunnels in Norway, J. of

Underground Space, 1984, 8(3):177-184.

Chaudhry M. Hanif, 1987. Applied Hydraulic Transients. Van Nostrand Reinhold

Company, 1987.

Christensen Bergh, 1982. Unlined Compressed Air Surge Chamber for 24 Atmospheres

Pressure at Jukla Power Station, Pressure Tunnels/Tunneling Machines/

Underground Storage, Vol. 2, No. 11, 1982, pp. 889-894.

Fan Boqin, Zhang Jian, Suo Lisheng, Liu Deyou, 2005. Stability analysis on small

fluctuation of the system with air-cushion surge chamber for hydropower station, J.

262
of Water Resources and Hydropower Engineering. Vol.36, No.7, 2005. (in Chinese)

Franzini Joseph, B., Finnemore E. John, 1997. Fluid Mechanics with Engineering

Application, 1997.

Gu Zhaoqi, 1986. Experiences in Norwegian Hydropower Engineers, TAPIR

PUBLISHERS Trondheim, Norway, 1986, pp. 55-82.

Gordon J. L., 1961. Determination of Generator Inertia, presented to Eastern Zone

Meeting Hydraulic Power Section, Canadian Electrical Association, Halifax,

January 1961.

Gordon J. L., Whitman D. H., 1985. Generator inertia for isolated hydropower

systems Canadian Journal of Civil Engineering 12(4): pp 814-820.

Hopkins John P., 1998. Valves for controlling pressure, flow and level., J. of Water

Engineering & Management; Aug. 1998; 135, 8; pg.42-44.

IEC 193, 1965. International Electrotechnical Commission: International code for model

acceptance test of hydraulic turbines. Publication 193, 1965.

Karney B.W. and McInnis D.A., 1992. Efficient calculation of transient flow in simple

pipe networks, J. of Hydr. Engrg., ASCE, 118(7), 1014-1030.

Karney B.W. and Simpson A.R., 2007. In-line Check Valves for Water Hammer Control,

Journal of Hydraulic Research, IAHR, 45:4:547.

263
Koelle E., 1992. Control valves inducing oscillatory flow in hydraulic networks,

International Conference on Unsteady Flow and Fluid Transients. Durham, United

Kingdom, 343-352.

Koelle E. & Poll H., 1992. Dynamic behavior of automatic control valves (ACV) in

hydraulic networks, 16th Symposium of the IAHR – Section on Hydraulic

Machinery and Cavitation, Sao Paulo, Brazil, 127-139.

Lauria J. & Koelle E., 1996. Model-based analysis of active PID-control of transient

flow in hydraulic networks, Hydraulic Machinery and Cavitation, 749-758, © 1996

Kluwer Academic Publishers. Printed in the Neitherlands.

Liu Deyou, Zhang Jian, Suo Lisheng, 2000. Advances in Research on Air-cushioned

Surge Chamber, International J. of Hydroelectric Energy, Vol.18, No.4, December

2000. (in Chinese)

Liu D., Wang F., et al., 2007. Field Measurement and Numerical Simulation of Hydraulic

Transients for Reversible Pump-Turbine Units with MGV Device.

COMPUTATIONAL MECHANICS. ISCM2007, July 30-August 1, 2007,

Beijing,China. ©2007 Tsinghua University Press & Springer.

Liu Q., et al., 1997. Hydropower Station, Hohai University Textbook, Version 3 (in

Chinese).

264
MNREG, India, 2008. Standards/Manuals/Guidelines for Small Hydro Development.

Sponsor: Ministry of New and Renewable Energy Government of India. Electro

Mechanical. Guide for selection of turbine and governing system for small

hydropower. Lead organization: Alternative Hydro Energy Center Indian Institute

of Technology, RooRkee. Draft May 19, 2008.

Pejović-Milić A., Pejović S, Karney B., 2003. Extreme Theoretical Pressure Oscillations

in Coronary Bypass, Proceedings of the International Conference on CSHS03,

Belgrade, 2003.

Pejovic S., Karney B., Zhang Q. (2004). Water Column Separation in Long Tailrace

Tunnel, Proceedings of HYDROTURBO 2004 international conference, Brno,

Czech Republic, Oct. 2004.

Pejovic S., Karney B., Zhang Q. and Kumar G. (2007). Small Hydro, Higher Risk, IEEE

Electrical Power Conference (EPC) 2007, Oct. 25-26, 2007 in Montreal, Canada.

Poll H., 2002. The importance of the unsteady friction term of the momentum equation

for hydraulic transients, Proceedings of Hydraulic Machinery and Systems, 21st

IAHR Symposium, Sept. 2002, Lausanne, Switzerland.

Ruus, E. & Karney, B.W., 1997. Applied Hydraulic Transients. Ruus Consulting Ltd.,

Ken Fench. British Columbia, Canada, 1997.

265
Suter P., 1966. Representation of Pumpcharacteristics for calculation of Water Hammer

Sulzer Technical Review. Research Number, 1966, (11):9-100.

Tan Kok Kiong, Wang Quing-Guo, Hang Chang Chieh with Tore J. Hagglund, 1999.

Advances in PID Control, Chapter 1. Springer-Verlag London Ltd. 1999.

Thorley, A.R.D. and Chaudry, A., 1966. Oumo Characteristics for Transient Flow

Analysis. Proc., 7th Int. Conf. On Pressure Surges and Fluid Transients in Pipelines

and Open Channels, Harrogate, UK, 16-18 April, 461-475.

Tong Jiandong, Zheng Naibo, Wang Xianhua, Hai Jing, Ding Huishen, 1997. Mini

Hydropower, Hangzhou Regional Centre for Small Hydro Power, Hangzhou, China,

JOHN WILEY & SONS.

Wylie E. B., Streeter V. L., with Suo L., 1993. Fluid Transients in Systems. Chapter 10,

Pg. 246. Prentice-Hall Inc., Englewood Cliffs, New Jersey.

Zaruba Josef, Vltava Basin, Prague, Czechoslovakia, 1993. Water Hammer in Pipe-line

Systems, P102-119.

Zhang Jian, Suo Lisheng, Liu Deyou, 2000. Deriving Hydraulically Stable Cross

Sectional Area of the Air-cushioned Surge Tank Based on Test Data of Throttled

Orifice, International J. of Hydroelectric Energy, Vol.18, No.3, Sept. 2000. (in

Chinese)

266
Zhang Jian, Suo Lisheng, Liu Deyou, 2001. Formulae for Calculating Surges in

Air-cushioned Surge Tank, J. of Hydroelectric Engineering, Vol. 19, No.1, 2001. (in

Chinese)

Zhang Jian, Suo Lisheng, Zheng Yuan, Liu Deyou, 2004. Study on gas characters of

air-cushioned surge tank, J. of Hydroelectric Engineering, Vol.23, No.4, Aug. 2004.

(in Chinese)

Zhang, Qinfen; Karney, Bryan; MacLean, Heather L. and Feng, Jingchun, 2007. Life

Cycle Inventory of Energy Use and GHG Emissions for Two Hydropower Projects

in China, J. of Infrastructure Systems, ASCE. Vol. 13, No. 4, p271-279, December

2007.

ZHANG Qinfen, SUO Lisheng, 2000. Application of Stochastic Analysis of

Water-hammer in Structural Design of Penstocks, J. of Hohai University (ISSN

1000-1980), 28(2), 17-22. (In Chinese)

ZHANG Qinfen, SUO Lisheng, GUO Wenzhu., 2000. A Further Study of Stochastic

Analysis of Water-hammer Pressure in a Reservoir-pipe-valve System, J. of

Hydroelectric Engineering (ISSN 1003-1243), 2, 56-63. (In Chinese)

ZHANG Qinfen, SUO Lisheng, 1998. Advances on Stochastic Analysis of Water-hammer

and Surge, J. of Advances in Science & Technology of Water Resources (ISSN

1006-7647), 18(3), 7-11. (In Chinese)

267
ZHANG Qinfen, SUO Lisheng, 1997. Stochastic Model for Transient Analysis in Surge

Tanks, J. of Hohai University (ISSN 1000-1980), 25(3), 41-45. (In Chinese)

268
APPENDICES:

AppendixⅠ: Turbine Characteristics Input Data Sample and Variable Index

Turbine Characteristics Input Data Sample:

| No. of turbines (NTURB)


1

| TURBINE-TYPE QR HR NR PR ETR Runner Dia.


| (cms) (m) (rpm) (KW) (%) (m)
'Francis' 191.5 142.0 187.5 250000 0.889 4.5

| No. of turbine wicket gate sets (Ndy)


10

| The full opening a00(mm)


23.0

| GATE:3mm
| No. Points (NPP) a0(mm)
7 3
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.81 0.1582 47.193
71.46 0.1567 42.275
73.97 0.1548 37.363
76.42 0.1535 30.857
85.04 0.1407 0.000
84.05 0.1420 0.086
82.58 0.1443 5.173

| GATE:5mm
| No. Points (NPP) a0(mm)
7 5
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.91 0.1981 56.090
71.40 0.1958 54.124
73.98 0.1943 50.944
76.43 0.1915 46.841
88.31 0.1774 7.021
88.84 0.1767 4.213
90.37 0.1718 0.000

269
| GATE:9mm
| No. Points (NPP) a0(mm)
10 9
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.85 0.3680 79.151
71.42 0.3636 79.872
73.98 0.3595 78.986
76.49 0.3534 76.914
64.09 0.3717 78.877
81.06 0.3471 72.030
90.31 0.3153 50.363
100.19 0.2858 14.472
103.53 0.2698 0.000
110.68 0.2114 0.000

| GATE:11mm
| No. Points (NPP) a0(mm)
10 11
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.87 0.4587 84.601
71.46 0.4532 85.307
73.93 0.4496 85.709
76.40 0.4429 84.643
64.10 0.4671 83.574
81.01 0.4346 82.295
91.43 0.3925 66.861
100.45 0.3707 34.993
110.06 0.3067 0.000
115.36 0.2460 0.000

| GATE:13mm
| No. Points (NPP) a0(mm)
10 13
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.82 0.5329 88.033
71.45 0.5255 88.833
73.94 0.5191 88.268
76.45 0.5144 87.652
64.13 0.5382 87.226
81.04 0.5033 85.594
90.10 0.5005 77.221
100.40 0.4737 60.334
109.84 0.4205 17.679

270
120.44 0.3034 0.000

| GATE:15mm
| No. Points (NPP) a0(mm)
10 15
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.91 0.6235 91.389
71.40 0.6165 91.244
73.88 0.6094 90.631
76.45 0.6041 90.210
63.99 0.6286 89.980
81.00 0.5893 87.944
90.51 0.5756 81.369
100.15 0.5548 66.185
109.78 0.4862 29.537
122.38 0.3954 0.000

| GATE:17mm
| No. Points (NPP) a0(mm)
12 17
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.85 0.7259 92.719
71.48 0.7216 92.959
72.51 0.7193 93.391
73.98 0.7177 93.187
76.39 0.7087 92.311
80.94 0.6966 89.988
63.98 0.7295 90.868
90.10 0.6623 80.822
100.26 0.6362 67.261
110.21 0.6005 48.442
120.61 0.5036 0.000
125.18 0.4767 0.000

| GATE:19mm
| No. Points (NPP) a0(mm)
11 19
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.82 0.7873 91.825
71.44 0.7833 92.210
73.93 0.7795 92.721
76.44 0.7770 92.683
64.10 0.7881 91.121

271
81.06 0.7614 91.678
90.42 0.7127 80.934
100.22 0.6833 67.873
110.37 0.6466 49.790
119.92 0.5522 1.524
126.96 0.5082 0.000

| GATE:21mm
| No. Points (NPP) a0(mmm)
14 21
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
64.07 0.8517 89.998
68.81 0.8440 90.682
71.40 0.8396 91.048
73.92 0.8337 91.502
76.41 0.8313 91.567
81.05 0.8223 90.789
90.14 0.7899 82.723
100.23 0.7522 68.686
110.43 0.7065 48.724
120.18 0.6672 23.940
121.47 0.6272 4.861
122.33 0.6190 0.653
123.26 0.6103 0.000
130.33 0.5591 0.000

| GATE:23mm
| No. Points (NPP) a0(mm)
14 23
| N11 Q11 EFFY
| (r/min) (m^3/s) (%)
68.92 0.9017 89.415
71.32 0.8984 89.596
73.94 0.8935 89.692
76.50 0.8859 89.311
64.08 0.9099 88.024
80.94 0.8762 88.195
124.68 0.6557 0.000
123.57 0.6640 0.684
122.83 0.6722 4.854
90.05 0.8588 81.104
100.40 0.8195 67.687
110.36 0.7642 47.491
120.27 0.7221 23.537
132.19 0.6194 0.000

272
Turbine Characteristics Input Variable Index:

KTCHAR—Turbine Characteristics (INTEGER) The number of sets of different head

and torque characteristics needed for the simulation. KTCHAR actually is the number of

different turbines in system, each of them with different characteristics. For the turbines

with the same head and torque characteristics, only one set of data is needed to input, but

the variable TCODE in *.DAT file should be specified to determine the order of sets in

TURB*.CHR for each turbine unit. The maximum number of KTCHAR is 40 as many as

the maximum turbines in a system.

TURBTYPE —Turbine Type Specifier (CHARACTER) Current version of TransAM

can deal with three types of turbines, that is, 1) ‘Fixed WG’ (no wicket gates or the wicket

gates fixed); 2) ‘Francis’ or ‘WG’ (turbines with wicket gates, Francis turbine is a typical

example); 3) ‘Kaplan’ (turbines with wicket gates and blades); and 4) ‘Pump-turbine’

(reversed turbine used in pump-storage hydropower plant). As for the Impulse turbine, it

is treated as an equivalent of BCTYPE=’Valve’. Actually, only Kaplan turbine needs the

exact input of ‘Kaplan’, which will infer other inputs for information of its blades.

In subroutine of solution of turbine boundary conditions, the type of turbine is determined

automatically by number of blade angle settings (NBlade) and the number of wicket-gate

settings (Ndy), and both of them are input from the turbine characteristics file

TURB*.CHR in BNDCND.f90.

QR—Rated Turbine Flow (REAL)

HR—Rated Turbine Head (REAL)

NR—Rated Turbine Speed (INTEGER)

273
ETR—Rated Turbine Efficiency (REAL)

D1—Diameter of Turbine Runner (REAL)

NBlade—Turbine Blades Angle Settings (INTEGER) Nblade>1 (up to 10) for Kaplan

turbine; NBlade=1 for other types of turbines which can be determined automatically by

program. Therefore, this input is needed only for Kaplan turbines.

Ndy—Turbine Wicket-gate Settings (INTEGER) The number of wicket-gate settings

between fully closed position y=0 to fully open position y=1.0, it is variable between (6-

15) depending on the available model test data, more data available higher accuracy in

modeling.

a00—Absolute Value of Full Wicket-gate Opening (REAL) Unit in mm

NPP—Input Data Points for Each Wicket-gate Position (INTEGER)

a0—Absolute Value of Each Wicket-gate Opening (REAL) Unit in mm

N11—Unit Speed of Model Turbine (REAL) Definition is found in chapter 2.

Q11—Unit Flow of Model Turbine (REAL) Definition can be found in chapter 2

EFFY—Efficiency of Model Turbine (REAL) The efficiency should be modified from

model to prototype.

274
AppendixⅡ: Turbine Characteristics Output Data Sample

-------------------------------------------------------------------------
| Total No. of Turbine Characteristics-KTCHAR
1
| TURBINE No. 1 (Starting with first turbine characteristics)

| Blade Angle set No.1 (Nblade=1 for Francis turbine)

| NBlade Ndy Ndx (Ndx – Number of points at each gate and same for all gates)
1 10 89

| Wicket-gate set No. 1


0.1304348 (This is relative wicket-gate opening, used in interpolation)

| Head-WH characteristic (89 points of Suter Head Parameter from x =–180˚to 180˚)
1.570 1.592 1.614 1.637 1.661 1.685 1.710 1.735 1.762 1.789
1.817 1.846 1.875 1.906 1.937 1.969 2.003 2.037 2.073 2.109
2.147 2.186 2.226 2.267 2.310 2.354 2.400 2.447 2.495 2.545
2.597 2.650 2.705 2.761 2.819 2.879 2.940 3.002 3.066 3.131
3.196 3.262 3.328 3.394 3.459 3.521 3.581 3.636 3.685 3.725
3.755 3.771 3.769 3.745 3.694 3.611 3.489 3.322 3.103 2.826
2.488 2.087 1.625 1.107 0.659 0.287 -0.096 -0.478 -0.846 -1.191
-1.504 -1.778 -2.013 -2.206 -2.361 -2.479 -2.565 -2.624 -2.659 -2.675
-2.674 -2.661 -2.637 -2.606 -2.568 -2.526 -2.481 -2.434 -2.386
| Torque-WB characteristic (89 points of Suter Torque Parameter from x =–180˚to
180˚)
0.481 0.487 0.494 0.501 0.508 0.515 0.523 0.530 0.538 0.546
0.555 0.563 0.572 0.581 0.590 0.600 0.610 0.620 0.631 0.641
0.652 0.664 0.676 0.688 0.700 0.713 0.726 0.740 0.754 0.769
0.783 0.799 0.814 0.830 0.847 0.864 0.881 0.898 0.916 0.934
0.952 0.969 0.987 1.004 1.021 1.036 1.050 1.063 1.073 1.080
1.083 1.081 1.073 1.058 1.034 0.999 0.952 0.890 0.811 0.715
0.600 0.466 0.313 0.145 -0.007 -0.241 -0.478 -0.709 -0.928 -1.129
-1.306 -1.458 -1.583 -1.682 -1.757 -1.810 -1.843 -1.860 -1.864 -1.857
-1.841 -1.819 -1.791 -1.760 -1.725 -1.689 -1.652 -1.615 -1.577
| Wicket-gate set No. 2
0.2173913
| Head-WH characteristic
1.158 1.174 1.191 1.208 1.225 1.243 1.262 1.281 1.300 1.320
1.341 1.362 1.384 1.407 1.430 1.454 1.479 1.504 1.531 1.558
1.586 1.615 1.644 1.675 1.707 1.740 1.774 1.809 1.845 1.882
1.921 1.960 2.001 2.044 2.087 2.132 2.177 2.224 2.272 2.321
2.370 2.420 2.470 2.520 2.570 2.617 2.663 2.705 2.743 2.775
2.799 2.813 2.813 2.797 2.760 2.698 2.607 2.481 2.314 2.103
1.844 1.536 1.179 0.791 0.532 0.342 0.143 -0.057 -0.251 -0.435

275
-0.603 -0.753 -0.882 -0.991 -1.079 -1.148 -1.201 -1.239 -1.264 -1.279
-1.285 -1.284 -1.278 -1.266 -1.252 -1.234 -1.215 -1.195 -1.173
| Torque-WB characteristic
0.388 0.393 0.398 0.404 0.410 0.415 0.422 0.428 0.434 0.441
0.447 0.454 0.461 0.469 0.476 0.484 0.492 0.500 0.509 0.517
0.526 0.536 0.545 0.555 0.565 0.576 0.586 0.598 0.609 0.621
0.633 0.645 0.658 0.671 0.684 0.698 0.712 0.726 0.741 0.755
0.770 0.785 0.799 0.813 0.827 0.840 0.851 0.862 0.870 0.876
0.879 0.878 0.871 0.859 0.839 0.810 0.771 0.719 0.653 0.572
0.475 0.362 0.233 0.089 -0.013 -0.064 -0.116 -0.167 -0.215 -0.258
-0.297 -0.329 -0.356 -0.377 -0.392 -0.403 -0.409 -0.412 -0.412 -0.410
-0.406 -0.401 -0.394 -0.387 -0.379 -0.371 -0.362 -0.354 -0.345
| Wicket-gate set No. 3
0.3913043
| Head-WH characteristic
0.648 0.657 0.666 0.676 0.686 0.696 0.706 0.717 0.728 0.740
0.751 0.764 0.776 0.789 0.802 0.816 0.830 0.845 0.860 0.875
0.891 0.908 0.925 0.943 0.961 0.980 1.000 1.020 1.041 1.063
1.085 1.109 1.133 1.158 1.183 1.210 1.237 1.265 1.294 1.323
1.353 1.384 1.415 1.445 1.476 1.505 1.534 1.560 1.584 1.603
1.617 1.624 1.621 1.606 1.575 1.524 1.449 1.344 1.205 1.027
0.820 0.653 0.532 0.433 0.376 0.312 0.244 0.173 0.101 0.032
-0.033 -0.093 -0.146 -0.193 -0.232 -0.265 -0.292 -0.314 -0.330 -0.343
-0.352 -0.358 -0.361 -0.363 -0.362 -0.361 -0.358 -0.355 -0.351
| Torque-WB characteristic
0.438 0.444 0.450 0.456 0.463 0.470 0.477 0.484 0.491 0.498
0.506 0.514 0.522 0.531 0.539 0.548 0.557 0.567 0.577 0.587
0.597 0.608 0.619 0.631 0.642 0.654 0.667 0.680 0.693 0.707
0.721 0.736 0.751 0.767 0.783 0.799 0.816 0.833 0.850 0.868
0.886 0.903 0.921 0.938 0.955 0.971 0.985 0.997 1.007 1.013
1.015 1.011 0.999 0.978 0.945 0.897 0.831 0.744 0.633 0.494
0.338 0.195 0.078 0.005 0.003 0.000 -0.002 -0.004 -0.007 -0.009
-0.011 -0.012 -0.014 -0.015 -0.016 -0.016 -0.017 -0.017 -0.017 -0.017
-0.017 -0.017 -0.017 -0.016 -0.016 -0.016 -0.016 -0.015 -0.015
| Wicket-gate set No. 4
0.4782609
| Head-WH characteristic
0.430 0.436 0.442 0.449 0.456 0.462 0.470 0.477 0.484 0.492
0.500 0.509 0.517 0.526 0.535 0.545 0.554 0.564 0.575 0.586
0.597 0.608 0.620 0.633 0.645 0.659 0.673 0.687 0.702 0.717
0.733 0.750 0.767 0.785 0.804 0.823 0.843 0.864 0.885 0.907
0.930 0.953 0.977 1.001 1.025 1.049 1.072 1.094 1.115 1.133
1.148 1.157 1.159 1.152 1.132 1.097 1.041 0.960 0.848 0.714
0.600 0.509 0.423 0.376 0.340 0.298 0.251 0.202 0.152 0.103
0.055 0.011 -0.029 -0.064 -0.095 -0.121 -0.143 -0.162 -0.176 -0.188
-0.197 -0.204 -0.209 -0.212 -0.214 -0.215 -0.215 -0.215 -0.213

276
| Torque-WB characteristic
0.383 0.389 0.394 0.400 0.406 0.412 0.418 0.424 0.431 0.438
0.444 0.452 0.459 0.467 0.474 0.482 0.491 0.499 0.508 0.517
0.527 0.536 0.546 0.557 0.568 0.579 0.590 0.602 0.615 0.627
0.641 0.654 0.668 0.683 0.698 0.713 0.729 0.745 0.762 0.779
0.796 0.813 0.830 0.847 0.864 0.880 0.894 0.907 0.918 0.925
0.928 0.925 0.913 0.892 0.858 0.807 0.736 0.640 0.515 0.376
0.249 0.137 0.037 0.004 0.002 0.000 -0.002 -0.004 -0.006 -0.008
-0.009 -0.011 -0.012 -0.013 -0.013 -0.014 -0.014 -0.014 -0.015 -0.015
-0.014 -0.014 -0.014 -0.014 -0.014 -0.013 -0.013 -0.013 -0.013
| Wicket-gate set No. 5
0.5652174
| Head-WH characteristic
0.375 0.380 0.386 0.391 0.397 0.403 0.409 0.416 0.423 0.429
0.437 0.444 0.451 0.459 0.467 0.476 0.484 0.493 0.502 0.512
0.522 0.532 0.542 0.553 0.565 0.577 0.589 0.602 0.615 0.629
0.643 0.658 0.674 0.690 0.706 0.724 0.742 0.760 0.780 0.800
0.820 0.841 0.862 0.884 0.906 0.928 0.949 0.969 0.987 1.003
1.015 1.022 1.021 1.010 0.985 0.942 0.876 0.781 0.674 0.572
0.458 0.392 0.355 0.326 0.289 0.248 0.203 0.155 0.107 0.060
0.016 -0.025 -0.061 -0.092 -0.119 -0.142 -0.160 -0.175 -0.187 -0.196
-0.202 -0.207 -0.210 -0.211 -0.212 -0.212 -0.211 -0.209 -0.207
| Torque-WB characteristic
0.393 0.398 0.404 0.410 0.416 0.422 0.429 0.435 0.442 0.449
0.456 0.463 0.471 0.479 0.487 0.495 0.504 0.513 0.522 0.532
0.541 0.552 0.562 0.573 0.584 0.596 0.608 0.621 0.634 0.647
0.661 0.675 0.690 0.705 0.721 0.737 0.754 0.771 0.789 0.807
0.825 0.844 0.862 0.880 0.898 0.915 0.931 0.944 0.955 0.961
0.962 0.956 0.940 0.911 0.865 0.797 0.702 0.574 0.432 0.299
0.168 0.068 0.020 0.002 -0.017 -0.038 -0.058 -0.077 -0.095 -0.110
-0.124 -0.135 -0.144 -0.150 -0.155 -0.157 -0.158 -0.158 -0.157 -0.155
-0.153 -0.150 -0.147 -0.143 -0.140 -0.136 -0.133 -0.129 -0.126
| Wicket-gate set No. 6
0.6521739
| Head-WH characteristic
0.296 0.301 0.305 0.310 0.314 0.319 0.324 0.330 0.335 0.340
0.346 0.352 0.358 0.365 0.371 0.378 0.385 0.392 0.400 0.407
0.415 0.424 0.433 0.442 0.451 0.461 0.471 0.482 0.493 0.504
0.516 0.528 0.541 0.555 0.569 0.584 0.599 0.615 0.631 0.648
0.666 0.684 0.703 0.722 0.741 0.761 0.780 0.798 0.814 0.829
0.839 0.845 0.843 0.831 0.804 0.757 0.685 0.604 0.532 0.445
0.386 0.350 0.312 0.266 0.214 0.157 0.096 0.035 -0.024 -0.080
-0.131 -0.175 -0.213 -0.245 -0.270 -0.290 -0.304 -0.314 -0.321 -0.325
-0.327 -0.326 -0.324 -0.321 -0.317 -0.313 -0.308 -0.302 -0.297
| Torque-WB characteristic
0.363 0.368 0.373 0.379 0.384 0.390 0.396 0.402 0.409 0.415

277
0.422 0.429 0.436 0.444 0.451 0.459 0.468 0.476 0.485 0.494
0.503 0.513 0.523 0.534 0.545 0.556 0.567 0.580 0.592 0.605
0.619 0.633 0.647 0.662 0.678 0.694 0.711 0.728 0.746 0.764
0.782 0.801 0.820 0.838 0.857 0.874 0.891 0.905 0.916 0.923
0.923 0.916 0.896 0.861 0.805 0.722 0.602 0.462 0.337 0.216
0.114 0.043 0.007 -0.032 -0.073 -0.116 -0.157 -0.196 -0.232 -0.263
-0.288 -0.308 -0.323 -0.333 -0.338 -0.341 -0.340 -0.337 -0.332 -0.326
-0.319 -0.311 -0.303 -0.295 -0.287 -0.279 -0.271 -0.263 -0.255

278
Appendix Ⅲ: Turbine Boundary Input Data Sample and Variable Index

A Sample of Input Data for Turbine Boundary Condition:

------------------------------------------------------------------------------------------
| DEVICE 4 IS A TURBINE UNIT

| BCTYPE NDN BCOUT


'TURB' 2 'OUTPUT'
| NLBC(1) NLBC(2) (Node list)
05 06
| TSID
'No.4 TURBINE'

| Zav(Htail0) Period(tail) AMP(tail)

| NValT Isyn Igov


0 1 2

| TTYPE TCODE
‘Francis’ 1
| TPHA0 TPHAmax TPHAmin

| NR QR HR PR N0 Q0 WRR Tg
| (rpm) (cms) (m) (Kw) (rpm) (cms) (Kg-m2) (s)
187.5 191.5 142.0 250000 187.5 191.5 72.5E+5 20

| NTy
4
| yG (I) I=1,NTy
0.904 0.904 0.715 0.0
| TyG(I) I=1,NTy
0.0 100.2 104.0 130.2

| NTP
3
| PG(1) PG(2)
250000 250000 0.0
| TPG(1) TPG(2)
0.0 100.0 100.1
------------------------------------------------------------------------------------

279
Input Variable Index for Turbine Boundary Condition:

1) General Data for Turbine Boundary Condition

BCTYPE (CHARACTER) The type of boundary condition, each turbine unit is

designed by ‘TURB’, and in one system can have up to 40 turbine units.

NDN (INTEGER) The number of nodes associated with a turbine boundary

condition. Usually, the turbine upstream is always linked to a penstock, in which

the pressure rise is the concern, but the turbine downstream depends. NDN=2 is set

for an in-line turbine unit, that is, the turbine discharges water to a pressured

conduit or pressured system; NDN=1 for an one-node (external) turbine unit, that is,

its short draft tube can be ignored, and the turbine downstream is considered to be

directly linked with tailrace or reservoir.

BCOUT (CHARACTER) Set BCOUT to ‘OUTPUT’ if time history results are

required. These will be logged every IPRINT time steps in *.BND file.

NLBC (INTEGER) This defines the point of attachment of the turbine unit to the

pipe network. For external turbines only a one-node identifier is required which is

the upstream end of the turbine. The in-line turbines are always defined by

upstream node and downstream node. The first node input in the list is assumed to

be the upstream side and the later the downstream side.

TSID—Turbine Unit Identifier (CHARACTER) Each turbine may be ‘named’

for easier reference in the output results. A string (enclosed by a pair of single

quotes) of up to 35 characters may be used.

NvalT—Total Number of Valves Attached to a Turbine Unit (INTEGER)

280
In current version of TransAM, the maximum 3 valves can be attached to one

turbine unit, that is, 1 Turbine Control Valve (‘TCV’), 1 Bypass Valve and 1 Surge

Valve. If a turbine has installed more than 3 valves or has more complicated valves,

we always can deal with any valve easily as a separate valve boundary condition by

adding one characteristic pipe section at each side of the valve.

Isyn—Identifier of Synchronous or Isolated Turbine Unit. (INTEGER) Isyn=0

for synchronous units, and Isyn=1 for isolated units.

2) Sinusoidal Tail Head For one-node (external) turbines, the downstream/tail water

head can be a constant or varies as a sine wave function because of the wind or other

influences. (Reference to P147, TransAM Manual)

Htail0 = (REAL) the mean piezometric downstream/tail head

Ptail = (REAL) the time required for a complete wave cycle

AMPtail = (REAL) the amplitude of the pressure head variation

3) Input Data for Turbine Attached Valves –These valves are typically used to limit

water hammer as auxiliary flow control devices. They are activated under emergency

situations, for example, when the wicket gate failed to be closed. They also function

as an additional means when the primary flow control system (wicket-gate, pin valve,

etc.) is not efficient enough. They are quite similar to those valves in pump station

(reference to P149 -153, TransAM Manual), and the first or last character ‘T’ in these

variables is used to indicate that they belong to turbine unit.

TValTYP—Type Specifier of Turbine Attached Valves (CHARACTER)

Following 3 types of valves have been included in modeling of turbine boundary

281
condition, any other more complicated valve can be treated as a separate boundary

condition.

‘TCV’—Turbine Flow Control Valve, which is installed at upstream side of

turbine. This valve follows the input Tau-curve.

‘BYPASS’—Turbine bypass valve, which is used as an additional device to

relieve turbine flow through a bypass pipe if without wicket gates or they

cannot to be closed effectively under load rejection.

‘SURGE’—Surge relief valve, which can be activated by time, upsurge at

upstream or downsurge at downstream of turbine. The valve executes its

opening when a setpoint is exceeded at first time, and then remains (fully) open,

i.e., the subsequent violations of a setpoint have no effect on the behavior of the

valve.

TES—Valve discharge Constant (REAL) The discharge through a valve is

expressed as: QV = (TES )τ H

TTFAC—Reverse Flow Factor (REAL) In current version, we only take the

default value TTFAC=0.0 for any turbine-attached valve, which means the valve will

permit flow only in the usual direction. TTFAC is a dimensionless parameter.

TTAU0—Initial Valve Setting (REAL) TAU0 is the value of τ at the beginning of

the valve motion; it is a dimensionless relative valve opening, from 0.0 to 1.0.

TTAUF—Final Valve Setting (REAL) TAUF is the value of τ at the conclusion of

the valve motion; it is a dimensionless relative valve opening, from 0.0 to 1.0.

282
TTV1—Segment 1 Valve Motion Duration (REAL) In seconds

TTV2—Valve motion Duration (REAL) In seconds

TSET0—Valve Setpoint (REAL)

TSET1—Valve Setpoint (REAL)

TSET2—Valve Setpoint (REAL)

NSTAUT—Starting Tau Index (INTEGER)

NENDTAUT—Ending Tau Index (INTEGER)

MMT—Number of Closure Points (INTEGER)

TAUTT—Tau Curve Points (REAL)

4) Turbine Data

TTYPE—Turbine Type Specifier (CHARACTER) Current TransAM can deal with

four types of turbines, that is, 1) ‘Fixed WG’ (no wicket gates or the wicket gates

fixed); 2) ‘Francis’ or ‘WG’ (turbines with wicket gates, Francis turbine is a typical

example); 3) ‘Kaplan’ (turbines with wicket gates and blades); and 4)’Pump-turbine’

(reversed turbine used in pump-storage hydropower plant). As for the Impulse

turbine, it is treated as an equivalent of BCTYPE=’Valve’. Actually, except for

pump-turbines, other types of turbine could be determined automatically by number

of blade angle settings (NBlade) and the number of wicket gate settings (Ndy), both

of them are input from the turbine characteristic data file in BCBND.f90.

TCODE—Turbine Type Identifier (INTEGER) An integer value acts as an index

describing the ordinal position of the particular turbine’s characteristics in the data

283
file TURB*.CHR. In other words, if the characteristics for a turbine are third data set

in TURB*.CHR, then Tcode for that turbine equals 3. The manner in which the

turbine characteristics are obtained for a given turbine and how these are assembled

into the file TURB*.CHR are explained in Chapter 4. Each turbine can have its own

specific characteristics in the file TURB*.CHR. If all turbines in the system could be

modeled using the same head and torque data, only a single set of characteristics

need be used and TCODE would equal 1 for all turbines. (The value of KTCHAR in

the file TURB*.CHR would be 1)

TPHA0—Initial Blades Angle Setting (REAL) Only for Kaplan turbines

TPHAmax—The Maximum Blades Angle Setting (REAL) Only for Kaplan

turbines

TPHAmin—The Minimum Blades Angle Setting (REAL) Only for Kaplan

turbines

TQR—Rated Turbine Flow (REAL)

THR—Rated Turbine Head (REAL)

TNR—Rated Turbine Speed (REAL) In rpm.

TQ0—Initial Turbine Flow (REAL)

TN0—Initial Turbine Speed (REAL)

TWRR—Turbine-Generator Unit Inertia (REAL) (Kg-m2 or lb-ft2) It is a

product of weight times the radius of gyration.

TPR—Rated Turbine Output Power (REAL)

284
TTg— Wicket-gate Closure Time at the Maximum Rate (REAL) In seconds, it is

actually the shortest time period to close the wicket-gate from full opening to full

closing.

5) Turbine Governor

Igov—Identifier of Turbine Governor Status. (INTEGER)

Igov=0 for turbine governor in normal operation.

Igov=1 for malfunctioned governor, the consequence is that the wicket- gate

failed to be closed under load rejection, and may cause turbine full runaway

speed if no other means to cut off the upstream coming flow.

Under this circumstance, the main valve at turbine upstream side should be

activated to control the flow rate of turbine, so this scenario may be combined

with an attached ‘CTV’.

Igov=2 for the turbine without any governor (or the governor malfunctioned) but

the wicket-gate is closed as a prescribed rate. This status may seldom occur in

practical applications; however, it is useful to analyze some extreme situations,

for instance, to simulate the scenario of sudden wicket- gate closing by setting

0.1s from full opening to full closing.

Following Governor Data only for Igov=0

GOVTYP—Governor Type Specifier (CHARACTER) GOVTYP=’PID’ for

proportional, differential, and integral governor; GOVTYP=’TEMP’ for

temporary-droop governor; and GOVTYP=’ACCE’ for accelerometric governor.

However, in current version, only GOVTYP=’PID’ is defined and solved. The

285
governor equation of ‘PID’ governor is referenced to P163 in Wylie’s book.

Following parameters are used in its equation.

SIGMA—Permanent Speed Drop (REAL)

DELTA—Temporary Speed Drop (REAL)

TAL—Promptitude Time Constant (REAL)

TD—Dashpot Time Constant (REAL)

6) Variation of Wicket gate Opening for Igov=2 Situation

NTY—Number of Points in y-curve (INTEGER)

YG—Sequence of Wicket gate Openings in y-curve (REAL)

TYG—Sequence of Time corresponding to YG in y-curve (REAL)

Wicket gate opening y is a dimensionless variable.

7) Load Variation

NTP—Number of Points in P-curve (INTEGER)

PG—Sequence of turbine loads in P-curve (REAL)

TPG—Sequence of Time corresponding to PG in P-curve (REAL)

The turbine load P is defined as the output power of turbine, which varies with the

load of unit. This set of data need to be input under any situation. If no load changes,

P-curve can be input as a straight line parallel to time axial.

286
Appendix Ⅳ: Turbine Boundary Output Data Sample and Variable Index

A Sample of Output Data for Turbine Boundary Condition:

-------------------------------------------------------------------------------------
|=======DEVICE No. 4 ==========|

TYPE IS in-line turbine


INCIDENT NODES: 05 06
OUTPUT
in-line turbine No.4 TURBINE

The Number of Valves with this Turbine = 0

Synchrounous Speed Specifier = 1

Governor Specifier = 2

DISCHARGE (m3/s) TURBINE HEAD (m )


Rated Initial Rated Initial
191.50 191.50 142.000 130.581

turbine No. SPEED (rpm) SPECIFIC RATED EFF. POWER WRR


Rated Initial SPEED (fraction) Rated (kW) (kg-m2)
No.4 187. 187. 190.49 0.930 250000.0 7250000.

TIME Ty v hh ALPH BETA Qturb Hturb Nturb Hsuc Hdis


(s) (m3/s) (m ) (rpm) (m ) (m )
===============================================================
=========
0.000 0.904 1.000 0.920 1.000 1.000 191.500 130.581 187.500 369.825 227.827
0.123 0.904 1.009 0.934 1.000 0.928 193.212 132.595 187.500 363.034 230.439
0.246 0.904 1.015 0.944 1.000 0.944 194.421 134.027 187.500 366.386 232.359
0.368 0.904 1.016 0.945 1.000 0.947 …….

Hsuc-NELEV HDIS-NELEV2 TAUC Hc Qc TAUB QB TAUS QS


(m ) (m ) (m ) (m3/s) (m3/s) (m3/s)
===============================================================
=========
160.825 18.827 0.000 0.000 --- 0.000 0.000 0.000 0.000
154.034 21.439 0.000 0.000 --- 0.000 0.000 0.000 0.000
157.386 23.359 0.000 0.000 --- 0.000 0.000 0.000 0.000
--------------------------------------------------------------------------------------------------

287
Turbine Boundary Condition Output Variable Index:

Ty — Relative wicket-gate opening

v — Relative/dimensionless turbine flow

hh — Relative/dimensionless turbine head

ALPH — Relative/dimensionless turbine speed

BETA — Relative/dimensionless turbine torque

Qturb —Absolute turbine flow

Hturb —Absolute turbine head

Nturb —Absolute turbine speed

Hsuc — Piezometric pressure head at turbine upstream node

Hdis — Piezometric pressure head at turbine downstream node

Hsuc-NELEV — Pressure head at turbine upstream node

Hdis-NELEV2 — Pressure head at turbine downstream node

TAUC — Relative opening of ‘TCV’

Hc — Head loss across ‘TCV’

Qc — Flow rate passing through ‘TCV’

TAUB — Relative opening of ‘Bypass Valve’

QB — Flow rate pass through ‘Bypass Valve’

TAUS — Relative opening of ‘Surge Valve’

QS — Flow rate pass through ‘Surge Valve’

288

You might also like