Вы находитесь на странице: 1из 62

Master of Science Thesis

Water Hammer Phenomenon Analysis


using the Method of Characteristics and
Direct Measurements using a ”stripped”
Electromagnetic Flow Meter

by
Joel Carlsson

August 2016

Division of Nuclear Reactor Technology, Department of Physics


Royal Institute of Technology, SE-106 91 Stockholm, Sweden

TRITA-FYS 2016:33
ISSN 0280-316X
ISRN KTH/FYS/–16:33SE
TITLE: Water Hammer Phenomenon Analysis using
the Method of Characteristics and Direct Mea-
surements using a ”stripped” Electromagnetic
Flow Meter

AUTHOR: Joel Carlsson

SUPERVISOR: Professor H. Anglart

PARTNER: Royal Institute of Technology


INSTITUTION: Department of Engineering Physics

INDUSTRIAL PARTNER: Vattenfall AB / YRLF


INDUSTRIAL SUPERVISORS: Dr. L. Facciolo and Dr. K. Angele

NUMBER OF PAGES: viii, 52


Abstract
This thesis deals with physical and mathematical models in order to simulate, explain
and reconstruct real world events of water hammer at Nuclear Power Plants (NPP’s).

The present thesis constitutes a part of Vattenfall’s Thermal Hydraulic Loads program
for development and experimental validation of RELAP5/MOD3 models for water
hammer transients. The RELAP5/MOD3 code is used at NPPs for load calculations,
which is a prerequisite for the license to operate

In the present thesis, the Method of Characteristics (MOC) was implemented as a com-
plement to the RELAP5/MOD3 code. As opposed to RELAP5/MOD3, the MOC-
code also includes the compressibility of H2 O and the deformation of the pipe. The
performance of the MOC code and the RELAP5/MOD3 code was evaluated against
water hammer experiments performed in a test rig in Älvkarleby, Sweden. The results
show that the MOC code produces more accurate results and that the RELAP5/MOD3
code underestimate the mass flow gradient during the transient and therefore the loads
on the pipe. An extended MOC (taking the FSI-effects into account) is necessary to
yield even better agreement with the experimental results.

In connection to the experiments, work has also been performed, within the framework
of this thesis, to test a possible direct fast transient mass flow measurement using a
”stripped” conventional electromagnetic flow meter. The initial results are promising
and indicate the possibility of measuring the mass flow at fast transients.

i
Acknowledgements
First and foremost I want to thank my industrial supervisors, Dr. Luca Facciolo and
Dr. Kristian Angele for their guidance, support, and most importantly encouragement
during my time at Vattenfall AB R&D. I appreciate all their contributions of time, ideas,
and knowledge to make this experience productive and stimulating.

I also gratefully acknowledge the support and guidance of the Engineering Physics De-
partment of the Royal Institute of Technology and supervisor Professor Henryk Anglart.
This project represents a six-months work under the financial support of Vattenfall AB
R&D. To the staff and other master thesis students, many thanks for your friendship
and good advice.

I would also like to express gratitude to Dr. Hans Henriksson and Eric Lillberg for
helping particularly in the beginning, to start up and guiding me through the course of
this thesis.

Lastly, I would like to thank my family and friends for all their love and support. For
my parents and older siblings who raised and supported me in all my pursuits.

ii
Contents

Abstract i

Acknowledgements ii

List of Figures v

List of Tables vi

Nomenclature vii

1 Introduction 1

2 Background and Theory 4


2.1 Classical Water Hammer Phenomenon and Theory . . . . . . . . . . . . . 4
2.1.1 Conservation of Mass . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.1.2 Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . . 7
2.1.3 The Pressure (Water Hammer) Wave Speed . . . . . . . . . . . . . 8
2.1.4 The Classical Water Hammer Equations . . . . . . . . . . . . . . . 11
2.1.5 Unsteady Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Direct Mass Flow Measurement . . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.1 Electromagnetic flow meters . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Gibson’s indirect method . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3 Method of Characteristics 16
3.1 Solution by Method of Characteristics . . . . . . . . . . . . . . . . . . . . 16
3.1.1 Different MOC schemes for more complex systems . . . . . . . . . 20
3.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

4 Experiments 25
4.1 Measuring Devices and System . . . . . . . . . . . . . . . . . . . . . . . . 27

5 Results and Discussion 30


5.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.1.1 Experimental results . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.1.1.1 Direct Mass Flow Measurement . . . . . . . . . . . . . . 32
5.1.2 Comparison between the RELAP and the MOC codes . . . . . . . 34

iii
Contents iv

5.1.3 Validation of the MOC and RELAP codes using experimental data 36
5.2 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
5.2.1 Extended MOC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

6 Conclusions 39

A Supplement on Water Hammer and the Friction Term 41

B Model Description 43

C Calibration protocol of electromagnetic flow meter 48

Bibliography 50
List of Figures

2.1 Water hammer transient . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5


2.2 Control volume diagram of a conduit for the continuity equation . . . . . 7
2.3 Control volume diagram of a conduit for the momentum equation . . . . . 8
2.4 Overview of the important principals of an EMF . . . . . . . . . . . . . . 14

3.1 Characteristic lines in x-t plane. . . . . . . . . . . . . . . . . . . . . . . . 18


3.2 Computational mesh/grid with index for MOC . . . . . . . . . . . . . . . 19
3.3 Boundary condition, Ideal Pipe Attachment to Pressure Vessel . . . . . . 22
3.4 Boundary condition, Valve at Discharge End of Pipe . . . . . . . . . . . . 23

4.1 Photograph of the test rig at Vattenfall R&D in Älvkarleby, Sweden . . . 25


4.2 Overview of the important components of the test rig . . . . . . . . . . . 26
4.3 The fixed valve in the test-rig. . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.4 Illustration of a pressure sensor model used . . . . . . . . . . . . . . . . . 28
4.5 The two electromagnetic flow meters at the DN50 pipe line between the
pump and the reservoir in the test-rig. . . . . . . . . . . . . . . . . . . . . 29

5.1 The positive pressure wave propagating upstream from the valve. . . . . . 31
5.2 The negative pressure wave propagating downstream from the valve. . . . 32
5.3 Voltage measurements proportional to the flow velocity together with
filtered and magnet field. . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
5.4 Voltage measurement together with the filtered one. . . . . . . . . . . . . 33
5.5 Filtered voltage together with time shifted pressure. . . . . . . . . . . . . 34
5.6 Pressure normalized by Joukowsky’s equation. . . . . . . . . . . . . . . . . 35
5.7 Pressure normalized by Joukowsky’s equation with MOCF SI . . . . . . . . 36
5.8 Flow rate normalized . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5.9 Pressure normalized by Joukowsky’s equation with experimental data. . . 37

B.1 Pressure change in the middle of the pipe until steady state was reached . 44
B.2 Mass flow in the middle of the pipe change until steady state was reached 45
B.3 The valve ratio change used in valve boundary condition . . . . . . . . . . 45

C.1 Calibration protocol of electromagnetic flow meter at SP, 2015-04-21 . . . 48


C.2 Calibration protocol of electromagnetic flow meter at SP, 2015-04-21 . . . 49

v
List of Tables

4.1 Pressure sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28


4.2 Electromagnetic Flow Meters . . . . . . . . . . . . . . . . . . . . . . . . . 28

B.1 Specification of parameters used in MOC . . . . . . . . . . . . . . . . . . 43

vi
Nomenclature

Roman
A cross-sectional area of the pipe [m2 ]
a flow area [m2 ]
B magnetic flux density [T ]
c pressure (water hammer) wave speed [m/s]
C contraction coefficient
Co Courant number
CS control surface [m2 ]
CV control volume [m3 ]
D diameter [m]
E Young’s elasticity modulus [P a] (or voltage [V ], see context)
f Darcy-Weisbach friction factor (or an adapted friction factor, see context)
F force [N ]
g gravitational acceleration [m/s2 ]
H piezometric head [m]
k dimensionless constant
K water bulk modulus [P a] (or loss coefficient, see context)
L length of pipe [m]
m mass [kg]
ṁ mass flow [kg/s]
M Mach number
n outward normal vector
p momentum [kg · m/s]
P piezometric pressure [P a]
t time [s]
u local freestream velocity [m/s]
V cross-sectional area average velocity [m/s] (or volume [m3 ], see context)
v fluid velocity [m/s]
Z elevation of the pipe center line from a given datum [m]

vii
Nomenclature viii

Greek
α angle between pipe and horizontal direction [rad]
β momentum correction coefficient
ε hoop strain
δ pipe wall thickness [m]
δx short x-directional length [m]
∆x spatial grid size [m]
∆t time step [s]
λ linear multiplier
µ viscosity [P a · s]
ψ nondimensional parameter
σ hoop (circumferential) stress [P a]
ρ fluid density [kg/m3 ]
τ shear stress [P a]

Sub and superscripts


A position, see context
B position, see context
e electrodes
end end of pipe
FSI fluid structure interaction
G position, see context
i space variable
j time variable
L length of pipe
S surface
Steady steady condition
Unsteady unsteady condition
v vena-contracta
w wall
Chapter 1

Introduction

Large pipe systems with long pipelines transporting fluids over great distances is a
reality in in modern society. The usage of small pipe diameters, high-velocity together
with sophisticated fluid control devices, many types of pumps and valves, coupled with
electronic sensors have increased the importance of correct design.

Nuclear power plants have systems with large networks of piping both for the production
of electricity and to ensure water cooling at all times in a reliable and safe way. The
water in nuclear power plants is often under high pressure and at high flow rate generated
by pumps. Pump failure, improper operations of valves and accidental events like power
losses and pipe ruptures create transient flows, which can lead to pressure waves through
the pipe system. A sudden change in flow like that generates a pressure pulse. This
phenomenon is called water hammer. Water hammer events rarely lead to pipe rupture
in nuclear power plants but due to conservative calculations together with limited models
many calculations are being made which costs time and money in the end.

Transient water hammer analysis is essential to verify design and operation of piping
systems to prevent damaging equipment and pipes. At this moment, 99 incidents induced
by water hammer exists in the IRS (Incident Reporting System), the nuclear safety
authorities reporting system, that is forward and collected in a database at the NEA
(Nuclear Energy Agency) and the IAEA (International Atomic Energy Agency). Thus,
correct water hammer analyses are essential for improving and withstanding a high level
of safety and availability of the electricity production from nuclear power plants.

Safer and more efficient reactors would not only be more profitable economically for
the nuclear power industry but it could also lead to better acceptance by the public
towards the nuclear industry, which could allow more research to improve and explore
the potential of nuclear power.

1
Chapter 1. Introduction 2

The majority owner of Swedish nuclear power, seven of the nine current existing elec-
tricity producing reactors in Sweden (four in Ringhals and three at Forsmark) [1], the
energy corporation Vattenfall received the request from Forsmark and Ringhals to vali-
date RELAP5/MOD3 (Reactor Excursion and Leak Analysis Program) models, a best
estimate transient simulation code, for water hammer transients. The code provides
a coupled thermal-hydrodynamic and point-kinetics capability with reactivity feedback
which makes it possible for example to simulate loss-of-coolant accidents (LOCAs) and
allows the study of various anticipated transients without scram (ATWS) [2], [3]. It is
mainly used for structural verification at Forsmark and Ringhals nuclear power plants.

The present thesis constitutes a part of Vattenfall’s Thermal Hydraulic Loads program
for experimental validation of RELAP5/MOD3 models for water hammer transients.
To succeed in full filling the validation objective, a test rig with a rapidly closing pneu-
matically controlled valve has been built at Vattenfall R&D in Älvkarleby, Sweden. The
rig is intended to be used for studies of water hammer transients and to generate val-
idation data. This Vattenfall project is large and comprehensive and there are parts
which will not be presented in detailed in this report but plays a major role towards the
validation. These are:

- The execution of the experiments.


- The RELAP5/MOD3 models of the test rig.
- A Computational Fluid Dynamic (CFD) model using the OpenFOAM software
of the test rig.

One drawback in the experiments is that the present electromagnetic flow meters does
not have fast enough time resolution to measure the transient flow rate which originates
from the water hammer. Therefore the mass flow rate data from RELAP5/MOD3
cannot be validated.

During the thesis studies, effort has been made to:

(1) solve the classic water hammer equations, including a small compressibility effect
of water and the deformation of the pipe which are not included in RELAP5/MOD3
simulations.

(2) identify and implement a possible direct method to measure the transient mass flow
rate in the test rig and

(3) investigate the usage of Gibsons indirect method for transient flow measurement
via transient pressure measurements provided from the test rig (Vattenfall Hydro power
provided a code using the Gibson method) .
Chapter 1. Introduction 3

All these are means to improve the reliability of the RELAP5/MOD3 models for water
hammer transients. This report will focus on (1) the classic water hammer equations
simulations.
Chapter 2

Background and Theory

This chapter describes the background and the theory which the present work is based
upon. It summarizes the classical water hammer equations and a possible approach to
directly measure the mass flow during the water hammer transient.

2.1 Classical Water Hammer Phenomenon and Theory

Any disturbances or change, planned or accidental, in a pipe system’s mean flow will
initiate pressure waves in the system. As the pressure waves propagate, they create
transient pressure and flow conditions. Potentially these transient conditions can have
serious consequences if not properly addressed by proper analysis, design and operational
considerations. For example, when a valve starts to close, the pressure increases until
it is completely closed. Depending on the time it takes for the valve to close, the first
pressure increase will have reached a position in space which will give the pressure a
wavelength. When the valve is fully closed, the pressure has reached maximum and the
wave propagate forward to even out the pressure difference.

Fig. 2.1 represents a water hammer transient of a long horizontal pipe filled with only
water. The left boundary is a reservoir with constant pressure and the right boundary
is a valve (gate) situated in the end of the figure. (a) A steady water flow is propagating
in the right direction of the figure. (b) The valve closes instantaneously (no wavelength)
initiating the water hammer phenomena, which is illustrated by the initiation of the
pressure increase in fig. 2.1.

(c) The positive water hammer propagates forward to even out the pressure difference.
The negative water hammer travels in the opposite direction of the positive. The spatial
extension continues in the flow direction to the right of the valve. (d) When the pressure

4
Chapter 2. Background and Theory 5

Figure 2.1: Water hammer transient

Source: Micro-hydropower sourcebook[4]

reaches the reservoir it reflects positively back and (e) propagate back towards the valve
leaving behind the initiate pressure. (f) The flow velocity is pushed in the opposite
direction of the pressure when it propagate back. (g) At the valve, the pressure will
reflect negatively (h) and propagate towards the reservoir again. (i) At negative pressure
the flow velocity will be zero. (j) When the negative pressure reflects again at the
reservoir (k) the pressure will return to initial condition (l) initiating the flow velocity
in same direction as the negative pressure propagte back.

In real life, (the valve cannot close instantaneously and) there will be a wavelength of
the pressure wave. When measuring the pressure at a distance from the reservoir within
the wave length of the pressure increase, a superposition of the wave with itself (the first
Chapter 2. Background and Theory 6

and the last part of the wave length) will result in lower amplitude than for a position
further away from the reservoir.

Joukowsky laid the foundation of the ”fundamental water hammer theory”, which de-
scribes the pressure amplitude.

c∆V
∆P = ±ρc∆V or ∆H = ± (2.1)
g

where c = the pressure (water hammer) wave speed, ρ = fluid density, g = acceleration
caused by gravity, P = ρg(H − Z) = piezometric pressure, Z = elevation of the pipe
R
center line from a given datum, H = piezometric head, V = A udA/A = cross-sectional
average velocity, u = local freestream velocity of the fluid and A = cross-sectional area
of the pipe. The head (H) is a physical parameter often used in hydraulics calculations
but the focus in this thesis will be on the pressure (P). The negative sign in Eq.2.1
describes a water hammer wave moving downstream while the positive sign describes
the water hammer wave moving upstream.

Water hammer transients are normally assumed to be axi-symmetric since the axial
changes of mass, momentum, and energy are a lot greater than their radial counterparts.
With the axi-symmetric assumption, the one-dimensional water hammer equations in
transient pipe flows are derived by applying the conservations of mass and momentum
in a control volume.

2.1.1 Conservation of Mass

The mass conservation principle results from the fact that mass cannot be created or
destroyed. Thus, the following equation is valid:

Dm
=0 (2.2)
dt

describing that the change of mass is equal to zero in the system. The mass conservation
principle related to a control volume CV and its control surface CS can be expressed
using the Reynolds transport theorem:

ZZZ ZZZ ZZ
Dm ∂
= ρ dV = ρ dV + ρvv · n dA = 0 (2.3)
dt CV ∂t CV CS
Chapter 2. Background and Theory 7

where n = outward normal vector to the control surface and v = fluid velocity vector.
Eq. 2.3 scaled down to one-dimension in a pipe with variable cross-section area, referring
to Fig. 2.2, can be written as follows:

Z x+δx ZZ

ρA dx + ρvv · n dA = 0 (2.4)
∂t x CS

The local form of Eq. 2.4, obtained by shrinking the length of the control volume to
zero (i.e. δx tends to zero), is Eq. 2.5.

∂ (ρA) ∂ (ρAV )
+ =0 (2.5)
∂t ∂x

Eq. 2.5, also called the ”continuity equation”, is the area-averaged mass balance equation
for one-dimensional unsteady and compressible fluids in a flexible pipe.

Figure 2.2: Control volume diagram of a conduit for the continuity equation

Source: A review of water hammer theory and practice[5]

2.1.2 Conservation of Momentum

Newton’s second law of motion states that the change of momentum of a system is equal
to the forces exerted on the system by its surrounding, as described in Eq. 2.6

Dpp X
= F (2.6)
Dt

By expanding the vector sum of the forces and applying it to a control volume, Eq. 2.7
yields

ZZZ ZZZ ZZ
X Dpp D ∂vv ρ
F = = v ρ dV = dV + v ρvv · n dA (2.7)
Dt Dt CV CV ∂t CS

Applying Eq. 2.7 to a specific control volume of a conduit, referring to Fig. 2.3; con-
sidering the gravity, wall shear and pressure gradient forces on the fluid and taking the
Chapter 2. Background and Theory 8

limit as done previously (i.e. δx tends to zero). The following local form of the axial
momentum equation is derived:

∂(ρAV ) ∂(βρAV 2 ) ∂P
+ = −A − πDτw − γAsinα (2.8)
∂t ∂x ∂x

where τw = shear stress at the pipe wall, γ = ρg which is the unit gravity force, α =
angle between the pipe and the horizontal direction and β = A u2 dA which is the
R

momentum correction coefficient. By using the product rule of differentiation to Eq.


2.8, and dividing everything by ρA yields Eq. 2.9:

∂V ∂V ∂(β − 1)ρAV 2 1 ∂P τw πD
+V + + + gsinα + =0 (2.9)
∂t ∂x ∂x ρ ∂x ρA

Figure 2.3: Control volume diagram of a conduit for the momentum equation

Source: A review of water hammer theory and practice [5]

2.1.3 The Pressure (Water Hammer) Wave Speed

The water hammer wave speed for a flow of slightly compressible fluid of a non-elastic
pipe is defined by Eq. 2.10

s
K
c= (2.10)
ρ

where K is the bulk modulus of elasticity of the fluid. When pipe walls are elastic,
the dynamic pressure waves are propagating both in the fluid and in the conduit walls.
The coupling between fluid and structure material, also known as the fluid structure
interaction (FSI), is typical for the water hammer phenomena in an elastic pipe. FSI
is defined by the interaction of some deformable structure (pipe) with an internal fluid
flow. It is necessary to take FSI into consideration which is done in the classical water
hammer equations by the definition of the water hammer wave speed which will be
derived below.
Chapter 2. Background and Theory 9

By expanding the terms in the parentheses and using the definition of total derivatives
the continuity equation, Eq. 2.5, becomes Eq. 2.11:

1 Dρ 1 DA ∂V
+ + =0 (2.11)
ρ Dt A Dt ∂x

Since the pressure changes will cause the changes of the cross-sectional area, it will be
desirable to express the time derivative of the cross-section area in terms of the pressure.
By using the linear elasticity model for a circular pipe with internal pressure and the
assumption that the axial stresses are zero, the strain-stress relationship can be written
as follows:

σ
ε= (2.12)
E

where ε = hoop strain, σ = hoop stress and E = Young’s elasticity modulus. The hoop
stress (or circumferential stress) in the following thin-walled pipe model can be written
as:

F PD
σ= = (2.13)
A 2δ

where P = internal pressure, D = internal pipe diameter and δ = pipe wall thickness.
The time derivative of Eq. 2.13 can be expressed as:

 
dσ d PD P dD D DP
= = + (2.14)
dt dt 2δ 2δ dt 2δ Dt

Combining Eqs. 2.12 and 2.14 yields,

dε P dD D DP
E = + (2.15)
dt 2δ dt 2δ Dt

dD dε
=D (2.16)
dt dt

Since Eq. 2.16 is true, the time derivative of the strain can be expressed as Eq. 2.17.

 
dε D DP PD
= E− (2.17)
dt 2δ Dt 2δ
Chapter 2. Background and Theory 10

The time derivative of the cross-section area can be expressed in terms of the strain as
in Eq. 2.18 which follows:

 
πD2
1 DA 1 d 4 1 dD dε
= πD2
= 2
2D =2 (2.18)
A Dt dt D dt dt
4

By combining Eq. 2.17 with Eq. 2.18 the following equation is created:

1 DA 2D DP
= (2.19)
A Dt 2δE − P D Dt

The deformations of the pipe (i.e. the cross-sectional area changes) in Eq. 2.19 are now
completely expressed by the pressure changes for a circular pipe with internal pressure.

To continue the rewriting of the continuity equation, the first term in Eq. 2.11 can be
expressed in terms of pressure time derivative as:

1 Dρ 1 dρ DP 1 1 DP 1 DP 1 DP
= = dP = dP
= (2.20)
ρ Dt ρ dP Dt ρ dρ Dt dρ/ρ
Dt K Dt

when Eqs. 2.11, 2.18 and 2.20 are combined, following Eq. 2.21 yields,

 
1 2D DP ∂V
+ · + =0 (2.21)
K 2δE − P D Dt ∂x

From Eq. 2.10 the water hammer wave speed in the case of an elastic pipe is defined by
Eq. 2.22

1 K
c2 =   =  (2.22)
1 2D K 2D
K + 2δE−P D ρ 1+ E 2δ−P D/E ρ

In normal engineering applications, P D/E  2δ implying that Eq. 2.22 may be written
as

K
ρ
c2 = (2.23)
+ KD

1 Eδ

In 1963, Halliwell [6] presented the following general expression for the classical water
hammer wave speed
Chapter 2. Background and Theory 11

s
K
c= (2.24)
ρ 1+ K

E ψ

where ψ is a nondimensional parameter that depends on the elastic properties of the


conduit. Expressions for ψ for various conditions important to this study are the fol-
lowing:

Rigid Conduit:

Fully rigid conduit that cannot move.

ψ=0 (2.25)

Two cases for the anchoring of the conduit against longitudinal movement are:

Thin-Walled Elastic Conduit:

(1) Conduit anchored against longitudinal movement throughout its length

D
ψ= (1 − ν 2 ) (2.26)
δ

and (2) the conduit with frequent expansion joints that was derived earlier.

D
ψ= (2.27)
δ

2.1.4 The Classical Water Hammer Equations

By combining the new defined wave speed from Eq. 2.23 with Eq. 2.21, following
equation yields,

DP ∂V
+ ρc2 · =0 (2.28)
Dt ∂x

Note that Eq. 2.28 is for a conduit with expansion joints due to the assumption that
the axial stresses are zero.

The usage of a small Mach number (M = V /c) approximation recognized by Allievi [7]
makes it possible for a final simplification of the momentum Eq. 2.9 and the continuity
Eq. 2.28 to the classical water hammer equations represented by Eqs. 2.29 and 2.30:
Chapter 2. Background and Theory 12

∂V 1 ∂P τw (t)πD
+ · + =0 (2.29)
∂t ρ ∂x ρA

∂P ∂V
+ ρc2 · =0 (2.30)
∂t ∂x

The classical water hammer equations are valid for axi-symmetric flow of a slightly
compressible fluid in a flexible pipe, where the Mach number is very small and τw (t) =
quasi-steady wall shear as a function of time [5].

More rigorous derivation of the classical water hammer equations than the one presented
here can be found in [8], [9] and [10].

2.1.5 Unsteady Friction

There exists different ways to model the friction between the pipe and the fluid. In
the third term of the momentum Eq. 2.29, it is valid to relate the wall shear to cross-
sectional average velocity in steady state condition. Weather it is valid under unsteady
condition, as for the water hammer transient, is questionable. The attenuation of the
water hammer is highly dependent on a correct friction between the pipe and fluid. It
has been shown in previous research that the use of steady state wall shear models for the
unsteady problem of water hammer is satisfactory for the first pressure pulse and in some
cases for the second but the oscillations coming afterwards show very slow attenuation
as compared to that measured in experiments. This fact does not cause problems for
determining the maximum or minimum pressures in the system. Since the interest for
this study lies in the approach to the first maximum pulse peak, it was decided to use
the Darcy-Weisbach equation. The Darcy-Weisbach equation is commonly used in water
hammer models and is described by Eq. 2.31,

ρf (t)|V (t)|V (t)


τw (t) = (2.31)
8

where f (t) = Darcy-Weisbach friction factor. Several unsteady friction models have
been proposed and tested for better agreement of the water hammer attenuation. More
about unsteady friction methods and models can be found in Appendix A.
Chapter 2. Background and Theory 13

2.2 Direct Mass Flow Measurement

The flow meters that are commercially available today has a typical time resolution
of 0.1 seconds or slower. The thesis included a literature review to identify a possible
direct method to measure the transient mass flow rate in the test rig. Already performed
RELAP5 simulations revealed an estimation of the time resolution needed, in order to
accurately measure the mass flow, to be faster than 0.1 ms, i.e. higher than 10kHz.

Since the interest in this research is the water hammer phenomenon, the measurement
method should preferably have no impact on the pressure wave propagation, regarding
both the reflection and transmission of the wave and its velocity. The literature review
did not reveal a way to measure the mass flow with the required time resolution.

The commercially electromagnetic flow meters (EMF) today have their main focus on
the uncertainty of the measurement not on the time resolution. In theory the physical
restriction of time resolution governing the EMF, (i.e. the speed of the moving electrons)
should not be a limitation.

2.2.1 Electromagnetic flow meters

The electromagnetic flow meter (or the magnetic flow meter) is based on Faraday’s law
of electromagnetic induction. Faraday’s law describes the way an electromagnetic field
may be generated by a changing magnetic environment. The measurement principal is
based on two coils creating a magnetic field perpendicular to the flow direction. The
magnetic flux density (B) have to be known. The magnetic field is generated in pulses.
For the magnetic flow meter to work it requires a conducting fluid, for example, in
the case of the experiment here, water that contains ions. The ions corresponds to
conductors, flowing through the magnetic field changing it and creating a voltage (ES ).
The induced voltage is created over the flow which can be detected by two measuring
electrodes. The electrodes read the induced voltage while the magnetic field is held
constant. A simplified expression for the voltage, proportional to the flow velocity, is
given by Eq. 2.32,

ES = kBDe v (2.32)

De = the distance between electrodes (inside diameter of the tube), v = mean velocity
of the liquid and k = dimensionless constant [11].
Chapter 2. Background and Theory 14

Figure 2.4: Overview of the important principals of an EMF

Source: Transactions Volume 4, Flow and Level measurement[11]

The commercially available electromagnetic flow meters do not have a fast enough time
resolution to measure the transient flow rate, which is generated by the water hammer.
This is due to an averaging of the original signal in the transmitter in order to reduce
the uncertainty. By bypassing the signal processing and taking the original raw signal,
a faster time resolution could possibly be achieved. This will increase the uncertainty of
the measurement but it is not clear how large the ”noise” level on the flow measurement
will be.

2.3 Gibson’s indirect method

The Gibson method, also known as the pressure-time method, for transient flow mea-
surement is an indirect measurement method for the mass flow. The Gibson method was
studied and considered for of the water hammer validation of the RELAP code. The
method calculates the mass flow transient via the time-history of pressure difference
measurements (provided from the experiment) by Eq. 2.33.

Z t2
A
Q0 = (∆p(t) + ∆Pf (t))dt + ql (2.33)
ρL t1

Q0 = the flow rate in the initial steady-state conditions, ∆p = the static pressure differ-
ence between the measuring sections, ∆Pf = the pressure drop between the measuring
sections caused by hydraulic resistance, ql = the flow rate in final steady-state conditions
(which is the leakage flow through the valve), t1 = the initial time limit of integration
and t2 = the final time limit of integration. Vattenfall Hydro power provided a code
using the Gibson method, which had been used successfully for slower water hammer
Chapter 2. Background and Theory 15

transients in hydro power plants applications. In slower water hammer transients, the
fluid compressibility and pipe deformations can be neglected. These assumptions are
made when deriving Eq. 2.33. In the fast water hammer transient considered in this the-
sis, there will be effects from pipe deformations and fluid compressibility on the pressure
oscillations. The simplifications done in the Gibson method are included in the classical
water hammer equations in a special manner through the water hammer wave speed
presented earlier. The Gibson theory and code was studied here but it was decided not
to be used for fast transients due to the above reasons [12], [13].
Chapter 3

Method of Characteristics

In this chapter the equations for simulating the transient flow behavior in pipes with
the method of characteristics are derived together with the boundary conditions used
for producing the results.

Water hammer models are improving and becoming more widely used. Today, several
commercially available water hammer software packages exist using different numerical
methods. The method of characteristics (MOC) is the most popular numerical method
when it comes to one dimensional, hydraulic transient problems, especially if the pressure
wave velocity is constant as in the present case. It is desirable due to its attributes of
accuracy, simplicity, numerical efficiency, and programming simplicity [5].

3.1 Solution by Method of Characteristics

The continuity Eq.3.1 and momentum Eq.3.2 together with Eq. 2.31 derived in the
previous chapter describe transient flows in closed conduits. In these equations, distance
x and time t are two independent variables and pressure p and flow velocity V are two
dependent variables. The other variables are the system parameters which are assumed
not to vary with time.

∂V 1 ∂P f |V |V
L1 = + · + =0 (3.1)
∂t ρ ∂x 2D

∂P ∂V
L2 = + ρc2 · =0 (3.2)
∂t ∂x

16
Chapter 3. Method of Characteristics 17

A linear combination of Eqs. 3.1 and 3.2 is L = L1 + λL2 , where λ is a linear multiplier.
By rearranging the linear combination the following equation yields,

   
∂V ∂V ∂P 1 ∂P f |V |V
+ ρc2 λ · +λ + + =0 (3.3)
∂t ∂x ∂t ρλ ∂x 2D

The total derivatives for the flow speed (V = V (t, x)) and pressure (P = P (t, x)), if
they depend on both the time and the position, as in this case, are represented by the
following equations,

dV ∂V ∂V dx
= + (3.4)
dt ∂t ∂x dt
and
dP ∂P ∂P dx
= + (3.5)
dt ∂t ∂x dt

From Eqs. 3.3, 3.4 and 3.5 the unknown multiplier is solved as follows,

1 dx
= = ρc2 λ (3.6)
ρλ dt
which is
1
λ=± (3.7)
ρc

By combining Eq. 3.3 with the other three Eqs. 3.4, 3.5 and 3.7 following equations are
derived,

dV 1 dP f |V |V
+ + =0 (3.8)
dt ρc dt 2D
if
dx
=c (3.9)
dt
and
dV 1 dP f |V |V
− + =0 (3.10)
dt ρc dt 2D
if
dx
= −c (3.11)
dt

What MOC basically does is that it eliminates the independent variable x and the
partial differential equations gets converted into ordinary differential equations (ODE:s)
in the dependent variable t. The ODE:s are Eq.3.8, 3.9, 3.10 and 3.11. They are called
compatibility equations. However, this simplification makes a restriction on where the
Chapter 3. Method of Characteristics 18

new ODE:s are valid. The original momentum and continuity equations were valid
everywhere in the x-t plane, but when it comes to the new ODE:s they are only valid
along the straight lines (if the wave speed is constant) described by the governing linear
equations, Eq.3.9 and Eq.3.11 (with the slopes plus/minus the wave speed) received
when the ODE:s were derived. These lines are called characteristic lines.

The derived ODE:s are valid along the pipe length (i.e. for 0 < x < L) and special
boundary conditions are required at the ends, i.e. at x = 0 and at x = L. Assume the
valve is rapidly closing. This reduces the flow at the valve to zero which increases the
pressure at the valve. Since the pressure increase, a positive pressure wave (the pressure
is higher behind the wave front than that in front) travels in the upstream direction.

Figure 3.1: Characteristic lines in x-t plane.

To compute the pressures and flow velocities, they need to be known at the initial
conditions or calculated during the previous time step. To compute the dependent
variables at next time step, which is going to be along a characteristic line, Eqs. 3.8
and 3.10 need to be multiplied by dt and integrated. The integration limit is going to
be two points along the positive characteristic line (A-G in Fig.3.1) for Eq.3.8 and two
points along the negative characteristic line (B-G in Fig.3.1) for Eq.3.10. The friction
losses in the third term are not linear and therefore cause a problem. By a first order
approximation, in other words, the flow velocity remains constant on the characteristic
line, the evaluation of the integral becomes possible. This first-order approximation
usually yields satisfactory results for engineering applications. If the results are not
sufficient, one can reduce the time step to a smaller value and improve the contribution
from the nonlinear part.

1 f
(VG − VA ) + (PG − PA ) + ∆tVA |VA | = 0 (3.12)
ρc 2D
Chapter 3. Method of Characteristics 19

and
1 f
(VG − VB ) − (PG − PB ) + ∆tVB |VB | = 0 (3.13)
ρc 2D

The ∆t in Eq.3.12 is replaced by Eq.3.9 (in finite form) and the ∆t in Eq.3.13 is replaced
by Eq.3.11 (in finite form) creating the following equations,

1 fc
(VG − VA ) + (PG − PA ) + ∆xVA |VA | = 0 (3.14)
ρc 2D
and
1 fc
(VG − VB ) − (PG − PB ) − ∆xVB |VB | = 0 (3.15)
ρc 2D

These equations above can be written in a more programming friendly manner, where
index i is the space variable and index j is the time variable as can be seen in Fig. 3.2

Figure 3.2: Computational mesh/grid with index for MOC

1 fc
Vi,j − Vi−1,j−1 + (Pi,j − Pi−1,j−1 ) + ∆xVi−1,j−1 |Vi−1,j−1 | = 0 (3.16)
ρc 2D
and

1 fc
Vi,j − Vi+1,j−1 − (Pi,j − Pi+1,j−1 ) − ∆xVi+1,j−1 |Vi+1,j−1 | = 0 (3.17)
ρc 2D
Chapter 3. Method of Characteristics 20

By rearranging Eq.3.16, it will constitute the boundary condition at x = L and by


rearranging Eq.3.17, it will constitute the boundary condition at x = 0. By combing
Eq.3.16 with Eq.3.17, all the nodes in between the boundaries can be calculated through
Eqs.3.18 and 3.19

1
Pi,j = (Pi−1,j−1 + Pi+1,j−1 )+
2
ρc ρc f c fc
(Vi−1,j−1 − Vi+1,j−1 ) − ( ∆xVi−1,j−1 |Vi−1,j−1 | + ∆xVi+1,j−1 |Vi+1,j−1 |)
2 2 2D 2D
(3.18)
and

1 1
Vi,j = (Pi−,j−1 − Pi+1,j−1 )+
2 ρc
1 1 fc fc
(Vi−1,j−1 + Vi+1,j−1 ) + ( ∆xVi+1,j−1 |Vi+1,j−1 | − ∆xVi−1,j−1 |Vi−1,j−1 |)
2 2 2D 2D
(3.19)

To summarize, the model utilized to produce results uses a fixed-grid MOC. The com-
putational grid uses a constant ∆x (here spatial grid size) and constant ∆t (time step)
throughout the numerical simulation together with constant pressure wave speeds prop-
agating along the characteristic lines [8].

3.1.1 Different MOC schemes for more complex systems

The fixed-grid MOC requires all pipes to satisfy the Courant condition in Eq.3.20. The
Courant number (Co), is defined as the ratio of the actual wave speed and the numerical
wave speed (∆x/∆t).

c c · ∆t ∆x
Co = = = {∆t = }=1 (3.20)
∆x/∆t ∆x c

In multiple pipe systems or systems with variable wave speeds, problems occurs. Wave-
speed or pipe-length adjustments and/or interpolations have to be applied to insure
reasonable results. The down side with the adjustments are the introduction of errors
that follows them. A wave speed adjustment can be thought of as an adjustment of one
or more of the physical parameters that are used to calculate the wave speed (Eq.2.24),
for example the wall thickness, the elastic modulus or the density. The adjustment
directly creates an artificially faster or slower wave speed, causing different timing of
interactions and reflections of the pressure waves in the system. In other words, it
misleads the physical characteristics of the problem [14].
Chapter 3. Method of Characteristics 21

There are different interpolation schemes that can be used with advantages and disad-
vantages. There exists both possibilities to make interpolations in the space-line and
in the time-line. Space-line interpolation approximates the solution to the fixed-grid
MOC. By interpolating the dependent variables (the pressures and flow velocities) from
adjacent grid locations to the point where the characteristic line intersects the previous
time step, the solution is given. The same procedure applies for the time-line.

Several technical articles have been written on the matter of interpolation to deal with
adjustable time steps and fluctuations of the wave speed. Different interpolations and
methods have been tested to decrease the errors that occur, linearly, implicit methods,
Hermite scheme, Holly-Preissmann schemes and spline interpolations [15], [16] and other
approaches as an algorithm to derive a flexible class of discretization approaches for
interpolating between grid points to obtain values of the dependent variables (V and P)
at the start of characteristic curves [17]. One of the articles highlight the adjustment of
the wave speed because of the fact that all interpolation methods alter the wave speed
of the water-hammer model [14].

3.2 Boundary Conditions

The simplest boundary conditions are specified values of the relevant variables, the
pressure and flow velocity in this case. However, many practical boundary conditions
are specified as time-dependent pressure, velocity or pressure-velocity functions. The
two boundary conditions used to produce the results are going to be derived in this part
of the report.

As mentioned before Eqs. 3.18 and 3.19 are valid along the pipe length (i.e., for 0<x<L)
and special boundary conditions are required at the ends, i.e., at x = 0 and at x = L.

The first boundary condition expressed in a more programming friendly manner that
more easily can be implemented in code, similar to Eqs. 3.16, 3.17, 3.18 and 3.19, is
represented by a tank with constant pressure.

The boundary condition is called Ideal Pipe Attachment to Pressure Vessel and is rep-
resented in Fig.3.3.

The equation describing the flow for the boundary condition is Eq.3.21


Vi,j
P0 = Pi,j + (3.21)
2

By rearranging Eq.3.17, Eq.3.22 yields


Chapter 3. Method of Characteristics 22

Figure 3.3: Boundary condition, Ideal Pipe Attachment to Pressure Vessel

fc
Pi,j = Pi+1,j−1 + ρc(Vi,j − Vi+1,j−1 ) − ρc ∆xVi+1,j−1 |Vi+1,j−1 | (3.22)
2D

By inserting Eq.3.22 into Eq.3.21 and then rearrange, Eq.3.23 yields

2 2 f c2
Vi,j + 2cVi,j − 2cVi+1,j−1 + (Pi+1,j−1 − P0 ) − ∆xVi+1,j−1 |Vi+1,j−1 | = 0 (3.23)
ρ D

By solving the second order equation that is Eq. 3.23, ignoring the negative square root
and after rearranging, the following equation yields

s
2Vi+1,j−1 2 f
Vi,j = −c + c 1+ − (Pi+1,j−1 − P0 ) + ∆xVi+1,j−1 |Vi+1,j−1 | (3.24)
c ρc D

Eq. 3.24 is used for the boundary at x = 0.

The second boundary condition derived for direct implementation in code, is representing
a valve. The name of the boundary condition is Valve at Discharge End of Pipe and is
illustrated in Fig.3.4
Chapter 3. Method of Characteristics 23

Figure 3.4: Boundary condition, Valve at Discharge End of Pipe

Fig.3.4 is described by Eq.3.25


Vi,j
Pi,j − Pend = KL (3.25)
2

Valve boundary conditions are often coupled to opening or closing transients which deter-
mine the loss coefficient (KL ) as a function of time. The loss coefficient is approximated
by Eq.3.26, which is the flow velocity in the full pipe area A.

 2
A
KL (t) ≈ −1 (3.26)
Cv av (t)

where Cv av (t) is the vena-contracta area and Cv the contraction coefficient which is
assumed to be equal to 1 for a gently curved entrance.

The same approach is used for this boundary condition as for the previous one. Eq.
3.16, can be written as Eq. 3.27

fc
Pi,j = Pi−1,j−1 − ρc(Vi,j − Vi−1,j−1 ) − ρc ∆xVi−1,j−1 |Vi−1,j−1 | (3.27)
2D
Chapter 3. Method of Characteristics 24

2 2c 2c 2 f c2
Vi,j + Vi,j − Vi−1,j−1 − (Pi−1,j−1 −Pend )+ ∆xVi−1,j−1 |Vi−1,j−1 | = 0
KL (t) KL (t) ρKL (t) DKL (t)
(3.28)

Solving the second order equation above, ignoring the negative square root and after
rearranging, the following equation yields

c c
Vi,j = − +
KL (t) KL (t)
s
2KL (t)Vi−1,j−1 2KL (t) f KL (t)
1+ − 2
(Pi+1,j−1 − P0 ) − ∆xVi−1,j−1 |Vi−1,j−1 | (3.29)
c ρc D

Eq. 3.29 is used for the boundary at x = L.


Chapter 4

Experiments

As mentioned in the introduction, an experimental facility for the study of water hammer
transients has been built at Vattenfall AB, Generation R&D in Älvkarleby, Sweden. To
quantify the water hammer propagation in the pipe accurate measurements of both the
pressure and the mass flow are required. Such data will be used for validation of the
RELAP and the MOC codes.

Figure 4.1: Photograph of the test rig at Vattenfall R&D in Älvkarleby, Sweden

25
Chapter 4. Experiments 26

The test-rig that is displayed in Fig. 4.1 is a closed loop system, which consists of two
long approximately 30 m straight pipes (DN100) connected with a 180 degrees bend.

Figure 4.2: Overview of the important components of the test rig

The flow is created by a centrifugal pump, which is rpm-regulated on one of the electro-
magnetic flow meters, assuring a constant flow rate. The flow runs counterclockwise in
the test rig, as illustrated in Fig. 4.2 by the big arrows on the left hand side.

Before the pump a rapidly closing valve is situated. It is kept open with pressurized air,
and closes mechanically with a spring. The generation of the water hammer appears
when the valve closes at the same time as there is a constant flow rate through the pipe
system. A positive pressure pulse is travelling left in the figure from the valve and a
negative to the right. The idea is that the straight pipe sections along with the 180
degree are long enough for the valve to be able to close before the reflection of the wave
comes back and interfere with the first pressure pulse. The valve has been shown to
have a minimum closing time of about 30ms. Slower closing times can be selected by
varying the pressure drop in the pneumatic system. The sonic speed in water calculated
from the experiments is about 1250m/s. The wave therefore travels about 37.5m before
the valve is fully closed, which implies that the requirement is reached (2 x 30m).

The valve has an asymmetric configuration: it has an inner diameter of 83 mm with two
different expansions that reach a diameter of 100 mm. In between the valve and the
pump there is a conical contraction from DN100 to DN50, illustrated in Fig. 4.2. The
valve can be set to different angles manually.
Chapter 4. Experiments 27

Figure 4.3: The fixed valve in the test-rig.

To be able to model the characteristics of the valve, the area as a function of the angular
position (α) was measured using a potentiometer connected to the valve. The angle α, is
the angle which the valve rotates around, and when it reaches approximately 75 degrees,
the valve is completely closed.

The pipe surface roughness has been measured and it is commercial standard about
3.5µm±10 %. The test water used in experiments is soften water. It is important for
the function of the flow meters that the water has a large enough conductivity. The
system pressure is produced and controlled by using Nitrogen gas (N2 ). The N2 -tube is
coupled to the top of the reservoir. The system pressure in the gas is applied on top of
the water level in the reservoir and hence in the entire pipe system.

The air initially contained in the reservoir and in the pipeline will be let out through
the degassing valves in the top of the upper pipe line during the filling of the system
with water. To achieve an effective degassing of the air dissolved in the water, a vacuum
pump, is connected to the top of the tank. A manual valve is used to close the line
between the tank and the vacuum pump.

4.1 Measuring Devices and System

The measuring system used is a 2 x 16 channels DAQ system to sample all the signals
from the sensors. Eight channels can sample at 50 kHz and the rest at 300 kHz. A
sampling frequency of 20 kHz (i.e. ∆t =0.05 ms) is sufficient. Then one can resolve a
Chapter 4. Experiments 28

pressure wave which occurs during 1 ms with 10 points. A LabView program is used to
collect all the signals.

The measuring devices consist of flow meters, pressure sensors and strain gauges (in-
strument for measuring strain).

The uncertainty of the pressure sensors is ±0.25% of full range and the frequency re-
sponse is stated to be 180 kHz according to the manufacturer [18]. The sensor positions
and the names of the positions are listed in Table 4.1

Figure 4.4: Illustration of a pressure sensor model used

Position Name of position


Between valve and pump P1
Upstreams from the valve P2
Before 180 degree bend P3
After 180 degree bend P4
Between flow meter and tank P5
Before the S-shaped pipe near the tank P6
Before the S-shaped pipe near the tank P7
In the middle of the above pipe P8

Table 4.1: Pressure sensors

There are two electromagnetic flow meters installed after the pump. One of the flow
meter’s purpose is as mentioned to regulate the flow speed generated by the pump. The
other one is used for measurements and also initially to study the possibility of measuring
the transient behavior of the flow. The flow meters has a measurement uncertainty of
about 0.5 % according to the recent calibration at SP, see Appendix C. The velocity in
the DN50 flow meter is about 1.2 m/s. At flow rates lower than that, the uncertainty
in the flow measurement will increase.

Position Name of position


Between pump and tank Q1
Right after Q1 Q2

Table 4.2: Electromagnetic Flow Meters


Chapter 4. Experiments 29

Figure 4.5: The two electromagnetic flow meters at the DN50 pipe line between the
pump and the reservoir in the test-rig.
Chapter 5

Results and Discussion

This chapter contain analyses of results and associated uncertainties followed by a discus-
sion. Since the RELAP and the MOC codes are using different equations for the analyses
of water hammer (RELAP is using weighted Navier-Stokes equations and MOC is using
the momentum and continuity equations adapted for conduits with a special wave speed
term) there will be differences in the results. The differences have been compared and
analysed as the results from the two codes are validated against the experimental data.
The general results from the experiment together with the direct mass flow measurement
results will also be presented.

The ultimate objective for performing these simulations are to predict the loads on the
pipes. The force is proportional to the gradient of the mass flow in each segment. The
reason why the focus is only on the first change in mass flow and pressure is that this
usually represents the largest force.

5.1 Results

5.1.1 Experimental results

The experimental pressure measurements are displayed relative to the local pressure at
stationary flow before the water hammer. The positive water hammer wave, displayed
in Fig. 5.1, will travel upstream until it is reflected in the reservoir. The time it takes
for the wave to get fully damped out was more than 30 seconds. The noise on top of the
first positive pressure peak in Fig. 5.1 is approximately ±10 %. Where the noise comes
from will be brought up in the discussion section.

30
Chapter 5. Results and Discussion 31

A pressure wave (for an instantaneous valve closure) calculated with Joukowsky’s fun-
damental equation of water hammer should theoretically produce a pressure amplitude
of ρ · ∆V · c = 999 · 0.31 · 1450 = 4.48 bar (Eq. 2.1). The lower amplitude observed in Fig.
5.1 can be explained by the lower speed of sound, i.e. the water hammer wave speed. If
the water hammer only affects the water, the wave speed would be the speed of sound
in water (1483.8 m/s). However, the elastic pipe expands due to the increased pressure,
sending out a wave in the structure as well. The structural wave is approximately four
times faster than the one in the fluid. This causes an interaction between the fluid
and the structure, resulting in a slower wave speed than the sonic speed in water, see
Appendix B. With a wave speed of 1255 m/s (see Fig. 5.1), the amplitude is calculated
to 3.8 bar, which corresponds quite well with what is displayed, not taking into account
the ”noise”, in Fig. 5.1.
6
P2

4
Pressure [bar]

-2

-4

-6
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Tid [s]

Figure 5.1: The positive pressure wave propagating upstream from the valve.

The results from Fig. 5.1 constitutes the validation data for the MOC simulation.

The negative water hammer wave continuing in the flow direction after the valve clo-
sure through the pump and gets damped faster than the positive wave, Fig. 5.2 Note
also that the negative pressure wave does not display almost any noise and has higher
amplitude, see Fig. 5.2. The distance between the valve and the tank is shorter than
the wavelength of the wave (approximately 9 m). Consequently, the reflected part of
the wave is superimposed with the part of the wave that have not yet been reflected
and the wave amplitude will therefore be higher and more ”sinusoidal” in its shape as
is illustrated in Fig. 5.2
Chapter 5. Results and Discussion 32

6
P1

Pressure [bar] 4

-2

-4

-6
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5
Tid [s]

Figure 5.2: The negative pressure wave propagating downstream from the valve.

5.1.1.1 Direct Mass Flow Measurement

The first measurements were taken in the DN50 pipe of the negative wave between the
valve and the reservoir (more exact between the pump and the reservoir). The EMF
used for the measurement was in position Q2.

The EMF shifted magnet field is to ensure that the parts do not get magnetized. This
shift is represented in Fig. 5.3 by the blue peaks. During the magnet field shift the volt-
age measurements proportional to the velocity of the fluid will be incorrect. Furthermore,
the measured voltage has an offset which will be needed to take into consideration when
calculating the corresponding flow velocity and mass flow.
0.3
Q2 original
Q2 filter
0.2 MagnField

0.1
Q2 [V]

-0.1

-0.2

-0.3
1.9 2 2.1 2.2 2.3 2.4 2.5
Time [s]

Figure 5.3: Voltage measurements proportional to the flow velocity together with
filtered and magnet field.
Chapter 5. Results and Discussion 33

0.03
Q2 original
0.02 Q2 filter

0.01

0
Q2 [V]

-0.01

-0.02

-0.03

-0.04

-0.05
1.9 2 2.1 2.2 2.3 2.4 2.5
Time [s]

Figure 5.4: Voltage measurement together with the filtered one.

Some embedded MATLAB function for filtering was tried before the filter used in Figs.
5.3 and 5.4 were chosen. The filtered version of the raw voltage signal is using a Savitzky-
Golay smoothing filter of the embedded MATLAB function sgolayfilt(x,k=3,f=41), where
x = the unfiltered data, k = The polynomial order and f = the frame size. In Fig. 5.4,
the magnetic field has been removed and the voltage measurement together with the fil-
tered voltage has been zoomed in around the time of the transient. The direct unfiltered
voltage is noisy. After filtering it appears less noisy, and the important first gradient
induced by the water hammer wave seem to be caught arounf t = 2.08 s. A deeper
investigation into weather the voltage measurement may be to rely on was done.

As stated in Table 4.1, P1 is in between the valve and the pump. Hence, there will be
a time shift in the pressure measurement compared to the EMF measurement from Q2.
This is why the P1 measurements has been shifted relative to Q2 so that the fist water
hammer wave ends up simultaneously for both as displayed in Fig. 5.5.

Fig. 5.5 display the filtered mass flow.

ṁ = V · A · ρ (5.1)

The procedure from the filtered voltage in Fig. 5.4 to the corrected mass flow in Fig. 5.5
is the following: The voltage is firstly corrected for the offset creating the symmetric shift
around zero. The negative voltage will then be flipped so that their will be a continuously
positive voltage. This will create an amplitude between zero and the positive voltage.
During steady state conditions, the amplitude will be approximately constant. The
direct measured mass flow from the other EMF (Q1) will produce the correct mass
Chapter 5. Results and Discussion 34

6
Q2 filter
P1
4

-2

-4

-6
2 2.05 2.1 2.15 2.2 2.25 2.3
Time [s]

Figure 5.5: Filtered voltage together with time shifted pressure.

flow and can with the amplitude therefore work as the proportionality for the voltage.
Eq. 2.32 explain the the relationship between the voltage and the flow velocity. Eq.
5.1 describes the relationship between flow velocity and mass flow. With the mean
velocity of the fluid (v) equal to the cross-sectional average velocity(V ) Eq. 5.1 can be
substituted into Eq. 2.32 and Eq. 5.2 yields.

kBDe
ES = kBDe v = ṁ = K ṁ (5.2)

Since A and ρ are assumed constant in Eq. 5.1, and B and D likewise are assumed
constant in Eq. 2.32 they will together form a constant (K) in Eq. 5.2.

A good correlation between the pressure and the velocity change can be seen. The
change in pressure happens almost simultaneously with the change in velocity. These
results indicate that it is possible to measure fast mass flow transients using EMF.

5.1.2 Comparison between the RELAP and the MOC codes

A code-to-code comparison was investigated. The same boundary conditions are used
for RELAP and MOC. The difference to address the water hammer phenomenon exist
within the equations and the numerical methods used by the different models.

The MOC-model that was used to produce results is presented in Appendix B together
with a flowchart for the code. The model was simplified with respect to the test rig
and consists of only a reservoir and a pipe segment connecting the reservoir with a fast
closing valve. Furthermore, to build up the initial values for the transient simulation, a
Chapter 5. Results and Discussion 35

pressure difference was set so that the reservoir corresponds to a position upstream and
the valve correspond to the downstream. Even though differences exist between the two
models, the results are expected to be similar.

The MOC model is using the same wave speed as the RELAP code calculated in its
simulation (see speed of sound in water in Table B.1). The results of the pressure for
the the MOC and the RELAP codes are shown in Fig. 5.6.

Figure 5.6: Pressure normalized by Joukowsky’s equation.

Two MOC models were implemented. The pipe deformations (FSI) and fluid com-
pressibility in the classical water hammer equations is implemented in a special manner
through the water hammer wave speed derived in Eq. 2.23. The MOCF SI model uses
the adjusted water hammer wave speed that was measured from the experiment (see
Appendix B).

Fig. 5.7 displays a small difference between the two MOC models for the first gradient
of the pressure.

The slopes of the flow rate curves in Fig. 5.8 which are all equally proportional to the
mass flow will give the force acting on the pipe. The following equation is used for
calculating the force on the pipe:

Z L
dṁ
F = dx (5.3)
0 dt

In the RELAP simulation, each segment will calculate a mass flow change and be
summed up creating the total force acting on the entire pipe as in Eq. 5.3.
Chapter 5. Results and Discussion 36

Figure 5.7: Pressure normalized by Joukowsky’s equation with MOCF SI .

Figure 5.8: Flow rate normalized

The graph in Fig. 5.8 clearly displays a faster flow rate change for the two MOC models
than for RELAP. Consequently, the force will be larger with the data from the MOC
code compared to the RELAP code.

5.1.3 Validation of the MOC and RELAP codes using experimental


data

Which one of the two methods MOC and RELAP that can produce results which rep-
resents real water hammer events best is further investigated below.
Chapter 5. Results and Discussion 37

A good indication of the validity of the model’s mass flow gradient is to compare the
pressure with the experimentally measured pressure. However, a better validation of
the forces would be to compare the simulated mass flow direct within the experimental
measured mass flow.

Figure 5.9: Pressure normalized by Joukowsky’s equation with experimental data.

The deviations in the data from the experiment and the simulation can partly be ex-
plained by the fact that the models do not include the 180 degree bend that would
induce most of the FSI (which is anyway not properly included in the present model),
insufficient degassing (there still exists some air in the pipe during experiments) and
finally the violent valve closure inducing FSI-effects, causing the ”noise” around t=0.1s.

5.2 Discussion

More about potential improvements of the model will be discussed in the section below.

5.2.1 Extended MOC

The classical water hammer equations used in the MOC model here can be extended to
better model specific physical phenomena and give better agreement with experimental
results. There exists four different phenomena which may affect the results: viscoelastic-
ity (VE), column separation (CS), unsteady friction (UF) and fluid structure interaction
(FSI). In the present case UF and FSI are relevant. FSI is present in the experiments
because of the elastic pipe, which is moreover not fully anchored to the ground. There
exists ”noise” in the pressure data resembles FSI effects shown in previous research.
Chapter 5. Results and Discussion 38

UF is induced due to the unsteady nature of water hammer transients. Furthermore,


the discrepancy between the long-time damping of the pressure in simulation and the
experiments is an effect of UF.

Consequently, fluid structure interaction and unsteady friction could be implemented to


improve the model and the results. Since this work focus on the first pressure gradient
UF is not relevant. UF does not have a large impact on the first pressure gradient but
is more important for the attenuation of the pressure wave in the long run. FSI on the
other hand has a quick response and is therefore relevant for the first pressure gradient.

A relevant FSI simulation for the present case would include Poisson coupling and Junc-
tion coupling. There also exists Friction coupling, but it is insignificant due to the fact
that it is much smaller than the other two, [19]. The Poisson coupling take in considera-
tion the physical phenomenon that originates from the radial expansion of the pipe due
to the force caused by the fluid pressure. The radial change of the structure translates
to waves propagating in the axial direction of the pipe. The fluid pressure gives rise to a
hoop stress in the pipe which is converted to axial stress waves by virtue of the Poisson
ratio coefficient.

The Junction coupling describes the physical phenomenon at discrete locations such as
elbows and valves during the transient. The local forces at these points give rise to
displacements of the structure. The created motions can be considered as ”pumping”
actions generating pressure waves in the fluid.

The modeling of these phenomena create two more unknown dependant variables, in
addition to the flow velocity and the pressure, together with two pipe equations for
axial motion from beam theory. The unknowns are the velocity of the pipe (the time
derivative of the pipe displacement) and the axial pipe stress [20], [21].

These four unknowns can all be solved by using MOC as well. Time-line interpolations
near boundary conditions would be necessary for an extended MOC simulation together
with an adjustment of the structural wave speed by changing the density of the pipe to
be able to fit the computational grid [22].

According to [23], Junction coupling has been proved to capture the most relevant of
FSI-effects. Junction coupling only exists at boundaries and since the valve and the
reservoir are fixed to the ground in the experiment, the only junction coupling here
would be in the 180 degree bend. This is the foundation of the belief that the best way
to improve the model is to build in a 180 degree bend with a junction coupling.
Chapter 6

Conclusions

The overall objective of Vattenfall’s thermohydraulic load program is to assess the valid-
ity of the RELAP5/MOD3 code for water hammer transients. As a part of the program
the classical water hammer equations are also solved with an implemented MOC code.
The performance and validation of both codes are scrutinized against new water ham-
mer experiments performed in a new test rig at Vattenfall R&D in Älvkarleby, Sweden.
Additionally, an investigation of direct fast transient mass flow measurements for the
validation purpose was also carried out. The following are the main conclusions:

The results indicate that the newly implemented MOC code produces more accurate
results and that the RELAP5/MOD3 code underestimates the mass flow gradient and
therefore the loads on the pipe.

The pipe deformations (FSI-effects) and fluid compressibility dealt with in the classical
water hammer equations through the adjustment of the water hammer wave speed in
the MOC code (which are not present in RELAP5/MOD3), is of less significance.

The deviations in the data MOC data from the experimental results can probably par-
tially be explained by the fact that the MOC models does not include the 180 degree
bend that would induce most of the FSI-effects (which is anyway not captured in the
present model), the insufficient degassing (there still exists some air in the pipe in the
experiments) and finally the violent valve closure inducing FSI-effects.

Further work is recommended for improvements of the present MOC model e.g. the ex-
tended MOC with ”real” FSI-effects. Poisson coupling and most importantly a Junction
coupling at a 180 degree bend should be implemented. These improvements appears to
be promising for an even better agreement with the experimental validation results.

The direct mass flow measurement results using a ”stripped” electromagnetic flow meter
indicate the possibility of measuring the mass flow at fast transients such as here. The
39
Chapter 6. Conclusions 40

drawbacks are the uncertainty produced by the ”noise” (which require filtering) and
that a sufficient enough flow speed is required for detecting the transient.

In view of the results obtained from the direct mass flow measurements, further work in
assessments of the uncertainties is recommended. In particular, our results indicate that
the uncertainties for smaller flow velocities will naturally be larger (but is then also less
challenging to the pipe system) and a deeper uncertainty analysis should be performed.
Appendix A

Supplement on Water Hammer


and the Friction Term

The unsteady friction effects occurring at water hammer transients is ongoing research.
Several methods have been proposed and validated by experiments. Besides unsteady
friction (UF), there are three different items which may affect the classical water ham-
mer results: viscoelasticity (VE), column separation (CS) and fluid structure interaction
(FSI). Each of these items has been investigated by researchers as well. Combinations
of two or more items in analysis have been studied but not all different existing com-
binations. The technical paper [24], has listed several research articles of the different
combinations.

The wall shear in the friction term is considered as a sum of the steady and unsteady
wall shear, presented by Eq. A.1

τw = τSteady + τU nsteady (A.1)

where τU nsteady is zero for steady flow. The unsteady friction term attempts to represent
the induced changes in the velocity profile by the water hammer transient.

The methods proposed for dealing with the UF effects can be classified by two different
categories.

(1) Empirical-Based Corrections to Quasi-Steady Wall Shear Models

Most of the research on this subject is governing the instantaneous acceleration-based


unsteady friction models. Brunone et al. suggested an important modification of in-
stantaneous acceleration-based unsteady friction models. The Brunone et al. model

41
Appendix A. Supplement on Water Hammer and the Friction Term 42

has become the most widely used modification in water hammer application due to its
simplicity and its ability to produce reasonable agreement with experimental results.

(2) Physically Based Wall Shear Models

This class is based on the analytical solution of the flow equations. These models have
provided accurate simulation results of the attenuation phenomena, which is normally
connected to the unsteady friction. However, they are computationally heavy, and thus
have been used primarily for simple piping systems.

For more detailed information see [5].


Appendix B

Model Description

A simplified model of the experimental test rig was set up for implementation of the
MOC equations in MATLAB. It consisted of a reservoir and a pipe segment connecting
the reservoir with a fast closing valve. Furthermore, to build up the initial values for the
transient simulation, a pressure difference was set so that the reservoir corresponds to a
position upstream and the pipe beyond the valve correspond to the downstream end.

Description Value
Length of pipe 60 [m]
Gravitational constant 9.82 [m/s2 ]
Pipe diameter 0.1 [m]
Fluid density 998.6 [kg/m3 ]
Initial upstream pressure 1000700 [Pa]
Initial downstream pressure 1000000 [Pa]
Fluid density 998.6 [kg/m3 ]
Speed of sound in water 1483.8 [m/s]
Measured wave speed 1248.4 [m/s]
Friction factor 0.023
Water bulk modulus 2.1·109 [Pa]
Young’s elasticity modulus (180-210)·109 [Pa]
Pipe thickness 0.003 [m]
Number of nodes in MOC 601
Increment of x (∆x) 0.1 [m]
Increment of t (∆t) 8.01 · 10−5 [s]

Table B.1: Specification of parameters used in MOC

The density, pressures and speed of sound in water in Table B.1 were taken from RELAP
simulations. The other parameters are from the test-rig. Young’s elasticity modulus (E)
is somewhere between 180 and 210 GPa because the actual value is uncertain.

43
Appendix B. Model Description 44

The correct value of the water hammer wave speed is important in the MOC equa-
tions. The way to calculate the wave speed is Eq. 2.23. By implementing the needed
parameters from Table B.1, the wave speed was calculated to be 1183-1263m/s.

The Darcy-Weisbach friction factor (f(t)) can be adapted to the experimental data. The
reason is for a better agreement of the first pressure pulse with the measured pulse from
the experiments, as was done in previous work [25].

The friction factor was adapted to the value displayed in Table B.1. The value of the
friction factor produced an initial flow velocity similar to RELAP’s simulation. The
pressure difference builds up a flow velocity of 0.3045 m/s in steady state.

The pressure and mass flow was saved every 0.5 seconds in Figs. B.1 and B.2. The
steady state was decided to have been reached after 120 seconds. The relative mass flow
change in the pipe at that time were less than 0.1%. After 120 seconds the transient
was initiated.

Figure B.1: Pressure change in the middle of the pipe until steady state was reached

Due to the asymmetry that exists in the real valve, corrections were needed in the
model. Therefore, when reaching steady state the valve closes a little bit to deal with
the asymmetric cross-section area. This is the reason why the valve ratio starts at 0.67
in Fig B.3. It takes approximately 0.1 seconds, from pushing the close valve button,
until the valve ratio is 0 and no more water is going through the valve. The valve ratio
corresponds to the characteristics of the ball valve. The valve ratio displayed in Fig.
B.3 was used in Eq. 3.26, for calculating the loss coefficient (KL ) as a function of time.
Appendix B. Model Description 45

Figure B.2: Mass flow in the middle of the pipe change until steady state was reached

0.7

0.6

0.5
Valve ratio

0.4

0.3

0.2

0.1

0
0 0.02 0.04 0.06 0.08 0.1 0.12
Time [s]

Figure B.3: The valve ratio change used in valve boundary condition
Appendix B. Model Description 46

Start

Computational Constants
N, M, ∆t

Physical Constants
Ai , Di , fi , ρi , ci
i = 1, 2,...,N

Inital Values
Pi,j and Vi,j
At j = 1, i = 1, 2,...,N

j=2

i=1

yes no yes
i=1? i=N?

Left Boundary Condition: Left Boundary Condition:


Vi,j from Eq. 3.24, no Vi,j from Eq. 3.29,
Pi,j from Eq. 3.22 Pi,j from Eq. 3.27

General Mesh Point


Vi,j and Pi,j
From Eqs. 3.18 and 3.19

no
i = i+1
yes yes no
Stop i=N? j=M? i=N?

no j = j+1
yes
Appendix B. Model Description 47

Another method to decide the wave speed is by utilizing the experimental pressure
sensors. By calculating the time difference between the start of the transient in the two
signals from the pressure sensors and by knowing the distance between the two sensors,
the wave speed can be calculated:

Distance between sensors


c= (B.1)
Time difference between the start of the transient in the sensors

By using Eq. B.1, the wave speed was calculated to 1248.4 m/s. This was the wave
speed value used in the MOCF SI model.
Appendix C

Calibration protocol of
electromagnetic flow meter

The flow meters has a measurement uncertainty of about 0,5 % according to the cali-
bration displayed below at SP.

Figure C.1: Calibration protocol of electromagnetic flow meter at SP, 2015-04-21

48
Appendix C. Calibration protocol of electromagnetic flow meter 49

Figure C.2: Calibration protocol of electromagnetic flow meter at SP, 2015-04-21


Bibliography

[1] Nuclear power at vattenfall. http://corporate.vattenfall.


com/about-energy/non-renewable-energy-sources/nuclear-power/
nuclear-power-at-vattenfall/?_t_id=1B2M2Y8AsgTpgAmY7PhCfg%
3d%3d&_t_q=nuclear+power&_t_tags=language%3aen%2csiteid%
3afa59d14b-8a3f-4bf7-93c4-94c8a99859e7&_t_ip=192.36.28.76&_t_
hit.id=Kwd_Kestrel_Library_Epi_Types_Pages_InformationPageType/
_d203bf19-ce58-4318-a1f9-2eb213d3c30b_en&_t_hit.pos=1. Accessed:
2015-02-15, Last updated: 2015-08-31 16:02.

[2] RELAP5/MOD3.3 CODE MANUAL VOLUME II: USERS GUIDE AND INPUT
REQUIREMENTS. Information Systems Laboratories, 2010.

[3] RELAP5/MOD3.3 CODE MANUAL VOLUME V: USERS GUIDELINES. Infor-


mation Systems Laboratories, 2010.

[4] Allen R Inversin. Micro-hydropower sourcebook. 1986.

[5] Mohamed S Ghidaoui, Ming Zhao, Duncan A McInnis, and David H Axworthy. A
review of water hammer theory and practice. Applied Mechanics Reviews, 58(1):
49–76, 2005.

[6] AR Halliwell. Velocity of a water-hammer wave in an elastic pipe. Journal of the


Hydraulics Division, 89(4):1–21, 1963.

[7] Lorenzo Allievi. Theory of Water-hammer, volume 1. Typography R. Garroni,


1925.

[8] M Hanif Chaudhry. Applied hydraulic transients. Technical report, Springer, 1979.

[9] E Benjamin Wylie, Victor Lyle Streeter, and Lisheng Suo. Fluid transients in
systems, volume 1. Prentice Hall Englewood Cliffs, NJ, 1993.

[10] Frederick J Moody and DE Winterbone. Introduction to unsteady thermofluid


mechanics, 1991.

50
Bibliography 51

[11] Transactions volume 4, flow and level measurement. http://www.omega.com/


literature/transactions/volume4/t9904-09-elec.html. Accessed: 2015-05-
16.

[12] W Adamkowski A. Janicki. A new approach to calculate the flow rate in the
pressure-time method - application of the method of characteristics. The Interna-
tional Conference HYDRO 2013 (Workshop: Turbine Flow Measurement), Inns-
bruck, Austria. Presented: 7-9 October 2013.

[13] Norman Rothwell Gibson. The Gibson method and apparatus for measuring the
flow of water in closed conduits. American Society of Mechanical Engineers, 1923.

[14] Mohamed S Ghidaoui and Bryan W Karney. Equivalent differential equations in


fixed-grid characteristics method. Journal of Hydraulic Engineering, 120(10):1159–
1175, 1994.

[15] IA Sibetheros, ER Holley, and JM Branski. Spline interpolations for water hammer
analysis. Journal of Hydraulic Engineering, 117(10):1332–1351, 1991.

[16] David E Goldberg and E Benjamin Wylie. Characteristics method using time-line
interpolations. Journal of Hydraulic Engineering, 109(5):670–683, 1983.

[17] Bryan W Karney and Mohamed S Ghidaoui. Flexible discretization algorithm for
fixed-grid moc in pipelines. Journal of Hydraulic Engineering, 123(11):1004–1011,
1997.

[18] Pressure sensor - xp5. http://www.meas-spec.com/product/pressure/XP5.


aspx. Accessed: 2015-05-25.

[19] David C Wiggert and Arris S Tijsseling. Fluid transients and fluid-structure in-
teraction in flexible liquid-filled piping. Applied Mechanics Reviews, 54(5):455–481,
2001.

[20] AGTJ Heinsbroek. Fluid-structure interaction in non-rigid pipeline systems. Nu-


clear engineering and design, 172(1):123–135, 1997.

[21] AS Tijsseling. Exact solution of linear hyperbolic four-equation system in axial


liquid-pipe vibration. Journal of Fluids and Structures, 18(2):179–196, 2003.

[22] CSW Lavooij and AS Tusseling. Fluid-structure interaction in liquid-filled piping


systems. Journal of fluids and structures, 5(5):573–595, 1991.

[23] AS Tijsseling and CSW Lavooij. Waterhammer with fluid-structure interaction.


Applied Scientific Research, 47(3):273–285, 1990.
Bibliography 52

[24] Alireza Keramat, AS Tijsseling, Qingzhi Hou, and Ahmad Ahmadi. Fluid–structure
interaction with pipe-wall viscoelasticity during water hammer. Journal of Fluids
and Structures, 28:434–455, 2012.

[25] Hyoung-Jin Kim. Numerical and Experimental Study on Dynamics of Unsteady


Pipe Flow Involving Backflow Prevention Assemblies. PhD thesis, University of
Southern California, 2012.

Вам также может понравиться