08.12.2012 Views

Pressure Waves in Pipelines and Impulse Pumping - NTNU

Pressure Waves in Pipelines and Impulse Pumping - NTNU

Pressure Waves in Pipelines and Impulse Pumping - NTNU

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es <strong>and</strong><br />

<strong>Impulse</strong> Pump<strong>in</strong>g<br />

Physical Pr<strong>in</strong>ciples,<br />

Model Development <strong>and</strong><br />

Numerical Simulation<br />

Benjam<strong>in</strong> Pierre<br />

December 2009


iii<br />

63 o N, 10 o E


Abstract<br />

The primary purpose of the work was to describe the physical pr<strong>in</strong>ciples of<br />

impulse pump<strong>in</strong>g <strong>and</strong> <strong>in</strong>vestigate its potential applications <strong>in</strong> petroleum eng<strong>in</strong>eer<strong>in</strong>g.<br />

Investigations of the potential applications of impulse pump<strong>in</strong>g were<br />

based on analysis of head, flow rate <strong>and</strong> efficiency performances for pump<strong>in</strong>g<br />

of hydrocarbon fluids.<br />

The secondary purpose of the work was to analyse pressure wave propagation<br />

<strong>and</strong> attenuation models <strong>in</strong> fluid-filled production tub<strong>in</strong>gs. Additionally,<br />

one dimensional numerical simulation methods were <strong>in</strong>vestigated for accurate<br />

modell<strong>in</strong>g of pressure transient events <strong>in</strong> pipel<strong>in</strong>es, such as water hammer<br />

phenomena.<br />

<strong>Pressure</strong> wave propagation, transmission <strong>and</strong> reflection characteristics were<br />

first studied. F<strong>in</strong>ite volume methods were also studied for numerical simulation<br />

of one-dimensional pressure transient phenomena. Accuracy of pressure<br />

wave attenuation models was then analysed us<strong>in</strong>g a contradiction method,<br />

rely<strong>in</strong>g on analysis of numerical simulation tools for pressure transient events.<br />

<strong>Impulse</strong> pump<strong>in</strong>g physical pr<strong>in</strong>ciples were next described us<strong>in</strong>g fundamentals<br />

of pressure wave propagation. A numerical simulation tool was then developed<br />

to reproduce impulse pump<strong>in</strong>g physical pr<strong>in</strong>ciples, us<strong>in</strong>g a f<strong>in</strong>ite volume<br />

numerical scheme.<br />

<strong>Impulse</strong> pump<strong>in</strong>g performances were modelled afterwards <strong>in</strong> terms of lift<strong>in</strong>g<br />

heights, flow rates <strong>and</strong> efficiency, based on results from the developed numerical<br />

simulation tool. Potential applications of impulse pump<strong>in</strong>g <strong>in</strong> petroleum<br />

eng<strong>in</strong>eer<strong>in</strong>g were then analysed us<strong>in</strong>g the performance model.<br />

<strong>Impulse</strong> pump<strong>in</strong>g generates flow from bottomhole to wellhead us<strong>in</strong>g pressure<br />

waves generated at wellhead. Fluid can be transported from bottomhole to<br />

wellhead without theoretical lift<strong>in</strong>g height limitations us<strong>in</strong>g impulse pump<strong>in</strong>g.<br />

<strong>Impulse</strong> pump<strong>in</strong>g performances were illustrated <strong>and</strong> a range of petroleum<br />

eng<strong>in</strong>eer<strong>in</strong>g applications was <strong>in</strong>vestigated.<br />

<strong>Impulse</strong> pump<strong>in</strong>g greatest efficiency is obta<strong>in</strong>ed for artificial lift of water from<br />

v


shallow wells. <strong>Impulse</strong> pump<strong>in</strong>g performances depend on pressure wave amplitude,<br />

wellhead pressure, lift<strong>in</strong>g height, fluid compressibility <strong>and</strong> volume<br />

occupied <strong>in</strong> the production tub<strong>in</strong>g.<br />

vi


Acknowledgments<br />

I would like to express my gratitude to my supervisor, Professor Jon Ste<strong>in</strong>ar<br />

Gudmundsson, for the <strong>in</strong>sightful discussions, encouragement, support <strong>and</strong><br />

guidance throughout the work. His thirst for knowledge <strong>and</strong> his joyful will<strong>in</strong>gness<br />

to challenge have <strong>in</strong>itiated countless ”th<strong>in</strong>k<strong>in</strong>g out loud” sessions that<br />

have enlightened the path lead<strong>in</strong>g to the PhD degree.<br />

I acknowledge Clavis <strong>Impulse</strong> Technology AS for their f<strong>in</strong>ancial support <strong>and</strong><br />

for giv<strong>in</strong>g us access to some of their experimental data <strong>and</strong> for permission to<br />

use them <strong>in</strong> the present thesis. For this I would like to thank Peter Grubyi,<br />

Magomet Sagov <strong>and</strong> Jim Viktor Paulsen.<br />

I also acknowledge the f<strong>in</strong>ancial support from the Norwegian Research Council,<br />

through the Petromaks program, project 196054/S60, without which this<br />

project could not have been completed.<br />

The support <strong>and</strong> consideration of the Technology Transfer Office at <strong>NTNU</strong> is<br />

deeply appreciated. For this I would like to thank Erik Wold, Karl Kl<strong>in</strong>gsheim<br />

<strong>and</strong> Krist<strong>in</strong> Jørstad.<br />

I would like to thank Dr. Donna Calhoun <strong>and</strong> her colleagues at Commisariat à<br />

l’ Energie Atomique (CEA) Saclay, for help<strong>in</strong>g me with source term modell<strong>in</strong>g<br />

<strong>in</strong> CLAWPACK.<br />

I would like to thank Professor Helena Ramos <strong>and</strong> her colleagues at the Technical<br />

University of Lisbon, for the pleasant stay <strong>and</strong> valuable help on modell<strong>in</strong>g<br />

of unsteady friction.<br />

The technical staff at the Department of Petroleum Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Applied<br />

Geophysics have been valuable <strong>in</strong> ma<strong>in</strong>tenance, repair <strong>and</strong> design of small scale<br />

experiments that have helped <strong>in</strong> better underst<strong>and</strong><strong>in</strong>g the physics of pressure<br />

waves <strong>in</strong> pipel<strong>in</strong>es. For this I would like to thank H˚akon Myhren, Roger<br />

Over˚a, ˚Age Sivertsen <strong>and</strong> Terje Bjerkan. I would also like to thank Amir<br />

Ghaderi, fellow PhD c<strong>and</strong>idate at <strong>NTNU</strong> IPT for help<strong>in</strong>g me with ultrasonic<br />

measurements.<br />

The cheerful help <strong>and</strong> support from the adm<strong>in</strong>istrative staff at the Department<br />

vii


of Petroleum Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Applied Geophysics are deeply appreciated.<br />

For this I would like to thank Marit Valle Raaness, Tone Sanne, Madele<strong>in</strong><br />

Wold, Anne Lise Brekken, Turid Halvorsen, Solveig Johnsen <strong>and</strong> Turid Ol<strong>in</strong>e<br />

Uvsløkk.<br />

Thanks to my friends at the Department of Petroleum Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Applied<br />

Geophysics for giv<strong>in</strong>g me an enjoyable time. I would like to thank Christian<br />

<strong>and</strong> Ivana for always enterta<strong>in</strong><strong>in</strong>g me. Pamela <strong>and</strong> Weider deserve a<br />

special thank for always be<strong>in</strong>g here for me.<br />

My family, I thank for care <strong>and</strong> support.<br />

Mércia, no matter how far we have come, I cannot wait to see tomorrow, with<br />

you.<br />

viii


List of Contents<br />

Abstract v<br />

Acknowledgments vii<br />

List of Contents ix<br />

List of Tables xiii<br />

List of Figures xv<br />

Nomenclature xxv<br />

1 Introduction 1<br />

2 <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es 5<br />

2.1 Modell<strong>in</strong>g Equations <strong>and</strong> Numerical Simulation . . . . . . . . . 5<br />

2.2 Source Term Modell<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . 6<br />

2.2.1 Local Flow History . . . . . . . . . . . . . . . . . . . . . 7<br />

2.2.2 Quasi Two-Dimensional Models . . . . . . . . . . . . . . 9<br />

2.2.3 Instantaneous Velocity <strong>and</strong> Acceleration . . . . . . . . . 10<br />

2.2.4 Local <strong>and</strong> Convective Acceleration . . . . . . . . . . . . 11<br />

2.2.5 Evaluation of Unsteady Friction Models . . . . . . . . . 12<br />

2.2.6 Pipel<strong>in</strong>e Viscoelasticity . . . . . . . . . . . . . . . . . . 13<br />

2.3 Leak Detection . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.3.1 Flow Measurements . . . . . . . . . . . . . . . . . . . . 14<br />

2.3.2 <strong>Pressure</strong> Wave Characteristics . . . . . . . . . . . . . . 15<br />

2.3.3 <strong>Pressure</strong> Signal Process<strong>in</strong>g . . . . . . . . . . . . . . . . . 15<br />

2.4 Flow Meter<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

3 Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation 19<br />

3.1 Propagation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

3.1.1 Modell<strong>in</strong>g Equations . . . . . . . . . . . . . . . . . . . . 19<br />

3.1.2 Transmission, Reflection . . . . . . . . . . . . . . . . . . 21<br />

ix


3.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.2.1 Static Boundary Conditions . . . . . . . . . . . . . . . . 26<br />

3.2.2 Dynamic Boundary Conditions . . . . . . . . . . . . . . 26<br />

3.3 <strong>Pressure</strong> Wave Superposition . . . . . . . . . . . . . . . . . . . 33<br />

3.3.1 St<strong>and</strong><strong>in</strong>g <strong>Waves</strong> . . . . . . . . . . . . . . . . . . . . . . . 39<br />

3.4 Two-Dimensional Modell<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . 41<br />

4 Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids 43<br />

4.1 Basic Relationships . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

4.1.1 Def<strong>in</strong>ition . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

4.1.2 Fluid Compressibility . . . . . . . . . . . . . . . . . . . 45<br />

4.2 Acoustic Velocities <strong>in</strong> Gases . . . . . . . . . . . . . . . . . . . . 46<br />

4.2.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

4.2.2 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

4.3 Acoustic Velocity <strong>in</strong> Liquid Oils . . . . . . . . . . . . . . . . . . 54<br />

4.3.1 Acoustic Velocity <strong>in</strong> Volatile Oil . . . . . . . . . . . . . 54<br />

4.3.2 Acoustic Velocity <strong>in</strong> Black Oil . . . . . . . . . . . . . . . 58<br />

4.3.3 Comparison Between Volatile Oils <strong>and</strong> Black Oils . . . . 62<br />

4.4 Acoustic Velocity <strong>in</strong> Liquid Water . . . . . . . . . . . . . . . . 62<br />

4.4.1 Acoustic Velocity . . . . . . . . . . . . . . . . . . . . . . 62<br />

4.4.2 Density of Water . . . . . . . . . . . . . . . . . . . . . . 64<br />

4.4.3 Isothermal Compressibility . . . . . . . . . . . . . . . . 65<br />

4.5 Acoustic Velocity <strong>in</strong> Fluid-Filled Pipe . . . . . . . . . . . . . . 66<br />

4.6 Acoustic Velocity Determ<strong>in</strong>ation . . . . . . . . . . . . . . . . . 71<br />

5 Numerical Simulation Methods 73<br />

5.1 Modell<strong>in</strong>g Equations . . . . . . . . . . . . . . . . . . . . . . . . 73<br />

5.2 Analytical Solutions . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.3 Method of Characteristics . . . . . . . . . . . . . . . . . . . . . 76<br />

5.4 F<strong>in</strong>ite Volume Methods . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.4.1 Description . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

5.4.2 Numerical Schemes . . . . . . . . . . . . . . . . . . . . . 79<br />

5.4.3 Test Cases . . . . . . . . . . . . . . . . . . . . . . . . . . 80<br />

5.4.4 Higher Resolution Method . . . . . . . . . . . . . . . . . 84<br />

5.4.5 CLAWPACK . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5.5 Numerical Scheme Selection . . . . . . . . . . . . . . . . . . . . 88<br />

6 <strong>Pressure</strong> Wave Attenuation 91<br />

6.1 Influence of Numerical Simulation Tool . . . . . . . . . . . . . . 91<br />

6.1.1 Unsteady Friction Modell<strong>in</strong>g Equations . . . . . . . . . 91<br />

6.1.2 Test Case <strong>and</strong> Numerical Simulation . . . . . . . . . . . 92<br />

6.1.3 Results of Friction Modell<strong>in</strong>g . . . . . . . . . . . . . . . 93<br />

x


6.1.4 F<strong>in</strong>ite Volume Method Accuracy . . . . . . . . . . . . . 96<br />

6.1.5 Pipel<strong>in</strong>e Viscoelasticity . . . . . . . . . . . . . . . . . . 97<br />

6.2 Acoustic Velocity . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

6.2.1 Ultrasonic Measurement . . . . . . . . . . . . . . . . . . 99<br />

6.2.2 Experimental Work . . . . . . . . . . . . . . . . . . . . 107<br />

6.3 Extended Acoustic Velocity Formulation . . . . . . . . . . . . . 111<br />

7 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments 115<br />

7.1 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

7.2 Incl<strong>in</strong>ed Experimental Set-Up <strong>and</strong> Data . . . . . . . . . . . . . 118<br />

7.3 Horizontal Experimental Set-Up <strong>and</strong> Data . . . . . . . . . . . . 121<br />

8 Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g 129<br />

8.1 Concept <strong>and</strong> Modell<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . . 129<br />

8.1.1 <strong>Impulse</strong> Generator . . . . . . . . . . . . . . . . . . . . . 129<br />

8.1.2 <strong>Pressure</strong> Wave Propagation . . . . . . . . . . . . . . . . 132<br />

8.1.3 Pipel<strong>in</strong>e Outlet . . . . . . . . . . . . . . . . . . . . . . . 133<br />

8.1.4 Pipel<strong>in</strong>e Inlet . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

8.2 Numerical Simulation Tool . . . . . . . . . . . . . . . . . . . . 135<br />

8.3 Modell<strong>in</strong>g Results . . . . . . . . . . . . . . . . . . . . . . . . . . 139<br />

8.4 Validation of <strong>Impulse</strong> Pump<strong>in</strong>g Model <strong>and</strong> Simulation Tool . . 143<br />

9 Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g 147<br />

9.1 Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147<br />

9.1.1 <strong>Impulse</strong> Generator . . . . . . . . . . . . . . . . . . . . . 147<br />

9.1.2 Non-Return Valves . . . . . . . . . . . . . . . . . . . . . 150<br />

9.2 Multiphase Fluid Flow . . . . . . . . . . . . . . . . . . . . . . . 152<br />

9.3 Horizontal Transport of Fluids . . . . . . . . . . . . . . . . . . 153<br />

9.3.1 In-Situ <strong>Pressure</strong> <strong>and</strong> <strong>Pressure</strong> Wave Amplitude . . . . . 153<br />

9.3.2 Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . 154<br />

9.3.3 Test Case . . . . . . . . . . . . . . . . . . . . . . . . . . 156<br />

9.4 Artificial Lift . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

9.4.1 Wellhead <strong>Pressure</strong> <strong>and</strong> <strong>Pressure</strong> Wave Amplitude . . . . 162<br />

9.4.2 Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . 164<br />

9.4.3 Test Case . . . . . . . . . . . . . . . . . . . . . . . . . . 165<br />

10 Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g 171<br />

10.1 Characteristic Curves . . . . . . . . . . . . . . . . . . . . . . . 171<br />

10.2 Comparison of Artificial Lift Techniques . . . . . . . . . . . . . 174<br />

10.2.1 Test Case Description . . . . . . . . . . . . . . . . . . . 174<br />

10.2.2 Determ<strong>in</strong>ation of <strong>Impulse</strong> Pump<strong>in</strong>g Parameters . . . . . 175<br />

10.2.3 Efficiency of <strong>Impulse</strong> Pump<strong>in</strong>g . . . . . . . . . . . . . . 176<br />

10.3 Possible Improvements of <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency . . . . . 183<br />

xi


10.4 Cost of <strong>Impulse</strong> Pump<strong>in</strong>g . . . . . . . . . . . . . . . . . . . . . 187<br />

11 Discussion 189<br />

12 Conclusions 191<br />

References 193<br />

A Technical Reports 201<br />

B Pump<strong>in</strong>g of Fluids Us<strong>in</strong>g <strong>Pressure</strong> <strong>Impulse</strong>s 205<br />

C Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation <strong>in</strong> Pipel<strong>in</strong>es: Steady-<br />

State <strong>and</strong> Transient Friction 217<br />

D Simulation of Rapid <strong>Pressure</strong> Transients <strong>in</strong><br />

Statfjord A Off-Load<strong>in</strong>g Pipel<strong>in</strong>e 235<br />

xii


List of Tables<br />

2.1 Theoretical Values of ̺T <strong>and</strong> ςT <strong>in</strong> Eq. 2.8 (Trikha, 1975) . . . 8<br />

2.2 Theoretical Values of ̺V <strong>and</strong> ςV <strong>in</strong> Eq. 2.10 for Particular f ×Re<br />

Conditions (Vardy <strong>and</strong> Hwang, 1991) . . . . . . . . . . . . . . . 9<br />

2.3 Theoretical Values of ςC <strong>and</strong> ̺C <strong>in</strong> Eq. 2.14 (Carstens <strong>and</strong><br />

Roller, 1959) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11<br />

3.1 Actuation Function Ξ for Circular Gate, Square Gate, Globe,<br />

Needle, Butterfly <strong>and</strong> Ball Valves for Dimensionless Actuation<br />

Time ξ (Eq. 3.20 to Eq. 3.23) (Wood <strong>and</strong> Jones, 1973, 1974) . 29<br />

4.1 Typical Natural Gas Molar Composition (Pedersen et al., 1989) 49<br />

4.2 Typical Volatile Oil Molar Composition, Pedersen et al. (1989) 55<br />

4.3 Typical Black Oil Molar Composition (Pedersen et al., 1989) . 59<br />

4.4 Structure Coefficient for Th<strong>in</strong>-Walled <strong>and</strong> Thick-Walled Pipel<strong>in</strong>es,<br />

S<strong>in</strong>gle-Sided Anchor or Double-Sided Anchor. Structure Fasten<strong>in</strong>gs<br />

Description <strong>in</strong> Figure 4.20 . . . . . . . . . . . . . . . . 68<br />

4.5 Typical Material Properties: Young’s Modulus <strong>and</strong> Poisson’s<br />

Ratio (Cheremis<strong>in</strong>off, 1996) . . . . . . . . . . . . . . . . . . . . 70<br />

4.6 Typical Fluid Densities <strong>and</strong> Compressibilities <strong>in</strong> Petroleum Eng<strong>in</strong>eer<strong>in</strong>g<br />

(McCa<strong>in</strong>, 1990) . . . . . . . . . . . . . . . . . . . . . 70<br />

5.1 Test Cases Description for Compar<strong>in</strong>g CTCS, Lax-Wendroff<br />

<strong>and</strong> ENO Methods . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

5.2 M<strong>in</strong>mod, Superbee, Monotonized-Centered <strong>and</strong> Van-Leer Flux<br />

Limiter Functions (Laney, 1998) . . . . . . . . . . . . . . . . . 84<br />

6.1 Test Case System Description <strong>and</strong> Numerical Simulation Parameters<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

6.2 Field A Molar Gas Composition (Jacobsen, 2008) . . . . . . . . 100<br />

6.3 Propeller Eng<strong>in</strong>e Voltage <strong>and</strong> Correspond<strong>in</strong>g Propeller Rotat<strong>in</strong>g<br />

Speed <strong>and</strong> L<strong>in</strong>ear Speed . . . . . . . . . . . . . . . . . . . . 108<br />

7.1 Conversion between Vertical Height <strong>and</strong> Pipel<strong>in</strong>e Length . . . . 118<br />

xiii


8.1 List of Subrout<strong>in</strong>es Included <strong>in</strong> the <strong>Impulse</strong> Pump<strong>in</strong>g Software 137<br />

8.2 List of Parameters Describ<strong>in</strong>g <strong>Impulse</strong> Pump<strong>in</strong>g . . . . . . . . . 138<br />

8.3 Computed Flow Rate <strong>in</strong> [l m<strong>in</strong> −1 ], for Different Inlet Non-Return<br />

Valve Crack<strong>in</strong>g <strong>Pressure</strong>s <strong>and</strong> <strong>Pressure</strong> Wave Peak-To-Peak Amplitudes<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146<br />

9.1 Description of Test Case for Numerical Investigation of <strong>Pressure</strong><br />

Waveform Impact on Reachable Flow Rate . . . . . . . . . . . 150<br />

9.2 Density, Acoustic Velocity, Bubble Po<strong>in</strong>t <strong>Pressure</strong>, Inlet Non-<br />

Return Valve Open<strong>in</strong>g <strong>Pressure</strong>s <strong>and</strong> Pipel<strong>in</strong>e In-Situ Static<br />

<strong>Pressure</strong> for <strong>Impulse</strong> Pump<strong>in</strong>g of Water, Volatile Oil <strong>and</strong> Black<br />

Oil at 100 o C . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157<br />

9.3 M<strong>in</strong>imum Production Tub<strong>in</strong>g Lengths L1 <strong>in</strong> Eq. 9.14 <strong>and</strong> L2 <strong>in</strong><br />

Eq. 9.15 for <strong>Impulse</strong> Pump<strong>in</strong>g of Water, Volatile Oil <strong>and</strong> Black<br />

Oil at 100 o C . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164<br />

9.4 <strong>Impulse</strong> Pump<strong>in</strong>g Parameters for Artificial Lift of Water, Volatile<br />

Oil <strong>and</strong> Black Oil at 100 o C, from 10000 m Deep Wellbore, 10<br />

<strong>in</strong>ches <strong>in</strong> Diameter, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation . . 166<br />

10.1 Desired Flow Rates, Wellbore <strong>Pressure</strong>s <strong>and</strong> Correspond<strong>in</strong>g Lift<strong>in</strong>g<br />

Heights for the Three Depressurization Phases of Statfjord<br />

Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175<br />

10.2 Description of Test Case for Numerical Investigation of <strong>Pressure</strong><br />

Waveform Impact on Reachable Flow Rate . . . . . . . . . . . 184<br />

D.1 Statfjord Oil Molar Composition at 5 ◦ C, 30 bara . . . . . . . . 238<br />

D.2 Geometry of Case 1 . . . . . . . . . . . . . . . . . . . . . . . . 251<br />

D.3 Geometry of Case 2 . . . . . . . . . . . . . . . . . . . . . . . . 251<br />

D.4 Geometry of Case 3 . . . . . . . . . . . . . . . . . . . . . . . . 252<br />

xiv


List of Figures<br />

2.1 Typical Water Hammer <strong>Pressure</strong> Trace, Reproduced from Holmboe<br />

<strong>and</strong> Rouleau (1967) . . . . . . . . . . . . . . . . . . . . . . 7<br />

3.1 Acoustic Interface Schematic Representation Between Media a<br />

<strong>and</strong> b . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

3.2 Influence of Acoustic Interface on <strong>Pressure</strong> Wave Propagation.<br />

Medium a between 0 <strong>and</strong> L/2. Medium b between L/2 <strong>and</strong> L.<br />

Za > Zb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

3.3 Influence of Acoustic Interface on <strong>Pressure</strong> Wave Propagation.<br />

Medium a between 0 <strong>and</strong> L/2. Medium b between L/2 <strong>and</strong> L.<br />

Za < Zb . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

3.4 Evolution of Transmission Coefficient Aga<strong>in</strong>st Ratio of Acoustic<br />

Impedances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.5 Evolution of Reflection Coefficient Aga<strong>in</strong>st Ratio of Acoustic<br />

Impedances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.6 <strong>Pressure</strong> <strong>and</strong> Fluid Velocity <strong>Waves</strong> Reflections From Open <strong>and</strong><br />

Close Boundary Conditions . . . . . . . . . . . . . . . . . . . . 27<br />

3.7 Schematic Representation of Check Valves. a) Circular Gate<br />

Valve. b) Square Gate Valve. c) Globe Valve. d) Needle Valve.<br />

e) Butterfly Valve. f) Ball Valve. . . . . . . . . . . . . . . . . . 28<br />

3.8 Valve Position Dur<strong>in</strong>g Uniform Closure . . . . . . . . . . . . . 31<br />

3.9 Valve Position Dur<strong>in</strong>g Accelerated Closure . . . . . . . . . . . . 31<br />

3.10 Flow Velocity Across Valve Dur<strong>in</strong>g Uniform Closure . . . . . . 32<br />

3.11 Flow Velocity Across Valve Dur<strong>in</strong>g Accelerated Closure . . . . 32<br />

3.12 Constructive Superposition of two <strong>Pressure</strong> <strong>Waves</strong> . . . . . . . 34<br />

3.13 Destructive Superposition of two <strong>Pressure</strong> <strong>Waves</strong> . . . . . . . . 35<br />

3.14 Constructive Superposition of a <strong>Pressure</strong> Wave on its own Reflection<br />

at a Closed Boundary . . . . . . . . . . . . . . . . . . . 37<br />

3.15 Destructive Superposition of a <strong>Pressure</strong> Wave on its own Reflection<br />

at an Open Boundary . . . . . . . . . . . . . . . . . . . 38<br />

xv


3.16 First <strong>and</strong> Third Harmonics for <strong>Pressure</strong> St<strong>and</strong><strong>in</strong>g Wave <strong>in</strong> a<br />

Pipel<strong>in</strong>e Closed at Both Ends . . . . . . . . . . . . . . . . . . . 40<br />

3.17 First <strong>and</strong> Second Harmonics for <strong>Pressure</strong> St<strong>and</strong><strong>in</strong>g Wave <strong>in</strong><br />

Pipel<strong>in</strong>e Open at Both Ends . . . . . . . . . . . . . . . . . . . . 40<br />

3.18 Two-Dimensional Lam<strong>in</strong>ar Velocity Profile Evolution <strong>in</strong> a Fluid-<br />

Filled Pipel<strong>in</strong>e at Zero Initial Velocity Dur<strong>in</strong>g One Angular<br />

<strong>Pressure</strong> Wave period . . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.19 Two-Dimensional Turbulent Velocity Profile Evolution <strong>in</strong> a Fluid-<br />

Filled Pipel<strong>in</strong>e at Zero Initial Velocity Dur<strong>in</strong>g One Angular<br />

<strong>Pressure</strong> Wave period . . . . . . . . . . . . . . . . . . . . . . . 42<br />

4.1 Typical Natural Gas Phase Envelope . . . . . . . . . . . . . . . 49<br />

4.2 Evolution of Typical Natural Gas Density with <strong>Pressure</strong> <strong>and</strong><br />

Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

4.3 Evolution of Typical Natural Gas Adiabatic Index with <strong>Pressure</strong><br />

<strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . . . . . . 50<br />

4.4 Evolution of Typical Natural Gas Z-Factor with <strong>Pressure</strong> <strong>and</strong><br />

Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

4.5 Evolution of Typical Natural Gas Isothermal Compressibility<br />

with <strong>Pressure</strong> <strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . 52<br />

4.6 Evolution of Typical Natural Gas Acoustic Velocity with <strong>Pressure</strong><br />

<strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . . . . . . 53<br />

4.7 Deviation Between Eq. 4.19 <strong>and</strong> Eq. 4.20 . . . . . . . . . . . . 53<br />

4.8 Typical Volatile Oil Phase Envelope . . . . . . . . . . . . . . . 55<br />

4.9 Evolution of Typical Volatile Oil Density with <strong>Pressure</strong> <strong>and</strong><br />

Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

4.10 Evolution of Typical Volatile Oil Isothermal Compressibility<br />

with <strong>Pressure</strong> <strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . 56<br />

4.11 Evolution of Acoustic Velocity <strong>in</strong> Typical Volatile Oil with <strong>Pressure</strong><br />

<strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . . . . . . 57<br />

4.12 Typical Black Oil Phase Envelope . . . . . . . . . . . . . . . . 59<br />

4.13 Evolution of Typical Black Oil Density with <strong>Pressure</strong> <strong>and</strong> Temperature<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

4.14 Evolution of Typical Black Oil Isothermal Compressibility with<br />

<strong>Pressure</strong> <strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . . . . 60<br />

4.15 Evolution of Acoustic Velocity <strong>in</strong> Typical Black Oil with <strong>Pressure</strong><br />

<strong>and</strong> Temperature . . . . . . . . . . . . . . . . . . . . . . . 61<br />

4.16 Evolution of Acoustic Velocity <strong>in</strong> Water with <strong>Pressure</strong> <strong>and</strong> Temperature<br />

at Zero Sal<strong>in</strong>ity (DelGrosso, 1974) . . . . . . . . . . . 64<br />

4.17 Evolution of Water Density with <strong>Pressure</strong> <strong>and</strong> Temperature at<br />

Zero Sal<strong>in</strong>ity (Michaelides, 1981) . . . . . . . . . . . . . . . . . 65<br />

xvi


4.18 Evolution of Water Isothermal Compressibility with <strong>Pressure</strong><br />

<strong>and</strong> Temperature at Zero Sal<strong>in</strong>ity . . . . . . . . . . . . . . . . . 66<br />

4.19 Water Isothermal Compressibility extrapolated from Acoustic<br />

Velocity <strong>and</strong> Density Correlations (DelGrosso, 1974; Michaelides,<br />

1981) <strong>and</strong> empirically correlated (Eisenberg <strong>and</strong> Kauzman, 1969) 67<br />

4.20 Illustration of Structure Fasten<strong>in</strong>gs (Wylie <strong>and</strong> Streeter, 1993) 68<br />

4.21 Acoustic Velocity <strong>in</strong> Fluid-Filled Pipel<strong>in</strong>es (Wylie <strong>and</strong> Streeter,<br />

1993) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.22 Acoustic Velocity <strong>in</strong> Fluid-Filled Non-Circular Pipel<strong>in</strong>es . . . . 72<br />

5.1 Typical Pipel<strong>in</strong>e System <strong>in</strong> <strong>Pressure</strong> Transient Analysis Schematic<br />

Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

5.2 Characteristic L<strong>in</strong>es Representation <strong>in</strong> St<strong>and</strong>ard X-T Plane Grid 76<br />

5.3 Illustration of a F<strong>in</strong>ite Volume Method for Updat<strong>in</strong>g the Cell<br />

Average Ωn i by Fluxes at the Cell Edges . . . . . . . . . . . . . 79<br />

5.4 Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical Schemes<br />

on Test Case 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.5 Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical Schemes<br />

on Test Case 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 82<br />

5.6 Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical Schemes<br />

on Test Case 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

5.7 Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical Schemes<br />

on Test Case 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 83<br />

5.8 Comparison of M<strong>in</strong>mod, Superbee, Monotonized Centered <strong>and</strong><br />

Van-Leer Flux Limiters Coupled to the Lax-Wendroff Numerical<br />

Scheme on Test Case 1 . . . . . . . . . . . . . . . . . . . . . 86<br />

5.9 Comparison of M<strong>in</strong>mod, Superbee, Monotonized Centered <strong>and</strong><br />

Van-Leer Flux Limiters Coupled to the Lax-Wendroff Numerical<br />

Scheme on Test Case 2 . . . . . . . . . . . . . . . . . . . . .<br />

5.10 Comparison of M<strong>in</strong>mod, Superbee, Monotonized Centered <strong>and</strong><br />

Van-Leer Flux Limiters Coupled to the Lax-Wendroff Numeri-<br />

86<br />

cal Scheme on Test Case 3 . . . . . . . . . . . . . . . . . . . . .<br />

5.11 Comparison of M<strong>in</strong>mod, Superbee, Monotonized Centered <strong>and</strong><br />

Van-Leer Flux Limiters Coupled to the Lax-Wendroff Numeri-<br />

87<br />

cal Scheme on Test Case 4 . . . . . . . . . . . . . . . . . . . . .<br />

5.12 Comparison of Method of Characteristics <strong>and</strong> F<strong>in</strong>ite Volume<br />

Method (Lax-Wendroff + Superbee Flux Limiter) on St<strong>and</strong>ard<br />

87<br />

<strong>Pressure</strong> Transient Test Case . . . . . . . . . . . . . . . . . . . 89<br />

6.1 Comparison of Method of Characteristics <strong>and</strong> F<strong>in</strong>ite Volume<br />

Method with Steady Friction Only. nX = 15. nT = 540. . . . . 94<br />

xvii


6.2 Comparison of Method of Characteristics <strong>and</strong> F<strong>in</strong>ite Volume<br />

Method with Brunone’s Unsteady Friction Model. nX = 15.<br />

nT = 540. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

6.3 Comparison of Steady Friction <strong>and</strong> Brunone’s Unsteady Friction<br />

Models with the Method of Characteristics. nX = 15.<br />

nT = 540. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

6.4 Comparison of Steady Friction <strong>and</strong> Brunone’s Unsteady Friction<br />

Models with the Selected F<strong>in</strong>ite Volume Method. nX = 15.<br />

nT = 540. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95<br />

6.5 Influence of Spatial Discretization on Water Hammer Calculation<br />

with the Selected F<strong>in</strong>ite Volume Method. Detail of <strong>Pressure</strong><br />

Trace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

6.6 Comparison of Investigated Unsteady Friction Models Us<strong>in</strong>g a<br />

Selected F<strong>in</strong>ite Volume Method. nX = 1000. nT = 540. . . . . 98<br />

6.7 Comparison of Water Hammer Simulation Us<strong>in</strong>g the Method of<br />

Characteristics <strong>and</strong> a F<strong>in</strong>ite Volume Method. MOC: nX = 15.<br />

nT = 540. FVM: nX = 1000. nT = 540. . . . . . . . . . . . . . 98<br />

6.8 Schematic Representation of an Ultrasonic Flow Meter . . . . . 100<br />

6.9 Field A Gas Phase Envelope (Jacobsen, 2008) . . . . . . . . . . 100<br />

6.10 Measured <strong>and</strong> Calculated Acoustic Velocities Function of <strong>Pressure</strong><br />

<strong>and</strong> Temperature. First Data Set . . . . . . . . . . . . . . 102<br />

6.11 Measured <strong>and</strong> Calculated Acoustic Velocities Function of <strong>Pressure</strong>.<br />

First Data Set . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

6.12 Measured <strong>and</strong> Calculated Acoustic Velocities Function of Temperature.<br />

First Data Set . . . . . . . . . . . . . . . . . . . . . . 103<br />

6.13 Measured <strong>and</strong> Calculated Acoustic Velocities Function of Flow<br />

Velocity. First Data Set . . . . . . . . . . . . . . . . . . . . . . 103<br />

6.14 Measured <strong>and</strong> Calculated Acoustic Velocities Function of <strong>Pressure</strong><br />

<strong>and</strong> Temperature. Second Data Set . . . . . . . . . . . . . 105<br />

6.15 Measured <strong>and</strong> Calculated Acoustic Velocities Function of <strong>Pressure</strong>.<br />

Second Data Set . . . . . . . . . . . . . . . . . . . . . . . 105<br />

6.16 Measured <strong>and</strong> Calculated Acoustic Velocities Function of Temperature.<br />

Second Data Set . . . . . . . . . . . . . . . . . . . . . 106<br />

6.17 Measured <strong>and</strong> Calculated Acoustic Velocities Function of Flow<br />

Velocity. Second Data Set . . . . . . . . . . . . . . . . . . . . . 106<br />

6.18 Schematic Representation of Experimental Apparatus for Measur<strong>in</strong>g<br />

Acoustic Velocity <strong>in</strong> Fluid Under Rotation . . . . . . . . 107<br />

6.19 Sound Beam Travel Times for Different Propeller Eng<strong>in</strong>e Voltage.<br />

Transducer 1 . . . . . . . . . . . . . . . . . . . . . . . . . 109<br />

6.20 Sound Beam Travel Times for Different Propeller Eng<strong>in</strong>e Voltage.<br />

Transducer 2 . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

xviii


6.21 Sound Beam Travel Times for Different Propeller Eng<strong>in</strong>e Voltage.<br />

Transducer 3 . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

6.22 Sound Beam Travel Times for Different Propeller Eng<strong>in</strong>e Voltage.<br />

Transducer 4 . . . . . . . . . . . . . . . . . . . . . . . . . 111<br />

6.23 Detail of Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of Flow Velocity - First Data Set Figure 6.13 - <strong>and</strong> Curve Fitt<strong>in</strong>g112<br />

6.24 <strong>Pressure</strong>, Fluid <strong>and</strong> Acoustic Velocity Gradients <strong>in</strong> Left-To-<br />

Right Propagat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> . . . . . . . . . . . . . . . . 113<br />

7.1 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Horizontal Configuration . . . . . 116<br />

7.2 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Incl<strong>in</strong>ed Configuration . . . . . . 116<br />

7.3 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Vertical Configuration . . . . . . 117<br />

7.4 Sample of <strong>Pressure</strong> Traces at Four Locations Along the Incl<strong>in</strong>ed<br />

Flow L<strong>in</strong>e. ˙γ = 2.0 Hz . . . . . . . . . . . . . . . . . . . . . . . 119<br />

7.5 Sample of <strong>Pressure</strong> Traces at Four Locations Along the Incl<strong>in</strong>ed<br />

Flow L<strong>in</strong>e. ˙γ = 3.0 Hz . . . . . . . . . . . . . . . . . . . . . . . 120<br />

7.6 Maximum <strong>Pressure</strong>s at Four Locations for Five Different Experiments<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120<br />

7.7 Sample of Piston Stroke Trace. ˙γ = 1.4 Hz, ∆l = 0.75 10 −3 m . 122<br />

7.8 Sample of Force on the Piston Trace. ˙γ = 1.4 Hz, ∆l =<br />

0.7510 −3 m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

7.9 Fourier Analysis of Piston Stroke Signal. ˙γ = 1.4 Hz, ∆l =<br />

0.7510 −3 m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

7.10 Sample of <strong>Pressure</strong> Trace at the Outlet Non-Return Valve. ˙γ =<br />

1.4 Hz, ∆l = 0.7510 −3 m . . . . . . . . . . . . . . . . . . . . . . 124<br />

7.11 Sample of <strong>Pressure</strong> Trace at the Inlet Non-Return Valve. ˙γ =<br />

1.4 Hz, ∆l = 0.7510 −3 m . . . . . . . . . . . . . . . . . . . . . . 124<br />

7.12 Sample of <strong>Impulse</strong> Pump<strong>in</strong>g Flow Rate Trace. ˙γ = 1.4 Hz,<br />

∆l = 0.7510 −3 m . . . . . . . . . . . . . . . . . . . . . . . . . . 125<br />

7.13 Sample of <strong>Pressure</strong> Trace at Outlet Non-Return Valve. ˙γ =<br />

1.4 Hz, ∆l = 1.2510 −3 m . . . . . . . . . . . . . . . . . . . . . . 126<br />

7.14 Sample of <strong>Pressure</strong> Trace at Inlet Non-Return Valve. ˙γ =<br />

1.4 Hz, ∆l = 1.2510 −3 m . . . . . . . . . . . . . . . . . . . . . . 126<br />

7.15 Maximum <strong>Pressure</strong> Amplitudes at Outlet Non-Return Valve<br />

Aga<strong>in</strong>st <strong>Impulse</strong> Generator Displacement for Different Frequencies<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

7.16 Maximum <strong>Pressure</strong> Amplitudes at Inlet Non-Return Valve Function<br />

of <strong>Impulse</strong> Generator Displacement for Different Frequencies128<br />

7.17 Experimental Flow Rates Function of <strong>Impulse</strong> Generator Displacement<br />

for Different Frequencies . . . . . . . . . . . . . . . . 128<br />

8.1 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Horizontal Configuration . . . . . 130<br />

xix


8.2 Illustration of <strong>Impulse</strong> Generator Generat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong><br />

<strong>in</strong> Diameter Greater than Flowl<strong>in</strong>e Dimensions . . . . . . . . . 131<br />

8.3 Illustration of Piston Back-And-Forth Oscillations for S<strong>in</strong>usoidal<br />

<strong>Pressure</strong> Wave Generation . . . . . . . . . . . . . . . . . . . . . 132<br />

8.4 Non-Return Valve Schematic Representations . . . . . . . . . . 134<br />

8.5 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus Applied to Artificial Lift . . . . . 136<br />

8.6 Spatial <strong>Pressure</strong> Traces at Different Instants Dur<strong>in</strong>g <strong>Impulse</strong><br />

Pump<strong>in</strong>g Process (In-Situ Static <strong>Pressure</strong> is Represented by 0<br />

<strong>Pressure</strong> L<strong>in</strong>e). Data Presented <strong>in</strong> Dimensionless Form. . . . . 141<br />

8.7 Influence of Water Hammer <strong>Pressure</strong> Wave on Synthetic <strong>Pressure</strong><br />

Spatial Trace . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

8.8 Influence of Wellbore <strong>Pressure</strong> on Synthetic <strong>Pressure</strong> Spatial<br />

Trace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

8.9 Synthetic Temporal <strong>Pressure</strong> Trace at Downhole Inlet Non-<br />

Return Valve . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

8.10 Synthetic Temporal <strong>Pressure</strong> Trace at Downhole Inlet Non-<br />

Return Valve. ˙γ = 1.5 Hz. L = 283m . . . . . . . . . . . . . . 144<br />

8.11 Synthetic Flow Rate for Diverse Inlet Non-Return Valve Crack<strong>in</strong>g<br />

<strong>Pressure</strong>s <strong>and</strong> <strong>Pressure</strong> Wave Peak-To-Peak Amplitudes.<br />

˙γ = 1.5 Hz. L = 283m . . . . . . . . . . . . . . . . . . . . . . . 146<br />

9.1 Evolution of Piston Stroke for Generat<strong>in</strong>g <strong>Pressure</strong> Wave 100<br />

bar <strong>in</strong> Amplitude <strong>in</strong> a 1000 m Long Pipel<strong>in</strong>e . . . . . . . . . . . 148<br />

9.2 Piston L<strong>in</strong>ear Velocity for Generat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 100 bar<br />

<strong>in</strong> Amplitude <strong>in</strong> a 1000 m Long Water-Filled Pipel<strong>in</strong>e at Frequencies<br />

From 1 to 8 Hz . . . . . . . . . . . . . . . . . . . . . . 149<br />

9.3 Piston L<strong>in</strong>ear Velocity for Generat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong><br />

Amplitude <strong>in</strong> a 1000 m Long Oil-Filled Pipel<strong>in</strong>e at Frequencies<br />

From 1 to 8 Hz . . . . . . . . . . . . . . . . . . . . . . . . . . . 149<br />

9.4 Comparison of Check-Valves Efficiencies for Uniform <strong>and</strong> Accelerated<br />

Actuations . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

9.5 Water Air Mixture Acoustic Velocity Aga<strong>in</strong>st Void Fraction at<br />

1 bara, 10 bara <strong>and</strong> 100 bara . . . . . . . . . . . . . . . . . . . 153<br />

9.6 Reachable Flow Rate for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of Water,<br />

Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e, Us<strong>in</strong>g<br />

<strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak Amplitude, 4 Hz <strong>in</strong><br />

Frequency, Aga<strong>in</strong>st Difference Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-<br />

Return Valve Open<strong>in</strong>g <strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158<br />

xx


9.7 Reachable Flow Rate for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of Water,<br />

Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e, Us<strong>in</strong>g<br />

<strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak Amplitude, 5 Hz <strong>in</strong><br />

Frequency, Aga<strong>in</strong>st Difference Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-<br />

Return Valve Open<strong>in</strong>g <strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

9.8 Inflow Efficiency for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of Water,<br />

Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e, Us<strong>in</strong>g <strong>Pressure</strong><br />

<strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak Amplitude, 4 Hz <strong>in</strong> Frequency,<br />

Aga<strong>in</strong>st Difference Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-<br />

Return Valve Open<strong>in</strong>g <strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160<br />

9.9 Inflow Efficiency for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of Water,<br />

Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e, Us<strong>in</strong>g <strong>Pressure</strong><br />

<strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak Amplitude, 5 Hz <strong>in</strong> Frequency,<br />

Aga<strong>in</strong>st Difference Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-<br />

Return Valve Open<strong>in</strong>g <strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161<br />

9.10 Reachable Flow Rates Aga<strong>in</strong>st Wellbore <strong>Pressure</strong>, for Vertical<br />

Transport of Water, S<strong>in</strong>gle Phase Volatile <strong>and</strong> Black Oils, at<br />

100 o C, from a 10 km Deep Wellbore, Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 4<br />

Hz <strong>in</strong> Frequency, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation . . . . 167<br />

9.11 Reachable Flow Rates Aga<strong>in</strong>st Wellbore <strong>Pressure</strong>, for Vertical<br />

Transport of Water, S<strong>in</strong>gle Phase Volatile <strong>and</strong> Black Oils, at<br />

100 o C, from a 10 km Deep Wellbore, Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 5<br />

Hz <strong>in</strong> Frequency, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation . . . . 168<br />

9.12 Theoretical <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Efficiency Aga<strong>in</strong>st Wellbore<br />

<strong>Pressure</strong>, for Vertical Transport of Water, S<strong>in</strong>gle Phase<br />

Volatile <strong>and</strong> Black Oils at 100 o C, from a 10 km Deep Wellbore,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong> Frequency, Neglect<strong>in</strong>g <strong>Pressure</strong><br />

Wave Attenuation . . . . . . . . . . . . . . . . . . . . . . . . . 169<br />

9.13 Theoretical <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Efficiency Aga<strong>in</strong>st Wellbore<br />

<strong>Pressure</strong>, for Vertical Transport of Water, S<strong>in</strong>gle Phase<br />

Volatile <strong>and</strong> Black Oils at 100 o C, from a 10 km Deep Wellbore,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 5 Hz <strong>in</strong> Frequency, Neglect<strong>in</strong>g <strong>Pressure</strong><br />

Wave Attenuation . . . . . . . . . . . . . . . . . . . . . . . . . 170<br />

10.1 Evolution of <strong>Impulse</strong> Pump<strong>in</strong>g Reachable Flow Rate Aga<strong>in</strong>st<br />

Production Tub<strong>in</strong>g Depth . . . . . . . . . . . . . . . . . . . . . 173<br />

10.2 Evolution of <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency Aga<strong>in</strong>st Production<br />

Tub<strong>in</strong>g Depth . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173<br />

xxi


10.3 Characteristic Curve of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong><br />

Wave, 1 Hz <strong>in</strong> Frequency, 200 bar <strong>in</strong> Peak-To-Peak Amplitude<br />

<strong>and</strong> pHH = 101bara . . . . . . . . . . . . . . . . . . . . 174<br />

10.4 <strong>Pressure</strong> Wave Peak-To-Peak Amplitude Aga<strong>in</strong>st <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Efficiency for Artificial Lift Dur<strong>in</strong>g Statfjord Depressurization<br />

Phase A, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong> Frequency<br />

(Table 10.1) . . . . . . . . . . . . . . . . . . . . . . . . 177<br />

10.5 Wellhead <strong>Pressure</strong> Aga<strong>in</strong>st <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency for Artificial<br />

Lift Dur<strong>in</strong>g Statfjord Depressurization Phase A, Us<strong>in</strong>g<br />

S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong> Frequency (Table 10.1) . . . 177<br />

10.6 M<strong>in</strong>imum Reachable Bottom Hole <strong>Pressure</strong> Aga<strong>in</strong>st <strong>Impulse</strong><br />

Pump<strong>in</strong>g Efficiency for Artificial Lift Dur<strong>in</strong>g Statfjord Depressurization<br />

Phase A, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong><br />

Frequency (Table 10.1) . . . . . . . . . . . . . . . . . . . . . . . 178<br />

10.7 Reachable Heights <strong>and</strong> Flow Rates for Depressurization of Statfjord<br />

Field, 2748 m Deep, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong> 412<br />

bar <strong>in</strong> Peak-To-Peak Amplitude Aga<strong>in</strong>st <strong>Pressure</strong> Wave Frequency178<br />

10.8 Reachable Inflow Efficiencies <strong>and</strong> Flow Rates for Depressurization<br />

of Statfjord Field, 2748 m Deep, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong><br />

<strong>Waves</strong> 412 bar <strong>in</strong> Peak-To-Peak Amplitude Aga<strong>in</strong>st <strong>Pressure</strong><br />

Wave Frequency . . . . . . . . . . . . . . . . . . . . . . . . . . 179<br />

10.9 <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Characteristics for 102 bara Wellhead<br />

<strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1 Hz <strong>in</strong> Frequency,<br />

52 bar <strong>in</strong> Peak-To-Peak Amplitude . . . . . . . . . . . . . . . . 181<br />

10.10<strong>Impulse</strong> Pump<strong>in</strong>g Inflow Characteristics for 104 bara Wellhead<br />

<strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1 Hz <strong>in</strong> Frequency,<br />

206 bar <strong>in</strong> Peak-To-Peak Amplitude . . . . . . . . . . . . . . . 181<br />

10.11<strong>Impulse</strong> Pump<strong>in</strong>g Inflow Characteristics for 207 bara Wellhead<br />

<strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1 Hz <strong>in</strong> Frequency,<br />

412 bar <strong>in</strong> Peak-To-Peak Amplitude . . . . . . . . . . . . . . . 182<br />

10.12Total <strong>Impulse</strong> Pump<strong>in</strong>g Characteristics for 207 bara Wellhead<br />

<strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1 Hz <strong>in</strong> Frequency,<br />

412 bar <strong>in</strong> Peak-To-Peak Amplitude . . . . . . . . . . . . . . . 183<br />

10.13Spatial Traces of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g S<strong>in</strong>usoidal Waveform 184<br />

10.14Spatial Traces of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g Square Waveform . . 185<br />

10.15Spatial Traces of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g Triangular Waveform 185<br />

10.16Reachable Flow Rates Aga<strong>in</strong>st <strong>Impulse</strong> Generator Frequency<br />

<strong>and</strong> <strong>Pressure</strong> Wave Waveforms . . . . . . . . . . . . . . . . . . 186<br />

10.17Total <strong>Impulse</strong> Pump<strong>in</strong>g Characteristics for 207 bara Wellhead<br />

<strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1 Hz <strong>in</strong> Frequency,<br />

412 bar <strong>in</strong> Peak-To-Peak Amplitude, Production Tub<strong>in</strong>g 1 <strong>in</strong>ch,<br />

2 <strong>in</strong>ch, 4 <strong>in</strong>ch, <strong>and</strong> 8 <strong>in</strong>ch <strong>in</strong> Diameter . . . . . . . . . . . . . . 188<br />

xxii


B.1 <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Horizontal Configuration . . . . . 213<br />

B.2 Representative Values of the Reflection Coefficient . . . . . . . 214<br />

B.3 <strong>Pressure</strong> Wave (a) After Pass<strong>in</strong>g a Non-return Outlet Valve, (b)<br />

After Reflect<strong>in</strong>g From a Closed Boundary, (c) After Reflect<strong>in</strong>g<br />

on an Open Boundary . . . . . . . . . . . . . . . . . . . . . . . 214<br />

B.4 Truncated <strong>Pressure</strong> Wave Travell<strong>in</strong>g (a) Downward, (b) After<br />

Reflect<strong>in</strong>g From a Non-Return Valve . . . . . . . . . . . . . . . 215<br />

B.5 Truncated <strong>Pressure</strong> Wave Travell<strong>in</strong>g After Reflect<strong>in</strong>g From a<br />

Non-Return Valve . . . . . . . . . . . . . . . . . . . . . . . . . 215<br />

C.1 Typical Water Hammer <strong>Pressure</strong> Trace . . . . . . . . . . . . . . 229<br />

C.2 Schematic Representation of Test Case System . . . . . . . . . 230<br />

C.3 Comparison Between the Method of Characteristics (dotted<br />

l<strong>in</strong>e) <strong>and</strong> the Selected F<strong>in</strong>ite Volume Method (solid l<strong>in</strong>e) . . . . 230<br />

C.4 Detail of Figure (C.3) . . . . . . . . . . . . . . . . . . . . . . . 231<br />

C.5 <strong>Pressure</strong> Trace at Non-Return Valve . . . . . . . . . . . . . . . 232<br />

C.6 <strong>Pressure</strong> Trace at the Valve . . . . . . . . . . . . . . . . . . . . 233<br />

C.7 Detail of the <strong>Pressure</strong> Trace at the Valve . . . . . . . . . . . . . 234<br />

D.1 Illustration of the Off-Load<strong>in</strong>g System between the Riser Base<br />

<strong>and</strong> the S<strong>in</strong>gle Path Coupler . . . . . . . . . . . . . . . . . . . . 237<br />

D.2 Fluid Velocity Response for 6 Different Types of Valves. Uniform<br />

Valve Closure (Wood <strong>and</strong> Jones, 1973) . . . . . . . . . . . 240<br />

D.3 Fluid Velocity Response for 6 Different Types of Valves. Accelerated<br />

Valve Closure (Wood <strong>and</strong> Jones, 1973) . . . . . . . . . . 240<br />

D.4 Maximum <strong>Pressure</strong> Achieved <strong>in</strong> the Oil Off-Load<strong>in</strong>g System <strong>in</strong><br />

3 seconds with asteel = 1000 ms <strong>and</strong> aflexible = 450 ms . . . . . 244<br />

D.5 Cumulated Time of the <strong>Pressure</strong> Surge be<strong>in</strong>g over 50 barg . . . 245<br />

D.6 Maximum <strong>Pressure</strong> Achieved <strong>in</strong> the Oil Off-Load<strong>in</strong>g System <strong>in</strong><br />

3 s with asteel = 1000 ms <strong>and</strong> aflexible = 350 ms . . . . . . . . 246<br />

D.7 Maximum <strong>Pressure</strong> Achieved <strong>in</strong> the Oil Off-Load<strong>in</strong>g System <strong>in</strong><br />

3 seconds with asteel = 1100 ms <strong>and</strong> aflexible = 450 ms . . . . . 247<br />

D.8 Illustration of the Three Cases Considered for the Modell<strong>in</strong>g . . 250<br />

xxiii


xxiv


Nomenclature<br />

Roman Symbols<br />

u Cross-Sectional Averaged Fluid Velocity [ms −1 ]<br />

A Cross Section Area [m 2 ]<br />

a Acoustic Velocity [m s −1 ]<br />

a Acoustic Velocity [ms −1 ]<br />

B Quadratic Matrix<br />

C + Backward Characteristics<br />

C + Forward Characteristics<br />

c1<br />

Structure Coefficient<br />

Cp Specific Heat Capacity at Constant <strong>Pressure</strong> [J K −1 ]<br />

cp<br />

Cv<br />

cv<br />

Heat Capacity at Constant <strong>Pressure</strong> [J K −1 ]<br />

Specific Heat Capacity at Constant Volume [J K −1 ]<br />

Heat Capacity at Constant Volume [J K −1 ]<br />

d Pipe Diameter [m]<br />

dp<br />

Piston Head Diameter [m]<br />

DS Dissipation Function [kg m s −2 ]<br />

E Energy [J]<br />

e Pipe Wall Thickness [m]<br />

f Friction Factor<br />

fq<br />

Quasi Steady Friction Factor<br />

xxv


fS<br />

Steady Friction Factor<br />

g Gravitational Acceleration [ms −2 ]<br />

H Vertical Lift Height [m]<br />

h Thermal Energy [J]<br />

I Unity Matrix<br />

J0<br />

J1<br />

First Order Bessel Function<br />

Second Order Bessel Function<br />

K Compressibility [Pa −1 ]<br />

k Adiabatic Index<br />

KΓ Isothermal Compressibility [Pa −1 ] (KFluid)<br />

KS Isentropic Compressibility [Pa −1 ]<br />

kA Empirical Constant <strong>in</strong> Axworthy et al.’s Model<br />

kB Empirical Constant <strong>in</strong> Brunone et al.’s Model<br />

KG Gas Compressibility [Pa −1 ]<br />

KO Oil Compressibility [Pa −1 ]<br />

KW Water Compressibility [Pa −1 ]<br />

KΓ Isothermal Compressibility [Pa −1 ] (KFluid)<br />

kB1 Empirical Constant <strong>in</strong> Brunone et al.’s Model<br />

KFlow Flow Structure Compressibility [Pa −1 ]<br />

KFluid Fluid Isothermal Compressibility [Pa −1 ] (KΓ)<br />

KPipe Pipel<strong>in</strong>e Compressibility [Pa −1 ]<br />

kRi Empirical Constant <strong>in</strong> Ramos et al.’s Model (i = 1, 2)<br />

kRi Empirical Constant <strong>in</strong> Ramos et al.’s Model (i = 1, 2)<br />

L Length / Depth [m]<br />

M Molar Mass [kg mol −1 ]<br />

Ma Mach Number<br />

xxvi


N Number of Sides<br />

n Number of Moles<br />

nT<br />

Number of Time Steps<br />

nX Number of Space Steps<br />

p <strong>Pressure</strong> [Pa]<br />

p0<br />

pI<br />

In-Situ Static <strong>Pressure</strong> [Pa]<br />

Incident <strong>Pressure</strong> [Pa]<br />

pR Reflected <strong>Pressure</strong> [Pa]<br />

pT<br />

pd<br />

pu<br />

Transmitted <strong>Pressure</strong> [Pa]<br />

Downstream <strong>Pressure</strong> [Pa]<br />

Upstream <strong>Pressure</strong> [Pa]<br />

pBP Bubble Po<strong>in</strong>t <strong>Pressure</strong> [Pa]<br />

pCI Inlet Non-Return Valve Crack<strong>in</strong>g <strong>Pressure</strong> [Pa]<br />

pCO Outlet Non-Return Valve Crack<strong>in</strong>g <strong>Pressure</strong> [Pa]<br />

pInlet Inlet Tank <strong>Pressure</strong> [Pa]<br />

pLift Vertical Lift <strong>Pressure</strong> [Pa]<br />

pMAWP Maximum Allowable Work<strong>in</strong>g <strong>Pressure</strong> [Pa]<br />

pm<strong>in</strong>BH M<strong>in</strong>imum Bottomhole <strong>Pressure</strong> [Pa]<br />

pm<strong>in</strong>IN M<strong>in</strong>imum Local <strong>Pressure</strong> at Inlet Non-Return Valve [Pa]<br />

pOpen<strong>in</strong>g Outlet Non-Return Valve Open<strong>in</strong>g <strong>Pressure</strong> [Pa]<br />

pOutlet Outlet Tank <strong>Pressure</strong> [Pa]<br />

pWave <strong>Pressure</strong> Wave Amplitude [Pa]<br />

pWB Wellbore <strong>Pressure</strong> [Pa]<br />

pWH Wellhead <strong>Pressure</strong> [Pa]<br />

q Flow Rate [l m<strong>in</strong> −1 ]<br />

qd<br />

Flow Rate Displaced by Piston Oscillations [l m<strong>in</strong> −1 ]<br />

xxvii


R Universal Gas Constant (R = 8.314 J mol −1 K −1 )<br />

Re Reynolds Number (Re = ρud/µ)<br />

s Source Term<br />

T Travel Period [s]<br />

t Time [s]<br />

t0<br />

Tv<br />

Instant of <strong>Pressure</strong> Wave Generation [s]<br />

Valve Full Open<strong>in</strong>g Time [s]<br />

TSim Simulated Period [s]<br />

u Fluid Velocity [ms −1 ]<br />

u0<br />

uI<br />

Initial Fluid Velocity [ms −1 ]<br />

Incident Fluid Velocity [ms −1 ]<br />

uR Reflected Fluid Velocity [ms −1 ]<br />

uT Transmitted Fluid Velocity [ms −1 ]<br />

ur<br />

u ′ r<br />

ux<br />

u ′ x<br />

Radial Fluid Velocity [ms −1 ]<br />

Variation of Radial Fluid Velocity [ms −1 ]<br />

Axial Fluid Velocity [ms −1 ]<br />

Variation of Axial Fluid Velocity [ms −1 ]<br />

uPiston Piston L<strong>in</strong>ear Velocity [ms −1 ]<br />

V Volume [m 3 ]<br />

V0<br />

Vd<br />

Volume of Fluid Orig<strong>in</strong>ally <strong>in</strong> Place [m 3 ]<br />

Volume Displaced by Piston Oscillation [m 3 ]<br />

X Outlet Valve Location [m]<br />

x Axial Coord<strong>in</strong>ate [m]<br />

Y Pipe Material Young’s Modulus [Pa]<br />

y Water Oil Mass Fraction<br />

z Z-Factor<br />

xxviii


Greek Symbols<br />

αp<br />

Thermal Expansion Coefficient [K −1 ]<br />

β Two-Root Parameter Def<strong>in</strong><strong>in</strong>g Characteristics<br />

∆A Change <strong>in</strong> Cross Section Area [m 2 ]<br />

∆ap Acoustic Velocity Correction for <strong>Pressure</strong> [ms −1 ]<br />

∆aT Acoustic Velocity Correction for Temperature [ms −1 ]<br />

∆aΣΓp Acoustic Velocity Correction for Sal<strong>in</strong>ity, Temperature <strong>and</strong><br />

<strong>Pressure</strong> [ms −1 ]<br />

∆aΣ Acoustic Velocity Correction for Sal<strong>in</strong>ity [ms −1 ]<br />

∆l <strong>Impulse</strong> Generator Stroke [m]<br />

∆p Change <strong>in</strong> <strong>Pressure</strong> [Pa]<br />

∆pf Frictional <strong>Pressure</strong> Drop [Pa]<br />

δ Stoke Layer Thickness [m]<br />

˙γ <strong>Impulse</strong> Generator Frequency [Hz]<br />

ǫG<br />

ǫL<br />

Gas - Liquid Volumetric Fraction<br />

Water - Oil Volumetric Fraction<br />

η Efficiency [%]<br />

Γ Temperature [K]<br />

κ Heat Transfer Coefficient [W m −2 K −1 ]<br />

Λ Source Term Vector<br />

λ Wavelength [m]<br />

µ Dynamic Viscosity [Pa s]<br />

ν K<strong>in</strong>ematic Viscosity [m 2 s −1 ]<br />

Ω Vector Variable<br />

ω Angular Frequency [rad s −1 ]<br />

∂ Partial Differential Operator<br />

xxix


Φ Average Flux Vector<br />

φ Theoretical Function<br />

Π Dimensionless Angular Frequency<br />

Ψ Constant<br />

ρ Density [kg m −3 ]<br />

ρG Gas Density [kg m −3 ]<br />

ρO Oil Density [kg m −3 ]<br />

ρW Water Density [kg m −3 ]<br />

Σ Sal<strong>in</strong>ity [ppm]<br />

τω<br />

τµ<br />

Wall Shear Stress [N m −2 ]<br />

Viscous Stress [N m −2 ]<br />

θ Incl<strong>in</strong>ation [rad]<br />

θm Valve Full Open<strong>in</strong>g Angle [rad]<br />

ΥT Weight<strong>in</strong>g Function <strong>in</strong> Trikha’s Model<br />

ΥV Weight<strong>in</strong>g Function <strong>in</strong> Vardy et al.’s Model<br />

ΥZ Weight<strong>in</strong>g Function <strong>in</strong> Zielke’s Model<br />

εG<br />

εL<br />

̺C<br />

Gas - Liquid Mass Fraction<br />

Water - Oil Mass Fraction<br />

Empirical Constant <strong>in</strong> Carstens <strong>and</strong> Roller’s Model<br />

̺D Empirical Constant <strong>in</strong> Daily et al.’s Model<br />

̺S<br />

̺T<br />

̺Z<br />

̺V i<br />

ςC<br />

ςS<br />

Empirical Constant <strong>in</strong> Safwat <strong>and</strong> Polder’s Model<br />

Theoretical Constant <strong>in</strong> Trikha’s Model (i = 1, 2, 3)<br />

Dimensionless Time <strong>in</strong> Zielke’s Model (̺Z = 4νt/d)<br />

Theoretical Constant <strong>in</strong> Vardy et al.’s Model (i = 1, 2)<br />

Velocity Profile Shape Parameter<br />

Empirical Constant <strong>in</strong> Safwat <strong>and</strong> Polder’s Model<br />

xxx


ςT<br />

ςV i<br />

Theoretical Constant <strong>in</strong> Trikha’s Model (i = 1, 2, 3)<br />

Theoretical Constant <strong>in</strong> Vardy et al.’s Model (i = 1, 2)<br />

�Φ Approximate Flux Function<br />

Ξ Valve Actuation Function<br />

ξ Valve Dimensionless Actuation Time<br />

Superscripts<br />

n Time Step Number<br />

Subscripts<br />

Γ Isothermal Condition<br />

S Isentropic Condition<br />

G Gas<br />

i Cell Number<br />

O Oil<br />

W Water<br />

Γ Isothermal Condition<br />

Other Symbols<br />

F Acceleration Due to External Forces [ms −2 ]<br />

Ra→b Fluid Velocity Reflection Coefficient From Medium a to Medium b<br />

Ta→b Fluid Velocity Transmission Coefficient From Medium a to Medium b<br />

FPiston Force Applied on the Piston [N]<br />

PG<strong>in</strong> Generated Power at Inlet Non-Return Valve [W]<br />

PGout Generated Power at Outlet Non-Return Valve [W]<br />

PG Generated Power [W]<br />

PR Required Power [W]<br />

Ra→b Reflection Coefficient From Medium a to Medium b<br />

Ta→b Transmission Coefficient From Medium a to Medium b<br />

xxxi


Z Acoustic Impedance [kg m −2 s −1 ]<br />

Za<br />

Zb<br />

Mediam a Acoustic Impedance [kg m −2 s −1 ]<br />

Mediam b Acoustic Impedance [kg m −2 s −1 ]<br />

H Enthalpy [J kg −1 ]<br />

S Entropy [J K −1 ]<br />

U Internal Energy [J]<br />

xxxii


Chapter 1<br />

Introduction<br />

<strong>Impulse</strong> pump<strong>in</strong>g is a novel pump<strong>in</strong>g method patented by Sagov (2004) <strong>and</strong> developed<br />

by Clavis <strong>Impulse</strong> Technology AS. <strong>Impulse</strong> pump<strong>in</strong>g has been demonstrated<br />

for horizontal <strong>and</strong> vertical transport of liquids. However, impulse<br />

pump<strong>in</strong>g physical pr<strong>in</strong>ciples <strong>and</strong> performances <strong>in</strong> terms of head, flow rate <strong>and</strong><br />

efficiency have never been described nor reported <strong>in</strong> the literature before.<br />

Analysis of physical pr<strong>in</strong>ciples <strong>and</strong> performances <strong>in</strong> terms of head, flow rate <strong>and</strong><br />

efficiency, are necessary to evaluate potential applications of impulse pump<strong>in</strong>g.<br />

Of particular <strong>in</strong>terest are potential applications <strong>in</strong> petroleum eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong><br />

<strong>in</strong>vestigations of correspond<strong>in</strong>g designs of an impulse pump<strong>in</strong>g apparatus <strong>and</strong><br />

system.<br />

<strong>Impulse</strong> pump<strong>in</strong>g is based on pressure wave propagation, transmission <strong>and</strong> reflection<br />

<strong>in</strong> pipel<strong>in</strong>es. Hence, study<strong>in</strong>g impulse pump<strong>in</strong>g provides an opportunity<br />

to better underst<strong>and</strong> the phenomena of pressure waves <strong>in</strong> pipel<strong>in</strong>es from<br />

a scientific po<strong>in</strong>t of view. Establish<strong>in</strong>g impulse pump<strong>in</strong>g performances <strong>and</strong><br />

potential applications represents also an alternative to conventional pump<strong>in</strong>g<br />

methods, from an eng<strong>in</strong>eer<strong>in</strong>g po<strong>in</strong>t of view.<br />

Conventional pump<strong>in</strong>g methods are responsible for 25 % of the world electricity<br />

dem<strong>and</strong> with an average efficiency of about 60 % (Hydraulic Institute,<br />

2004). In addition, pump<strong>in</strong>g of deep wells or difficult fluids such as viscous oil<br />

us<strong>in</strong>g conventional methods still rema<strong>in</strong> challenges for petroleum eng<strong>in</strong>eers.<br />

Therefore, study<strong>in</strong>g <strong>in</strong>novative technologies offers opportunities to overcome<br />

pump<strong>in</strong>g issues with lower energy requirements.<br />

<strong>Impulse</strong> pump<strong>in</strong>g is based on pressure wave propagation <strong>in</strong> fluid-filled pipel<strong>in</strong>es<br />

that has been extensively <strong>in</strong>vestigated <strong>in</strong> hydraulic eng<strong>in</strong>eer<strong>in</strong>g. Of particular<br />

<strong>in</strong>terest is the considerable literature on water hammer phenomena <strong>in</strong> hydraulic<br />

pipel<strong>in</strong>es, with emphasis on propagation <strong>and</strong> attenuation modell<strong>in</strong>g<br />

that recently focused on transient friction <strong>and</strong> pipel<strong>in</strong>e viscoelasticity.<br />

1


Chapter 1. Introduction<br />

Practical applications us<strong>in</strong>g pressure waves <strong>in</strong> pipel<strong>in</strong>es, such as monitor<strong>in</strong>g of<br />

leakages <strong>and</strong> deposits, or flow meter<strong>in</strong>g, have also received particular attention<br />

over the last twenty years. Nevertheless, state of the art one dimensional<br />

numerical simulations still use the st<strong>and</strong>ard method of characteristics that<br />

has been developed for pipel<strong>in</strong>e modell<strong>in</strong>g more than fifty years ago; that is,<br />

at the beg<strong>in</strong>n<strong>in</strong>g of the computer age.<br />

The present thesis work analyses the pr<strong>in</strong>ciples of impulse pump<strong>in</strong>g, establishes<br />

a mechanistic model that substantiates a performance model, <strong>and</strong> explores<br />

applications of impulse pump<strong>in</strong>g <strong>in</strong> production <strong>and</strong> process<strong>in</strong>g of petroleum.<br />

In the present work, a pressure wave propagation model was <strong>in</strong>vestigated first<br />

based on hydraulic studies of water hammer events. <strong>Pressure</strong> wave propagation<br />

was then modelled <strong>in</strong> fluid-filled pipel<strong>in</strong>es, with particular focus on pressure<br />

wave characteristics dur<strong>in</strong>g superposition, reflection <strong>and</strong> transmission at<br />

acoustic <strong>in</strong>terfaces <strong>in</strong> pipel<strong>in</strong>es.<br />

A mechanistic impulse pump<strong>in</strong>g model was then derived based on pressure<br />

wave propagation <strong>and</strong> observation of a limited number of experimental data<br />

made available to <strong>NTNU</strong> by Clavis <strong>Impulse</strong> Technology AS. Of particular<br />

<strong>in</strong>terest was modell<strong>in</strong>g of pressure wave reflections at dynamic boundaries<br />

such as non-return valves.<br />

Numerical simulations of pressure wave propagation us<strong>in</strong>g f<strong>in</strong>ite volume numerical<br />

schemes were also <strong>in</strong>vestigated. Accuracy of f<strong>in</strong>ite volume methods<br />

was <strong>in</strong>vestigated first. Then, results of pressure transient calculations us<strong>in</strong>g a<br />

selected f<strong>in</strong>ite volume method <strong>and</strong> the well-established method of characteristics<br />

were compared.<br />

<strong>Pressure</strong> wave attenuation was <strong>in</strong>vestigated based on numerical simulation <strong>and</strong><br />

several models were compared. A new approach to pressure wave attenuation<br />

was also <strong>in</strong>vestigated based on experimental data made available to <strong>NTNU</strong><br />

dur<strong>in</strong>g the course of the PhD work.<br />

The impulse pump<strong>in</strong>g mechanistic model was then programmed us<strong>in</strong>g the<br />

numerical simulation tool based on a f<strong>in</strong>ite volume numerical scheme. Results<br />

from numerical simulations were then compared with experimental data made<br />

available to <strong>NTNU</strong> by Clavis <strong>Impulse</strong> Technology AS.<br />

<strong>Impulse</strong> pump<strong>in</strong>g performance characteristics for a range of petroleum eng<strong>in</strong>eer<strong>in</strong>g<br />

applications were f<strong>in</strong>ally <strong>in</strong>vestigated based on numerical simulation<br />

of different test-cases. The thesis is divided <strong>in</strong>to 10 chapters <strong>and</strong> 4 appendices<br />

listed below:<br />

• Chapter 2 conta<strong>in</strong>s literature review of basic pressure wave propagation<br />

pr<strong>in</strong>ciples <strong>and</strong> advanced models of pressure wave attenuation mecha-<br />

2


nisms. The literature review also describes applications based on pressure<br />

wave propagation properties.<br />

• Chapter 3 details modell<strong>in</strong>g of pressure wave propagation, reflection <strong>and</strong><br />

transmission at <strong>in</strong>terfaces <strong>and</strong> boundary conditions.<br />

• Chapter 4 reviews acoustic velocity calculations <strong>in</strong> s<strong>in</strong>gle-phase petroleum<br />

fluids.<br />

• Chapter 5 describes f<strong>in</strong>ite volume numerical methods that can be used<br />

for one-dimensional simulation of pressure transients <strong>in</strong> pipel<strong>in</strong>es, <strong>in</strong> replacement<br />

to st<strong>and</strong>ard numerical methods.<br />

• Chapter 6 analyses pressure wave attenuation models based on transient<br />

friction <strong>and</strong> pipel<strong>in</strong>e viscoelasticity, <strong>and</strong> <strong>in</strong>troduces a new approach to<br />

model pressure wave attenuation <strong>in</strong> pipel<strong>in</strong>es.<br />

• Chapter 7 describes an impulse pump apparatus <strong>and</strong> illustrates pressure<br />

measurements obta<strong>in</strong>ed from experimental work dur<strong>in</strong>g impulse pump<strong>in</strong>g<br />

of water.<br />

• Chapter 8 def<strong>in</strong>es the impulse pump<strong>in</strong>g concept <strong>and</strong> presents a mechanistic<br />

model <strong>and</strong> the correspond<strong>in</strong>g numerical model <strong>and</strong> numerical<br />

simulation tool.<br />

• Chapter 9 def<strong>in</strong>es an impulse pump<strong>in</strong>g design criteria <strong>and</strong> a performance<br />

model for horizontal <strong>and</strong> vertical transport of liquids.<br />

• Chapter 10 analyses impulse pump<strong>in</strong>g characteristic curves <strong>and</strong> <strong>in</strong>vestigates<br />

the range of petroleum eng<strong>in</strong>eer<strong>in</strong>g applications where impulse<br />

pump<strong>in</strong>g is the most efficient. A real practical case is also studied to<br />

compare impulse pump<strong>in</strong>g with st<strong>and</strong>ard artificial lift techniques.<br />

• Appendix A presents the list of technical reports written as part of the<br />

PhD work.<br />

• Appendix B conta<strong>in</strong>s the first scientific paper published on impulse<br />

pump<strong>in</strong>g.<br />

• Appendix C conta<strong>in</strong>s a scientific publication compar<strong>in</strong>g several transient<br />

friction models us<strong>in</strong>g a f<strong>in</strong>ite volume method.<br />

• Appendix C conta<strong>in</strong>s a scientific paper deal<strong>in</strong>g with pressure transient<br />

<strong>in</strong> multi section pipel<strong>in</strong>es, such as platform off-load<strong>in</strong>g pipel<strong>in</strong>es.<br />

3


Chapter 2<br />

<strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

2.1 Modell<strong>in</strong>g Equations <strong>and</strong> Numerical Simulation<br />

”<strong>Pressure</strong> waves are disturbances that transmit energy <strong>and</strong> momentum from<br />

one po<strong>in</strong>t to another through a medium without significant displacement of<br />

matter between the two po<strong>in</strong>ts” (Walski et al., 2003). Another way to express<br />

this is that pressure waves are disturbances that conta<strong>in</strong> <strong>in</strong>formation about<br />

the propagation medium <strong>and</strong> its surround<strong>in</strong>gs.<br />

<strong>Pressure</strong> waves can be destructive or <strong>in</strong>formative. On one h<strong>and</strong>, water hammer<br />

events are commonly associated with accidents that generally result from<br />

sudden valve actuation, or pump failures, for <strong>in</strong>stance. Water hammer events<br />

can therefore be seen as destructive. On the other h<strong>and</strong>, pressure pulses are<br />

commonly associated with purposely generated pressure waves, for collect<strong>in</strong>g<br />

<strong>in</strong>formation on pipel<strong>in</strong>e <strong>in</strong>tegrity <strong>and</strong> flow conditions, for <strong>in</strong>stance. Hence,<br />

pressure pulses can be seen as <strong>in</strong>formative.<br />

Water hammer <strong>and</strong> pressure pulses are both pressure waves. Historically, water<br />

hammer was def<strong>in</strong>ed as the propagation of pressure waves <strong>in</strong> compressible<br />

media <strong>in</strong> rigid tubes, whereas pressure pulses were def<strong>in</strong>ed as the propagation<br />

of pressure waves <strong>in</strong> <strong>in</strong>compressible media <strong>in</strong> elastic tubes. Both phenomena<br />

were then studied separately. The equivalence between water hammer <strong>and</strong><br />

pressure pulses was first observed by Kries (Tijssel<strong>in</strong>g <strong>and</strong> Anderson, 2007).<br />

<strong>Pressure</strong> wave propagation <strong>in</strong> general <strong>and</strong> water hammer <strong>in</strong> particular have<br />

been extensively studied over the years. Countless scientific publications <strong>and</strong><br />

books are available, from pressure wave propagation to modell<strong>in</strong>g of transient<br />

friction, <strong>and</strong> to case studies. Nevertheless, pressure wave propagation<br />

modell<strong>in</strong>g can be summarized by mass, momentum <strong>and</strong> energy conservation<br />

equations. Energy conservation is however commonly neglected due to the<br />

rapidness of transient phenomena.<br />

5


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

After simplification, one-dimensional mass <strong>and</strong> momentum conservation equations<br />

for study<strong>in</strong>g pressure wave propagation <strong>in</strong> terms of pressure, p, <strong>and</strong> fluid<br />

velocity, u, are, respectively (Ghidaoui et al., 2005):<br />

∂p<br />

+ ρa2∂u = 0 (2.1)<br />

∂t ∂x<br />

∂u<br />

∂t<br />

1 ∂p<br />

+<br />

ρ ∂x<br />

−1<br />

= fu|u| + F (2.2)<br />

2d<br />

where, a, def<strong>in</strong>es acoustic velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e, f, friction factor,<br />

<strong>and</strong>, F, acceleration due to external forces. Initial water hammer pressure<br />

amplitude <strong>in</strong> case of sudden valve actuation can be related to <strong>in</strong>itial fluid<br />

velocity, fluid density <strong>and</strong> acoustic velocity <strong>in</strong> the fluid-filled pipel<strong>in</strong>e us<strong>in</strong>g<br />

Joukowski’s equation. The theoretical equation results from simplification of<br />

mass <strong>and</strong> momentum conservation equations. Joukowski’s equation can be<br />

expressed by (Tijssel<strong>in</strong>g <strong>and</strong> Anderson, 2007):<br />

∆p = ρa∆u (2.3)<br />

Numerical simulations of pressure waves <strong>and</strong> water hammer phenomena started<br />

with the advent of computer age. The method of characteristics, based on<br />

mass <strong>and</strong> momentum hyperbolic system characteristics received particular attention<br />

due to its ease of programm<strong>in</strong>g, its low need <strong>in</strong> computer memory,<br />

<strong>and</strong> its overall efficiency. The eng<strong>in</strong>eer friendly method of characteristics was<br />

described <strong>and</strong> further detailed by Wylie <strong>and</strong> Streeter (1993).<br />

2.2 Source Term Modell<strong>in</strong>g<br />

Recent developments of pressure wave modell<strong>in</strong>g focus on momentum equation<br />

right-h<strong>and</strong> term, called source term. The source term corresponds to pressure<br />

wave energy dissipation mechanisms <strong>and</strong> attenuation. <strong>Pressure</strong> wave attenuation<br />

can be observed experimentally by damp<strong>in</strong>g <strong>and</strong> smooth<strong>in</strong>g of pressure<br />

signals. A typical pressure wave temporal trace for a water hammer event,<br />

result<strong>in</strong>g from sudden valve actuation, is shown <strong>in</strong> Figure 2.1.<br />

Frictional pressure drop was the first attenuation mechanism <strong>in</strong>cluded <strong>in</strong> pressure<br />

wave modell<strong>in</strong>g. Steady state friction factor was first calculated us<strong>in</strong>g the<br />

Darcy-Weisbach equation (Moody, 1944):<br />

fS = d 2<br />

L u2∆pf (2.4)<br />

where, fS, represents Darcy-Weisbach friction factor, d, pipel<strong>in</strong>e diameter, L,<br />

pipel<strong>in</strong>e length, <strong>and</strong>, ∆pf, frictional pressure drop. Numerical simulation of<br />

6


<strong>Pressure</strong><br />

2 4 6 8 10 12<br />

2.2. Source Term Modell<strong>in</strong>g<br />

Time [2L/a]<br />

Figure 2.1: Typical Water Hammer <strong>Pressure</strong> Trace, Reproduced from<br />

Holmboe <strong>and</strong> Rouleau (1967)<br />

water hammer events <strong>in</strong>clud<strong>in</strong>g steady state frictional pressure drop however<br />

failed to reproduce accurately experimental data behaviour. Thus, further<br />

modell<strong>in</strong>g of pressure wave attenuation was required. Unsteady friction received<br />

particular attention for expla<strong>in</strong><strong>in</strong>g damp<strong>in</strong>g <strong>and</strong> smooth<strong>in</strong>g.<br />

Unsteady friction models are composed of the st<strong>and</strong>ard steady state frictional<br />

pressure drop, <strong>and</strong> a frequency dependent term that is the unsteady friction<br />

factor per se. Frequency is not explicitly <strong>in</strong>cluded <strong>in</strong> unsteady friction models.<br />

Nevertheless, unsteady friction factor models were derived by tak<strong>in</strong>g <strong>in</strong>to<br />

account that pressure waves high order harmonics attenuate faster than low order<br />

harmonics. Unsteady friction is therefore implicitly a frequency dependent<br />

term. Unsteady friction has first been modelled based on local flow history,<br />

then based on flow velocity profiles, <strong>and</strong> lately based on flow accelerations.<br />

2.2.1 Local Flow History<br />

Evolution of past local flow history was first used for modell<strong>in</strong>g unsteady<br />

friction (Zielke, 1966, 1968). Unsteady friction was def<strong>in</strong>ed as a frequency<br />

dependent pressure wave attenuation mechanism, not<strong>in</strong>g that high order harmonics<br />

attenuate faster than low frequency components. Based on experimental<br />

observations, Zielke derived theoretically an unsteady friction factor that<br />

considered pressure transient frequency b<strong>and</strong>, <strong>and</strong> fluid flow history.<br />

Zielke derived the dynamic component of the wall shear stress us<strong>in</strong>g convolution<br />

<strong>in</strong>tegrals of temporal acceleration. In addition, weight<strong>in</strong>g functions were<br />

<strong>in</strong>troduced to <strong>in</strong>clude history of flow velocity <strong>in</strong> unsteady friction modell<strong>in</strong>g.<br />

Zielke’s wall shear stress, τω, <strong>and</strong> weight functions, ΥZ, were, respectively:<br />

τω(t) = 8ρν<br />

� t 4ρν ∂u(τ)<br />

u(t) +<br />

d d 0 ∂t ΥZ (t − τ) dτ (2.5)<br />

7


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

ΥZ<br />

�<br />

̺Z = 4ν<br />

⎧<br />

� ⎪⎨ exp(−773.7091̺Z) for ̺Z > 0.02<br />

t =<br />

d2 0.282095̺<br />

⎪⎩<br />

−1/2<br />

Z − 1.25 + 1.057855̺1/2 Z + 0.9375̺Z + ...<br />

...0.396696̺ 3/2<br />

Z − 0.351563̺2 Z for ̺Z < 0.02<br />

(2.6)<br />

Zielke developed the first unsteady friction model for lam<strong>in</strong>ar flow <strong>and</strong> tested<br />

it us<strong>in</strong>g the method of characteristics. Calculation of Zielke’s convolution<br />

<strong>in</strong>tegrals were expensive <strong>in</strong> terms of computer resources (CPU time <strong>and</strong> memory<br />

usage). Thus, Trikha (1975) approximated Zielke’s model by simplify<strong>in</strong>g<br />

weight<strong>in</strong>g functions <strong>and</strong> reduc<strong>in</strong>g computer costs. Trikha’s model was also<br />

valid for lam<strong>in</strong>ar flow only, <strong>and</strong> was expressed by:<br />

τω(t) = 8ρν 4ρν<br />

u(t) +<br />

d d (ΥT1(t) + ΥT2(t) + ΥT3(t)) (2.7)<br />

ΥTi=1,2,3(t + ∆t) = ΥTi(t)exp −̺Ti<br />

4ν<br />

d<br />

2 t<br />

+ςTi (u (t + ∆t) − u (t)) (2.8)<br />

where, ̺T, <strong>and</strong>, ςT, are theoretical constants. Values of ̺T <strong>and</strong> ςT are listed<br />

<strong>in</strong> Table 2.1.<br />

Table 2.1: Theoretical Values of ̺T <strong>and</strong> ςT <strong>in</strong> Eq. 2.8 (Trikha, 1975)<br />

̺T ςT<br />

i = 1 40 -8000<br />

i = 2 8.1 -200<br />

i = 3 1 -26.4<br />

Both Zielke’s <strong>and</strong> Trikha’s models of frequency dependent friction were derived<br />

for lam<strong>in</strong>ar flow conditions. Vardy <strong>and</strong> Hwang (1991) then derived an unsteady<br />

friction model for a wider range of Reynolds number based on decomposition<br />

of pipel<strong>in</strong>es <strong>in</strong>to concentric annuli surround<strong>in</strong>g a uniform core. Wall shear<br />

stress <strong>and</strong> weight<strong>in</strong>g functions were then expressed by:<br />

τω(t) = 8ρν 4ρν<br />

+<br />

d2 d2 (ΥV 1(t) + ΥV 2(t)) (2.9)<br />

ΥV i=1,2(t + ∆t) = ΥV i(t)exp −̺V i<br />

4ν<br />

d<br />

2 t<br />

+ςV i (u (t + ∆t) − u (t)) (2.10)<br />

where, ̺V , <strong>and</strong>, ςV , are theoretical constants. Values of ̺V <strong>and</strong> ςV are listed<br />

for particular values of the product of friction factor f <strong>and</strong> Reynolds number<br />

Re <strong>in</strong> Table 2.2. Vardy <strong>and</strong> Hwang’s unsteady friction model was also less<br />

expensive <strong>in</strong> computer resources than the orig<strong>in</strong>al model by Zielke.<br />

8


2.2. Source Term Modell<strong>in</strong>g<br />

Table 2.2: Theoretical Values of ̺V <strong>and</strong> ςV <strong>in</strong> Eq. 2.10 for Particular<br />

f × Re Conditions (Vardy <strong>and</strong> Hwang, 1991)<br />

f Re ̺V 1 ςV 1 ̺V 2 ςV 2<br />

69.5 365 1.2 10 5 35 5 10 3<br />

250 250 4.4 10 5 74 4.2 10 4<br />

500 260 5.6 10 5 65 1.15 10 5<br />

1000 350 9.8 10 5 65 4.12 10 5<br />

2000 470 2.8 10 6 65 1.62 10 6<br />

2.2.2 Quasi Two-Dimensional Models<br />

Vardy <strong>and</strong> Hwang’s unsteady friction model was based on decomposition of<br />

a pipel<strong>in</strong>e <strong>in</strong>to concentric annuli surround<strong>in</strong>g a uniform core (Vardy et al.,<br />

1993). Vardy <strong>and</strong> Hwang’s model was therefore two-dimensional <strong>in</strong> essence<br />

<strong>and</strong> mass <strong>and</strong> momentum conservation equations were solved <strong>in</strong> cyl<strong>in</strong>drical<br />

coord<strong>in</strong>ates. However, Vardy <strong>and</strong> Hwang used the one-dimensional method<br />

of characteristics to simulate their model. Of particular <strong>in</strong>terest with twodimensional<br />

modell<strong>in</strong>g is the prediction of flow cross-sectional velocity profile.<br />

Silva-Araya <strong>and</strong> Chaudry (1997) also proposed a quasi two-dimensional model<br />

for computation of unsteady friction <strong>and</strong> prediction of flow velocity profiles.<br />

Silva-Araya <strong>and</strong> Chaudry’s model was based on one-dimensional mass <strong>and</strong><br />

momentum conservation equations, but <strong>in</strong>cluded a two-dimensional dissipation<br />

function to account for viscous <strong>and</strong> turbulent stresses. The dissipation function<br />

was expressed by:<br />

DS = ∂ux<br />

∂r<br />

�<br />

µ ∂ux<br />

∂r − ρu′ xu ′ �<br />

r<br />

(2.11)<br />

where, ux, def<strong>in</strong>es axial flow velocity, ur, radial flow velocity, µ, dynamic<br />

viscosity, <strong>and</strong>, ρu ′ xu ′ r, Reynolds stress tensor. Silva-Araya <strong>and</strong> Chaudry’s<br />

model then required a turbulent model for approximat<strong>in</strong>g <strong>in</strong>stantaneous pressure<br />

gradients, velocity profiles <strong>and</strong> energy dissipation dur<strong>in</strong>g transient events.<br />

Pezz<strong>in</strong>ga (1999) then developed a quasi two-dimensional model for unsteady<br />

flow that also required modell<strong>in</strong>g of turbulence.<br />

Pezz<strong>in</strong>ga based his quasi two-dimensional model on mass <strong>and</strong> momentum conservation<br />

equations <strong>in</strong> cyl<strong>in</strong>drical coord<strong>in</strong>ates. Pezz<strong>in</strong>ga’s model worked for<br />

pipel<strong>in</strong>e networks <strong>and</strong> calculated flow velocity profiles, thus allow<strong>in</strong>g a more<br />

accurate evaluation of unsteady flow resistances. Velocity profile predictions<br />

permitted calculat<strong>in</strong>g wall shear stresses with greater accuracy. However,<br />

Pezz<strong>in</strong>ga noted that quasi two-dimensional modell<strong>in</strong>g was more expensive <strong>in</strong><br />

9


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

terms of computer resources <strong>and</strong> was barely justified by the result improvements.<br />

Correct approximation of flow velocity profiles dur<strong>in</strong>g transient events is nonetheless<br />

of great importance as it is directly related to wall shear stress calculations.<br />

Velocity profiles have been measured experimentally dur<strong>in</strong>g transient events<br />

by Brunone et al. (2000). Brunone et al. observed that the complexity of<br />

experimental results were not completely reflected by available source term<br />

modell<strong>in</strong>gs, even the more extended.<br />

2.2.3 Instantaneous Velocity <strong>and</strong> Acceleration<br />

Relations between flow structures <strong>and</strong> pressure losses were first <strong>in</strong>vestigated<br />

experimentally based on turbulent flows. Daily et al. (1956) proposed a friction<br />

factor formulation decomposed <strong>in</strong>to a steady part <strong>and</strong> an unsteady part that<br />

depends on the flow <strong>in</strong>stantaneous acceleration. Daily et al.’s friction factor<br />

was expressed by:<br />

2d<br />

f = fS + ̺D<br />

u2 ∂u<br />

∂t<br />

(2.12)<br />

where, ̺D, is an empirical constant equal to 0.01 <strong>and</strong> 0.62 for accelerat<strong>in</strong>g <strong>and</strong><br />

decelerat<strong>in</strong>g flows, respectively. Carstens <strong>and</strong> Roller (1959) proposed a similar<br />

expression based on the power law of velocity distribution for turbulent flow<br />

<strong>in</strong> a smooth pipe. The power law of velocity was expressed by:<br />

u<br />

u = (2ςC + 1)(ςC + 1)<br />

2ς2 �<br />

1 −<br />

C<br />

2r<br />

�1/ςC (2.13)<br />

d<br />

where, ςC, characterizes the shape of the velocity profile <strong>and</strong> is equal to 7 for<br />

Reynolds number less than 10 5 . Carstens <strong>and</strong> Roller’s friction factor formulation<br />

was then expressed by:<br />

d<br />

f = fS + ̺C<br />

u2 du<br />

dt<br />

(2.14)<br />

Values of ̺C for correspond<strong>in</strong>g values of ςC are listed <strong>in</strong> Table 2.3. H<strong>in</strong>o<br />

et al. (1976) also studied flow velocity profiles <strong>and</strong> their relation with the<br />

friction factor. H<strong>in</strong>o et al. proposed an empirical friction factor function of<br />

the Stoke layer thickness for Reynolds number from 105 to 5830. The Stokes<br />

layer thickness is the viscous boundary layer for an oscillatory flow <strong>and</strong> is<br />

def<strong>in</strong>ed by::<br />

δ =<br />

� �1/2 2ν<br />

ω<br />

10<br />

(2.15)


2.2. Source Term Modell<strong>in</strong>g<br />

Table 2.3: Theoretical Values of ςC <strong>and</strong> ̺C <strong>in</strong> Eq. 2.14 (Carstens <strong>and</strong><br />

Roller, 1959)<br />

ςC ̺ C<br />

7 0.449<br />

8 0.391<br />

9 0.346<br />

10 0.310<br />

where, δ, def<strong>in</strong>es Stokes layer thickness, ν, k<strong>in</strong>ematic viscosity, <strong>and</strong>, ω, angular<br />

frequency. H<strong>in</strong>o et al.’s relation between friction <strong>and</strong> <strong>in</strong>stantaneous velocity<br />

was expressed by:<br />

� �<br />

4L<br />

f = 0.1888 u<br />

νπa<br />

� −1/2.85<br />

(2.16)<br />

H<strong>in</strong>o et al.’s friction factor model was not decomposed <strong>in</strong>to steady <strong>and</strong> unsteady<br />

components. Safwat <strong>and</strong> Polder (1973) also proposed a new wall shear<br />

stress formulation that did not explicitly <strong>in</strong>clude steady <strong>and</strong> unsteady components.<br />

Safwat <strong>and</strong> Polder’s model was based on experimental work on oscillatory<br />

flow <strong>in</strong> U-tube <strong>and</strong> was expressed by:<br />

du<br />

τω = ̺Su + ςS<br />

dt<br />

(2.17)<br />

where, ̺S <strong>and</strong> ςS, were two empirical frequency dependent coefficients. Experimental<br />

work on wall shear stress also led Shuy (1996) to propose a new<br />

friction factor formulation. Shuy observed that unsteady friction <strong>in</strong>creased<br />

<strong>in</strong> decelerat<strong>in</strong>g flows <strong>and</strong> decreased <strong>in</strong> accelerat<strong>in</strong>g flows; thus contradict<strong>in</strong>g<br />

previous observations (Daily et al., 1956). Shuy’s friction factor is similar to<br />

Eq. 2.14, with ̺C equal to -0.33 <strong>and</strong> -0.52 for accelerat<strong>in</strong>g <strong>and</strong> decelerat<strong>in</strong>g<br />

flows, respectively.<br />

2.2.4 Local <strong>and</strong> Convective Acceleration<br />

Recent models of unsteady friction <strong>in</strong>cluded local <strong>and</strong> convective accelerations<br />

<strong>in</strong> the unsteady friction component. Local <strong>and</strong> convective acceleration were<br />

<strong>in</strong>cluded to account for local <strong>in</strong>ertia <strong>and</strong> wall shear stress. A first model us<strong>in</strong>g<br />

both accelerations was proposed by Brunone et al. (1991) <strong>and</strong> was expressed<br />

by:<br />

f = fS + kBd<br />

� �<br />

∂u<br />

− a∂u<br />

u|u| ∂t ∂x<br />

11<br />

(2.18)


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

where, kB, is an empirical constant that varies between 0.03 <strong>and</strong> 0.01 (Brunone<br />

et al., 1995), but a common value of 0.02 is normally used (Szymkiewicz <strong>and</strong><br />

Mitosek, 2007). Brunone et al. used the method of characteristics for simulat<strong>in</strong>g<br />

water hammer events <strong>in</strong>clud<strong>in</strong>g unsteady friction. Axworthy et al. (2000)<br />

proposed a second unsteady friction formulation for improv<strong>in</strong>g results from<br />

Brunone et al.’s model. Axworthy et al.’s model was expressed by:<br />

f = fS + kAd<br />

� � ��<br />

∂u u �<br />

+ a �<br />

∂u�<br />

�<br />

u|u| ∂t |u| �∂x<br />

�<br />

(2.19)<br />

where, kA, is a theoretical frequency dependent relaxation time, derived from<br />

irreversible thermodynamics. Axworthy et al. presented his model as an<br />

improvement of Brunone et al.’s model. Of particular <strong>in</strong>terest was the direction<br />

of propagation that was considered by Axworthy et al.. Ramos et al. (2004)<br />

also considered direction of propagation <strong>and</strong> proposed an improved unsteady<br />

friction formulation:<br />

f = fS + d<br />

u|u|<br />

�<br />

kR1<br />

∂u<br />

∂t<br />

+ kR2a u<br />

|u|<br />

� ��<br />

�<br />

�<br />

∂u�<br />

�<br />

�∂x<br />

�<br />

(2.20)<br />

where, kR1 <strong>and</strong>, kR2, are two empirical constants with values of 0.003 <strong>and</strong><br />

0.04, respectively. Brunone et al., Axworthy et al. <strong>and</strong> Ramos et al. (2004)<br />

used the method of characteristics for simulat<strong>in</strong>g water hammer events with<br />

unsteady friction modell<strong>in</strong>g.<br />

2.2.5 Evaluation of Unsteady Friction Models<br />

Unsteady friction models based on local flow history, <strong>in</strong>stantaneous velocity<br />

<strong>and</strong> acceleration, local <strong>and</strong> convective acceleration, <strong>and</strong> quasi two-dimensional<br />

models have been proposed over the years. Adamkowski <strong>and</strong> Lew<strong>and</strong>owski<br />

(2006) compared several unsteady friction models <strong>and</strong> observed that local flow<br />

history <strong>and</strong> quasi two-dimensional models were most accurate at low Reynolds<br />

number. Modell<strong>in</strong>g based on local <strong>and</strong> convective acceleration were however<br />

more accurate at high Reynolds number.<br />

Nixon <strong>and</strong> Ghidaoui (2007) compared several unsteady friction models <strong>in</strong><br />

pipel<strong>in</strong>e networks, for leakage detection. Nixon <strong>and</strong> Ghidaoui concluded that<br />

relevance of unsteady friction decreased as leakage sizes <strong>in</strong>creased <strong>and</strong> could<br />

be neglected for leak rates greater than 30 % of the <strong>in</strong>itial flow rate. Nixon<br />

<strong>and</strong> Ghidaoui also stated that <strong>in</strong>troduction of unsteady friction modell<strong>in</strong>g for<br />

prediction of transient events <strong>in</strong> pipel<strong>in</strong>e networks was not of particular importance<br />

for <strong>in</strong>creas<strong>in</strong>g calculations accuracy.<br />

Vitkovsky et al. (2006) also compared several unsteady friction models based<br />

on experimental considerations. Vitkovsky et al. noticed that agreements<br />

12


2.2. Source Term Modell<strong>in</strong>g<br />

between results from experimental work <strong>and</strong> numerical simulations were due<br />

to the <strong>in</strong>herent dissipation of the method of characteristics. Szymkiewicz<br />

<strong>and</strong> Mitosek (2007) concluded similarly based on mathematical study of the<br />

modell<strong>in</strong>g equations <strong>and</strong> compar<strong>in</strong>g the method of characteristics to a third<br />

order accurate f<strong>in</strong>ite element numerical scheme (Szymkiewicz <strong>and</strong> Mitosek,<br />

2005).<br />

2.2.6 Pipel<strong>in</strong>e Viscoelasticity<br />

Source term modell<strong>in</strong>g focused first on unsteady friction with four different<br />

approaches proposed <strong>in</strong> the literature: evolution of local flow history, quasi<br />

two-dimensional model, evolution of <strong>in</strong>stantaneous velocity <strong>and</strong> acceleration<br />

<strong>and</strong> evolution of local <strong>and</strong> convective accelerations, as review above. However,<br />

recent pressure wave propagation <strong>and</strong> attenuation models have <strong>in</strong>troduced<br />

pipel<strong>in</strong>e material viscoelastic behaviour for expla<strong>in</strong><strong>in</strong>g damp<strong>in</strong>g <strong>and</strong> smooth<strong>in</strong>g<br />

of pressure traces. Pipel<strong>in</strong>e viscoelastic behaviour has received particular<br />

attention due to the <strong>in</strong>creas<strong>in</strong>g use of plastic pipel<strong>in</strong>es for water distribution<br />

networks, but also for hydrocarbon transport systems.<br />

Bergant <strong>and</strong> Tijssel<strong>in</strong>g (2001) <strong>and</strong> Bergant et al. (2003a,b) <strong>in</strong>vestigated effects<br />

of pipe wall viscoelasticity <strong>and</strong> noted that the s<strong>in</strong>gle effect could not reproduced<br />

experimental results. However, comparison of experimental data <strong>and</strong><br />

numerical simulations by the method of characteristics comb<strong>in</strong><strong>in</strong>g pipel<strong>in</strong>e<br />

viscoelasticity <strong>and</strong> unsteady friction gave satisfactory results. Pipel<strong>in</strong>e wall<br />

behaviour dur<strong>in</strong>g transient events was decomposed <strong>in</strong>to an <strong>in</strong>stantaneous pipe<br />

wall response followed by a delayed response dependent on pipel<strong>in</strong>e material,<br />

use <strong>and</strong> fatigue.<br />

Covas (2003) <strong>and</strong> Covas et al. (2004, 2005) also <strong>in</strong>vestigated pipe wall viscoelastic<br />

behaviour <strong>in</strong> plastic pipel<strong>in</strong>es where unsteady friction alone could<br />

not expla<strong>in</strong> observed pressure damp<strong>in</strong>g. The acoustic velocity calculation was<br />

modified to account for temporal evolution of pipel<strong>in</strong>e material Young’s modulus.<br />

Comparison between experimental data <strong>and</strong> results from numerical simulation<br />

us<strong>in</strong>g the method of characteristics were <strong>in</strong> good agreement.<br />

Pipel<strong>in</strong>e viscoelasticity modell<strong>in</strong>g results were obta<strong>in</strong>ed by us<strong>in</strong>g the method<br />

of characteristics. In addition, pipel<strong>in</strong>e viscoelasticity was often coupled with<br />

unsteady friction. Numerical dissipation us<strong>in</strong>g the method of characteristics<br />

was not considered by the different authors, therefore improvements offered by<br />

pipe wall viscoelasticity modell<strong>in</strong>g are questionable. <strong>Pressure</strong> wave attenuation<br />

rema<strong>in</strong>s today an important field of research with no satisfactory modell<strong>in</strong>g.<br />

13


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

2.3 Leak Detection<br />

<strong>Pressure</strong> wave propagate, transmit <strong>and</strong> reflect accord<strong>in</strong>g to elastic properties of<br />

the propagation medium <strong>and</strong> its environment. Reflections <strong>and</strong> transmissions<br />

of pressure waves take place at acoustic <strong>in</strong>terfaces where elastic properties<br />

change. Difference <strong>in</strong> acoustic properties can be due to changes <strong>in</strong> fluids,<br />

flows, pipel<strong>in</strong>e structure or geometry. Informative pressure waves offer then<br />

valuable <strong>in</strong>formation relative to pipel<strong>in</strong>e structure, <strong>in</strong>tegrity <strong>and</strong> flow.<br />

2.3.1 Flow Measurements<br />

Transient mass flow measurements were <strong>in</strong>vestigated by Liggett <strong>and</strong> Chen<br />

(1994) for detect<strong>in</strong>g leakages <strong>in</strong> pipel<strong>in</strong>e networks. Inverse modell<strong>in</strong>g was performed<br />

from outlet flow rate measurements <strong>and</strong> calculated <strong>in</strong>let flow rate conditions<br />

were then compared with actual measurements. Deviations between<br />

calculated <strong>and</strong> measured <strong>in</strong>let flow conditions therefore <strong>in</strong>dicated presence of<br />

leaks <strong>in</strong> the network. Only detection was available from transient mass flow<br />

measurement, <strong>and</strong> the method relied heavily on the accuracy of transient flow<br />

modell<strong>in</strong>g.<br />

Transient flow rate measurements were also <strong>in</strong>vestigated by Erickson <strong>and</strong><br />

Twaite (1996) for detect<strong>in</strong>g leakages <strong>in</strong> multiphase pipel<strong>in</strong>e networks. Direct<br />

modell<strong>in</strong>g was performed from <strong>in</strong>let flow rate measurements <strong>and</strong> calculated<br />

outlet flow rate conditions were then compared with actual measurements. Deviations<br />

between calculated <strong>and</strong> measured outlet flow rates <strong>in</strong>dicated presence<br />

of leaks. Erickson <strong>and</strong> Twaite’s leak detection method relied on a simplified<br />

multiphase transient modell<strong>in</strong>g, thus was not accurate.<br />

Transient momentum balance was first <strong>in</strong>vestigated by Scott <strong>and</strong> Yi (1998) for<br />

detect<strong>in</strong>g, locat<strong>in</strong>g <strong>and</strong> dimension<strong>in</strong>g of leakages <strong>in</strong> s<strong>in</strong>gle pipel<strong>in</strong>es. Momentum<br />

measurements presented the advantage of gett<strong>in</strong>g rid of unreliable <strong>in</strong>let<br />

flow meter<strong>in</strong>g <strong>and</strong> calibration uncerta<strong>in</strong>ties. Leakages as low as 5 % of the<br />

outlet flow rate <strong>and</strong> more than 50 % of the outlet flow rate were located <strong>and</strong><br />

dimensioned for distant leaks <strong>and</strong> leaks with<strong>in</strong> the first 10 % of the pipel<strong>in</strong>e,<br />

respectively.<br />

Transient <strong>and</strong> momentum balances were among the first techniques for detect<strong>in</strong>g<br />

leakages <strong>in</strong> s<strong>in</strong>gle pipel<strong>in</strong>es <strong>and</strong> pipel<strong>in</strong>e networks. Actual flow rate<br />

measurements were used directly without signal treatment nor <strong>in</strong>terpretation.<br />

Flow measurements were used <strong>in</strong>stead of more reliable pressure signals (Abhulimen<br />

<strong>and</strong> Susu, 2004). Leak detection relied greatly on transient flow simulations,<br />

<strong>and</strong> pressure wave properties such as acoustic velocity <strong>and</strong> reflection<br />

characteristics were ignored.<br />

14


2.3.2 <strong>Pressure</strong> Wave Characteristics<br />

2.3. Leak Detection<br />

Contractor (1965) <strong>in</strong>vestigated water hammer pressure waves <strong>in</strong> pipe systems<br />

<strong>and</strong> observed wave reflections from pipel<strong>in</strong>e features such as valves, orifices<br />

<strong>and</strong> junctions. Contractor concluded that pipel<strong>in</strong>e features must be treated as<br />

boundary conditions <strong>in</strong> numerical simulations. Contractor’s <strong>in</strong>vestigation was<br />

the start<strong>in</strong>g po<strong>in</strong>t for Jonsson <strong>and</strong> Larson (1992) who analysed pressure traces<br />

<strong>and</strong> their frequency content. Jonsson <strong>and</strong> Larson showed that propagation<br />

characteristics conta<strong>in</strong>ed valuable <strong>in</strong>formation.<br />

<strong>Pressure</strong> wave propagation characteristics were then considered by Brunone<br />

(1999) for detect<strong>in</strong>g <strong>and</strong> locat<strong>in</strong>g leakages <strong>in</strong> s<strong>in</strong>gle pipel<strong>in</strong>es. A water hammer<br />

event was purposely generated by sudden valve actuation, <strong>and</strong> was then<br />

monitored at one end of a pipel<strong>in</strong>e. Deviations between experimental pressure<br />

measurements with <strong>and</strong> without leak were compared. Then, leakage location<br />

was predicted from pressure wave travel time <strong>and</strong> calculated acoustic velocity<br />

<strong>in</strong> the s<strong>in</strong>gle phase fluid-filled pipel<strong>in</strong>e.<br />

<strong>Pressure</strong> wave propagation characteristics were also <strong>in</strong>vestigated by Gudmundsson<br />

et al. (2000) for detect<strong>in</strong>g, locat<strong>in</strong>g <strong>and</strong> monitor<strong>in</strong>g deposits <strong>in</strong> multiphase<br />

s<strong>in</strong>gle pipel<strong>in</strong>es. A water hammer event was purposely generated by sudden<br />

valve actuation, <strong>and</strong> was then monitored at one end of the pipel<strong>in</strong>e. Experimental<br />

pressure data were then compared to synthetic pressure traces<br />

simulated for a leak-free system. Deposits were then located <strong>and</strong> monitored<br />

from pressure wave travel time <strong>and</strong> acoustic velocity <strong>in</strong> multiphase fluid-filled<br />

pipel<strong>in</strong>e (Dong <strong>and</strong> Gudmundsson, 1993).<br />

Studies of pressure wave characteristics helped improv<strong>in</strong>g leakage detection<br />

techniques by allow<strong>in</strong>g localization <strong>and</strong> monitor<strong>in</strong>g of leaks. Of particular<br />

<strong>in</strong>terest were the repeatability of transient tests for monitor<strong>in</strong>g <strong>and</strong> calibrat<strong>in</strong>g<br />

numerical simulations. However, leakage <strong>and</strong> deposit detection techniques<br />

proposed by Brunone <strong>and</strong> Gudmundsson et al. relied on water hammer events<br />

that required flowl<strong>in</strong>es full stoppage while measur<strong>in</strong>g. Nonetheless, both techniques<br />

relied on simple <strong>and</strong> direct <strong>in</strong>terpretation of pressure traces.<br />

2.3.3 <strong>Pressure</strong> Signal Process<strong>in</strong>g<br />

<strong>Pressure</strong> traces were <strong>in</strong>terpreted <strong>and</strong> analysed by Wang et al. (2002) for detect<strong>in</strong>g,<br />

locat<strong>in</strong>g <strong>and</strong> dimension<strong>in</strong>g leakages <strong>in</strong> s<strong>in</strong>gle pipel<strong>in</strong>es. A water hammer<br />

was purposely generated <strong>and</strong> monitored. Experimental pressure traces were<br />

then transposed <strong>in</strong> the frequency doma<strong>in</strong> by Fourier transform. Leakages were<br />

located <strong>and</strong> monitored by observ<strong>in</strong>g that damp<strong>in</strong>g of leak-free systems was <strong>in</strong>dependent<br />

on harmonics whereas damp<strong>in</strong>g <strong>in</strong> systems with leaks depended on<br />

frequency range. Leakages as small as 0.1 % of pipel<strong>in</strong>e’s cross sectional area<br />

15


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

were detected <strong>and</strong> located us<strong>in</strong>g numerical simulations.<br />

<strong>Pressure</strong> traces Fourier analysis as well as wavelet analysis were <strong>in</strong>vestigated by<br />

Ferrante <strong>and</strong> Brunone (2003a,b) for detect<strong>in</strong>g <strong>and</strong> locat<strong>in</strong>g leakages <strong>in</strong> s<strong>in</strong>gle<br />

pipel<strong>in</strong>es. Both analysis proved to be efficient for detect<strong>in</strong>g, locat<strong>in</strong>g <strong>and</strong> monitor<strong>in</strong>g<br />

leakages. However, modell<strong>in</strong>g of pressure wave damp<strong>in</strong>g <strong>in</strong> fluid-filled<br />

pipel<strong>in</strong>e was questioned s<strong>in</strong>ce leak characteristics determ<strong>in</strong>ation depended on<br />

numerical simulation results. Both Wang et al. <strong>and</strong> Ferrante <strong>and</strong> Brunone<br />

used the method of characteristics for numerical simulation of pressure transients.<br />

Leak detection <strong>and</strong> localization from numerical modell<strong>in</strong>g us<strong>in</strong>g the method<br />

of characteristics was <strong>in</strong>vestigated further by Wang <strong>and</strong> Carroll (2007). Of<br />

particular <strong>in</strong>terest were pipel<strong>in</strong>e description <strong>and</strong> discretization. Wang <strong>and</strong><br />

Carroll observed that imprecise pipel<strong>in</strong>e modell<strong>in</strong>g <strong>and</strong> configuration, as well<br />

as numerical model discretization could lead to absence of leakage detection<br />

or greatly <strong>in</strong>accurate calculations. The study relied also on the method of<br />

characteristics for pressure wave modell<strong>in</strong>g.<br />

Leaks <strong>and</strong> other pipel<strong>in</strong>e features effects on pressure traces were <strong>in</strong>vestigated<br />

by Beck et al. (2005) us<strong>in</strong>g cross correlation. Effects of leakages as well as<br />

pipel<strong>in</strong>e junctions were observed from cross-correlation of reflected pressure<br />

waves. Although be<strong>in</strong>g a repetition of well established pressure wave behaviours,<br />

the paper highlighted detection of pipel<strong>in</strong>e elements by simple pressure<br />

trace physical analysis <strong>and</strong> signal treatment. The technique relied on<br />

physical underst<strong>and</strong><strong>in</strong>g of pressure wave propagation.<br />

Lee et al. (2007) <strong>in</strong>terpreted <strong>and</strong> analysed pressure pulse traces for locat<strong>in</strong>g<br />

leaks <strong>in</strong> s<strong>in</strong>gle pipel<strong>in</strong>es. <strong>Pressure</strong> pulses were generated <strong>and</strong> monitored at<br />

one end of the pipel<strong>in</strong>e. <strong>Pressure</strong> traces were then transposed <strong>in</strong> frequency<br />

doma<strong>in</strong> by Fourier transform. Sharp pressure pulses of wide frequency range<br />

were used to m<strong>in</strong>imize signal distortion. Hence improved leaks localization <strong>and</strong><br />

monitor<strong>in</strong>g. <strong>Pressure</strong> pulses did not require complete stoppage of the flowl<strong>in</strong>e<br />

compared to water hammer events. Hence pressure pulses were presented as<br />

less <strong>in</strong>trusive than water hammer for detection, localization <strong>and</strong> monitor<strong>in</strong>g.<br />

Pothof <strong>and</strong> Clemens (2008) analyzed pressure traces for detect<strong>in</strong>g, locat<strong>in</strong>g<br />

<strong>and</strong> monitor<strong>in</strong>g gas pockets <strong>in</strong> multiphase s<strong>in</strong>gle pipel<strong>in</strong>es. A water hammer<br />

event was purposely generated <strong>and</strong> monitored at one end of the pipel<strong>in</strong>e. The<br />

pressure trace was then transposed <strong>in</strong> frequency doma<strong>in</strong> by Fourier transform.<br />

Pothof <strong>and</strong> Clemens observed that the second harmonic <strong>in</strong> the frequency<br />

doma<strong>in</strong> <strong>in</strong>dicated location <strong>and</strong> dimensions of a gas pocket. However,<br />

the method’s poor accuracy proved to be detrimental.<br />

Water hammer events <strong>and</strong> pressure pulses have been <strong>in</strong>vestigated for detect-<br />

16


2.4. Flow Meter<strong>in</strong>g<br />

<strong>in</strong>g, locat<strong>in</strong>g <strong>and</strong> monitor<strong>in</strong>g pipel<strong>in</strong>e features. Methods are divided <strong>in</strong> two<br />

categories, the first one be<strong>in</strong>g a direct <strong>in</strong>terpretation of pressure traces based<br />

on physical underst<strong>and</strong><strong>in</strong>g of pressure wave properties; the second be<strong>in</strong>g mathematical<br />

analysis of pressure traces <strong>and</strong> correlat<strong>in</strong>g experimental results with<br />

observations. Information from pressure pulses <strong>and</strong> water hammer events are<br />

identical, however, pressure pulses do not require complete stoppage of flowl<strong>in</strong>es.<br />

2.4 Flow Meter<strong>in</strong>g<br />

<strong>Pressure</strong> waves propagation velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e depends on elastic<br />

properties of the propagation medium. Acoustic velocity calculations require<br />

fluid density <strong>and</strong> isothermal compressibility, as well as pipel<strong>in</strong>e dimensions<br />

<strong>and</strong> material Young’s modulus. Therefore, acoustic velocity measurements can<br />

provide valuable <strong>in</strong>formation about pipel<strong>in</strong>e structure <strong>and</strong> fluid composition.<br />

<strong>Pressure</strong> waves propagation velocity <strong>in</strong> multiphase flowl<strong>in</strong>es was <strong>in</strong>vestigated<br />

by Gudmundsson et al. (1992) for meter<strong>in</strong>g fluid composition. Pulses were<br />

generated at one location <strong>and</strong> four pressure transducers were monitor<strong>in</strong>g pulse<br />

propagation. Cross correlation of pressure traces at the four locations <strong>and</strong><br />

acoustic velocity def<strong>in</strong>ition <strong>in</strong> multiphase flow (Dong <strong>and</strong> Gudmundsson, 1993)<br />

were used for calculat<strong>in</strong>g void fraction. The technique was repeatable <strong>and</strong> did<br />

not require complete stoppage of the flowl<strong>in</strong>e.<br />

Multiphase flow meter<strong>in</strong>g was further developed <strong>and</strong> applied to gas lift wells<br />

by Gudmundsson et al. (2001). Only two pressure transducers were used for<br />

monitor<strong>in</strong>g a purposely generated water hammer event. Cross correlation <strong>and</strong><br />

travel time determ<strong>in</strong>ation were then used to calculate multiphase flow void<br />

fraction. Direct <strong>in</strong>terpretation of pressure traces was used. However, water<br />

hammer required a complete stoppage of the flowl<strong>in</strong>e dur<strong>in</strong>g test<strong>in</strong>g.<br />

<strong>Pressure</strong> wave propagation velocity <strong>in</strong> gas pipel<strong>in</strong>es was analysed <strong>in</strong> AGA-9<br />

(2007). <strong>Pressure</strong> pulses are generated at one emitter <strong>and</strong> received at a distant<br />

receiver. <strong>Pressure</strong> waves travel times were measured <strong>and</strong> wave propagation<br />

velocities deduced from distance between transducers. Stream <strong>and</strong> counterstream<br />

wave propagation velocities were used for calculat<strong>in</strong>g flow<strong>in</strong>g velocities.<br />

Large b<strong>and</strong>width pulses improved travel time detection at receivers.<br />

Gas flow meter<strong>in</strong>g can also be performed analytically by calculat<strong>in</strong>g the acoustic<br />

velocity us<strong>in</strong>g AGA-10 (2003) <strong>and</strong> accurate gas composition. <strong>Pressure</strong><br />

pulses travel times as well as flow velocities were then calculated from acoustic<br />

velocities. Deviation between calculated <strong>and</strong> measured acoustic velocities<br />

must be with<strong>in</strong> ±0.2 % <strong>and</strong> must be verified dur<strong>in</strong>g calibration of the apparatus.<br />

17


Chapter 2. <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Pipel<strong>in</strong>es<br />

Distances between emitters <strong>and</strong> receivers are the ma<strong>in</strong> uncerta<strong>in</strong>ties <strong>in</strong> gas ultrasonic<br />

flow meter<strong>in</strong>g s<strong>in</strong>ce pressure <strong>and</strong> temperature <strong>in</strong>fluence pipel<strong>in</strong>e materials,<br />

thus modify<strong>in</strong>g distances between transducers. However, gas ultrasonic<br />

flow meter<strong>in</strong>g resolution is 0.001 ms −1 . Multiphase flow <strong>and</strong> gas flow meter<strong>in</strong>g<br />

are based on acoustic velocity determ<strong>in</strong>ation from pressure wave travel time.<br />

Large b<strong>and</strong>width pressure pulses improve travel time measurements, hence<br />

acoustic velocity determ<strong>in</strong>ations.<br />

Berrebi et al. (2004) studied ultrasonic flow meter<strong>in</strong>g accuracy <strong>and</strong> <strong>in</strong>vestigated<br />

effects of pulsat<strong>in</strong>g flow on measurements. Pulsat<strong>in</strong>g flows were generated from<br />

dynamic <strong>in</strong>stallations such as pumps or valves. Pulsat<strong>in</strong>g flows generated oscillation<br />

of pressure wave signals <strong>and</strong> affected travel time determ<strong>in</strong>ation, hence<br />

acoustic <strong>and</strong> flow velocity measurements. Berrebi et al. proposed a simple<br />

filter for reduc<strong>in</strong>g ultrasonic flow measurement errors. Berrebi et al. also concluded<br />

that no relevant modell<strong>in</strong>g of pulsations effect on wave propagation was<br />

available.<br />

18


Chapter 3<br />

Modell<strong>in</strong>g of <strong>Pressure</strong> Wave<br />

Propagation<br />

3.1 Propagation<br />

3.1.1 Modell<strong>in</strong>g Equations<br />

Rapid pressure transients can be generated by sudden valve actuation or pump<br />

failure, for <strong>in</strong>stance. The rapidness of transient phenomenon is relative to the<br />

travel period of the pressure wavefront, or to the pressure transient wavelength.<br />

Wavelength def<strong>in</strong>es distance between wavefront <strong>and</strong> wavetail <strong>and</strong> depends on<br />

the angular frequency, ω, <strong>and</strong> propagation velocity, a, of the transient phenomenon.<br />

The wavelength is expressed by (Lighthill, 1978):<br />

λ = ω<br />

a (3.1)<br />

2π<br />

Water hammer describes transient events <strong>in</strong>duced by sudden valve actuation.<br />

The travel period is the difference <strong>in</strong> time between the beg<strong>in</strong>n<strong>in</strong>g of the actuation<br />

<strong>and</strong> the <strong>in</strong>stant the wavefront has travelled back for the first time to the<br />

valve after reflection from the end element. The pressure transient wavelength<br />

is then twice the distance between the valve <strong>and</strong> the pipel<strong>in</strong>e end element.<br />

Water hammer phenomena are rapid if the valve actuation time is greatly<br />

shorter than one travel period.<br />

In an <strong>in</strong>duced pressure wave event generated by piston back <strong>and</strong> forth movement,<br />

the wavelength is the product of the acoustic velocity <strong>in</strong> the fluid-filled<br />

pipel<strong>in</strong>e, <strong>and</strong> the frequency of the piston. Wavelengths are long when greater<br />

than the pipel<strong>in</strong>e length, <strong>and</strong> short otherwise. A pressure wave can be overlooked<br />

as a succession of two water hammer phenomena generated by opposite<br />

valve actuations. Therefore, water hammer <strong>and</strong> pressure wave analysis can be<br />

performed us<strong>in</strong>g the same modell<strong>in</strong>g.<br />

19


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

<strong>Pressure</strong> wave propagation <strong>in</strong> fluid-filled pipel<strong>in</strong>e modell<strong>in</strong>g is based on the<br />

st<strong>and</strong>ard mass, momentum <strong>and</strong> energy conservation equations. Hydrocarbonfilled<br />

pipel<strong>in</strong>es are commonly several to tens of <strong>in</strong>ches <strong>in</strong> diameter <strong>and</strong> can<br />

cover hundreds of kilometers <strong>in</strong> length. Thus only one-dimensional modell<strong>in</strong>g is<br />

practically used. One-dimensional flow of a viscous <strong>and</strong> heat conduct<strong>in</strong>g fluid,<br />

is characterized by the follow<strong>in</strong>g mass, momentum <strong>and</strong> energy conservation<br />

equations, respectively (Chevray <strong>and</strong> Mathieu, 1993):<br />

∂ρ ∂<br />

+ (ρu) = 0 (3.2)<br />

∂t ∂x<br />

∂ ∂ −∂p ∂τµ<br />

(ρu) + (ρuu) = +<br />

∂t ∂x ∂x ∂x + τωπd + F (3.3)<br />

∂ ∂ ∂<br />

(ρE) + (ρuE) =<br />

∂t ∂x ∂x [(−p + τµ)u] − ∂h<br />

∂x<br />

(3.4)<br />

where, t, represents time, x, pipel<strong>in</strong>e axial direction, ρ, density, u, fluid velocity,<br />

p, pressure, τµ, viscous stress, τω, wall shear stress, d, pipel<strong>in</strong>e diameter,<br />

E, energy, h, heat transfer, <strong>and</strong>, F, external forces. Right h<strong>and</strong> terms <strong>in</strong> the<br />

momentum conservation equation, Eq. 3.3, def<strong>in</strong>e a source term, responsible<br />

for pressure wave attenuation. Attenuation mechanisms <strong>and</strong> the source term<br />

are attentively <strong>in</strong>vestigated <strong>in</strong> Chapter 6 <strong>and</strong> will not be detailed here<strong>in</strong>.<br />

Wall shear stress, viscous stress <strong>and</strong> heat transfer <strong>in</strong> one-dimensional modell<strong>in</strong>g<br />

are expressed by, respectively (Chevray <strong>and</strong> Mathieu, 1993):<br />

τω = −1<br />

ρfu|u| (3.5)<br />

2d<br />

τµ = 4<br />

3 µ∂u<br />

∂x<br />

h = −κ ∂Γ<br />

∂x<br />

(3.6)<br />

(3.7)<br />

where, µ, represents fluid dynamic viscosity, Γ, temperature, <strong>and</strong>, κ, heat<br />

transfer coefficient. Hydrocarbon liquids heat conductivity are generally low<br />

<strong>and</strong> rapid pressure transient processes are commonly assumed to be isothermal.<br />

The energy conservation equation is then omitted. Shear stress is also<br />

commonly neglected s<strong>in</strong>ce pressure transients are longitud<strong>in</strong>al waves. The<br />

system of equations, Eq. 3.2, Eq. 3.3 <strong>and</strong> Eq. 3.4, then becomes:<br />

∂ρ<br />

+ ρ∂u + u∂ρ = 0 (3.8)<br />

∂t ∂x ∂x<br />

ρ ∂u<br />

∂t<br />

−∂p 1<br />

+ ρu∂u = − ρfu|u| + F (3.9)<br />

∂x ∂x 2d<br />

20


3.1. Propagation<br />

Flow velocities <strong>in</strong> oil <strong>and</strong> gas transport pipel<strong>in</strong>es typically range from 2 ms −1<br />

to 4 ms −1 <strong>and</strong> 20 ms −1 to 40 ms −1 , respectively. Acoustic velocities <strong>in</strong> oilfilled<br />

<strong>and</strong> gas-filled pipel<strong>in</strong>es are typically of order 850 ms −1 <strong>and</strong> 300 ms −1 ,<br />

respectively. Therefore, the Mach number, Ma = u/a, where a is the acoustic<br />

velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e, is much less than unity <strong>in</strong> both oil <strong>and</strong> gas<br />

transport pipel<strong>in</strong>es. The acoustic velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e can be def<strong>in</strong>ed<br />

by (Lighthill, 1978):<br />

a =<br />

�<br />

∂p<br />

∂ρ<br />

(3.10)<br />

An order-of-magnitude analysis of Eq. 3.8 <strong>and</strong> Eq. 3.9 establishes that advective<br />

terms (u ∂ ·/∂x) <strong>in</strong> both equations are negligible because of the low Mach<br />

number. The low Mach number approximation <strong>and</strong> substitution of density by<br />

pressure, us<strong>in</strong>g the acoustic velocity def<strong>in</strong>ition Eq. 3.10, then reduces Eq. 3.8<br />

<strong>and</strong> Eq. 3.9 for pressure wave analysis to (Ghidaoui et al., 2005):<br />

∂p<br />

+ ρa2∂u = 0 (3.11)<br />

∂t ∂x<br />

∂u<br />

∂t<br />

1 ∂p<br />

+<br />

ρ ∂x<br />

−1<br />

= fu|u| + F (3.12)<br />

2d<br />

3.1.2 Transmission, Reflection<br />

<strong>Pressure</strong> waves propagate, transmit <strong>and</strong> reflect <strong>in</strong> fluid-filled pipel<strong>in</strong>es at the<br />

speed of sound. Practically, the acoustic velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>es depends<br />

on the fluid density <strong>and</strong> isothermal compressibility, KΓ, <strong>and</strong> on the<br />

pipel<strong>in</strong>e diameter <strong>and</strong> wall thickness, e, <strong>and</strong> the wall material Young’s modulus,<br />

E. The acoustic velocity <strong>in</strong> a fluid-filled pipel<strong>in</strong>e is expressed by (Pa<strong>in</strong>,<br />

2005):<br />

�<br />

�<br />

� 1<br />

a = � �<br />

�<br />

ρ KΓ + d<br />

� (3.13)<br />

Y e<br />

The product of the fluid density <strong>and</strong> the wave acoustic velocity <strong>in</strong> the fluidfilled<br />

pipel<strong>in</strong>e def<strong>in</strong>es the propagation medium acoustic impedance, Z (Lighthill,<br />

1978):<br />

Z = ρa (3.14)<br />

Acoustic <strong>in</strong>terfaces are virtual borderl<strong>in</strong>es between two pipel<strong>in</strong>e sections of different<br />

acoustic impedances. Changes <strong>in</strong> impedance can orig<strong>in</strong>ate from changes<br />

21


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

<strong>in</strong> fluid density, due to pressure or temperature, or acoustic velocity, due to<br />

pipel<strong>in</strong>e geometry or material properties. Changes <strong>in</strong> acoustic impedance can<br />

also result from a comb<strong>in</strong>ation of both changes <strong>in</strong> density <strong>and</strong> acoustic velocity.<br />

Reflection <strong>and</strong> transmission coefficients for a pressure wave propagat<strong>in</strong>g<br />

from a medium a to a medium b, as illustrated <strong>in</strong> Figure 3.1, are, respectively<br />

(Pa<strong>in</strong>, 2005):<br />

Ra→b = pR<br />

pI<br />

Ta→b = pT<br />

pI<br />

= Zb − Za<br />

Zb + Za<br />

= 2Zb<br />

Zb + Za<br />

pI<br />

pR<br />

a b<br />

Figure 3.1: Acoustic Interface Schematic Representation Between<br />

Media a <strong>and</strong> b<br />

pT<br />

(3.15)<br />

(3.16)<br />

The reflection coefficient is the ratio of the reflected pressure wave amplitude,<br />

pR, <strong>and</strong> the <strong>in</strong>cident pressure wave amplitude, pI. The transmission coefficient<br />

is the ratio of the transmitted pressure wave amplitude, pT, <strong>and</strong> the <strong>in</strong>cident<br />

pressure wave amplitude. Reflection <strong>and</strong> transmission coefficients, Eq. 3.15<br />

<strong>and</strong> Eq. 3.16, are def<strong>in</strong>ed as pressure ratios. However, a second set of reflection<br />

<strong>and</strong> transmission coefficients relative to fluid velocity can also be derived <strong>and</strong><br />

are expressed by (Pa<strong>in</strong>, 2005):<br />

Ra→b = uR<br />

uI<br />

Ta→b = uT<br />

uI<br />

= −Ra→b<br />

= Za<br />

Ta→b<br />

Zb<br />

(3.17)<br />

(3.18)<br />

Influence of an acoustic <strong>in</strong>terface on a square pressure wave propagat<strong>in</strong>g from a<br />

high acoustic impedance medium a to a low acoustic impedance b is illustrated<br />

<strong>in</strong> Figure 3.2. Both transmitted <strong>and</strong> reflected waves conserve square waveforms<br />

although the transmitted pressure wave has a lower amplitude <strong>and</strong> shorter<br />

22


3.1. Propagation<br />

wavelength than the <strong>in</strong>cident pressure wave (T < 1). The reflected pressure<br />

wave has the same wavelength as the <strong>in</strong>cident pressure wave but has a negative<br />

amplitude (R < 0).<br />

Changes <strong>in</strong> wavelength are due to the difference <strong>in</strong> acoustic impedance before<br />

<strong>and</strong> after the <strong>in</strong>terface. The pressure wave front propagates <strong>in</strong> medium b with<br />

a lower impedance than medium a, therefore propagat<strong>in</strong>g slower <strong>in</strong> medium b<br />

than <strong>in</strong> medium a <strong>and</strong> with a lower amplitude.<br />

Influence of an acoustic <strong>in</strong>terface on a square pressure wave propagat<strong>in</strong>g from a<br />

low acoustic impedance medium a to a high acoustic impedance b is illustrated<br />

<strong>in</strong> Figure 3.3. Both transmitted <strong>and</strong> reflected waves conserve square waveforms<br />

although the transmitted pressure wave has a greater amplitude <strong>and</strong> longer<br />

wavelength than the <strong>in</strong>cident pressure wave (T > 1). The reflected pressure<br />

wave has the same wavelength as the <strong>in</strong>cident pressure wave but has a lower<br />

amplitude (R < 1).<br />

Changes <strong>in</strong> wavelength are due to the difference <strong>in</strong> acoustic impedance before<br />

<strong>and</strong> after the <strong>in</strong>terface. The pressure wave front propagates <strong>in</strong> medium b with<br />

a greater impedance than medium a, therefore propagat<strong>in</strong>g faster <strong>in</strong> medium<br />

b than <strong>in</strong> medium a <strong>and</strong> with a greater amplitude.<br />

Evolution of transmission coefficient with the ratio of acoustic impedances<br />

between media a <strong>and</strong> b is shown <strong>in</strong> Figure 3.4. Transmission coefficient values<br />

range from 0 to +2, asymptotically. The transmission coefficient is 0 when<br />

the acoustic impedance <strong>in</strong> medium b is negligible compared to the acoustic<br />

impedance <strong>in</strong> medium a. The transmission coefficient is +2 when the acoustic<br />

impedance <strong>in</strong> medium b is much greater compared to the acoustic impedance<br />

<strong>in</strong> medium a.<br />

Evolution of reflection coefficient with the ratio of acoustic impedances between<br />

media a <strong>and</strong> b is shown <strong>in</strong> Figure 3.5. Reflection coefficient values<br />

range from −1 to +1, asymptotically. The reflection coefficient is −1 when<br />

the acoustic impedance <strong>in</strong> medium b is negligible compared to the acoustic<br />

impedance <strong>in</strong> medium a. The reflection coefficient is +1 when the acoustic<br />

impedance <strong>in</strong> medium b is much greater compared to the acoustic impedance<br />

<strong>in</strong> medium a.<br />

A pressure wave transmits totally from medium a to medium b when acoustic<br />

impedances are equal, Figure 3.4. The reflection coefficient equals zero <strong>in</strong><br />

such case, Figure 3.5. Reflected pressure waves relative pressure can be with<br />

or without phase change, but their amplitudes are limited by the <strong>in</strong>cident<br />

pressure wave amplitude. However, transmitted pressure waves always have<br />

the same phase as the <strong>in</strong>cident pressure wave, but their amplitude can be up<br />

to twice the <strong>in</strong>cident amplitude.<br />

23


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Incident →<br />

−0.5<br />

0 L/2 L<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

← Reflected<br />

Transmitted →<br />

−0.5<br />

0 L/2<br />

Distance [m]<br />

L<br />

Figure 3.2: Influence of Acoustic Interface on <strong>Pressure</strong> Wave<br />

Propagation. Medium a between 0 <strong>and</strong> L/2. Medium b<br />

between L/2 <strong>and</strong> L. Za > Zb<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Incident →<br />

−0.5<br />

0 L/2 L<br />

1.5<br />

1.0<br />

0.5<br />

← Reflected<br />

0.0<br />

−0.5<br />

0 L/2<br />

Distance [m]<br />

Transmitted →<br />

L<br />

Figure 3.3: Influence of Acoustic Interface on <strong>Pressure</strong> Wave<br />

Propagation. Medium a between 0 <strong>and</strong> L/2. Medium b<br />

between L/2 <strong>and</strong> L. Za < Zb<br />

24


Ta→b<br />

2.00<br />

1.75<br />

1.50<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

10 −3<br />

10 −2<br />

10 −1<br />

10 0<br />

Zb/Za<br />

10 1<br />

10 2<br />

3.1. Propagation<br />

Figure 3.4: Evolution of Transmission Coefficient Aga<strong>in</strong>st Ratio of<br />

Acoustic Impedances<br />

Ra→b<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

−0.25<br />

−0.50<br />

−0.75<br />

−1.00<br />

10 −3<br />

10 −2<br />

10 −1<br />

10 0<br />

Zb/Za<br />

Figure 3.5: Evolution of Reflection Coefficient Aga<strong>in</strong>st Ratio of<br />

Acoustic Impedances<br />

25<br />

10 1<br />

10 2<br />

10 3<br />

10 3


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

3.2 Boundary Conditions<br />

3.2.1 Static Boundary Conditions<br />

Pipe walls, tanks or pumps represent static boundary conditions. Pumps <strong>and</strong><br />

tanks are characterized by fixed pressure values after flow establishment. Pipe<br />

walls are characterized by a zero fluid velocity. Static boundary conditions are<br />

categorized <strong>in</strong>to open <strong>and</strong> close by the reflection coefficient value. Another way<br />

to categorize boundary conditions is by soft wall or hard wall, correspond<strong>in</strong>g<br />

to open <strong>and</strong> close boundary, respectively.<br />

Reflections of a square <strong>in</strong>cident pressure wave at closed <strong>and</strong> open boundaries<br />

are illustrated <strong>in</strong> Figure 3.6. A pressure wave reflects totally without phase<br />

change at a close boundary (R = 1, Figure 3.5), or hard wall, as shown<br />

<strong>in</strong> Figure 3.6 b). A pressure wave reflects totally with phase change at an<br />

open boundary (R = −1, Figure 3.5), or soft wall, as shown <strong>in</strong> Figure 3.6 c).<br />

Fluid velocity reflection coefficient, Eq. 3.17, is opposite of pressure reflection<br />

coefficient. Thus, fluid velocity signals are phase shifted compared to pressure<br />

signals as observed <strong>in</strong> Figure 3.6 d) <strong>and</strong> e).<br />

The transmission coefficient for a closed boundary (R = 1, Figure 3.5) equals<br />

+2 (Figure 3.4). However, the fluid medium does not exist beyond the boundary.<br />

Another way to express this is that a pressure wave of the same amplitude<br />

as the <strong>in</strong>cident pressure value will reflect totally <strong>in</strong>to the fluid-filled pipel<strong>in</strong>e<br />

whereas a pressure wave twice as large will propagate <strong>in</strong>to the physical boundary.<br />

<strong>Pressure</strong> waves generated <strong>in</strong> fluid-filled pipel<strong>in</strong>e can propagate further <strong>in</strong>to<br />

the support<strong>in</strong>g structure <strong>and</strong> weaken mechanical parts.<br />

3.2.2 Dynamic Boundary Conditions<br />

Tanks <strong>and</strong> pumps represent dynamic boundary conditions dur<strong>in</strong>g flow establishment,<br />

or transient phases. Tanks <strong>and</strong> pumps transients are characterized<br />

by changes <strong>in</strong> pressure <strong>and</strong> fluid velocity. <strong>Pressure</strong> or fluid velocity gradients<br />

can be smooth such that boundary conditions are characterized by a temporal<br />

function, or can be abrupt, thus requir<strong>in</strong>g transient calculation. Check valves<br />

are also dynamic boundaries dur<strong>in</strong>g actuation <strong>and</strong> represent one of the ma<strong>in</strong><br />

causes for rapid pressure transients.<br />

Water hammer pressure waves are created by check valves sudden closure or<br />

open<strong>in</strong>g. The pressure transient amplitude can be approximated by Joukowski’s<br />

expression (Wylie <strong>and</strong> Streeter, 1993):<br />

∆p = ρa∆u (3.19)<br />

where, ∆p, is pressure change generated by a flow velocity change, ∆u.<br />

26


(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

a)<br />

1.0<br />

0.0<br />

−1.0<br />

Incident →<br />

3.2. Boundary Conditions<br />

0 L/2 L<br />

b)<br />

1.0<br />

Closed Boundary<br />

0.0<br />

← Reflected<br />

−1.0<br />

c)<br />

1.0<br />

0.0<br />

−1.0<br />

0 L/2 L<br />

Open Boundary<br />

← Reflected<br />

0 L/2<br />

Distance [m]<br />

L<br />

(u − u 0 ) / u Wave<br />

(u − u 0 ) / u Wave<br />

d)<br />

1.0<br />

← Reflected<br />

0.0<br />

Closed Boundary<br />

−1.0<br />

−1.0<br />

0 L/2 L<br />

e)<br />

1.0<br />

Open Boundary<br />

0.0<br />

← Reflected<br />

0 L/2<br />

Distance [m]<br />

L<br />

Figure 3.6: <strong>Pressure</strong> <strong>and</strong> Fluid Velocity <strong>Waves</strong> Reflections From<br />

Open <strong>and</strong> Close Boundary Conditions<br />

a): Incident <strong>Pressure</strong> Wave.<br />

b): Reflected <strong>Pressure</strong> Wave from a Close Boundary.<br />

c): Reflected <strong>Pressure</strong> Wave from an Open Boundary.<br />

d): Reflected Velocity Wave from a Close Boundary.<br />

e): Reflected Velocity Wave from an Open Boundary.<br />

Joukowski’s equation approximates pressure transient amplitudes for complete<br />

closure or open<strong>in</strong>g. However, partial valve actuations are also responsible for<br />

transient phenomena. Time dependent pressure transient amplitudes can be<br />

calculated from valve actuation functions derived from geometrical considerations<br />

<strong>and</strong> valve actuation time (Wood <strong>and</strong> Jones, 1973).<br />

Circular gate, square gate, globe, needle, butterfly <strong>and</strong> ball valves are studied<br />

for uniform <strong>and</strong> accelerated actuation functions. Geometries <strong>and</strong> important<br />

dimensions of the six check valves are represented <strong>in</strong> Figure 3.7. Actuation<br />

functions are given <strong>in</strong> Table 3.1 where, Ξ, is the ratio of open area <strong>and</strong> total<br />

possible flow area, <strong>and</strong>, ξ, the ratio of displacement <strong>and</strong> total stroke for circular<br />

gate, square gate, globe <strong>and</strong> needle valves, or the open<strong>in</strong>g angle for butterfly<br />

<strong>and</strong> ball valves (Wood <strong>and</strong> Jones, 1973, 1974).<br />

27


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

x<br />

e)<br />

c)<br />

a)<br />

θ<br />

X<br />

x<br />

Figure 3.7: Schematic Representation of Check Valves.<br />

a) Circular Gate Valve. b) Square Gate Valve.<br />

c) Globe Valve. d) Needle Valve.<br />

e) Butterfly Valve. f) Ball Valve.<br />

X<br />

28<br />

X<br />

x<br />

B<br />

f)<br />

d)<br />

b)<br />

b<br />

X<br />

x<br />

θ


3.2. Boundary Conditions<br />

Table 3.1: Actuation Function Ξ for Circular Gate, Square Gate,<br />

Globe, Needle, Butterfly <strong>and</strong> Ball Valves for Dimensionless<br />

Actuation Time ξ (Eq. 3.20 to Eq. 3.23) (Wood <strong>and</strong><br />

Jones, 1973, 1974)<br />

Valve Type Actuation Function<br />

Circular Gate 1 − 2<br />

�<br />

arccos(ξ) − ξ<br />

π<br />

� 1 − ξ2 �<br />

1 −<br />

Square Gate<br />

1<br />

�<br />

�<br />

arccos(2ξ − 1) − (2ξ − 1) 1 − (2ξ − 1)<br />

π<br />

2<br />

�<br />

Globe ξ<br />

Needle 2ξ − ξ 2<br />

Butterfly<br />

cos (ξ)<br />

π<br />

1 + cos(ξ)<br />

+<br />

2<br />

�<br />

arcs<strong>in</strong>(−x) + 1<br />

�<br />

s<strong>in</strong> (2arcs<strong>in</strong>(−x)) −<br />

2<br />

�<br />

1<br />

arcs<strong>in</strong>(x) +<br />

π<br />

1<br />

2 s<strong>in</strong>(2arcs<strong>in</strong>(x))<br />

�<br />

�<br />

�B �2 s<strong>in</strong>(ξ) − 1<br />

b<br />

where x =<br />

1 + cos (ξ)<br />

�<br />

π<br />

�<br />

Ball 1 − cos − ξ<br />

2<br />

29


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

Uniform <strong>and</strong> accelerated actuation functions for l<strong>in</strong>early open<strong>in</strong>g valves are<br />

expressed by, respectively:<br />

ξ = 1 − t<br />

Tv<br />

� �2 t<br />

ξ = 1 −<br />

Tv<br />

(3.20)<br />

(3.21)<br />

Uniform <strong>and</strong> accelerated actuation functions for rotary open<strong>in</strong>g valves are<br />

expressed by, respectively:<br />

� �<br />

t<br />

ξ = θm<br />

(3.22)<br />

ξ = θm<br />

Tv<br />

�2 � t<br />

Tv<br />

(3.23)<br />

where, Tv, is valve actuation time, <strong>and</strong>, θm, complete actuation angle. Valve<br />

uniform <strong>and</strong> accelerated closures are shown <strong>in</strong> Figure 3.8 <strong>and</strong> Figure 3.9, respectively.<br />

For both closure functions, the circular gate valve flow area gradient<br />

evolves slowly dur<strong>in</strong>g the first 50 % of closure before <strong>in</strong>creas<strong>in</strong>g, whereas<br />

the flow area gradient for a ball valve is important dur<strong>in</strong>g the first 50 % of<br />

closure before decreas<strong>in</strong>g. Uniform closure of a globe valve is characterized by<br />

a constant gradient.<br />

Fluid velocity reductions for uniform closure of the six valves considered are<br />

shown <strong>in</strong> Figure 3.10. The fluid velocity reduces slowly until 80 % of the fluid<br />

<strong>in</strong>itial velocity, the fluid velocity then decreases rapidly until full closure. The<br />

fluid velocity reaches 80 % of the <strong>in</strong>itial velocity at 87 %, 81 %, 76 %, 71 %,<br />

65 % <strong>and</strong> 55 % of the actuation time for needle, circular gate, globe, square<br />

gate, ball <strong>and</strong> butterfly valves, respectively.<br />

Fluid velocity reductions for accelerated closure of the six valves considered<br />

are shown <strong>in</strong> Figure 3.11. The fluid velocity reduces slowly until 80 % of the<br />

fluid <strong>in</strong>itial velocity, the fluid velocity then decreases rapidly until full closure.<br />

The fluid velocity reaches 80 % of the <strong>in</strong>itial velocity at 93 %, 90 %, 87 %,<br />

81 %, 80 % <strong>and</strong> 74 % of the actuation time for needle, circular gate, globe,<br />

square gate, ball <strong>and</strong> butterfly valves, respectively.<br />

Practically, pressure waves orig<strong>in</strong>at<strong>in</strong>g from sudden valve actuations are generated<br />

ma<strong>in</strong>ly dur<strong>in</strong>g the last 20 % of valve uniform actuation, or the last<br />

10 % of valve accelerated actuation, as observed from Figure 3.10 <strong>and</strong> Figure<br />

3.11. Fluid velocity gradients determ<strong>in</strong>e also pressure waveforms. Needle<br />

valves generate sharper waveforms than circular gate, globe, square gate, ball<br />

<strong>and</strong> butterfly valves.<br />

30


χ [%]<br />

χ [%]<br />

100<br />

80<br />

60<br />

40<br />

3.2. Boundary Conditions<br />

20<br />

Needle Valve<br />

Circular Gate Valve<br />

Square Gate Valve<br />

Globe Valve<br />

Ball Valve<br />

Butterfly Valve<br />

0<br />

0 20 40 60 80 100<br />

t / T [%]<br />

v<br />

Figure 3.8: Valve Position Dur<strong>in</strong>g Uniform Closure<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Needle Valve<br />

Circular Gate Valve<br />

Square Gate Valve<br />

Globe Valve<br />

Ball Valve<br />

Butterfly Valve<br />

0<br />

0 20 40 60 80 100<br />

t / T [%]<br />

v<br />

Figure 3.9: Valve Position Dur<strong>in</strong>g Accelerated Closure<br />

31


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

u(t) / u 0 [%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Needle<br />

Gate − Circular<br />

Globe<br />

Gate − Square<br />

Ball<br />

Butterfly<br />

0<br />

0 20 40 60 80 100<br />

t / T [%]<br />

v<br />

Figure 3.10: Flow Velocity Across Valve Dur<strong>in</strong>g Uniform Closure<br />

u(t) / u 0 [%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Needle<br />

Gate − Circular<br />

Globe<br />

Gate − Square<br />

Ball<br />

Butterfly<br />

0<br />

0 20 40 60 80 100<br />

t / T [%]<br />

v<br />

Figure 3.11: Flow Velocity Across Valve Dur<strong>in</strong>g Accelerated Closure<br />

32


3.3 <strong>Pressure</strong> Wave Superposition<br />

3.3. <strong>Pressure</strong> Wave Superposition<br />

Several pressure waves orig<strong>in</strong>at<strong>in</strong>g from multiple sources can propagate simultaneously<br />

<strong>in</strong> a fluid-filled pipel<strong>in</strong>e. Orig<strong>in</strong>s of multiple pressure waves <strong>in</strong> a<br />

pipel<strong>in</strong>e can be multiple sources or multiple reflections from changes <strong>in</strong> acoustic<br />

impedance. Wave superposition can then take place <strong>in</strong> a pipel<strong>in</strong>e when<br />

several pressure waves arrive at a particular location at the exact same <strong>in</strong>stant.<br />

Wave superposition can also occur when a pressure wave superimposes<br />

on its own reflection at boundaries.<br />

<strong>Pressure</strong> wave superposition can be constructive or destructive (Pa<strong>in</strong>, 2005).<br />

Constructive wave superpositions take place when two waves of same phase<br />

encounter, or at closed boundaries when a pressure wave superimposes on<br />

its own reflection. Destructive wave superpositions occur when two waves of<br />

opposite phase encounter, or at open boundaries when a pressure superimposes<br />

on its own reflection. Here<strong>in</strong> presented pressure signals were obta<strong>in</strong>ed from<br />

numerical simulation us<strong>in</strong>g a f<strong>in</strong>ite volume scheme detailed <strong>and</strong> selected <strong>in</strong><br />

Chapter 5.<br />

Constructive wave superposition of two pressure waves of same amplitude <strong>and</strong><br />

phase <strong>and</strong> different waveforms is illustrated <strong>in</strong> Figure 3.12. The two pressure<br />

waves of same phase propagate <strong>in</strong> opposite direction, a), then superimpose,<br />

b), before cont<strong>in</strong>u<strong>in</strong>g separately unaffected by the superposition, c). Dur<strong>in</strong>g<br />

superposition, pressure wave amplitudes add to each other, thus creat<strong>in</strong>g a<br />

pressure peak of greater amplitude than either of the pressure waves.<br />

Destructive wave superposition of two pressure waves of same amplitude <strong>and</strong><br />

different waveforms <strong>and</strong> phases is illustrated <strong>in</strong> Figure 3.13. The two pressure<br />

waves of opposite phase propagate <strong>in</strong> opposite direction, a), then superimpose,<br />

b), before cont<strong>in</strong>u<strong>in</strong>g separately unaffected by the superposition, c). Dur<strong>in</strong>g<br />

superposition, pressure wave amplitudes subtract to each other, thus creat<strong>in</strong>g<br />

a pressure peak of lower amplitude than either of the pressure waves.<br />

33


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

a)<br />

2<br />

1<br />

0<br />

b)<br />

2<br />

1<br />

0<br />

c)<br />

2<br />

1<br />

0<br />

→ ←<br />

0 L/2 L<br />

←<br />

0 L/2 L<br />

← →<br />

0 L/2<br />

Distance [m]<br />

L<br />

Figure 3.12: Constructive Superposition of two <strong>Pressure</strong> <strong>Waves</strong><br />

a): Propagation <strong>in</strong> Opposite Direction Before Superposition<br />

b): Superposition, Additive Amplitudes<br />

c) Propagation <strong>in</strong> Opposite Direction After Superposition<br />

34<br />


(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

a)<br />

1<br />

0<br />

−1<br />

b)<br />

1<br />

0<br />

−1<br />

c)<br />

1<br />

0<br />

−1<br />

→<br />

3.3. <strong>Pressure</strong> Wave Superposition<br />

0 L/2 L<br />

←<br />

0 L/2 L<br />

←<br />

0 L/2<br />

Distance [m]<br />

L<br />

Figure 3.13: Destructive Superposition of two <strong>Pressure</strong> <strong>Waves</strong><br />

a: Propagation <strong>in</strong> Opposite Direction Before Superposition<br />

b): Superposition, Subtractive Amplitudes<br />

c): Propagation <strong>in</strong> Opposite Direction After Superposition<br />

35<br />

→<br />

←<br />


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

Constructive wave superposition of a pressure wave <strong>and</strong> its own reflection at a<br />

closed boundary is illustrated <strong>in</strong> Figure 3.14. The pressure wave hits the closed<br />

boundary <strong>and</strong> reflects without phase change (R = 1, Figure 3.5), Figure 3.14<br />

b); until half the wave length has reflected <strong>and</strong> superimposes on the <strong>in</strong>cident<br />

pressure wave, Figure 3.14 c). Dur<strong>in</strong>g maximum superposition, the pressure<br />

amplitude is twice the <strong>in</strong>cident pressure wave amplitude. The pressure wave<br />

then cont<strong>in</strong>ues reflect<strong>in</strong>g, Figure 3.14 d), until be<strong>in</strong>g totally reflected, Figure<br />

3.14 e).<br />

Destructive wave superposition of a pressure wave <strong>and</strong> its own reflection at an<br />

open boundary is illustrated <strong>in</strong> Figure 3.15. The pressure wave hits the open<br />

boundary <strong>and</strong> reflects with phase change (R = −1, Figure 3.5), Figure 3.15<br />

b); until half the wave length has reflected <strong>and</strong> superimposes on the <strong>in</strong>cident<br />

pressure wave, Figure 3.15 c). Dur<strong>in</strong>g maximum superposition, the pressure<br />

amplitude is zero as the two amplitudes cancel each other. The pressure wave<br />

then cont<strong>in</strong>ues reflect<strong>in</strong>g, Figure 3.15 d), until be<strong>in</strong>g totally reflected, Figure<br />

3.15 e).<br />

Practically, pressure transducers must be located at distances greater than half<br />

a wavelength from boundaries or <strong>in</strong>terfaces for captur<strong>in</strong>g pressure waves <strong>in</strong>dividually<br />

without superposition on their own reflections. <strong>Pressure</strong> traces would<br />

otherwise be difficult to <strong>in</strong>terpret <strong>and</strong> could lead to mislead<strong>in</strong>g conclusions.<br />

36


(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

a)<br />

2<br />

1<br />

0<br />

b)<br />

2<br />

1<br />

0<br />

e)<br />

2<br />

1<br />

0<br />

Incident →<br />

3.3. <strong>Pressure</strong> Wave Superposition<br />

0 L/2 L<br />

→<br />

←<br />

0 L/2 L<br />

c)<br />

2<br />

1<br />

0<br />

Closed Boundary<br />

→←<br />

0 L/2 L<br />

d)<br />

2<br />

1<br />

0<br />

→<br />

←<br />

0 L/2 L<br />

← Reflected<br />

0 L/2<br />

Distance [m]<br />

L<br />

Figure 3.14: Constructive Superposition of a <strong>Pressure</strong> Wave on its<br />

own Reflection at a Closed Boundary<br />

a): Incident <strong>Pressure</strong> Wave<br />

b): Beg<strong>in</strong>n<strong>in</strong>g of <strong>Pressure</strong> Wave Superposition on its Own Reflection<br />

c): Total Superposition of <strong>Pressure</strong> Wave <strong>and</strong> its Own Reflection.<br />

Additive Amplitudes<br />

d): End of <strong>Pressure</strong> Wave Superposition on its Own Reflection<br />

e): Reflected <strong>Pressure</strong> Wave<br />

37


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

a)<br />

1<br />

0<br />

−1<br />

b)<br />

1<br />

0<br />

−1<br />

e)<br />

1<br />

0<br />

−1<br />

Incident →<br />

0 L/2 L<br />

←<br />

→<br />

0 L/2 L<br />

c)<br />

1<br />

0<br />

−1<br />

Open Boundary<br />

→<br />

←<br />

0 L/2 L<br />

d)<br />

1<br />

0<br />

−1<br />

→<br />

←<br />

0 L/2 L<br />

← Reflected<br />

0 L/2<br />

Distance [m]<br />

L<br />

Figure 3.15: Destructive Superposition of a <strong>Pressure</strong> Wave on its<br />

own Reflection at an Open Boundary<br />

a): Incident <strong>Pressure</strong> Wave<br />

b): Beg<strong>in</strong>n<strong>in</strong>g of <strong>Pressure</strong> Wave Superposition on its Own Reflection<br />

c): Total Superposition of <strong>Pressure</strong> Wave <strong>and</strong> its Own Reflection.<br />

Subtractive Amplitudes<br />

d): End of <strong>Pressure</strong> Wave Superposition on its Own Reflection<br />

e): Reflected <strong>Pressure</strong> Wave<br />

38


3.3.1 St<strong>and</strong><strong>in</strong>g <strong>Waves</strong><br />

3.3. <strong>Pressure</strong> Wave Superposition<br />

St<strong>and</strong><strong>in</strong>g waves arise when <strong>in</strong>cident <strong>and</strong> reflected pressure waves superimpose.<br />

St<strong>and</strong><strong>in</strong>g wave patterns are obta<strong>in</strong>ed for longitud<strong>in</strong>al waves such as pressure<br />

waves that reflect completely from static boundaries. Closed <strong>and</strong> open boundary<br />

conditions give dist<strong>in</strong>ct patterns <strong>and</strong> sets of nodal po<strong>in</strong>ts. Nodal po<strong>in</strong>ts<br />

are locations where superposition of <strong>in</strong>cident <strong>and</strong> reflected waves are zero <strong>in</strong><br />

amplitude.<br />

First <strong>and</strong> third st<strong>and</strong><strong>in</strong>g wave harmonics from a pressure wave propagat<strong>in</strong>g <strong>in</strong><br />

a pipel<strong>in</strong>e closed at both ends are shown <strong>in</strong> Figure 3.16. Incident <strong>and</strong> reflected<br />

waves from closed boundaries have same phases (R = 1, Figure 3.5), thus<br />

maximum amplitude of st<strong>and</strong><strong>in</strong>g waves are twice the <strong>in</strong>cident wave amplitude.<br />

Boundaries are anti-nodal po<strong>in</strong>ts. Another way to express this is that pressure<br />

wave superpositions with their own reflections at boundaries generate local<br />

pressure peaks.<br />

First <strong>and</strong> second st<strong>and</strong><strong>in</strong>g wave harmonics from a pressure wave propagat<strong>in</strong>g<br />

<strong>in</strong> a pipel<strong>in</strong>e open at both ends are shown <strong>in</strong> Figure 3.17. Incident <strong>and</strong> reflected<br />

waves from open boundaries have opposite phases (R = −1, Figure<br />

3.5), thus maximum amplitude of st<strong>and</strong><strong>in</strong>g waves is the same as the <strong>in</strong>cident<br />

wave amplitude. Boundaries are nodal po<strong>in</strong>ts. Another way to express this is<br />

that <strong>in</strong>cident pressure waves cancel their own reflection at open boundaries.<br />

Check valves <strong>and</strong> non-return valves are neither closed nor open boundary conditions.<br />

Therefore, no st<strong>and</strong>ard st<strong>and</strong><strong>in</strong>g wave patterns can illustrate pressure<br />

wave superpositions. Nodal po<strong>in</strong>ts obta<strong>in</strong>ed from valve characteristics are not<br />

fixed neither. Instead, nodal po<strong>in</strong>ts move <strong>and</strong> are limited by the fully open<br />

<strong>and</strong> fully closed conditions.<br />

39


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

(p − p 0 ) / p Wave<br />

2<br />

1<br />

0<br />

−1<br />

−2<br />

n = 1<br />

n = 3<br />

0 L/2<br />

x / L<br />

L<br />

Figure 3.16: First <strong>and</strong> Third Harmonics for <strong>Pressure</strong> St<strong>and</strong><strong>in</strong>g Wave<br />

<strong>in</strong> a Pipel<strong>in</strong>e Closed at Both Ends<br />

(p − p 0 ) / p Wave<br />

1<br />

0<br />

−1<br />

n = 1<br />

n = 2<br />

0 L/2<br />

x / L<br />

L<br />

Figure 3.17: First <strong>and</strong> Second Harmonics for <strong>Pressure</strong> St<strong>and</strong><strong>in</strong>g<br />

Wave <strong>in</strong> Pipel<strong>in</strong>e Open at Both Ends<br />

40


3.4 Two-Dimensional Modell<strong>in</strong>g<br />

3.4. Two-Dimensional Modell<strong>in</strong>g<br />

<strong>Pressure</strong> transient modell<strong>in</strong>g generally focuses on pressure amplitudes <strong>and</strong> onedimensional<br />

modell<strong>in</strong>g is preferred because of pipel<strong>in</strong>e lengths. However, twodimensional<br />

pressure transient modell<strong>in</strong>g can also be performed <strong>and</strong> focuses<br />

on fluid velocity by establish<strong>in</strong>g velocity profiles <strong>in</strong> pipel<strong>in</strong>e cross sections.<br />

Assum<strong>in</strong>g axisymmetrical homogeneous flow <strong>in</strong> a circular pipel<strong>in</strong>e, where the<br />

velocity radial component is small compared to the longitud<strong>in</strong>al component,<br />

<strong>and</strong> where the pressure is averaged <strong>in</strong> the cross-section, mass <strong>and</strong> momentum<br />

conservation equations are, respectively (Nathan et al., 1988):<br />

∂p<br />

∂t<br />

∂p<br />

+ ux + ρa2∂ux<br />

∂x ∂x<br />

∂ux<br />

∂t +ux<br />

∂ux<br />

∂x<br />

∂p<br />

= −1<br />

p ∂x<br />

= 0 (3.24)<br />

∂z<br />

−g<br />

∂x +µ<br />

�<br />

∂2ux ∂r<br />

�<br />

1 ∂ux<br />

+ − 2 r ∂r<br />

1 ∂<br />

r ∂r<br />

�<br />

ru ′<br />

xu ′ �<br />

r (3.25)<br />

where, u ′ xu ′ r, is Reynolds tensor. Thus, turbulence modell<strong>in</strong>g is required <strong>in</strong><br />

two-dimensional pressure wave modell<strong>in</strong>g. However, the coupl<strong>in</strong>g between<br />

pressure <strong>and</strong> fluid velocity is of particular <strong>in</strong>terest <strong>in</strong> two-dimensional modell<strong>in</strong>g<br />

s<strong>in</strong>ce fluid velocity profiles can be determ<strong>in</strong>ed dur<strong>in</strong>g rapid pressure<br />

events. Dur<strong>in</strong>g a pressure transient event, the flow pulsates <strong>and</strong> the velocity<br />

profile can be described by (Uchida, 1956; Yamanaka et al., 2002)<br />

ur = u0<br />

⎛<br />

⎝1 − 2<br />

i 3/2 Π<br />

J1<br />

�<br />

i3/2 �<br />

Π2r/d<br />

�<br />

i3/2 ⎞<br />

� ⎠<br />

Π<br />

J0<br />

−1 ⎛<br />

⎝1 −<br />

J0<br />

�<br />

i3/2 �<br />

Π2r/d<br />

�<br />

i3/2 ⎞<br />

� ⎠ e<br />

Π<br />

iωt (3.26)<br />

where, i, is the imag<strong>in</strong>ary number, J0 <strong>and</strong> J1, Bessel functions of first <strong>and</strong><br />

second order, respectively, <strong>and</strong> Π, dimensionless angular frequency expressed<br />

by (Uchida, 1956):<br />

�<br />

ω<br />

Π = d<br />

(3.27)<br />

µ<br />

Piston back <strong>and</strong> forth movements <strong>in</strong> a fluid-filled pipel<strong>in</strong>e <strong>in</strong>duce pressure<br />

waves <strong>and</strong> pulsat<strong>in</strong>g flows. A lam<strong>in</strong>ar pulsat<strong>in</strong>g flow generated by a piston<br />

back <strong>and</strong> forth mov<strong>in</strong>g piston are illustrated <strong>in</strong> Figure 3.18 <strong>and</strong> Figure 3.19 for<br />

lam<strong>in</strong>ar <strong>and</strong> turbulent flow, respectively. In both cases, flow velocity changes<br />

are driven by outer boundary regions. The fluid flow <strong>in</strong>creases first, then<br />

reaches an extreme before <strong>in</strong>vert<strong>in</strong>g, reach<strong>in</strong>g the opposite extreme <strong>and</strong> com<strong>in</strong>g<br />

back to rest aga<strong>in</strong>.<br />

41<br />

J0


Chapter 3. Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

2 r / d<br />

2 r / d<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

φ = 0<br />

−1.0<br />

−1 0 1<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

φ = π<br />

−1.0<br />

−1 0<br />

u / u<br />

x 0<br />

1<br />

φ = π / 4<br />

−1 0 1<br />

φ = 5 π / 4<br />

−1 0 1<br />

u x / u 0<br />

φ = π / 2<br />

−1 0 1<br />

φ = 3 π / 2<br />

−1 0 1<br />

u x / u 0<br />

φ = 3 π / 4<br />

−1 0 1<br />

φ = 7 π /4<br />

−1 0 1<br />

u x / u 0<br />

Figure 3.18: Two-Dimensional Lam<strong>in</strong>ar Velocity Profile Evolution <strong>in</strong><br />

a Fluid-Filled Pipel<strong>in</strong>e at Zero Initial Velocity Dur<strong>in</strong>g<br />

One Angular <strong>Pressure</strong> Wave period<br />

2 r / d<br />

2 r / d<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

φ = 0<br />

−1.0<br />

−1 0 1<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

φ = π<br />

−1.0<br />

−1 0<br />

u / u<br />

x 0<br />

1<br />

φ = π / 4<br />

−1 0 1<br />

φ = 5 π / 4<br />

−1 0 1<br />

u x / u 0<br />

φ = π / 2<br />

−1 0 1<br />

φ = 3 π / 2<br />

−1 0 1<br />

u x / u 0<br />

φ = 3 π / 4<br />

−1 0 1<br />

φ = 7 π /4<br />

−1 0 1<br />

u x / u 0<br />

Figure 3.19: Two-Dimensional Turbulent Velocity Profile Evolution<br />

<strong>in</strong> a Fluid-Filled Pipel<strong>in</strong>e at Zero Initial Velocity Dur<strong>in</strong>g<br />

One Angular <strong>Pressure</strong> Wave period<br />

42


Chapter 4<br />

Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle<br />

Phase Fluids<br />

4.1 Basic Relationships<br />

4.1.1 Def<strong>in</strong>ition<br />

A wave is a disturbance that transports energy <strong>in</strong> a medium. A wave propagates,<br />

transmits <strong>and</strong> reflects accord<strong>in</strong>g to the elastic properties of the propagation<br />

medium. In case of a pressure wave, the disturbance amplitude is<br />

pressure, p, <strong>and</strong> the wave propagation speed is acoustic velocity, a, <strong>in</strong> the<br />

propagation medium. A pressure wave that propagates through a three dimensional<br />

homogeneous medium, at a s<strong>in</strong>gle wave speed, <strong>and</strong> that conserves<br />

energy, can be described by the l<strong>in</strong>ear wave equation (Lighthill, 1978):<br />

∂2p ∂t2 = a2 ∂2p ∂x2 (4.1)<br />

The pressure wave propagation process described by Eq. 4.1, conserves energy,<br />

is <strong>in</strong>dependent of waveform or direction of propagation. Another way to<br />

express this is that the wave propagation process is adiabatic <strong>and</strong> reversible;<br />

that is isentropic. Thus, the acoustic velocity relates to pressure waves characterized<br />

by small gradients such that entropy, S, is constant. The acoustic<br />

velocity is then expressed by (Lighthill, 1978):<br />

��∂p �<br />

a =<br />

∂ρ<br />

S<br />

(4.2)<br />

where, a, is acoustic velocity, p, is pressure, ρ, is density, <strong>and</strong> subscript S, denotes<br />

isentropic conditions. Entropy is a thermodynamic function that can not<br />

43


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

be measured <strong>and</strong> depends on heat <strong>and</strong> temperature. Therefore, thermal conditions<br />

can substitute entropic conditions s<strong>in</strong>ce state variables are dependent.<br />

Such substitution requires the enthalpy <strong>and</strong> <strong>in</strong>ternal energy thermodynamic<br />

equations <strong>and</strong> are expressed by the second law of thermodynamics (Chevray<br />

<strong>and</strong> Mathieu, 1993):<br />

dH = ΓdS + V dp (4.3)<br />

dU = ΓdS − pdV (4.4)<br />

where, H, is enthalpy, Γ, temperature, V , volume <strong>and</strong>, U, <strong>in</strong>ternal energy. For<br />

gases, changes <strong>in</strong> enthalpy <strong>and</strong> <strong>in</strong>ternal energy are also equal to the product<br />

of changes <strong>in</strong> temperatures <strong>and</strong> heat capacities at constant pressure, cp, <strong>and</strong><br />

constant volume, cv, respectively. For gases, Eq. 4.3 <strong>and</strong> Eq. 4.4 can be<br />

reformulated (Chevray <strong>and</strong> Mathieu, 1993):<br />

dH = cpdΓ (4.5)<br />

dU = cvdΓ (4.6)<br />

For gases, comb<strong>in</strong><strong>in</strong>g Eq. 4.3 <strong>and</strong> Eq. 4.5, <strong>and</strong> Eq. 4.4 <strong>and</strong> Eq. 4.6, for<br />

an isothermal process (dΓ = 0) <strong>and</strong> for an isentropic process (dS = 0) give,<br />

respectively:<br />

−V<br />

p<br />

−V<br />

p<br />

dp<br />

dV<br />

dp<br />

dV<br />

= 1 (4.7)<br />

= cp<br />

cv<br />

(4.8)<br />

The ratio of heat capacities at constant pressure, cp, <strong>and</strong> constant volume, cv,<br />

def<strong>in</strong>es a gas adiabatic <strong>in</strong>dex, k. A gas adiabatic <strong>in</strong>dex is function of pressure,<br />

temperature <strong>and</strong> fluid composition. Of particular <strong>in</strong>terest, gas adiabatic<br />

<strong>in</strong>dexes relate gas isentropic <strong>and</strong> isothermal processes such that the acoustic<br />

velocity expression, Eq. 4.2 can be reformulated for gas:<br />

a =<br />

�<br />

k<br />

� �<br />

∂p<br />

∂ρ Γ<br />

(4.9)<br />

The above equation, Eq. 4.9, expresses the acoustic velocity <strong>in</strong> gases only.<br />

Equivalent of Eq. 4.3 <strong>and</strong> Eq. 4.4 for liquids can not be expressed simply.<br />

Thus, relation between isothermal <strong>and</strong> isentropic processes lead<strong>in</strong>g to a simple<br />

acoustic velocity expression for liquids can not be simply derived. Nonetheless,<br />

a common acoustic velocity expression can be formulated for both gases <strong>and</strong><br />

liquids when relat<strong>in</strong>g the isentropic partial derivative of pressure with density<br />

<strong>in</strong> Eq. 4.2 with fluid compressibility.<br />

44


4.1.2 Fluid Compressibility<br />

4.1. Basic Relationships<br />

Fluid compressibility is a measure of relative changes <strong>in</strong> volume for changes <strong>in</strong><br />

pressure. Because density equals the ratio of mass <strong>and</strong> volume occupied, <strong>and</strong><br />

that a fluid mass does not change with pressure, fluid compressibility can also<br />

be expressed as a measure of changes <strong>in</strong> density under pressure constra<strong>in</strong>ts.<br />

Fluid compressibility, K, depends on fluid composition, pressure, temperature<br />

<strong>and</strong> thermodynamic process associated, <strong>and</strong> is expressed by (Leighton, 1994):<br />

K = 1<br />

ρ<br />

� �<br />

∂ρ<br />

∂p<br />

(4.10)<br />

Comb<strong>in</strong><strong>in</strong>g Eq. 4.2 <strong>and</strong> Eq. 4.10 for an isentropic process, a fluid acoustic<br />

velocity is then def<strong>in</strong>ed by, (Lighthill, 1978):<br />

a =<br />

� 1<br />

ρKS<br />

(4.11)<br />

For gases, semi-empirical equations of state provide explicit relationships between<br />

pressure <strong>and</strong> density. Therefore changes <strong>in</strong> density with pressure as<br />

expressed by the partial derivative of density with pressure can be evaluated<br />

theoretically. In addition, isothermal <strong>and</strong> isentropic processes for gases are<br />

related by the adiabatic <strong>in</strong>dex k. The acoustic velocity expression, Eq. 4.11<br />

can then be reformulated for gases by (Lighthill, 1978):<br />

a =<br />

�<br />

k<br />

ρKΓ<br />

(4.12)<br />

For liquids, empirical correlations provide relationships between isothermal<br />

<strong>and</strong> isentropic compressibilities, as well as tables of isothermal compressibilities.<br />

For liquids, isothermal <strong>and</strong> isentropic compressibilities are related by<br />

(Stallard et al., 1969):<br />

KΓ = KS + Γα2 p<br />

ρCp<br />

(4.13)<br />

where, Cp, represents specific heat capacity at constant pressure, <strong>and</strong>, αp,<br />

liquid thermal expansion coefficient at constant pressure. Liquid isothermal<br />

compressibility is commonly approximated by isentropic compressibility (Stallard<br />

et al., 1969). For example, at 25 o C, water thermal expansion coefficient<br />

<strong>and</strong> specific heat capacity are 69 10 −6 K −1 <strong>and</strong> 4.18 10 −3 J kg −1 K −1 , respectively<br />

(Eisenberg <strong>and</strong> Kauzman, 1969). The thermal expansion correction is<br />

45


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

then 4.35 10 −12 Pa −1 whereas water isothermal compressibility is 435 10 −12<br />

Pa −1 (McCa<strong>in</strong>, 1990).<br />

Thermal expansion correction term is negligible <strong>in</strong> Eq. 4.13 <strong>and</strong> uncerta<strong>in</strong>ties<br />

have low impact on isothermal compressibilities calculations (Stallard et al.,<br />

1969). Thus, acoustic velocity <strong>in</strong> liquids can be calculated us<strong>in</strong>g an expression<br />

similar to gases with k = 1. Acoustic velocity <strong>in</strong> liquids is then expressed by<br />

(Leighton, 1994):<br />

�<br />

1<br />

a =<br />

(4.14)<br />

ρKΓ<br />

Acoustic velocities <strong>in</strong> both gases <strong>and</strong> liquids are generally calculated us<strong>in</strong>g<br />

Eq. 4.12 with k = 1 for liquids. Acoustic velocity then depends on density,<br />

pressure <strong>and</strong> temperature. Isothermal compressibilities can be derived from<br />

semi-empirical equations of state for gases, while empirical correlations are<br />

necessary to evaluate liquid compressibilities.<br />

4.2 Acoustic Velocities <strong>in</strong> Gases<br />

4.2.1 Derivation<br />

Acoustic velocity <strong>in</strong> gases, Eq. 4.12, depends on density, pressure <strong>and</strong> temperature.<br />

Gases thermodynamical properties can be theoretically derived from<br />

semi-empirical equations of state that relate state variables. Each equation of<br />

state focuses on different characteristics <strong>and</strong> corresponds to different assumptions.<br />

Commonly used equations of state for gases are the ideal gas law <strong>and</strong><br />

the real gas law. Peng-Rob<strong>in</strong>son equation of state is also used even though<br />

other equations of state are reported better for gas.<br />

The ideal gas law correlates density, pressure <strong>and</strong> temperature for a perfect<br />

gas that assumes <strong>in</strong>f<strong>in</strong>itely small molecules without <strong>in</strong>termolecular forces. The<br />

ideal gas law predicts physical <strong>and</strong> thermodynamical properties of gases close<br />

to atmospheric pressure at moderate temperatures with 95 % accuracy. The<br />

ideal gas law relates pressure <strong>and</strong> temperature with the volume occupied by n<br />

molecules, us<strong>in</strong>g a universal gas constant, R. The ideal gas law is (Katz <strong>and</strong><br />

Lee, 1990):<br />

pV = nRΓ (4.15)<br />

The real gas law correlates density, pressure <strong>and</strong> temperature for a natural<br />

gas that considers molecule sizes <strong>and</strong> <strong>in</strong>termolecular forces us<strong>in</strong>g a correction<br />

factor, z. The z-factor is an empirical function of pressure, temperature <strong>and</strong><br />

gas composition. The real gas law predicts physical <strong>and</strong> thermodynamical<br />

46


4.2. Acoustic Velocities <strong>in</strong> Gases<br />

properties of gases at pressures greater than atmospheric pressure <strong>and</strong> a wider<br />

range of temperature than the ideal gas law. The real gas law is (Katz <strong>and</strong><br />

Lee, 1990):<br />

pV = znRΓ (4.16)<br />

The ideal gas law <strong>and</strong> real gas law can be reformulated by substitut<strong>in</strong>g volume<br />

by density. The ideal gas law can be obta<strong>in</strong>ed from the real gas law for z = 1,<br />

so properties will be derived from the real gas law only. The number of moles,<br />

n, is the ratio of a gas mass, m, <strong>and</strong> its molecular mass, M; <strong>and</strong> density is the<br />

ratio of mass <strong>and</strong> volume. After substitution, the real gas law becomes:<br />

p = zρ R<br />

Γ (4.17)<br />

M<br />

The isothermal compressibility required partial derivative of density with pressure<br />

at constant temperature can be obta<strong>in</strong>ed directly from Eq. 4.17. Both<br />

density <strong>and</strong> z-factor are functions of pressure <strong>and</strong> temperature such that the<br />

isothermal compressibility of a gas follow<strong>in</strong>g the real gas law becomes:<br />

KΓ = 1<br />

� �<br />

1 ∂z<br />

− (4.18)<br />

p z ∂p<br />

Γ<br />

The first term <strong>in</strong> Eq. 4.18 is the ideal gas isothermal compressibility s<strong>in</strong>ce<br />

the z-factor is constant <strong>and</strong> equals unity. The second term <strong>in</strong> Eq. 4.18 is<br />

then a correction factor that accounts for the changes <strong>in</strong> molecule sizes <strong>and</strong><br />

molecular forces with pressure at constant temperature. Acoustic velocities <strong>in</strong><br />

natural gases can be approximated by neglect<strong>in</strong>g the second term <strong>in</strong> Eq. 4.18.<br />

However, the true acoustic velocity can be obta<strong>in</strong>ed from comb<strong>in</strong><strong>in</strong>g Eq. 4.12<br />

<strong>and</strong> Eq. 4.18:<br />

�<br />

�<br />

� k<br />

a = � � � � � (4.19)<br />

� 1 1 ∂z<br />

ρ −<br />

p z ∂p<br />

Γ<br />

The above acoustic velocity formula, Eq. 4.19, is to be compared with a<br />

st<strong>and</strong>ard calculation provided by the American Gas Association (AGA-10,<br />

2003). AGA is a trade organization that promotes the use of natural gas <strong>and</strong><br />

that provides st<strong>and</strong>ards for calculat<strong>in</strong>g natural gas properties such as z-factor<br />

(AGA-8, 1992), or acoustic velocity (AGA-10, 2003). In the latter report,<br />

acoustic velocity is derived from thermodynamical relations <strong>and</strong> is expressed<br />

by (Savidge et al., 1988):<br />

��cp �� �� � � �<br />

RΓ ∂z<br />

a =<br />

z + ρ<br />

(4.20)<br />

cv M ∂ρ Γ<br />

47


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Acoustic velocity expressions for natural gas, Eq. 4.19 <strong>and</strong> Eq. 4.20, are<br />

different. The first expressions is directly deduced from the real gas equation<br />

of state, whereas the second expression <strong>in</strong>cludes a high accuracy equation<br />

of state for the z-factor. The present study relies on values computed with<br />

the Peng-Rob<strong>in</strong>son fluid package <strong>in</strong>cluded <strong>in</strong> HYSYS, a process simulation<br />

software. Both formulations are here<strong>in</strong> numerically compared.<br />

The Peng-Rob<strong>in</strong>son equation of state is a cubic, two-constant equation of<br />

state, of the van der Walls type (Peng <strong>and</strong> Rob<strong>in</strong>son, 1976). The Peng-<br />

Rob<strong>in</strong>son equation of state is commonly used <strong>in</strong> petroleum eng<strong>in</strong>eer<strong>in</strong>g for<br />

its accurate prediction of vapor pressure, liquid densities <strong>and</strong> phase equilibria<br />

for multi-component fluids. The Peng-Rob<strong>in</strong>son equation of state is particularly<br />

accurate near the critical po<strong>in</strong>t (Katz <strong>and</strong> Lee, 1990).<br />

4.2.2 Example<br />

A typical natural gas molar composition is given <strong>in</strong> Table 4.1 (Pedersen et al.,<br />

1989). The correspond<strong>in</strong>g phase envelope is shown <strong>in</strong> Figure 4.1. Two phases<br />

existence envelope range from -155 o C to -20 o C, <strong>in</strong> temperature, <strong>and</strong> from 1<br />

bara to 80 bara, <strong>in</strong> pressure. Gas density, adiabatic <strong>in</strong>dex <strong>and</strong> z-factor, are<br />

obta<strong>in</strong>ed us<strong>in</strong>g the Peng-Rob<strong>in</strong>son fluid package <strong>in</strong>cluded <strong>in</strong> HYSYS. First<br />

order partial derivatives of z-factor with pressure or density <strong>in</strong> Eq. 4.18, Eq.<br />

4.19 <strong>and</strong> Eq. 4.20, are approximated by f<strong>in</strong>ite differences.<br />

Evolution of natural gas density with pressure <strong>and</strong> temperature is shown <strong>in</strong><br />

Figure 4.2. The gas density <strong>in</strong>creases with pressure <strong>and</strong> decreases with temperature.<br />

At 25 o C, the gas density <strong>in</strong>creases from 0.75 kg m −3 at 1 bara, to<br />

42.72 kg m −3 at 50 bara, <strong>and</strong> to 95.92 kg m −3 at 100 bara. At 100 bara, the<br />

density decreases from 95.92 kg m −3 at 25 o C, to 82.11 kg m −3 at 50 o C, <strong>and</strong><br />

to 65.30 kg m −3 at 100 o C.<br />

Evolution of natural gas adiabatic <strong>in</strong>dex with pressure <strong>and</strong> temperature is<br />

shown <strong>in</strong> Figure 4.3. The gas adiabatic <strong>in</strong>dex <strong>in</strong>creases with pressure <strong>and</strong><br />

decreases with temperature. At 25 o C, the gas adiabatic <strong>in</strong>dex <strong>in</strong>creases from<br />

1.28 at 1 bara, to 1.49 at 50 bara, <strong>and</strong> to 1.77 at 100 bara. At 100 bara, the<br />

gas adiabatic <strong>in</strong>dex decreases from 1.77 at 25 o C, to 1.59 at 50 o C, <strong>and</strong> to 1.42<br />

at 100 o C.<br />

Evolution of natural gas z-factor with pressure <strong>and</strong> temperature is shown<br />

<strong>in</strong> Figure 4.4. The gas z-factor decreases with pressure <strong>and</strong> <strong>in</strong>creases with<br />

temperature. At 25 o C, the z-factor decreases from 1 at 1 bara, to 0.87 at 50<br />

bara, <strong>and</strong> to 0.77 at 100 bara. At 100 bara, the z-factor <strong>in</strong>creases from 0.77 at<br />

25 o C, to 0.83 at 50 o C, <strong>and</strong> to 0.91 at 100 o C. The isothermal compressibility,<br />

KΓ, can be calculated from the pressure <strong>and</strong> the gradient <strong>in</strong> z-factor.<br />

48


4.2. Acoustic Velocities <strong>in</strong> Gases<br />

Table 4.1: Typical Natural Gas Molar Composition (Pedersen et al.,<br />

1989)<br />

<strong>Pressure</strong> [bara]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Component Mole Fraction<br />

N2 0.30<br />

CO2 1.10<br />

C1 90.00<br />

C2 4.90<br />

C3 1.90<br />

C4 (i + n) 1.10<br />

C5 (i + n) 0.40<br />

C6 (i + n) 0.30<br />

Critical Po<strong>in</strong>t<br />

Bubble Po<strong>in</strong>t<br />

Dew Po<strong>in</strong>t<br />

0<br />

−160 −140 −120 −100 −80 −60 −40 −20 0<br />

Temperature [ o C]<br />

Figure 4.1: Typical Natural Gas Phase Envelope<br />

Composition given <strong>in</strong> Table 4.1<br />

Gas Properties Calculated Us<strong>in</strong>g Peng-Ronb<strong>in</strong>son Equation of State<br />

49


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Density [kg m −3 ]<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

Γ = 100 o C<br />

0<br />

0 50 100 150<br />

<strong>Pressure</strong> [bara]<br />

Figure 4.2: Evolution of Typical Natural Gas Density with <strong>Pressure</strong><br />

<strong>and</strong> Temperature<br />

Adiabatic Index<br />

2.0<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

Γ = 100 o C<br />

1.0<br />

0 50 100 150<br />

<strong>Pressure</strong> [bara]<br />

Figure 4.3: Evolution of Typical Natural Gas Adiabatic Index with<br />

<strong>Pressure</strong> <strong>and</strong> Temperature<br />

Composition given <strong>in</strong> Table 4.1<br />

Gas Properties Calculated Us<strong>in</strong>g Peng-Ronb<strong>in</strong>son Equation of State<br />

50


4.2. Acoustic Velocities <strong>in</strong> Gases<br />

Evolution of natural gas isothermal compressibility with pressure <strong>and</strong> temperature<br />

is shown <strong>in</strong> Figure 4.5. The gas isothermal compressibility decreases<br />

with pressure <strong>and</strong> temperature. At 25 o C, the gas compressibility decreases<br />

from 10 10 −6 Pa −1 , to 0.227 10 −6 Pa −1 at 50 bara, <strong>and</strong> to 0.117 10 −6 Pa −1<br />

at 100 bara. At 100 bara, the isothermal compressibility goes from 0.117 10 −6<br />

Pa −1 at 25 o C, to 0.111 10 −6 Pa −1 at 50 o C, <strong>and</strong> to 0.105 10 −6 Pa −1 at 100 o C.<br />

The greater the pressure, the stiffer the gas becomes.<br />

Evolution of natural gas acoustic velocity with pressure <strong>and</strong> temperature, computed<br />

from Eq. 4.19 <strong>and</strong> Eq. 4.20 are shown <strong>in</strong> Figure 4.6. In both cases,<br />

the acoustic velocity <strong>in</strong>creases with temperature, <strong>and</strong> decreases with pressure<br />

down to a m<strong>in</strong>imum, before <strong>in</strong>creas<strong>in</strong>g aga<strong>in</strong>. At 25 o C, the acoustic velocity<br />

goes from 414 ms −1 at 1 bara, to 392 ms −1 at 50 bara, <strong>and</strong> to 397 ms −1 at<br />

100 bara. At 100 bara, the acoustic velocity <strong>in</strong>creases from 397 ms −1 at 25 o C<br />

to 418 ms −1 at 50 o C, <strong>and</strong> to 455 ms −1 at 100 o C.<br />

Acoustic velocities calculated with Eq. 4.19 <strong>and</strong> reference Eq. 4.20 are almost<br />

identical. However, results deviate slightly from each other, as observed <strong>in</strong><br />

Figure 4.7. Deviations decrease with temperature, <strong>and</strong> <strong>in</strong>crease with pressure<br />

up to a maximum before decreas<strong>in</strong>g aga<strong>in</strong>. A maximum deviation of 0.1 % is<br />

obta<strong>in</strong>ed at 25 o C, 70 bara which is reasonable for most of the cases, except<br />

for ultrasonic measurements.<br />

51


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

z − Factor<br />

1.00<br />

0.95<br />

0.90<br />

0.85<br />

0.80<br />

0.75<br />

Γ = 100 o C<br />

Γ = 50 o C<br />

Γ = 25<br />

0.70<br />

0 50 100 150<br />

<strong>Pressure</strong> [bara]<br />

o C<br />

Figure 4.4: Evolution of Typical Natural Gas Z-Factor with <strong>Pressure</strong><br />

<strong>and</strong> Temperature<br />

Compressibility [10 −6 Pa −1 ]<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

Γ = 100 o C<br />

0<br />

0 50 100 150<br />

<strong>Pressure</strong> [bara]<br />

Figure 4.5: Evolution of Typical Natural Gas Isothermal<br />

Compressibility with <strong>Pressure</strong> <strong>and</strong> Temperature<br />

Composition given <strong>in</strong> Table 4.1<br />

Gas Properties Calculated Us<strong>in</strong>g Peng-Ronb<strong>in</strong>son Equation of State<br />

52


Acoustic Velocity [m s −1 ]<br />

475<br />

450<br />

425<br />

400<br />

4.2. Acoustic Velocities <strong>in</strong> Gases<br />

Γ = 100 o C<br />

Γ = 50 o C<br />

Γ = 25<br />

375<br />

0 50 100 150<br />

<strong>Pressure</strong> [bara]<br />

o C<br />

Def<strong>in</strong>ition<br />

AGA 10<br />

Figure 4.6: Evolution of Typical Natural Gas Acoustic Velocity with<br />

<strong>Pressure</strong> <strong>and</strong> Temperature<br />

Error [%]<br />

0.10<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

Γ = 100 o C<br />

0.00<br />

0 50 100 150<br />

<strong>Pressure</strong> [bara]<br />

Figure 4.7: Deviation Between Eq. 4.19 <strong>and</strong> Eq. 4.20<br />

Composition given <strong>in</strong> Table 4.1<br />

Gas Properties Calculated Us<strong>in</strong>g Peng-Ronb<strong>in</strong>son Equation of State<br />

53


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

4.3 Acoustic Velocity <strong>in</strong> Liquid Oils<br />

4.3.1 Acoustic Velocity <strong>in</strong> Volatile Oil<br />

A typical volatile oil composition molar composition is given <strong>in</strong> Table 4.2 <strong>and</strong><br />

the correspond<strong>in</strong>g phase envelope is shown <strong>in</strong> Figure 4.8. Two phases existence<br />

envelope range from -150 o C to 250 o C, <strong>in</strong> temperature, <strong>and</strong> from 0 bara to 250<br />

bara, <strong>in</strong> pressure. Volatile oil density is obta<strong>in</strong>ed us<strong>in</strong>g the Peng-Rob<strong>in</strong>son<br />

fluid package <strong>in</strong>cluded <strong>in</strong> HYSYS. Isothermal compressibility is approximated<br />

by f<strong>in</strong>ite difference (Eq. 4.10) <strong>and</strong> neglect<strong>in</strong>g thermal expansion (Eq. 4.13).<br />

Evolution of volatile oil density with pressure <strong>and</strong> temperature is shown <strong>in</strong><br />

Figure 4.9. The density <strong>in</strong>creases with pressure <strong>and</strong> decreases with temperature.<br />

At 25 o C, the density <strong>in</strong>creases from 721 kg m −3 , to 727 kg m −3 , <strong>and</strong><br />

to 732 kg m −3 at 1 bara, 50 bara <strong>and</strong> 100 bara, respectively. At 100 bara,<br />

the density decreases from 732 kg m −3 , to 720 kg m −3 , <strong>and</strong> to 695 kg m −3 at<br />

25 o C, 50 o C <strong>and</strong> 100 o C, respectively.<br />

Evolution of volatile oil isothermal compressibility with pressure <strong>and</strong> temperature<br />

is shown <strong>in</strong> Figure 4.10. The isothermal compressibility decreases<br />

with pressure <strong>and</strong> <strong>in</strong>creases with temperature. At 25 o C, the isothermal compressibility<br />

decreases from 1.67 10 −9 Pa −1 , to 1.56 10 −9 Pa −1 , <strong>and</strong> to 1.46<br />

10 −9 Pa −1 at 1 bara, 50 bara <strong>and</strong> 100 bara, respectively. At 100 bara, the<br />

isothermal compressibility <strong>in</strong>creases from 1.46 10 −9 Pa −1 , to 1.60 10 −9 Pa −1 ,<br />

<strong>and</strong> to 1.98 10 −9 Pa −1 at 25 o C, 50 o C <strong>and</strong> 100 o C, respectively.<br />

Evolution of acoustic velocity with pressure <strong>and</strong> temperature, <strong>in</strong> volatile oil,<br />

calculated with Eq. 4.14, is shown <strong>in</strong> Figure 4.11. The acoustic velocity<br />

<strong>in</strong>creases with pressure <strong>and</strong> decreases with temperature. At 25 o C, the acoustic<br />

velocity <strong>in</strong>creases from 911 ms −1 , to 940 ms −1 , <strong>and</strong> to 968 ms −1 at 1 bara, 50<br />

bara <strong>and</strong> 100 bara, respectively. At 100 bara, the acoustic velocity decreases<br />

from 968 ms −1 , to 930 ms −1 , <strong>and</strong> to 853 ms −1 at 25 o C, 50 o C <strong>and</strong> 100 o C,<br />

respectively.<br />

54


4.3. Acoustic Velocity <strong>in</strong> Liquid Oils<br />

Table 4.2: Typical Volatile Oil Molar Composition, Pedersen et al.<br />

(1989)<br />

Component Mole Fraction Component Mole Fraction<br />

N2 1.67 CO2 2.18<br />

C1 60.51 C7 2.45<br />

C2 7.52 C8 2.41<br />

C3 4.74 C9 1.69<br />

C4 (i + n) 4.12 C10 1.42<br />

C5 (i + n) 2.97 C11 1.02<br />

C6 (i + n) 1.99 C12 (12+) 5.31<br />

<strong>Pressure</strong> [bara]<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

Critical po<strong>in</strong>t<br />

Bubble Po<strong>in</strong>t<br />

Dew Po<strong>in</strong>t<br />

0<br />

−200 −100 0 100<br />

Temperature [<br />

200 300 400<br />

o C]<br />

Figure 4.8: Typical Volatile Oil Phase Envelope<br />

Composition given <strong>in</strong> Table 4.2 from Pedersen et al. (1989)<br />

55


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Density [kg m −3 ]<br />

760<br />

740<br />

720<br />

700<br />

680<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

660<br />

0 20 40 60 80<br />

Γ = 100<br />

100<br />

<strong>Pressure</strong> [bara]<br />

o C<br />

Figure 4.9: Evolution of Typical Volatile Oil Density with <strong>Pressure</strong><br />

<strong>and</strong> Temperature<br />

Compressibility [10 −9 Pa −1 ]<br />

2.40<br />

2.20<br />

2.00<br />

1.80<br />

1.60<br />

1.40<br />

1.20<br />

Γ = 100 o C<br />

Γ = 50 o C<br />

1.00<br />

0<br />

Γ = 25<br />

20 40 60 80 100<br />

<strong>Pressure</strong> [bara]<br />

o C<br />

Figure 4.10: Evolution of Typical Volatile Oil Isothermal<br />

Compressibility with <strong>Pressure</strong> <strong>and</strong> Temperature<br />

Composition given <strong>in</strong> Table 4.2 from Pedersen et al. (1989)<br />

Volatile Oil Properties Calculated Us<strong>in</strong>g Peng-Rob<strong>in</strong>son Equation of State<br />

56


Acoustic Velocity [m s −1 ]<br />

1150<br />

1100<br />

1050<br />

1000<br />

950<br />

900<br />

850<br />

800<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

Γ = 100 o C<br />

4.3. Acoustic Velocity <strong>in</strong> Liquid Oils<br />

750<br />

0 20 40 60 80 100<br />

<strong>Pressure</strong> [bara]<br />

Figure 4.11: Evolution of Acoustic Velocity <strong>in</strong> Typical Volatile Oil<br />

with <strong>Pressure</strong> <strong>and</strong> Temperature<br />

Composition given <strong>in</strong> Table 4.2 from Pedersen et al. (1989)<br />

Volatile Oil Properties Calculated Us<strong>in</strong>g Peng-Rob<strong>in</strong>son Equation of State<br />

57


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

4.3.2 Acoustic Velocity <strong>in</strong> Black Oil<br />

A typical black oil composition molar composition is given <strong>in</strong> Table 4.3 <strong>and</strong> the<br />

correspond<strong>in</strong>g phase envelope is shown <strong>in</strong> Figure 4.12. Two phases existence<br />

envelope range from -150 o C to 400 o C, <strong>in</strong> temperature, <strong>and</strong> from 0 bara to 170<br />

bara, <strong>in</strong> pressure. Volatile oil density is obta<strong>in</strong>ed us<strong>in</strong>g the Peng-Rob<strong>in</strong>son<br />

fluid package <strong>in</strong>cluded <strong>in</strong> HYSYS. First order partial derivatives <strong>in</strong> Eq. 4.10<br />

is approximated by f<strong>in</strong>ite difference <strong>and</strong> thermal expansion <strong>in</strong> Eq. 4.13 is<br />

neglected.<br />

Evolution of black oil density with pressure <strong>and</strong> temperature is shown <strong>in</strong> Figure<br />

4.13. The density <strong>in</strong>creases with pressure <strong>and</strong> decreases with temperature. At<br />

25 o C, the density <strong>in</strong>creases from 747 kg m −3 at 1 bara, to 751 kg m −3 at 50<br />

bara, <strong>and</strong> to 756 kg m −3 at 100 bara. At 100 bara, the density decreases from<br />

756 kg m −3 at 25 o C, to 742 kg m −3 at 50 bara, <strong>and</strong> to 717 kg m −3 at 100 o C.<br />

Evolution of black oil isothermal compressibility with pressure <strong>and</strong> temperature<br />

is shown <strong>in</strong> Figure 4.14. The isothermal compressibility decreases with<br />

pressure <strong>and</strong> <strong>in</strong>creases with temperature. At 25 o C, the isothermal compressibility<br />

decreases from 1.35 10 −9 Pa −1 at 1 bara, to 1.27 10 −9 Pa −1 at 50 bara,<br />

<strong>and</strong> to 1.19 10 −9 Pa −1 at 100 bara. At 100 bara, the isothermal compressibility<br />

<strong>in</strong>creases from 1.19 10 −9 Pa −1 at 25 o C, to 1.28 10 −9 Pa −1 at 50 o C, <strong>and</strong> to<br />

1.46 10 −9 Pa −1 at 100 o C.<br />

Evolution of acoustic velocity with pressure <strong>and</strong> temperature, <strong>in</strong> black oil,<br />

calculated with Eq. 4.14, is shown <strong>in</strong> Figure 4.15. The acoustic velocity<br />

<strong>in</strong>creases with pressure <strong>and</strong> decreases with temperature. At 25 o C, the acoustic<br />

velocity <strong>in</strong>creases from 994 ms −1 at 1 bara, to 1024 ms −1 at 50 bara, <strong>and</strong><br />

to 1053 ms −1 at 100 bara. At 100 bara, the acoustic velocity decreases from<br />

1053 ms −1 at 25 o C, to 1027 ms −1 at 50 o C, <strong>and</strong> to 977 ms −1 at 100 o C.<br />

58


4.3. Acoustic Velocity <strong>in</strong> Liquid Oils<br />

Table 4.3: Typical Black Oil Molar Composition (Pedersen et al.,<br />

1989)<br />

Component Mole Fraction Component Mole Fraction<br />

N2 0.67 CO2 2.11<br />

C1 34.93 C11 1.72<br />

C2 7.00 C12 1.74<br />

C3 7.82 C13 1.74<br />

C4 (i + n) 5.48 C14 1.35<br />

C5 (i + n) 3.80 C15 1.34<br />

C6 (i + n) 3.04 C16 1.06<br />

C7 4.39 C17 1.02<br />

C8 4.71 C18 1.00<br />

C9 3.21 C19 0.90<br />

C10 1.79 C20 (20+) 9.18<br />

<strong>Pressure</strong> [bara]<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

Critical Po<strong>in</strong>t<br />

Bubble Po<strong>in</strong>t<br />

Dew Po<strong>in</strong>t<br />

0<br />

−200 −100 0 100<br />

Temperature [<br />

200 300 400<br />

o C]<br />

Figure 4.12: Typical Black Oil Phase Envelope<br />

Composition given <strong>in</strong> Table 4.3 from Pedersen et al. (1989)<br />

59


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Density [kg m −3 ]<br />

760<br />

740<br />

720<br />

700<br />

680<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

660<br />

0 20 40 60 80<br />

Γ = 100<br />

100<br />

<strong>Pressure</strong> [bara]<br />

o C<br />

Figure 4.13: Evolution of Typical Black Oil Density with <strong>Pressure</strong><br />

<strong>and</strong> Temperature<br />

Compressibility [10 −9 Pa −1 ]<br />

2.40<br />

2.20<br />

2.00<br />

1.80<br />

1.60<br />

1.40<br />

1.20<br />

Γ = 100 o C<br />

Γ = 50 o C<br />

Γ = 25 o C<br />

1.00<br />

0 20 40 60 80 100<br />

<strong>Pressure</strong> [bara]<br />

Figure 4.14: Evolution of Typical Black Oil Isothermal<br />

Compressibility with <strong>Pressure</strong> <strong>and</strong> Temperature<br />

Composition given <strong>in</strong> Table 4.3 from Pedersen et al. (1989)<br />

Oil Properties Calculated Us<strong>in</strong>g Peng-Ronb<strong>in</strong>son Equation of State<br />

60


Acoustic Velocity [m s −1 ]<br />

1150<br />

1100<br />

1050<br />

1000<br />

950<br />

900<br />

850<br />

800<br />

4.3. Acoustic Velocity <strong>in</strong> Liquid Oils<br />

Γ = 25 o C<br />

Γ = 50 o C<br />

750<br />

0 20 40 60 80<br />

Γ = 100<br />

100<br />

<strong>Pressure</strong> [bara]<br />

o C<br />

Figure 4.15: Evolution of Acoustic Velocity <strong>in</strong> Typical Black Oil with<br />

<strong>Pressure</strong> <strong>and</strong> Temperature<br />

Composition given <strong>in</strong> Table 4.3 from Pedersen et al. (1989)<br />

Oil Properties Calculated Us<strong>in</strong>g Peng-Ronb<strong>in</strong>son Equation of State<br />

61


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

4.3.3 Comparison Between Volatile Oils <strong>and</strong> Black Oils<br />

Volatile oils conta<strong>in</strong> more light molecules (C1 to C6) than black oils as observed<br />

from the molar compositions, Tables 4.2 <strong>and</strong> 4.3. Therefore, black oils are<br />

denser <strong>and</strong> less compressible than volatile oils. At 50 o C, 50 bara, black oil<br />

<strong>and</strong> volatile oil densities are 738 kg m −3 <strong>and</strong> 714 kg m −3 , respectively (Figure<br />

4.9 <strong>and</strong> Figure 4.13). At 50 o C, 50 bara, black oil <strong>and</strong> volatile oil isothermal<br />

compressibilities are 1.36 10 −9 Pa −1 <strong>and</strong> 1.73 10 −9 Pa −1 , respectively (Figure<br />

4.10 <strong>and</strong> Figure 4.14).<br />

Black oils greater densities <strong>and</strong> compressibilities <strong>in</strong>duce greater acoustic velocities<br />

than <strong>in</strong> volatile oils. At 50 o C, 50 bara, acoustic velocities <strong>in</strong> black <strong>and</strong><br />

volatile oils are 998 ms −1 <strong>and</strong> 900 ms −1 , respectively (Figures 4.11 <strong>and</strong> 4.15).<br />

Currently, no theory can predict accurate acoustic velocities <strong>in</strong> oils, ma<strong>in</strong>ly due<br />

to the diversity of hydrocarbons. However, empirical correlations can predict<br />

acoustic velocities function of molecular weight with<strong>in</strong> 3 % accuracy (Wang<br />

et al., 1990).<br />

Critical temperatures are lower for volatile oils than for black oils. The studied<br />

volatile oil <strong>and</strong> black oil critical temperatures are 170 o C <strong>and</strong> 356 o C, respectively<br />

(Figure 4.8 <strong>and</strong> Figure 4.12). Nevertheless, critical pressures are greater<br />

for volatile oils than for black oils. The studied volatile oil <strong>and</strong> black oil critical<br />

pressures are 170 bara <strong>and</strong> 107 bara, respectively (Figure 4.8 <strong>and</strong> Figure<br />

4.12). Volatile oil critical pressures are generally close to reservoir conditions.<br />

4.4 Acoustic Velocity <strong>in</strong> Liquid Water<br />

4.4.1 Acoustic Velocity<br />

Unlike liquid oils, water is a s<strong>in</strong>gle molecule fluid. Thus, thorough experimental<br />

work results are available through steam tables for <strong>in</strong>stance. Density<br />

<strong>and</strong> acoustic velocity are two parameters extensively <strong>in</strong>vestigated. Accurate<br />

empirical correlations for the acoustic velocity <strong>in</strong> natural water relate wave<br />

propagation speeds to pressure, p, temperature, Γ, <strong>and</strong> sal<strong>in</strong>ity, Σ. One accurate<br />

correlation has been proposed by DelGrosso (1974):<br />

a(Σ, Γ, p) = a0,0,0 + ∆aΣ + ∆aΓ + ∆ap + ∆aΣΓp<br />

(4.21)<br />

where, a0,0,0, is the acoustic velocity reference value at zero sal<strong>in</strong>ity, Σ = 0 ppm,<br />

zero temperature, Γ = 0 o C <strong>and</strong> atmospheric pressure. Functions ∆aΣ, ∆aΓ,<br />

∆ap <strong>and</strong> ∆aΣΓp, are correction operators for sal<strong>in</strong>ity, temperature, pressure<br />

<strong>and</strong> <strong>in</strong>teraction between these three parameters, respectively. Correction functions<br />

are (DelGrosso, 1974):<br />

62


∆aΣ = 0.132952290781 10 1 Σ<br />

+ 0.128955756844 10 −3 Σ 2<br />

∆aΓ = 0.501109398873 10 1 (Γ − 273.15)<br />

− 0.550946843172 10 −1 (Γ − 273.15) 2<br />

+ 0.221535969240 10 −3 (Γ − 273.15) 3<br />

∆ap = 0.156059257041 10 0 Ψp<br />

+ 0.244998688441 10 −4 (Ψp) 2<br />

− 0.883392332513 10 −8 (Ψp) 3<br />

∆aΣΓp = −0.127562783426 10 −1 (Γ − 273.15)Σ<br />

+ 0.635191613389 10 −2 (Γ − 273.15)Ψp<br />

+ 0.265484716608 10 −7 ((Γ − 273.15)Ψp) 2<br />

− 0.159349479045 10 −5 (Γ − 273.15)(Ψp) 2<br />

+ 0.522116437235 10 −9 (Γ − 273.15)(Ψp) 3<br />

− 0.438031096213 10 −6 (Γ − 273.15) 3 Ψp<br />

− 0.161674495909 10 −8 (ΣΨp) 2<br />

+ 0.968403156410 10 −4 (Γ − 273.15) 2 Σ<br />

+ 0.485639620015 10 −5 (Γ − 273.15)ΣΨp<br />

4.4. Acoustic Velocity <strong>in</strong> Liquid Water<br />

(4.22)<br />

(4.23)<br />

(4.24)<br />

− 0.340597039004 10 −3 (Γ − 273.15)ΣΨp (4.25)<br />

where, Ψ, is constant equal to 1.019716213 used for convert<strong>in</strong>g pressures from<br />

bara to kg cm −2 . Evolution of acoustic velocity with pressure <strong>and</strong> temperature,<br />

<strong>in</strong> water at zero sal<strong>in</strong>ity, is shown <strong>in</strong> Figure 4.16. The acoustic velocity<br />

<strong>in</strong>creases with pressure <strong>and</strong> temperature. At 1 bara, the acoustic velocity <strong>in</strong>creases<br />

from 1426 ms −1 at 5 o C to 1497 ms −1 at 25 o C, <strong>and</strong> to1543 ms −1 at<br />

50 o C. At 25 o C, the acoustic velocity <strong>in</strong>creases from 1497 ms −1 to 1509 ms −1<br />

at 40 bara, <strong>and</strong> to 1534 ms −1 at 120 bara.<br />

63


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Acoustic Velocity [m s −1 ]<br />

1600<br />

1575<br />

1550<br />

1525<br />

1500<br />

1475<br />

1450<br />

1425<br />

1400<br />

0 10 20 30 40 50<br />

Temperature [ o p = 120 bara<br />

p = 80 bara<br />

p = 40 bara<br />

p = 20 bara<br />

p = 10 bara<br />

p = 1 bara<br />

C]<br />

Figure 4.16: Evolution of Acoustic Velocity <strong>in</strong> Water with <strong>Pressure</strong><br />

<strong>and</strong> Temperature at Zero Sal<strong>in</strong>ity (DelGrosso, 1974)<br />

4.4.2 Density of Water<br />

Water density is thoroughly tabulated function of pressure, temperature <strong>and</strong><br />

sal<strong>in</strong>ity. Accurate knowledge of water density is of particular <strong>in</strong>terest for predict<strong>in</strong>g<br />

geothermal system performances, for <strong>in</strong>stance. Based on steam tables,<br />

empirical correlations def<strong>in</strong>e density as a function of temperature <strong>and</strong> molality<br />

of salt <strong>in</strong> pure water (Michaelides, 1981). Density prediction <strong>in</strong>volves calculations<br />

of enthalpy, temperature elevation <strong>and</strong> specific volume. Calculations<br />

<strong>and</strong> results are detailed <strong>in</strong> Michaelides (1981).<br />

Evolution of water density with pressure <strong>and</strong> temperature, us<strong>in</strong>g Michaelides<br />

correlation <strong>and</strong> us<strong>in</strong>g the ASME Steam Fluid package <strong>in</strong>cluded <strong>in</strong> HYSYS,<br />

is shown <strong>in</strong> Figure 4.17. Density decreases with temperature <strong>and</strong> <strong>in</strong>creases<br />

with pressure. At 1 bara, the density calculated us<strong>in</strong>g Michaelides correlation<br />

decreases from 999.7 ms −1 at 5 o C, to 996.9 ms −1 at 25 o C, <strong>and</strong> to 989.8<br />

ms −1 at 50 o C. At 25 o C, the density calculated us<strong>in</strong>g Michaelides correlation<br />

<strong>in</strong>creases from 997.3 ms −1 at 1 bara, to 998.6 ms −1 at 40 bara, <strong>and</strong> to 1003<br />

ms −1 at 120 bara.<br />

Densities predicted by Michaelides correlation <strong>and</strong> by the ASME Steam fluid<br />

package <strong>in</strong>cluded <strong>in</strong> HYSYS are similar at low pressures <strong>and</strong> low temperatures,<br />

Figure 4.17. However, the maximum discrepancy is less than 0.10 % <strong>and</strong> is<br />

64


Density [kg m −3 ]<br />

1006<br />

1004<br />

1002<br />

1000<br />

998<br />

996<br />

4.4. Acoustic Velocity <strong>in</strong> Liquid Water<br />

994<br />

992<br />

990<br />

988<br />

986<br />

0 10 20 30 40 50<br />

Temperature [ o p = 120 bara<br />

p = 80 bara<br />

p = 40 bara<br />

p = 20 bara<br />

p = 10 bara<br />

p = 1 bara<br />

HYSYS<br />

Michaelides<br />

C]<br />

Figure 4.17: Evolution of Water Density with <strong>Pressure</strong> <strong>and</strong><br />

Temperature at Zero Sal<strong>in</strong>ity (Michaelides, 1981)<br />

therefore acceptable s<strong>in</strong>ce empirical correlations rely on experimental data<br />

with <strong>in</strong>accuracy greater than 0.1 %.<br />

4.4.3 Isothermal Compressibility<br />

Evolution of water isothermal compressibility with pressure <strong>and</strong> temperature<br />

can be deduced from acoustic velocity <strong>and</strong> water density evolution with pressure<br />

<strong>and</strong> temperature us<strong>in</strong>g Eq. 4.14.<br />

Evolution of water isothermal compressibility with pressure <strong>and</strong> temperature<br />

is shown <strong>in</strong> Figure 4.18. Water isothermal compressibility decreases with<br />

temperature <strong>and</strong> pressure. At 1 bara, the isothermal compressibility decreases<br />

from 0.492 10 −9 Pa −1 at 5 o C, to 0.448 10 −9 Pa −1 at 25 o C, <strong>and</strong> to<br />

0.425 10 −9 Pa −1 at 50 o C. At 25 o C, the isothermal compressibility decreases<br />

from 0.425 10 −9 Pa −1 at 1 bara, to 0.440 10 −9 Pa −1 at 40 bara, <strong>and</strong> to 0.424<br />

10 −9 Pa −1 at 120 bara. A value of 0.435 10 −9 Pa −1 (3 10 −6 psi −1 ) is a commonly<br />

admitted value for water compressibility (Dake, 1978).<br />

The water isothermal compressibility calcaulted above at atmospheric pressure<br />

can be compared to empirical correlations (Eisenberg <strong>and</strong> Kauzman, 1969).<br />

The comparison of the extrapolated isothermal compressibility from acoustic<br />

velocity <strong>and</strong> density <strong>and</strong> empirically correlated isothermal compressibility<br />

65


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Compressibility [10 −9 Pa −1 ]<br />

0.52<br />

0.50<br />

0.48<br />

0.46<br />

0.44<br />

0.42<br />

0.40<br />

p = 1 bara<br />

p = 10 bara<br />

p = 20 bara<br />

p = 40 bara<br />

p = 80 bara<br />

p = 120 bara<br />

0.38<br />

0 10 20 30 40 50<br />

Temperature [ o C]<br />

Figure 4.18: Evolution of Water Isothermal Compressibility with<br />

<strong>Pressure</strong> <strong>and</strong> Temperature at Zero Sal<strong>in</strong>ity<br />

is shown <strong>in</strong> Figure 4.19. Deduced <strong>and</strong> empirical isothermal compressibilities<br />

co<strong>in</strong>cide at low temperature but diverge at high temperature. At 50 o C, calculated<br />

isothermal compressibility is 0.425 10 −9 Pa −1 <strong>and</strong> empirically correlated<br />

is 0.442 10 −9 Pa −1 .<br />

4.5 Acoustic Velocity <strong>in</strong> Fluid-Filled Pipe<br />

<strong>Pressure</strong> wave propagation velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e depends on fluid<br />

elastic properties, as well as pipel<strong>in</strong>e geometry <strong>and</strong> spatial description. <strong>Pressure</strong><br />

waves used <strong>in</strong> seismic exploration are generated at sea-surface, propagate<br />

through sea <strong>and</strong> Earth. In such case, the medium is described as <strong>in</strong>f<strong>in</strong>ite <strong>and</strong><br />

acoustic velocity calculations follow aforementioned formulas. However, pressure<br />

waves <strong>in</strong> petroleum eng<strong>in</strong>eer<strong>in</strong>g propagate <strong>in</strong> pipel<strong>in</strong>es. Acoustic velocity<br />

<strong>in</strong> liquid-filled pipel<strong>in</strong>e can be calculated us<strong>in</strong>g (Wylie <strong>and</strong> Streeter, 1993):<br />

�<br />

�<br />

� 1<br />

a = � �<br />

�<br />

ρ KΓ + 1<br />

� (4.26)<br />

∆A<br />

A ∆p<br />

where, A, is pipel<strong>in</strong>e cross-sectional area, ∆A, is change <strong>in</strong> cross-sectional<br />

area for change <strong>in</strong> pressure, ∆p. Changes <strong>in</strong> cross-sectional area for a given<br />

66


Compressibility [10 −9 Pa −1 ]<br />

0.515<br />

0.490<br />

0.465<br />

0.440<br />

4.5. Acoustic Velocity <strong>in</strong> Fluid-Filled Pipe<br />

ASME Steam Tables<br />

Eisenberg Empirical Correlation<br />

0.415<br />

0 10 20 30 40 50<br />

Temperature [ o C]<br />

Figure 4.19: Water Isothermal Compressibility extrapolated from<br />

Acoustic Velocity <strong>and</strong> Density Correlations (DelGrosso,<br />

1974; Michaelides, 1981) <strong>and</strong> empirically correlated<br />

(Eisenberg <strong>and</strong> Kauzman, 1969)<br />

pressure gradient <strong>in</strong>cluded <strong>in</strong> Eq. 4.26 mirrors Eq. 4.10 <strong>and</strong> def<strong>in</strong>es the <strong>in</strong>verse<br />

of a pipel<strong>in</strong>e compressibility. The greater changes <strong>in</strong> pressure, the greater a<br />

pipel<strong>in</strong>e cross-sectional deformation.<br />

Changes <strong>in</strong> cross-sectional area depend on a pipel<strong>in</strong>e material, geometry <strong>and</strong><br />

anchor<strong>in</strong>g. A pipel<strong>in</strong>e can be th<strong>in</strong> or thick-walled, circular or non-circular<br />

<strong>and</strong> can be anchored at one side only, or at both sides. In the latter case,<br />

the pipel<strong>in</strong>e can be restricted <strong>in</strong> axial movement or <strong>in</strong>clude expansion jo<strong>in</strong>ts,<br />

Figure 4.20. The acoustic velocity for a circular pipel<strong>in</strong>e, when consider<strong>in</strong>g<br />

wall thickness <strong>and</strong> anchor<strong>in</strong>g can be expressed <strong>in</strong> a general form by (Wylie<br />

<strong>and</strong> Streeter, 1993):<br />

�<br />

�<br />

� 1<br />

a = � �<br />

�<br />

ρ KΓ + d<br />

� (4.27)<br />

c1<br />

e Y<br />

where, d, is diameter, e, wall-thickness, Y , pipel<strong>in</strong>e material Young’s modulus,<br />

<strong>and</strong>, c1, structure coefficient that depends on the pipel<strong>in</strong>e fasten<strong>in</strong>gs. Structure<br />

coefficients are presented <strong>in</strong> Table 4.4.<br />

67


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

3<br />

2<br />

1<br />

Figure 4.20: Illustration of Structure Fasten<strong>in</strong>gs (Wylie <strong>and</strong> Streeter,<br />

1993)<br />

Type 1: Pipe Anchored at One End Only.<br />

Type 2: Pipe Anchored at Both Ends.<br />

Type 3: Pipe Anchored at Both Ends with Expansion Jo<strong>in</strong>ts.<br />

Table 4.4: Structure Coefficient for Th<strong>in</strong>-Walled <strong>and</strong> Thick-Walled<br />

Pipel<strong>in</strong>es, S<strong>in</strong>gle-Sided Anchor or Double-Sided Anchor.<br />

Structure Fasten<strong>in</strong>gs Description <strong>in</strong> Figure 4.20<br />

Structure Th<strong>in</strong>-Walled Thick-Walled<br />

Fasten<strong>in</strong>g Pipel<strong>in</strong>e Pipel<strong>in</strong>e<br />

Type c1 c1 c2<br />

1 1 − µ<br />

2<br />

2 1 − µ2<br />

c2 + d<br />

d + e<br />

�<br />

1 − µ<br />

2<br />

c2 + D � 1 − µ 2�<br />

d + e<br />

3 1 c2 + d<br />

d + e<br />

68<br />

�<br />

2e<br />

d<br />

(1 + µ)


4.5. Acoustic Velocity <strong>in</strong> Fluid-Filled Pipe<br />

Pipel<strong>in</strong>e are typically th<strong>in</strong>-walled, anchored at both ends <strong>and</strong> <strong>in</strong>clude expansion<br />

jo<strong>in</strong>ts. Acoustic velocity calculations then follow Korteweg’s formulation<br />

(Tijssel<strong>in</strong>g <strong>and</strong> Anderson, 2007)<br />

�<br />

�<br />

� 1<br />

a = � �<br />

�<br />

ρ KΓ + d<br />

� (4.28)<br />

Y e<br />

Evolution of acoustic velocity with pipel<strong>in</strong>e compressibility for three different<br />

fluids <strong>and</strong> two different pipel<strong>in</strong>e materials is shown <strong>in</strong> Figure 4.21. The three<br />

considered fluids are water, a typical oil <strong>and</strong> a typical gas. Studied fluid densities<br />

<strong>and</strong> isothermal compressibilities are listed <strong>in</strong> Table 4.6 (McCa<strong>in</strong>, 1990).<br />

Listed density <strong>and</strong> compressibility values are <strong>in</strong> accordance with previously<br />

calculated values. The two considered pipel<strong>in</strong>e materials are steel <strong>and</strong> PVC.<br />

Studied materials Young’s moduli <strong>and</strong> Poisson’s ratio are listed <strong>in</strong> Table 4.5<br />

(Cheremis<strong>in</strong>off, 1996).<br />

Acoustic velocity calculations us<strong>in</strong>g Eq. 4.28 are for liquids only. However,<br />

the same equation can be used for gas when the z-factor value is neglected.<br />

Another way to express this is that Eq. 4.28 can be used for gas when pipel<strong>in</strong>e<br />

compressibility effect is much greater than the z-factor effect on the acoustic<br />

velocity calculation. In addition, calculated acoustic velocities are used only<br />

for representation of pipel<strong>in</strong>e compressibility effect.<br />

The acoustic velocity <strong>in</strong>creases with a pipel<strong>in</strong>e wall thickness <strong>and</strong> <strong>in</strong>ternal<br />

diameter ratio, <strong>and</strong> with Young’s modulus, as shown <strong>in</strong> Figure 4.21. For a<br />

water-filled th<strong>in</strong> PVC pipel<strong>in</strong>e, the acoustic velocity <strong>in</strong>creases from 358 ms −1<br />

at 5 % e/d ratio, to 919 ms −1 at 50 % e/d ratio, <strong>and</strong> to 1107 ms −1 at 100<br />

% e/d ratio. The acoustic velocity <strong>in</strong>creases with a pipel<strong>in</strong>e material Young’s<br />

modulus as shown <strong>in</strong> Figure 4.21. For a 50 % e/d ratio, the acoustic velocity<br />

<strong>in</strong>creases from 919 ms −1 for a water-filled th<strong>in</strong> PVC pipel<strong>in</strong>e, to 1477 ms −1<br />

for a water-filled th<strong>in</strong> steel pipel<strong>in</strong>e.<br />

Acoustic velocities for th<strong>in</strong>-walled <strong>and</strong> thick-walled steel pipel<strong>in</strong>es are similar<br />

for all fluids, but are dist<strong>in</strong>ct for PVC pipel<strong>in</strong>es, as shown <strong>in</strong> Figure 4.21. For<br />

a 50 % e/d ratio <strong>and</strong> an oil-filled PVC pipel<strong>in</strong>e, calculated acoustic velocities<br />

are 616 ms −1 us<strong>in</strong>g the th<strong>in</strong>-walled formula, Eq. 4.28, <strong>and</strong> 699 ms −1 us<strong>in</strong>g<br />

the thick-walled formula, Eq. 4.27 <strong>and</strong> Table 4.4. For a 50 % e/d ratio<br />

<strong>and</strong> a water-filled PVC pipel<strong>in</strong>e, calculated acoustic velocities are 711 ms −1<br />

us<strong>in</strong>g the th<strong>in</strong>-walled formula, Eq. 4.28, <strong>and</strong> 919 ms −1 us<strong>in</strong>g the thick-walled<br />

formula, Eq. 4.27 <strong>and</strong> Table 4.4.<br />

69


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Acoustic Velocity [m s −1 ]<br />

1500<br />

1250<br />

1000<br />

750<br />

500<br />

Water <strong>in</strong> Steel<br />

Oil <strong>in</strong> Steel<br />

Gas <strong>in</strong> Steel<br />

Water <strong>in</strong> PVC<br />

250<br />

Oil <strong>in</strong> PVC<br />

Gas <strong>in</strong> PVC<br />

Th<strong>in</strong>−Walled Pipel<strong>in</strong>e<br />

Thick−Walled Pipel<strong>in</strong>e<br />

0<br />

0 25 50 75 100<br />

Wall Thickness / Diameter [%]<br />

Figure 4.21: Acoustic Velocity <strong>in</strong> Fluid-Filled Pipel<strong>in</strong>es (Wylie <strong>and</strong><br />

Streeter, 1993)<br />

Table 4.5: Typical Material Properties: Young’s Modulus <strong>and</strong><br />

Poisson’s Ratio (Cheremis<strong>in</strong>off, 1996)<br />

Material Young’s Modulus [Pa] Poisson’s Ratio<br />

Steel 200 10 9 0.30<br />

PVC 2.8 10 9 0.41<br />

Table 4.6: Typical Fluid Densities <strong>and</strong> Compressibilities <strong>in</strong><br />

Petroleum Eng<strong>in</strong>eer<strong>in</strong>g (McCa<strong>in</strong>, 1990)<br />

Fluid Density [kg m −3 ] Compressibility [Pa −1 ]<br />

Water 1030 0.435 10 −9<br />

Oil 760 1.98 10 −9<br />

Gas 60 188 10 −9<br />

70


4.6. Acoustic Velocity Determ<strong>in</strong>ation<br />

Acoustic velocity values are determ<strong>in</strong>ed either by fluid compressibility or by<br />

pipel<strong>in</strong>e compressibility. The lowest compressibility determ<strong>in</strong>es the acoustic<br />

velocity. Water compressibility is 100 times greater than a steel pipel<strong>in</strong>e compressibility.<br />

Hence, acoustic velocity <strong>in</strong> water-filled pipel<strong>in</strong>e is limited by the<br />

pipel<strong>in</strong>e compressibility. On the other h<strong>and</strong>, gas compressibility are commonly<br />

100 times lower than a PVC pipel<strong>in</strong>e compressibility. Thus, acoustic velocity<br />

<strong>in</strong> gas-filled pipel<strong>in</strong>e is limited by the gas compressibility. Oil <strong>and</strong> steel compressibilities<br />

are of the same order of magnitude, therefore acoustic velocity<br />

<strong>in</strong> oil-filled pipel<strong>in</strong>e is greatly dependent on the ration e/d.<br />

Fluid transport can also take place <strong>in</strong> non-circular pipel<strong>in</strong>es. Such pipel<strong>in</strong>es are<br />

not usual <strong>in</strong> petroleum eng<strong>in</strong>eer<strong>in</strong>g, however the acoustic velocity formulation<br />

for non-circular pipel<strong>in</strong>es is given for completeness. For a non-circular pipel<strong>in</strong>e,<br />

changes <strong>in</strong> cross-sectional area with pressure depend on the number of sides,<br />

N, mak<strong>in</strong>g the polygonal shape. Changes <strong>in</strong> cross-sectional area to be <strong>in</strong>cluded<br />

<strong>in</strong> the acoustic velocity formulation, Eq. 4.26, is (Thorley, 1991):<br />

1 ∆A d<br />

=<br />

A ∆p Y e<br />

⎛<br />

⎜<br />

tan<br />

⎜<br />

⎝1 +<br />

4<br />

� �<br />

180<br />

N<br />

15<br />

� d<br />

e<br />

� 2<br />

⎞<br />

⎟<br />

⎠<br />

(4.29)<br />

Acoustic velocities for circular, square, hexagonal <strong>and</strong> octagonal water-filled<br />

steel pipel<strong>in</strong>es, are shown <strong>in</strong> Figure 4.22. The acoustic velocity <strong>in</strong>creases with<br />

the number of sides mak<strong>in</strong>g the polygonal cross-section. For a 50 % d/e<br />

ratio, the acoustic velocity <strong>in</strong>creases from 151 ms −1 for a square pipe, to 427<br />

ms −1 for an hexagonal pipe, <strong>and</strong> to 713 ms −1 for an octagonal pipel<strong>in</strong>e. The<br />

acoustic velocity <strong>in</strong> a water- filled steel circular pipel<strong>in</strong>e is 1191 ms −1 for a 50<br />

% e/d ratio.<br />

4.6 Acoustic Velocity Determ<strong>in</strong>ation<br />

The acoustic velocity <strong>in</strong> a fluid-filled pipel<strong>in</strong>e depends on fluid density <strong>and</strong><br />

isothermal compressibility, <strong>and</strong> on pipel<strong>in</strong>e geometry <strong>and</strong> material. Acoustic<br />

velocity <strong>in</strong> a fluid-filled pipel<strong>in</strong>e can be calculated us<strong>in</strong>g Eq. 4.28 or Eq. 4.19,<br />

for <strong>in</strong>stance. But the acoustic velocity can also be measured experimentally<br />

from pressure data recorded by two transducers set apart along a pipel<strong>in</strong>e<br />

dur<strong>in</strong>g a pressure wave propagation experiment.<br />

Previously def<strong>in</strong>ed acoustic velocities are phase velocities while measured acoustic<br />

velocities are group velocities. Phase velocity refers to a theoretical s<strong>in</strong>gle<br />

frequency wave, whereas experimental pressure waves consist of a broad range<br />

71


Chapter 4. Acoustic Velocity <strong>in</strong> S<strong>in</strong>gle Phase Fluids<br />

Acoustic Velocity [m s −1 ]<br />

1500<br />

1250<br />

1000<br />

750<br />

500<br />

250<br />

Circular Pipe<br />

4 − Sides<br />

6 − Sides<br />

8 − Sides<br />

0<br />

0 25 50 75 100<br />

Diameter / Wall Thickness [%]<br />

Figure 4.22: Acoustic Velocity <strong>in</strong> Fluid-Filled Non-Circular Pipel<strong>in</strong>es<br />

of frequencies that def<strong>in</strong>e a group of s<strong>in</strong>gle frequency waves. Frequency dependent<br />

acoustic velocities can however be determ<strong>in</strong>ed from group velocity <strong>and</strong><br />

are expressed by (Tijssel<strong>in</strong>g <strong>and</strong> Anderson, 2007):<br />

�<br />

� ⎛ �<br />

�<br />

�<br />

a(ω) = a�2<br />

⎝1 + 1 +<br />

� 32ν<br />

dω<br />

� 2<br />

⎞<br />

⎠<br />

−1<br />

(4.30)<br />

where, ω, represents circular frequency, ν, k<strong>in</strong>ematic viscosity, a(ω), s<strong>in</strong>gle<br />

frequency phase velocity, <strong>and</strong>, a, measured group velocity. The measured<br />

group velocity must be comparable with the calculated value us<strong>in</strong>g Eq. 4.28.<br />

When fluid compressibilities are of the same order of magnitude as pipel<strong>in</strong>e<br />

compressibilities, then the thick-walled pipel<strong>in</strong>e formulation should be used,<br />

Table 4.4.<br />

72


Chapter 5<br />

Numerical Simulation<br />

Methods<br />

5.1 Modell<strong>in</strong>g Equations<br />

Analysis of pressure wave propagation <strong>in</strong> fluid-filled pipel<strong>in</strong>es is performed<br />

us<strong>in</strong>g the fundamental conservation equations. Heat transfer is commonly neglected<br />

<strong>in</strong> pressure transient analysis. Thus the energy conservation equation<br />

is excluded from the system of equations to be solved. One-dimensional mass<br />

<strong>and</strong> momentum equations for viscous fluids are then, respectively (Chevray<br />

<strong>and</strong> Mathieu, 1993):<br />

∂ρ ∂<br />

+ (ρu) = 0 (5.1)<br />

∂t ∂x<br />

∂ ∂<br />

(ρu) +<br />

∂t ∂x (ρu2 ) = −∂p ∂<br />

+<br />

∂x ∂x<br />

�<br />

4<br />

3 µ∂u<br />

�<br />

+<br />

∂x<br />

4τω<br />

πd<br />

(5.2)<br />

where, p, represents pressure, u, fluid velocity, ρ, fluid density, µ, fluid dynamic<br />

viscosity, <strong>and</strong>, τω, wall shear stress. Wall shear stress <strong>in</strong>cludes frictional pressure<br />

drop <strong>and</strong> is conventionally expressed by the Darcy-Weisbach formulation<br />

(Ghidaoui et al., 2005). Simplifications of the above system of two equations,<br />

Eq. 5.1 <strong>and</strong> Eq. 5.2 are obta<strong>in</strong>ed from order-of-magnitude analysis, neglect<strong>in</strong>g<br />

viscosity <strong>and</strong> restat<strong>in</strong>g the system <strong>in</strong> terms of pressure <strong>and</strong> fluid velocity<br />

variables. The system of equations then becomes (Wylie <strong>and</strong> Streeter, 1993):<br />

1<br />

ρa2 ∂p ∂u<br />

+ = 0 (5.3)<br />

∂t ∂x<br />

∂u<br />

∂t<br />

1 ∂p<br />

+<br />

ρ ∂x<br />

−1 f<br />

=<br />

2 d u2 − gs<strong>in</strong>(θ) (5.4)<br />

73


Chapter 5. Numerical Simulation Methods<br />

where, f, represents the Darcy-Weisbach friction factor, d, pipel<strong>in</strong>e diameter,<br />

a, pressure wave propagation velocity, <strong>and</strong>, θ, pipel<strong>in</strong>e <strong>in</strong>cl<strong>in</strong>ation from the<br />

horizontal. The wave propagation velocity can be expressed by (Lighthill,<br />

1978):<br />

�<br />

∂p<br />

a =<br />

(5.5)<br />

∂ρ<br />

The above system of two first-order partial differential equations, Eq. 5.3 <strong>and</strong><br />

Eq. 5.4 can be restated <strong>in</strong> a conservative form <strong>and</strong> <strong>in</strong> a matrix form. Eq. 5.3<br />

<strong>and</strong> Eq. 5.4 <strong>in</strong> conservative form become (Laney, 1998):<br />

∂Ω<br />

∂t<br />

+ ∂Φ(Ω)<br />

∂x<br />

= Λ (5.6)<br />

consist<strong>in</strong>g of the vector variable, Ω, the flux vector, Φ, <strong>and</strong> the source term<br />

vector, Λ:<br />

� � � � � �<br />

p ρa2u 0<br />

Ω = Φ(Ω) = Λ =<br />

(5.7)<br />

u<br />

p s<br />

Matrix form of Eq. 5.3 <strong>and</strong> Eq. 5.4 is expressed by (Laney, 1998):<br />

I ∂Ω<br />

∂t<br />

− B∂Ω<br />

∂x<br />

= Λ (5.8)<br />

where, Ω <strong>and</strong> Λ are vectors def<strong>in</strong>ed <strong>in</strong> Eq. 5.7, I, represents the unity matrix,<br />

<strong>and</strong> B is a quadratic matrix def<strong>in</strong>ed by:<br />

� �<br />

0 ρa2 B =<br />

(5.9)<br />

1 0<br />

Explicit analytical solutions of the system of two first-order partial differential<br />

equations, Eq. 5.3 <strong>and</strong> Eq. 5.4, or Eq. 5.6, or Eq. 5.8, are rarely possible.<br />

However, simple assumptions can lead to simplifications of the two equations<br />

to be solved, <strong>and</strong> first approximations can be postulated. Nevertheless, several<br />

numerical methods can be used to perform more complete transient analysis.<br />

74


5.2 Analytical Solutions<br />

5.2. Analytical Solutions<br />

Rigid water column is an analytical method for calculat<strong>in</strong>g pressure values <strong>and</strong><br />

transient times <strong>in</strong> simplified water-hammer problems. Simplifications assume<br />

fully compressible fluid, fully rigid pipel<strong>in</strong>e walls <strong>and</strong> a long transient generation.<br />

The transient generation is relative to a pressure wave travel period<br />

<strong>in</strong> a fluid-filled pipel<strong>in</strong>e. <strong>Pressure</strong> transients are rapid if the generation time<br />

is much less than the travel period. <strong>Pressure</strong> transients are considered long<br />

otherwise. The travel period is def<strong>in</strong>ed by (Wylie <strong>and</strong> Streeter, 1993):<br />

T = 2L<br />

a<br />

(5.10)<br />

where, L, represents pipel<strong>in</strong>e length. A schematic representation of a pipel<strong>in</strong>e<br />

system typically used <strong>in</strong> pressure transient analysis is shown <strong>in</strong> Figure 5.1.<br />

Figure 5.1: Typical Pipel<strong>in</strong>e System <strong>in</strong> <strong>Pressure</strong> Transient Analysis<br />

Schematic Representation<br />

The system is composed of a pipel<strong>in</strong>e connected to a tank at one end, <strong>and</strong> to<br />

a quick-act<strong>in</strong>g valve at the other end. The valve closure generates a pressure<br />

transient that propagates back <strong>and</strong> forth <strong>in</strong> the fluid-filled pipel<strong>in</strong>e. Us<strong>in</strong>g rigid<br />

water column approximations, the time to reach a new steady-state condition<br />

after transient <strong>in</strong>itiation, <strong>and</strong> the steady state pressure are (Thorley, 1991):<br />

t = −ρLu<br />

∆p<br />

pu = pd + 1 L<br />

ρf<br />

2 d u20 L<br />

d<br />

(5.11)<br />

(5.12)<br />

where, pu is the upstream pressure <strong>and</strong> pd, the downstream pressure. Information<br />

provided by the rigid water column approximations are <strong>in</strong>sufficient <strong>in</strong><br />

conventional pressure transient analysis s<strong>in</strong>ce multiple transmissions <strong>and</strong> reflections<br />

can take place along the fluid-filled pipel<strong>in</strong>e. However, the analytical<br />

solution is practical for a first estimation.<br />

75


Chapter 5. Numerical Simulation Methods<br />

A graphical method was the first method to solve the complete system of equations<br />

Eq. 5.3 <strong>and</strong> Eq. 5.4. Characteristics were calculated <strong>and</strong> represented<br />

<strong>in</strong> a spatial-temporal graph (x-t plane). The graphical method transformed<br />

<strong>in</strong>to the method of characteristics when computer calculation permitted to<br />

compute characteristics more efficiently than h<strong>and</strong>-written methods.<br />

5.3 Method of Characteristics<br />

The l<strong>in</strong>ear hyperbolic system of two first order partial differential equations<br />

formed by Eq. 5.3 <strong>and</strong> Eq. 5.4 relates pressure, p, <strong>and</strong> fluid velocity, u. The<br />

l<strong>in</strong>ear hyperbolic system is solved <strong>in</strong> the discretized x-t plane. Grid po<strong>in</strong>ts<br />

spatial <strong>and</strong> temporal distances are ∆x <strong>and</strong> ∆t, respectively. The grid can be<br />

rectangular, diamond-shaped staggered, or composed by the l<strong>in</strong>ear hyperbolic<br />

system characteristics. The rectangular double grid (Figure 5.2) is the most<br />

commonly used (Wylie <strong>and</strong> Streeter, 1993).<br />

∆t<br />

t<br />

∆x<br />

C + C −<br />

Figure 5.2: Characteristic L<strong>in</strong>es Representation <strong>in</strong> St<strong>and</strong>ard X-T<br />

Plane Grid<br />

The hyperbolic system has two dist<strong>in</strong>ct real characteristics. Characteristic<br />

l<strong>in</strong>es <strong>in</strong> the x-t plane form a new system of coord<strong>in</strong>ates where the hyperbolic<br />

system Eq. 5.6 transforms <strong>in</strong>to four first-order ord<strong>in</strong>ary differential equations<br />

that can be solved numerically by f<strong>in</strong>ite difference approximations. Characteristics<br />

are easily obta<strong>in</strong>ed from the eigenvalues of the system <strong>in</strong> its matrix<br />

form, Eq. 5.8, <strong>and</strong> are calculated by (Wylie <strong>and</strong> Streeter, 1993):<br />

|I − βB| = 0 (5.13)<br />

where, β, is a parameter that has two roots. Values of β def<strong>in</strong>e the characteristics:<br />

dx<br />

= β (5.14)<br />

dt<br />

76<br />

x


5.3. Method of Characteristics<br />

Characteristics of Eq. 5.3 <strong>and</strong> Eq. 5.4 are then ±a. The source term, s, does<br />

not <strong>in</strong>fluence the characteristics of the system. Characteristic l<strong>in</strong>es <strong>in</strong> the x-t<br />

plane are l<strong>in</strong>es of constant slope that def<strong>in</strong>e a new coord<strong>in</strong>ate system. In the<br />

new coord<strong>in</strong>ate system, the two first-order partial differential equations transform<br />

<strong>in</strong>to four first-order ord<strong>in</strong>ary differential equations that can be grouped<br />

<strong>in</strong> two characteristic paths, (Wylie <strong>and</strong> Streeter, 1993):<br />

C + ⎧<br />

⎪⎨<br />

dp<br />

+ adu = −as<br />

= dt dt<br />

⎪⎩<br />

dx<br />

= +a<br />

dt<br />

C − ⎧<br />

⎪⎨<br />

dp<br />

− adu = as<br />

= dt dt<br />

⎪⎩<br />

dx<br />

= −a<br />

dt<br />

(5.15)<br />

(5.16)<br />

where, C + , represents the forward characteristic path, <strong>and</strong>, C − , the backward<br />

characteristics path. Each group of equations can be solved <strong>in</strong>to the characteristic<br />

system of coord<strong>in</strong>ates such that pressure <strong>and</strong> fluid velocity values<br />

can be calculated at each <strong>in</strong>tersection of two characteristic l<strong>in</strong>es. A typical<br />

representation of a x-t plane <strong>and</strong> the correspond<strong>in</strong>g system characteristics for<br />

a simple pipel<strong>in</strong>e system problem is shown <strong>in</strong> Figure 5.2.<br />

The Courant condition def<strong>in</strong>es the maximum distances between grid po<strong>in</strong>ts.<br />

The Courant condition is expressed by:<br />

a∆t ≤ ∆x (5.17)<br />

Characteristic l<strong>in</strong>es <strong>in</strong>tersections <strong>and</strong> x-t plane nodes do not necessarily co<strong>in</strong>cide<br />

as represented <strong>in</strong> Figure 5.2. Hence, pressure <strong>and</strong> fluid velocity values at<br />

x-t plane nodes must be <strong>in</strong>terpolated (Vardy, 2008). Interpolations decrease<br />

calculation accuracy. A mov<strong>in</strong>g x-t plane grid can be used <strong>in</strong>stead, where ∆x<br />

<strong>and</strong> ∆t become δx <strong>and</strong> δt, <strong>and</strong> vary with time <strong>and</strong> space. However, adaptive<br />

grid methods are more expensive <strong>in</strong> computer resources <strong>and</strong> time, <strong>and</strong> mesh<br />

ref<strong>in</strong>ement close to boundaries is also necessary.<br />

Boundary conditions such as pumps, tanks or valves, require pressure <strong>and</strong><br />

fluid velocity values for the problem to be solvable. A problem is underdef<strong>in</strong>ed<br />

when hyper-static. Another way to express this is that a problem is<br />

def<strong>in</strong>ed when the number of parameters equals the number of equations <strong>and</strong><br />

sufficient <strong>in</strong>itial conditions are known. Interpolations are generally required<br />

at boundaries s<strong>in</strong>ce characteristic l<strong>in</strong>es <strong>in</strong>tersections barely co<strong>in</strong>cide with grid<br />

nodes, even us<strong>in</strong>g an adaptive grid method.<br />

77


Chapter 5. Numerical Simulation Methods<br />

The method of characteristics offers an eng<strong>in</strong>eer-friendly method for comput<strong>in</strong>g<br />

pressure <strong>and</strong> fluid velocity values <strong>in</strong> a pipel<strong>in</strong>e dur<strong>in</strong>g pressure transient<br />

propagation. The method of characteristics is first order <strong>in</strong> both space <strong>and</strong><br />

time because of f<strong>in</strong>ite difference approximations of first-order ord<strong>in</strong>ary differential<br />

equations. Nonetheless, the method of characteristics provide fast<br />

solution with an acceptable accuracy that can be improved by grid ref<strong>in</strong>ement<br />

<strong>and</strong> high order <strong>in</strong>terpolations.<br />

Numerical errors <strong>and</strong> large deviations from experimental data are often reported<br />

<strong>in</strong> pressure transient analysis. Of particular <strong>in</strong>terest is the non-physical<br />

numerical attenuation due to <strong>in</strong>terpolation to obta<strong>in</strong> pressure <strong>and</strong> fluid velocity<br />

values at grid po<strong>in</strong>ts. Despite its eas<strong>in</strong>ess <strong>and</strong> efficiency to compute results<br />

<strong>in</strong> pipel<strong>in</strong>e systems, research tends to focus on different numerical techniques<br />

<strong>and</strong> schemes such as f<strong>in</strong>ite volume methods (Soares et al., 2008).<br />

5.4 F<strong>in</strong>ite Volume Methods<br />

5.4.1 Description<br />

F<strong>in</strong>ite volume methods are based on the <strong>in</strong>tegral form of Eq. 5.6 <strong>in</strong> order to<br />

use grid cells <strong>in</strong>stead of grid po<strong>in</strong>ts. Total <strong>in</strong>tegrals of vector variable Ω are<br />

now evaluated over each grid cells. Grid cell Q values evolve with time by the<br />

flux over the edges of each grid cells. Flux functions <strong>and</strong> grid cells averag<strong>in</strong>g<br />

methods def<strong>in</strong>e several numerical schemes. The <strong>in</strong>tegral form of Eq. 5.6 is<br />

(LeVeque, 2002):<br />

�<br />

d<br />

�<br />

Ω(x, t)dx = Φ<br />

dt ζi<br />

ϕ(x i− 1<br />

2<br />

� �<br />

, t) − Φ<br />

ϕ(x i+ 1<br />

2<br />

�<br />

, t)<br />

(5.18)<br />

where, ζ, is the spatial doma<strong>in</strong>, <strong>and</strong> the above <strong>in</strong>tegration def<strong>in</strong>es the average<br />

value over the ith cell. The spatial doma<strong>in</strong> is divided <strong>in</strong>to grid cells (or f<strong>in</strong>ite<br />

volumes) of length:<br />

∆x = x 1<br />

i+ − x 1<br />

i− 2 2<br />

(5.19)<br />

The vector variable Ω evolves with space <strong>and</strong> time. The cell average at each<br />

position <strong>and</strong> each time depends on the flux from the adjacent cells as illustrated<br />

<strong>in</strong> Figure 5.3. A cell average value after <strong>in</strong>crement<strong>in</strong>g the time step is expressed<br />

by (LeVeque, 2002):<br />

Ω n+1<br />

i<br />

= Ωni − ∆t<br />

� �<br />

�Φ n<br />

Ωi , Ω<br />

∆x<br />

n � �<br />

i+1 − Φ� n<br />

Ωi−1, Ω n� i<br />

�<br />

78<br />

(5.20)


tn+1<br />

tn<br />

Φ n<br />

i− 1<br />

2<br />

Ω n+1<br />

i<br />

Φ n<br />

i+ 1<br />

2<br />

Ω n i−1 Ω n i Ω n i+1<br />

5.4. F<strong>in</strong>ite Volume Methods<br />

Figure 5.3: Illustration of a F<strong>in</strong>ite Volume Method for Updat<strong>in</strong>g the<br />

Cell Average Ωn i by Fluxes at the Cell Edges<br />

where, i, represents the space <strong>in</strong>dex, n, the time <strong>in</strong>dex, <strong>and</strong>, � Φ, a numerical<br />

flux function that approximates the average flux function Φ. Different f<strong>in</strong>ite<br />

volume numerical schemes correspond to different formulation of � Φ <strong>in</strong> Eq. 5.20.<br />

Available f<strong>in</strong>ite volume methods try to <strong>in</strong>clude a Riemann solver for captur<strong>in</strong>g<br />

wave propagation. The Riemann problem occurs at a grid edge where the data<br />

is piecewise constant with a s<strong>in</strong>gle jump discont<strong>in</strong>uity. The Riemann problem<br />

is expressed by (Laney, 1998):<br />

Ω(xi, tn) =<br />

�<br />

Ωl if x < 0<br />

Ωr if x > 0<br />

(5.21)<br />

where, Ωl <strong>and</strong> Ωr, are solutions just before <strong>and</strong> just after the grid node x = 0,<br />

respectively.<br />

5.4.2 Numerical Schemes<br />

F<strong>in</strong>ite volume numerical schemes differ from one another by the flux average<br />

approximation function, � Φ. Three different numerical schemes are here<strong>in</strong> studied<br />

based on three different flux averag<strong>in</strong>g techniques. The <strong>in</strong>vestigated numerical<br />

scheme are Central Time Central Space (CTCS), Lax-Wendroff (LW),<br />

<strong>and</strong> Essentially Non-Oscillatory (ENO). The three schemes are explicit <strong>and</strong><br />

second-order accurate <strong>in</strong> both space <strong>and</strong> time (Laney, 1998). The three selected<br />

schemes represent three different mathematical concept.<br />

The CTCS method is a conservative numerical scheme based on f<strong>in</strong>ite difference<br />

approximations of first-order derivatives. The CTCS scheme is also<br />

referred to as the Leapfrog method. The CTCS numerical scheme locates<br />

shocks accurately s<strong>in</strong>ce it is a conservative scheme. However, shock shapes are<br />

not correctly predicted because of the odd-even time <strong>and</strong> space decoupl<strong>in</strong>g.<br />

79


Chapter 5. Numerical Simulation Methods<br />

The CTCS numerical scheme is centered, l<strong>in</strong>ear, <strong>and</strong> expressed by (Laney,<br />

1998):<br />

Ω n+1<br />

i<br />

= Ωn−1<br />

i<br />

∆t � n<br />

− 2 Φ(Ω<br />

∆x<br />

i+1) − Φ(Ω n i−1) �<br />

(5.22)<br />

The Lax-Wendroff numerical scheme is a l<strong>in</strong>ear, centered method, based on<br />

decomposition of Eq. 5.6 <strong>in</strong> Taylor series. Several other methods are derived<br />

from the Lax-Wendroff technique of Taylor series decomposition by chang<strong>in</strong>g<br />

flux averag<strong>in</strong>g. The Lax-Wendroff numerical scheme <strong>in</strong> its conservative form<br />

is expressed by (Laney, 1998):<br />

Ω n+1<br />

i<br />

ˆΦ n<br />

i+ 1<br />

2<br />

a n<br />

i+ 1<br />

2<br />

= Ωn i − ∆t<br />

�<br />

ˆΦ n<br />

i+ 1 −<br />

2<br />

ˆ Φ n<br />

i− 1<br />

�<br />

2<br />

∆x<br />

= 1 � n<br />

Φ(Ω<br />

2<br />

i+1) + Φ(Ω n i ) � − ∆t<br />

2∆x an<br />

i+ 1<br />

� n<br />

Φ(Ωi+1) − Φ(Ω<br />

2<br />

n i ) �<br />

⎧<br />

⎪⎨ Φ(Ω<br />

=<br />

⎪⎩<br />

n i+1 ) + Φ(Ωni )<br />

Ωn i+1 − Ωn for Ω<br />

i<br />

n i �= Ωn i+1<br />

Φ(Ω n i ) for Ωn i = Ωn i+1<br />

(5.23)<br />

The ENO numerical scheme is a piecewise-polynomial reconstruction-evolution<br />

method based on polynomial <strong>in</strong>terpolation. The ENO scheme reconstructs<br />

solutions from cell-<strong>in</strong>tegral averages. The numerical scheme is a high resolution<br />

method of second order but less stable than the Lax-Wendroff scheme for<br />

<strong>in</strong>stance. The ENO numerical scheme is not detailed here but can be found<br />

<strong>in</strong> st<strong>and</strong>ard textbooks such as (Laney, 1998).<br />

5.4.3 Test Cases<br />

Four test cases were used to assess strengths <strong>and</strong> weaknesses of the three<br />

studied numerical schemes. The four test cases descriptions are given <strong>in</strong> Table<br />

5.1. The first test case illustrates phase <strong>and</strong> amplitude error on a completely<br />

smooth solution. The second test case illustrates progressive contact smear<strong>in</strong>g<br />

<strong>and</strong> dispersion with two jump discont<strong>in</strong>uities. The third test case is similar<br />

to the second but illustrates the importance of the number of grid po<strong>in</strong>ts.<br />

The fourth test case illustrates sonic po<strong>in</strong>t importance <strong>in</strong> the Burger equation<br />

expressed by (Laney, 1998):<br />

∂ϕ<br />

∂t<br />

+ ∂<br />

∂x<br />

� 1<br />

2 ϕ2<br />

�<br />

80<br />

(5.24)


Table 5.1: Test Cases Description for Compar<strong>in</strong>g CTCS,<br />

Lax-Wendroff <strong>and</strong> ENO Methods<br />

Case Equation<br />

5.4. F<strong>in</strong>ite Volume Methods<br />

Initial Conditions Grid<br />

ϕ(x,0) Po<strong>in</strong>ts<br />

1<br />

−s<strong>in</strong>(πx)<br />

40<br />

2<br />

∂ϕ ∂ϕ<br />

+ = 0<br />

∂t ∂x ⎧<br />

⎪⎨ 1<br />

⎪⎩<br />

for |x| < 1<br />

0<br />

3<br />

for 1<br />

3<br />

4<br />

�<br />

∂ϕ ∂ 1<br />

+<br />

∂t ∂x 2<br />

< |x| ≤ 1<br />

3<br />

600<br />

ϕ2<br />

�<br />

= 0 40<br />

Comparison of the three numerical schemes on test case 1 is shown <strong>in</strong> Figure<br />

5.4. The three numerical schemes capture the s<strong>in</strong>usoidal shape. Lax-Wendroff<br />

<strong>and</strong> CTCS calculated solutions are almost superimposed. Both CTCS <strong>and</strong><br />

Lax-Wendroff predict the s<strong>in</strong>usoidal amplitude correctly but the solutions lag<br />

beh<strong>in</strong>d the exact solution. The ENO method predicts phase correctly but the<br />

s<strong>in</strong>usoidal wave is smeared at both extremes.<br />

Comparison of the three numerical schemes on test case 2 is shown <strong>in</strong> Figure<br />

5.5. Both CTCS <strong>and</strong> Lax-Wendroff methods locate shocks properly. However,<br />

predicted solutions by CTCS <strong>and</strong> Lax-Wendroff numerical schemes undershoot<br />

<strong>and</strong> overshoot the exact solution, <strong>and</strong> both result behaviours are greatly oscillatory.<br />

The ENO method captures the square wave shape properly <strong>and</strong> locates<br />

shocks correctly with moderate smear<strong>in</strong>g.<br />

Comparison of the three numerical schemes on test case 3 is shown <strong>in</strong> Figure<br />

5.6. Increas<strong>in</strong>g the number of grid po<strong>in</strong>ts improves prediction of shock<br />

locations. However, both CTCS <strong>and</strong> Lax-Wendroff methods exhibit a highly<br />

oscillatory behaviour that overshoots <strong>and</strong> undershoots the exact solution. On<br />

the other h<strong>and</strong>, the ENO numerical scheme smear<strong>in</strong>g of the square wave observed<br />

<strong>in</strong> Figure 5.5 is now reduced.<br />

Comparison of the three numerical schemes on test case 4 is shown <strong>in</strong> Figure<br />

5.7. The CTCS numerical scheme fails to predict correct location or amplitude<br />

of the exact solution. The Lax-wendroff locates accurately the shock <strong>and</strong> the<br />

expansion fan but still undershoots <strong>and</strong> overshoots the exact solution, especially<br />

close to the sonic po<strong>in</strong>t. The ENO scheme predicts accurately location<br />

<strong>and</strong> amplitude of the expansion fan <strong>and</strong> the shock.<br />

81


Chapter 5. Numerical Simulation Methods<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

−0.25<br />

−0.50<br />

−0.75<br />

−1.00<br />

Exact Solution<br />

CTCS<br />

Lax−Wendroff<br />

ENO<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.4: Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical<br />

Schemes on Test Case 1<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

Exact Solution<br />

CTCS<br />

Lax−Wendroff<br />

ENO<br />

−0.25<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.5: Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical<br />

Schemes on Test Case 2<br />

82


1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

5.4. F<strong>in</strong>ite Volume Methods<br />

Exact Solution<br />

CTCS<br />

Lax−Wendroff<br />

ENO<br />

−0.25<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.6: Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical<br />

Schemes on Test Case 3<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

Exact Solution<br />

CTCS<br />

Lax−Wendroff<br />

ENO<br />

−0.25<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.7: Comparison of CTCS, Lax-Wendroff <strong>and</strong> ENO Numerical<br />

Schemes on Test Case 4<br />

83


Chapter 5. Numerical Simulation Methods<br />

The three second-order accurate methods <strong>in</strong> both space <strong>and</strong> time do not perform<br />

identically. The ENO scheme performs better globally than both CTCS<br />

<strong>and</strong> Lax-Wendroff. Lax-Wendroff numerical scheme captures shocks accurately<br />

but presents artificial oscillations. Nevertheless, the Lax-Wendroff numerical<br />

scheme is of particular <strong>in</strong>terest because flux limiter functions can be<br />

added to <strong>in</strong>crease the resolution of the method.<br />

5.4.4 Higher Resolution Method<br />

Flux limiter coupled to second-order accurate numerical schemes <strong>in</strong>creases<br />

the resolution of the methods. The Lax-Wendroff numerical results can be<br />

improved by add<strong>in</strong>g a flux limiter to the flux function. The numerical scheme<br />

is still second-order accurate <strong>in</strong> both space <strong>and</strong> time, but the resolution is<br />

<strong>in</strong>creased, especially close to shocks. The Lax-Wendroff numerical scheme<br />

coupled to a flux limiter function is expressed by (Laney, 1998):<br />

Ω n+1<br />

i<br />

= Ωni − ∆t<br />

∆x ˆ �<br />

φ(θ) ˆF n<br />

i+ 1 −<br />

2<br />

ˆ F n<br />

i− 1<br />

�<br />

2<br />

(5.25)<br />

where, ˆ φ, represents a flux-limiter function, <strong>and</strong>, ˆ Φ, has been def<strong>in</strong>ed <strong>in</strong><br />

Eq. 5.23. Four flux limiter functions are here<strong>in</strong> studied: m<strong>in</strong>mod, superbee,<br />

monotonized-centered <strong>and</strong> Van-Leer flux limiter. The four flux limiter<br />

functions are given <strong>in</strong> Table 5.2. Comparison of the four flux limiter functions<br />

was performed us<strong>in</strong>g the four test cases described previously.<br />

Table 5.2: M<strong>in</strong>mod, Superbee, Monotonized-Centered <strong>and</strong> Van-Leer<br />

Flux Limiter Functions (Laney, 1998)<br />

Flux Limiter Function<br />

M<strong>in</strong>mod<br />

Superbee<br />

ˆ φ(θ) = max (0, m<strong>in</strong>(1, θ))<br />

ˆ φ(θ) = max (0, m<strong>in</strong>(1, 2θ), m<strong>in</strong>(2, θ))<br />

Monotonized-Centered ˆ � � ��<br />

1 + θ<br />

φ(θ) = max 0, m<strong>in</strong> , 2, 2θ<br />

2<br />

Van-Leer<br />

84<br />

ˆ φ(θ) = θ + |θ|<br />

1 + |θ|


5.4. F<strong>in</strong>ite Volume Methods<br />

Comparison of the four flux limiters on test case 1 is shown <strong>in</strong> Figure 5.8.<br />

All four flux limiter functions improve the phase prediction compare to the<br />

st<strong>and</strong>ard Lax-Wendroff numerical scheme. However, all functions smear the<br />

s<strong>in</strong>usoidal shape. Nevertheless, the Superbee flux limiter smear<strong>in</strong>g is negligible<br />

while both m<strong>in</strong>mod <strong>and</strong> Van-Leer functions smear<strong>in</strong>g are 25 % of the signal<br />

amplitude.<br />

Comparison of the four flux limiters on test case 2 is shown <strong>in</strong> Figure 5.9. All<br />

four flux limiter functions improve shock locations <strong>and</strong> reduce Lax-Wendroff<br />

numerical scheme oscillatory behaviour. Both m<strong>in</strong>mod <strong>and</strong> Van-Leer flux<br />

limiter functions underestimate slightly the square wave amplitude, while Superbee<br />

<strong>and</strong> monotonized-centered functions capture the wave amplitude accurately.<br />

Comparison of the four flux limiters on test case 3 is shown <strong>in</strong> Figure 5.10.<br />

Increas<strong>in</strong>g the number of grid po<strong>in</strong>ts improves prediction of shock locations<br />

compare to test case 2. The square wave amplitude is now also well-captured<br />

by all four flux limiter functions. However, <strong>in</strong>creas<strong>in</strong>g the number of grid<br />

po<strong>in</strong>ts <strong>in</strong>crease computational time <strong>and</strong> necessary resources.<br />

Comparison of the four flux limiters on test case 4 is shown <strong>in</strong> Figure 5.11.<br />

All four flux limiter functions improve shock <strong>and</strong> expansion fan locations,<br />

but most importantly, the wave shape is now captured without undershoot<strong>in</strong>g<br />

or overshoot<strong>in</strong>g close to sonic po<strong>in</strong>ts compare to the st<strong>and</strong>ard Lax-Wendroff<br />

method. All four flux limiter functions give superimposed solutions.<br />

Coupl<strong>in</strong>g a flux limiter function to the st<strong>and</strong>ard Lax-Wendroff numerical<br />

scheme improves shock <strong>and</strong> expansion fan locations at convex corners, <strong>and</strong> better<br />

captures wave shape. Increas<strong>in</strong>g the number of grid po<strong>in</strong>ts globally helps<br />

improv<strong>in</strong>g the accuracy of the solution. The Superbee flux limiter function<br />

performs globally better than the three other flux limiter functions studied,<br />

as seen <strong>in</strong> Figures 5.8 to 5.11.<br />

5.4.5 CLAWPACK<br />

CLAWPACK - Conservation LAW PACKage - is a software package of wave<br />

propagation methods for hyperbolic equations (LeVeque, 2002). Several numerical<br />

schemes are <strong>in</strong>cluded <strong>in</strong> form of subrout<strong>in</strong>es. Available numerical<br />

schemes <strong>in</strong>clude CTCS, Lax-Wendroff <strong>and</strong> the four flux limiter functions here<strong>in</strong><br />

studied. Of particular <strong>in</strong>terest, the Lax-Wendroff numerical scheme with the<br />

Superbee flux limiter was used <strong>in</strong> the rest of the thesis for pressure wave simulation.<br />

85


Chapter 5. Numerical Simulation Methods<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

−0.25<br />

−0.50<br />

−0.75<br />

−1.00<br />

Exact Solution<br />

Lax−Wendroff<br />

LW + M<strong>in</strong>mod<br />

LW + Superbee<br />

LW + MC<br />

LW + Van−Leer<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.8: Comparison of M<strong>in</strong>mod, Superbee, Monotonized<br />

Centered <strong>and</strong> Van-Leer Flux Limiters Coupled to the<br />

Lax-Wendroff Numerical Scheme on Test Case 1<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

Exact Solution<br />

Lax−Wendroff<br />

LW + M<strong>in</strong>mod<br />

LW + Superbee<br />

LW + MC<br />

LW + Van−Leer<br />

−0.25<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.9: Comparison of M<strong>in</strong>mod, Superbee, Monotonized<br />

Centered <strong>and</strong> Van-Leer Flux Limiters Coupled to the<br />

Lax-Wendroff Numerical Scheme on Test Case 2<br />

86


1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

5.4. F<strong>in</strong>ite Volume Methods<br />

Exact Solution<br />

Lax−Wendroff<br />

LW + M<strong>in</strong>mod<br />

LW + Superbee<br />

LW + MC<br />

LW + Van−Leer<br />

−0.25<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.10: Comparison of M<strong>in</strong>mod, Superbee, Monotonized<br />

Centered <strong>and</strong> Van-Leer Flux Limiters Coupled to the<br />

Lax-Wendroff Numerical Scheme on Test Case 3<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

Exact Solution<br />

Lax−Wendroff<br />

LW + M<strong>in</strong>mod<br />

LW + Superbee<br />

LW + MC<br />

LW + Van−Leer<br />

−0.25<br />

−1.00 −0.75 −0.50 −0.25 0.00 0.25 0.50 0.75 1.00<br />

Figure 5.11: Comparison of M<strong>in</strong>mod, Superbee, Monotonized<br />

Centered <strong>and</strong> Van-Leer Flux Limiters Coupled to the<br />

Lax-Wendroff Numerical Scheme on Test Case 4<br />

87


Chapter 5. Numerical Simulation Methods<br />

5.5 Numerical Scheme Selection<br />

Selection of a numerical scheme is performed us<strong>in</strong>g a st<strong>and</strong>ard test case composed<br />

of a 100 meter long pipel<strong>in</strong>e, connected to a 50 bara constant pressure<br />

tank on the left-h<strong>and</strong> side, <strong>and</strong> to a quick-act<strong>in</strong>g valve on the right side (Figure<br />

5.1). Water flows from the tank <strong>in</strong>to the pipel<strong>in</strong>e <strong>and</strong> flows at 2 ms −1<br />

constant velocity. The acoustic velocity <strong>in</strong> the fluid-filled pipel<strong>in</strong>e is fixed at<br />

1435 ms −1 .<br />

The test case is modelled us<strong>in</strong>g both the method of characteristics <strong>and</strong> the<br />

Lax-Wendroff scheme coupled to Superbee flux limiter. Frictional pressure<br />

drop <strong>and</strong> heat transfer are not considered. Another way to express this is that<br />

the source term Λ <strong>in</strong> Eq. 5.8 equals zero. The valve is suddenly actuated such<br />

that a rapid pressure transient is created at the valve <strong>and</strong> propagates through<br />

the pipel<strong>in</strong>e. 4 seconds are simulated for observ<strong>in</strong>g pressure waves reflections.<br />

The method of characteristics models the test case with 1000 space grid po<strong>in</strong>ts<br />

<strong>and</strong> 30873 space steps for ensur<strong>in</strong>g a Courant number equal to one for limit<strong>in</strong>g<br />

numerical dissipation. The Lax-Wendroff numerical scheme coupled to Superbee<br />

flux limiter <strong>in</strong>cluded <strong>in</strong> CLAWPACK, models the test case with 1000 grid<br />

po<strong>in</strong>ts <strong>and</strong> 1000 space steps. Each time step is iterated to ensure a Courant<br />

number equal to one.<br />

The modell<strong>in</strong>g results by the method of characteristics <strong>and</strong> the f<strong>in</strong>ite volume<br />

method are showed <strong>in</strong> Figure 5.12. The pressure trace obta<strong>in</strong>ed from both<br />

methods are identical, confirm<strong>in</strong>g that both method are equivalent for sourceless<br />

pressure transient modell<strong>in</strong>g. Therefore, only the Lax-Wendroff numerical<br />

scheme coupled to Superbee flux limiter is used <strong>in</strong> the follow<strong>in</strong>g of the thesis.<br />

88


<strong>Pressure</strong> [bara]<br />

100<br />

75<br />

50<br />

25<br />

5.5. Numerical Scheme Selection<br />

MOC<br />

FVM<br />

0<br />

0.0 0.5 1.0 1.5 2.0<br />

Time [s]<br />

2.5 3.0 3.5 4.0<br />

Figure 5.12: Comparison of Method of Characteristics <strong>and</strong> F<strong>in</strong>ite<br />

Volume Method (Lax-Wendroff + Superbee Flux<br />

Limiter) on St<strong>and</strong>ard <strong>Pressure</strong> Transient Test Case<br />

89


Chapter 6<br />

<strong>Pressure</strong> Wave Attenuation<br />

6.1 Influence of Numerical Simulation Tool<br />

6.1.1 Unsteady Friction Modell<strong>in</strong>g Equations<br />

<strong>Pressure</strong> wave modell<strong>in</strong>g is based on mass <strong>and</strong> momentum conservation equations<br />

where the source term is responsible for smooth<strong>in</strong>g <strong>and</strong> damp<strong>in</strong>g of pressure<br />

waves. Steady state frictional pressure drop was the first attenuation<br />

mechanism <strong>in</strong>cluded, <strong>and</strong> state of the art pressure wave modell<strong>in</strong>g now <strong>in</strong>cludes<br />

unsteady friction factors <strong>in</strong> addition. Various unsteady friction models<br />

have been proposed s<strong>in</strong>ce the first model by Zielke (1966). Of particular <strong>in</strong>terest<br />

are unsteady friction models proposed by Brunone et al. (1995) <strong>and</strong> Ramos<br />

et al. (2004).<br />

Brunone et al.’s <strong>and</strong> Ramos et al.’s unsteady friction models decompose the<br />

friction factor <strong>in</strong> two parts. The first part is the quasi steady friction factor, fq,<br />

<strong>and</strong> is commonly calculated us<strong>in</strong>g Darcy-Weisbach friction factor formulation,<br />

for <strong>in</strong>stance. The second part takes <strong>in</strong>to account local <strong>and</strong> <strong>in</strong>stantaneous<br />

convective acceleration of flow velocities. Brunone et al. (1995) <strong>and</strong> Ramos<br />

et al. (2004) unsteady friction models are, respectively:<br />

f = fq + kB1d<br />

� �<br />

∂u<br />

− a∂u<br />

(6.1)<br />

u|u| ∂t ∂x<br />

f = fq + d<br />

u|u|<br />

�<br />

kR1<br />

∂u<br />

∂t<br />

u<br />

+ kR2<br />

|u| a<br />

� ��<br />

�<br />

�<br />

∂u�<br />

�<br />

�∂x<br />

�<br />

(6.2)<br />

where, kB1, kR1 <strong>and</strong> kR2, are empirical constants that can be tuned for match<strong>in</strong>g<br />

experimental data with numerical simulations. St<strong>and</strong>ard values are 0.02,<br />

0.003 <strong>and</strong> 0.04, for kB1, kR1 <strong>and</strong> kR2, respectively. Simulations of water hammer<br />

events <strong>in</strong>clud<strong>in</strong>g unsteady friction modell<strong>in</strong>g were performed by Brunone<br />

et al. <strong>and</strong> Ramos et al. us<strong>in</strong>g the method of characteristics.<br />

91


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

6.1.2 Test Case <strong>and</strong> Numerical Simulation<br />

A water hammer test case was simulated us<strong>in</strong>g the method of characteristics<br />

<strong>and</strong> the previously selected f<strong>in</strong>ite volume scheme. The method of characteristics<br />

is <strong>in</strong>cluded <strong>in</strong> a commercial software, Bentley Hammer V 8i, (Haestad<br />

Methods Solution Center, 2008). Whereas the f<strong>in</strong>ite volume method is <strong>in</strong>cluded<br />

<strong>in</strong> an <strong>in</strong> house numerical simulation tool for pressure transient simulations<br />

<strong>in</strong> s<strong>in</strong>gle pipel<strong>in</strong>es. The f<strong>in</strong>ite volume method is based on the Lax-<br />

Wendroff numerical scheme coupled with Superbee flux limiter function (Chapter<br />

5).<br />

The water hammer test case <strong>in</strong>cludes a steel horizontal pipel<strong>in</strong>e, 1000 m <strong>in</strong><br />

length, 10 <strong>in</strong>ches <strong>in</strong> diameter, <strong>and</strong> liquid-water-filled. The fluid density is<br />

1000 kg m −3 , <strong>and</strong> the acoustic velocity <strong>in</strong> the fluid-filled pipel<strong>in</strong>e is set at 1200<br />

ms −1 . The pipel<strong>in</strong>e is connected to a 50 bara constant pressure reservoir on its<br />

left-h<strong>and</strong> side, <strong>and</strong> to a quick act<strong>in</strong>g check valve on its right-h<strong>and</strong> side. Water<br />

was <strong>in</strong>itially flow<strong>in</strong>g at 2 ms −1 before valve actuation, <strong>and</strong> the steady friction<br />

factor was set to 0.01. The valve actuation is assumed ideal <strong>and</strong> <strong>in</strong>stantaneous.<br />

The water hammer test case is first simulated us<strong>in</strong>g the method of characteristics<br />

<strong>in</strong>cluded <strong>in</strong> Bentley Hammer V 8i commercial software. The Courant<br />

number, a dt/dx, was set to 1 to limit the method of characteristis <strong>in</strong>herent<br />

numerical dissipation. The water hammer test case is simulated us<strong>in</strong>g<br />

steady state friction factor only, <strong>and</strong> then <strong>in</strong>clud<strong>in</strong>g Brunone et al.’s unsteady<br />

friction factor. Ramos et al.’s unsteady friction factor is not <strong>in</strong>cluded <strong>in</strong><br />

Bentley Hammer V 8i commercial software.<br />

The water hammer test case was then simulated us<strong>in</strong>g the selected f<strong>in</strong>ite volume<br />

method. The Courant number was also set to 1 <strong>and</strong> other simulation<br />

parameters from Bentley Hammer V 8i commercial software were used as set<br />

parameters <strong>in</strong> the more flexible <strong>in</strong> house f<strong>in</strong>ite volume method numerical code.<br />

Of particular <strong>in</strong>terest, 15 spatial nodes were used. Comparison between the<br />

two numerical methods is then performed based on numerical scheme only.<br />

Complete test case description <strong>and</strong> numerical simulation parameters are summarized<br />

<strong>in</strong> Table 6.1.<br />

Table 6.1: Test Case System Description <strong>and</strong> Numerical Simulation<br />

Parameters<br />

L [m] d [m] p0 [bara] u0 [ms −1 ] a [ms −1 ] nX nT TSim [s]<br />

100 0.254 50 2 1200 15 540 30<br />

92


6.1.3 Results of Friction Modell<strong>in</strong>g<br />

6.1. Influence of Numerical Simulation Tool<br />

Simulation results us<strong>in</strong>g the method of characteristics <strong>and</strong> the selected f<strong>in</strong>ite<br />

volume method, <strong>in</strong>clud<strong>in</strong>g steady friction only, are shown <strong>in</strong> Figure 6.1.<br />

Both numerical method predicts correctly the first pressure transient amplitude<br />

accord<strong>in</strong>g to the Joukowski equation. However, predicted pressure trace<br />

is square-shaped with MOC <strong>and</strong> round-shaped with FVM. Predicted pressure<br />

amplitudes also decrease faster with FVM than with MOC. In addition, pressure<br />

wave propagation velocity is accurately with MOC but under-estimated<br />

with FVM.<br />

Simulation results us<strong>in</strong>g the method of characteristics <strong>and</strong> the selected f<strong>in</strong>ite<br />

volume method, <strong>in</strong>clud<strong>in</strong>g Brunone et al.’s unsteady friction factor, are shown<br />

<strong>in</strong> Figure 6.2. Both numerical methods predict correctly the first pressure<br />

transient amplitude accord<strong>in</strong>g to the Joukowski equation. However, predicted<br />

pressure traces are square-shaped with MOC <strong>and</strong> round-shaped with FVM.<br />

<strong>Pressure</strong> amplitudes are first lower with MOC until 15 s, then pressure amplitudes<br />

decrease faster with FVM. <strong>Pressure</strong> wave propagation velocity is aga<strong>in</strong><br />

accurately predicted with MOC but under-estimated with FVM.<br />

Simulation results us<strong>in</strong>g the method of characteristics, <strong>and</strong> <strong>in</strong>clud<strong>in</strong>g steady<br />

friction factor, <strong>and</strong> Brunone et al.’s unsteady friction factor, are repeated <strong>in</strong><br />

Figure 6.3. As previously observed, pressure traces are square-shaped <strong>and</strong><br />

pressure wave propagation velocities are accurately predicted. Both models<br />

predict the same first pressure wave amplitude accord<strong>in</strong>g to the Joukowski’s<br />

equation. However, pressure amplitudes decrease faster with Brunone et al.’s<br />

unsteady friction modell<strong>in</strong>g as reported <strong>in</strong> the literature (Brunone et al., 1995).<br />

Simulation results us<strong>in</strong>g the selected f<strong>in</strong>ite volume method, <strong>and</strong> <strong>in</strong>clud<strong>in</strong>g<br />

steady friction factor, <strong>and</strong> Brunone et al.’s unsteady friction factor, are repeated<br />

<strong>in</strong> Figure 6.4. As previously observed, pressure traces are roundshaped<br />

<strong>and</strong> pressure wave propagation velocity are under-estimated. Both<br />

models predict the same pressure trace, that is shape, travel time, damp<strong>in</strong>g<br />

<strong>and</strong> smooth<strong>in</strong>g. Another way to express this is that Brunone et al.’s unsteady<br />

friction model does not improve water hammer calculation when simulated<br />

with FVM.<br />

Simulation results from the method of characteristics show that unsteady friction<br />

<strong>in</strong>creases pressure damp<strong>in</strong>g as claimed by Brunone et al. (1995). However,<br />

simulation results from the f<strong>in</strong>ite volume method show that unsteady friction<br />

does not <strong>in</strong>crease pressure damp<strong>in</strong>g compare to steady state friction. But,<br />

simulation results from the f<strong>in</strong>ite volume method show that the pressure wave<br />

propagation velocity is under-estimated, thus <strong>in</strong>dicat<strong>in</strong>g poor numerical resolution<br />

<strong>and</strong> mesh<strong>in</strong>g for the f<strong>in</strong>ite volume method. F<strong>in</strong>ite volume mesh<strong>in</strong>g can<br />

however be adjusted.<br />

93


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

<strong>Pressure</strong> [bara]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

MOC Steady<br />

FVM Steady<br />

20<br />

0 5 10 15<br />

Time [s]<br />

20 25 30<br />

Figure 6.1: Comparison of Method of Characteristics <strong>and</strong> F<strong>in</strong>ite<br />

Volume Method with Steady Friction Only. nX = 15.<br />

nT = 540.<br />

<strong>Pressure</strong> [bara]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

MOC Unsteady<br />

FVM Unsteady<br />

20<br />

0 5 10 15<br />

Time [s]<br />

20 25 30<br />

Figure 6.2: Comparison of Method of Characteristics <strong>and</strong> F<strong>in</strong>ite<br />

Volume Method with Brunone’s Unsteady Friction<br />

Model. nX = 15. nT = 540.<br />

94


<strong>Pressure</strong> [bara]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

6.1. Influence of Numerical Simulation Tool<br />

MOC Steady<br />

MOC Unsteady<br />

20<br />

0 5 10 15<br />

Time [s]<br />

20 25 30<br />

Figure 6.3: Comparison of Steady Friction <strong>and</strong> Brunone’s Unsteady<br />

Friction Models with the Method of Characteristics.<br />

nX = 15. nT = 540.<br />

<strong>Pressure</strong> [bara]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

FVM Steady<br />

FVM Unsteady<br />

20<br />

0 5 10 15<br />

Time [s]<br />

20 25 30<br />

Figure 6.4: Comparison of Steady Friction <strong>and</strong> Brunone’s Unsteady<br />

Friction Models with the Selected F<strong>in</strong>ite Volume Method.<br />

nX = 15. nT = 540.<br />

95


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

6.1.4 F<strong>in</strong>ite Volume Method Accuracy<br />

The water hammer test case can be simulated us<strong>in</strong>g the selected f<strong>in</strong>ite volume<br />

method <strong>and</strong> <strong>in</strong>creas<strong>in</strong>g numerical resolution <strong>and</strong> spatial mesh<strong>in</strong>g. Simulation<br />

results us<strong>in</strong>g the f<strong>in</strong>ite volume method <strong>and</strong> steady state friction factor are<br />

shown for <strong>in</strong>creas<strong>in</strong>g number of spatial nodes <strong>in</strong> Figure 6.5. The pressure<br />

trace becomes sharper with <strong>in</strong>creas<strong>in</strong>g number of spatial nodes, <strong>and</strong> propagation<br />

velocity are correctly captured for nX = 1000, therefore <strong>in</strong>creas<strong>in</strong>g the<br />

simulation accuracy. Another way to express this is that the selected f<strong>in</strong>ite<br />

volume method accuracy <strong>in</strong>creases with the number of spatial nodes.<br />

<strong>Pressure</strong> [bara]<br />

70<br />

68<br />

66<br />

64<br />

62<br />

60<br />

58<br />

n X = 1000<br />

n X = 200<br />

n X = 100<br />

n X = 50<br />

n X = 15<br />

56<br />

19.5 20.0 20.5 21.0<br />

Time [s]<br />

21.5 22.0 22.5<br />

Figure 6.5: Influence of Spatial Discretization on Water Hammer<br />

Calculation with the Selected F<strong>in</strong>ite Volume Method.<br />

Detail of <strong>Pressure</strong> Trace<br />

Simulation results us<strong>in</strong>g the selected f<strong>in</strong>ite volume method, <strong>and</strong> <strong>in</strong>clud<strong>in</strong>g<br />

steady friction factor, Brunone et al.’s <strong>and</strong> Ramos et al.’s unsteady friction<br />

factors, with nX = 1000, are shown <strong>in</strong> Figure 6.6. The three simulated pressure<br />

traces are merged, thus <strong>in</strong>dicat<strong>in</strong>g that unsteady friction models do not<br />

<strong>in</strong>crease pressure wave damp<strong>in</strong>g. In simple words, unsteady friction models<br />

do not improve water hammer prediction compare to steady state friction. In<br />

addition, no pressure trace smooth<strong>in</strong>g is observed.<br />

Simulation results us<strong>in</strong>g the method of characteristics, <strong>in</strong>clud<strong>in</strong>g steady state<br />

<strong>and</strong> Brunone et al.’s unsteady friction factors, <strong>and</strong> us<strong>in</strong>g the selected f<strong>in</strong>ite<br />

volume method are repeated <strong>in</strong> Figure 6.7. Both numerical schemes predict<br />

96


6.1. Influence of Numerical Simulation Tool<br />

accurately the first pressure transient amplitude <strong>and</strong> pressure wave propagation<br />

velocity. However, the method of characteristics predicts faster damp<strong>in</strong>g<br />

than the f<strong>in</strong>ite volume method, thus illustrat<strong>in</strong>g that the method of characteristics<br />

<strong>in</strong>herent dissipation is non-physical.<br />

Unsteady friction factor terms <strong>in</strong>crease convective acceleration. However, convective<br />

terms are physically negligible <strong>in</strong> accordance with the low Mach number<br />

approximation. Unsteady friction factors proposed by Brunone et al. (1995)<br />

<strong>and</strong> Ramos et al. (2004) are therefore artifacts that <strong>in</strong>crease numerical dissipation<br />

<strong>and</strong> that by co<strong>in</strong>cidence mimics pressure wave attenuation mechanisms.<br />

In addition, unsteady friction factor models do not change the modell<strong>in</strong>g equations<br />

hyperbolic nature, <strong>and</strong> therefore can not reproduce pressure trace damp<strong>in</strong>g.<br />

Such result is <strong>in</strong> accordance with Szymkiewicz <strong>and</strong> Mitosek (2007).<br />

The selected f<strong>in</strong>ite volume method is therefore more accurate <strong>and</strong> appropriate<br />

than the method of characteristics for simulat<strong>in</strong>g pressure wave transients,<br />

provided that spatial discretization is sufficient enough. In addition, water<br />

hammer simulation us<strong>in</strong>g a f<strong>in</strong>ite volume method <strong>in</strong>dicates that current one<br />

dimensional pressure wave attenuation mechanisms are <strong>in</strong>accurate <strong>and</strong> do not<br />

capture the physics of the phenomenon. New models need therefore to be<br />

<strong>in</strong>vestigated, modelled, <strong>and</strong> simulated.<br />

6.1.5 Pipel<strong>in</strong>e Viscoelasticity<br />

Pipel<strong>in</strong>e material viscoelastic behaviour has received much attention <strong>in</strong> the<br />

past two decades due to the <strong>in</strong>creas<strong>in</strong>g use of plastic pipel<strong>in</strong>es. Pipel<strong>in</strong>e viscoelasticity<br />

models relate to pipel<strong>in</strong>e material deformation properties dur<strong>in</strong>g<br />

pressure transient events. Pipel<strong>in</strong>e materials Young’s moduli change with time<br />

dur<strong>in</strong>g pressure transient event. Pipel<strong>in</strong>e materials first have an <strong>in</strong>stantaneous<br />

response to a pressure transient, then a delayed response that depends on material<br />

characteristics <strong>and</strong> fatigue (Covas, 2003).<br />

Viscoelasticity modell<strong>in</strong>g uses pipel<strong>in</strong>e decomposition <strong>in</strong> Kelv<strong>in</strong>-Voigt elements<br />

that can be represented by a purely viscous damper <strong>and</strong> a purely elastic<br />

str<strong>in</strong>g. The number of elements, as well as retardation time <strong>and</strong> tensile moduli<br />

of each element are modell<strong>in</strong>g parameters. Numerical schemes <strong>in</strong>herent<br />

dissipation <strong>and</strong> characteristics need therefore to be taken <strong>in</strong>to consideration<br />

(We<strong>in</strong>erowska-Bords, 2007).<br />

Pipel<strong>in</strong>e viscoelasticity modell<strong>in</strong>g has been simulated us<strong>in</strong>g the method of<br />

characteristics. Hence, the method of characteristics <strong>in</strong>herent numerical dissipation<br />

also <strong>in</strong>duces greater pressure wave damp<strong>in</strong>g than <strong>in</strong> practice. Thus,<br />

pipel<strong>in</strong>e viscoelasticity modell<strong>in</strong>g results <strong>and</strong> accuracy are questionable. In<br />

addition, pipel<strong>in</strong>e viscoelasticity is negligible <strong>in</strong> steel pipel<strong>in</strong>es, thus pressure<br />

wave attenuation <strong>in</strong> steel pipel<strong>in</strong>es is still <strong>in</strong>accurately estimated.<br />

97


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

<strong>Pressure</strong> [bara]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

Steady Friction<br />

Brunone<br />

Ramos<br />

20<br />

0 5 10 15<br />

Time [s]<br />

20 25 30<br />

Figure 6.6: Comparison of Investigated Unsteady Friction Models<br />

Us<strong>in</strong>g a Selected F<strong>in</strong>ite Volume Method. nX = 1000.<br />

nT = 540.<br />

<strong>Pressure</strong> [bara]<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

MOC Steady<br />

MOC Unsteady<br />

FVM Unsteady<br />

20<br />

0 5 10 15<br />

Time [s]<br />

20 25 30<br />

Figure 6.7: Comparison of Water Hammer Simulation Us<strong>in</strong>g the<br />

Method of Characteristics <strong>and</strong> a F<strong>in</strong>ite Volume Method.<br />

MOC: nX = 15. nT = 540. FVM: nX = 1000. nT = 540.<br />

98


6.2 Acoustic Velocity<br />

6.2. Acoustic Velocity<br />

Current pressure wave attenuation mechanisms focus on the modell<strong>in</strong>g equations<br />

source term, <strong>and</strong> viscoelasticity models focus on <strong>in</strong>clud<strong>in</strong>g pipel<strong>in</strong>e elastic<br />

properties <strong>in</strong> acoustic velocity calculations. Current pressure wave propagation<br />

velocity models <strong>in</strong>clude fluids <strong>and</strong> pipel<strong>in</strong>e properties. Acoustic velocity<br />

calculations then require fluid density <strong>and</strong> isothermal compressibility, <strong>and</strong><br />

pipel<strong>in</strong>e compressibility that depends on pipel<strong>in</strong>e dimensions, materials <strong>and</strong><br />

fasten<strong>in</strong>gs. The acoustic velocity of a pressure wave <strong>in</strong> fluid-filled pipel<strong>in</strong>e<br />

therefore be expressed by:<br />

a =<br />

�<br />

1<br />

ρ (KFluid + KPipe)<br />

(6.3)<br />

where, ρ, is fluid density, KFluid, fluid isothermal compressibility, <strong>and</strong>, KPipe,<br />

pipel<strong>in</strong>e compressibility. The <strong>in</strong>fluence of every component of the acoustic<br />

velocity calculation have been detailed <strong>in</strong> Chapter 4. It is also noted that<br />

the state of the fluid is accounted by density <strong>and</strong> isothermal compressibility<br />

changes. However, flow structures, such as turbulent eddies, are not considered<br />

by current calculation models.<br />

6.2.1 Ultrasonic Measurement<br />

<strong>Pressure</strong> wave propagation velocity <strong>in</strong> gas flowl<strong>in</strong>es can be obta<strong>in</strong>ed experimentally<br />

us<strong>in</strong>g ultrasonic flow meters. A schematic representation of an ultrasonic<br />

flow meter is shown <strong>in</strong> Figure 6.8. An emitter generates a sonic beam that<br />

propagates to a receiver. Flow velocity <strong>and</strong> acoustic velocity are then calculated<br />

from pressure wave travel time between emitter <strong>and</strong> receiver. Ultrasonic<br />

measurements are performed with great accuracy <strong>and</strong> therefore provide valuable<br />

data.<br />

Three different ultrasonic measurement data sets from one 24 <strong>in</strong>ches FMC<br />

apparatus, class 600, located on a natural gas transport pipel<strong>in</strong>e at Field A,<br />

were made available to <strong>NTNU</strong> by Jacobsen (2008). Gas composition from field<br />

A is given <strong>in</strong> Table 6.2, <strong>and</strong> the gas phase envelope is shown <strong>in</strong> Figure 6.9.<br />

Data sets <strong>in</strong>clude pressure <strong>and</strong> temperature conditions, <strong>and</strong> measured flow<br />

<strong>and</strong> acoustic velocities. Gas properties are calculated from gas composition,<br />

pressure <strong>and</strong> temperature conditions us<strong>in</strong>g the Peng-Rob<strong>in</strong>son fluid package<br />

<strong>in</strong>cluded <strong>in</strong> Hysys.<br />

99


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Emitter<br />

δ<br />

Receiver<br />

Figure 6.8: Schematic Representation of an Ultrasonic Flow Meter<br />

Table 6.2: Field A Molar Gas Composition (Jacobsen, 2008)<br />

<strong>Pressure</strong> [bara]<br />

Component C1 C2 C3 iC4 nC4<br />

% 85.823 8.271 2.335 0.233 0.255<br />

Component iC5 nC5 C6 CO2 N2<br />

% 0.029 0.029 0.038 1.825 1.150<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Critical Po<strong>in</strong>t<br />

Bubble Po<strong>in</strong>t<br />

Dew Po<strong>in</strong>t<br />

0<br />

−160 −140 −120 −100 −80 −60 −40 −20<br />

Temperature [ o C]<br />

Figure 6.9: Field A Gas Phase Envelope (Jacobsen, 2008)<br />

100


6.2. Acoustic Velocity<br />

At experimental pressure <strong>and</strong> temperature conditions, averaged density <strong>and</strong><br />

dynamic viscosity are 55.6 kg m −3 <strong>and</strong> 1.24 10 −2 cP, respectively. Theoretically<br />

calculated values of density, adiabatic <strong>in</strong>dex <strong>and</strong> z-factor are then used<br />

to calculate acoustic velocity <strong>in</strong> the gas us<strong>in</strong>g:<br />

�<br />

�<br />

� k<br />

a = � � � �<br />

� 1 1 ∂z<br />

ρ −<br />

p z ∂p<br />

Γ<br />

� (6.4)<br />

Evolutions of measured <strong>and</strong> calculated acoustic velocities with pressure <strong>and</strong><br />

temperature are shown <strong>in</strong> Figure 6.10. Calculated values are about 2 ms −1<br />

lower <strong>in</strong> average than measured values. Curve shapes are however similar, with<br />

approximately constant acoustic velocity for decreas<strong>in</strong>g pressures at constant<br />

temperature, <strong>and</strong> for decreas<strong>in</strong>g temperatures at constant pressure. Differences<br />

between measured <strong>and</strong> calculated values can be expla<strong>in</strong>ed by the use of<br />

the Peng-Rob<strong>in</strong>son equation of state for predict<strong>in</strong>g gas properties.<br />

Evolution of measured <strong>and</strong> calculated acoustic velocities with pressure <strong>and</strong><br />

temperature are also shown separately <strong>in</strong> Figure 6.11 <strong>and</strong> Figure 6.12, respectively.<br />

Measured acoustic velocity decreases from 374 ms −1 at 57.98 bara<br />

to 373.7 ms −1 at 58.12 bara. Whereas calculated acoustic velocity oscillates<br />

slightly around 372 ms −1 for the range of measured pressures. Measured<br />

acoustic velocity <strong>in</strong>creases from 374 ms −1 at 8.87 o C, to 374.5 ms −1 at 9.07 o C.<br />

Whereas calculated values <strong>in</strong>crease from 371.8 ms −1 to 372.3 ms −1 for the<br />

range of measured temperatures.<br />

Measured <strong>and</strong> calculated acoustic velocities evolution with gas flow velocity<br />

is shown <strong>in</strong> Figure 6.13. Calculated values are slightly oscillat<strong>in</strong>g around 372<br />

ms −1 , whereas measured acoustic velocities <strong>in</strong>crease from 373.7 ms −1 at 1<br />

ms −1 to 374.5 ms −1 at 20 ms −1 . Acoustic velocity dependence on flow<br />

velocity is observed experimentally but unexpected from a modell<strong>in</strong>g po<strong>in</strong>t of<br />

view, <strong>and</strong> is not taken <strong>in</strong>to consideration by current models (AGA-10, 2003).<br />

Two other data set were analysed similarly. Both data sets are similar <strong>and</strong><br />

<strong>in</strong> the same range of pressure <strong>and</strong> temperature, therefore only one data set is<br />

here<strong>in</strong> presented. Measured <strong>and</strong> calculated acoustic velocities evolution with<br />

pressure <strong>and</strong> temperature is shown <strong>in</strong> Figure 6.14. Calculated values are 2<br />

ms −1 lower than measured values, <strong>in</strong> average. The large acoustic velocity<br />

drop at 4.5 o C is however not predicted. Curve shapes are however globally<br />

similar. Differences between measured <strong>and</strong> calculated values can aga<strong>in</strong> be<br />

expla<strong>in</strong>ed by the use of the Peng-Rob<strong>in</strong>son equation of state for predict<strong>in</strong>g gas<br />

properties.<br />

101


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Acoustic Velocity [m s −1 ]<br />

373.5<br />

373.0<br />

372.5<br />

372.0<br />

371.5<br />

9.00<br />

8.95<br />

8.90<br />

Temperature [ o C]<br />

8.85<br />

8.80<br />

57.95<br />

58.00<br />

58.05<br />

Measured<br />

Calculated<br />

58.10<br />

<strong>Pressure</strong> [bara]<br />

58.15<br />

Figure 6.10: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of <strong>Pressure</strong> <strong>and</strong> Temperature. First Data Set<br />

Acoustic Velocity [m s −1 ]<br />

375.0<br />

374.5<br />

374.0<br />

373.5<br />

373.0<br />

372.5<br />

372.0<br />

Measured<br />

Calculated<br />

371.5<br />

57.95 58.00 58.05<br />

<strong>Pressure</strong> [bara]<br />

58.10 58.15<br />

Figure 6.11: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of <strong>Pressure</strong>. First Data Set<br />

102


Acoustic Velocity [m s −1 ]<br />

375.0<br />

374.5<br />

374.0<br />

373.5<br />

373.0<br />

372.5<br />

6.2. Acoustic Velocity<br />

372.0<br />

371.5<br />

8.80 8.85 8.90 8.95<br />

Temperature [<br />

9.00 9.05 9.10<br />

o Measured<br />

Calculated<br />

C]<br />

Figure 6.12: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of Temperature. First Data Set<br />

Acoustic Velocity [m s −1 ]<br />

375.0<br />

374.5<br />

374.0<br />

373.5<br />

373.0<br />

372.5<br />

372.0<br />

Measured<br />

Calculated<br />

371.5<br />

0 5 10 15 20 25<br />

Flow Velocity [m s −1 ]<br />

Figure 6.13: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of Flow Velocity. First Data Set<br />

103


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Measured <strong>and</strong> calculated acoustic velocities evolution with pressure <strong>and</strong> temperature<br />

are also shown separately <strong>in</strong> Figure 6.15 <strong>and</strong> Figure 6.16, respectively.<br />

Measured <strong>and</strong> calculated acoustic velocities decrease l<strong>in</strong>early between<br />

50.5 bara <strong>and</strong> 56 bara from 373 ms −1 to 371 ms −1 , <strong>and</strong> from 371 ms −1 to<br />

369 ms −1 , respectively. Measured <strong>and</strong> calculated acoustic velocities <strong>in</strong>crease<br />

l<strong>in</strong>early between 3.9 o C <strong>and</strong> 5.3 o C from 369 ms −1 to 371 ms −1 , <strong>and</strong> from<br />

367 ms −1 to 369 ms −1 , respectively. In both evolutions, the large acoustic<br />

velocity drop observed is not predicted by calculations.<br />

Measured <strong>and</strong> calculated acoustic velocities evolution with gas flow velocity<br />

is shown <strong>in</strong> Figure 6.17. Both measured <strong>and</strong> calculated acoustic velocities<br />

decrease with flow velocity between 3 ms −1 <strong>and</strong> 21 ms −1 , from 372 ms −1 to<br />

369 ms −1 , <strong>and</strong> from 369 ms −1 to 367 ms −1 , respectively. Measured acoustic<br />

velocity is therefore not dependent upon gas flow velocity as opposed to the<br />

first data set. Measured acoustic velocity from the third data set is also<br />

<strong>in</strong>dependent of gas flow velocity.<br />

Measured acoustic velocity changed with gas flow velocity <strong>in</strong> the three ultrasonic<br />

measurement sets provided. The acoustic velocity evolution could not<br />

be predicted by theoretical calculations <strong>and</strong> pressure <strong>and</strong> temperature adjustments<br />

<strong>in</strong> the first case, whereas acoustic velocity evolution could be calculated<br />

theoretically <strong>in</strong> the two other cases. However, it is noticeable that temperature<br />

is about 8 o C <strong>in</strong> the first case <strong>and</strong> only 4 o C <strong>in</strong> the two other cases. <strong>Pressure</strong><br />

values are <strong>in</strong> the same range for all three cases.<br />

Decrease of acoustic velocity with flow velocity at low temperature can be<br />

expla<strong>in</strong>ed by the presence of liquid droplets. Liquid droplets can form <strong>in</strong> gasfilled<br />

pipel<strong>in</strong>e at low temperature, <strong>and</strong> stick to the pipel<strong>in</strong>e wall. Increase <strong>in</strong><br />

gas flow velocity can then overcome wall shear stress for droplets <strong>and</strong> acoustic<br />

velocity is then measured <strong>in</strong> a two phase fluid. Small amount of droplets can<br />

decrease dramatically acoustic velocity <strong>in</strong> gas-filled pipel<strong>in</strong>es. Thus expla<strong>in</strong><strong>in</strong>g<br />

the decrease <strong>in</strong> acoustic velocity with gas flow velocity.<br />

On the other h<strong>and</strong>, <strong>in</strong>crease <strong>in</strong> acoustic velocity is unexpected <strong>and</strong> has not been<br />

observed previously. Increase <strong>in</strong> velocity can also be observed as contrary to<br />

what could be expected s<strong>in</strong>ce the system could have gradually emptied for<br />

adhered liquid dur<strong>in</strong>g the test<strong>in</strong>g. Thus, the observed phenomenon calls for<br />

confirmation <strong>and</strong> additional experimental work.<br />

104


Acoustic Velocity [m s −1 ]<br />

374<br />

372<br />

370<br />

368<br />

366<br />

364<br />

362<br />

5.5<br />

5.0<br />

4.5<br />

Temperature [ o C]<br />

4.0<br />

3.5<br />

50<br />

52<br />

6.2. Acoustic Velocity<br />

54<br />

Measured<br />

Calculated<br />

56<br />

<strong>Pressure</strong> [bara]<br />

Figure 6.14: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of <strong>Pressure</strong> <strong>and</strong> Temperature. Second Data Set<br />

Acoustic Velocity [m s −1 ]<br />

374<br />

372<br />

370<br />

368<br />

366<br />

364<br />

Measured<br />

Calculated<br />

362<br />

50 51 52 53 54 55 56 57<br />

<strong>Pressure</strong> [bara]<br />

Figure 6.15: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of <strong>Pressure</strong>. Second Data Set<br />

105<br />

58


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Acoustic Velocity [m s −1 ]<br />

374<br />

372<br />

370<br />

368<br />

366<br />

364<br />

362<br />

3.8 4.0 4.2 4.4 4.6 4.8 5.0 5.2 5.4<br />

Temperature [ o Measured<br />

Calculated<br />

C]<br />

Figure 6.16: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of Temperature. Second Data Set<br />

Acoustic Velocity [m s −1 ]<br />

374<br />

372<br />

370<br />

368<br />

366<br />

364<br />

Measured<br />

Calculated<br />

362<br />

0 5 10 15 20 25<br />

Flow Velocity [m s −1 ]<br />

Figure 6.17: Measured <strong>and</strong> Calculated Acoustic Velocities Function<br />

of Flow Velocity. Second Data Set<br />

106


6.2.2 Experimental Work<br />

6.2. Acoustic Velocity<br />

A small scale experiment has been set-up at <strong>NTNU</strong> to observe the hypothesis<br />

validity of acoustic velocity be<strong>in</strong>g dependent on flow velocity. A vertical plastic<br />

tube, 60 cm <strong>in</strong> depth, <strong>and</strong> 3.9 cm <strong>in</strong> diameter, is filled with tap water at<br />

atmospheric pressure <strong>and</strong> room temperature. A small diameter propeller is<br />

located at the surface to rotate the fluid <strong>in</strong> the water column. Eight emittertransmitter<br />

transducers are located equidistantly along the water column. A<br />

schematic representation of the apparatus is shown <strong>in</strong> Figure 6.18.<br />

600 mm<br />

1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

39 mm<br />

Figure 6.18: Schematic Representation of Experimental Apparatus<br />

for Measur<strong>in</strong>g Acoustic Velocity <strong>in</strong> Fluid Under<br />

Rotation<br />

107<br />

30 mm<br />

60 mm<br />

95 mm


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Transducers measured sound beam travel times <strong>in</strong> the fluid with a 500 kHz<br />

sampl<strong>in</strong>g frequency. The first transducer is located 95 mm from the surface,<br />

then transducers are equidistant, 60 mm apart from each others. The bottom<br />

of the propeller is located 30 mm from the surface. The experimental tube<br />

is flexible with an average diameter of 39 mm. However, distances between<br />

emitters <strong>and</strong> receivers are not accurately determ<strong>in</strong>ed <strong>and</strong> can evolve, therefore<br />

results are presented <strong>in</strong> form of travel time <strong>and</strong> are relative to each transducer.<br />

The propeller rotat<strong>in</strong>g velocity <strong>and</strong> direction are controlled by adjust<strong>in</strong>g the<br />

supplied voltage. The rotat<strong>in</strong>g speed is limited by vibrations of the propeller<br />

eng<strong>in</strong>e <strong>and</strong> its support. Maximum velocity is therefore reached for 4.5 V .<br />

The rotat<strong>in</strong>g speed <strong>in</strong>creases l<strong>in</strong>early with voltage, <strong>and</strong> equals 1280 rpm at 5<br />

V , without charge. The propeller radius is 19 mm. Voltage, correspond<strong>in</strong>g<br />

rotat<strong>in</strong>g velocities <strong>and</strong> l<strong>in</strong>ear velocities are listed <strong>in</strong> Table 6.3. Rotat<strong>in</strong>g speed<br />

dur<strong>in</strong>g experiment are slightly lower due to the resistant torque <strong>in</strong>duced by<br />

water.<br />

Table 6.3: Propeller Eng<strong>in</strong>e Voltage <strong>and</strong> Correspond<strong>in</strong>g Propeller<br />

Rotat<strong>in</strong>g Speed <strong>and</strong> L<strong>in</strong>ear Speed<br />

Voltage [V ] 0 2.5 3.5 4.5<br />

Rotat<strong>in</strong>g Speed [rpm] 0 640 896 1152<br />

L<strong>in</strong>ear Speed [ms −1 ] 0 1.27 1.78 2.29<br />

<strong>Pressure</strong> traces from transducers 1 for four different propeller eng<strong>in</strong>e voltages<br />

are shown <strong>in</strong> Figure 6.19. Travel times are measured for 0 V amplitude. Travel<br />

times are similar <strong>and</strong> equal 31.6 µs for 0 V <strong>and</strong> 2.5 V propeller eng<strong>in</strong>e voltages.<br />

Then, travel times equal 31.4 µs for both 3.5 V <strong>and</strong> 4.5 V propeller eng<strong>in</strong>e<br />

voltages. Decrease <strong>in</strong> travel times <strong>in</strong>dicate that pressure waves propagate<br />

faster when the fluid rotat<strong>in</strong>g velocity <strong>in</strong>creases. For a 39 mm diameter, the<br />

maximum <strong>in</strong>crease <strong>in</strong> pressure wave propagation velocity equals 8 ms −1 .<br />

<strong>Pressure</strong> traces from transducer 2 for four different propeller eng<strong>in</strong>e voltages<br />

are shown <strong>in</strong> Figure 6.20. Travel times are similar <strong>and</strong> equal 31.1 µs for 0 V<br />

<strong>and</strong> 2.5 V propeller eng<strong>in</strong>e voltages. Then travel times decrease to 30.86 µs<br />

<strong>and</strong> to 30.84 µs for 3.5 V <strong>and</strong> 4.5 V , respectively. Decrease <strong>in</strong> travel times<br />

<strong>in</strong>dicate once aga<strong>in</strong> that pressure wave propagate faster when the fluid rotat<strong>in</strong>g<br />

velocity <strong>in</strong>creases. For a 39 mm diameter, the maximum <strong>in</strong>crease <strong>in</strong> pressure<br />

wave propagation velocity equals 11 ms −1 .<br />

<strong>Pressure</strong> traces from transducer 3 for four different propeller eng<strong>in</strong>e voltages<br />

are shown <strong>in</strong> Figure 6.21. Travel times are similar <strong>and</strong> equal 31.93 µs for 0<br />

V <strong>and</strong> 2.5 V propeller eng<strong>in</strong>e voltages. Then travel times decrease to 31.9<br />

µs for both 3.5 V <strong>and</strong> 4.5 V . Decrease <strong>in</strong> travel times <strong>in</strong>dicate once aga<strong>in</strong><br />

108


6.2. Acoustic Velocity<br />

that pressure wave propagate faster when the fluid rotat<strong>in</strong>g velocity <strong>in</strong>creases.<br />

For a 39 mm diameter, the maximum <strong>in</strong>crease <strong>in</strong> pressure wave propagation<br />

velocity equals 2 ms −1 .<br />

<strong>Pressure</strong> traces from transducer 4 for four different propeller eng<strong>in</strong>e voltages<br />

are shown <strong>in</strong> Figure 6.22. Travel times are approximately equal to 31.72 µs<br />

<strong>in</strong> the four different cases. Travel times recorded from transducers 5, 6, 7 <strong>and</strong><br />

8 are equal <strong>in</strong>dependently of the propeller rotat<strong>in</strong>g speed <strong>and</strong> are not shown<br />

here<strong>in</strong>. Such observation <strong>in</strong>dicates that the pressure wave propagation velocity<br />

changes with fluid velocity.<br />

The propeller maximum l<strong>in</strong>ear velocity equals 2.29 ms −1 whereas a difference<br />

<strong>in</strong> pressure wave propagation velocity approximat<strong>in</strong>g 10 ms −1 is observed.<br />

Thus <strong>in</strong>dicat<strong>in</strong>g that fluid velocity affects acoustic velocity. Fluid velocity<br />

changes due to the propeller are greatest the closest to the propeller. However,<br />

the propeller proximity to the surface can <strong>in</strong>duce some air bubbles dur<strong>in</strong>g<br />

the rotation, hence expla<strong>in</strong><strong>in</strong>g a greater maximum <strong>in</strong>crease <strong>in</strong> pressure wave<br />

propagation <strong>in</strong> transducer 2 than <strong>in</strong> transducer 1.<br />

Experimental <strong>in</strong>accuracies due to vibrations, <strong>in</strong>appropriate control of propeller<br />

rotat<strong>in</strong>g velocity <strong>and</strong> fluid properties <strong>in</strong>duce that obta<strong>in</strong>ed results are only<br />

<strong>in</strong>dicative <strong>and</strong> not conclusive. However, experimental results tend to confirm<br />

that acoustic velocity depends on flow conditions <strong>and</strong> structure.<br />

Amplitude [V]<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

−0.01<br />

V = 4.5 V<br />

V = 3.5 V<br />

V = 2.5 V<br />

V = 0.0 V<br />

−0.02<br />

31.0 31.2 31.4 31.6 31.8 32.0<br />

Time [µs]<br />

Figure 6.19: Sound Beam Travel Times for Different Propeller<br />

Eng<strong>in</strong>e Voltage. Transducer 1<br />

109


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Amplitude [V]<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

−0.01<br />

V = 4.5 V<br />

V = 3.5 V<br />

V = 2.5 V<br />

V = 0.0 V<br />

−0.02<br />

30.6 30.8 31.0 31.2<br />

Time [µs]<br />

Figure 6.20: Sound Beam Travel Times for Different Propeller<br />

Eng<strong>in</strong>e Voltage. Transducer 2<br />

Amplitude [V]<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

−0.01<br />

V = 4.5 V<br />

V = 3.5 V<br />

V = 2.5 V<br />

V = 0.0 V<br />

−0.02<br />

31.8 31.9 32.0 32.1<br />

Time [µs]<br />

Figure 6.21: Sound Beam Travel Times for Different Propeller<br />

Eng<strong>in</strong>e Voltage. Transducer 3<br />

110


Amplitude [V]<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

−0.01<br />

V = 4.5 V<br />

V = 3.5 V<br />

V = 2.5 V<br />

V = 0.0 V<br />

6.3. Extended Acoustic Velocity Formulation<br />

−0.02<br />

31.5 31.6 31.7<br />

Time [µs]<br />

31.8 31.9<br />

Figure 6.22: Sound Beam Travel Times for Different Propeller<br />

Eng<strong>in</strong>e Voltage. Transducer 4<br />

6.3 Extended Acoustic Velocity Formulation<br />

Acoustic velocity of pressure wave propagat<strong>in</strong>g <strong>in</strong> fluid-filled pipel<strong>in</strong>e depends<br />

on fluid <strong>and</strong> pipel<strong>in</strong>e elastic properties, <strong>and</strong> experimental observations <strong>in</strong>dicate<br />

that flow velocity is also an important parameter. Based on the decomposition<br />

of acoustic velocity formulation <strong>in</strong> fluid isothermal compressibility <strong>and</strong> pipel<strong>in</strong>e<br />

compressibility, a new expression is proposed for determ<strong>in</strong><strong>in</strong>g acoustic velocity<br />

<strong>in</strong> s<strong>in</strong>gle phase fluid-filled pipel<strong>in</strong>e. The acoustic velocity can now be expressed<br />

by:<br />

�<br />

1<br />

a =<br />

(6.5)<br />

ρ (KFluid + KPipe − KFlow)<br />

where, ρ, is fluid density, KFluid, fluid isothermal compressibility, KPipe,<br />

pipel<strong>in</strong>e compressibility, <strong>and</strong>, KFlow, flow structure compressibility. Compressibilities<br />

are expressed <strong>in</strong> Pa −1 , that is <strong>in</strong>verse of pressure. No appropriate<br />

flow structure compressibility expression has yet been found. Nonetheless,<br />

flow structure compressibility should decrease when fluid velocity <strong>in</strong>creases.<br />

Another way to express this is that the fluid becomes stiffer when flow<strong>in</strong>g<br />

velocity <strong>in</strong>creases.<br />

111


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

Based on the first ultrasonic measurement data set (Figure 6.13), a fluid structure<br />

compressibility <strong>in</strong> the range 10 9 Pa −1 is calculated for a flow velocity of<br />

8 ms −1 <strong>and</strong> a Reynolds number about 21 10 9 . Curve fitt<strong>in</strong>g of acoustic velocity<br />

function of flow velocity with an exponential function is shown <strong>in</strong> Figure<br />

6.23. The curve fitt<strong>in</strong>g matches experimental data with a l<strong>in</strong>ear regression<br />

coefficient of 0.9973, <strong>and</strong> is expressed by:<br />

a = 0.010506 exp(0.26739u) + 373.75 (6.6)<br />

Acoustic Velocity [m s −1 ]<br />

374.4<br />

374.3<br />

374.2<br />

374.1<br />

374.0<br />

373.9<br />

373.8<br />

Experiment<br />

Fit<br />

373.7<br />

0 2 4 6 8 10 12 14 16<br />

Flow Velocity [m s −1 ]<br />

Figure 6.23: Detail of Measured <strong>and</strong> Calculated Acoustic Velocities<br />

Function of Flow Velocity - First Data Set Figure 6.13 -<br />

<strong>and</strong> Curve Fitt<strong>in</strong>g<br />

Acoustic velocity dependence on flow velocity can be responsible for pressure<br />

wave attenuation. <strong>Pressure</strong> <strong>and</strong> flow velocity are dependent variables, <strong>and</strong><br />

pressure waves are coupled with fluid velocity waves. Another way to express<br />

this is that fluid velocity changes as a pressure wave propagates. Oscillations of<br />

fluid velocity have been <strong>in</strong>vestigated by Yamanaka et al. (2002) <strong>and</strong> illustrated<br />

<strong>in</strong> Chapter 3. Changes <strong>in</strong> pressure <strong>and</strong> fluid velocity along a pressure wave<br />

then <strong>in</strong>duce acoustic velocity gradients along the pressure wave.<br />

112


6.3. Extended Acoustic Velocity Formulation<br />

Acoustic velocity gradients along a pressure wave then <strong>in</strong>duce acoustic impedance<br />

gradients along the pressure wave. Acoustic impedance gradients are responsible<br />

for pressure wave transmission <strong>and</strong> reflection. Thus, acoustic impedance<br />

gradients along a pressure wave <strong>in</strong>duce multiple transmission <strong>and</strong> reflection of<br />

the pressure wave itself as it propagates. <strong>Pressure</strong>, fluid velocity <strong>and</strong> acoustic<br />

velocity gradients for left-to-right propagat<strong>in</strong>g pressure waves are illustrated<br />

<strong>in</strong> Figure 6.24.<br />

∆p or ∆u or ∆a<br />

a)<br />

1<br />

0<br />

−1<br />

a2 > a1<br />

⇒<br />

a2 < a1<br />

Distance / Wavelength<br />

b)<br />

1<br />

0<br />

−1<br />

a2 < a1<br />

⇒<br />

a2 > a1<br />

Distance / Wavelength<br />

Figure 6.24: <strong>Pressure</strong>, Fluid <strong>and</strong> Acoustic Velocity Gradients <strong>in</strong><br />

Left-To-Right Propagat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong><br />

Left-to-right propagat<strong>in</strong>g positive pressure waves attenuate because the transmission<br />

coefficient is greater than 1, T < 1, Figure 6.24a). Reflected rightto-left<br />

propagat<strong>in</strong>g wave will be negative, R < 0, <strong>in</strong> such case. Left-to-right<br />

propagat<strong>in</strong>g negative pressure waves do not attenuate because the transmission<br />

coefficient is greater than 1, T > 1, Figure 6.24b), thus <strong>in</strong>duc<strong>in</strong>g a greater<br />

pressure wave negative amplitude. Reflected right-to-left propagat<strong>in</strong>g wave<br />

will be negative, R > 0, <strong>in</strong> such case.<br />

<strong>Pressure</strong> wave attenuation due to gradients of acoustic impedance along the<br />

pressure wave itself can be expla<strong>in</strong>ed <strong>in</strong> case of positive pressure pulse. Of particular<br />

<strong>in</strong>terest is pressure wave attenuation dur<strong>in</strong>g total reflection at boundary<br />

conditions. A reflected pressure wave superposition on itself dur<strong>in</strong>g reflection<br />

generates a pressure wave double <strong>in</strong> amplitude for a closed boundary, <strong>and</strong> cancels<br />

itself for a soft boundary. Acoustic impedance gradients along the wave<br />

113


Chapter 6. <strong>Pressure</strong> Wave Attenuation<br />

are therefore greater dur<strong>in</strong>g reflection than dur<strong>in</strong>g propagation. Another way<br />

to express this is that a pressure wave would attenuate more dur<strong>in</strong>g reflection<br />

than dur<strong>in</strong>g regular propagation process.<br />

However, negative pressure waves attenuation can not be expla<strong>in</strong>ed <strong>in</strong> the<br />

same way. In simple words, the proposed explanation needs to be extensively<br />

<strong>in</strong>vestigated based on more detailed <strong>and</strong> accurate experimental results, both<br />

<strong>in</strong>volv<strong>in</strong>g positive <strong>and</strong> negative pressure pulses, that are pressure waves that<br />

<strong>in</strong>crease <strong>and</strong> decrease locally <strong>and</strong> temporarily the <strong>in</strong>-situ static pressure, respectively.<br />

114


Chapter 7<br />

<strong>Impulse</strong> Pump<strong>in</strong>g Apparatus<br />

<strong>and</strong> Experiments<br />

7.1 Apparatus<br />

<strong>Impulse</strong> pump<strong>in</strong>g is a new concept for transport of fluid <strong>in</strong> pipel<strong>in</strong>es, patented<br />

by Sagov (2004), <strong>and</strong> developed by Clavis <strong>Impulse</strong> Technology AS. In impulse<br />

pump<strong>in</strong>g, the fluid source can be located far away from the external energy<br />

driven mechanism, <strong>and</strong> the fluid s<strong>in</strong>k can be located anywhere along the flowl<strong>in</strong>e.<br />

<strong>Impulse</strong> pump<strong>in</strong>g is based on the dynamics of pressure wave propagation<br />

while conventional pumps, such as centrifugal equipments, <strong>in</strong>crease the<br />

pipel<strong>in</strong>e <strong>in</strong>-situ static pressure for transport<strong>in</strong>g fluids.<br />

An impulse pump<strong>in</strong>g apparatus comprises an impulse generator, an outlet nonreturn<br />

valve, an <strong>in</strong>let non-return valve, <strong>and</strong> a fluid-filled flowl<strong>in</strong>e. Inlet <strong>and</strong><br />

outlet non-return valves connect the fluid-filled flowl<strong>in</strong>e to fluid source <strong>and</strong> s<strong>in</strong>k<br />

tanks, respectively. Inlet <strong>and</strong> outlet non-return valves are located at distance<br />

L <strong>and</strong> X of the impulse generator, respectively. The impulse generator is the<br />

only component that requires an external energy supply.<br />

<strong>Impulse</strong> pump<strong>in</strong>g can be applied to horizontal, <strong>in</strong>cl<strong>in</strong>ed <strong>and</strong> vertical situations,<br />

schematically represented <strong>in</strong> Figure 7.1, Figure 7.2, <strong>and</strong> figure 7.3, respectively.<br />

In vertical situation, impulse pump<strong>in</strong>g def<strong>in</strong>es a new artificial lift technique<br />

where fluid is pumped from downhole to wellhead. The impulse generator<br />

is located at wellhead, thus the only external energy requirement is at the<br />

wellhead. The <strong>in</strong>let <strong>and</strong> outlet non-return valves are located at distance L<br />

<strong>and</strong> X from the impulse generator, respectively.<br />

115


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

In pr<strong>in</strong>ciple, the impulse generator produces a pressure wave that propagates<br />

<strong>in</strong> the fluid-filled pipel<strong>in</strong>e. <strong>Pressure</strong> waves then actuate distant outlet nonreturn<br />

valve such that fluid enters the flowl<strong>in</strong>e from the <strong>in</strong>let tank. <strong>Pressure</strong><br />

waves also actuate outlet non-return valve such that fluid exits the flowl<strong>in</strong>e <strong>in</strong>to<br />

the outlet non-return tan. Therefore, pressure wave propagation generates a<br />

pump<strong>in</strong>g action where fluid flows from an <strong>in</strong>let tank to an outlet tank through<br />

a fluid-filled pipel<strong>in</strong>e.<br />

1<br />

X<br />

3<br />

2<br />

Figure 7.1: <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Horizontal Configuration<br />

1: <strong>Impulse</strong> Generator<br />

2: Outlet Non-Return Valve<br />

3: Outlet Tank<br />

4: Inlet Non-Return Valve<br />

5: Inlet Tank<br />

1<br />

2<br />

3<br />

X<br />

Figure 7.2: <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Incl<strong>in</strong>ed Configuration<br />

116<br />

L<br />

L<br />

4<br />

4<br />

5<br />

5


L<br />

X<br />

5<br />

4<br />

1<br />

2<br />

3<br />

7.1. Apparatus<br />

Figure 7.3: <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Vertical Configuration<br />

1: <strong>Impulse</strong> Generator<br />

2: Outlet Non-Return Valve<br />

3: Outlet Tank<br />

4: Inlet Non-Return Valve<br />

5: Inlet Tank<br />

117


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

7.2 Incl<strong>in</strong>ed Experimental Set-Up <strong>and</strong> Data<br />

Laboratory experimental data from an early impulse pump prototype for <strong>in</strong>cl<strong>in</strong>ed<br />

pump<strong>in</strong>g were obta<strong>in</strong>ed from Clavis <strong>Impulse</strong> Technology AS. An <strong>in</strong>cl<strong>in</strong>ed<br />

liquid-water-filled pipel<strong>in</strong>e, 75 m <strong>in</strong> length, 27 m <strong>in</strong> height, was connected to<br />

an impulse generator at its upper po<strong>in</strong>t, <strong>and</strong> to a source tank at its lower part,<br />

as shown <strong>in</strong> Figure 7.2. A s<strong>in</strong>k tank was located nearby the impulse generator.<br />

Inlet <strong>and</strong> outlet non-return valves separated the flowl<strong>in</strong>e from source <strong>and</strong> s<strong>in</strong>k<br />

tanks, respectively. Fluid was <strong>in</strong>itially at rest <strong>and</strong> atmospheric temperature.<br />

Experiments were performed at impulse generator frequencies from 2 Hz to 4<br />

Hz, with 0.5 Hz <strong>in</strong>crements. The impulse generator prototype used for vertical<br />

experiments was different from the equipment used for horizontal experiments.<br />

In addition, impulse generator frequency was the only user-set parameter.<br />

Experimental data sets <strong>in</strong>clude pressure traces at the impulse generator, at<br />

13 m depth, at 21 m depth, <strong>and</strong> 27 m depth, that is at the <strong>in</strong>let non-return<br />

valve. Correspond<strong>in</strong>g pipel<strong>in</strong>e lengths are listed <strong>in</strong> Table 7.1. <strong>Pressure</strong> signals<br />

were acquired with a 800 Hz sampl<strong>in</strong>g frequency.<br />

Table 7.1: Conversion between Vertical Height <strong>and</strong> Pipel<strong>in</strong>e Length<br />

Height [m] Length [m]<br />

13 36.3<br />

21 58.6<br />

27 75<br />

<strong>Pressure</strong> traces at four locations for a 2 Hz impulse generator frequency are<br />

shown <strong>in</strong> Figure 7.4. <strong>Pressure</strong> signals are cyclic, with m<strong>in</strong>imum pressure approximately<br />

0 bara. Maximum pressure amplitudes <strong>in</strong>crease with depth. Maximum<br />

pressures are 1.8 bara, 3.3 bara, 4.4 bara <strong>and</strong> 5 bara at 0 m, 13 m, 21<br />

m <strong>and</strong> 27 m depth, respectively. Increase <strong>in</strong> pressure with depth due to gravity<br />

is approximately 1 bara per 10 meters. <strong>Pressure</strong> waves attenuation with<br />

distance is not directly observed.<br />

<strong>Pressure</strong> traces at 13 m, 21 m <strong>and</strong> 27 m, are divided <strong>in</strong> three parts. The<br />

pressure approximately at 0 bara first <strong>in</strong>creases rapidly to reach a maximum<br />

value. The pressure then goes down before reach<strong>in</strong>g a new maximum value.<br />

The short decl<strong>in</strong>e <strong>in</strong> pressure is not observed close to the impulse generator,<br />

therefore <strong>in</strong>dicates that the decrease <strong>in</strong> pressure is <strong>in</strong>duced by an element<br />

further along the pipel<strong>in</strong>e, such as pipel<strong>in</strong>e components or non-return valves.<br />

118


<strong>Pressure</strong> [bara]<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

27 meter deep<br />

21 meter deep<br />

13 meter deep<br />

0 meter deep<br />

7.2. Incl<strong>in</strong>ed Experimental Set-Up <strong>and</strong> Data<br />

−1<br />

9.6 9.8 10 10.2<br />

Time [s]<br />

10.4 10.6<br />

Figure 7.4: Sample of <strong>Pressure</strong> Traces at Four Locations Along the<br />

Incl<strong>in</strong>ed Flow L<strong>in</strong>e. ˙γ = 2.0 Hz<br />

<strong>Pressure</strong> traces at four locations for a 3 Hz impulse generator frequency are<br />

shown <strong>in</strong> Figure 7.5. <strong>Pressure</strong> signals are cyclic, with m<strong>in</strong>imum pressure approximately<br />

0 bara. Maximum pressure amplitudes <strong>in</strong>crease with depth. Maximum<br />

pressures are 2.5 bara, 7.4 bara, 11.2 bara <strong>and</strong> 12 bara at 0 m, 13 m,<br />

21 m <strong>and</strong> 27 m depth, respectively. Increase <strong>in</strong> pressure with depth due to<br />

gravity is approximately 1 bara per 10 meters. Observed <strong>in</strong>crease <strong>in</strong> pressure<br />

amplitudes about 10 bara per 30 meters.<br />

<strong>Pressure</strong> maximum at four locations along the <strong>in</strong>cl<strong>in</strong>ed pipel<strong>in</strong>e are shown<br />

aga<strong>in</strong>st impulse generator frequency <strong>in</strong> Figure 7.6. Maximum pressure values<br />

<strong>in</strong>crease with impulse generator frequency. At the impulse generator, the<br />

maximum pressure <strong>in</strong>creases from 2 bara at 2 Hz, to 3 bara at 3 Hz, <strong>and</strong><br />

to 5.5 bara at 4 Hz. At the <strong>in</strong>let non-return valve, the maximum pressure<br />

<strong>in</strong>creases from 4.5 bara at 2 Hz, to 12.9 bara at 3 Hz, <strong>and</strong> to 13.7 bara at 4<br />

Hz.<br />

Maximum pressures at 13 m depth first <strong>in</strong>crease with frequency, then decreases<br />

slightly before stabiliz<strong>in</strong>g around 8 bara. Frequency spectra of pressure signals<br />

are not centered on the user-set impulse generator frequency. Another way<br />

to express this is that measured frequency are <strong>in</strong> disagreements with user set<br />

impulse generator frequency reported <strong>in</strong> data logs. Such observation <strong>in</strong>dicates<br />

that the impulse generator frequency calibration was <strong>in</strong>correctly performed.<br />

However, pressure measurements are not affected by such <strong>in</strong>accuracy.<br />

119


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

<strong>Pressure</strong> [bara]<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

27 meter deep<br />

21 meter deep<br />

13 meter deep<br />

0 meter deep<br />

−2<br />

9.6 9.8 10 10.2<br />

Time [s]<br />

10.4 10.6<br />

Figure 7.5: Sample of <strong>Pressure</strong> Traces at Four Locations Along the<br />

Incl<strong>in</strong>ed Flow L<strong>in</strong>e. ˙γ = 3.0 Hz<br />

<strong>Pressure</strong> [bara]<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

27 meter deep<br />

21 meter deep<br />

13 meter deep<br />

0 meter deep<br />

0<br />

2 2.5 3<br />

Frequency [Hz]<br />

3.5 4<br />

Figure 7.6: Maximum <strong>Pressure</strong>s at Four Locations for Five Different<br />

Experiments<br />

120


7.3. Horizontal Experimental Set-Up <strong>and</strong> Data<br />

7.3 Horizontal Experimental Set-Up <strong>and</strong> Data<br />

Laboratory experiments with an impulse pump prototype were carried out<br />

<strong>in</strong> March 2007, by <strong>NTNU</strong>, at Clavis <strong>Impulse</strong> Technology AS <strong>in</strong> Hokksund<br />

(Norway). An horizontal liquid-water-filled pipel<strong>in</strong>e, 283 m <strong>in</strong> length, 1 <strong>in</strong>ch<br />

<strong>in</strong> diameter, <strong>and</strong> 1 10 −3 m <strong>in</strong> wall thickness, was connected to an impulse<br />

generator at one end, <strong>and</strong> to a source tank at the other end. A s<strong>in</strong>k tank was<br />

approximately located 1 m from the impulse generator. Inlet <strong>and</strong> outlet nonreturn<br />

valves separated the flowl<strong>in</strong>e from source <strong>and</strong> s<strong>in</strong>k tanks, respectively.<br />

Fluid was <strong>in</strong>itially at rest <strong>and</strong> room temperature.<br />

Experiments were performed at impulse generator frequencies from 1.3 Hz<br />

to 1.7 Hz, with 0.1 Hz steps, <strong>and</strong> piston displacements from 0.5 10 −3 m to<br />

1.5 10 −3 m, with 0.25 10 −3 m steps. Experimental data sets <strong>in</strong>clude impulse<br />

generator force <strong>and</strong> piston displacement signals, flow rates measured at the<br />

outlet, <strong>and</strong> pressure traces at the outlet non-return valve, <strong>and</strong> at the <strong>in</strong>let<br />

non-return valve. Data were acquired with a 1200 Hz sampl<strong>in</strong>g frequency.<br />

Data here<strong>in</strong> shown for presentation purposes were obta<strong>in</strong>ed from a particular<br />

experiment at 1.4 Hz <strong>and</strong> 0.75 10 −3 m piston frequency <strong>and</strong> displacement,<br />

respectively. Other data correspond<strong>in</strong>g to the range of frequencies <strong>and</strong> piston<br />

displacements tested dur<strong>in</strong>g experimental work, exhibited similar properties<br />

as data shown hereunder. Therefore only one data set is presented <strong>in</strong> the<br />

follow<strong>in</strong>g of the chapter.<br />

The impulse generator comprises a back <strong>and</strong> forth mov<strong>in</strong>g piston <strong>and</strong> a freetranslat<strong>in</strong>g<br />

disk. The back <strong>and</strong> forth mov<strong>in</strong>g piston is mechanically actuated<br />

at user set frequency <strong>and</strong> stroke. Piston displacements are transmitted to the<br />

free translat<strong>in</strong>g disk through a compressible fluid. Displacement ratio between<br />

the piston <strong>and</strong> the disk was measured prior to the experimental work <strong>and</strong> was<br />

equal to 12. Such configuration allowed the mechanically actuated piston to<br />

be separated from the flowl<strong>in</strong>e, hence facilitat<strong>in</strong>g operation <strong>and</strong> ma<strong>in</strong>tenance<br />

dur<strong>in</strong>g experimental work.<br />

The back <strong>and</strong> forth mov<strong>in</strong>g piston was fitted with a stroke gauge. The s<strong>in</strong>usoidal<br />

piston displacement, 9 10 −3 m <strong>in</strong> peak to peak amplitude for this<br />

particular experiment, is shown <strong>in</strong> Figure 7.7. The result<strong>in</strong>g free-translat<strong>in</strong>g<br />

disk displacement is then 0.75 10 −3 m <strong>in</strong> amplitude. The back <strong>and</strong> forth mov<strong>in</strong>g<br />

piston was also fitted with a force gauge. The force signal is a s<strong>in</strong>usoid<br />

truncated on its upper part. The truncated s<strong>in</strong>usoidal force on the piston signal,<br />

2.5 10 3 N <strong>in</strong> amplitude for this particular experiment, is shown <strong>in</strong> Figure<br />

7.8.<br />

121


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

Stroke [10 −3 m]<br />

4<br />

2<br />

0<br />

−2<br />

−4<br />

−6<br />

−8<br />

75 76 77 78 79 80<br />

Time [s]<br />

Figure 7.7: Sample of Piston Stroke Trace. ˙γ = 1.4 Hz,<br />

∆l = 0.75 10 −3 m<br />

Force [10 3 N]<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

75 76 77 78 79 80<br />

Time [s]<br />

Figure 7.8: Sample of Force on the Piston Trace. ˙γ = 1.4 Hz,<br />

∆l = 0.7510 −3 m<br />

122


Normalized Amplitude<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

7.3. Horizontal Experimental Set-Up <strong>and</strong> Data<br />

0.0<br />

0 0.5 1<br />

Frequency [Hz]<br />

1.5 2<br />

Figure 7.9: Fourier Analysis of Piston Stroke Signal. ˙γ = 1.4 Hz,<br />

∆l = 0.7510 −3 m<br />

Frequency spectrum of the stroke signal us<strong>in</strong>g Fourier transform conta<strong>in</strong>s a<br />

s<strong>in</strong>gle 1.4 Hz frequency, as shown <strong>in</strong> Figure 7.9. Hence, experimental data<br />

correspond to user-set impulse generator frequency <strong>and</strong> displacement parameters.<br />

A first pressure transducer was located <strong>in</strong> the fluid-filled pipel<strong>in</strong>e just upstream<br />

of the outlet non-return valve. The outlet pressure signal is a s<strong>in</strong>usoid<br />

truncated on its upper part. The truncated s<strong>in</strong>usoidal pressure signal, approximately<br />

4 bara <strong>in</strong> amplitude for this particular experiment, is shown <strong>in</strong><br />

Figure 7.10. M<strong>in</strong>imum <strong>and</strong> maximum pressure values at the outlet are 0.75<br />

bara <strong>and</strong> 4.4 bara, approximately. The truncated upper part of the signal<br />

conta<strong>in</strong>s pressure oscillations that correspond to oscillations observed on the<br />

force signal, Figure 7.8.<br />

A second pressure transducer was located <strong>in</strong> the fluid-filled pipel<strong>in</strong>e just downstream<br />

of the <strong>in</strong>let non-return valve. The <strong>in</strong>let pressure signal is cyclic, with<br />

maximum pressure 8.4 bara <strong>in</strong> amplitude, as shown <strong>in</strong> Figure 7.11. M<strong>in</strong>imum<br />

pressure values are -0.1 bara, <strong>in</strong>dicat<strong>in</strong>g that pressure transducer calibration<br />

was <strong>in</strong>accurate <strong>and</strong> pressure data are reported with an offset. The <strong>in</strong>let nonreturn<br />

valve is located further from the impulse generator than the outlet<br />

non-return valve. However, pressure amplitudes observed are greater at the<br />

<strong>in</strong>let non-return valve than at the outlet non-return valve.<br />

123


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

<strong>Pressure</strong> [bara]<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

75 76 77 78 79 80<br />

Time [s]<br />

Figure 7.10: Sample of <strong>Pressure</strong> Trace at the Outlet Non-Return<br />

Valve. ˙γ = 1.4 Hz, ∆l = 0.7510 −3 m<br />

<strong>Pressure</strong> [bara]<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

75 76 77 78 79 80<br />

Time [s]<br />

Figure 7.11: Sample of <strong>Pressure</strong> Trace at the Inlet Non-Return<br />

Valve. ˙γ = 1.4 Hz, ∆l = 0.7510 −3 m<br />

124


Flow Rate [l m<strong>in</strong> −1 ]<br />

1.95<br />

1.90<br />

1.85<br />

1.80<br />

1.75<br />

7.3. Horizontal Experimental Set-Up <strong>and</strong> Data<br />

1.70<br />

75 76 77 78 79 80<br />

Time [s]<br />

Figure 7.12: Sample of <strong>Impulse</strong> Pump<strong>in</strong>g Flow Rate Trace.<br />

˙γ = 1.4 Hz, ∆l = 0.7510 −3 m<br />

An electromagnetic flowmeter was located at the outlet downstream of the<br />

non-return valve. Flow rate data are shown <strong>in</strong> Figure 7.12. Presented flow<br />

rate data are averaged over 50 po<strong>in</strong>ts for clarity. The measured flow rate<br />

fluctuates around a 1.81 l m<strong>in</strong> −1 ±0.1 l m<strong>in</strong> −1 average value, <strong>in</strong>dicat<strong>in</strong>g that<br />

impulse pump<strong>in</strong>g <strong>in</strong>duces pulsat<strong>in</strong>g flow.<br />

Outlet <strong>and</strong> <strong>in</strong>let pressure traces for 1.4 Hz <strong>and</strong> 1.25 10 −3 m impulse generator<br />

frequency <strong>and</strong> displacement, are shown <strong>in</strong> Figure 7.13 <strong>and</strong> Figure 7.14,<br />

respectively. The outlet maximum pressure is aga<strong>in</strong> approximately 4.4 bara<br />

while the maximum <strong>in</strong>let pressure is now 10 bara, thus confirm<strong>in</strong>g that the<br />

pressure amplitudes are greater at the <strong>in</strong>let non-return valve than at the outlet<br />

non-return valve close to the impulse generator.<br />

Maximum pressure evolution at the outlet non-return valve, function of the<br />

impulse generator displacement for different frequencies is shown <strong>in</strong> Figure<br />

7.15. Maximum pressure values at the outlet non-return valve <strong>in</strong>crease with<br />

displacement <strong>and</strong> decrease with frequency. At 1.6 Hz frequency, maximum<br />

pressure amplitude <strong>in</strong>creases from 1.75 bara at 0.5 10 −3 m, to 4.3 bara at 1<br />

10 −3 m, to 5.4 bara at 1.75 10 −3 m. At 0.75 10 −3 m displacement, maximum<br />

pressure amplitude decreases from 5 bara at 1.3 Hz, to 4 bara at 1.5 Hz, <strong>and</strong><br />

to 3 bara at 1.6 Hz.<br />

125


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

<strong>Pressure</strong> [bara]<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

248 249 250 251 252 253<br />

Time [s]<br />

Figure 7.13: Sample of <strong>Pressure</strong> Trace at Outlet Non-Return Valve.<br />

˙γ = 1.4 Hz, ∆l = 1.2510 −3 m<br />

<strong>Pressure</strong> [bara]<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

248 249 250 251 252 253<br />

Time [s]<br />

Figure 7.14: Sample of <strong>Pressure</strong> Trace at Inlet Non-Return Valve.<br />

˙γ = 1.4 Hz, ∆l = 1.2510 −3 m<br />

126


<strong>Pressure</strong> [bara]<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

γ = 1.6 Hz<br />

γ = 1.5 Hz<br />

γ = 1.4 Hz<br />

γ = 1.3 Hz<br />

7.3. Horizontal Experimental Set-Up <strong>and</strong> Data<br />

0<br />

0 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2<br />

Stroke [10 −3 m]<br />

Figure 7.15: Maximum <strong>Pressure</strong> Amplitudes at Outlet Non-Return<br />

Valve Aga<strong>in</strong>st <strong>Impulse</strong> Generator Displacement for<br />

Different Frequencies<br />

Maximum pressure evolution at the <strong>in</strong>let non-return valve, function of the<br />

impulse generator displacement for different frequencies is shown <strong>in</strong> Figure<br />

7.16. Maximum pressure values at the outlet non-return valve <strong>in</strong>crease with<br />

displacement <strong>and</strong> decrease with frequency. At 1.5 Hz frequency, maximum<br />

pressure amplitude <strong>in</strong>creases from 7.5 bara at 0.5 10 −3 m, to 10.8 bara at 1<br />

10 −3 m, to 11.2 bara at 1.75 10 −3 m. At 0.75 10 −3 m displacement, maximum<br />

pressure amplitude decreases from 8.9 bara at 1.3 Hz, to 7.8 bara at 1.6 Hz.<br />

Flow rate evolutions for the set of experimental data, from 1.3 Hz to 1.6 Hz<br />

<strong>and</strong> from 0.5 10 −3 m to 1.5 10 −3 m impulse generator frequencies <strong>and</strong> displacements,<br />

ares shown <strong>in</strong> Figure 7.17. The flow rate <strong>in</strong>creases with impulse<br />

generator displacement <strong>and</strong> frequency. For a 1 10 −3 m impulse generator displacement,<br />

the flow rate <strong>in</strong>creases from 3.13 l m<strong>in</strong> −1 at 1.3 Hz, to 3.29 l m<strong>in</strong> −1<br />

at 1.4 Hz, <strong>and</strong> to 3.83 l m<strong>in</strong> −1 at 1.6 Hz. For a 1.6 Hz impulse generator frequency,<br />

the flow rate <strong>in</strong>creases from 1.1 l m<strong>in</strong> −1 at 0.5 10 −3 m to 4.87 l m<strong>in</strong> −1<br />

at 1.25 10 −3 m.<br />

The 1.5 Hz experimental flow rate lies below the other experimental flow rates,<br />

with a drop at 1.75 10 −3 ms −1 (or 1.25 10 −3 m impulse generator displacement).<br />

The 1.5 Hz experiment disparity with other experimental flow rates<br />

can <strong>in</strong>dicate experimental measurement errors or be a characteristics l<strong>in</strong>ked<br />

to impulse pump<strong>in</strong>g.<br />

127


Chapter 7. <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus <strong>and</strong> Experiments<br />

<strong>Pressure</strong> [bara]<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

γ = 1.6 Hz<br />

γ = 1.5 Hz<br />

γ = 1.4 Hz<br />

γ = 1.3 Hz<br />

0<br />

0 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2<br />

Stroke [10 −3 m]<br />

Figure 7.16: Maximum <strong>Pressure</strong> Amplitudes at Inlet Non-Return<br />

Valve Function of <strong>Impulse</strong> Generator Displacement for<br />

Different Frequencies<br />

Flow Rate [l m<strong>in</strong> −1 ]<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

γ = 1.6 Hz<br />

γ = 1.5 Hz<br />

γ = 1.4 Hz<br />

γ = 1.3 Hz<br />

0<br />

0 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2<br />

Stroke [10 −3 m]<br />

Figure 7.17: Experimental Flow Rates Function of <strong>Impulse</strong><br />

Generator Displacement for Different Frequencies<br />

128


Chapter 8<br />

Concept, Modell<strong>in</strong>g <strong>and</strong><br />

Simulation of <strong>Impulse</strong><br />

Pump<strong>in</strong>g<br />

8.1 Concept <strong>and</strong> Modell<strong>in</strong>g<br />

<strong>Impulse</strong> pump<strong>in</strong>g is a new pump<strong>in</strong>g concept based on pressure wave propagation<br />

(Sagov, 2004). An impulse pump<strong>in</strong>g apparatus comprises an impulse<br />

generator, an outlet non-return valve <strong>and</strong> an <strong>in</strong>let non-return valve that connect<br />

a fluid-filled pipel<strong>in</strong>e to outlet <strong>and</strong> <strong>in</strong>let tanks, respectively. A schematic<br />

representation of a horizontal impulse pump<strong>in</strong>g apparatus is shown <strong>in</strong> Figure<br />

8.1, where <strong>in</strong>let <strong>and</strong> outlet non-return valves are located at distance L <strong>and</strong> X<br />

from the impulse generator, respectively.<br />

8.1.1 <strong>Impulse</strong> Generator<br />

The impulse generator oscillates back <strong>and</strong> forth to produce a pressure wave<br />

that propagates <strong>in</strong> the fluid-filled pipel<strong>in</strong>e on top of the <strong>in</strong>-situ static pressure.<br />

Forward oscillations generate pressure pulses that <strong>in</strong>crease temporarily the<br />

local pressure. Whereas backward oscillations generate pressure pulses that<br />

decrease temporarily the local pressure. Forward <strong>and</strong> backward oscillations<br />

then generate positive <strong>and</strong> negative pressure pulses relative to local <strong>in</strong>-situ<br />

static pressures, respectively.<br />

The impulse generator can be seen as a back <strong>and</strong> forth mov<strong>in</strong>g piston. The<br />

oscillation function def<strong>in</strong>es a pressure waveform that depends on frequency<br />

<strong>and</strong> stroke length. For illustration purposes, a s<strong>in</strong>usoidal pressure waveform<br />

similar to the experimental piston stroke signal is modelled (Figure 7.7). The<br />

impulse generator behaves as a mov<strong>in</strong>g boundary condition whose movement is<br />

129


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

1<br />

X<br />

3<br />

2<br />

Figure 8.1: <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Horizontal Configuration<br />

1: <strong>Impulse</strong> Generator<br />

2: Outlet Non-Return Valve<br />

3: Outlet Tank<br />

4: Inlet Non-Return Valve<br />

5: Inlet Tank<br />

modelled by a time dependent fluid velocity. The impulse generator boundary<br />

condition can be expressed by:<br />

L<br />

u (0, t) = pWave<br />

ρa s<strong>in</strong>(ω (t − t0)) (8.1)<br />

where, u, is fluid velocity, pWave, pressure wave amplitude, ρ, fluid density, a,<br />

acoustic velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e, ω, angular frequency, <strong>and</strong>, t0, <strong>in</strong>itial<br />

time of pressure wave generation. The generated pressure wave then propagates<br />

on top of the <strong>in</strong>-situ static pressure. At maximum pressure amplitude<br />

(s<strong>in</strong>(ω (t − t0)) = 1), the above equation Eq. 8.1 is similar to the Joukowski<br />

equation:<br />

pWave = ρau (8.2)<br />

Back <strong>and</strong> forth piston oscillations <strong>in</strong>duce changes <strong>in</strong> volume <strong>in</strong> the fluid-filled<br />

pipel<strong>in</strong>e, hence changes <strong>in</strong> pressure that propagate <strong>in</strong> the form of a pressure<br />

wave. A pressure wave can consequently be decomposed <strong>in</strong> countless <strong>in</strong>f<strong>in</strong>itesimal<br />

displacements of the impulse generator. Changes <strong>in</strong> pressure <strong>and</strong> impulse<br />

generator displacements, thus changes <strong>in</strong> volume, are related by fluid compressibility<br />

(Leighton, 1994):<br />

K = −1 ∂V<br />

V0 ∂p<br />

130<br />

4<br />

5<br />

(8.3)


8.1. Concept <strong>and</strong> Modell<strong>in</strong>g<br />

where, K, represents fluid compressibility, <strong>and</strong>, V0, volume of fluid orig<strong>in</strong>ally<br />

<strong>in</strong> place. For <strong>in</strong>f<strong>in</strong>itesimal changes <strong>in</strong> volume, the first order partial derivative<br />

of volume with pressure <strong>in</strong>cluded <strong>in</strong> the fluid compressibility def<strong>in</strong>ition Eq. 8.3<br />

can be approximated by:<br />

K ≈ 1<br />

V0<br />

∆V<br />

∆p<br />

(8.4)<br />

Back-<strong>and</strong>-forth oscillations of the impulse generator can be modelled by displacements<br />

of a piston. <strong>Pressure</strong> waves can be generated directly <strong>in</strong> the fluidfilled<br />

pipel<strong>in</strong>e, L <strong>in</strong> length, <strong>and</strong> d <strong>in</strong> diameter, or <strong>in</strong> a connected pipel<strong>in</strong>e of<br />

greater diameter dp, as illustrated <strong>in</strong> Figure 8.2. The fluid compressibility (Eq.<br />

8.3) can then be <strong>in</strong>tegrated over a piston stroke ∆l:<br />

K =<br />

� dp<br />

d<br />

� 2 1<br />

L<br />

� ∆l<br />

0<br />

∂x<br />

∂p<br />

φ dp<br />

Figure 8.2: Illustration of <strong>Impulse</strong> Generator Generat<strong>in</strong>g <strong>Pressure</strong><br />

<strong>Waves</strong> <strong>in</strong> Diameter Greater than Flowl<strong>in</strong>e Dimensions<br />

φ d<br />

(8.5)<br />

The piston stroke required for generat<strong>in</strong>g a pressure wave, pWave <strong>in</strong> amplitude,<br />

can now be calculated us<strong>in</strong>g Eq. 8.5 assum<strong>in</strong>g fluid compressibility changes<br />

negligible with pressure:<br />

� �2 d<br />

∆l = KL<br />

(8.6)<br />

dp<br />

pWave<br />

The piston stroke <strong>in</strong> Eq. 8.6 is calculated from a central position of the<br />

piston. Another way to express this is that a pressure 2pWave <strong>in</strong> peak-to-peak<br />

amplitude can be decomposed <strong>in</strong> one ∆l forward movement, followed by two<br />

∆l backward movements <strong>and</strong> one ∆l forward movement, as illustrated for a<br />

s<strong>in</strong>usoidal pressure wave <strong>in</strong> Figure 8.3. <strong>Pressure</strong> wave generation therefore<br />

requires four piston strokes per wavelength. Hence, the piston l<strong>in</strong>ear velocity,<br />

uPiston, correspond<strong>in</strong>g to oscillation frequency, ˙γ, can now be expressed by:<br />

uPiston = 4∆l ˙γ (8.7)<br />

131


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

(p − p 0 ) / p Wave<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

−0.2<br />

−0.4<br />

−0.6<br />

−0.8<br />

↑<br />

Forward<br />

Backward<br />

↓<br />

−1.0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Distance / Wavelength<br />

↓<br />

Backward<br />

Forward<br />

Figure 8.3: Illustration of Piston Back-And-Forth Oscillations for<br />

S<strong>in</strong>usoidal <strong>Pressure</strong> Wave Generation<br />

One wavelength decomposition <strong>in</strong> 4 piston displacements is theoretical <strong>and</strong> is<br />

here<strong>in</strong> used <strong>in</strong> order to relate piston stroke length <strong>and</strong> velocity with pressure<br />

wave peak-to-peak semi amplitude. Practically, only two phases make for one<br />

pressure wave cycle: a 2 ∆l forward movement followed by a 2 ∆l backward<br />

displacement.<br />

8.1.2 <strong>Pressure</strong> Wave Propagation<br />

Generated pressure waves propagate on top of the <strong>in</strong>-situ static pressure <strong>in</strong><br />

the fluid-filled pipel<strong>in</strong>e. Acoustic velocity <strong>in</strong> hydrocarbon pipel<strong>in</strong>es are <strong>in</strong> the<br />

range 300 ms −1 for gases, <strong>and</strong> 750 ms −1 for s<strong>in</strong>gle phase liquid oils. Fluid is<br />

<strong>in</strong>itially at rest <strong>and</strong> back <strong>and</strong> forth impulse generator oscillations <strong>in</strong>duce low<br />

velocity pulsat<strong>in</strong>g flow. The correspond<strong>in</strong>g Mach number is then very low.<br />

Hence, assumptions commonly used for water hammer modell<strong>in</strong>g can also be<br />

applied for impulse pump<strong>in</strong>g modell<strong>in</strong>g. Mass <strong>and</strong> momentum conservation<br />

equations are then, respectively (Ghidaoui et al., 2005):<br />

∂p<br />

+ ρa2∂u = 0 (8.8)<br />

∂t ∂x<br />

∂u<br />

∂t<br />

1 ∂p<br />

+<br />

ρ ∂x<br />

−1<br />

= fu|u| + F (8.9)<br />

2d<br />

132<br />


8.1. Concept <strong>and</strong> Modell<strong>in</strong>g<br />

where, f, is friction factor, <strong>and</strong>, F, acceleration due to external forces. Unsteady<br />

friction <strong>and</strong> pipel<strong>in</strong>e viscoelasticity are omitted for simplicity <strong>and</strong> due<br />

to the uncerta<strong>in</strong>ty of their validity. Therefore, steady-state frictional pressure<br />

drop is the only pressure wave attenuation mechanism <strong>in</strong>cluded <strong>in</strong> modell<strong>in</strong>g<br />

of impulse pump<strong>in</strong>g. <strong>Pressure</strong> wave amplitudes are then overestimated <strong>and</strong><br />

waveform changes not accurately modelled. Nonetheless, here<strong>in</strong> modell<strong>in</strong>g of<br />

impulse pump<strong>in</strong>g uses s<strong>in</strong>usoidal pressure waves, that is without steep gradients<br />

<strong>in</strong> amplitude. Thus, changes <strong>in</strong> pressure waveform are not of importance.<br />

<strong>Pressure</strong> waves propagation velocity is the acoustic velocity <strong>in</strong> the fluid-filled<br />

pipel<strong>in</strong>e. The acoustic velocity depends on fluid density, ρ, <strong>and</strong> isothermal<br />

compressibility, KΓ, <strong>and</strong> pipel<strong>in</strong>e diameter, d, wall-thickness, e, <strong>and</strong> material<br />

Young’s modulus, Y . The fluid-filled pipel<strong>in</strong>e is fastened at one end by the<br />

impulse generator <strong>and</strong> at the other end by the <strong>in</strong>let tank. Acoustic velocity<br />

for a fluid-filled th<strong>in</strong>-walled pipel<strong>in</strong>e is then expressed by (Wylie <strong>and</strong> Streeter,<br />

1993):<br />

�<br />

�<br />

� 1<br />

a = � �<br />

�<br />

ρ KΓ + d<br />

� (8.10)<br />

Y e<br />

8.1.3 Pipel<strong>in</strong>e Outlet<br />

A generated pressure wave then propagates from the impulse generator to<br />

the outlet non-return valve. The outlet non-return valve can open when the<br />

local pressure <strong>in</strong> the fluid-filled pipel<strong>in</strong>e, p0, is greater than the valve open<strong>in</strong>g<br />

pressure. The outlet non-return valve open<strong>in</strong>g pressure is def<strong>in</strong>ed as the sum<br />

of the outlet tank pressure, pOutlet, <strong>and</strong> the valve mechanical resistance, or<br />

crack<strong>in</strong>g pressure, pCO. A schematic representation of the outlet non-return<br />

valve is shown <strong>in</strong> Figure 8.4(a).<br />

Positive pressure pulses <strong>in</strong>crease temporarily the local pressure <strong>in</strong> the fluidfilled<br />

pipel<strong>in</strong>e at the outlet non-return valve. The outlet non-return valve can<br />

then open <strong>and</strong> fluid flows from the fluid-filled pipel<strong>in</strong>e <strong>in</strong>to the outlet tank.<br />

The outlet non-return valve open<strong>in</strong>g then limits the pressure <strong>in</strong>side the fluidfilled<br />

pipel<strong>in</strong>e by truncat<strong>in</strong>g the positive pressure pulse upper part. The outlet<br />

non-return valve open<strong>in</strong>g condition, thus the maximum local pressure <strong>in</strong>side<br />

the fluid-filled pipel<strong>in</strong>e can be expressed by:<br />

p0 (X) > pOutlet + pCO<br />

(8.11)<br />

The fluid <strong>in</strong>itially at rest <strong>in</strong> the fluid-filled pipel<strong>in</strong>e at the outlet non-return<br />

valve starts flow<strong>in</strong>g out when the valve opens. Consequently, a negative water<br />

hammer pressure wave propagates <strong>in</strong>side the fluid-filled pipel<strong>in</strong>e. The local<br />

133


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

p0 (L)<br />

pOutlet<br />

p0 (X)<br />

pCO<br />

(a) Outlet Non-Return Valve<br />

pCI<br />

(b) Inlet Non-Return Valve<br />

pInlet<br />

Figure 8.4: Non-Return Valve Schematic Representations<br />

pressure <strong>in</strong> the fluid-filled pipel<strong>in</strong>e at the valve then decreases, thus counteract<strong>in</strong>g<br />

the positive pressure pulse open<strong>in</strong>g action. Local pressure fluctuations<br />

<strong>in</strong>duce non-return valve vibrations that generate pressure oscillations of low<br />

amplitude <strong>and</strong> high frequency.<br />

The outlet non-return valve closes when the <strong>in</strong>equality Eq. 8.11 is no longer<br />

verified. Truncated pressure waves then propagate from the outlet non-return<br />

valve to the distant <strong>in</strong>let non-return valve.<br />

8.1.4 Pipel<strong>in</strong>e Inlet<br />

The <strong>in</strong>let non-return valve can open when the local pressure <strong>in</strong> the fluid-filled<br />

pipel<strong>in</strong>e, p0 (L), is lower than the valve open<strong>in</strong>g pressure. The <strong>in</strong>let valve<br />

open<strong>in</strong>g pressure is def<strong>in</strong>ed as the <strong>in</strong>let tank pressure, pInlet, decreased by the<br />

valve crack<strong>in</strong>g pressure, pCI. A schematic representation of an <strong>in</strong>let non-return<br />

valve is shown <strong>in</strong> Figure 8.4(b).<br />

134


8.2. Numerical Simulation Tool<br />

Negative pressure pulses decrease temporarily the local pressure <strong>in</strong> the fluidfilled<br />

pipel<strong>in</strong>e at the <strong>in</strong>let non-return valve. The <strong>in</strong>let non-return valve can<br />

then open <strong>and</strong> fluid flows from the <strong>in</strong>let tank <strong>in</strong>to the fluid-filled pipel<strong>in</strong>e.<br />

Negative pressure pulses then reflect totally with phase change at an open<br />

boundary condition. Another way to express this is that negative pressure<br />

pulses reflect as positive pressure pulses at an open boundary condition. The<br />

<strong>in</strong>let non-return valve open<strong>in</strong>g condition can be expressed by:<br />

p0 (L) < pInlet − pCI<br />

(8.12)<br />

The fluid <strong>in</strong>itially at rest <strong>in</strong> the fluid-filled pipel<strong>in</strong>e at the <strong>in</strong>let non-return valve<br />

starts flow<strong>in</strong>g when the valve opens. A positive water hammer pressure wave<br />

therefore propagates <strong>in</strong>side the fluid-filled pipel<strong>in</strong>e. The local pressure <strong>in</strong> the<br />

fluid-filled pipel<strong>in</strong>e at the valve then <strong>in</strong>creases, thus counteract<strong>in</strong>g the negative<br />

pressure pulse open<strong>in</strong>g action. Local pressure fluctuations <strong>in</strong>duce non-return<br />

valve vibrations that generate pressure oscillations of low amplitude <strong>and</strong> high<br />

frequency.<br />

The <strong>in</strong>let non-return valve closes when the <strong>in</strong>equality Eq. 8.12 is no longer<br />

valid. Positive pressure pulses <strong>in</strong>crease temporarily the local pressure <strong>in</strong> the<br />

fluid-filled pipel<strong>in</strong>e at the <strong>in</strong>let non-return valve. The valve then rema<strong>in</strong>s<br />

closed <strong>and</strong> positive pressure pulses reflect totally at a close boundary. Another<br />

way to express this is that positive pressure pulses reflect as positive pressure<br />

pulses. <strong>Pressure</strong> waves with truncated positive pressure pulses <strong>and</strong> reflected<br />

phase-changed negative pressure pulses then propagate back to the impulse<br />

generator.<br />

In simple words, a pressure wave that <strong>in</strong>cludes a positive pulse <strong>and</strong> a negative<br />

pulse propagates from the impulse generator <strong>in</strong> the fluid-filled pipel<strong>in</strong>e. The<br />

positive pressure pulse actuates the outlet non-return valve <strong>and</strong> fluid flows<br />

<strong>in</strong>to the outlet tank. The negative pressure pulse actuates the <strong>in</strong>let nonreturn<br />

valve <strong>and</strong> fluid flows from the <strong>in</strong>let tank. Thus, pressure waves produce<br />

a cyclic pump<strong>in</strong>g action characterized by a pulsat<strong>in</strong>g flow.<br />

8.2 Numerical Simulation Tool<br />

<strong>Impulse</strong> pump<strong>in</strong>g is based on pressure wave propagation that is modelled us<strong>in</strong>g<br />

the one dimensional hyperbolic system of mass <strong>and</strong> momentum conservation<br />

equations. The common low Mach number approximation used <strong>in</strong> pressure<br />

transient modell<strong>in</strong>g is also applied <strong>in</strong> modell<strong>in</strong>g of impulse pump<strong>in</strong>g. <strong>Impulse</strong><br />

pump<strong>in</strong>g process can then be simulated us<strong>in</strong>g the Lax-Wendroff numerical<br />

scheme coupled with the Superbee flux limiter selected <strong>in</strong> Chapter 5. Simulated<br />

output variables are then fluid velocity <strong>and</strong> pressure at every grid po<strong>in</strong>ts<br />

<strong>and</strong> time steps.<br />

135


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

A numerical simulation tool was developed based on a Riemann solver with<br />

the Lax-Wendroff numerical scheme <strong>and</strong> the Superbee flux limiter <strong>in</strong>cluded<br />

<strong>in</strong> CLAWPACK. Specific subrout<strong>in</strong>es are implemented to reproduce impulse<br />

pump<strong>in</strong>g apparatus characteristics <strong>and</strong> behaviour, tak<strong>in</strong>g <strong>in</strong>to account fluid<br />

properties <strong>and</strong> system configuration, dimensions <strong>and</strong> geometry. The developed<br />

software is <strong>in</strong>tended for petroleum eng<strong>in</strong>eers <strong>and</strong> is therefore best described <strong>in</strong><br />

artificial lift applications, that is vertical well as illustrated <strong>in</strong> Figure 8.5.<br />

3<br />

2<br />

1<br />

X XF<br />

Formation<br />

Figure 8.5: <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus Applied to Artificial Lift<br />

1: <strong>Impulse</strong> Generator<br />

2: Outlet Non-Return Valve<br />

3: Outlet Tank<br />

136<br />

L


8.2. Numerical Simulation Tool<br />

Specific subrout<strong>in</strong>es were developed around the Riemann solver to reproduce<br />

impulse pump<strong>in</strong>g characteristics for diverse applications. Each apparatus component<br />

work<strong>in</strong>gs as function of pressure <strong>and</strong> fluid velocity are programmed<br />

accord<strong>in</strong>g to the developed model. In addition, dedicated subrout<strong>in</strong>es are implemented<br />

to compute fluid properties such as pressure gradient <strong>in</strong>side the<br />

production tub<strong>in</strong>g, or flow rate, <strong>and</strong> impulse pump<strong>in</strong>g efficiency. Subrout<strong>in</strong>e<br />

names <strong>and</strong> short descriptions are listed <strong>in</strong> Table 8.1.<br />

Table 8.1: List of Subrout<strong>in</strong>es Included <strong>in</strong> the <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Software<br />

Subrout<strong>in</strong>e Name Description<br />

fluid Fluid Properties<br />

acoustic Acoustic Properties<br />

pulse <strong>Impulse</strong> Generator<br />

outlet Outlet Valve Condition<br />

<strong>in</strong>let Inlet Valve Condition<br />

flowgrad Inlet Flow Gradient<br />

claw1 CLAWPACK Ma<strong>in</strong> Rout<strong>in</strong>e<br />

attenuation <strong>Pressure</strong> Wave Attenuation<br />

rp1 Riemann Solver<br />

out1 Output Results<br />

flowrate <strong>Impulse</strong> Pump<strong>in</strong>g Flow Rate<br />

efficiency <strong>Impulse</strong> Pump Efficiency<br />

The ”rp1” rout<strong>in</strong>e is the Riemann solver <strong>in</strong>cluded <strong>in</strong> CLAWPACK <strong>and</strong> corresponds<br />

to pressure wave propagation modell<strong>in</strong>g while the ”attenuation” subrout<strong>in</strong>e<br />

<strong>in</strong>tegrates the source term. Subrout<strong>in</strong>es ”<strong>in</strong>let” <strong>and</strong> ”outlet” follow<br />

<strong>in</strong>let <strong>and</strong> outlet non-return valve work<strong>in</strong>gs, respectively. Subrout<strong>in</strong>es ”fluid”<br />

<strong>and</strong> ”acoustic” calculate pressure, density <strong>and</strong> acoustic velocity gradients <strong>in</strong>side<br />

the production tub<strong>in</strong>g. Subrout<strong>in</strong>es are called successively by the ma<strong>in</strong><br />

rout<strong>in</strong>e ”claw1” that <strong>in</strong>terfaces user-set parameters <strong>and</strong> numerical calculations.<br />

Simulated processes depend on user-set fluid properties, system configuration,<br />

application characteristics <strong>and</strong> impulse pump<strong>in</strong>g parameters. Fluid properties<br />

are <strong>in</strong>formed at impulse pump<strong>in</strong>g pressure <strong>and</strong> temperature conditions.<br />

Properties such as density, viscosity or heat capacities can be obta<strong>in</strong>ed from<br />

hydrocarbon molar composition us<strong>in</strong>g the Peng-Rob<strong>in</strong>son fluid package <strong>in</strong>cluded<br />

<strong>in</strong> Hysys. Isothermal compressibilities are calculated by the ratio of<br />

density <strong>and</strong> pressure gradients at constant temperature.<br />

137


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

System configuration <strong>in</strong>cludes pipel<strong>in</strong>e length, diameter, wall thickness <strong>and</strong><br />

roughness, material Young’s modulus, non-return valve locations, types <strong>and</strong><br />

crack<strong>in</strong>g pressures. Application characteristics are wellbore pressure <strong>and</strong> pressure<br />

loss coefficient, <strong>and</strong> well <strong>in</strong>dex productivity. In addition, spatial <strong>and</strong><br />

temporal meshes are user-set parameters of the numerical simulation tool.<br />

User-set parameters for the system <strong>and</strong> application description, <strong>and</strong> computational<br />

parameters, are listed <strong>in</strong> Table 8.2.<br />

Table 8.2: List of Parameters Describ<strong>in</strong>g <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Characteristic Variable Unit<br />

Annulus Fluid Level XF [m]<br />

Fluid Isothermal Compressibility KΓ [Pa −1 ]<br />

Fluid Density ρ [kg m −3 ]<br />

Fluid Heat Transfer Coefficient κ [W m −2 K −1 ]<br />

Fluid <strong>Pressure</strong> p [bara]<br />

Fluid Temperature Γ [ o C]<br />

Fluid Viscosity ν [ o C]<br />

<strong>Impulse</strong> Generator Frequency ˙γ [Hz]<br />

<strong>Impulse</strong> Generator Stroke ∆l [m]<br />

Inlet Valve Crack<strong>in</strong>g <strong>Pressure</strong> pCI [bar]<br />

Inlet Valve Open<strong>in</strong>g Time topen<strong>in</strong>let [s]<br />

Inlet Valve Type InletType<br />

Outlet Tank <strong>Pressure</strong> pOutlet [bara]<br />

Outlet Valve Crack<strong>in</strong>g <strong>Pressure</strong> pCO [bar]<br />

Outlet Valve Location X [m]<br />

Outlet Valve Open<strong>in</strong>g Time topenoutlet [s]<br />

Outlet Valve Type OutletType<br />

Pipel<strong>in</strong>e Diameter d [m]<br />

Pipel<strong>in</strong>e Length L [m]<br />

Pipel<strong>in</strong>e Material Young’s Modulus Y [Pa]<br />

Pipel<strong>in</strong>e Wall Roughness k<br />

Pipel<strong>in</strong>e Wall Thickness e [m]<br />

Wellbore <strong>Pressure</strong> pWB [bara]<br />

Wellbore <strong>Pressure</strong> Loss Coefficient kv<br />

Well Productivity Index PI<br />

<strong>Pressure</strong> wave propagation characteristics are def<strong>in</strong>ed from fluid properties<br />

<strong>and</strong> system configuration. Acoustic velocity <strong>in</strong> fluid-filled pipel<strong>in</strong>e is calculated<br />

from fluid <strong>and</strong> pipel<strong>in</strong>e properties us<strong>in</strong>g Eq. 8.10. Acoustic velocity<br />

dependence on pressure <strong>and</strong> temperature is considered by calculat<strong>in</strong>g pressure<br />

<strong>and</strong> temperature gradients <strong>in</strong> the production tub<strong>in</strong>g from gravity <strong>and</strong> heat<br />

138


8.3. Modell<strong>in</strong>g Results<br />

transfer, respectively. <strong>Pressure</strong> wave attenuation is calculated us<strong>in</strong>g steady<br />

friction factor <strong>in</strong>clud<strong>in</strong>g pipel<strong>in</strong>e wall roughness effects.<br />

Application characteristics <strong>and</strong> impulse pump<strong>in</strong>g parameters also affect the<br />

pump<strong>in</strong>g action. Application characteristics are wellbore pressure <strong>and</strong> pressure<br />

loss coefficient, <strong>and</strong> well productivity <strong>in</strong>dex. <strong>Impulse</strong> pump<strong>in</strong>g parameters<br />

are impulse generator frequency <strong>and</strong> stroke. Def<strong>in</strong><strong>in</strong>g parameters, application<br />

description <strong>and</strong> impulse pump<strong>in</strong>g characteristics are <strong>in</strong>put parameters of the<br />

simulation tool <strong>and</strong> are listed <strong>in</strong> Table 8.2. Simulation parameters are number<br />

of grid po<strong>in</strong>ts <strong>and</strong> simulation period of pressure wave propagation process.<br />

Valve actuation functions <strong>in</strong>vestigated <strong>in</strong> Chapter 3 are <strong>in</strong>cluded <strong>in</strong> subrout<strong>in</strong>es<br />

<strong>in</strong>let <strong>and</strong> outlet. However, a needle valve corresponds to the experimental setup.<br />

Simulation tool output parameters are pressure <strong>and</strong> fluid velocity along<br />

the pipel<strong>in</strong>e dur<strong>in</strong>g the simulated period. Such outputs allow for visualization<br />

of impulse pump<strong>in</strong>g process. The reachable flow rate is also an output of the<br />

program.<br />

8.3 Modell<strong>in</strong>g Results<br />

The impulse pump<strong>in</strong>g concept <strong>and</strong> modell<strong>in</strong>g can now be illustrated <strong>and</strong> <strong>in</strong>vestigated<br />

us<strong>in</strong>g the impulse pump<strong>in</strong>g simulation tool. Here<strong>in</strong> presented illustrations<br />

were obta<strong>in</strong>ed from simulations of a 100 meter long vertical liquid-waterfilled<br />

pipel<strong>in</strong>e, 1 <strong>in</strong>ch <strong>in</strong> diameter, <strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation.<br />

Simulated pressure waves are 2.5 Hz <strong>in</strong> frequency, <strong>and</strong> their amplitudes equal<br />

wellhead pressure. Liquid water density <strong>and</strong> acoustic velocity are 1000 kg m −3<br />

<strong>and</strong> 1400 ms −1 , respectively. <strong>Pressure</strong> waves are then 560 m <strong>in</strong> wavelength.<br />

Spatial traces of a s<strong>in</strong>gle pressure wave dur<strong>in</strong>g impulse pump<strong>in</strong>g action are<br />

synthetically illustrated <strong>in</strong> Figure 8.6. A left-to-right truncated s<strong>in</strong>usoidal<br />

pressure wave, 2 pWave <strong>in</strong> peak-to-peak amplitude before outlet non-return<br />

valve, propagates to the downhole closed <strong>in</strong>let non-return valve, Figure 8.6 (a).<br />

The <strong>in</strong>let-non-return valve is assumed to open as soon as the local pressure<br />

at the valve <strong>in</strong> the fluid-filled pipel<strong>in</strong>e becomes lower than the <strong>in</strong>-situ static<br />

pressure, represented by the 0 pressure l<strong>in</strong>e <strong>in</strong> Figure 8.6.<br />

The negative pressure pulse then reaches the boundary <strong>and</strong> decreases temporarily<br />

the local pressure at the <strong>in</strong>let non-return valve. The valve then opens<br />

<strong>and</strong> the negative pulse reflects with phase change from an open boundary<br />

condition. Another way to express this is that the negative pressure pulse<br />

becomes positive after reflection at the open <strong>in</strong>let non-return valve. The negative<br />

pressure pulse then superimposes on its own phase-changed reflection <strong>and</strong><br />

cancels it temporarily, Figure 8.6 (b).<br />

139


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

The <strong>in</strong>let non-return valve closes as soon as the local pressure <strong>in</strong> the fluidfilled<br />

pipel<strong>in</strong>e equals the <strong>in</strong>-situ static pressure. Another way to express this is<br />

that the <strong>in</strong>let non-return valve closes after total reflection of the negative pressure<br />

pulse. The right-to-left mov<strong>in</strong>g reflected phase-changed negative pressure<br />

pulse then superimposes on the left-to-right mov<strong>in</strong>g positive pressure pulse,<br />

Figure 8.6(c). The local pressure amplitude is then temporarily the sum of<br />

the two pressure pulses amplitudes. For the illustrated case, the pressure<br />

amplitude is temporarily about twice the orig<strong>in</strong>al pressure pulse amplitude.<br />

The positive pressure pulse then reflects totally at the closed <strong>in</strong>let non-return<br />

valve; that is, without phase change. Another way to express this is that the<br />

positive pressure pulse reflects as a positive pressure pulse of same amplitude.<br />

After complete reflection at the closed <strong>in</strong>let non-return valve, the right-to-left<br />

mov<strong>in</strong>g pressure wave then propagates from downhole back to wellhead, Figure<br />

8.6 (d). The pump<strong>in</strong>g action occurs only while the <strong>in</strong>let non-return valve is<br />

open. Thus, impulse pump<strong>in</strong>g <strong>in</strong>duces pulsat<strong>in</strong>g flows.<br />

The water hammer pressure wave due to the <strong>in</strong>let non-return valve open<strong>in</strong>g<br />

was not represented for simplification. The water hammer effect can nonetheless<br />

be considered as illustrated <strong>in</strong> Figure 8.7. Maximum pressure waves when<br />

consider<strong>in</strong>g <strong>and</strong> neglect<strong>in</strong>g the water hammer effect are obta<strong>in</strong>ed at the phasechanged<br />

reflected negative pressure pulse, <strong>and</strong> are 1.12 pWave <strong>and</strong> pWave, respectively.<br />

The water hammer pressure wave then changes the phase-changed<br />

reflected negative pressure pulse waveform <strong>and</strong> <strong>in</strong>creases its amplitude.<br />

Wellbore <strong>and</strong> downhole pressures were first assumed to be equal, <strong>and</strong> the <strong>in</strong>let<br />

non-return valve was assumed to open <strong>in</strong>stantaneously without mechanical<br />

resistance for simplification. However, the valve open<strong>in</strong>g pressure can be lower<br />

than the downhole pressure. As the negative pulse reflects without phase<br />

changed at a closed <strong>in</strong>let non-return valve, onto itself, to decrease the local<br />

pressure further, there comes a po<strong>in</strong>t where the downhole pressure is less than<br />

the wellbore pressure. The <strong>in</strong>let non-return valve will then open <strong>and</strong> fluid will<br />

flow <strong>in</strong>to the pipel<strong>in</strong>e.<br />

Influence of wellbore static pressure on the spatial trace of a s<strong>in</strong>gle pressure<br />

wave after reflection at the <strong>in</strong>let non-return valve is illustrated <strong>in</strong> Figure 8.8.<br />

Maximum amplitudes, reported from the partially phase-changed reflected<br />

negative pressure pulse, are 1.12 pWave, 0.75 pWave, 0.25 pWave <strong>and</strong> -0.25<br />

pWave, for 0, 0.5 pWave, pWave <strong>and</strong> 1.5 pWave pressure differences between<br />

downhole <strong>and</strong> wellbore, respectively. The lower the wellbore static pressure,<br />

the less time the <strong>in</strong>let non-return valve opens <strong>and</strong> the less fluid flows <strong>in</strong>to the<br />

production tub<strong>in</strong>g.<br />

140


(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

(p − p 0 ) / p Wave<br />

a)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

8.3. Modell<strong>in</strong>g Results<br />

−1.0<br />

0 0.2 0.4 0.6 0.8 1<br />

b)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

−1.0<br />

0 0.2 0.4 0.6 0.8 1<br />

c)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

−1.0<br />

0 0.2 0.4 0.6 0.8 1<br />

d)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

−1.0<br />

0 0.2 0.4 0.6 0.8 1<br />

Distance / Length<br />

Figure 8.6: Spatial <strong>Pressure</strong> Traces at Different Instants Dur<strong>in</strong>g<br />

<strong>Impulse</strong> Pump<strong>in</strong>g Process (In-Situ Static <strong>Pressure</strong> is<br />

Represented by 0 <strong>Pressure</strong> L<strong>in</strong>e). Data Presented <strong>in</strong><br />

Dimensionless Form.<br />

a): Left-to-Right Mov<strong>in</strong>g Truncated S<strong>in</strong>usoidal <strong>Pressure</strong> Wave.<br />

b): Superposition of Negative <strong>Pressure</strong> Pulse with its Own Phase-Changed<br />

Reflection at Open Valve.<br />

c): Superposition of Positive Pulse with Phase-Changed Reflected <strong>Pressure</strong><br />

Pulse.<br />

d): Right-to-Left Mov<strong>in</strong>g Truncated Partially Phase-Changed <strong>Pressure</strong><br />

Wave.<br />

141


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

(p − p 0 ) / p Wave<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

−0.25<br />

−0.50<br />

−0.75<br />

−1.00<br />

With Water Hammer<br />

No Water Hammer<br />

−1.25<br />

0 0.2 0.4 0.6 0.8 1<br />

Distance / Length<br />

Figure 8.7: Influence of Water Hammer <strong>Pressure</strong> Wave on Synthetic<br />

<strong>Pressure</strong> Spatial Trace<br />

(p − p 0 ) / p Wave<br />

1.25<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0.00<br />

−0.25<br />

−0.50<br />

−0.75<br />

p Open<strong>in</strong>g = 0<br />

p Open<strong>in</strong>g = −0.5 p Wave<br />

p Open<strong>in</strong>g = −1.0 p Wave<br />

−1.00<br />

p = −1.5 p<br />

Open<strong>in</strong>g Wave<br />

p = −2.0 p<br />

Open<strong>in</strong>g Wave<br />

−1.25<br />

0 0.2 0.4 0.6 0.8 1<br />

Distance / Length<br />

Figure 8.8: Influence of Wellbore <strong>Pressure</strong> on Synthetic <strong>Pressure</strong><br />

Spatial Trace<br />

142


8.4. Validation of <strong>Impulse</strong> Pump<strong>in</strong>g Model <strong>and</strong> Simulation Tool<br />

Experimental data are temporal traces from pressure transducers at particular<br />

locations. Synthetic temporal pressure traces at the <strong>in</strong>let non-return valve<br />

for two different open<strong>in</strong>g pressures are shown <strong>in</strong> Figure 8.9. Maximum pressure<br />

amplitudes for both <strong>in</strong>let non-return valve open<strong>in</strong>g pressures are 2 pWave,<br />

whereas m<strong>in</strong>imum pressure amplitude is approximately downhole pressure <strong>and</strong><br />

zero absolute pressure for 0 <strong>and</strong> pWave open<strong>in</strong>g pressures, respectively. Another<br />

way to express this is that m<strong>in</strong>imum pressure amplitudes are downhole<br />

pressure decreased by open<strong>in</strong>g pressures.<br />

(p − p 0 ) / p Wave<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

−0.5<br />

−1.0<br />

p Open<strong>in</strong>g = 0<br />

−1.5<br />

0 0.2 0.4 0.6<br />

p = −1 p<br />

Open<strong>in</strong>g Wave<br />

0.8 1<br />

t / T<br />

0<br />

Figure 8.9: Synthetic Temporal <strong>Pressure</strong> Trace at Downhole Inlet<br />

Non-Return Valve<br />

8.4 Validation of <strong>Impulse</strong> Pump<strong>in</strong>g Model <strong>and</strong><br />

Simulation Tool<br />

Model validation was performed by compar<strong>in</strong>g experimental pressure <strong>and</strong> flow<br />

rate measurements with results from numerical simulations. Synthetic temporal<br />

pressure traces were computed at the <strong>in</strong>let non-return valve <strong>and</strong> compared<br />

with experimental data. The experimental set-up was a horizontal steel<br />

pipel<strong>in</strong>e, 283 meter <strong>in</strong> length, 1 <strong>in</strong>ch <strong>in</strong> diameter, <strong>and</strong> 1 mm <strong>in</strong> wall thickness.<br />

Experimental fluid was tap water at atmospheric temperature, that conta<strong>in</strong>ed<br />

a small amount of air bubbles. Therefore, density <strong>and</strong> acoustic velocity <strong>in</strong> the<br />

fluid-filled pipel<strong>in</strong>e are assumed 1000 kg m −3 , <strong>and</strong> 1200 ms −1 , respectively.<br />

143


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

In-situ static pressure <strong>in</strong> the fluid-filled pipel<strong>in</strong>e <strong>and</strong> the <strong>in</strong>let tank were assumed<br />

3 bara <strong>and</strong> 1 bara, respectively. <strong>Impulse</strong> generator frequency was assumed<br />

1.5 Hz. The outlet non-return valve crack<strong>in</strong>g pressure was assumed<br />

4.2 bara. Several simulations with different <strong>in</strong>let non-return valve crack<strong>in</strong>g<br />

pressure <strong>and</strong> pressure wave peak-to-peak amplitude are used to model experimental<br />

uncerta<strong>in</strong>ties related to these two parameters. Steady frictional<br />

pressure loss is the only pressure wave attenuation mechanism <strong>in</strong>cluded <strong>in</strong> the<br />

modell<strong>in</strong>g <strong>and</strong> the simulations. Numerical simulations are run for one second<br />

of a s<strong>in</strong>gle pressure wave propagation.<br />

<strong>Pressure</strong> values are calculated at grid po<strong>in</strong>ts <strong>and</strong> the temporal trace at the <strong>in</strong>let<br />

non-return valve is shown <strong>in</strong> Figure 8.10 for 0.2 bar <strong>in</strong>let non-return crack<strong>in</strong>g<br />

pressure, <strong>and</strong> 4.5 bar pressure wave peak-to-peak amplitude. A maximum<br />

pressure value of 14 bara is observed. The pressure peak is due to superposition<br />

of multiple pressure wave; namely, the positive pressure pulse superposition<br />

with its own reflection, <strong>and</strong> the phase-changed reflected negative pressure pulse<br />

reflected at the closed boundary impulse generator.<br />

<strong>Pressure</strong> [bara]<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

−2<br />

0 0.2 0.4 0.6 0.8 1<br />

Time [s]<br />

Figure 8.10: Synthetic Temporal <strong>Pressure</strong> Trace at Downhole Inlet<br />

Non-Return Valve. ˙γ = 1.5 Hz. L = 283m<br />

144


8.4. Validation of <strong>Impulse</strong> Pump<strong>in</strong>g Model <strong>and</strong> Simulation Tool<br />

Synthetic maximum pressure values are overestimated due to the lack of accurate<br />

modell<strong>in</strong>g of pressure wave attenuation. In addition, more oscillations are<br />

expected at the maximum pressure peak due to the outlet non-return valve<br />

oscillations. However, such oscillations were idealized <strong>in</strong> the modell<strong>in</strong>g. Nevertheless,<br />

synthetic pressure data are <strong>in</strong> good agreement with experimental<br />

pressure signals, Figure 7.11. <strong>Pressure</strong> traces are therefore correctly predicted<br />

by the impulse pump<strong>in</strong>g model <strong>and</strong> the numerical simulation tool developed.<br />

Computed flow rates for the various <strong>in</strong>let non-return valve crack<strong>in</strong>g pressure<br />

<strong>and</strong> pressure wave amplitude conditions are shown <strong>in</strong> Table 8.3 <strong>and</strong> illustrated<br />

<strong>in</strong> Figure 8.11. The reachable flow rate depends greatly on both the <strong>in</strong>let nonreturn<br />

valve crack<strong>in</strong>g pressure <strong>and</strong> the pressure wave peak-to-peak amplitude.<br />

The maximum flow rate of 2.97 l m<strong>in</strong> −1 is reached for a 4.5 bar pressure wave<br />

peak-to-peak amplitude, <strong>and</strong> 0.4 bar <strong>in</strong>let non-return valve crack<strong>in</strong>g pressure.<br />

Flow rate fluctuations are expla<strong>in</strong>ed by the water hammer pressure wave generated<br />

dur<strong>in</strong>g <strong>in</strong>let non-return valve open<strong>in</strong>g.<br />

The water hammer pressure wave <strong>in</strong>duced by <strong>in</strong>let non-return valve open<strong>in</strong>g<br />

counteracts the negative pressure pulse open<strong>in</strong>g action. A low amplitude water<br />

hammer is then required to limit pressure oscillations at the <strong>in</strong>let non-return<br />

valve. However, a low flow rate arises from a low amplitude water hammer<br />

pressure wave. Therefore a compromise must be reached to maximize the<br />

reachable flow rate dur<strong>in</strong>g impulse pump<strong>in</strong>g. Such compromise must be analysed<br />

for every <strong>in</strong>dividual applications.<br />

The computed flow rate varies between 2 <strong>and</strong> 3 l m<strong>in</strong> −1 for the <strong>in</strong>let non-return<br />

valve crack<strong>in</strong>g pressures <strong>and</strong> pressure wave peak-to-peak amplitude simulated.<br />

Simulated results are therefore <strong>in</strong> the same range as experimental flow rate<br />

between 1 <strong>and</strong> 4 l m<strong>in</strong> −1 for the same impulse generator frequency, Figure<br />

7.17. Thus, impulse pump<strong>in</strong>g is correctly modelled, <strong>and</strong> simulated results are<br />

close to experimental measurements.<br />

145


Chapter 8. Concept, Modell<strong>in</strong>g <strong>and</strong> Simulation of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Table 8.3: Computed Flow Rate <strong>in</strong> [l m<strong>in</strong> −1 ], for Different Inlet<br />

Non-Return Valve Crack<strong>in</strong>g <strong>Pressure</strong>s <strong>and</strong> <strong>Pressure</strong> Wave<br />

Peak-To-Peak Amplitudes<br />

Flow Rate [l m<strong>in</strong> −1 ]<br />

3.00<br />

2.75<br />

2.50<br />

2.25<br />

pCI[bar]<br />

2pWave [bar]<br />

4.5 5.0 5.5<br />

0.1 2.19 2.53 2.34<br />

0.2 2.52 2.47 2.38<br />

0.4 2.97 2.47 2.63<br />

p CI = 0.1 bar<br />

p CI = 0.2 bar<br />

p CI = 0.4 bar<br />

2.0<br />

4.50 4.75 5.00<br />

p [bar]<br />

Wave<br />

5.25 5.50<br />

Figure 8.11: Synthetic Flow Rate for Diverse Inlet Non-Return Valve<br />

Crack<strong>in</strong>g <strong>Pressure</strong>s <strong>and</strong> <strong>Pressure</strong> Wave Peak-To-Peak<br />

Amplitudes. ˙γ = 1.5 Hz. L = 283m<br />

146


Chapter 9<br />

Design, Applications <strong>and</strong><br />

Performance of <strong>Impulse</strong><br />

Pump<strong>in</strong>g<br />

9.1 Design<br />

9.1.1 <strong>Impulse</strong> Generator<br />

<strong>Impulse</strong> pump<strong>in</strong>g is based on pressure wave propagation <strong>in</strong>duced by back<strong>and</strong>-forth<br />

oscillations of an impulse generator that can be modelled by piston<br />

displacements. A direct relation can therefore be derived between piston displacements<br />

<strong>and</strong> pressure wave amplitude (Eq. 8.6). A second equation can<br />

also relate piston velocity <strong>and</strong> pressure wave amplitude (Eq. 8.7). <strong>Impulse</strong><br />

generator characteristics can consequently be derived theoretically accord<strong>in</strong>g<br />

to impulse pump<strong>in</strong>g conditions.<br />

Evolutions of piston stroke for generat<strong>in</strong>g pressure changes, 100 bar <strong>in</strong> amplitude,<br />

that is a 200 bar peak-to-peak pressure wave, <strong>in</strong> 1000 m long water-filled<br />

<strong>and</strong> oil-filled pipel<strong>in</strong>es are shown <strong>in</strong> Figure 9.1 aga<strong>in</strong>st ratio of piston head<br />

<strong>and</strong> pipel<strong>in</strong>e diameters (dp/d <strong>in</strong> Figure 8.2). Investigated water <strong>and</strong> oil compressibilities<br />

are 0.435 10 −9 Pa −1 <strong>and</strong> 1.98 10 −9 Pa −1 , respectively (McCa<strong>in</strong>,<br />

1990). Calculated strokes <strong>in</strong> water-filled pipel<strong>in</strong>e decrease from 4.35 m, to<br />

0.27 m <strong>and</strong> to 6.8 10 −3 m for dp/d ratios equal to 1, 4 <strong>and</strong> 8, respectively.<br />

Calculated strokes <strong>in</strong> an oil-filled pipel<strong>in</strong>e decrease from 19.8 m, to 1.24 m <strong>and</strong><br />

to 0.31 m for dp/d ratios equal to 1, 4 <strong>and</strong> 8, respectively.<br />

Evolutions of piston l<strong>in</strong>ear velocities <strong>in</strong> a water-filled pipel<strong>in</strong>e for piston frequencies<br />

between 1 Hz <strong>and</strong> 8 Hz are shown <strong>in</strong> Figure 9.2. Velocity decreases<br />

with <strong>in</strong>crease <strong>in</strong> ratio dp/d. At 1 Hz, velocity decreases from 17.4 ms −1 to<br />

1.1 ms −1 , <strong>and</strong> to 0.27 ms −1 for dp/d ratios equal to 1, 4 <strong>and</strong> 8, respectively.<br />

147


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Stroke [m]<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 −1<br />

Oil<br />

Water<br />

10<br />

0 2 4 6 8 10<br />

−2<br />

d / d<br />

p<br />

Figure 9.1: Evolution of Piston Stroke for Generat<strong>in</strong>g <strong>Pressure</strong> Wave<br />

100 bar <strong>in</strong> Amplitude <strong>in</strong> a 1000 m Long Pipel<strong>in</strong>e<br />

Piston velocity <strong>in</strong>creases with frequency. For a ratio dp/d equal to 4, velocity<br />

<strong>in</strong>creases from 1.1 ms −1 , to 2.2 ms −1 , <strong>and</strong> to 8.7 ms −1 for 1 Hz, 2 Hz, <strong>and</strong><br />

8 Hz frequencies, respectively.<br />

Evolutions of piston l<strong>in</strong>ear velocities <strong>in</strong> an oil-filled pipel<strong>in</strong>e for piston frequencies<br />

between 1 Hz <strong>and</strong> 8 Hz are shown <strong>in</strong> Figure 9.3. Velocity decreases with<br />

<strong>in</strong>crease <strong>in</strong> ratio dp/d. At 1 Hz, velocity decreases from 79.2 ms −1 to 4.95<br />

ms −1 , <strong>and</strong> to 1.24 ms −1 for dp/d ratios equal to 1, 4 <strong>and</strong> 8, respectively.<br />

Piston velocity <strong>in</strong>creases with frequency. For a ratio dp/d equal to 4, velocity<br />

<strong>in</strong>creases from 4.95 ms −1 , to 9.9 ms −1 , <strong>and</strong> to 39.6 ms −1 for 1 Hz, 2 Hz,<br />

<strong>and</strong> 8 Hz frequencies, respectively.<br />

Generation of pressure wave us<strong>in</strong>g pistons is limited due to reachable strokes<br />

<strong>and</strong> l<strong>in</strong>ear velocities. Low frequency pressure waves require lower piston velocity.<br />

Great dp/d ratios also <strong>in</strong>duce shorter strokes <strong>and</strong> lower piston velocities.<br />

148


Piston Velocity [m s −1 ]<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

8 Hz<br />

4 Hz<br />

2 Hz<br />

1 Hz<br />

10<br />

0 2 4 6 8 10<br />

−1<br />

d / d<br />

p<br />

Figure 9.2: Piston L<strong>in</strong>ear Velocity for Generat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong><br />

100 bar <strong>in</strong> Amplitude <strong>in</strong> a 1000 m Long Water-Filled<br />

Pipel<strong>in</strong>e at Frequencies From 1 to 8 Hz<br />

Piston Velocity [m s −1 ]<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

8 Hz<br />

4 Hz<br />

2 Hz<br />

1 Hz<br />

10<br />

0 2 4 6 8 10<br />

−1<br />

d / d<br />

p<br />

Figure 9.3: Piston L<strong>in</strong>ear Velocity for Generat<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong><br />

100 bar <strong>in</strong> Amplitude <strong>in</strong> a 1000 m Long Oil-Filled<br />

Pipel<strong>in</strong>e at Frequencies From 1 to 8 Hz<br />

149<br />

9.1. Design


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

9.1.2 Non-Return Valves<br />

Reachable flow rates theoretically depend on the <strong>in</strong>let non-return valve type.<br />

Check valves actuation functions <strong>and</strong> flow rate evolutions with actuation are<br />

detailed <strong>in</strong> Chapter 3 <strong>and</strong> are used here<strong>in</strong> to evaluate flow rates aga<strong>in</strong>st nonreturn<br />

valve type. Needle valves are the only non-return valves <strong>in</strong>vestigated<br />

<strong>in</strong> Chapter 3. However, a flap valve with the same actuation function as a<br />

butterfly valve, Figure 3.7 e), but with a different mechanical design is also<br />

<strong>in</strong>vestigated.<br />

Reachable flow rates are calculated us<strong>in</strong>g needle <strong>and</strong> butterfly non-return valve<br />

types. Calculations are based on numerical simulations of impulse pump<strong>in</strong>g of<br />

water from a 1000 m deep wellbore at 75 bara constant pressure (correspond<strong>in</strong>g<br />

to 235 m lift<strong>in</strong>g height) us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 5 Hz <strong>in</strong> frequency, 200<br />

bar <strong>in</strong> peak-to-peak amplitude (Table 9.1). No noticeable differences can be<br />

observed between numerical simulation with needle valve model <strong>and</strong> butterfly<br />

valve model. Reachable flow rates depend neither on the <strong>in</strong>let non-return<br />

valve type or time for full open<strong>in</strong>g, nor on the <strong>in</strong>let non-return valve actuation<br />

pattern, uniform or accelerated.<br />

Table 9.1: Description of Test Case for Numerical Investigation of<br />

<strong>Pressure</strong> Waveform Impact on Reachable Flow Rate<br />

L [m] d [<strong>in</strong>ch] pWH [bara] pWB [bara] pWave [bara]<br />

1000 10 100 75 100<br />

˙γ [Hz] TSim [s] a ρ Valve Type<br />

5 1.5 1200 1000 Needle<br />

nX<br />

nT<br />

10000 1000<br />

Practically, reachable flow rates do not depend on <strong>in</strong>let non-return valve type.<br />

Nevertheless, <strong>in</strong>let <strong>and</strong> outlet non-return valve types can be selected based<br />

on mechanical efficiency. Non-return valve mechanical efficiency can be calculated<br />

us<strong>in</strong>g actuation functions <strong>and</strong> flow rate evolutions studied <strong>in</strong> Chapter 3,<br />

by compar<strong>in</strong>g flow evolutions dur<strong>in</strong>g real <strong>and</strong> <strong>in</strong>stantaneous valve actuation.<br />

Mechanical efficiency calculations here<strong>in</strong> assume symmetry between open<strong>in</strong>g<br />

<strong>and</strong> closure functions, no mechanical resistance <strong>and</strong> same actuation time for<br />

every non-return valves <strong>and</strong> also check-valves <strong>in</strong>vestigated.<br />

Mechanical efficiencies for uniform <strong>and</strong> accelerated actuations for the six checkvalves<br />

considered are shown <strong>in</strong> Figure 9.4. Uniform actuation efficiency is<br />

approximately 2 % greater than accelerated actuation for every check-valve<br />

<strong>in</strong>vestigated. Mechanical efficiencies for uniform actuation are 91.3 %, 87.6 %,<br />

84.6 %, 80.6 %, 76.8 %, <strong>and</strong> 68.4 % for needle valve, circular gate valve, globe<br />

150


Efficiency [%]<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Uniform<br />

Accelerated<br />

Needle Circular Globe Square Ball Butterfly<br />

Valve Type<br />

9.1. Design<br />

Figure 9.4: Comparison of Check-Valves Efficiencies for Uniform <strong>and</strong><br />

Accelerated Actuations<br />

valve, square gate valve, ball valve <strong>and</strong> butterfly valve, respectively. For nonreturn<br />

valves only, needle valves are more efficient mechanically than butterfly<br />

valves.<br />

<strong>Impulse</strong> pump<strong>in</strong>g is theoretically applicable for h<strong>and</strong>l<strong>in</strong>g difficult fluids such<br />

as highly viscous or non-Newtonian fluids. Non-return valve open<strong>in</strong>g is theoretically<br />

slowed down by fluid viscosity. However, non-return valve closure is<br />

also slowed down by fluid viscosity s<strong>in</strong>ce non-return valve actuation is here<strong>in</strong><br />

assumed symmetric. Highly viscous fluids are therefore assumed h<strong>and</strong>led as<br />

st<strong>and</strong>ard fluids <strong>in</strong> impulse pump<strong>in</strong>g applications.<br />

151


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

9.2 Multiphase Fluid Flow<br />

<strong>Impulse</strong> pump<strong>in</strong>g is theoretically applicable to s<strong>in</strong>gle phase <strong>and</strong> two-phase<br />

flows; that is gas-liquid, <strong>and</strong> liquid-liquid. Gas-Liquid two-phase flows are<br />

here<strong>in</strong> characterized by small amount of gas <strong>in</strong> liquid that decrease pressure<br />

wave propagation velocity <strong>and</strong> <strong>in</strong>crease pressure wave attenuation, as compared<br />

to s<strong>in</strong>gle phase fluids. <strong>Pressure</strong> wave propagation velocity can be calculated<br />

by (Dong <strong>and</strong> Gudmundsson, 1993):<br />

a =<br />

⎛<br />

⎜<br />

⎝<br />

(εGcpG + (1 − εG)(εLcpW + (1 − εL)cpO))<br />

(εGcvG + (1 − εG)(εLcvW + (1 − εL)cvO))<br />

(ǫGρG + (1 − ǫG)(ǫLρW + (1 − ǫL)ρO))<br />

(ǫGKΓG + (1 − ǫG) (ǫLKΓW + (1 − ǫL)KΓO))<br />

⎞<br />

⎟<br />

⎠<br />

1/2<br />

(9.1)<br />

where, εG, def<strong>in</strong>es gas-liquid mass fraction, εL, water-oil mass fraction, ǫG,<br />

gas-liquid volumetric fraction, ǫL, water-oil volumetric fraction, ρ, density,<br />

KΓ, isothermal compressibility, cp <strong>and</strong> cv, heat capacity at constant pressure<br />

<strong>and</strong> constant volume, respectively, <strong>and</strong> <strong>in</strong>dexes G, W <strong>and</strong> O, gas, water <strong>and</strong><br />

oil, respectively.<br />

Evolution of acoustic velocity with void fraction is illustrated for a waterair<br />

mixture at 1 bara, 10 bara <strong>and</strong> 100 bara <strong>in</strong> Figure 9.5. Acoustic velocity<br />

decreases dramatically for small amount of gas bubbles present <strong>in</strong> the mixture.<br />

At 1 bara, acoustic velocity decreases from 1516 ms −1 to 141 ms −1 <strong>and</strong> to<br />

52 ms −1 for 0, 0.01 <strong>and</strong> 0.08 void fractions. Acoustic velocity <strong>in</strong>creases with<br />

pressure. At 0.5 void fraction, air-water mixture acoustic velocity <strong>in</strong>creases<br />

from 28.2 ms −1 to 73.9 ms −1 , <strong>and</strong> to 223.7 ms −1 for 1 bara, 10 bara <strong>and</strong><br />

100 bara, respectively.<br />

Small amount of gas bubbles can generate cavitation that can be detrimental<br />

<strong>and</strong> damage irreversibly equipments such as non-return valves. Cavitation<br />

should therefore be avoided dur<strong>in</strong>g impulse pump<strong>in</strong>g. Above bubble po<strong>in</strong>t<br />

pressures, hydrocarbon fluids are s<strong>in</strong>gle phase <strong>in</strong> nature. Thus, local pressures<br />

<strong>in</strong>side the pipel<strong>in</strong>e should not be lowered below bubble po<strong>in</strong>t pressure dur<strong>in</strong>g<br />

impulse pump<strong>in</strong>g, particularly dur<strong>in</strong>g propagation of the negative pulse of a<br />

pressure wave, or dur<strong>in</strong>g reflection of the negative part at a closed <strong>in</strong>let nonreturn<br />

valve. Non-return valves are idealized <strong>and</strong> Net Positive Suction Head<br />

(NPSH) is not considered.<br />

The phase envelope correspond<strong>in</strong>g to the volatile oil composition given <strong>in</strong> Table<br />

4.2 is illustrated <strong>in</strong> Figure 4.8. The phase envelope correspond<strong>in</strong>g to the black<br />

oil composition given <strong>in</strong> Table 4.3 is illustrated <strong>in</strong> Figure 4.12. At 100 o C,<br />

bubble po<strong>in</strong>t pressures are 240 bara <strong>and</strong> 145 bara for volatile <strong>and</strong> black oil,<br />

152


9.3. Horizontal Transport of Fluids<br />

respectively. Ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g local pressure <strong>in</strong>side a pipel<strong>in</strong>e above bubble po<strong>in</strong>t<br />

pressure ensures s<strong>in</strong>gle phase flow without cavitation.<br />

<strong>Impulse</strong> pump<strong>in</strong>g of gas-liquid two-phase flow can be managed without caus<strong>in</strong>g<br />

irreversible damages by ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g <strong>in</strong>side local pressures above bubble po<strong>in</strong>t<br />

pressures. Water is easier than volatile or black oils to h<strong>and</strong>le s<strong>in</strong>ce bubble<br />

po<strong>in</strong>t pressure is below atmospheric pressure.<br />

<strong>Impulse</strong> pump<strong>in</strong>g of liquid-liquid two-phase flows can also be obta<strong>in</strong>ed without<br />

caus<strong>in</strong>g irreversible damages to the production-tub<strong>in</strong>g.<br />

Acoustic Velocity [m s −1 ]<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

100 bara<br />

10 bara<br />

1 bara<br />

0<br />

0 0.2 0.4 0.6 0.8 1<br />

Void Fraction<br />

Figure 9.5: Water Air Mixture Acoustic Velocity Aga<strong>in</strong>st Void<br />

Fraction at 1 bara, 10 bara <strong>and</strong> 100 bara<br />

9.3 Horizontal Transport of Fluids<br />

9.3.1 In-Situ <strong>Pressure</strong> <strong>and</strong> <strong>Pressure</strong> Wave Amplitude<br />

Horizontal or quasi horizontal transport of hydrocarbons <strong>and</strong> water are common<br />

<strong>in</strong> <strong>in</strong>dustry. Pipel<strong>in</strong>e pressures must be kept above fluid bubble po<strong>in</strong>t<br />

pressures <strong>in</strong> order to avoid cavitation dur<strong>in</strong>g impulse pump<strong>in</strong>g of multiphase<br />

fluids (gas - liquid). In-situ local pressure must consequently be selected for<br />

local pressure not to be lowered further than bubble po<strong>in</strong>t pressures, even dur<strong>in</strong>g<br />

propagation of negative pressure pulses, or dur<strong>in</strong>g reflection of negative<br />

pressure pulses at a closed <strong>in</strong>let non-return valve.<br />

153


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Reachable m<strong>in</strong>imum pressures <strong>in</strong> a horizontal pipel<strong>in</strong>e dur<strong>in</strong>g impulse pump<strong>in</strong>g<br />

can take place at the <strong>in</strong>let non-return valve dur<strong>in</strong>g reflection of a negative<br />

pressure pulse at a closed boundary. For pressure waves 2 pWave <strong>in</strong> peakto-peak<br />

amplitude, maximum decrease <strong>in</strong> pressure dur<strong>in</strong>g superposition of<br />

a negative pressure pulse on its own reflection equals 2 pWave; that is, the<br />

pressure wave amplitude. The m<strong>in</strong>imum local pressure then becomes:<br />

pm<strong>in</strong>IN = p0 (L) − 2pWave<br />

(9.2)<br />

where, p0 (L), is <strong>in</strong>-situ local static pressure at the <strong>in</strong>let non-return valve. The<br />

m<strong>in</strong>imum local pressure at the <strong>in</strong>let non-return valve ensures no flow from the<br />

<strong>in</strong>let <strong>in</strong>side the production tub<strong>in</strong>g. Another way to express this is that the<br />

m<strong>in</strong>imum pressure will never be reached if pump<strong>in</strong>g of liquids is obta<strong>in</strong>ed.<br />

In-situ local pressure <strong>and</strong> pressure wave amplitude must therefore be selected<br />

to ensure <strong>in</strong>let non-return valve open<strong>in</strong>g <strong>and</strong> avoid cavitation <strong>in</strong>side the horizontal<br />

pipel<strong>in</strong>e. For <strong>in</strong>stance, at 100 o C, horizontal impulse pump<strong>in</strong>g of volatile<br />

oil can be obta<strong>in</strong>ed with pipel<strong>in</strong>e <strong>in</strong>-situ static pressure of 300 bara, pressure<br />

wave peak-to-peak amplitude of 120 bar, <strong>and</strong> <strong>in</strong>let non-return valve open<strong>in</strong>g<br />

pressure greater than 240 bara. At 100 o C, horizontal impulse pump<strong>in</strong>g of<br />

black oil can be obta<strong>in</strong>ed with pipel<strong>in</strong>e <strong>in</strong>-situ static pressure of 300 bara,<br />

pressure wave peak-to-peak amplitude of 310 bar, <strong>and</strong> <strong>in</strong>let non-return valve<br />

open<strong>in</strong>g pressure greater than 145 bara.<br />

Horizontal impulse pump<strong>in</strong>g of water can be obta<strong>in</strong>ed for almost any pipel<strong>in</strong>e<br />

<strong>in</strong>-situ static pressure <strong>and</strong> pressure wave peak-to-peak amplitudes lower than<br />

two times the <strong>in</strong>-situ static pressure <strong>in</strong> the pipel<strong>in</strong>e.<br />

9.3.2 Efficiency<br />

<strong>Impulse</strong> pump<strong>in</strong>g efficiency <strong>in</strong> horizontal pipel<strong>in</strong>e or quasi horizontal pipel<strong>in</strong>e<br />

can be calculated directly from impulse generator characteristics <strong>and</strong> reachable<br />

flow rates. The impulse generator is the only part of an impulse pump<strong>in</strong>g apparatus<br />

that requires external energy supply. Hence, impulse generator characteristics<br />

<strong>and</strong> generated pressure wave amplitude def<strong>in</strong>e the required power.<br />

Modell<strong>in</strong>g an impulse generator by a back <strong>and</strong> forth oscillat<strong>in</strong>g piston, the<br />

required power can be calculated by:<br />

PRequired = FPistonuPiston<br />

(9.3)<br />

where, FPiston, represents force applied on the mov<strong>in</strong>g piston, <strong>and</strong>, uPiston,<br />

piston l<strong>in</strong>ear velocity. The force on the piston dur<strong>in</strong>g one cycle directly relates<br />

to the pressure wave amplitude:<br />

FPiston = πd 2 ppWave<br />

154<br />

(9.4)


9.3. Horizontal Transport of Fluids<br />

Us<strong>in</strong>g the piston velocity calculated <strong>in</strong> Eq. 8.7, Eq. 9.3 becomes:<br />

PR = 4πKL˙γ (d pWave) 2<br />

(9.5)<br />

where, L, is pipel<strong>in</strong>e length, ˙γ, piston oscillation frequency, <strong>and</strong>, d, pipel<strong>in</strong>e<br />

diameter. Power generated by impulse pump<strong>in</strong>g can be calculated by reachable<br />

flow rate <strong>and</strong> frictional pressure drop <strong>in</strong> the pipel<strong>in</strong>e.<br />

Generated power def<strong>in</strong>ition is then:<br />

PG = qpf<br />

(9.6)<br />

where, q, characterizes reachable flow rate, <strong>and</strong>, pf, frictional pressure drop<br />

def<strong>in</strong>ed by:<br />

pf = 1 L<br />

f<br />

2 d ρu2<br />

(9.7)<br />

where, f, is friction factor. Power generated at the <strong>in</strong>let non-return valve by<br />

impulse pump<strong>in</strong>g can then be simplified as:<br />

PG<strong>in</strong> = 8 L<br />

π2f d5ρq3 (9.8)<br />

<strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiency for horizontal or quasi horizontal pipel<strong>in</strong>es is<br />

the ratio of generated <strong>and</strong> required powers. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiency<br />

can after simplifications be expressed by:<br />

η = PG<strong>in</strong><br />

PR<br />

= 2<br />

π 3<br />

ρfq 3<br />

K ˙γd 7 (pWave) 2<br />

(9.9)<br />

Inflow efficiency of impulse pump<strong>in</strong>g for horizontal or quasi horizontal transport<br />

of fluids depends on fluid density <strong>and</strong> compressibility, impulse generator<br />

frequency, pressure wave amplitude <strong>and</strong> reachable flow rate. <strong>Pressure</strong> wave<br />

attenuation is taken <strong>in</strong>to account by the steady state friction factor, f. Another<br />

way to express this is that estimated efficiency is optimistic s<strong>in</strong>ce steady<br />

friction only does not model accurately pressure wave attenuation (Chapter<br />

6).<br />

155


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

9.3.3 Test Case<br />

<strong>Impulse</strong> pump<strong>in</strong>g of water, volatile <strong>and</strong> black oils at 100 o C, is considered <strong>in</strong><br />

a 1000 m long pipel<strong>in</strong>e, connected to a tank at one end, <strong>and</strong> to an impulse<br />

generator at the other end. In-situ static pressures are set to 100 bar above<br />

bubble po<strong>in</strong>t pressures <strong>and</strong> <strong>in</strong>let non-return valve open<strong>in</strong>g pressures are <strong>in</strong>creased<br />

from 10 bar above bubble po<strong>in</strong>t pressure to 1 bar below <strong>in</strong>-situ static<br />

pressure. <strong>Pressure</strong> waves 200 bar <strong>in</strong> peak-to-peak amplitude are used <strong>in</strong> each<br />

case. <strong>Pressure</strong> conditions for water, volatile oil <strong>and</strong> black oil are listed <strong>in</strong> Table<br />

9.2.<br />

Reachable flow rates aga<strong>in</strong>st pressure gradient between pipel<strong>in</strong>e <strong>and</strong> <strong>in</strong>let nonreturn<br />

valve open<strong>in</strong>g pressure for pressure waves 4 Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency,<br />

<strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation, are shown <strong>in</strong> Figure 9.6 <strong>and</strong> Figure<br />

9.7, respectively. Reachable flow rates us<strong>in</strong>g pressure waves 4 Hz or 5 Hz<br />

<strong>in</strong> frequency are the greatest with volatile oil, <strong>and</strong> the lowest with water.<br />

Reachable flow rates globally decrease with <strong>in</strong>crease <strong>in</strong> pressure difference<br />

between pipel<strong>in</strong>e <strong>in</strong>-situ static pressure <strong>and</strong> <strong>in</strong>let tank.<br />

Reachable flow rates of volatile oil us<strong>in</strong>g pressure waves 4 Hz <strong>in</strong> frequency <strong>and</strong><br />

neglect<strong>in</strong>g pressure wave attenuation are 95 m 3 m<strong>in</strong> −1 , 75 m 3 m<strong>in</strong> −1 <strong>and</strong> 33<br />

m 3 m<strong>in</strong> −1 for pressure differences between pipel<strong>in</strong>e <strong>and</strong> <strong>in</strong>let tank of 1 bara,<br />

40 bara <strong>and</strong> 90 bara, respectively. Reachable flow rates of volatile oil, black oil<br />

<strong>and</strong> water, us<strong>in</strong>g pressure waves 4 Hz <strong>in</strong> frequency, for pipel<strong>in</strong>e <strong>in</strong>-situ static<br />

pressure about 45 bar greater than <strong>in</strong>let tank pressure, are 75 m 3 m<strong>in</strong> −1 , 22.2<br />

m 3 m<strong>in</strong> −1 <strong>and</strong> 10.4 m 3 m<strong>in</strong> −1 , respectively.<br />

Reachable flow rates of volatile oil us<strong>in</strong>g pressures 5 Hz <strong>in</strong> frequency <strong>and</strong><br />

neglect<strong>in</strong>g pressure wave attenuation are 60.7 m 3 m<strong>in</strong> −1 , 45.7 m 3 m<strong>in</strong> −1 <strong>and</strong><br />

36 m 3 m<strong>in</strong> −1 for pressure differences between pipel<strong>in</strong>e <strong>and</strong> <strong>in</strong>let tank of 1 bara,<br />

40 bara <strong>and</strong> 90 bara, respectively. Reachable flow rates of volatile oil, black oil<br />

<strong>and</strong> water, us<strong>in</strong>g pressure waves 4 Hz <strong>in</strong> frequency, for pipel<strong>in</strong>e <strong>in</strong>-situ static<br />

pressure about 45 bar greater than <strong>in</strong>let tank pressure, are 45.7 m 3 m<strong>in</strong> −1 ,<br />

23.4 m 3 m<strong>in</strong> −1 <strong>and</strong> 10.2 m 3 m<strong>in</strong> −1 , respectively.<br />

Differences between reachable flow rates us<strong>in</strong>g pressure waves 4 Hz <strong>and</strong> 5<br />

Hz <strong>in</strong> frequency arise from <strong>in</strong>let non-return valve oscillations. Water hammer<br />

pressure waves <strong>in</strong>duced by <strong>in</strong>let non-return valve open<strong>in</strong>g superimpose on<br />

top of negative pressure pulses that open the <strong>in</strong>let non-return valve. Hence<br />

counteract the open<strong>in</strong>g action <strong>and</strong> are responsible for pressure wave oscillations.<br />

Reachable flow rates are therefore not predictable easily without help<br />

of numerical simulations.<br />

<strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiencies calculated us<strong>in</strong>g Eq. 9.9 aga<strong>in</strong>st pressure<br />

differences between pipel<strong>in</strong>e <strong>and</strong> <strong>in</strong>let tank, assum<strong>in</strong>g a 0.01 friction factor, are<br />

156


9.3. Horizontal Transport of Fluids<br />

Table 9.2: Density, Acoustic Velocity, Bubble Po<strong>in</strong>t <strong>Pressure</strong>, Inlet<br />

Non-Return Valve Open<strong>in</strong>g <strong>Pressure</strong>s <strong>and</strong> Pipel<strong>in</strong>e In-Situ<br />

Static <strong>Pressure</strong> for <strong>Impulse</strong> Pump<strong>in</strong>g of Water, Volatile<br />

Oil <strong>and</strong> Black Oil at 100 o C<br />

Based on Data From Figure 4.8, Figure 4.12 <strong>and</strong> McCa<strong>in</strong> (1990)<br />

Fluid<br />

ρ a pBP pOpen<strong>in</strong>g p0<br />

[kg m −3 ] [ms −1 ] [bara] [bara] [bara]<br />

Water 1000 1400 1 11 101<br />

Volatile Oil 700 900 240 250 340<br />

Black Oil 720 1000 145 155 245<br />

shown for pressure waves 4 Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency <strong>in</strong> Figure 9.8 <strong>and</strong> Figure<br />

9.9, respectively. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiency is greater for low pressure<br />

difference between pipel<strong>in</strong>e <strong>and</strong> <strong>in</strong>let tank. However, maximum reachable impulse<br />

pump<strong>in</strong>g efficiency for the simulated test case is 3.25 % <strong>and</strong> is obta<strong>in</strong>ed<br />

for pump<strong>in</strong>g of volatile oil. <strong>Impulse</strong> pump<strong>in</strong>g efficiencies of water <strong>and</strong> black<br />

oils oscillate between 0.1 <strong>and</strong> 0.5 %.<br />

Horizontal transport of fluids us<strong>in</strong>g impulse pump<strong>in</strong>g can be obta<strong>in</strong>ed with<br />

low <strong>in</strong>flow efficiency. In addition, impulse pump<strong>in</strong>g efficiency must be considered<br />

jo<strong>in</strong>tly with additional equipment efficiencies for <strong>in</strong>creas<strong>in</strong>g <strong>in</strong>let tank<br />

pressure, if necessary. In-situ static pressure <strong>in</strong>side the fluid-filled pipel<strong>in</strong>e can<br />

be regulated by the outlet non-return valve.<br />

<strong>Impulse</strong> pump<strong>in</strong>g can be of <strong>in</strong>terest for horizontal transport of fluids despite<br />

its low efficiency. Of particular <strong>in</strong>terest is the possibility of separat<strong>in</strong>g the<br />

impulse generator mechanical part from the ma<strong>in</strong> flow l<strong>in</strong>e, hence facilitat<strong>in</strong>g<br />

ma<strong>in</strong>tenance. Of <strong>in</strong>terest is also the high pressure fluid exit<strong>in</strong>g the pipel<strong>in</strong>e.<br />

Such high pressure can help further <strong>in</strong> process<strong>in</strong>g of petroleum.<br />

157


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Flow Rate [10 3 l m<strong>in</strong> −1 ]<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Volatile Oil<br />

Black Oil<br />

Water<br />

0<br />

0 20 40 60 80 100<br />

p − p [bara]<br />

0 Open<strong>in</strong>g<br />

Figure 9.6: Reachable Flow Rate for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of<br />

Water, Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak<br />

Amplitude, 4 Hz <strong>in</strong> Frequency, Aga<strong>in</strong>st Difference<br />

Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-Return Valve Open<strong>in</strong>g<br />

<strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

158


Flow Rate [10 3 l m<strong>in</strong> −1 ]<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

9.3. Horizontal Transport of Fluids<br />

Volatile Oil<br />

Black Oil<br />

Water<br />

0<br />

0 20 40 60 80 100<br />

p − p [bara]<br />

0 Open<strong>in</strong>g<br />

Figure 9.7: Reachable Flow Rate for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of<br />

Water, Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak<br />

Amplitude, 5 Hz <strong>in</strong> Frequency, Aga<strong>in</strong>st Difference<br />

Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-Return Valve Open<strong>in</strong>g<br />

<strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

159


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Efficiency [%]<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

Volatile Oil<br />

Black Oil<br />

Water<br />

0.0<br />

0 20 40 60 80 100<br />

p − p [bara]<br />

0 Open<strong>in</strong>g<br />

Figure 9.8: Inflow Efficiency for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of<br />

Water, Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak<br />

Amplitude, 4 Hz <strong>in</strong> Frequency, Aga<strong>in</strong>st Difference<br />

Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-Return Valve Open<strong>in</strong>g<br />

<strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

160


Efficiency [%]<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

9.3. Horizontal Transport of Fluids<br />

Volatile Oil<br />

Black Oil<br />

Water<br />

0.0<br />

0 20 40 60 80 100<br />

p − p [bara]<br />

0 Open<strong>in</strong>g<br />

Figure 9.9: Inflow Efficiency for Horizontal <strong>Impulse</strong> Pump<strong>in</strong>g of<br />

Water, Volatile <strong>and</strong> Black Oils <strong>in</strong> 1000 m Long Pipel<strong>in</strong>e,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 100 bar <strong>in</strong> Peak-To-Peak<br />

Amplitude, 5 Hz <strong>in</strong> Frequency, Aga<strong>in</strong>st Difference<br />

Between Pipel<strong>in</strong>e <strong>and</strong> Inlet Non-Return Valve Open<strong>in</strong>g<br />

<strong>Pressure</strong>s, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

161


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

9.4 Artificial Lift<br />

9.4.1 Wellhead <strong>Pressure</strong> <strong>and</strong> <strong>Pressure</strong> Wave Amplitude<br />

Artificial lift is theoretically possible with an impulse pump located at the wellhead<br />

s<strong>in</strong>ce the vertical lift depends directly on the wellhead pressure. M<strong>in</strong>imum<br />

reachable pressures are aga<strong>in</strong> bubble po<strong>in</strong>t pressures <strong>in</strong> order to avoid cavitation.<br />

In addition, maximum pressure are limited to a maximum allowable<br />

work<strong>in</strong>g pressure, pMAWP, of 6000 psia (414 bara), st<strong>and</strong>ardized by ASME<br />

B31 (2003). Superposition of pressure waves <strong>and</strong> <strong>in</strong>-situ static wellhead pressure<br />

must therefore be ma<strong>in</strong>ta<strong>in</strong>ed between the bubble po<strong>in</strong>t pressure <strong>and</strong> the<br />

maximum allowable work<strong>in</strong>g pressure.<br />

In vertical production tub<strong>in</strong>gs, m<strong>in</strong>imum pressures occur at wellhead dur<strong>in</strong>g<br />

propagation of negative pressure pulses, <strong>and</strong> at bottomhole dur<strong>in</strong>g reflection of<br />

negative pressure pulses at a closed <strong>in</strong>let non-return valve. For pressure waves<br />

2 pWave <strong>in</strong> peak-to-peak amplitude, maximum decrease <strong>in</strong> pressure dur<strong>in</strong>g<br />

superposition of a negative pressure pulse on its own reflection equals 2 pWave;<br />

that is the pressure wave amplitude. The m<strong>in</strong>imum bottomhole local pressure<br />

can then be expressed by:<br />

pm<strong>in</strong>BH = pWH + ρgL − 2pWave<br />

(9.10)<br />

where, pWH, is wellhead pressure, g, gravitational constant, <strong>and</strong>, L, production<br />

tub<strong>in</strong>g depth. The m<strong>in</strong>imum bottomhole pressure at the <strong>in</strong>let non-return valve<br />

ensures no <strong>in</strong>flow from the wellbore. Another way to express this is that the<br />

m<strong>in</strong>imum bottomhole pressure will never be reached if a pump<strong>in</strong>g action is<br />

obta<strong>in</strong>ed.<br />

Wellhead pressures <strong>and</strong> pressure wave amplitude must therefore be selected to<br />

ensure <strong>in</strong>let non-return valve open<strong>in</strong>g <strong>and</strong> avoid cavitation <strong>in</strong>side the pipel<strong>in</strong>e.<br />

The wellbore pressure must also be greater than the bubble po<strong>in</strong>t pressure.<br />

Optimised wellhead pressure <strong>and</strong> pressure wave peak-to-peak amplitude can<br />

then be calculated based on bubble po<strong>in</strong>t pressure <strong>and</strong> maximum allowable<br />

work<strong>in</strong>g pressure. The optimised wellhead pressure can be expressed by:<br />

pWH = pMAWP + pBP<br />

2<br />

(9.11)<br />

The correspond<strong>in</strong>g optimised pressure wave peak-to-peak amplitude is then:<br />

2pWave = pMAWP − pBP<br />

162<br />

(9.12)


9.4. Artificial Lift<br />

Pump<strong>in</strong>g of fluids from wellbore to wellhead <strong>in</strong> petroleum production is required<br />

if the production tub<strong>in</strong>g bottomhole pressure is lower than the gravitational<br />

pressure <strong>in</strong>duced by the equivalent ocean water column. In addition,<br />

m<strong>in</strong>imum bottomhole pressure must be greater than the fluid bubble po<strong>in</strong>t<br />

pressure at considered temperature conditions. M<strong>in</strong>imum bottomhole pressure<br />

lower <strong>and</strong> upper limits can then be expressed by:<br />

pBP < pm<strong>in</strong>BH < ρWgL (9.13)<br />

where, ρW, is water density. Both <strong>in</strong>equalities <strong>in</strong> Eq. 9.13 express m<strong>in</strong>imum<br />

production tub<strong>in</strong>g depth conditions. After simplification, the m<strong>in</strong>imum production<br />

tub<strong>in</strong>g depth expressed by the first <strong>in</strong>equality <strong>in</strong> Eq. 9.13 becomes:<br />

L1 > pMAWP − pBP<br />

2<br />

1<br />

ρg<br />

(9.14)<br />

where, ρ, represents fluid density. After simplification, the m<strong>in</strong>imum production<br />

tub<strong>in</strong>g depth expressed by the second <strong>in</strong>equality <strong>in</strong> Eq. 9.13 becomes:<br />

L2 > −pMAWP + 3pBP<br />

2<br />

1<br />

(ρW − ρ)g<br />

(9.15)<br />

Production tub<strong>in</strong>g depths are then limited by the greater value between L1<br />

<strong>and</strong> L2. M<strong>in</strong>imum production tub<strong>in</strong>g depths for impulse pump<strong>in</strong>g of water,<br />

volatile oil <strong>and</strong> black oil at 100 o C are listed <strong>in</strong> Table 9.3. Only L1 can be<br />

calculated for water. Production tub<strong>in</strong>g depths are limited by L2 value of<br />

5181 m <strong>and</strong> L1 value of 1911 m for impulse pump<strong>in</strong>g of volatile <strong>and</strong> black oil,<br />

respectively. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong> production tub<strong>in</strong>gs shorter than m<strong>in</strong>imum<br />

calculated values can not be obta<strong>in</strong>ed without <strong>in</strong>duc<strong>in</strong>g cavitation that can<br />

cause damages to the equipment.<br />

M<strong>in</strong>imum production tub<strong>in</strong>g depths are evaluated based on optimised wellhead<br />

pressure <strong>and</strong> pressure wave peak-to-peak amplitude. However, wellhead<br />

pressure <strong>and</strong> pressure wave peak-to-peak amplitude can take other values. Of<br />

particular <strong>in</strong>terest is impulse pump<strong>in</strong>g of water with low bubble po<strong>in</strong>t pressure<br />

where wellhead pressure can be decreased dramatically <strong>and</strong> allow impulse<br />

pump<strong>in</strong>g from short depths. For <strong>in</strong>stance, a 10 bara wellhead pressure <strong>and</strong><br />

a 20 bar pressure wave peak-to-peak amplitude can be generated for impulse<br />

pump<strong>in</strong>g of water from a 100 m deep water-filled wellbore.<br />

163


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Table 9.3: M<strong>in</strong>imum Production Tub<strong>in</strong>g Lengths L1 <strong>in</strong> Eq. 9.14 <strong>and</strong><br />

L2 <strong>in</strong> Eq. 9.15 for <strong>Impulse</strong> Pump<strong>in</strong>g of Water, Volatile Oil<br />

<strong>and</strong> Black Oil at 100 o C<br />

9.4.2 Efficiency<br />

Fluid<br />

pBP<br />

[bara]<br />

ρ<br />

[kg m<br />

L1 L2<br />

−3 ] [m] [m]<br />

Water 1 1000 2110 ×<br />

Volatile Oil 240 700 1274 5181<br />

Black Oil 145 720 1911 364<br />

<strong>Impulse</strong> pump<strong>in</strong>g efficiency <strong>in</strong> vertical production tub<strong>in</strong>gs can be calculated<br />

directly from impulse generator characteristics <strong>and</strong> reachable flow rates. The<br />

impulse generator is the only part of an impulse pump<strong>in</strong>g apparatus that<br />

requires external energy supply. Hence, impulse generator characteristics <strong>and</strong><br />

generated pressure wave amplitude def<strong>in</strong>e the required power. Modell<strong>in</strong>g an<br />

impulse generator by a back <strong>and</strong> forth oscillat<strong>in</strong>g piston, the required power<br />

can be calculated by:<br />

PR = 4πKL˙γd 2 p 2 Wave<br />

(9.16)<br />

Power generated by impulse pump<strong>in</strong>g applied to artificial lift can be def<strong>in</strong>ed<br />

by flow rate <strong>and</strong> reached vertical lift. Generated power at the <strong>in</strong>let non-return<br />

valve can then be expressed by:<br />

PG<strong>in</strong> = qpLift<br />

(9.17)<br />

where the vertical lift pressure, pLift, can directly be calculated from the<br />

difference <strong>in</strong> pressure between the wellbore <strong>and</strong> the production tub<strong>in</strong>g depth<br />

equivalent water column. The vertical lift pressure can then be expressed by:<br />

pLift = ρgL − pWB<br />

(9.18)<br />

After simplifications, <strong>in</strong>flow efficiency of impulse pump<strong>in</strong>g applied to vertical<br />

lift can be calculated by:<br />

η = PG<strong>in</strong><br />

PR<br />

= q (ρgL − pWB)<br />

4πKL˙γd 2 p 2 Wave<br />

164<br />

(9.19)


9.4. Artificial Lift<br />

Inflow efficiency of impulse pump<strong>in</strong>g applied to vertical lift depends on wellbore<br />

pressure, fluid density <strong>and</strong> compressibility, pressure wave peak-to-peak<br />

amplitude, production tub<strong>in</strong>g depth <strong>and</strong> reachable flow rate. Of particular<br />

<strong>in</strong>terest is the lack of parameter represent<strong>in</strong>g directly pressure wave attenuation<br />

<strong>in</strong> the previous equation Eq. 9.19 compared to impulse pump<strong>in</strong>g <strong>in</strong>flow<br />

efficiency for horizontal transport of fluids Eq. 9.9. <strong>Pressure</strong> wave attenuation<br />

is only represented by the reachable flow rate.<br />

9.4.3 Test Case<br />

Reachable flow rates <strong>and</strong> impulse pump<strong>in</strong>g <strong>in</strong>flow efficiencies are calculated for<br />

artificial lift of hydrocarbons from a 10 km deep wellbore through a production<br />

tub<strong>in</strong>g 10 <strong>in</strong>ches <strong>in</strong> diameter. <strong>Impulse</strong> pump<strong>in</strong>g parameters for artificial lift<br />

of water, volatile oil <strong>and</strong> black oil are listed <strong>in</strong> Table 9.4. Reachable flow rates<br />

are simulated us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves 4 Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency,<br />

<strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation.<br />

Evolution of reachable flow rates aga<strong>in</strong>st wellbore pressure for s<strong>in</strong>usoidal pressure<br />

waves 4 Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency, neglect<strong>in</strong>g pressure wave attenuation,<br />

are shown <strong>in</strong> Figure 9.10 <strong>and</strong> Figure 9.11, respectively. Reachable flow rates<br />

us<strong>in</strong>g pressure waves 4 Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency are the greatest for black oil.<br />

Reachable flow rates are greater for water than for volatile oil until approximately<br />

930 bara wellbore pressure, then reachable flow rates are the lowest<br />

with water. Reachable flow rates <strong>in</strong>crease with wellbore pressure.<br />

Reachable flow rates of black oil us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 4 Hz <strong>in</strong><br />

frequency <strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation, are 36 m 3 m<strong>in</strong> −1 , 92.7<br />

m 3 m<strong>in</strong> −1 <strong>and</strong> 155 m 3 m<strong>in</strong> −1 for 875 bara, 925 bara <strong>and</strong> 975 bara wellbore<br />

pressures, respectively. Reachable flow rates of black oil, volatile oil <strong>and</strong> water,<br />

us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 4 Hz <strong>in</strong> frequency, for a 950 bara wellbore<br />

pressure are 119 m 3 m<strong>in</strong> −1 , 41.7 m 3 m<strong>in</strong> −1 <strong>and</strong> 20.2 m 3 m<strong>in</strong> −1 , respectively.<br />

Reachable flow rates of black oil us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 5 Hz <strong>in</strong><br />

frequency <strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation, are 27.1 m 3 m<strong>in</strong> −1 , 55.2<br />

m 3 m<strong>in</strong> −1 <strong>and</strong> 156 m 3 m<strong>in</strong> −1 for 875 bara, 925 bara <strong>and</strong> 975 bara wellbore<br />

pressures, respectively. Reachable flow rates of black oil, volatile oil <strong>and</strong> water<br />

us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 5 Hz <strong>in</strong> frequency, for a 950 bara wellbore<br />

pressure are 112 m 3 m<strong>in</strong> −1 , 34.1 m 3 m<strong>in</strong> −1 <strong>and</strong> 56 m 3 m<strong>in</strong> −1 , respectively.<br />

Differences between reachable flow rates us<strong>in</strong>g pressure waves 4 Hz <strong>and</strong> 5 Hz<br />

<strong>in</strong> frequency arise from <strong>in</strong>let non-return valve oscillations.<br />

Evolution of impulse pump<strong>in</strong>g <strong>in</strong>flow efficiency for s<strong>in</strong>usoidal pressure waves 4<br />

Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency, neglect<strong>in</strong>g pressure wave attenuation, are shown <strong>in</strong><br />

Figure 9.12 <strong>and</strong> Figure 9.13, respectively. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiency<br />

us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves 4 Hz <strong>and</strong> 5 Hz <strong>in</strong> frequency is the greatest<br />

165


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

for artificial lift of black oil. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiency is greater for<br />

water than for volatile oil until approximately 930 bara wellbore pressure, then<br />

impulse pump<strong>in</strong>g <strong>in</strong>flow efficiency is the lowest for water.<br />

<strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiencies for artificial lift of black oil, us<strong>in</strong>g s<strong>in</strong>usoidal<br />

pressure waves 4 Hz <strong>in</strong> frequency <strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation<br />

are 0.74 %, 1.0 % <strong>and</strong> 0.18 % for 875 bara, 925 bara <strong>and</strong> 975 bara<br />

wellbore pressures, respectively. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiencies for artificial<br />

lift of black oil, volatile oil <strong>and</strong> water, us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves 4<br />

Hz <strong>in</strong> frequency, for a 950 bara wellbore pressure are 0.72 %, 0.34 % <strong>and</strong> 0.17<br />

%, respectively.<br />

<strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiencies for artificial lift of black oil, us<strong>in</strong>g s<strong>in</strong>usoidal<br />

pressure waves 5 Hz <strong>in</strong> frequency <strong>and</strong> neglect<strong>in</strong>g pressure wave attenuation<br />

are 0.45 %, 0.48 % <strong>and</strong> 0.15 % for 875 bara, 925 bara <strong>and</strong> 975 bara<br />

wellbore pressures, respectively. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow efficiencies for artificial<br />

lift of black oil, volatile oil <strong>and</strong> water, us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves 4<br />

Hz <strong>in</strong> frequency, for a 95 bara wellbore pressure are 0.54 %, 0.27 % <strong>and</strong> 0.08<br />

%, respectively.<br />

Artificial lift of hydrocarbons us<strong>in</strong>g impulse pump<strong>in</strong>g is obta<strong>in</strong>ed with a very<br />

low <strong>in</strong>flow efficiency due to the required power to generate pressure waves us<strong>in</strong>g<br />

a back <strong>and</strong> forth mov<strong>in</strong>g piston. In addition, impulse pump<strong>in</strong>g of hydrocarbon<br />

requires s<strong>in</strong>gle phase fluids <strong>in</strong> order to avoid cavitation that could result <strong>in</strong><br />

irreversible damages to the apparatus. Wellbore pressures required to ma<strong>in</strong>ta<strong>in</strong><br />

fluids <strong>in</strong> s<strong>in</strong>gle phase state are very seldom <strong>in</strong> practice. Artificial <strong>in</strong>crease <strong>in</strong><br />

wellbore pressure is possible, yet goes aga<strong>in</strong>st the advantage of hav<strong>in</strong>g energy<br />

supply requirement close to wellhead only with impulse pump<strong>in</strong>g. <strong>Impulse</strong><br />

pump<strong>in</strong>g applied to artificial lift is consequently possible, but not practical.<br />

Table 9.4: <strong>Impulse</strong> Pump<strong>in</strong>g Parameters for Artificial Lift of Water,<br />

Volatile Oil <strong>and</strong> Black Oil at 100 o C, from 10000 m Deep<br />

Wellbore, 10 <strong>in</strong>ches <strong>in</strong> Diameter, Neglect<strong>in</strong>g <strong>Pressure</strong><br />

Wave Attenuation<br />

ρ pWH 2 pWave p m<strong>in</strong>(BH)<br />

[kg m −3 ] [bara] [bar] [bara]<br />

Water<br />

1000 207 412 775<br />

Volatile Oil<br />

700 327.5 175 839<br />

Black Oil<br />

720 280 270 716<br />

166


Flow Rate [10 3 l m<strong>in</strong> −1 ]<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Black Oil<br />

Water<br />

Volatile Oil<br />

9.4. Artificial Lift<br />

0<br />

850 875 900 925 950 975 1000<br />

Wellbore <strong>Pressure</strong> [bara]<br />

Figure 9.10: Reachable Flow Rates Aga<strong>in</strong>st Wellbore <strong>Pressure</strong>, for<br />

Vertical Transport of Water, S<strong>in</strong>gle Phase Volatile <strong>and</strong><br />

Black Oils, at 100 o C, from a 10 km Deep Wellbore,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong> Frequency, Neglect<strong>in</strong>g<br />

<strong>Pressure</strong> Wave Attenuation<br />

167


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Flow Rate [10 3 l m<strong>in</strong> −1 ]<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Black Oil<br />

Water<br />

Volatile Oil<br />

0<br />

850 875 900 925 950 975 1000<br />

Wellbore <strong>Pressure</strong> [bara]<br />

Figure 9.11: Reachable Flow Rates Aga<strong>in</strong>st Wellbore <strong>Pressure</strong>, for<br />

Vertical Transport of Water, S<strong>in</strong>gle Phase Volatile <strong>and</strong><br />

Black Oils, at 100 o C, from a 10 km Deep Wellbore,<br />

Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 5 Hz <strong>in</strong> Frequency, Neglect<strong>in</strong>g<br />

<strong>Pressure</strong> Wave Attenuation<br />

168


Efficiency [%]<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Black Oil<br />

Water<br />

Volatile Oil<br />

9.4. Artificial Lift<br />

0.0<br />

850 875 900 925 950 975 1000<br />

Wellbore <strong>Pressure</strong> [bara]<br />

Figure 9.12: Theoretical <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Efficiency Aga<strong>in</strong>st<br />

Wellbore <strong>Pressure</strong>, for Vertical Transport of Water,<br />

S<strong>in</strong>gle Phase Volatile <strong>and</strong> Black Oils at 100 o C, from a<br />

10 km Deep Wellbore, Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong><br />

Frequency, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

169


Chapter 9. Design, Applications <strong>and</strong> Performance of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Efficiency [%]<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Black Oil<br />

Water<br />

Volatile Oil<br />

0.0<br />

850 875 900 925 950 975 1000<br />

Wellbore <strong>Pressure</strong> [bara]<br />

Figure 9.13: Theoretical <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Efficiency Aga<strong>in</strong>st<br />

Wellbore <strong>Pressure</strong>, for Vertical Transport of Water,<br />

S<strong>in</strong>gle Phase Volatile <strong>and</strong> Black Oils at 100 o C, from a<br />

10 km Deep Wellbore, Us<strong>in</strong>g <strong>Pressure</strong> <strong>Waves</strong> 5 Hz <strong>in</strong><br />

Frequency, Neglect<strong>in</strong>g <strong>Pressure</strong> Wave Attenuation<br />

170


Chapter 10<br />

Characteristic Curves of<br />

<strong>Impulse</strong> Pump<strong>in</strong>g<br />

10.1 Characteristic Curves<br />

Pump characteristic curves illustrate performances of an equipment <strong>in</strong> terms of<br />

reachable vertical lift <strong>and</strong> efficiency aga<strong>in</strong>st flow rate. Characteristic curves of<br />

impulse pump<strong>in</strong>g applied to artificial lift depend on production tub<strong>in</strong>g depth,<br />

wellhead pressure <strong>and</strong> pressure wave peak-to-peak amplitude. Characteristic<br />

curves are here<strong>in</strong> presented for particular test cases that illustrate impulse<br />

pump<strong>in</strong>g of aquifer from 10 m deep to 5000 m deep.<br />

<strong>Impulse</strong> pump<strong>in</strong>g of aquifer, 1000 kg m −3 <strong>in</strong> density, us<strong>in</strong>g a production tub<strong>in</strong>g<br />

0.124 m <strong>in</strong> diameter is here<strong>in</strong> <strong>in</strong>vestigated through two test cases. The first<br />

test case is characterized by a 21 bara wellhead pressure, a 40 bar pressure<br />

wave peak-to-peak amplitude <strong>and</strong> a constant 70 m lift<strong>in</strong>g height. The second<br />

test case is characterized by a 207 bara wellhead pressure, a 412 bar pressure<br />

wave peak-to-peak amplitude <strong>and</strong> a constant 700 m lift<strong>in</strong>g height. S<strong>in</strong>usoidal<br />

pressure waves, 1 Hz <strong>in</strong> frequency are used <strong>in</strong> both cases.<br />

Characteristic curves for heights between 10 m to 1000 m <strong>and</strong> between 1000 m<br />

<strong>and</strong> 5000 m are obta<strong>in</strong>ed us<strong>in</strong>g the first <strong>and</strong> the second test case, respectively.<br />

Reachable flow rates are simulated us<strong>in</strong>g the impulse pump<strong>in</strong>g simulation tool<br />

described <strong>in</strong> Chapter 8. Required power <strong>and</strong> <strong>in</strong>let generated power are calculated<br />

us<strong>in</strong>g Eq. 9.16 <strong>and</strong> Eq. 9.17, respectively. Power generated at the outlet<br />

non-return valve due to outflow of high pressure liquid is added to the <strong>in</strong>let<br />

generated power to calculate the total efficiency. The generated power at the<br />

outlet non-return valve can be expressed by:<br />

PGout = q(pWH − pOutlet) (10.1)<br />

171


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

where, pWH, def<strong>in</strong>es wellhead pressure, q, reachable flow rate, <strong>and</strong>, pOutlet,<br />

outlet tank pressure. The total efficiency can therefore be expressed by:<br />

η = PG<strong>in</strong> + PGout<br />

PR<br />

(10.2)<br />

Evolutions of reachable flow rate aga<strong>in</strong>st production tub<strong>in</strong>g depth for both test<br />

cases are shown <strong>in</strong> Figure 10.1. Reachable flow rates <strong>in</strong>crease with production<br />

tub<strong>in</strong>g depth for both test cases. Reachable flow rate <strong>in</strong>creases from 151<br />

l m<strong>in</strong> −1 at 100 m depth, to 508 l m<strong>in</strong> −1 at 600 m depth, <strong>and</strong> to 598 l m<strong>in</strong> −1<br />

at 800 m depth for the first test case. Reachable flow rate <strong>in</strong>creases from 956<br />

l m<strong>in</strong> −1 at 100 m depth, to 4010 l m<strong>in</strong> −1 at 1750 m depth, <strong>and</strong> to 5870 l m<strong>in</strong> −1<br />

at 3750 m depth for the second test case. Reachable flow rate fluctuations are<br />

due to <strong>in</strong>let non-return valve oscillations.<br />

Evolutions of impulse pump<strong>in</strong>g total efficiency aga<strong>in</strong>st production tub<strong>in</strong>g depth<br />

for both test cases are shown <strong>in</strong> Figure 10.2 us<strong>in</strong>g a 1 bara outlet pressure.<br />

<strong>Impulse</strong> pump<strong>in</strong>g total efficiency decreases when production tub<strong>in</strong>g depth <strong>in</strong>creases.<br />

<strong>Impulse</strong> pump<strong>in</strong>g total efficiency decreases from 20.1 % at 100 m<br />

depth, to 11.3 % at 600 m depth, <strong>and</strong> to 10 % at 800 m depth for the first<br />

test case. <strong>Impulse</strong> pump<strong>in</strong>g total efficiency decreases when production tub<strong>in</strong>g<br />

depth <strong>in</strong>creases. <strong>Impulse</strong> pump<strong>in</strong>g total efficiency oscillates from 1.2 % at 1000<br />

m depth, to 2.9 % at 1750 m depth, <strong>and</strong> to 1.8 % at 3750 m depth for the<br />

second test case.<br />

<strong>Impulse</strong> pump<strong>in</strong>g is therefore most efficient for artificial lift of shallow water.<br />

Performance of impulse pump<strong>in</strong>g for artificial lift of shallow water can be<br />

illustrated by its characteristic curve. The characteristic curve is obta<strong>in</strong>ed<br />

us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 1 Hz <strong>in</strong> frequency, 200 bar <strong>in</strong> peak-to-peak<br />

amplitude, propagat<strong>in</strong>g <strong>in</strong> a production tub<strong>in</strong>g, 0.124 m <strong>in</strong> diameter, with<br />

101 bara wellhead pressure, <strong>and</strong> 1 bara wellbore pressure. The correspond<strong>in</strong>g<br />

characteristic curve is shown <strong>in</strong> Figure 10.3.<br />

Increase <strong>in</strong> lift<strong>in</strong>g height <strong>in</strong>duces decrease <strong>in</strong> reachable flow rate <strong>and</strong> total<br />

impulse impulse pump<strong>in</strong>g efficiency. Reachable flow rates decreases from 1170<br />

l m<strong>in</strong> −1 at 90 m, to 63.1 l m<strong>in</strong> −1 at 990 m. <strong>Impulse</strong> pump<strong>in</strong>g total efficiency<br />

decreases from 25.2 % at 90 m, to 0.25 % at 990 m. Oscillations <strong>in</strong> reachable<br />

flow rate, hence <strong>in</strong> impulse pump<strong>in</strong>g total efficiency, are due to fluctuations<br />

of pressure <strong>in</strong> the vic<strong>in</strong>ity of the <strong>in</strong>let non-return valve. <strong>Impulse</strong> pump<strong>in</strong>g<br />

greatest efficiency is then obta<strong>in</strong>ed for artificial lift of shallow water.<br />

172


Flow Rate [l m<strong>in</strong> −1 ]<br />

6000<br />

5000<br />

4000<br />

3000<br />

2000<br />

1000<br />

Case 1<br />

Case 2<br />

10.1. Characteristic Curves<br />

0<br />

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000<br />

Depth [m]<br />

Figure 10.1: Evolution of <strong>Impulse</strong> Pump<strong>in</strong>g Reachable Flow Rate<br />

Aga<strong>in</strong>st Production Tub<strong>in</strong>g Depth<br />

Efficiency [%]<br />

25.0<br />

22.5<br />

20.0<br />

17.5<br />

15.0<br />

12.5<br />

10.0<br />

7.5<br />

5.0<br />

2.5<br />

Case 1<br />

Case 2<br />

0.0<br />

0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000<br />

Depth [m]<br />

Figure 10.2: Evolution of <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency Aga<strong>in</strong>st<br />

Production Tub<strong>in</strong>g Depth<br />

173


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Height [m]<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

Head<br />

Efficiency<br />

0<br />

0 200 400 600 800 1000 1200<br />

Flow Rate [l m<strong>in</strong> −1 0 200 400 600 800 1000 1200<br />

]<br />

0<br />

0 200 400 600 800 1000 1200<br />

0<br />

Figure 10.3: Characteristic Curve of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g<br />

S<strong>in</strong>usoidal <strong>Pressure</strong> Wave, 1 Hz <strong>in</strong> Frequency, 200 bar<br />

<strong>in</strong> Peak-To-Peak Amplitude <strong>and</strong> pHH = 101bara<br />

10.2 Comparison of Artificial Lift Techniques<br />

Jet pumps, electrical submersible pumps (ESP), <strong>and</strong> gas lift, are three st<strong>and</strong>ard<br />

techniques used for artificial lift of water <strong>and</strong> hydrocarbons <strong>in</strong> production<br />

tub<strong>in</strong>g. <strong>Impulse</strong> pump<strong>in</strong>g provides a new artificial lift technique that can be<br />

compared to st<strong>and</strong>ard artificial lift by consider<strong>in</strong>g a real test case application.<br />

The considered test case is the depressurization of the Statfjord oil field (Boge<br />

et al., 2005).<br />

10.2.1 Test Case Description<br />

The Statfjord field, located on the Norwegian cont<strong>in</strong>ental shelf <strong>in</strong> the Tampen<br />

area, is the largest produc<strong>in</strong>g oil field <strong>in</strong> Europe. Oil production from Statfjord<br />

has been optimised over the years by pressure ma<strong>in</strong>tenance, drill<strong>in</strong>g of deviated<br />

<strong>and</strong> horizontal wells, cont<strong>in</strong>ued <strong>in</strong>fill drill<strong>in</strong>g, water alternat<strong>in</strong>g gas <strong>in</strong>jection.<br />

The Statfjord field oil production is to be shifted to ma<strong>in</strong>ly gas production by<br />

depressuriz<strong>in</strong>g the field. Depressurization is accelerated by pump<strong>in</strong>g aquifer<br />

<strong>and</strong> <strong>in</strong>jected water from the reservoir. Gas will then be liberated from the<br />

rema<strong>in</strong><strong>in</strong>g oil <strong>and</strong> migrate towards the top of the reservoir.<br />

174<br />

30<br />

24<br />

18<br />

12<br />

6<br />

Efficiency [%]


10.2. Comparison of Artificial Lift Techniques<br />

Table 10.1: Desired Flow Rates, Wellbore <strong>Pressure</strong>s <strong>and</strong><br />

Correspond<strong>in</strong>g Lift<strong>in</strong>g Heights for the Three<br />

Depressurization Phases of Statfjord Field<br />

Case Stage pWB q [Sm 3 s −1 ] q [Sm 3 m<strong>in</strong> −1 ] H [m]<br />

A Early 200 bara 0.0579 3.747 709.3<br />

B Middle 100 bara 0.0434 2.604 1728.6<br />

C Late 70 bara 0.0289 1.734 2034.4<br />

The depressurization cycle can be divided <strong>in</strong> three cases that correspond to<br />

different phase of the process. The three stages are early, middle <strong>and</strong> late<br />

stage of depressurization, <strong>and</strong> are characterised by three different reservoir<br />

pressures. Wellbore pressures are assumed equal to reservoir pressure for simplifications.<br />

Water is then pumped from wellbore to wellhead us<strong>in</strong>g a 2748<br />

m production tub<strong>in</strong>g, 0.124 m <strong>in</strong> diameter (Pedersen, June 2007). Wellbore<br />

pressures correspond<strong>in</strong>g to the three depressurization stages <strong>in</strong>vestigated, <strong>and</strong><br />

required flow rates are listed <strong>in</strong> Table 10.1.<br />

10.2.2 Determ<strong>in</strong>ation of <strong>Impulse</strong> Pump<strong>in</strong>g Parameters<br />

<strong>Pressure</strong> wave peak-to-peak amplitude can be calculated for all three depressurization<br />

phased function of impulse pump<strong>in</strong>g efficiency us<strong>in</strong>g Eq. 9.19 <strong>and</strong><br />

test case parameters. <strong>Pressure</strong> wave peak-to-peak amplitudes are calculated<br />

us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 4 Hz <strong>in</strong> frequency for depressurization phase<br />

A, correspond<strong>in</strong>g to a wellbore pressure of 200 bara. Evolutions of pressure<br />

wave peak-to-peak amplitude, wellhead pressure <strong>and</strong> m<strong>in</strong>imum reachable bottomhole<br />

pressure aga<strong>in</strong>st impulse pump<strong>in</strong>g efficiency for various flow rate conditions<br />

are shown <strong>in</strong> Figure 10.4, Figure 10.5 <strong>and</strong> Figure 10.6, respectively.<br />

<strong>Pressure</strong> wave peak-to-peak amplitude decreases with <strong>in</strong>crease <strong>in</strong> impulse<br />

pump<strong>in</strong>g efficiency. For the required flow rate (Table 10.1), pressure wave<br />

peak-to-peak amplitude decreases from 228.7 bar, to 72.3 bar, <strong>and</strong> to 22.9 bar<br />

for 1 %, 10 % <strong>and</strong> 100 % impulse pump<strong>in</strong>g efficiency, respectively. <strong>Pressure</strong><br />

wave peak-to-peak amplitude <strong>in</strong>creases with flow rate. At 50 % impulse pump<strong>in</strong>g<br />

efficiency, pressure wave peak-to-peak amplitude <strong>in</strong>creases from 32.4 bar,<br />

to 45.8 bar, to 64.7 bar <strong>and</strong> to 91.5 bar for qRequired, 2 qRequired, 4 qRequired<br />

<strong>and</strong> 8 qRequired flow rates, respectively.<br />

Wellhead pressure decreases with <strong>in</strong>crease <strong>in</strong> impulse pump<strong>in</strong>g efficiency. For<br />

the required flow rate, wellhead pressure decreases from 114.4 bara, to 36.2<br />

bara, <strong>and</strong> to 11.4 bara for 1 %, 10 % <strong>and</strong> 100 % impulse pump<strong>in</strong>g efficiency,<br />

respectively. Wellhead pressure <strong>in</strong>creases with flow rate. At 50 % impulse<br />

pump<strong>in</strong>g efficiency, wellhead pressure <strong>in</strong>creases from 16.2 bara, to 22.9 bara,<br />

175


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

to 32.4 bara <strong>and</strong> to 45.8 bara for qRequired, 2 qRequired, 4 qRequired <strong>and</strong> 8 qRequired<br />

flow rates, respectively.<br />

M<strong>in</strong>imum reachable bottomhole pressure <strong>in</strong>creases with impulse pump<strong>in</strong>g efficiency.<br />

For the required flow rate (Table 10.1), m<strong>in</strong>imum reachable bottomhole<br />

pressure <strong>in</strong>creases from 155.2 bara, to 233.4 bara, <strong>and</strong> to 258.1 bara for 1 %,<br />

10 % <strong>and</strong> 100 % impulse pump<strong>in</strong>g efficiency, respectively. M<strong>in</strong>imum reachable<br />

bottomhole pressure decreases with flow rate. At 50 % impulse pump<strong>in</strong>g efficiency,<br />

m<strong>in</strong>imum reachable bottomhole pressure decreases from 253.4 bara, to<br />

246.7 bara, to 237.2 bara <strong>and</strong> to 223.8 bara for qRequired, 2 qRequired, 4 qRequired<br />

<strong>and</strong> 8 qRequired flow rates, respectively.<br />

<strong>Impulse</strong> pump<strong>in</strong>g is applicable to Statfjord depressurization phase A either<br />

with a very low efficiency or with a flow rate several times greater than the<br />

required flow rate. A 20 % impulse pump<strong>in</strong>g efficiency can theoretically be<br />

reached by pump<strong>in</strong>g at 8 times the required flow rates. The required flow<br />

rate can theoretically be <strong>in</strong>creased by decreas<strong>in</strong>g the number of wells used for<br />

artificial lift while keep<strong>in</strong>g the field total flow rate the same. Required flow rate<br />

must be compared with reachable flow rate <strong>in</strong> order to assess the applicability<br />

of impulse pump<strong>in</strong>g.<br />

<strong>Impulse</strong> pump<strong>in</strong>g reachable heights with 207 bara wellhead pressure, us<strong>in</strong>g<br />

s<strong>in</strong>usoidal pressure waves 412 bar <strong>in</strong> peak-to-peak amplitude, are shown for<br />

different frequencies <strong>in</strong> Figure 10.7. Reachable heights decrease with <strong>in</strong>crease<br />

<strong>in</strong> flow rate. For pressure waves 1 Hz <strong>in</strong> frequency, height decrease follow<strong>in</strong>g<br />

a straight l<strong>in</strong>e from 2238 m to 0 m for 0 m 3 m<strong>in</strong> −1 <strong>and</strong> 5.37 m 3 m<strong>in</strong> −1 flow<br />

rates, respectively. Height is almost <strong>in</strong>dependent of pressure wave frequency,<br />

except at high flow regime where <strong>in</strong>let non-return valve oscillations <strong>in</strong>duce<br />

irregular flow rates.<br />

10.2.3 Efficiency of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Inflow efficiency of impulse pump<strong>in</strong>g us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 1 Hz <strong>in</strong><br />

frequency, 412 bar <strong>in</strong> peak-to-peak amplitude, with 207 bara wellhead pressure,<br />

are shown for different frequencies <strong>in</strong> Figure 10.8. <strong>Impulse</strong> pump<strong>in</strong>g efficiency<br />

first <strong>in</strong>crease with flow rates <strong>and</strong> then decrease. For pressure waves 1 Hz<br />

<strong>in</strong> frequency, impulse pump<strong>in</strong>g efficiency <strong>in</strong>creases from 0.25 % to 0.53 %,<br />

<strong>and</strong> then decreases to 0.16 % for 0.25 m 3 m<strong>in</strong> −1 , 2.63 m 3 m<strong>in</strong> −1 <strong>and</strong> 4.91<br />

m 3 m<strong>in</strong> −1 flow rates, respectively. <strong>Impulse</strong> pump<strong>in</strong>g efficiency decreases with<br />

<strong>in</strong>crease <strong>in</strong> pressure wave frequency. For about 2.63 m 3 m<strong>in</strong> −1 reachable flow<br />

rate, impulse pump<strong>in</strong>g decreases from 0.53 % to 0.27 %, <strong>and</strong> to 0.13 % for<br />

pressure waves 1 Hz, 2 Hz <strong>and</strong> 4 Hz <strong>in</strong> frequency, respectively.<br />

176


2 p Wave [bar]<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

10.2. Comparison of Artificial Lift Techniques<br />

q = 8 * q Requested<br />

q = 4 * q Requested<br />

q = 2 * q Requested<br />

q = 1 * q Requested<br />

0<br />

0 20 40 60 80 100<br />

Efficiency [%]<br />

Figure 10.4: <strong>Pressure</strong> Wave Peak-To-Peak Amplitude Aga<strong>in</strong>st<br />

<strong>Impulse</strong> Pump<strong>in</strong>g Efficiency for Artificial Lift Dur<strong>in</strong>g<br />

Statfjord Depressurization Phase A, Us<strong>in</strong>g S<strong>in</strong>usoidal<br />

<strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong> Frequency (Table 10.1)<br />

Wellhead <strong>Pressure</strong> [bara]<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

q = 8 * q Requested<br />

q = 4 * q Requested<br />

q = 2 * q Requested<br />

q = 1 * q Requested<br />

0<br />

0 20 40 60 80 100<br />

Efficiency [%]<br />

Figure 10.5: Wellhead <strong>Pressure</strong> Aga<strong>in</strong>st <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency<br />

for Artificial Lift Dur<strong>in</strong>g Statfjord Depressurization<br />

Phase A, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong><br />

Frequency (Table 10.1)<br />

177


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

M<strong>in</strong>imum Bottomhole <strong>Pressure</strong> [bara]<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

q = 1 * q<br />

Requested<br />

q = 2 * q<br />

Requested<br />

−50<br />

q = 4 * q<br />

Requested<br />

q = 8 * q<br />

Requested<br />

−100<br />

0 20 40 60 80 100<br />

Efficiency [%]<br />

Figure 10.6: M<strong>in</strong>imum Reachable Bottom Hole <strong>Pressure</strong> Aga<strong>in</strong>st<br />

<strong>Impulse</strong> Pump<strong>in</strong>g Efficiency for Artificial Lift Dur<strong>in</strong>g<br />

Statfjord Depressurization Phase A, Us<strong>in</strong>g S<strong>in</strong>usoidal<br />

<strong>Pressure</strong> <strong>Waves</strong> 4 Hz <strong>in</strong> Frequency (Table 10.1)<br />

Head [m]<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

1 Hz<br />

2 Hz<br />

4 Hz<br />

8 Hz<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

Flow Rate [l m<strong>in</strong> −1 ]<br />

Figure 10.7: Reachable Heights <strong>and</strong> Flow Rates for Depressurization<br />

of Statfjord Field, 2748 m Deep, Us<strong>in</strong>g S<strong>in</strong>usoidal<br />

<strong>Pressure</strong> <strong>Waves</strong> 412 bar <strong>in</strong> Peak-To-Peak Amplitude<br />

Aga<strong>in</strong>st <strong>Pressure</strong> Wave Frequency<br />

178


Efficiency [%]<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

10.2. Comparison of Artificial Lift Techniques<br />

1 Hz<br />

2 Hz<br />

4 Hz<br />

8 Hz<br />

0.0<br />

0 1000 2000 3000 4000 5000 6000<br />

Flow Rate [l m<strong>in</strong> −1 ]<br />

Figure 10.8: Reachable Inflow Efficiencies <strong>and</strong> Flow Rates for<br />

Depressurization of Statfjord Field, 2748 m Deep, Us<strong>in</strong>g<br />

S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong> 412 bar <strong>in</strong> Peak-To-Peak<br />

Amplitude Aga<strong>in</strong>st <strong>Pressure</strong> Wave Frequency<br />

Reachable lift<strong>in</strong>g heights are almost <strong>in</strong>dependent of pressure wave frequency.<br />

However, impulse pump<strong>in</strong>g efficiency is directly proportional to the <strong>in</strong>verse of<br />

pressure wave frequency. Consequently, low pressure wave frequencies are preferred<br />

s<strong>in</strong>ce lift<strong>in</strong>g heights rema<strong>in</strong> unchanged, <strong>and</strong> impulse pump<strong>in</strong>g efficiency<br />

is then maximized.<br />

Volume of water displaced by the back <strong>and</strong> forth mov<strong>in</strong>g piston dur<strong>in</strong>g one<br />

pressure wave can be calculated by:<br />

Vd = π d2p ∆l (10.3)<br />

4<br />

where, ∆l, represents piston stroke <strong>and</strong> is calculated us<strong>in</strong>g Eq. 8.6. After<br />

simplifictaions, the displaced flow rate can be expressed by:<br />

qd = π d2<br />

4 KL˙γpWave<br />

(10.4)<br />

The volume of water displaced by the back <strong>and</strong> forth mov<strong>in</strong>g piston, <strong>in</strong> a<br />

production tub<strong>in</strong>g 0.124 m <strong>in</strong> diameter, 2748 m <strong>in</strong> depth, us<strong>in</strong>g pressure<br />

179


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

waves 1 Hz <strong>in</strong> frequency, 412 bar <strong>in</strong> peak-to-peak amplitude then equals 17.8<br />

m 3 m<strong>in</strong> −1 ; that is more than the volume of fluid pumped. Therefore, impulse<br />

pump<strong>in</strong>g is not a volumetric pump<strong>in</strong>g method.<br />

<strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow characteristics that correspond to <strong>in</strong>flow efficiency<br />

only are obta<strong>in</strong>ed from numerical simulations, for selected wellhead pressures<br />

<strong>and</strong> pressure wave peak-to-peak amplitudes, us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves,<br />

1 Hz <strong>in</strong> frequency. <strong>Impulse</strong> pump<strong>in</strong>g <strong>in</strong>flow characteristics for 52 bara, 104<br />

bara <strong>and</strong> 207 bara wellhead pressures are shown <strong>in</strong> Figure 10.9, Figure 10.10<br />

<strong>and</strong> Figure 10.11, respectively. Correspond<strong>in</strong>g pressure wave peak-to-peak<br />

amplitudes are 102 bar, 206 bar <strong>and</strong> 410 bar, respectively.<br />

Maximum reachable lift<strong>in</strong>g heights <strong>in</strong>crease with wellhead pressure. Maximum<br />

reachable heights us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 1 Hz <strong>in</strong> frequency, 102 bar,<br />

206 bar <strong>and</strong> 410 bar <strong>in</strong> peak-to-peak amplitude, are 709 m, 1219 m <strong>and</strong> 2238<br />

m, respectively. Lift<strong>in</strong>g heights decrease with flow rates. Maximum reachable<br />

flow rates are, 2.42 m 3 m<strong>in</strong> −1 , 3.68 m 3 m<strong>in</strong> −1 <strong>and</strong> 5.37 m 3 m<strong>in</strong> −1 , for 52 bara,<br />

104 bara <strong>and</strong> 207 bara wellhead pressures, respectively.<br />

Maximum impulse pump<strong>in</strong>g efficiency decreases with <strong>in</strong>crease <strong>in</strong> wellhead pressure.<br />

Maximum impulse pump<strong>in</strong>g efficiency calculated us<strong>in</strong>g s<strong>in</strong>usoidal pressure<br />

waves, 1 Hz <strong>in</strong> frequency, 102 bar <strong>in</strong> peak-to-peak amplitude, is 0.99 %,<br />

obta<strong>in</strong>ed for 1.82 m 3 m<strong>in</strong> −1 flow rate <strong>and</strong> 199.6 m head (Figure 10.9). Maximum<br />

impulse pump<strong>in</strong>g efficiency calculated us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves,<br />

1 Hz <strong>in</strong> frequency, 102 bar <strong>in</strong> peak-to-peak amplitude, is 0.69 %, obta<strong>in</strong>ed for<br />

1.45 m 3 m<strong>in</strong> −1 flow rate <strong>and</strong> 709.3 m lift<strong>in</strong>g height (Figure 10.10). Maximum<br />

impulse pump<strong>in</strong>g efficiency calculated us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 1 Hz<br />

<strong>in</strong> frequency, 207 bar <strong>in</strong> peak-to-peak amplitude, is 0.69 %, obta<strong>in</strong>ed for 2.63<br />

m 3 m<strong>in</strong> −1 flow rate <strong>and</strong> 1219 m lift<strong>in</strong>g height (Figure 10.11).<br />

180


Head [m]<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

10.2. Comparison of Artificial Lift Techniques<br />

Head<br />

Efficiency<br />

0.4<br />

0.2<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

Flow Rate [l m<strong>in</strong> −1 0 1000 2000 3000 4000 5000 6000<br />

]<br />

0.0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.0<br />

Figure 10.9: <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Characteristics for 102 bara<br />

Wellhead <strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1<br />

Hz <strong>in</strong> Frequency, 52 bar <strong>in</strong> Peak-To-Peak Amplitude<br />

Head [m]<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

Head<br />

Efficiency<br />

0.4<br />

0.2<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

Flow Rate [l m<strong>in</strong> −1 0 1000 2000 3000 4000 5000 6000<br />

]<br />

0.0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.0<br />

Figure 10.10: <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Characteristics for 104 bara<br />

Wellhead <strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1<br />

Hz <strong>in</strong> Frequency, 206 bar <strong>in</strong> Peak-To-Peak Amplitude<br />

181<br />

Efficiency [%]<br />

Efficiency [%]


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Head [m]<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

Head<br />

Efficiency<br />

2.0<br />

1.6<br />

1.2<br />

0.8<br />

0.5<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

Flow Rate [l m<strong>in</strong> −1 0 1000 2000 3000 4000 5000 6000<br />

]<br />

0.0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.0<br />

Figure 10.11: <strong>Impulse</strong> Pump<strong>in</strong>g Inflow Characteristics for 207 bara<br />

Wellhead <strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1<br />

Hz <strong>in</strong> Frequency, 412 bar <strong>in</strong> Peak-To-Peak Amplitude<br />

<strong>Impulse</strong> pump<strong>in</strong>g of aquifer with vertical lift of 709.3 m, 1728.6 m, <strong>and</strong> 2034.4<br />

m is theoretically possible us<strong>in</strong>g s<strong>in</strong>usoidal pressure waves, 1 Hz <strong>in</strong> frequency,<br />

412 bar <strong>in</strong> peak-to-peak amplitude with a 207 bara wellhead pressure (Figure<br />

10.11). Reachable flow rates are approximately 3.81 m 3 m<strong>in</strong> −1 , 1.43 m 3 m<strong>in</strong> −1<br />

<strong>and</strong> 0.65 m 3 m<strong>in</strong> −1 , respectively. Correspond<strong>in</strong>g efficiencies are 0.45 %, 0.41<br />

% <strong>and</strong> 0.24 %, respectively. Calculated flow rate for phase A is approximately<br />

the requested flow rate, whereas calculated flow rates for phases B <strong>and</strong> C are<br />

lower than requested (Table 10.1).<br />

<strong>Impulse</strong> pump<strong>in</strong>g can consequently be applied to Statfjord depressurization<br />

plan. <strong>Impulse</strong> pump<strong>in</strong>g greatest <strong>in</strong>flow efficiency is obta<strong>in</strong>ed us<strong>in</strong>g 207 bara<br />

wellhead pressure, <strong>and</strong> s<strong>in</strong>usoidal pressure waves, 1 Hz <strong>in</strong> frequency, 412 bar<br />

<strong>in</strong> peak-to-peak amplitude. Total efficiency of impulse pump<strong>in</strong>g with such<br />

values can then be calculated us<strong>in</strong>g Eq. 10.2 with a 1 bara outlet pressure,<br />

<strong>and</strong> reachable flow rates calculated previously.<br />

<strong>Impulse</strong> pump<strong>in</strong>g total efficiency for phase A of depressurization of Statfjord<br />

is illustrated <strong>in</strong> Figure 10.12. <strong>Impulse</strong> pump<strong>in</strong>g total efficiency <strong>in</strong>creases with<br />

reachable flow rate, thus with decrease <strong>in</strong> lift<strong>in</strong>g height. <strong>Impulse</strong> pump<strong>in</strong>g<br />

total efficiency <strong>in</strong>creases from 0.5 % at 1983 m, to 1.4 % at 1219 m, <strong>and</strong> to<br />

1.9 % at 5.9 m. <strong>Impulse</strong> pump<strong>in</strong>g is theoretically applicable for the three<br />

182<br />

Efficiency [%]


Head [m]<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

10.3. Possible Improvements of <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency<br />

Head<br />

Efficiency<br />

2.0<br />

1.6<br />

1.2<br />

0.8<br />

0.5<br />

0<br />

0 1000 2000 3000 4000 5000 6000<br />

Flow Rate [l m<strong>in</strong> −1 0 1000 2000 3000 4000 5000 6000<br />

]<br />

0.0<br />

0 1000 2000 3000 4000 5000 6000<br />

0.0<br />

Figure 10.12: Total <strong>Impulse</strong> Pump<strong>in</strong>g Characteristics for 207 bara<br />

Wellhead <strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1<br />

Hz <strong>in</strong> Frequency, 412 bar <strong>in</strong> Peak-To-Peak Amplitude<br />

depressurization phases of the Statfjord field. Nevertheless, the very low total<br />

efficiency makes impulse pump<strong>in</strong>g impractical for deep water.<br />

10.3 Possible Improvements of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Efficiency<br />

<strong>Impulse</strong> pump<strong>in</strong>g reachable flow rates <strong>and</strong> efficiency calculations assume s<strong>in</strong>usoidal<br />

pressure waves. However, other pressure wave waveforms are theoretically<br />

possible. S<strong>in</strong>usoidal, triangular <strong>and</strong> square waveforms are here<strong>in</strong><br />

compared <strong>in</strong> terms of reachable flow rates dur<strong>in</strong>g impulse pump<strong>in</strong>g of water<br />

from a 75 bara constant pressure wellbore, 1000 m deep, us<strong>in</strong>g a water-filled<br />

production tub<strong>in</strong>g, 10 <strong>in</strong>ches <strong>in</strong> diameter. Test case characteristics <strong>and</strong> numerical<br />

simulation parameters are listed <strong>in</strong> Table 10.2.<br />

Spatial traces of s<strong>in</strong>usoidal, square <strong>and</strong> triangular pressure waves propagat<strong>in</strong>g<br />

from wellhead to bottomhole are shown <strong>in</strong> Figure 10.13a), Figure 10.14a),<br />

<strong>and</strong> Figure 10.15a), respectively. Spatial traces of s<strong>in</strong>usoidal, square <strong>and</strong> triangular<br />

pressure waves propagat<strong>in</strong>g back from bottomhole to wellhead after<br />

impulse pump<strong>in</strong>g action are shown <strong>in</strong> Figure 10.13b), Figure 10.14b), <strong>and</strong> Figure<br />

10.15b), respectively. Negative pressure pulses are reflected with a partial<br />

183<br />

Efficiency [%]


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

phase change, <strong>and</strong> a waveform change, due to the <strong>in</strong>let non-return valve open<strong>in</strong>g.<br />

Table 10.2: Description of Test Case for Numerical Investigation of<br />

<strong>Pressure</strong> Waveform Impact on Reachable Flow Rate<br />

L [m] d [<strong>in</strong>ch] pWH [bara] pWB [bara] pWave [bara]<br />

1000 10 100 75 100<br />

˙γ [Hz] TSim [s] a ρ Valve Type<br />

5 1.5 1200 1000 Needle<br />

nX<br />

nT<br />

10000 1000<br />

Depth [m]<br />

a)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

600<br />

700<br />

800<br />

900<br />

1000<br />

0 100 200 300<br />

<strong>Pressure</strong> [bara]<br />

b)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

600<br />

700<br />

800<br />

900<br />

1000<br />

0 100 200 300<br />

<strong>Pressure</strong> [bara]<br />

Figure 10.13: Spatial Traces of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g S<strong>in</strong>usoidal<br />

Waveform<br />

a) Propagat<strong>in</strong>g From Wellhead to Bottomhole<br />

b) Propagat<strong>in</strong>g from Bottomhole to Wellhead<br />

184


Depth [m]<br />

a)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

600<br />

700<br />

800<br />

900<br />

10.3. Possible Improvements of <strong>Impulse</strong> Pump<strong>in</strong>g Efficiency<br />

1000<br />

0 100 200 300<br />

<strong>Pressure</strong> [bara]<br />

b)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

600<br />

700<br />

800<br />

900<br />

1000<br />

0 100 200 300<br />

<strong>Pressure</strong> [bara]<br />

Figure 10.14: Spatial Traces of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g Square<br />

Waveform<br />

Depth [m]<br />

a)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

600<br />

700<br />

800<br />

900<br />

1000<br />

0 100 200 300<br />

<strong>Pressure</strong> [bara]<br />

b)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

600<br />

700<br />

800<br />

900<br />

1000<br />

0 100 200 300<br />

<strong>Pressure</strong> [bara]<br />

Figure 10.15: Spatial Traces of <strong>Impulse</strong> Pump<strong>in</strong>g Us<strong>in</strong>g Triangular<br />

Waveform<br />

a) Propagat<strong>in</strong>g From Wellhead to Bottomhole<br />

b) Propagat<strong>in</strong>g from Bottomhole to Wellhead<br />

185


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Flow Rate [10 3 l m<strong>in</strong> −1 ]<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

Square<br />

S<strong>in</strong>us<br />

Triangle<br />

2<br />

0 2 4 6 8 10<br />

Frequency [Hz]<br />

Figure 10.16: Reachable Flow Rates Aga<strong>in</strong>st <strong>Impulse</strong> Generator<br />

Frequency <strong>and</strong> <strong>Pressure</strong> Wave Waveforms<br />

Test Case Characteristics Described <strong>in</strong> Table 10.2<br />

Evolutions of flow rate us<strong>in</strong>g s<strong>in</strong>usoidal, square <strong>and</strong> triangular waveforms, for<br />

frequency between 1 Hz <strong>and</strong> 8 Hz are shown <strong>in</strong> Figure 10.16. Reachable<br />

flow rates are the greatest with square pressure waves <strong>and</strong> the lowest with<br />

triangular pressure waves, except at 1 Hz where reachable flow rate is 30.2<br />

m 3 m<strong>in</strong> −1 <strong>and</strong> 80.3 m 3 m<strong>in</strong> −1 for s<strong>in</strong>usoidal <strong>and</strong> triangular pressure waves, respectively.<br />

Reachable flow rates with square pressure waves are 189 m 3 m<strong>in</strong> −1 ,<br />

187 m 3 m<strong>in</strong> −1 , 185 m 3 m<strong>in</strong> −1 <strong>and</strong> 176 m 3 m<strong>in</strong> −1 , at 1 Hz, 2 Hz, 4 Hz <strong>and</strong> 8<br />

Hz, respectively.<br />

Evolution of reachable flow rate aga<strong>in</strong>st impulse generator frequency for s<strong>in</strong>usoidal<br />

pressure waves shows that maximum flow rates are obta<strong>in</strong>ed for<br />

steep wavefronts. Reachable flow rates with s<strong>in</strong>usoidal pressure waves are<br />

30.2 m 3 m<strong>in</strong> −1 , 103 m 3 m<strong>in</strong> −1 , 150 m 3 m<strong>in</strong> −1 <strong>and</strong> 149 m 3 m<strong>in</strong> −1 , at 1 Hz,<br />

2 Hz, 4 Hz <strong>and</strong> 8 Hz, respectively. Greater impulse generator frequencies<br />

<strong>in</strong>duce steeper s<strong>in</strong>usoidal pressure wavefronts. Another way to express this<br />

is that s<strong>in</strong>usoidal pressure wavefronts are more moderate when wavelength<br />

<strong>in</strong>creases.<br />

A drop <strong>in</strong> reachable flow rate for the considered test case is observed for<br />

impulse pump<strong>in</strong>g with s<strong>in</strong>usoidal pressure waves 5 Hz <strong>in</strong> frequency <strong>in</strong> Figure<br />

10.16. Reachable flow rate goes from to 150 m 3 m<strong>in</strong> −1 at 4 Hz, to 80<br />

m 3 m<strong>in</strong> −1 at 6 Hz <strong>and</strong> to 150 m 3 m<strong>in</strong> −1 at 8 Hz. Such drop <strong>in</strong> reachable<br />

186


10.4. Cost of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

flow rate is observed only with s<strong>in</strong>usoidal pressure waves <strong>and</strong> can be expla<strong>in</strong>ed<br />

by oscillations of the <strong>in</strong>let non-return valve when the <strong>in</strong>flow water hammer<br />

pressure wave counteracts the negative pressure pulse open<strong>in</strong>g effect.<br />

Greater impulse generator frequency <strong>and</strong> square shaped pressure waves can<br />

give rise to greater reachable flow rates. Hence, to greater impulse pump<strong>in</strong>g<br />

efficiency. Different waveforms can be obta<strong>in</strong>ed us<strong>in</strong>g a back-<strong>and</strong>-forth oscillat<strong>in</strong>g<br />

piston as an impulse generator by chang<strong>in</strong>g the piston velocity comm<strong>and</strong>.<br />

Ideal square velocity comm<strong>and</strong> are not atta<strong>in</strong>able <strong>in</strong> practice. However, velocity<br />

comm<strong>and</strong>s approach<strong>in</strong>g square shaped wavefronts can be implemented<br />

practically.<br />

10.4 Cost of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Efficiency is only one criteria for select<strong>in</strong>g artificial lift technique. Pedersen<br />

(June 2007) concluded that ESP were the most efficient artificial lift technique<br />

for the three depressurization phases of the Statfjord field, with efficiency of<br />

38 % for phase A, <strong>and</strong> 44 % for phases B <strong>and</strong> C. Jet pump efficiencies for the<br />

same conditions were 31 % for phase A, 26 % for phase B <strong>and</strong> 24 % for phase<br />

C. However, ESP apparatus need to be shifted every two years on average,<br />

whereas jet pumps can operate for longer periods without shift<strong>in</strong>g out. In<br />

addition, ESP apparatus are more expensive to buy than jet pumps.<br />

Capital <strong>and</strong> operational expenditures must be considered dur<strong>in</strong>g selection of<br />

an artificial lift technique. Capital expenditures, CAPEX, relate to buy<strong>in</strong>g<br />

costs whereas operational expenditures, OPEX, relate to ma<strong>in</strong>tenance <strong>and</strong><br />

shift<strong>in</strong>g of apparatus. OPEX related to impulse pump<strong>in</strong>g are lower than with<br />

ESP or jet pumps s<strong>in</strong>ce external energy supply requirements are located at<br />

wellhead only. In addition, fluid does not go through the impulse generator;<br />

that is, shift<strong>in</strong>g of impulse generator can be performed without shutt<strong>in</strong>g the<br />

production tub<strong>in</strong>g or chang<strong>in</strong>g pressure conditions.<br />

<strong>Impulse</strong> pump<strong>in</strong>g greatest efficiency is obta<strong>in</strong>ed for artificial lift of shallow<br />

water. <strong>Impulse</strong> pump<strong>in</strong>g applied to deep water is theoretically possible despite<br />

its low efficiency. Nevertheless, impulse pump<strong>in</strong>g low efficiency can be<br />

compensated by low OPEX <strong>and</strong> especially advantages related to ma<strong>in</strong>tenance.<br />

Additionally, cost analysis can help compromis<strong>in</strong>g on impulse pump<strong>in</strong>g design.<br />

Production tub<strong>in</strong>g parameter is an important impulse pump<strong>in</strong>g design criteria.<br />

<strong>Impulse</strong> pump<strong>in</strong>g flow rate is directly proportional to production tub<strong>in</strong>g<br />

diameter, but efficiency is proportional to flow rate, <strong>and</strong> <strong>in</strong>versely proportional<br />

to production tub<strong>in</strong>g square. <strong>Impulse</strong> pump<strong>in</strong>g total efficiency for phase A of<br />

depressurization of Statfjord studied above is illustrated for production tub<strong>in</strong>g<br />

1 <strong>in</strong>ch, 2 <strong>in</strong>ch, 4 <strong>in</strong>ch, <strong>and</strong> 8 <strong>in</strong>ch <strong>in</strong> diameter <strong>in</strong> Figure 10.17.<br />

187


Chapter 10. Characteristic Curves of <strong>Impulse</strong> Pump<strong>in</strong>g<br />

Efficiency [%]<br />

1.0<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

d = 1 <strong>in</strong>ch<br />

d = 2 <strong>in</strong>ch<br />

d = 4 <strong>in</strong>ch<br />

d = 8 <strong>in</strong>ch<br />

0.0<br />

0 20 40 60 80 100<br />

Flow Rate [10 3 l m<strong>in</strong> −1 ]<br />

Figure 10.17: Total <strong>Impulse</strong> Pump<strong>in</strong>g Characteristics for 207 bara<br />

Wellhead <strong>Pressure</strong>, Us<strong>in</strong>g S<strong>in</strong>usoidal <strong>Pressure</strong> <strong>Waves</strong>, 1<br />

Hz <strong>in</strong> Frequency, 412 bar <strong>in</strong> Peak-To-Peak Amplitude,<br />

Production Tub<strong>in</strong>g 1 <strong>in</strong>ch, 2 <strong>in</strong>ch, 4 <strong>in</strong>ch, <strong>and</strong> 8 <strong>in</strong>ch<br />

<strong>in</strong> Diameter<br />

Maximum flow rate <strong>in</strong>creases l<strong>in</strong>early from 1100 l m<strong>in</strong> −1 , to 8800 l m<strong>in</strong> −1<br />

for production tub<strong>in</strong>g 1 <strong>in</strong>ch <strong>and</strong> 8 <strong>in</strong>ch <strong>in</strong> diameter, respectively. Whereas<br />

impulse pump<strong>in</strong>g total efficiency decreases from 0.9 % to 0.11 %, for production<br />

tub<strong>in</strong>g 1 <strong>in</strong>ch <strong>and</strong> 8 <strong>in</strong>ch <strong>in</strong> diameter, respectively. A compromise between flow<br />

rate <strong>and</strong> efficiency must therefore be reached, <strong>and</strong> cost analysis can provide a<br />

selection criteria.<br />

188


Chapter 11<br />

Discussion<br />

<strong>Impulse</strong> pump<strong>in</strong>g physical pr<strong>in</strong>ciples are based on well established pressure<br />

wave propagation, transmission <strong>and</strong> reflection characteristics. The description<br />

of impulse pump<strong>in</strong>g physical pr<strong>in</strong>ciples consequently constitutes a solid<br />

mechanistic model. However, a number of uncerta<strong>in</strong>ties arise from modell<strong>in</strong>g<br />

details <strong>in</strong> general, <strong>and</strong> numerical simulation <strong>in</strong> particular.<br />

Inlet <strong>and</strong> outlet non-return valve mechanical characteristics were assumed from<br />

geometrical considerations <strong>and</strong> static pressure on the upstream. However, the<br />

<strong>in</strong>let non-return valve actuation result from reflection of a negative pressure<br />

pulse <strong>and</strong> <strong>in</strong>duces a positive water hammer pressure wave. Another way to<br />

express this is that the <strong>in</strong>let non-return valve upstream pressure is dynamic<br />

<strong>and</strong> depends on the pressure gradient across the valve.<br />

Non-return valve mechanical characteristics uncerta<strong>in</strong>ties <strong>in</strong>fluence impulse<br />

pump<strong>in</strong>g flow rate estimations. <strong>Impulse</strong> pump<strong>in</strong>g flow rates are estimated<br />

from numerical simulation results, based on uncerta<strong>in</strong> non-return valve mechanical<br />

characteristics. However, comparisons between numerical simulation<br />

<strong>and</strong> experimental data for horizontal transport of water performed with an<br />

unknown valve design, proved to be <strong>in</strong> good agreement.<br />

<strong>Pressure</strong> wave attenuation is another uncerta<strong>in</strong>ty that <strong>in</strong>fluences impulse pump<strong>in</strong>g<br />

flow rate estimations. <strong>Impulse</strong> pump<strong>in</strong>g flow rate depends on negative pressure<br />

pulse amplitude at bottomhole; that is, after propagation <strong>in</strong> the fluid-filled<br />

production tub<strong>in</strong>g. However, numerical simulation was performed <strong>in</strong>clud<strong>in</strong>g<br />

steady state friction only s<strong>in</strong>ce other attenuation models are <strong>in</strong>accurate. Another<br />

way to express this is that estimated flow rate values are optimistic.<br />

<strong>Impulse</strong> pump<strong>in</strong>g critical parameters are pressure wave amplitude, wellhead<br />

pressure, lift<strong>in</strong>g height, fluid compressibility <strong>and</strong> volume occupied by the fluid.<br />

However, the fluid volume <strong>in</strong>fluences differently the flow rate <strong>and</strong> the impulse<br />

pump<strong>in</strong>g efficiency. Increase <strong>in</strong> diameter <strong>in</strong>duces a greater flow rate, but the<br />

189


Chapter 11. Discussion<br />

greater the volume, the greater the energy required to generate a pressure wave<br />

at the wellhead. <strong>Impulse</strong> pump<strong>in</strong>g efficiency is proportional to the <strong>in</strong>verse of<br />

the square diameter.<br />

Production tub<strong>in</strong>g diameter is an important impulse pump<strong>in</strong>g design criteria.<br />

However, a compromise must be found between flow rate <strong>and</strong> efficiency. Such<br />

compromise depends on every situation where impulse pump<strong>in</strong>g is applied,<br />

<strong>and</strong> represents also an economical criteria.<br />

Comparison between experimental data <strong>and</strong> results from numerical simulations<br />

were <strong>in</strong> good agreement. Nevertheless, comparisons were based on a<br />

limited number of experimental data for horizontal transport of water. More<br />

experimental work are therefore necessary for artificial lift situations. Experimental<br />

work is also required to <strong>in</strong>vestigate pressure wave attenuation <strong>in</strong><br />

production tub<strong>in</strong>gs.<br />

A new approach to pressure wave attenuation based on acoustic velocity<br />

changes with fluid velocity with<strong>in</strong> a pressure wave has been proposed. However,<br />

uncerta<strong>in</strong>ties l<strong>in</strong>ked to the proposed new approach must be assessed both<br />

experimentally <strong>and</strong> us<strong>in</strong>g numerical simulations.<br />

190


Chapter 12<br />

Conclusions<br />

<strong>Impulse</strong> pump<strong>in</strong>g generates flow from bottomhole to wellhead us<strong>in</strong>g pressure<br />

waves generated at wellhead. A positive pressure pulse first actuates a near<br />

wellhead non-return valve that <strong>in</strong>duces outflow from the production tub<strong>in</strong>g.<br />

A negative pressure pulse then actuates a bottomhole <strong>in</strong>let non-return valve<br />

that <strong>in</strong>duces <strong>in</strong>flow to the production tub<strong>in</strong>g. <strong>Impulse</strong> pump<strong>in</strong>g def<strong>in</strong>es a new<br />

pump<strong>in</strong>g method <strong>in</strong> addition to dynamic <strong>and</strong> volumetric pump<strong>in</strong>g methods.<br />

Artificial lift of s<strong>in</strong>gle phase liquids is theoretically possible without lift<strong>in</strong>g<br />

height limitation us<strong>in</strong>g impulse pump<strong>in</strong>g. The lift<strong>in</strong>g height is directly proportional<br />

to the negative pressure wave amplitude that depends on wellhead<br />

pressure. Thus, a one bar <strong>in</strong>crease <strong>in</strong> negative peak pressure that can be <strong>in</strong>duced<br />

by a one bara <strong>in</strong>crease <strong>in</strong> wellhead pressure, can generate a 10 m water<br />

lift<strong>in</strong>g height <strong>in</strong>crease.<br />

<strong>Impulse</strong> pump<strong>in</strong>g greatest efficiency is obta<strong>in</strong>ed for artificial lift of shallow water.<br />

An estimated 25 % efficiency can be reached for 100 m water lift<strong>in</strong>g height<br />

accord<strong>in</strong>g to the presented performance model. <strong>Impulse</strong> pump<strong>in</strong>g performance<br />

depends on pressure wave amplitude, wellhead pressure, lift<strong>in</strong>g height, fluid<br />

compressibility <strong>and</strong> volume occupied <strong>in</strong> the production tub<strong>in</strong>g. The only effect<br />

of frequency is on impulse pump<strong>in</strong>g efficiency.<br />

<strong>Impulse</strong> pump<strong>in</strong>g of two-phase oil <strong>and</strong> gas fluids, or s<strong>in</strong>gle phase live oils, are<br />

not recommended due to cavitation. <strong>Impulse</strong> pump<strong>in</strong>g of two-phase water<br />

<strong>and</strong> oil fluids, or dead oils, are possible, yet with low efficiency. <strong>Impulse</strong><br />

pump<strong>in</strong>g of deep water wells is also possible, yet with low efficiency due to<br />

power requirements for generat<strong>in</strong>g pressure waves <strong>in</strong> nearly <strong>in</strong>compressible<br />

fluids.<br />

One-dimensional pressure wave attenuation models based on transient friction<br />

<strong>and</strong> pipel<strong>in</strong>e viscoelasticity are <strong>in</strong>accurate. Modell<strong>in</strong>g of transient friction are<br />

not physically correct <strong>and</strong> matches between experimental data <strong>and</strong> results<br />

191


Chapter 12. Conclusions<br />

from numerical simulations us<strong>in</strong>g the method of characteristics are purely<br />

co<strong>in</strong>cidental.<br />

F<strong>in</strong>ite volume methods can simulate pressure transient events with greater<br />

accuracy than the well established method of characteristics. F<strong>in</strong>ite volume<br />

methods have less <strong>in</strong>herent numerical dissipation than the method of characteristics.<br />

<strong>Pressure</strong> signals <strong>and</strong> flow velocities dur<strong>in</strong>g impulse pump<strong>in</strong>g can<br />

therefore be simulated us<strong>in</strong>g a f<strong>in</strong>ite volume numerical scheme.<br />

192


References<br />

Abhulimen, K. E. <strong>and</strong> Susu, A. A. (2004). Liquid Pipel<strong>in</strong>e Leak Detection System:<br />

Model Development <strong>and</strong> Numerical Simulation, Chemical Eng<strong>in</strong>eer<strong>in</strong>g<br />

Journal 97(1): 47–67.<br />

Adamkowski, A. <strong>and</strong> Lew<strong>and</strong>owski, M. (2006). Experimental Exam<strong>in</strong>ation of<br />

Unsteady Friction Models for Transient Pipe Flow Simulation, Journal of<br />

Fluids Eng<strong>in</strong>eer<strong>in</strong>g 128: 1351–1363.<br />

AGA-10 (2003). Report No. 10: Speed of Sound <strong>in</strong> Natural Gas <strong>and</strong> Other<br />

Related Hydrocarbon Gases, American Gas Association. Lori Traweek <strong>and</strong><br />

Ali Quraishi.<br />

AGA-8 (1992). Report No. 8: Compressibility Factors of Natural Gas <strong>and</strong><br />

Other Related Hydrocarbon Gases, American Gas Association. Jeffrey L.<br />

Savidge <strong>and</strong> Kenneth E. Starl<strong>in</strong>g.<br />

AGA-9 (2007). Report No. 9: Measurement of Gas by Multipath Ultrasonic<br />

Meters, American Gas Association. Lori Traweek <strong>and</strong> Ali Quraishi.<br />

ASME B31 (2003). Process Pip<strong>in</strong>g, ASME Code for <strong>Pressure</strong> Pip<strong>in</strong>g B31,<br />

American Society of Mechanical Eng<strong>in</strong>eer.<br />

Axworthy, D. H., Ghidaoui, M. S. <strong>and</strong> McInnis, D. A. (2000). Extended<br />

Thermodynamics Derivation of Energy Dissipation <strong>in</strong> Unsteady Pipe Flow,<br />

Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 126(4): 276–287.<br />

Beck, S. B., Curren, M. D., Sims, N. D. <strong>and</strong> Stanway, R. (2005). Pipel<strong>in</strong>e<br />

Network Features <strong>and</strong> Leak Detection by Cross-Correlation Analysis of Reflected<br />

<strong>Waves</strong>, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 131(8): 715–723.<br />

Bergant, A. <strong>and</strong> Tijssel<strong>in</strong>g, A. (2001). Parameters Affect<strong>in</strong>g Water Hammer<br />

Wave Attenuation, Shape <strong>and</strong> Tim<strong>in</strong>g, IAHR 10th International Meet<strong>in</strong>g<br />

of the Work Group on the Behaviour of Hydraulic Mach<strong>in</strong>ery Under Steady<br />

Oscillatory Conditions, Trondheim, Norway.<br />

193


Bergant, A., Tijssel<strong>in</strong>g, A., Vitkovsky, J., Covas, D., Simpson, A. <strong>and</strong> Lambert,<br />

M. (2003a). Further Investigation of Parameters Affect<strong>in</strong>g Water Hammer<br />

Wave Attenuation, Shape <strong>and</strong> Tim<strong>in</strong>g - Part 1: Mathematical Tools,<br />

11th International Meet<strong>in</strong>g of the IAHR Work Group on the Behaviour of<br />

Hydraulic Mach<strong>in</strong>ery Under Steady Oscillatory Conditions, Stuttgart, Germany.<br />

Bergant, A., Tijssel<strong>in</strong>g, A., Vitkovsky, J., Covas, D., Simpson, A. <strong>and</strong> Lambert,<br />

M. (2003b). Further Investigation of Parameters Affect<strong>in</strong>g Water Hammer<br />

Wave Attenuation, Shape <strong>and</strong> Tim<strong>in</strong>g - Part 2: Case Study, 11th International<br />

Meet<strong>in</strong>g of the IAHR Work Group on the Behaviour of Hydraulic<br />

Mach<strong>in</strong>ery Under Steady Oscillatory Conditions, Stuttgart, Germany.<br />

Berrebi, J., Mart<strong>in</strong>sson, P.-E., Willatzen, M. <strong>and</strong> Dels<strong>in</strong>g, J. (2004). Ultrasonic<br />

Flow Meter<strong>in</strong>g Errors due to Pulsat<strong>in</strong>g Flow, Flow Measurement <strong>and</strong><br />

Instrumentation 15(3): 179–185.<br />

Boge, R., Lien, S. K., Gjesdal, A. <strong>and</strong> Hansen, A. G. (2005). Turn<strong>in</strong>g a North<br />

Sea Oil Giant Into a Gas Field - Depressurization of the Statfjord Field,<br />

Offshore Europe, Aberdeen, Scotl<strong>and</strong>. SPE 96403.<br />

Brunone, B. (1999). Transient Test-Based Technique for Leak Detection<br />

<strong>in</strong> Outfall Pipes, Journal of Water Resources Plann<strong>in</strong>g <strong>and</strong> Management<br />

125(5): 302–306.<br />

Brunone, B., Golia, U. M. <strong>and</strong> Greco, M. (1991). Modell<strong>in</strong>g of Fast Transients<br />

by Numerical Methods, International Meet<strong>in</strong>g on Hydraulic Transients with<br />

Column Separation. 9th Round Table, IAHR, Valencia, Spa<strong>in</strong>, Valencia,<br />

Spa<strong>in</strong>.<br />

Brunone, B., Golia, U. M. <strong>and</strong> Greco, M. (1995). Effects of Two Dimensionality<br />

on Pipe Transients Modell<strong>in</strong>g, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g<br />

121(12): 906–912.<br />

Brunone, B., Karney, B. W., Mecarelli, M. <strong>and</strong> Ferrante, M. (2000). Velocity<br />

Profile <strong>and</strong> Unsteady Pipe Friction <strong>in</strong> Transient Flow, Journal of Water<br />

Resources Plann<strong>in</strong>g <strong>and</strong> Management 126(4): 236–244.<br />

Carstens, M. R. <strong>and</strong> Roller, J. E. (1959). Boundary Shear Stress <strong>in</strong> Unsteady<br />

Turbulent Pipe Flow, Journal of the Hydraulics Division 85(2): 67–81.<br />

Cheremis<strong>in</strong>off, N. P. (1996). Materials Selection Deskbook, William Andrew<br />

Publish<strong>in</strong>g.<br />

Chevray, R. <strong>and</strong> Mathieu, J. (1993). Topics <strong>in</strong> Fluid Mechanics, Cambridge<br />

University Press.<br />

194


Contractor, D. N. (1965). The Reflection of Waterhammer <strong>Pressure</strong> <strong>Waves</strong><br />

From M<strong>in</strong>or Losses, Journal of Basic Eng<strong>in</strong>eer<strong>in</strong>g 87: 445–452.<br />

Covas, D. (2003). Inverse Transient Analysis for Leak Detection <strong>and</strong> Calibration<br />

of Water Pipe Systems Modell<strong>in</strong>g Special Dynamic Effects, PhD thesis,<br />

Department of Civil Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Environmental Eng<strong>in</strong>eer<strong>in</strong>g- Imperial<br />

College of Science, Technology <strong>and</strong> Medic<strong>in</strong>e.<br />

Covas, D., Stoianov, I., Mano, J. F., Ramos, H., Graham, N. <strong>and</strong> Maksimovic,<br />

C. (2004). The Dynamic Effect of Pipe-Wall Viscoelasticity <strong>in</strong> Hydraulic<br />

Transients. Part 1: Experimental Analysis <strong>and</strong> Creep Characterization,<br />

Journal of Hydraulic Research 42(5): 516–530.<br />

Covas, D., Stoianov, I., Mano, J. F., Ramos, H., Graham, N. <strong>and</strong> Maksimovic,<br />

C. (2005). The Dynamic Effect of Pipe-Wall Viscoelasticity <strong>in</strong> Hydraulic<br />

Transients. Part 2: Model Development, Calibration <strong>and</strong> Verification, Journal<br />

of Hydraulic Research 43(1): 56–70.<br />

Daily, J. W., Hankey, W. L., Olive, R. W. <strong>and</strong> Jordaan, J. M. (1956). Resistance<br />

Coefficients for Accelerated <strong>and</strong> Decelerated Flows Through Smooth<br />

Tubes <strong>and</strong> Orifices, Journal of Basic Eng<strong>in</strong>eer<strong>in</strong>g 78(7): 1071–1077.<br />

Dake, L. P. (1978). Fundamentals of Reservoir Eng<strong>in</strong>eer<strong>in</strong>g, Elsevier.<br />

DelGrosso, V. A. (1974). New Equation for the Speed of Sound <strong>in</strong> Natural Waters<br />

(with Comparisons to Other Equations), The Journal of the Acoustical<br />

Society of America 56(4): 1084–1091.<br />

Dong, L. <strong>and</strong> Gudmundsson, J. S. (1993). Model for Sound Speed <strong>in</strong> Multiphase<br />

Mixtures, 3rd Lerkendal Petroleum Eng<strong>in</strong>eer<strong>in</strong>g Workshop, Trondheim,<br />

Norway.<br />

Eisenberg, D. <strong>and</strong> Kauzman, W. (1969). The Structure <strong>and</strong> Properties of<br />

Water, Oxford University Press.<br />

Erickson, D. <strong>and</strong> Twaite, D. (1996). Pipel<strong>in</strong>e Integrity Monitor<strong>in</strong>g System<br />

for Leak Detection, Control, <strong>and</strong> Optimization of Wet Gas Pipel<strong>in</strong>es, SPE<br />

Annual Technical Conference <strong>and</strong> Exhibition, Denver, Colorado. SPE 36607.<br />

Ferrante, M. <strong>and</strong> Brunone, B. (2003a). Pipe System Diagnosis <strong>and</strong> Leak<br />

Detection by Unsteady-State Tests. 1: Harmonic Analysis, Advances <strong>in</strong><br />

Water Ressources 26(1): 95–105.<br />

Ferrante, M. <strong>and</strong> Brunone, B. (2003b). Pipe System Diagnosis <strong>and</strong> Leak Detection<br />

by Unsteady-State Tests. 2: Wavelet Analysis, Advances <strong>in</strong> Water<br />

Ressources 26(1): 107–116.<br />

195


Ghidaoui, M. S., Zhao, M., McInnis, D. A. <strong>and</strong> Axworthy, D. H. (2005). A<br />

Review of Water Hammer Theory <strong>and</strong> Practice, Applied Mechanics Review<br />

58(1): 49–76.<br />

Gudmundsson, J. S., Dong, L. <strong>and</strong> Rodoy, F. (1992). Two Phase Flow Meter<strong>in</strong>g<br />

by <strong>Pressure</strong> Pulse Propagation, 67th Annual Technical Conference <strong>and</strong><br />

Exhibition, Wash<strong>in</strong>gton D.C.<br />

Gudmundsson, J. S., Durgut, I., Celius, H. K. <strong>and</strong> Korsan, K. (2000). Detection<br />

<strong>and</strong> Monitor<strong>in</strong>g of Deposits <strong>in</strong> Multiphase Flow Pipel<strong>in</strong>es Us<strong>in</strong>g <strong>Pressure</strong><br />

Pulse Technology, 12th International Oil Field Chemistry Symposium,<br />

Geilo, Norway.<br />

Gudmundsson, J. S., Durgut, I., Ronnevig, J., Korsan, K. <strong>and</strong> Celius, H. K.<br />

(2001). <strong>Pressure</strong> Pulse Analysis of Gas Lift Wells, ASME/API Gas Lift<br />

Workshop, Aberdeen, Scotl<strong>and</strong>.<br />

Haestad Methods Solution Center (2008). Bentley HAMMER V8i User’s<br />

Guide.<br />

H<strong>in</strong>o, M., Sawamoto, M. <strong>and</strong> Takasu, S. (1976). Experiments on Transition<br />

to Turbulence <strong>in</strong> an Oscillatory Pipe Flow, Journal of Fluid Mechanics<br />

75(2): 193–207.<br />

Holmboe, E. L. <strong>and</strong> Rouleau, W. T. (1967). The Effect of Viscous Shear on<br />

Transients <strong>in</strong> Liquid L<strong>in</strong>es, Journal of Basic Eng<strong>in</strong>eer<strong>in</strong>g 89(1): 174–180.<br />

Hydraulic Institute (2004). Optimiz<strong>in</strong>g Pump<strong>in</strong>g Systems, Pump Systems Matter<br />

<strong>and</strong> Hydraulic Institute.<br />

Jacobsen, E. (2008). Private Communication. StatoilHydro ASA, Fiscal Meter<strong>in</strong>g<br />

<strong>and</strong> Field Allocation.<br />

Jonsson, L. <strong>and</strong> Larson, M. (1992). Leak Detection Through Hydraulic Transient<br />

Analysis, Pipel<strong>in</strong>e Systems (-): 273–286.<br />

Katz, D. L. <strong>and</strong> Lee, R. L. (1990). Natural Gas Eng<strong>in</strong>eer<strong>in</strong>g, Production <strong>and</strong><br />

Storage, McGraw-Hill.<br />

Laney, C. B. (1998). Computational Gasdynamics, Cambridge University<br />

Press.<br />

Lee, P. J., Vitkovsky, J. P., Lambert, M. F., Simpson, A. R. <strong>and</strong> Liggett, J.<br />

(2007). Leak Location <strong>in</strong> Pipel<strong>in</strong>es Us<strong>in</strong>g the <strong>Impulse</strong> Response Function,<br />

Journal of Hydraulic Research 45(5): 643–652.<br />

Leighton, T. G. (1994). The Acoustic Bubble, Academic Press.<br />

196


LeVeque, R. J. (2002). F<strong>in</strong>ite Volume Methods for Hyperbolic Problems, Cambridge<br />

University Press.<br />

Liggett, J. A. <strong>and</strong> Chen, L.-C. (1994). Inverse Transient Analysis <strong>in</strong> Pipe<br />

Networks, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 120(8): 934–955.<br />

Lighthill, J. (1978). <strong>Waves</strong> <strong>in</strong> Fluids, Cambridge University Press.<br />

McCa<strong>in</strong>, W. D. (1990). The Properties of Petroleum Fluids, Pennwell Publish<strong>in</strong>g<br />

Company.<br />

Michaelides, E. E. (1981). Thermodynamic Properties of Geothermal Fluids,<br />

Geothermal Resources Council 5: 361–364.<br />

Moody, L. F. (1944). Friction Factors for Pipe Flow, Transactions of the<br />

A.S.M.E. 66(8): 671–677.<br />

Nathan, G. K., Tan, J. K. <strong>and</strong> Ng, K. C. (1988). Two-Dimensional Analysis<br />

of <strong>Pressure</strong> Transients <strong>in</strong> Pipel<strong>in</strong>es, International Journal for Numerical<br />

Methods <strong>in</strong> Fluids 8(3): 339–349.<br />

Nixon, W. <strong>and</strong> Ghidaoui, M. S. (2007). Numerical Sensitivity Study of Unsteady<br />

Friction <strong>in</strong> Simple Systems with External Flows, Journal of Hydraulic<br />

Eng<strong>in</strong>eer<strong>in</strong>g 133(7): 736–749.<br />

Pa<strong>in</strong>, H. J. (2005). The Physics of Vibrations <strong>and</strong> <strong>Waves</strong>, John Wiley & Sons.<br />

Pedersen, K. S., Fredenslund, A. <strong>and</strong> Thomassen, P. (1989). Properties of Oils<br />

<strong>and</strong> Natural Gases, Gulf Publish<strong>in</strong>g Company.<br />

Pedersen, P. J. (June 2007). Artificial Water Lift at Statfjord - A Case Study of<br />

Offshore Artificial Lift Methods, Master’s thesis, Department of Petroleum<br />

Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Applied Geophysics, <strong>NTNU</strong>, Trondheim, Norway.<br />

Peng, D.-Y. <strong>and</strong> Rob<strong>in</strong>son, D. B. (1976). A New Two-Constant Equation of<br />

State, Industrial & Eng<strong>in</strong>eer<strong>in</strong>g Chemistry Fundamentals 15(1): 59–64.<br />

Pezz<strong>in</strong>ga, G. (1999). Quasi-2D Model for Unsteady Flow <strong>in</strong> Pipe Networks,<br />

Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 125(7): 676–685.<br />

Pothof, I. <strong>and</strong> Clemens, F. (2008). Detection of Gas Pockets <strong>in</strong> Dendritic<br />

Wastewater Ma<strong>in</strong>s, 10th International Conference on <strong>Pressure</strong> Surges, Ed<strong>in</strong>burgh,<br />

Scotl<strong>and</strong>. BHR Group.<br />

Ramos, H., Covas, D., Borga, A. <strong>and</strong> Loureiro, D. (2004). Surge Damp<strong>in</strong>g<br />

Analysis <strong>in</strong> Pipe Systems: Modell<strong>in</strong>g <strong>and</strong> Experiments, Journal of Hydraulic<br />

Research 42(4): 413–425.<br />

197


Safwat, H. <strong>and</strong> Polder, J. (1973). Friction Frequency Dependence for<br />

Oscillatory Flows <strong>in</strong> Circular Pipes, Journal of the Hydraulics Division<br />

99(1): 1933–1945.<br />

Sagov, M. (2004). Method <strong>and</strong> Apparatus for Transport<strong>in</strong>g Fluid <strong>in</strong> a Conduit,<br />

Norwegian Patent No. 20045382. <strong>in</strong> Norwegian.<br />

Savidge, J. L., Starl<strong>in</strong>g, K. E. <strong>and</strong> McFall, R. L. (1988). Sound Speed of<br />

Natural Gas, SPE Gas Technology Symposium, SPE, Dalas, Texas. SPE<br />

18396.<br />

Scott, S. L. <strong>and</strong> Yi, J. (1998). Detection of Critical Flow Leaks <strong>in</strong> Deepwater<br />

Gas Flowl<strong>in</strong>es, SPE Annual Technical Conference <strong>and</strong> Exhibition, New<br />

Orleans, Louisiana. SPE 49310.<br />

Shuy, E. B. (1996). Wall Shear Stress <strong>in</strong> Accelerat<strong>in</strong>g <strong>and</strong> Decelerat<strong>in</strong>g Turbulent<br />

Pipe Flows, Journal of Hydraulic Research 34(2): 173–183.<br />

Silva-Araya, W. F. <strong>and</strong> Chaudry, M. H. (1997). Computation of Energy Dissipation<br />

<strong>in</strong> Transient Flow, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 123(2): 108–<br />

115.<br />

Soares, A. K., Melo, N., Goncalves, F. V., Covas, D. I. <strong>and</strong> Ramos, H. M.<br />

(2008). Hydraulic Modell<strong>in</strong>g of <strong>Pressure</strong> Surges <strong>in</strong> a Water Pump<strong>in</strong>g System,<br />

24th Symposium on Hydraulic Mach<strong>in</strong>ery <strong>and</strong> Systems, IAHR, pp. 1–16.<br />

Stallard, J. M., Rosenbaum, I. J. <strong>and</strong> Davis, C. M. (1969). Ultrasonic Method<br />

for Determ<strong>in</strong><strong>in</strong>g Thermodynamic Properties of Liquids with Results for Mercury,<br />

The Journal of the Acoustical Society of America 45(3): 583–586.<br />

Szymkiewicz, R. <strong>and</strong> Mitosek, M. (2005). Analysis of Unsteady Pipe Flow<br />

Us<strong>in</strong>g the Modified F<strong>in</strong>ite Element Method, Communications <strong>in</strong> Numerical<br />

Methods <strong>in</strong> Eng<strong>in</strong>eer<strong>in</strong>g 21: 183–199.<br />

Szymkiewicz, R. <strong>and</strong> Mitosek, M. (2007). Numerical Aspects of Improvement<br />

of the Unsteady Pipe Flow Equations, International Journal for Numerical<br />

Methods <strong>in</strong> Fluids 55: 1039–1058.<br />

Thorley, A. D. (1991). Fluid Transients <strong>in</strong> Pipel<strong>in</strong>e Systems, D. & L. George<br />

Ltd., Herts, Engl<strong>and</strong>.<br />

Tijssel<strong>in</strong>g, A. <strong>and</strong> Anderson, A. (2007). Johannes von Kries <strong>and</strong> the History<br />

of Water Hammer, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 133(1): 1–8.<br />

Trikha, A. K. (1975). An Efficient Method for Simulat<strong>in</strong>g Frequency-<br />

Dependent Friction <strong>in</strong> Transient Liquid Flow, Journal of Fluids Eng<strong>in</strong>eer<strong>in</strong>g<br />

97(1): 97–105.<br />

198


Uchida, S. (1956). The Pulsat<strong>in</strong>g Viscous Flow Superposed on the Steady<br />

Lam<strong>in</strong>ar Motion of Incompressible Fluid <strong>in</strong> a Circular Pipe, Journal of Applied<br />

Mathematics <strong>and</strong> Physics ZAMP 7(5): 403–422.<br />

Vardy, A. E. (2008). Method of Characteristics <strong>in</strong> Quasi-Steady Compressible<br />

Flows, Surge Analysis - System Design, Simulation, Monitor<strong>in</strong>g <strong>and</strong><br />

Control. 10th International Conference on <strong>Pressure</strong> Surges, BHR Group,<br />

pp. 505–518.<br />

Vardy, A. E. <strong>and</strong> Hwang, K.-L. (1991). A Characteristics Model of Transient<br />

Friction <strong>in</strong> Pipes, Journal of Hydraulic Research 29(5): 669–684.<br />

Vardy, A. E., Hwang, K.-L. <strong>and</strong> Brown, J. M. (1993). A Weight<strong>in</strong>g Function<br />

Model of transient Turbulent Pipe Friction, Journal of Hydraulic Research<br />

31(4): 533–548.<br />

Vitkovsky, J. P., Bergant, A., Simpson, A. R. <strong>and</strong> Lambert, M. F. (2006).<br />

Systematic Evaluation of One-Dimensional Unsteady Friction Models <strong>in</strong><br />

Simple Pipel<strong>in</strong>es, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 132(7): 696–708.<br />

Walski, T. M., Chase, D. V., Savic, D. A., Grayman, W., Beckwith, S. <strong>and</strong><br />

Koelle, E. (2003). Advanced Water Distribution Model<strong>in</strong>g <strong>and</strong> Management,<br />

Haestad Methods Inc.<br />

Wang, S. <strong>and</strong> Carroll, J. J. (2007). Leak Detection for Gas <strong>and</strong> Liquid Pipel<strong>in</strong>es<br />

by Onl<strong>in</strong>e Model<strong>in</strong>g, SPE Projects, Facilities & Construction 2(2): 1–9.<br />

Wang, X.-J., Lambert, M. F., Simpson, A. R., Liggett, J. A. <strong>and</strong> Vitkovsky,<br />

J. P. (2002). Leak Detection <strong>in</strong> Pipel<strong>in</strong>es Us<strong>in</strong>g the Damp<strong>in</strong>g of Fluid Transients,<br />

Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g 128(7): 697–711.<br />

Wang, Z., Nur, A. M. <strong>and</strong> Batzle, M. L. (1990). Acoustic Velocities <strong>in</strong><br />

Petroleum Oils, Journal of Petroleum Technology 42(2): 192–200.<br />

We<strong>in</strong>erowska-Bords, K. (2007). Accuracy <strong>and</strong> Parameter Estimation of Elastic<br />

<strong>and</strong> Viscoelastic Models of the Water Hammer, Task Quarterly: : Scientific<br />

Bullet<strong>in</strong> of Academic Computer Centre <strong>in</strong> Gdansk 11(4): 383–395.<br />

Wood, D. J. <strong>and</strong> Jones, S. E. (1973). Water-Hammer Charts for Various<br />

Types of Valves, Journal of the Hydraulics Division 99(1): 167–178.<br />

Wood, D. J. <strong>and</strong> Jones, S. E. (1974). Water-Hammer Charts for Various<br />

Types of Valves, Journal of the Hydraulics Division 100(9): 1272–1274.<br />

Wylie, E. B. <strong>and</strong> Streeter, V. L. (1993). Fluid Transients <strong>in</strong> Systems, Prentice<br />

Hall.<br />

199


Yamanaka, G., Kikura, H., Takeda, Y. <strong>and</strong> Aritomi, M. (2002). Flow Measurement<br />

on an Oscillat<strong>in</strong>g Pipe Flow Near the Entrance Us<strong>in</strong>g the UVP<br />

Method, Experiment <strong>in</strong> Fluids 32(2): 212–220.<br />

Zielke, W. (1966). Frequency Dependent Friction <strong>in</strong> Transient Pipe Flow, PhD<br />

thesis, Department of Civil Eng<strong>in</strong>eer<strong>in</strong>g - University of Michigan.<br />

Zielke, W. (1968). Frequency-Dependent Friction <strong>in</strong> Transient Pipe Flow,<br />

Journal of Basic Eng<strong>in</strong>eer<strong>in</strong>g 90(1): 109–115.<br />

200


Appendix A<br />

Technical Reports<br />

Seven technical reports were written as part of the thesis work. A list of titles<br />

<strong>and</strong> tables of content is given below:<br />

• Propagation of <strong>Pressure</strong> <strong>Waves</strong> <strong>in</strong> Liquid-Filled Pipes (Confidential),<br />

December 2007<br />

1. Introduction<br />

2. Experimental Apparatus<br />

3. Experimental Data<br />

4. Analysis of Incl<strong>in</strong>ed Set-Up Data<br />

5. Analysis of Horizontal Set-Up Data<br />

6. <strong>Impulse</strong> Propagation<br />

7. St<strong>and</strong><strong>in</strong>g Wave<br />

8. Flow Rate<br />

9. Discussion<br />

• Simulation of Valve Conditions Us<strong>in</strong>g CLAWPACK (Confidential),<br />

January 2008<br />

1. Introduction<br />

2. One Dimensional Wave Equation<br />

3. Numerical Schemes<br />

4. St<strong>and</strong>ard Test-Cases<br />

5. Numerical Scheme for Valve Model<br />

6. Conclusions<br />

• Numerical Schemes, Compressibility <strong>and</strong> Pulsat<strong>in</strong>g Flow (Confidential),<br />

August 2008<br />

201


Appendix A<br />

1. Introduction<br />

2. Further Work on Numerical Schemes<br />

3. Modell<strong>in</strong>g of Fluid Exit Po<strong>in</strong>t<br />

4. Velocity Profile from Piston Movement<br />

5. Conclusions<br />

• Water Hammer Simulation of <strong>Pressure</strong> <strong>in</strong> the Failed Flexible Hose for<br />

Off-Load<strong>in</strong>g Oil at Statfjord A, October 2008<br />

1. Introduction<br />

2. Wave Transmission at an Interface<br />

3. Simulation Cases<br />

4. Conclusions<br />

• Parameters Affect<strong>in</strong>g the Speed of Sound <strong>in</strong> Fluids (Confidential),<br />

December 2008<br />

1. Introduction<br />

2. Speed of Sound Def<strong>in</strong>ition<br />

3. Experimental Work<br />

4. Ultrasonic Meter Data Analysis<br />

5. Extended Speed of Sound Def<strong>in</strong>ition<br />

6. Conclusions<br />

• Horizontal <strong>Impulse</strong> Pump Performance (Confidential), January 2009<br />

1. Introduction<br />

2. <strong>Impulse</strong> Pump<strong>in</strong>g<br />

3. <strong>Impulse</strong> Pump Performance<br />

4. Conclusions<br />

• Vertical <strong>Impulse</strong> Pump Modell<strong>in</strong>g (Confidential), February 2009<br />

1. Introduction<br />

2. <strong>Impulse</strong> Pump<strong>in</strong>g<br />

3. Non-Return Valves<br />

4. Vertical <strong>Impulse</strong> Pump<strong>in</strong>g Pr<strong>in</strong>ciples<br />

5. <strong>Pressure</strong> <strong>and</strong> <strong>Pressure</strong> Gradients<br />

6. Vertical <strong>Impulse</strong> Pump<strong>in</strong>g Cases<br />

202


7. Examples<br />

8. Conclusions<br />

Appendix A<br />

• <strong>Impulse</strong> Pump<strong>in</strong>g Design, Model <strong>and</strong> Simulation (Confidential), December<br />

2009<br />

1. Introduction<br />

2. <strong>Impulse</strong> Generator<br />

3. Outlet Non-Return Valve<br />

4. Inlet Non-Return Valve<br />

5. Characteristic Curves<br />

6. Test Case<br />

7. Conclusions<br />

203


204


Appendix B<br />

Pump<strong>in</strong>g of Fluids Us<strong>in</strong>g<br />

<strong>Pressure</strong> <strong>Impulse</strong>s<br />

Presented at SPE EUROPEC, Amsterdam , June 2009. Paper SPE 120896<br />

reformatted for the present thesis.<br />

Abstract<br />

<strong>Impulse</strong> pump<strong>in</strong>g is a new method to pump fluids. The method is described<br />

briefly <strong>and</strong> then numerical simulations were carried out to illustrate the nature<br />

of impulse pump<strong>in</strong>g. The simulations are based on the l<strong>in</strong>ear ideal wave equation<br />

<strong>and</strong> the numerical approach is based on a second-order Godunov scheme.<br />

The work illustrates how s<strong>in</strong>usoidal pressure waves are reflected at open <strong>and</strong><br />

closed boundaries, represent<strong>in</strong>g open <strong>and</strong> closed non-return valves <strong>in</strong> impulse<br />

pump<strong>in</strong>g. An impulse pump<strong>in</strong>g example was presented for 40 m lift<strong>in</strong>g of water.<br />

It was shown how the wellhead pressure used determ<strong>in</strong>es the lift<strong>in</strong>g depth<br />

<strong>and</strong> flow rate achievable; the greater the wellhead pressure, the greater the lift<strong>in</strong>g<br />

<strong>and</strong> rate. In impulse pump<strong>in</strong>g the wellhead pressure is controlled by the<br />

crack<strong>in</strong>g pressure of a non-return outlet valve. Fluid enters the production<br />

tub<strong>in</strong>g through a non-return <strong>in</strong>let valve downhole. The stroke length of a pressure<br />

wave generator determ<strong>in</strong>es the wellhead pressures achievable <strong>in</strong> impulse<br />

pump<strong>in</strong>g. It was suggested that the water hammer equation can be used to<br />

estimate the flow rate <strong>in</strong> impulse pump<strong>in</strong>g.<br />

Introduction<br />

The pump<strong>in</strong>g of fluids is an important unit operation worldwide, <strong>in</strong>clud<strong>in</strong>g<br />

the petroleum <strong>in</strong>dustry. Volumetric <strong>and</strong> dynamic pumps f<strong>in</strong>d different appli-<br />

205


Appendix B<br />

cations. They have <strong>in</strong> common to be located where fluids are pumped from.<br />

In addition to location, the pump<strong>in</strong>g of difficult fluids is a challenge. The<br />

efficiency <strong>and</strong> reliability of pump<strong>in</strong>g systems is of considerable importance<br />

(Hydraulic Institute 2008).<br />

A novel pump<strong>in</strong>g concept has been proposed by Sagov (2004), here called<br />

impulse pump<strong>in</strong>g. The actuat<strong>in</strong>g unit of an impulse pump<strong>in</strong>g system does not<br />

need to be located where the fluids are pumped from. Instead, an <strong>in</strong>let nonreturn<br />

valve is placed at the distant-end of a pipel<strong>in</strong>e or production tub<strong>in</strong>g. In<br />

the case of oil <strong>and</strong> water wells, the <strong>in</strong>let non-return valve is located downhole,<br />

while the actuat<strong>in</strong>g unit <strong>and</strong> associated power source (motor) are place at the<br />

wellhead.<br />

Accord<strong>in</strong>g to the description of impulse pump<strong>in</strong>g by Sagov (2004), a s<strong>in</strong>usoidal<br />

pressure wave is generated at the wellhead. In an example provided, the<br />

frequency range was from 0.5 to 3.5 Hz <strong>and</strong> a typical stroke length of the<br />

actuat<strong>in</strong>g unit was given as 5 mm. The pressure values reported were <strong>in</strong><br />

the range 0 to 10 bara. These are typical values presented by Sagov (2004)<br />

to illustrate one possible embodiment of an impulse pump<strong>in</strong>g system. Other<br />

wave forms are possible, other frequencies are possible <strong>and</strong> other stroke lengths<br />

are possible.<br />

While further experimental data are not available, it is of <strong>in</strong>terest to consider<br />

the theoretical basis of impulse pump<strong>in</strong>g us<strong>in</strong>g established pressure wave<br />

physics <strong>and</strong> numerical simulations. In this paper the impulse pump<strong>in</strong>g method<br />

will be described based on the <strong>in</strong>formation provided by Sagov (2004). Furthermore,<br />

the propagation (transmission <strong>and</strong> reflection) of pressure waves <strong>in</strong><br />

an impulse pump<strong>in</strong>g system will be modelled. Simple models to estimate the<br />

pressure wave amplitude <strong>and</strong> flow rate are also suggested.<br />

<strong>Pressure</strong> Wave Propagation<br />

A simplified representation of an impulse pump is shown <strong>in</strong> Figure 1 based<br />

on the orig<strong>in</strong>al draw<strong>in</strong>g provided by Sagov (2004). At the left-h<strong>and</strong>-side an<br />

impulse generator (also called an oscillator) with a back-<strong>and</strong>-forth mov<strong>in</strong>g element<br />

generates a cont<strong>in</strong>uous s<strong>in</strong>usoidal pressure wave, which is superimposed<br />

on the <strong>in</strong>-situ static pressure. Nearby a non-return outlet valve is located.<br />

The crack<strong>in</strong>g (open<strong>in</strong>g) pressure of the valve will be high <strong>and</strong> can be adjusted.<br />

In fact, this valve controls the wellhead pressure. At the distant end (righth<strong>and</strong>-side<br />

of draw<strong>in</strong>g) of the pipe a non-return <strong>in</strong>let valve is located. This<br />

valve needs to have a low <strong>and</strong> fixed crack<strong>in</strong>g pressure. Crack<strong>in</strong>g pressure is<br />

the pressure difference required to open a non-return valve.<br />

The s<strong>in</strong>usoidal pressure wave propagates at the speed of sound from the im-<br />

206


Appendix B<br />

pulse generator to the outlet valve. In this paper the wave pressure higher<br />

(above) than the <strong>in</strong>-situ static pressure will be referred to as a positive pressure<br />

<strong>and</strong> the pressure lower (below) than the <strong>in</strong>-situ static pressure will be<br />

referred to as a negative pressure. When a positive pressure wave arrives at<br />

the non-return outlet valve, the valve will open <strong>and</strong> close at the set crack<strong>in</strong>g<br />

pressure. The pressure wave moves further <strong>in</strong> the direction of the distant end.<br />

When a negative pressure wave arrives at the non-return <strong>in</strong>let valve at the<br />

distant end, the valve will open <strong>and</strong> fluid enters the impulse pump<strong>in</strong>g system.<br />

The nature of pressure waves <strong>in</strong> fluids has been described by Pa<strong>in</strong> (2005). The<br />

propagation (transmission <strong>and</strong> reflection) of pressure waves depends on the<br />

acoustic impedance:<br />

z = ρa (B.1)<br />

where, ρ, is fluid density <strong>and</strong>, a, the acoustic velocity. The reflection, R, <strong>and</strong><br />

transmission, T, coefficients for a pressure wave at an impedance <strong>in</strong>terface are,<br />

respectively:<br />

R = pR<br />

pI<br />

T = pT<br />

pI<br />

= z2 − z1<br />

z2 + z1<br />

= 2z2<br />

z2 + z1<br />

(B.2)<br />

(B.3)<br />

For the nomenclature used a pressure wave arrives from medium 1 <strong>and</strong> is<br />

reflected back <strong>in</strong>to medium 1 <strong>and</strong> transmitted <strong>in</strong>to medium 2. The subscript,<br />

I, st<strong>and</strong>s for <strong>in</strong>cident.<br />

Representative values of the reflection coefficient, R, are illustrated <strong>in</strong> Figure<br />

2. The coefficient is plotted aga<strong>in</strong>st the speed of sound ratio a2/a1 from 0 to 1.<br />

When the ratio is 0 the coefficient R = −1 <strong>and</strong> the boundary condition can be<br />

described as open, <strong>and</strong> a pressure wave has total reflection <strong>and</strong> phase change.<br />

When the speed of sound ratio is 1 the coefficient R = 0 <strong>and</strong> the boundary<br />

condition can be described as non-exist<strong>in</strong>g; that is, the same situation as an<br />

open pipe. Not shown <strong>in</strong> Figure 2 is when the reflection coefficient R = 1<br />

where the speed of sound ratio tends to <strong>in</strong>f<strong>in</strong>ity <strong>and</strong> the boundary condition<br />

can be described as closed.<br />

<strong>Pressure</strong> Wave Simulation<br />

The l<strong>in</strong>ear ideal wave equation was used to simulate the propagation of pressure<br />

waves <strong>in</strong> a fluid <strong>in</strong> a pipe. The simulation method was based on a secondorder<br />

Godunov numerical scheme implemented <strong>in</strong> CLAWPACK (LeVque 2002).<br />

207


Appendix B<br />

The typical time step was 1 s <strong>and</strong> the typical space step was 0.20 m such that<br />

the Courant number was <strong>in</strong> the range 0.95 <strong>and</strong> 1. In the simulations the fluid<br />

density was set to 1000 kg m −3 <strong>and</strong> the speed of sound was set to 1400 ms −1 .<br />

Attenuation mechanisms are not <strong>in</strong>cluded <strong>in</strong> the l<strong>in</strong>ear ideal wave equation.<br />

Such a simplification has no practical significance for the present work; the<br />

focus of the work has been on illustrat<strong>in</strong>g basic pr<strong>in</strong>ciples.<br />

Simulations were carried out to illustrate the nature of reflections from closed<br />

<strong>and</strong> open boundaries <strong>in</strong> a horizontal pipe. A closed boundary represents a<br />

solid boundary such as a closed valve. An open boundary correspond<strong>in</strong>gly<br />

represents an open valve. The grid chosen was 1 m long <strong>and</strong> a s<strong>in</strong>usoidal wave<br />

10 bar positive <strong>and</strong> 10 bar negative from the pressure wave generator was<br />

used. In the modell<strong>in</strong>g the positive pressure wave was truncated by 2.5 bar to<br />

illustrate the effect of an outlet non-return valve. The simulation results are<br />

shown <strong>in</strong> Figure 3.<br />

Figure 3a illustrates a pressure wave travell<strong>in</strong>g from left-to-right after hav<strong>in</strong>g<br />

passed an outlet non-return valve. Two cases of reflections are illustrated<br />

<strong>in</strong> (b) <strong>and</strong> (c). Figure 3b illustrates the shape of the same pressure wave<br />

after hav<strong>in</strong>g reflected from a closed boundary, mov<strong>in</strong>g from right-to-left <strong>in</strong> the<br />

figure. It shows that the pressure wave reflects without phase change. Figure<br />

3c illustrates the shape of the same pressure wave after reflect<strong>in</strong>g from an<br />

open boundary, aga<strong>in</strong> mov<strong>in</strong>g from right-to-left <strong>in</strong> the figure. It shows that<br />

the pressure wave reflects with phase change.<br />

Simulations were then carried out to illustrate the nature of reflections at open<br />

<strong>and</strong> closed boundaries <strong>in</strong> a vertical pipe. The grid chosen was 100 m long<br />

(deep) <strong>and</strong> a static pressure of 10 bara at the pressure generator. In other<br />

words, 10 bara wellhead pressure. A s<strong>in</strong>gle superimposed s<strong>in</strong>usoidal pressure<br />

wave 10 bar positive <strong>and</strong> 10 bar negative was chosen. Because the simulations<br />

were carried out for illustration purposes, a high pressure wave frequency was<br />

used for convenience. The simulation results are shown <strong>in</strong> Figure 4 at two<br />

times/locations for a propagat<strong>in</strong>g pressure wave.<br />

Figure 4a illustrates a s<strong>in</strong>gle truncated pressure wave travell<strong>in</strong>g downward.<br />

It is superimposed on the sum of the wellhead pressure <strong>and</strong> the hydrostatic<br />

pressure. Figure 4b illustrates the same pressure wave after reflection from<br />

an open non-return <strong>in</strong>let valve at 100 m depth (downhole). In the numerical<br />

simulations the non-return <strong>in</strong>let valve can be modelled as open<strong>in</strong>g immediately<br />

when the pressure wave starts to become negative. It can also be modelled<br />

as open<strong>in</strong>g at different percentages of the negative pressure wave. In the case<br />

illustrated <strong>in</strong> Figure 4b the valve was modelled to open when the negative<br />

pressure wave reached 25 percent of its maximum value of 10 bara. The valve<br />

was modelled to close on the reverse cycle of the negative pressure wave, aga<strong>in</strong><br />

208


at the 25 percent value.<br />

Appendix B<br />

The above simulations say noth<strong>in</strong>g about the pump<strong>in</strong>g action of an impulse<br />

pump. The simulations just illustrate how a pressure wave reflects from a<br />

non-return <strong>in</strong>let valve when closed <strong>and</strong> when open, irrespective of how <strong>and</strong><br />

why it opens.<br />

Downhole Pump<strong>in</strong>g Action<br />

In the text above the nature of pressure wave reflections at open <strong>and</strong> closed<br />

boundaries were illustrated; an open boundary represents an open valve <strong>and</strong><br />

a closed boundary represents a closed valve. What rema<strong>in</strong>s to be illustrated is<br />

how the downhole non-return <strong>in</strong>let valve allows pump<strong>in</strong>g action to take place.<br />

Fluid will flow from the wellbore <strong>in</strong>to the production tub<strong>in</strong>g when the outside<br />

pressure is higher than the <strong>in</strong>side pressure. A non-return <strong>in</strong>let valve should<br />

preferably have a low crack<strong>in</strong>g pressure.<br />

The static downhole pressure <strong>in</strong> the tub<strong>in</strong>g will be the sum of the wellhead<br />

pressure <strong>and</strong> the hydrostatic pressure. The pressure wave generated at the<br />

wellhead will be superimposed on the static pressure. To create a low pressure<br />

<strong>in</strong>side the tub<strong>in</strong>g downhole, the wellhead pressure needs to be equally large.<br />

At the wellhead, the pressure wave generator generates a pressure wave with a<br />

positive part <strong>and</strong> a negative part. The negative part of the pressure wave can<br />

never lower the static pressure below zero absolute pressure. It follows that<br />

the negative part of a pressure wave can never be larger than the wellhead<br />

pressure.<br />

Assume a water well 100 m deep with a wellhead pressure of 10 bara. The<br />

downhole pressure (at the <strong>in</strong>side tub<strong>in</strong>g <strong>in</strong>let) will be 20 bara. Assume a<br />

pressure wave is generated at the wellhead with amplitude of 20 bar such<br />

that the negative part of the wave is 10 bar. When this negative part of<br />

the pressure wave arrives at a non-return <strong>in</strong>let valve downhole, the valve is<br />

closed <strong>and</strong> the pressure will reflect without phase change. In other words, the<br />

negative pressure will reflect as negative pressure, hence further decreas<strong>in</strong>g the<br />

pressure.<br />

As the reflected negative pressure reflects onto itself to decrease the pressure<br />

further, there comes a po<strong>in</strong>t where the <strong>in</strong>side tub<strong>in</strong>g pressure is less than the<br />

outside wellbore pressure. When that po<strong>in</strong>t is reached, the non-return valve<br />

will open <strong>and</strong> fluid will flow <strong>in</strong>to the production tub<strong>in</strong>g. While the non-return<br />

valve is open, the negative pressure wave will no longer reflect as negative,<br />

but it will reflect positively because of phase change. Thereafter there comes<br />

a po<strong>in</strong>t when the low pressure <strong>in</strong>side the tub<strong>in</strong>g is not sufficient to keep the<br />

non-return valve open. The flow stops <strong>and</strong> the non-return valve rema<strong>in</strong>s closed<br />

209


Appendix B<br />

until the next negative pressure wave arrives downhole.<br />

For the assumed 100 m deep water well above, the pressure wave has the<br />

potential to decrease the downhole pressure by maximum 20 bar; that is,<br />

twice the negative part of the pressure wave. Assum<strong>in</strong>g that the wellbore<br />

pressure outside the production tub<strong>in</strong>g is 6 bara, a non-return <strong>in</strong>let valve with<br />

1 bar crack<strong>in</strong>g pressure will open when 75% of the negative pressure wave<br />

has reflected. The pressure difference between the outside wellbore <strong>and</strong> the<br />

<strong>in</strong>side production tub<strong>in</strong>g will be 1 bar (pressure <strong>in</strong>side production tub<strong>in</strong>g will<br />

be 5 bara). It is this pressure difference that makes fluid flow from outside to<br />

<strong>in</strong>side. The valve stays open until the negative pressure wave sw<strong>in</strong>gs back to<br />

three-quarters (75%) of its value.<br />

The pressure with depth <strong>in</strong> the assumed 100 m deep water well is shown <strong>in</strong><br />

Figure 5. Superimposed on the pressure with depth is the pressure wave reflected<br />

(travell<strong>in</strong>g from downhole to wellhead) from the non-return <strong>in</strong>let valve.<br />

The vertical lift <strong>in</strong> the assumed well is 40 m. For the assumed wellhead pressure<br />

<strong>and</strong> non-return <strong>in</strong>let valve crack<strong>in</strong>g pressure, the theoretical maximum<br />

vertical lift is 90 m for the case presented. However, for such vertical lift the<br />

pump<strong>in</strong>g rate will be zero. For vertical lift less than 90 m but greater than 40<br />

m, the pump<strong>in</strong>g rate will be less than for 40 m. This because the non-return<br />

<strong>in</strong>let valve will be open for a shorter time (dur<strong>in</strong>g the full cycle of the negative<br />

pressure wave) the greater the vertical lift.<br />

Pump Performance<br />

It is of <strong>in</strong>terest to estimate the performance of impulse pump<strong>in</strong>g <strong>in</strong> terms<br />

of pressure <strong>and</strong> flow rate. First, a relationship between the stroke length of<br />

the actuator <strong>and</strong> the pressure wave amplitude will be established. Second, a<br />

relationship between the crack<strong>in</strong>g pressure of the non-return <strong>in</strong>let valve <strong>and</strong><br />

the flow rate through the valve will be suggested.<br />

The isothermal compressibility, K, of fluids is given by the equation<br />

K = −1 dV<br />

V dp<br />

(B.4)<br />

where V is volume <strong>and</strong> p is pressure. In impulse pump<strong>in</strong>g the system volume<br />

V is the fluid volume <strong>in</strong> the pipe between the generator <strong>and</strong> the non-return<br />

<strong>in</strong>let valve, depend<strong>in</strong>g on the pipe length L <strong>and</strong> diameter d. And the change <strong>in</strong><br />

fluid volume dV is that result<strong>in</strong>g from the compression exerted on the system<br />

by the pressure wave generator, depend<strong>in</strong>g on the stroke length x <strong>and</strong> effective<br />

diameter. Assum<strong>in</strong>g that the actuator <strong>and</strong> pipe have the same diameter, it<br />

210


Appendix B<br />

can be shown (us<strong>in</strong>g the compressibility equation) that the amplitude of a<br />

pressure wave generated can be expressed by the relationship:<br />

∆p<strong>Impulse</strong> = x 1<br />

L K<br />

(B.5)<br />

The isothermal compressibility of water at st<strong>and</strong>ard conditions is reported as<br />

4.65 10 −10 Pa −1 (Yaws 1999). For a pipe length of 100 m <strong>and</strong> stroke length<br />

of 5 mm, the pressure amplitude was calculated 10.75 bar. We observe that<br />

the maximum pressure reported by Sagov (2004) was about 10 bara.<br />

In the present paper, the water hammer equation is assumed to control the<br />

fluid velocity through the non-return <strong>in</strong>let valve. The equation, also called the<br />

Joukowski equation, can be expressed as:<br />

∆pHammer = ρa∆u (B.6)<br />

where ∆u represents the change <strong>in</strong> fluid velocity from zero to the velocity<br />

through the non-return <strong>in</strong>let valve. If the water hammer pressure drop is<br />

assumed equal to the crack<strong>in</strong>g pressure of the non-return <strong>in</strong>let valve, the effective<br />

fluid velocity can be calculated. Us<strong>in</strong>g speed of sound 1400 m/s <strong>and</strong><br />

water density of 1000 kg/m3 <strong>and</strong> a crack<strong>in</strong>g pressure of 1 bar, the flow velocity<br />

becomes 0.0714 ms −1 .<br />

Fluid will flow <strong>in</strong>to the production tub<strong>in</strong>g only while the non-return valve is<br />

open. The valve will open periodically. This period can be expressed as a<br />

fraction of the negative pressure wave. In the example above the non-return<br />

valve opened when the pressure reach 75% of the negative pressure wave.<br />

Approximat<strong>in</strong>g a s<strong>in</strong>us wave by a triangle the non-return valve will be opened<br />

25% of the full wave cycle. Therefore, the volumetric flow rate, q, through the<br />

non-return valve can be expressed by the relationship:<br />

q = αuA (B.7)<br />

where, α is the cycle fraction <strong>and</strong> A the flow area. The flow rate depends on<br />

the diameter of the production tub<strong>in</strong>g. For example, for 1 <strong>in</strong>ch ID the flow<br />

rate will be 0.55 litres per m<strong>in</strong>ute. Similarly, for 2 <strong>in</strong>ch <strong>and</strong> 4 <strong>in</strong>ch ID the<br />

volumetric flow rate will be 2.2 litres per m<strong>in</strong>ute <strong>and</strong> 4.4 litres per m<strong>in</strong>ute,<br />

respectively. The flow rates are ideal because pressure losses due to valve<br />

mechanics, fluid acceleration/deceleration <strong>and</strong> wall friction <strong>in</strong> the production<br />

tub<strong>in</strong>g have not been <strong>in</strong>cluded.<br />

211


Appendix B<br />

Conclusions<br />

Numerical simulations were used to illustrate the nature of pressure wave<br />

propagation <strong>in</strong> impulse pump<strong>in</strong>g systems. In particular how pressure waves<br />

reflect at open <strong>and</strong> closed boundaries.<br />

Analytical considerations presented illustrate how the wellhead pressure determ<strong>in</strong>es<br />

the vertical lift capability of liquid filled wells. The higher the wellhead<br />

pressure, the greater the vertical lift <strong>and</strong> flow rate achievable.<br />

One particular example was presented where liquid water was lifted 40 m. The<br />

wellhead pressure assumed was 10 bara, while no lift<strong>in</strong>g would be achieved with<br />

a wellhead pressure of 6 bara.<br />

The stroke length of the pressure wave generator determ<strong>in</strong>es the amplitude of<br />

the pressure wave superimposed on the static pressure.<br />

The flow rate <strong>in</strong> impulse pump<strong>in</strong>g can be related to the water hammer equation.<br />

For the example presented, assum<strong>in</strong>g 4 <strong>in</strong>ch production tub<strong>in</strong>g, the flow<br />

rate was 4.4 litres per m<strong>in</strong>ute.<br />

Acknowledgement<br />

We thank Clavis <strong>Impulse</strong> Technology AS for provid<strong>in</strong>g a Ph.D. scholarship to<br />

Benjam<strong>in</strong> Pierre. However, the ideas <strong>and</strong> results presented <strong>in</strong> the paper are<br />

solely the responsibility of the authors.<br />

References<br />

LeVeque R.J. (2002): F<strong>in</strong>ite Volume Methods for Hyperbolic Problems, Cambrigde<br />

University Press.<br />

Pa<strong>in</strong> H.J. (2005): The Physics of Vibrations <strong>and</strong> <strong>Waves</strong>, Wiley.<br />

Sagov M. (2004): Method <strong>and</strong> Apparatus for Transport<strong>in</strong>g Fluid <strong>in</strong> a Conduit,<br />

Norwegian Patent No. 20045382 (<strong>in</strong> Norwegian).<br />

Yaws C.L. (1999): Chemical Properties H<strong>and</strong>book, Mac-Graw Hill.<br />

212


Nomenclature<br />

a Acoustic Velocity [ms−1 ]<br />

A Flow Area [m2 ]<br />

K Fluid Compressibility [Pa−1 ]<br />

L Pipe Length [m]<br />

p <strong>Pressure</strong> [Pa]<br />

q Volumetric Flow Rate [l m<strong>in</strong>−1 ]<br />

u Fluid Velocity [ms−1 ]<br />

V Fluid Volume [m3 ]<br />

x Stroke Length [m]<br />

z Acoustic Impedance [kg s−1 m−2 ]<br />

z1 Acoustic Impedance <strong>in</strong> Medium 1 [kg s−1 m−2 ]<br />

z2 Acoustic Impedance <strong>in</strong> Medium 2 [kg s−1 m−2 ]<br />

∆pHammer Water Hammer Amplitude [Pa]<br />

∆p<strong>Impulse</strong> Half of the <strong>Pressure</strong> Wave Amplitude [Pa]<br />

α Cycle Fraction<br />

ρ Fluid Density [kg m−3 ]<br />

Figures<br />

1<br />

h<br />

3<br />

2<br />

L<br />

4<br />

Appendix B<br />

Figure B.1: <strong>Impulse</strong> Pump<strong>in</strong>g Apparatus, Horizontal Configuration<br />

1: <strong>Impulse</strong> Generator<br />

2: Outlet Non-Return Valve<br />

3: Outlet Tank<br />

4: Inlet Non-Return Valve<br />

5: Inlet Tank<br />

213<br />

5


Appendix B<br />

Reflection Coefficient<br />

0<br />

−0.1<br />

−0.2<br />

−0.3<br />

−0.4<br />

−0.5<br />

−0.6<br />

−0.7<br />

−0.8<br />

−0.9<br />

−1<br />

0 0.1 0.2 0.3 0.4 0.5<br />

z / z<br />

2 1<br />

0.6 0.7 0.8 0.9 1<br />

Figure B.2: Representative Values of the Reflection Coefficient<br />

10<br />

0<br />

−10<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Distance [m]<br />

10<br />

p [bar] a)<br />

0<br />

−10<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Distance [m]<br />

10<br />

p [bar] b)<br />

0<br />

−10<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Distance [m]<br />

p [bar] c)<br />

Figure B.3: <strong>Pressure</strong> Wave (a) After Pass<strong>in</strong>g a Non-return Outlet<br />

Valve, (b) After Reflect<strong>in</strong>g From a Closed Boundary, (c)<br />

After Reflect<strong>in</strong>g on an Open Boundary<br />

214


Depth [m]<br />

a)<br />

0<br />

10<br />

20<br />

30<br />

40<br />

50<br />

60<br />

70<br />

80<br />

90<br />

100<br />

0 10 20 30<br />

<strong>Pressure</strong> [bara]<br />

b)<br />

0<br />

10<br />

20<br />

30<br />

40<br />

50<br />

60<br />

70<br />

80<br />

90<br />

100<br />

0 10 20 30<br />

<strong>Pressure</strong> [bara]<br />

Appendix B<br />

Figure B.4: Truncated <strong>Pressure</strong> Wave Travell<strong>in</strong>g (a) Downward, (b)<br />

After Reflect<strong>in</strong>g From a Non-Return Valve<br />

Depth [m]<br />

0<br />

10<br />

20<br />

30<br />

40<br />

50<br />

60<br />

70<br />

80<br />

90<br />

100<br />

0 5 10 15 20 25 30<br />

<strong>Pressure</strong> [bara]<br />

Figure B.5: Truncated <strong>Pressure</strong> Wave Travell<strong>in</strong>g After Reflect<strong>in</strong>g<br />

From a Non-Return Valve<br />

215


216


Appendix C<br />

Modell<strong>in</strong>g of <strong>Pressure</strong> Wave<br />

Propagation <strong>in</strong> Pipel<strong>in</strong>es:<br />

Steady-State <strong>and</strong> Transient<br />

Friction<br />

Presented at MekIT’09, Fifth National Conference on Computational Mechanics,<br />

Trondheim, 26-27 May 2009. Paper reformatted for the present thesis.<br />

Abstract<br />

The physics of the water hammer effect are used to model rapid pressure transients<br />

<strong>in</strong> wells <strong>and</strong> pipel<strong>in</strong>es <strong>and</strong> more generally <strong>in</strong> fluid transport systems.<br />

Traditionally, the Method of Characteristics (MOC) is used to solve numerically<br />

the govern<strong>in</strong>g partial differential equations. In addition to the water<br />

hammer effect, effects due to l<strong>in</strong>e pack<strong>in</strong>g, pressure attenuation <strong>and</strong> transient<br />

friction need to be <strong>in</strong>cluded <strong>in</strong> comprehensive modell<strong>in</strong>g. Recent works have<br />

<strong>in</strong>dicated that the errors <strong>in</strong>herent <strong>in</strong> the use of the MOC tend to mimic the<br />

errors <strong>in</strong> the formulation of the transient friction factor. Numerical simulations<br />

have been carried out us<strong>in</strong>g several transient friction factor formulations<br />

<strong>in</strong> the hyperbolic system of equations describ<strong>in</strong>g rapid pressure transients. The<br />

results argue for the necessity to develop new formulations that accurately represent<br />

the physics of rapid pressure transient propagation <strong>in</strong> fluid transport<br />

systems. The MOC needs also to be replaced by higher resolutions numerical<br />

methods such as a second order f<strong>in</strong>ite volume method.<br />

217


Appendix C<br />

Introduction<br />

Fluid transport systems, for <strong>in</strong>stance transport <strong>and</strong> distribution systems of<br />

oil <strong>and</strong> gas, are subject to rapid pressure transients. Rapid pressure transients<br />

arise when a quick-act<strong>in</strong>g valve is actuated, or when a pump is suddenly<br />

stopped. Rapid pressure transient <strong>in</strong> the form of water hammer can lead<br />

to pipe rupture (Pierre <strong>and</strong> Gudmunsson, 2008). Therefore, rapid pressure<br />

transients need to be accurately predicted.<br />

Rapid pressure transients <strong>in</strong> the form of pressure wave can be used <strong>in</strong> fluid<br />

transport systems. In impulse pump<strong>in</strong>g (Pierre <strong>and</strong> Gudmundsson, 2009),<br />

pressure waves are used to pump liquids from bottom-hole to wellhead. In<br />

pressure pulse technology (Gudmundsson <strong>and</strong> Celius, 1999), pressure wave<br />

propagation properties are used to f<strong>in</strong>d the mass flow rate, gas-oil ratio <strong>and</strong><br />

deposits on pipe wall. Predict<strong>in</strong>g the attenuation of a pressure wave due to<br />

wall friction is therefore important to evaluate how far a pressure wave will<br />

travel <strong>and</strong> therefore, assess the technologies limitations.<br />

Numerical methods have long been used to predict rapid pressure transients.<br />

The method of characteristics (MOC) presents an eng<strong>in</strong>eer-friendly method<br />

for solv<strong>in</strong>g a simplified one dimensional unsteady flow of compressible fluid<br />

system of equations. Recent model improvements for the govern<strong>in</strong>g system<br />

of equations were presented Brunone et al (1991) <strong>and</strong> Ramos et al. (2004),<br />

where the source term <strong>in</strong>cludes an unsteady friction factor <strong>in</strong> addition to the<br />

steady-state Darcy-Weisbach friction factor. The proposed models of transient<br />

friction factors were numerically studied by the method of characteristics.<br />

However, a recent study by Szymkiewicz <strong>and</strong> Mitosek (2007) showed that<br />

the improvements generated by the addition of transient friction factors were<br />

negligible. Szymkiewicz <strong>and</strong> Mitosek showed also that the agreement between<br />

experimental results <strong>and</strong> numerical simulations were not reproducible if not<br />

us<strong>in</strong>g the method of characteristics for solv<strong>in</strong>g the govern<strong>in</strong>g equations. In<br />

particular, the damp<strong>in</strong>g <strong>and</strong> smooth<strong>in</strong>g of a pressure wave <strong>in</strong> fluid transport<br />

pipel<strong>in</strong>e are <strong>in</strong>accurately predicted.<br />

In this paper, a second order f<strong>in</strong>ite volume method will be used for solv<strong>in</strong>g<br />

the commonly used govern<strong>in</strong>g equations. The f<strong>in</strong>ite volume method will be<br />

compared with the st<strong>and</strong>ard method of characteristics <strong>in</strong> order to justify its<br />

use. Two transient friction factor models will then be <strong>in</strong>cluded <strong>in</strong> the selected<br />

f<strong>in</strong>ite volume numerical scheme <strong>and</strong> the results will be compared with the<br />

steady friction modell<strong>in</strong>g. The adequacy of the two models will be evaluated<br />

<strong>and</strong> the results will be compared to Szymkiewicz <strong>and</strong> Mitosek observations.<br />

218


<strong>Pressure</strong> Wave <strong>in</strong> Pipel<strong>in</strong>es<br />

Appendix C<br />

Oil <strong>and</strong> gas pipel<strong>in</strong>es are used to transport hydrocarbons over long distances.<br />

Transport pipel<strong>in</strong>es are generally a few <strong>in</strong>ches to tens of <strong>in</strong>ches <strong>in</strong> diameter,<br />

<strong>and</strong> cover hundreds of kilometers. For <strong>in</strong>stance, the Europipe gas transport<br />

pipel<strong>in</strong>e between Draupner <strong>and</strong> Dornum measures 40 <strong>in</strong>ches <strong>in</strong> diameter <strong>and</strong><br />

660 kilometres <strong>in</strong> length. Flow prediction <strong>in</strong> transport pipel<strong>in</strong>es is therefore<br />

best assessed us<strong>in</strong>g one-dimensional models.<br />

Oil <strong>and</strong> gas transport systems <strong>in</strong>clude quick act<strong>in</strong>g valves. The actuation of a<br />

quick act<strong>in</strong>g valve generates a pressure surge, also called water hammer. The<br />

pressure surge generated propagates along the transmission l<strong>in</strong>e as a pressure<br />

wave. The water hammer pressure amplitude is calculated us<strong>in</strong>g the<br />

Joukowsky equation (Gudmundsson <strong>and</strong> Celius, 1999):<br />

∆p = ρa∆u (C.1)<br />

where, ρ, is fluid density, a, speed of sound <strong>in</strong> the fluid-pipe system, ∆u,<br />

change <strong>in</strong> fluid velocity, <strong>and</strong>, ∆p, water hammer pressure amplitude.<br />

The pressure surge generated by the actuation of a quick act<strong>in</strong>g valve propagates<br />

along the transmission l<strong>in</strong>e at the speed of sound. The speed of sound<br />

depends on the fluid properties; that is the fluid density, ρ, <strong>and</strong> the fluid compressibility,<br />

K. The speed of sound depends also on the pipe material; that is<br />

the Young’s modulus, E, <strong>and</strong> on the pipe geometry; that is the diameter, D,<br />

<strong>and</strong> the wall thickness, e. The speed of sound is calculated us<strong>in</strong>g Korteweg’s<br />

formula (Tijssel<strong>in</strong>g <strong>and</strong> Anderson, 2007):<br />

1<br />

a = � �<br />

ρ K + D<br />

�<br />

Ee<br />

(C.2)<br />

The water hammer <strong>and</strong> pressure wave propagation, <strong>in</strong> fluid transport pipel<strong>in</strong>es,<br />

are modelled us<strong>in</strong>g the mass balance equation, <strong>and</strong> the momentum equation,<br />

for 1D unsteady <strong>and</strong> compressible fluids <strong>in</strong> flexible pipel<strong>in</strong>es. The mass balance<br />

<strong>and</strong> momentum equations form the follow<strong>in</strong>g system of equations:<br />

∂u<br />

∂t<br />

+ u∂u<br />

∂x<br />

1 ∂p<br />

ρ ∂x<br />

1<br />

ρa2 Dp ∂u<br />

+ = 0<br />

Dt ∂x<br />

1 ∂ (β − 1)ρAu<br />

+<br />

ρA<br />

2<br />

+<br />

∂x<br />

τwπD<br />

+ gs<strong>in</strong>(α) + = 0<br />

ρA<br />

219<br />

(C.3)


Appendix C<br />

where, A, is pipe cross-sectional area, g, gravitational acceleration, α, pipel<strong>in</strong>e<br />

<strong>in</strong>cl<strong>in</strong>ation, <strong>and</strong>, τw, wall shear stress. The pressure wave propagation <strong>in</strong> fluid<br />

transport generates an unsteady flow of a compressible fluid. Typical fluid<br />

velocities range from 10 m/s to 20 m/s for gas transport <strong>and</strong> from 2 m/s to 4<br />

m/s for oil transport. Typical speeds of sound <strong>in</strong> gas <strong>and</strong> oil are of order 300<br />

m/s <strong>and</strong> 1200 m/s, respectively. The flow Mach number, def<strong>in</strong>ed by the ratio<br />

of the fluid velocity <strong>and</strong> the acoustic velocity, is therefore much less than 1,<br />

both <strong>in</strong> oil <strong>and</strong> gas pipel<strong>in</strong>es.<br />

The relative magnitude of the various terms <strong>in</strong> (C.4) is <strong>in</strong>vestigated for a full<br />

wave scale; that is four times the travel time of a pressure wave propagat<strong>in</strong>g<br />

through the fluid transport pipel<strong>in</strong>e. The low Mach number approximation<br />

is also used <strong>in</strong> the magnitude analysis. In addition, changes <strong>in</strong> density <strong>in</strong><br />

unsteady compressible flows are of the order of magnitude of the Mach number.<br />

The simplifications applied on (C.4) <strong>and</strong> assum<strong>in</strong>g a horizontal pipel<strong>in</strong>e read:<br />

∂p<br />

+ ρa2∂u = 0<br />

∂t ∂x<br />

∂u<br />

∂t<br />

∂p<br />

+ = s<br />

∂x<br />

(C.4)<br />

where, s, represents a source term account<strong>in</strong>g for the energy losses. The<br />

above system of equations is the one used <strong>in</strong> the <strong>in</strong>dustry for prediction of<br />

pressure transients. The method of characteristics is commonly used to solve<br />

numerically the above system of equations (Wylie et al., 1993). A typical<br />

experimental result of the water hammer effect <strong>in</strong> a pipel<strong>in</strong>e is shown <strong>in</strong> Figure<br />

C.1. The maximum pressure value decays <strong>and</strong> the pressure trace smooths with<br />

time.<br />

Source Term<br />

The source term, s, <strong>in</strong> the system of equations (C.5), accounts for the energy<br />

losses <strong>in</strong> the fluid transport system. External forces on the fluid are assumed<br />

negligible. Heat transfer between the fluid <strong>and</strong> its surround<strong>in</strong>g is assumed<br />

negligible. The pipe visco-elastic behaviour is not considered <strong>in</strong> the present<br />

study, neither <strong>in</strong> the speed of sound formulation (C.2) nor <strong>in</strong> the mass conservation<br />

equation. The source term, s, is then reduced to the frictional pressure<br />

loss of the flow along the pipel<strong>in</strong>e <strong>and</strong> is expressed by:<br />

s = −f<br />

ρu | u | (C.5)<br />

2D<br />

220


Appendix C<br />

where, f, is Darcy-Weisbach friction factor. The friction factor can be accurately<br />

estimated by Haal<strong>and</strong>’s explicit formula (Haal<strong>and</strong>, 1983):<br />

1<br />

√ =<br />

f −1.8<br />

n log<br />

��6.9�n � � �<br />

1.1n<br />

k<br />

+<br />

(C.6)<br />

Re 3.75D<br />

where, Re, def<strong>in</strong>es flow Reynolds number <strong>and</strong>, n, a constant equal to 3 for a<br />

gas <strong>and</strong> 1 for a liquid.<br />

The aforementioned system of equations for predict<strong>in</strong>g pressure transient propagation<br />

is generally solved numerically by the methods of characteristics. The<br />

Darcy-Weisbach frictional pressure loss is the only energy loss mechanism<br />

taken <strong>in</strong>to account. The computed solution does not predict accurately pressure<br />

wave behaviour <strong>in</strong> fluid transport pipel<strong>in</strong>es (Szymkiewicz <strong>and</strong> Mitosek,<br />

2007). The pressure wave decay is underestimated, <strong>and</strong> the pressure wave is<br />

not smoothed correctly; therefore the computed pressure traces do not correspond<br />

to experimental data. Several models have therefore been proposed to<br />

<strong>in</strong>crease the accuracy of computer models.<br />

The models proposed <strong>in</strong> the literature rely on a decomposition of the frictional<br />

pressure drop <strong>in</strong> the source term, s. The frictional pressure drop is split <strong>in</strong><br />

two terms. The first term is the steady frictional pressure drop calculated<br />

us<strong>in</strong>g the st<strong>and</strong>ard Darcy-Weisbach friction factor, fq. The second term is<br />

a transient friction term, fu. The transient friction factor is time dependent<br />

<strong>and</strong> is related to the flow parameters. The transient friction factor <strong>in</strong>creases<br />

locally the frictional pressure drop at the wave front, where the flow gradient<br />

is important. The selected models for the present work are express<strong>in</strong>g the<br />

transient friction factor value.<br />

Two proposed models of transient friction factors are here<strong>in</strong> studied. The first<br />

transient friction factor model studied was proposed by Brunone et al. (1991).<br />

The formulation proposed by Brunone <strong>in</strong>cludes the local <strong>and</strong> <strong>in</strong>stantaneous<br />

convective acceleration of the fluid. The second transient friction factor model<br />

studied was proposed by Ramos et al. (2004). The formulation proposed by<br />

Ramos is an improvement of Brunone’s formulation aim<strong>in</strong>g at <strong>in</strong>creas<strong>in</strong>g the<br />

accuracy of the solution. Brunone’s transient friction factor formulation is<br />

expressed by:<br />

fu = k0D<br />

u | u |<br />

� ∂u<br />

∂t<br />

�<br />

− a∂u<br />

∂x<br />

(C.7)<br />

where, k0, is an empirical constant related to the Reynolds number. k0 is of<br />

order 0.02-0.03. Ramos’ transient friction factor formulation is expressed by:<br />

fu = D<br />

�<br />

� ��<br />

∂u u �<br />

k1 − k2a �<br />

∂u�<br />

�<br />

u | u | ∂t | u | �∂x<br />

�<br />

(C.8)<br />

221


Appendix C<br />

where, k1, is a constant <strong>and</strong> of order 0.003, <strong>and</strong>, k2, a constant of order<br />

0.04. Both transient friction factor formulations <strong>in</strong>clude the local <strong>and</strong> the<br />

<strong>in</strong>stantaneous convective accelerations which were neglected <strong>in</strong> (4) because of<br />

the low Mach Number <strong>and</strong> the magnitude analysis. It is noted that the system<br />

of equations <strong>in</strong>clud<strong>in</strong>g a transient friction factor has been studied <strong>and</strong> solved<br />

numerically by their respective authors us<strong>in</strong>g the method of characteristics.<br />

Modell<strong>in</strong>g Equations<br />

The system of equations (C.5) is a hyperbolic system of two first-order partial<br />

differential equations. The transient friction factors (C.7) <strong>and</strong> (C.8) are<br />

characterised by first-order partial differential equations. The one-dimensional<br />

compressible flow cont<strong>in</strong>uity <strong>and</strong> mass equations (C.5) for horizontal conduits<br />

may be written <strong>in</strong> their conservative form:<br />

∂Q<br />

∂t<br />

+ ∂F (Q)<br />

∂x<br />

= S (C.9)<br />

consist<strong>in</strong>g of the vector variable, Q, the flux vector, F, <strong>and</strong> the source term<br />

vector, S:<br />

� � � � � �<br />

p<br />

ρa2u 0<br />

Q = F(Q) = S =<br />

(C.10)<br />

u<br />

p s<br />

The well-established method of characteristics is based on f<strong>in</strong>ite difference<br />

approximations of the partial derivations. The method of characteristics is<br />

therefore first order. Szymkiewicz <strong>and</strong> Mitosek (2007) used a third order<br />

f<strong>in</strong>ite element method to model the govern<strong>in</strong>g system (9). We chose to apply<br />

a second order f<strong>in</strong>ite volume method <strong>in</strong> order to compare the results based on<br />

the <strong>in</strong>volved physical phenomena <strong>and</strong> not mathematical considerations.<br />

The modell<strong>in</strong>g <strong>and</strong> numerical simulation <strong>in</strong> the present work is based on a<br />

f<strong>in</strong>ite volume method for hyperbolic systems <strong>in</strong>cluded <strong>in</strong> CLAWPACK (LeVeque,<br />

2002). CLAWPACK is a software package of high-resolution, multidimensional<br />

wave propagation algorithms, for time dependent hyperbolic problems.<br />

CLAWPACK <strong>in</strong>cludes Riemann solvers of first <strong>and</strong> second order <strong>in</strong> time<br />

<strong>and</strong> space. Flux limiters are then applied to the result<strong>in</strong>g waves, <strong>in</strong>creas<strong>in</strong>g<br />

the accuracy of the solution.<br />

The Lax-Wendroff numerical scheme with Van-Leer flux limiter was selected<br />

for solv<strong>in</strong>g the system of equations here<strong>in</strong> studied. The homogeneous system<br />

is first solved us<strong>in</strong>g the Lax-Wendroff numerical scheme. The Van-Leer flux<br />

limiter is applied directly on the flux vector F. Then, the source term approximated<br />

by f<strong>in</strong>ite differences is applied to the solution. The Lax-Wendroff<br />

222


Appendix C<br />

flux limited method applied to a homogeneous system takes the form (Laney,<br />

1998):<br />

if Cr > 0<br />

Q n+1<br />

i<br />

= Qn i − Cr(Q n i − Q n i−1)−<br />

Cr<br />

(1 − Cr)<br />

2<br />

if Cr < 0<br />

Q n+1<br />

i<br />

�<br />

Φ(θ n i+1/2 )(Qn i+1 − Qn i ) − Φ(θn i−1/2 )(Qn i − Qn i−1 )<br />

= Qn i − Cr(Q n i+1 − Q n i )−<br />

Cr<br />

(1 + Cr)<br />

2<br />

�<br />

Φ(θ n i+1/2 )(Qn i+1 − Qn i ) − Φ(θn i−1/2 )(Qn i − Qn i−1 )<br />

�<br />

�<br />

(C.11)<br />

where, Cr, represents Courant number, Φ(θ), flux limiter function, subscript<br />

i, computational node <strong>and</strong>, superscript n, computational time step. The calculated<br />

Courant number is relative to each solution wave <strong>and</strong> its direction of<br />

propagation; that is the Courant number is positive when the wave propagates<br />

<strong>in</strong> the positive direction +x, <strong>and</strong> negative otherwise. The Van-Leer flux<br />

limiter function takes the form (Laney, 1998):<br />

Φ(θ) =<br />

θ + |θ|<br />

1 + |θ|<br />

where the variable θ is def<strong>in</strong>ed by:<br />

if Cr > 0<br />

θ n i−1/2 = Qn i−1 − Qn i−2<br />

Q n i − Qn i−1<br />

if Cr < 0<br />

θ n i−1/2 = Qn i+1 − Qn i<br />

Q n i − Qn i−1<br />

(C.12)<br />

(C.13)<br />

The Lax-Wendroff numerical scheme is explicit, centered, second order accurate<br />

<strong>in</strong> both time <strong>and</strong> space, l<strong>in</strong>early stable provided that the Courant number<br />

is less than one. The transient friction factor equations (7) <strong>and</strong> (8) are first<br />

order partial differential equations <strong>and</strong> are solved us<strong>in</strong>g f<strong>in</strong>ite differences.<br />

223


Appendix C<br />

Modell<strong>in</strong>g Results<br />

A test case is used to compare the two transient friction factors studied <strong>in</strong> this<br />

work. The test case is a water pipel<strong>in</strong>e, 1 <strong>in</strong>ch <strong>in</strong> diameter, 100 meter <strong>in</strong> length.<br />

The speed of sound is 1400 m/s. The <strong>in</strong>itial fluid velocity is 2 m/s. Upstream;<br />

the transport pipel<strong>in</strong>e is connected to a reservoir at constant pressure, 50 bara.<br />

Downstream, the pipel<strong>in</strong>e is connected to a quick act<strong>in</strong>g valve. A schematic<br />

representation of the transport system is shown <strong>in</strong> Figure C.2. The upstream<br />

reservoir is shown on the left-h<strong>and</strong> side, <strong>and</strong> the quick-act<strong>in</strong>g valve is shown<br />

on the right-h<strong>and</strong> side. The water transport system is meshed us<strong>in</strong>g 10,000<br />

grid po<strong>in</strong>ts. The Courant number is set to 0.99.<br />

The flow is fully established, then the quick-act<strong>in</strong>g valve is actuated at time<br />

t=0.01s. The quick act<strong>in</strong>g of the check valve is ideal. In other word, the<br />

actuation of the valve assumes no energy loss. The quick act<strong>in</strong>g of the valve<br />

generates a pressure wave that propagates along the transmission l<strong>in</strong>e as a<br />

pressure wave. Follow<strong>in</strong>g the Joukowsky equation (C.1), the pressure surge<br />

generated is 28 bar <strong>in</strong> amplitude. The pressure surge is superimposed on top<br />

of the <strong>in</strong>-situ static pressure. The Reynolds number is 50800. The flow is<br />

therefore fully turbulent before the valve closure.<br />

The selected f<strong>in</strong>ite volume method is first compared to the st<strong>and</strong>ard method of<br />

characteristics. The comparison is performed by simulat<strong>in</strong>g the pressure wave<br />

successive transmissions <strong>and</strong> reflections without the Darcy-Weisbach steadystate<br />

friction factor or a transient friction factor; that is, the pressure wave<br />

propagates without energy loss. The result is shown <strong>in</strong> Figure C.3. Both<br />

models predict the same pressure wave amplitude accord<strong>in</strong>g to (C.1). However,<br />

the MOC result decays with time as shown <strong>in</strong> Figure C.4, while the result<br />

simulated by the f<strong>in</strong>ite volume method rema<strong>in</strong>s periodic without noticeable<br />

decay over the simulated period of time. The MOC shows discrepancy with<br />

the expected result while the selected numerical scheme decay can be neglected<br />

for the period of time simulated.<br />

The method of characteristics <strong>in</strong>troduces numerical friction <strong>in</strong>to the solution.<br />

That is the damp<strong>in</strong>g is greater than what the physical model describes. Therefore,<br />

the predicted results do not correspond entirely to the physics of the problem<br />

<strong>and</strong> discrepancy are observed between experimental data <strong>and</strong> computed<br />

solutions. Such observations suit the modell<strong>in</strong>g of the system of equations<br />

when <strong>in</strong>clud<strong>in</strong>g a source term or not. Therefore, the transient friction factor is<br />

not modelled <strong>in</strong> the present work with the method of characteristics. Instead,<br />

the present work focuses on the effect of the transient friction factor on the<br />

solution computed with the selected f<strong>in</strong>ite volume method.<br />

The pressure wave transmission <strong>and</strong> reflection are modelled successively us<strong>in</strong>g<br />

224


Appendix C<br />

four different expressions of the source term. The first expression of the source<br />

term does not <strong>in</strong>clude any frictional pressure loss; that is, the system is ideal<br />

without energy loss. The second expression of the source term <strong>in</strong>cludes only<br />

Darcy-Weisbach steady friction factor. The third expression of the source<br />

term adds Brunone’s transient friction factor to the steady friction factor.<br />

The fourth expression of the source term adds Ramos’ transient friction factor<br />

to the steady friction factor.<br />

The pressure trace at the downstream valve is computed <strong>in</strong> the four cases.<br />

The pressure traces are shown <strong>in</strong> Figure C.5 <strong>and</strong> Figure C.6. The pressure<br />

wave propagates on top of the <strong>in</strong>-situ static pressure. The water hammer<br />

pressure amplitude is accurately predicted accord<strong>in</strong>g to (1). The wave propagation<br />

speed is respected. The modell<strong>in</strong>g results are <strong>in</strong> good agreement with<br />

the physics of the water hammer phenomenon. However, the steady friction<br />

model solution <strong>and</strong> the two transient friction models solution are almost superimposed.<br />

It means that the proposed models by Brunone et al. <strong>and</strong> Ramos<br />

et al. do not improve significantly the predicted solution. The result is similar<br />

to observations of Szymkiewicz <strong>and</strong> Mitosek (2007).<br />

A detail of the pressure trace at the valve is shown <strong>in</strong> Figure C.7. The solutions<br />

us<strong>in</strong>g only the steady friction factor <strong>and</strong> us<strong>in</strong>g the steady friction factor <strong>and</strong><br />

Brunone’s transient friction factor are superimposed. Therefore, Brunone’s<br />

proposed model does not improve the prediction. The solution us<strong>in</strong>g the steady<br />

friction factor with Ramos transient friction factor formulation is slightly lower<br />

than the two aforementioned solutions. Especially, the wave front pressure is<br />

lower. However, the difference between the Ramos et al. ’s proposed model<br />

<strong>and</strong> Brunone et al. ’s proposed model is negligible. Consequently, none of the<br />

proposed model used <strong>in</strong> the present work is evaluated as be<strong>in</strong>g satisfactory.<br />

Discussion<br />

The test case has been modelled <strong>in</strong> four different ways. None of the solution<br />

exhibits the same characteristics as the typical water hammer pressure trace<br />

(Figure C.1). The transient friction factors <strong>in</strong>cluded <strong>in</strong> the source term modell<strong>in</strong>g<br />

did not improve significantly the accuracy of the computed solution. The<br />

result from the present work confirms Szymkiewicz <strong>and</strong> Mitosek observations<br />

(2007).<br />

Historically, the first source term model us<strong>in</strong>g transient effects was developed<br />

by Zielke (1966). In Zielke’s model, the source term <strong>in</strong>cluded the past history<br />

of the fluid velocity. The transient friction models proposed by Brunone et al.<br />

(1991) <strong>and</strong> Ramos et al. (2004) use the past history of the flow by tak<strong>in</strong>g <strong>in</strong>to<br />

account the convective acceleration of the fluid. All transient friction mod-<br />

225


Appendix C<br />

els developed were tested us<strong>in</strong>g the method of characteristics. The method<br />

of characteristics does not ensure mass conservation, only momentum conservation.<br />

The method of characteristics is also diffusive <strong>and</strong> dissipative; that<br />

is the pressure wave decay <strong>in</strong>cluded <strong>in</strong> the mathematical model is <strong>in</strong>creased<br />

by an artificial decay due to the numerical method. The <strong>in</strong>herent properties<br />

of the method of characteristics are consequently <strong>in</strong>appropriate for accurate<br />

unsteady flow predictions.<br />

Szymkiewicz <strong>and</strong> Mitosek used a modified f<strong>in</strong>ite element method with adjustable<br />

accuracy, up to the third order <strong>in</strong> space <strong>and</strong> time (Szymkiewicz <strong>and</strong><br />

Mitosek, 2005) for modell<strong>in</strong>g the mass conservation <strong>and</strong> momentum equations.<br />

Szymkiewicz <strong>and</strong> Mitosek results are <strong>in</strong> good agreement with the results<br />

here<strong>in</strong> presented; that is the transient friction factor does not improve<br />

the prediction. They demonstrated that the mathematical functions developed<br />

for modell<strong>in</strong>g the transient friction factor does not change the hyperbolic nature<br />

of the system of equations. Szymkiewicz <strong>and</strong> Mitosek suggested that<br />

an advection-diffusion transport can expla<strong>in</strong>s the smooth<strong>in</strong>g. They concluded<br />

that the current models are not suitable for predict<strong>in</strong>g the damp<strong>in</strong>g <strong>and</strong> the<br />

smooth<strong>in</strong>g of a pressure wave.<br />

The present work uses a f<strong>in</strong>ite volume method, second order accurate <strong>in</strong> both<br />

space <strong>and</strong> time. The presented results are not compared with experimental<br />

data, however, the conclusions are similar to (Szymkiewicz <strong>and</strong> Mitosek,<br />

2007); that is the current transient friction models do not improve the accuracy<br />

of a pressure wave behaviour. While the pressure wave is damped, it<br />

is not smoothed. Besides, it is observed that the transient friction formulation<br />

is <strong>in</strong> contradiction with the low Mach number approximation used for<br />

establish<strong>in</strong>g the mass <strong>and</strong> momentum conservation equations. Such approximation<br />

m<strong>in</strong>imizes the effect of the convective acceleration. Nevertheless, the<br />

transient friction models are centred on the local <strong>and</strong> <strong>in</strong>stantaneous convective<br />

acceleration terms.<br />

The transient friction models are <strong>in</strong>adequate for expla<strong>in</strong><strong>in</strong>g both the damp<strong>in</strong>g<br />

<strong>and</strong> the smooth<strong>in</strong>g of the pressure wave along the pipel<strong>in</strong>e. Besides work on<br />

the mathematics of the source term <strong>and</strong> the system of equations to be solved,<br />

additional fundamental work should be performed on the pressure wave propagation<br />

phenomena. Ramos et al. (2004) added the visco-elastic behaviour<br />

of the pipe material to the mass conservation equation. The proposed model<br />

was aga<strong>in</strong> tested us<strong>in</strong>g the well-established method of characteristics.<br />

Better modell<strong>in</strong>g of pressure wave propagation is still to be established. Further<br />

modell<strong>in</strong>g of pressure wave propagation will have to <strong>in</strong>clude new physical<br />

statements not <strong>in</strong>cluded <strong>in</strong> the current models; that is the modell<strong>in</strong>g with the<br />

st<strong>and</strong>ard assumptions of low Mach number <strong>and</strong> simplifications after order of<br />

226


Appendix C<br />

magnitude analysis, as used <strong>in</strong> the present work. Moreover, further models<br />

will have to be tested us<strong>in</strong>g several numerical methods to demonstrate their<br />

<strong>in</strong>dependence with respect to the simulation tool.<br />

Conclusions<br />

A second order f<strong>in</strong>ite volume has been applied successfully to model one dimensional<br />

compressible mass conservation <strong>and</strong> momentum equations. Two<br />

different expressions have been applied to model the source term; that is<br />

steady-state friction <strong>and</strong> transient friction.<br />

The present work confirms that the transient friction factor <strong>in</strong> the literature<br />

do not improve the pressure wave predictions currently used. Of particular<br />

<strong>in</strong>terest, it is shown that the transient friction factor is not responsible for the<br />

smooth<strong>in</strong>g <strong>and</strong> damp<strong>in</strong>g of a pressure wave propagat<strong>in</strong>g <strong>in</strong> a fluid transport<br />

pipel<strong>in</strong>e.<br />

Further study of pressure wave propagation should focus on the fundamentals.<br />

Assumptions currently used to establish the govern<strong>in</strong>g system of equations<br />

will have to be re-evaluated. New models will have to be tested with diverse<br />

numerical methods to ensure their reliability.<br />

Acknowledgment<br />

The authors want to thank Clavis <strong>Impulse</strong> Technology AS for provid<strong>in</strong>g a<br />

Ph.D scholarship to Benjam<strong>in</strong> Pierre.<br />

References<br />

Brunone B., Golia U. <strong>and</strong> Greco M. (1991): Some Remarks on the Momentum<br />

Equation for Fast Transients, Int. Meet<strong>in</strong>g on Hydraulic Transients with<br />

Column Separation, volume 9, pages 140-148.<br />

Gudmundsson J.S. <strong>and</strong> Celius H.K. (1999): Gas-Liquid Meter<strong>in</strong>g Us<strong>in</strong>g <strong>Pressure</strong>-<br />

Pulse Technology, SPE Annual Technical Conference <strong>and</strong> Exhibition, Houston,<br />

SPE 56584.<br />

Haal<strong>and</strong> S. (1983): Simple <strong>and</strong> Explicit Formulas for the Friction Factor <strong>in</strong><br />

Turbulent Pipe Flow, Journal of Fluids Eng<strong>in</strong>eer<strong>in</strong>g, volume 105(1).<br />

Laney C. (1998): Computational Gasdynamics, Cambridge University Press.<br />

227


Appendix C<br />

LeVeque R.J. (2002): F<strong>in</strong>ite Volume Methods for Hyperbolic Problems, Cambridge<br />

University Press.<br />

Pierre B. <strong>and</strong> Gudmundsson J.S. (2008): Water Hammer Simulation of <strong>Pressure</strong><br />

<strong>in</strong> the Failed Flexible Hose for Off-Load<strong>in</strong>g Oil at Statfjord A, Department<br />

of Petroleum Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Applied Geophysics, <strong>NTNU</strong>, 7491 Trondheim,<br />

Technical Report.<br />

Pierre B. <strong>and</strong> Gudmundsson J.S. (2009): Pump<strong>in</strong>g of Fluids Us<strong>in</strong>g <strong>Pressure</strong><br />

<strong>Impulse</strong>s, SPE EUROPEC, Amsterdam, SPE 120896.<br />

Ramos H., Covas D., Borga A. <strong>and</strong> Loureiro D. (2004): Surge Damp<strong>in</strong>g Analysis<br />

<strong>in</strong> Pipe Systems - Modell<strong>in</strong>g <strong>and</strong> Experiments, Journal of Hydraulic Research,<br />

Volume 42(4).<br />

Szymkiewicz R. <strong>and</strong> Mitosek M. (2005): Analysis of Unsteady Pipe Flow Us<strong>in</strong>g<br />

the Modified F<strong>in</strong>ite Element Method, Communications <strong>in</strong> Numerical Methods<br />

<strong>in</strong> Eng<strong>in</strong>eer<strong>in</strong>g, volume 21(4).<br />

Szymkiewicz R. <strong>and</strong> Mitosek M. (2007): Numerical Aspects of Improvement of<br />

the Unsteady Pipe Flow Equations, International Journal for Numerical Methods<br />

<strong>in</strong> Fluids, volume 55(11).<br />

Tijssel<strong>in</strong>g A. <strong>and</strong> Anderson A. (2007): Johannes Von Kries <strong>and</strong> the History<br />

of Water Hammer, Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g, volume 133(1).<br />

Wylie E.B., Streeter V.L. <strong>and</strong> Suo L. (1993): Fluid Transients <strong>in</strong> Systems,<br />

Prentice Hall.<br />

Zielke W. (1966): Frequency Dependent Friction <strong>in</strong> Transient Pipe Flow, Industry<br />

Program of the College of Eng<strong>in</strong>eer<strong>in</strong>g, University of Michigan.<br />

228


Nomenclature<br />

a Acoustic Velocity [ms −1 ]<br />

A Flow Area [m 2 ]<br />

Cr Courant Number<br />

D Diameter [m]<br />

f Friction Factor<br />

g Gravitational Acceleration [ms −2 ]<br />

K Fluid Compressibility [Pa −1 ]<br />

L Pipe Length [m]<br />

p <strong>Pressure</strong> [Pa]<br />

Re Reynolds Number<br />

u Fluid Velocity [ms −1 ]<br />

α Angle [rad]<br />

ρ Fluid Density [kg m −3 ]<br />

Figures<br />

<strong>Pressure</strong><br />

2 4 6 8 10 12<br />

Figure C.1: Typical Water Hammer <strong>Pressure</strong> Trace<br />

229<br />

Appendix C<br />

Time [2L/a]


Appendix C<br />

L = 100m<br />

Figure C.2: Schematic Representation of Test Case System<br />

Figure C.3: Comparison Between the Method of Characteristics<br />

(dotted l<strong>in</strong>e) <strong>and</strong> the Selected F<strong>in</strong>ite Volume Method<br />

(solid l<strong>in</strong>e)<br />

Ideal <strong>Pressure</strong> Wave Propagation; that is Without Energy Loss.<br />

230<br />

D


Figure C.4: Detail of Figure (C.3)<br />

Solid L<strong>in</strong>e: F<strong>in</strong>ite Volume Method<br />

Dotted L<strong>in</strong>e: Method of Characteristics<br />

231<br />

Appendix C


Appendix C<br />

Figure C.5: <strong>Pressure</strong> Trace at Non-Return Valve<br />

Dotted L<strong>in</strong>e: No Source Term<br />

Dashed L<strong>in</strong>e: Source Term with Steady-State Friction Factor Only<br />

Plus Sign: Source Term with Steady-State Friction <strong>and</strong> Brunone’s Transient<br />

Friction factor<br />

232


Appendix C<br />

Figure C.6: <strong>Pressure</strong> Trace at the Valve<br />

Dotted L<strong>in</strong>e: No Source Term<br />

Dashed L<strong>in</strong>e: Source Term with Steady-State Friction Factor Only<br />

Cross Sign: Source Term with Steady-State Friction <strong>and</strong> Ramos Transient<br />

Friction factor<br />

233


Appendix C<br />

Figure C.7: Detail of the <strong>Pressure</strong> Trace at the Valve<br />

Dotted L<strong>in</strong>e: No Source Term<br />

Dashed L<strong>in</strong>e: Source Term with Steady-State Friction Factor Only<br />

Plus Sign: Source Term with Steady-State Friction <strong>and</strong> Brunone’s Transient<br />

Friction factor<br />

Cross Sign: Source Term with Steady-State Friction <strong>and</strong> Ramos Transient<br />

Friction factor<br />

234


Appendix D<br />

Simulation of Rapid <strong>Pressure</strong><br />

Transients <strong>in</strong><br />

Statfjord A Off-Load<strong>in</strong>g<br />

Pipel<strong>in</strong>e<br />

Paper to be submitted.<br />

Abstract<br />

The second largest oil spillage <strong>in</strong> Norwegian petroleum history occurred on<br />

December 12, 2007, when an off-load<strong>in</strong>g pipel<strong>in</strong>e ruptured. Results of numerical<br />

simulations of pressure transients <strong>in</strong> the off-load<strong>in</strong>g system are presented.<br />

Of particular <strong>in</strong>terest is the <strong>in</strong>fluence of acoustic properties of pipel<strong>in</strong>e sections<br />

<strong>and</strong> equipment on the pressure transient propagation. The damaged Statfjord A<br />

off-load<strong>in</strong>g system geometry <strong>and</strong> characteristics as well as the acoustic properties<br />

are described. The paper describes numerical simulations of pressure transients<br />

<strong>in</strong> the off-load<strong>in</strong>g pipel<strong>in</strong>e. The paper shows particularly how changes<br />

<strong>in</strong> acoustic properties generate pressure wave reflections that superimpose <strong>and</strong><br />

create rapid <strong>and</strong> local pressure peaks.<br />

Introduction<br />

Hydrocarbon transport pipel<strong>in</strong>es are subject to rapid pressure transients <strong>in</strong> the<br />

form of water-hammer events. Rapid pressure transients can arise when quickact<strong>in</strong>g<br />

valves are suddenly activated or after failure of a pump. Consequences<br />

235


Appendix D<br />

of rapid pressure transients <strong>in</strong> hydrocarbon transport pipel<strong>in</strong>es range from<br />

operational damages such as leakages (Moura et al., 2004; Wang et al., 2004),<br />

to pipe rupture (Misiunas et al.; 2005, NPD, 2008). Therefore, rapid pressure<br />

transients need to be accurately predicted <strong>in</strong> hydrocarbon transport pipel<strong>in</strong>es.<br />

Fluid transport systems consist typically of many <strong>in</strong>dividual pipe sections <strong>and</strong><br />

pieces of equipment. Valves <strong>and</strong> pumps are among the pieces of equipment<br />

<strong>in</strong>tegrated <strong>in</strong>to a fluid transport system. Calculations <strong>and</strong> modell<strong>in</strong>g of rapid<br />

pressure transient events need to account for all the sections <strong>and</strong> pieces of<br />

equipment <strong>in</strong>cluded <strong>in</strong> the system.<br />

The Statfjord A off-load<strong>in</strong>g system is a 2,808 m long oil transport system,<br />

made of 38 sections. The off-load<strong>in</strong>g system connects a storage tank located<br />

on a production platform, to a s<strong>in</strong>gle path coupler located on a tanker. The<br />

Statfjord A off-load<strong>in</strong>g system ruptured on December 12, 2007, due to a rapid<br />

pressure transient event generated by sudden actuation of the s<strong>in</strong>gle path<br />

coupler (Pierre <strong>and</strong> Gudmundsson, 2008).<br />

Rupture of the Statfjord A off-load<strong>in</strong>g system caused 4,400 m 3 of oil be<strong>in</strong>g<br />

pumped <strong>in</strong>to the sea; thus generat<strong>in</strong>g the second largest spillage <strong>in</strong> the history<br />

of oil activities <strong>in</strong> Norway. Predictions of rapid pressure transients need to be<br />

accurate <strong>in</strong> order to prevent future accidents.<br />

Predictions of rapid pressure transients with <strong>and</strong> without consider<strong>in</strong>g all the<br />

sections produce different results. Of particular <strong>in</strong>terest is the <strong>in</strong>fluence of the<br />

many sections geometrical <strong>and</strong> material properties on the maximum pressure<br />

reached.<br />

Statfjord Off-Load<strong>in</strong>g System<br />

Statfjord is an oil field <strong>in</strong> the Tampen area <strong>in</strong> the northern part of the North<br />

Sea (NPD, 2008). Statfjord is one of the oldest produc<strong>in</strong>g field on the Norwegian<br />

cont<strong>in</strong>ental shelf. The Statfjord reservoirs lie between 2,500 <strong>and</strong> 3,000<br />

m deep. The sea depth <strong>in</strong> the area is about 150 m. A total of approximately<br />

47,000 barrels of oil were produced per day <strong>in</strong> 2008 from the Statfjord field.<br />

The field has been developed with three fully <strong>in</strong>tegrated facilities: Statfjord<br />

A, Statfjord B <strong>and</strong> Statfjord C. Statfjord A was the first facility <strong>in</strong> place <strong>and</strong><br />

is centrally positioned <strong>in</strong> the field.<br />

The produced oil is stored <strong>in</strong> cells (tanks) at each facility. Storage cells are<br />

located <strong>in</strong>side the platform structure at the bottom of the sea. The storage<br />

cell on Statfjord A holds 206,000 st<strong>and</strong>ard cubic metres; that is 1.3 million<br />

barrels of oil. Oil from the storage cell is loaded to tankers approximately<br />

every two days. Tankers connect to the off-load<strong>in</strong>g system at a safe distance<br />

236


Appendix D<br />

Figure D.1: Illustration of the Off-Load<strong>in</strong>g System between the Riser<br />

Base <strong>and</strong> the S<strong>in</strong>gle Path Coupler<br />

away from the platform. The off-load<strong>in</strong>g system is the whole system from the<br />

storage cell to the off-load<strong>in</strong>g po<strong>in</strong>t <strong>and</strong> comprises two load<strong>in</strong>g pumps located<br />

near the platform (PSA, 2008).<br />

The Statfjord A off-load<strong>in</strong>g system is made of several sections as partly shown<br />

<strong>in</strong> Figure D.1 (Sangesl<strong>and</strong>, 2008). The first section, not shown <strong>in</strong> Figure D.1,<br />

is a 2,516 m long steel pipe connect<strong>in</strong>g the storage cells on the production<br />

platform, to a riser. A 80 m long flexible hose then connects the base of<br />

the riser to a swivel. A sequence of flexible hoses <strong>in</strong>ter-connected by steel<br />

components l<strong>in</strong>ks the swivel to a s<strong>in</strong>gle path coupler located on a tanker. The<br />

whole system is 2,808 m <strong>in</strong> length.<br />

The composition of the Statfjord A processed oil determ<strong>in</strong>es such properties<br />

as density, viscosity <strong>and</strong> compressibility. An exact composition of the relevant<br />

oil is not available <strong>in</strong> the literature. Therefore, an early composition of the<br />

Statfjord A reservoir oil was used to f<strong>in</strong>d the properties of the processed oil.<br />

The processed oil composition from Statfjord is given <strong>in</strong> Table D.1 (Johnsen,<br />

2008). The reservoir hydrocarbon has been flashed down to 1 bara, 5 ◦ C before<br />

be<strong>in</strong>g pressurized to 30 bara.<br />

Hydrocarbon properties have been obta<strong>in</strong>ed us<strong>in</strong>g the Peng Rob<strong>in</strong>son fluid<br />

package <strong>in</strong>cluded <strong>in</strong> HYSYS. The fluid flow<strong>in</strong>g <strong>in</strong> the transport pipel<strong>in</strong>e be-<br />

237


Appendix D<br />

Table D.1: Statfjord Oil Molar Composition at 5 ◦ C, 30 bara<br />

Component N2 CO2 C1 C2 C3 iC4<br />

% mol 0.04 0.19 9.32 6.25 10.27 1.81<br />

Component nC4 iC5 nC5 C6 C7 C8<br />

% mol 6.25 2.09 3.32 5.33 7.44 6.87<br />

Component C9 C10 C11 C12<br />

% mol 5.61 4.43 3.57 27.19<br />

tween the storage cells (tanks) <strong>and</strong> the tanker is approximately at 30 bara<br />

pressure <strong>and</strong> at 5 ◦ C temperature. For such conditions, a 705 kg m −3 density<br />

<strong>and</strong> a 0.635 mPas viscosity are obta<strong>in</strong>ed. A 1.42 10 −9 Pa −1 isothermal<br />

compressibility, KΓ, has been approximated us<strong>in</strong>g:<br />

KΓ = 1<br />

� �<br />

∂ρ<br />

ρ ∂p<br />

Γ<br />

(D.1)<br />

where, K, def<strong>in</strong>es compressibility, ρ, fluid density, p, pressure, <strong>and</strong> subscript<br />

Γ, isothermal conditions. The partial derivation of density with pressure <strong>in</strong><br />

Eq. D.1 has been approximated by first order f<strong>in</strong>ite difference.<br />

The complete Statfjord A off-load<strong>in</strong>g system is made of several sections of<br />

different sizes, lengths <strong>and</strong> materials. Components of the Statfjord A offload<strong>in</strong>g<br />

system are either steel sections or flexible hose sections. Steel has a<br />

Young’s modulus of 200 10 9 Pa. A flexible hose is essentially a rubber hose<br />

re<strong>in</strong>forced with steel elements. Rubber has a Young’s modulus of 0.1 10 9 Pa<br />

(Cheremis<strong>in</strong>off, 1996). However, wired steel elements <strong>in</strong>crease the resistance to<br />

stress (Horn <strong>and</strong> Kuipers, 1988). A 1.8 10 9 Pa Young’s modulus was selected<br />

for the re<strong>in</strong>forced rubber material.<br />

Every change <strong>in</strong> geometry <strong>and</strong>/or material creates an <strong>in</strong>terface between two<br />

media of different acoustic impedances where pressure waves can transmit<br />

<strong>and</strong> reflect. Acoustic impedance is def<strong>in</strong>ed by the product of fluid density<br />

<strong>and</strong> acoustic velocity. Isothermal acoustic velocities can be experimentally<br />

measured or calculated us<strong>in</strong>g (Wylie et al., 1993):<br />

1<br />

a = � �<br />

ρ KΓ + d<br />

�<br />

Y e<br />

(D.2)<br />

where, a, represents isothermal acoustic velocity <strong>in</strong> the fluid-filled pipel<strong>in</strong>e,<br />

d, pipel<strong>in</strong>e diameter, e, pipel<strong>in</strong>e wall thickness, <strong>and</strong>, Y , pipel<strong>in</strong>e material’s<br />

Young modulus. A constant ratio e/d along the pipel<strong>in</strong>e of 0.1 was selected<br />

238


Appendix D<br />

for calculations of acoustic velocities. Calculated isothermal acoustic velocities<br />

<strong>in</strong> steel <strong>and</strong> flexible elements were then 1000 ms −1 <strong>and</strong> 450 ms −1 , respectively.<br />

Isothermal acoustic impedances can now be expressed by (Pa<strong>in</strong>, 1993):<br />

Z = ρa (D.3)<br />

where, Z, represents isothermal acoustic impedance. Calculated isothermal<br />

acoustic impedances <strong>in</strong> steel <strong>and</strong> flexible elements were then 705,000 <strong>and</strong><br />

317,250 <strong>in</strong> steel <strong>and</strong> flexible elements, respectively.<br />

<strong>Pressure</strong> Transients<br />

Failure of the Statford off-load<strong>in</strong>g system was <strong>in</strong>duced by sudden actuation of<br />

the s<strong>in</strong>gle path coupler located on the tanker. Sudden flow stoppage generates<br />

a pressure wave whose amplitude can be calculated by the Joukowsky’s<br />

equation (Wylie et al., 1993):<br />

∆p = ρa∆u (D.4)<br />

where, ∆p, represents pressure wave amplitude, <strong>and</strong>, ∆u, reduction <strong>in</strong> fluid<br />

velocity.<br />

The fluid velocity reduction ∆u depends on the type of valve (Thorley, 1987).<br />

The ratio of fluid velocity reduction <strong>and</strong> <strong>in</strong>itial fluid velocity for a valve closure<br />

is shown with time <strong>in</strong> Figure D.2 for 6 types of valves (Wood <strong>and</strong> Jones, 1973).<br />

First the fluid velocity decreases slowly until 80 % of the <strong>in</strong>itial velocity value,<br />

u0. Then, the fluid velocity decreases rapidly until full closure of the valve.<br />

The valve closure function def<strong>in</strong>es the shape of the front of the pressure wave.<br />

The pressure wave front is shaped ma<strong>in</strong>ly dur<strong>in</strong>g the last 20 % of the valve<br />

closure process, where 80 % of the fluid velocity is lost. Dur<strong>in</strong>g the last 20 % of<br />

the valve closure, the fluid velocity decreases faster for a needle valve than for<br />

a circular gate-valve, as observed <strong>in</strong> Figure D.2. The result<strong>in</strong>g pressure wave<br />

front will then be sharper if <strong>in</strong>duced by a needle valve. Practically, most of<br />

the pressure surge is created dur<strong>in</strong>g the last 20 % of the valve closure action.<br />

The fluid velocity reduction <strong>in</strong> Figure D.2 is theoretical <strong>and</strong> based on flow<br />

area calculation (Wood <strong>and</strong> Jones, 1973). The above result is obta<strong>in</strong>ed for<br />

a uniform valve closure. However the same result is obta<strong>in</strong>ed for a l<strong>in</strong>early<br />

accelerated valve closure, Figure D.3. In case of an accelerated valve closure,<br />

the 80 % fluid velocity reduction takes place <strong>in</strong> a shorter time than a uniform<br />

valve closure; therefore creat<strong>in</strong>g a sharper pressure wave front.<br />

239


Appendix D<br />

u(t) / u 0 [%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Needle<br />

Gate − Circular<br />

Globe<br />

Gate − Square<br />

Ball<br />

Butterfly<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

τ<br />

0.6 0.7 0.8 0.9 1<br />

Figure D.2: Fluid Velocity Response for 6 Different Types of Valves.<br />

Uniform Valve Closure (Wood <strong>and</strong> Jones, 1973)<br />

τ: non-dimensional closure time<br />

u(t) / u 0 [%]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Needle<br />

Gate − Circular<br />

Globe<br />

Gate − Square<br />

Ball<br />

Butterfly<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5<br />

τ<br />

0.6 0.7 0.8 0.9 1<br />

Figure D.3: Fluid Velocity Response for 6 Different Types of Valves.<br />

Accelerated Valve Closure (Wood <strong>and</strong> Jones, 1973)<br />

τ: non-dimensional closure time<br />

240


Appendix D<br />

Calculations of pressure transients must be performed assum<strong>in</strong>g a worst case<br />

scenario. A pressure transient worst case scenario can be obta<strong>in</strong>ed assum<strong>in</strong>g<br />

<strong>in</strong>stantaneous <strong>and</strong> ideal valve actuation. Another way to express this is that<br />

the sharper the pressure wave front, the worse the consequences.<br />

<strong>Pressure</strong> transients propagate <strong>in</strong> fluid-filled pipel<strong>in</strong>es <strong>in</strong> the form of pressure<br />

waves. <strong>Pressure</strong> wave propagation can be modelled based on the onedimensional<br />

mass <strong>and</strong> momentum conservation equations. After order of magnitude<br />

analysis, mass <strong>and</strong> momentum conservation equations can be expressed<br />

by (Wylie et al., 1993):<br />

∂p<br />

∂t + ∂ � ρa2u �<br />

= 0<br />

∂x<br />

(D.5)<br />

∂u<br />

∂t<br />

∂p<br />

+ = s<br />

∂x<br />

where, p, characterizes pressure, u, fluid velocity, t, time, x, axial direction,<br />

<strong>and</strong>, s, source term. The source term accounts for pressure wave attenuation<br />

mechanisms. Only steady state friction is here<strong>in</strong> considered due to the lack<br />

of physical mean<strong>in</strong>g <strong>and</strong> <strong>in</strong>accuracies of unsteady friction models (Pierre <strong>and</strong><br />

Gudmundsson, 2009).<br />

Propagation of a pressure wave <strong>in</strong> a fluid-filled pipel<strong>in</strong>e is def<strong>in</strong>ed as an adiabatic<br />

process. However, propagation of pressure waves <strong>in</strong> liquids can be<br />

approximated by an isothermal process due to the rapidness of the event.<br />

Acoustic velocities used <strong>in</strong> calculations are then isothermal acoustic velocities<br />

<strong>and</strong> can be calculated us<strong>in</strong>g Eq. D.2.<br />

In conservative forms, the hyperbolic system Eq. D.6 can be expressed by<br />

∂Q<br />

∂t<br />

+ ∂F(Q)<br />

∂x<br />

= S (D.6)<br />

consist<strong>in</strong>g of the vector variable, Q, the flux vector, F, <strong>and</strong> the source term<br />

vector, S:<br />

Q =<br />

� p<br />

u<br />

�<br />

F(Q) =<br />

� ρa 2 u<br />

p<br />

�<br />

S =<br />

� 0<br />

s<br />

�<br />

(D.7)<br />

Propagation of pressure transients <strong>in</strong> the Statfjord A off-load<strong>in</strong>g system was<br />

modelled <strong>and</strong> numerically simulated us<strong>in</strong>g a f<strong>in</strong>ite volume method (Pierre <strong>and</strong><br />

Gudmundsson, 2009). The Lax-Wendroff numerical scheme with Superbee<br />

flux limiter function were selected. The numerical scheme is therefore second<br />

241


Appendix D<br />

order accurate <strong>in</strong> both space <strong>and</strong> time, with <strong>in</strong>creased resolution. The Lax-<br />

Wendroff numerical scheme applied to a homogeneous system can be expressed<br />

by (Laney, 1998):<br />

Ω n+1<br />

i<br />

ˆΦ n<br />

i+ 1<br />

2<br />

a n<br />

i+ 1<br />

2<br />

= Ωn i − ∆t<br />

�<br />

ˆΦ n<br />

i+ 1 −<br />

2<br />

ˆ Φ n<br />

i− 1<br />

�<br />

2<br />

∆x<br />

= 1 � n<br />

Φ(Ω<br />

2<br />

i+1) + Φ(Ω n i ) � − ∆t<br />

2∆x an<br />

i+ 1<br />

� n<br />

Φ(Ωi+1) − Φ(Ω<br />

2<br />

n i ) �<br />

⎧<br />

⎪⎨ Φ(Ω<br />

=<br />

⎪⎩<br />

n i+1 ) + Φ(Ωni )<br />

Ωn i+1 − Ωn for Ω<br />

i<br />

n i �= Ωn i+1<br />

Φ(Ω n i ) for Ωn i = Ωn i+1<br />

(D.8)<br />

The Lax-Wendroff numerical scheme coupled to a flux limiter function is expressed<br />

by (Laney, 1998)<br />

Ω n+1<br />

i<br />

= Ωni − ∆t<br />

∆x ˆ �<br />

φ(θ) ˆF n<br />

i+ 1 −<br />

2<br />

ˆ F n<br />

i− 1<br />

�<br />

2<br />

The Superbee flux limiter function takes the form (Laney, 1998):<br />

(D.9)<br />

ˆφ(θ) = max (0, m<strong>in</strong>(1, 2θ), m<strong>in</strong>(2, θ)) (D.10)<br />

where the variable θ is def<strong>in</strong>ed by:<br />

Cr > 0<br />

θ n i−1/2 = Qn i−1 − Qn i−2<br />

Q n i − Qn i−1<br />

Cr < 0<br />

θ n i−1/2 = Qn i+1 − Qn i<br />

Q n i − Qn i−1<br />

(D.11)<br />

The Lax-Wendroff numerical scheme is explicit, centred, second order accurate<br />

<strong>in</strong> both time <strong>and</strong> space, l<strong>in</strong>early stable provided that the Courant number is<br />

less than one.<br />

Cases<br />

Three cases were modelled to assess the importance of the sequence of offload<strong>in</strong>g<br />

system components on rapid pressure transient events. The first case<br />

describes the off-load<strong>in</strong>g system without consider<strong>in</strong>g steel components at each<br />

242


Appendix D<br />

end of a flexible hose. The second case describes the off-load<strong>in</strong>g system tak<strong>in</strong>g<br />

<strong>in</strong>to consideration steel components at each end of a flexible hose. The third<br />

case considers flexible hoses twice their orig<strong>in</strong>al length with steel components<br />

at each end. The three cases represent a similar system with the same length.<br />

Differences lay <strong>in</strong> the sequence of sections.<br />

The three cases are illustrated <strong>in</strong> Figure D.8, <strong>and</strong> the correspond<strong>in</strong>g lengths of<br />

every element are reported <strong>in</strong> Tables D.2, D.3 <strong>and</strong> D.4. We observe that case 2<br />

corresponds to the actual geometry of the Statfjord A off-load<strong>in</strong>g system while<br />

cases 1 <strong>and</strong> 3 are used for analysis. Purpose of cases 1 <strong>and</strong> 3 is to evaluate<br />

the importance of the flexible sections lengths <strong>in</strong> the sequence of the whole<br />

off-load<strong>in</strong>g pipel<strong>in</strong>e. Distances are measured from the storage cell. The 2,808<br />

metres distance corresponds to the s<strong>in</strong>gle path coupler.<br />

In the three cases, the 2,808 metres long off-load<strong>in</strong>g system is modelled us<strong>in</strong>g<br />

22,463 grid cells; that is, with a 12.5 centimetres spatial resolution. The<br />

pressure wave propagation is simulated over 3 seconds with 28,800 steps; that<br />

is, with a 0.1 millisecond temporal resolution. The 3 seconds simulated time<br />

corresponds approximately to the travel time of the pressure wave propagat<strong>in</strong>g<br />

from the s<strong>in</strong>gle path coupler to the storage cells. Dur<strong>in</strong>g the 3 seconds, load<strong>in</strong>g<br />

pumps effects are assumed negligible.<br />

The emergency shut-down valve connected to the s<strong>in</strong>gle path coupler was suddenly<br />

activated on December 12, 2007, generat<strong>in</strong>g a rapid pressure transient.<br />

Before sudden valve actuation, the oil flow rate <strong>in</strong> the off-load<strong>in</strong>g system was<br />

6,000 m 3 hr −1 . The correspond<strong>in</strong>g fluid velocity <strong>in</strong> a 20 <strong>in</strong>ches pipel<strong>in</strong>e is 8.2<br />

ms −1 . The flow is fully turbulent with a 4.62 10 6 Reynolds number. A 0.02<br />

steady-state friction factor is <strong>in</strong>cluded as a source term <strong>in</strong> (D.6). The pressure<br />

evolution <strong>in</strong> time is computed at every grid cell of the model dur<strong>in</strong>g 3 seconds.<br />

Every time step is iterated such that the Courant number is 0.9.<br />

The emergency shut-down valve was assumed to close <strong>in</strong>stantaneously, therefore<br />

creat<strong>in</strong>g a sharp pressure wave front. Such assumption corresponds to<br />

the worst case of valve closure. <strong>Pressure</strong> surge predictions aimed at <strong>in</strong>creas<strong>in</strong>g<br />

safety <strong>and</strong> prevent damages to the fluid transport pipel<strong>in</strong>e. Therefore, worst<br />

case modell<strong>in</strong>g was to be used.<br />

Results<br />

The three cases were solved numerically with the Lax-Wendroff numerical<br />

scheme, <strong>in</strong>clud<strong>in</strong>g the Superbee flux limiter. The three cases were modelled<br />

with three sets of acoustic velocities. The first set of acoustic velocities <strong>in</strong> the<br />

steel sections <strong>and</strong> flexible sections was a 1000 ms −1 <strong>and</strong> 450 ms −1 , respectively.<br />

The second set was 1000 ms −1 <strong>and</strong> 350 ms −1 , respectively. F<strong>in</strong>ally<br />

243


Appendix D<br />

<strong>Pressure</strong> [bara]<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 500 1000 1500<br />

Distance [m]<br />

2000<br />

Case 1<br />

Case 2<br />

Case 3<br />

2500<br />

Figure D.4: Maximum <strong>Pressure</strong> Achieved <strong>in</strong> the Oil Off-Load<strong>in</strong>g<br />

System <strong>in</strong> 3 seconds with asteel = 1000 ms <strong>and</strong><br />

aflexible = 450 ms<br />

the third set was 1100 ms −1 <strong>and</strong> 450 ms −1 , respectively.<br />

The first set of acoustic velocities corresponds to the st<strong>and</strong>ard speeds of sound<br />

described <strong>and</strong> calculated previously. The second <strong>and</strong> third set of acoustic<br />

velocities were used for analysis. Purpose of acoustic velocities sets 2 <strong>and</strong> 3 is<br />

to evaluate the importance of the acoustic velocities on the maximum pressure<br />

reached. For all three sets of acoustic velocities, all three cases were modelled<br />

<strong>and</strong> results are shown <strong>in</strong> Figures D.4, D.5, D.6 <strong>and</strong> D.7.<br />

The pressure wave generated at the s<strong>in</strong>gle path coupler propagates towards<br />

the storage cells. Interfaces between pipel<strong>in</strong>e elements with different acoustic<br />

impedances <strong>in</strong>duce multiple pressure wave reflections along the oil off-load<strong>in</strong>g<br />

system. Superpositions of pressure waves reflections generate pressure peaks.<br />

Locations of pressure peaks depend on the off-load<strong>in</strong>g system geometry <strong>and</strong><br />

on the acoustic velocities <strong>in</strong> every element.<br />

Maximum pressures reached along the off-load<strong>in</strong>g pipel<strong>in</strong>e for the first set<br />

of acoustic velocities, dur<strong>in</strong>g the first three seconds after valve closure, are<br />

shown <strong>in</strong> Figure D.4 for the three cases. The maximum pressure reached is<br />

the greatest for case 2. The maximum pressure reached is the lowest for case<br />

1. Another way to express the result is that the more numerous changes <strong>in</strong><br />

244


Cumulated Time [s]<br />

x 10−5<br />

1.8<br />

1.6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Case 1<br />

Case 2<br />

Case 3<br />

0<br />

0 500 1000 1500<br />

Distance [m]<br />

2000 2500<br />

Figure D.5: Cumulated Time of the <strong>Pressure</strong> Surge be<strong>in</strong>g over 50<br />

barg<br />

Appendix D<br />

acoustic impedances along the transport pipel<strong>in</strong>e, the greater the maximum<br />

pressure reached.<br />

The maximum pressure reached <strong>in</strong>creases with the number of changes <strong>in</strong> acoustic<br />

impedance. Interfaces between media of different acoustic impedances generate<br />

pressure wave reflections that travel separately of the orig<strong>in</strong>al pressure<br />

wave. The superposition of two or more pressure waves generate pressure<br />

peaks that are responsible for high maximum pressure reached along the offload<strong>in</strong>g<br />

system for case 2.<br />

Calculated pressures are gauge pressures. Another way to express this is that<br />

the calculated pressures are on top of the <strong>in</strong>-situ static pressure. The maximum<br />

calculated pressure reached <strong>in</strong> the oil off-load<strong>in</strong>g system for case 2 is 58<br />

barg <strong>and</strong> takes place approximately 80 m from the s<strong>in</strong>gle path coupler. The<br />

maximum calculated pressures are always less than 60 barg for all three cases.<br />

<strong>Pressure</strong> wave superpositions are ma<strong>in</strong>ly responsible for the high maximum<br />

pressure values. It means that maximum pressure values are due to pressure<br />

peaks short <strong>in</strong> time. Repetitive pressure peaks <strong>in</strong>duce fatigue <strong>in</strong> pipel<strong>in</strong>e elements<br />

<strong>and</strong> decrease the resistance to burst<strong>in</strong>g. The cumulated time of pressure<br />

surge be<strong>in</strong>g over 50 barg is shown <strong>in</strong> Figure D.5. The pressure surge is greater<br />

than 50 barg dur<strong>in</strong>g 1.62 s at 80 m from the s<strong>in</strong>gle path coupler.<br />

245


Appendix D<br />

<strong>Pressure</strong> [bara]<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 500 1000 1500<br />

Distance [m]<br />

2000<br />

Case 1<br />

Case 2<br />

Case 3<br />

2500<br />

Figure D.6: Maximum <strong>Pressure</strong> Achieved <strong>in</strong> the Oil Off-Load<strong>in</strong>g<br />

System <strong>in</strong> 3 s with asteel = 1000 ms <strong>and</strong><br />

aflexible = 350 ms<br />

Flexible hoses <strong>and</strong> steel elements <strong>in</strong> the Statfjord A off-load<strong>in</strong>g system have<br />

a 95 bara resistance to burst<strong>in</strong>g (PSA, 2008); that is almost 40 barg over the<br />

calculated maximum pressure of 58 barg at 80 m from the s<strong>in</strong>gle path coupler.<br />

However, the calculated maximum pressure do not consider the <strong>in</strong>-situ static<br />

pressure. Besides, the repetitive pressure peaks mentioned previously decrease<br />

the resistance to burst<strong>in</strong>g of the structure.<br />

The calculated maximum pressure was located 80 m from the s<strong>in</strong>gle path coupler.<br />

However, the predicted location <strong>and</strong> amplitude depend on the acoustic<br />

velocity <strong>in</strong> the flexible hose elements of the off-load<strong>in</strong>g system. The same three<br />

cases were then modelled with the second <strong>and</strong> third sets of acoustic velocities.<br />

Case 2 was the worst case for both sets of acoustic velocities as shown <strong>in</strong><br />

Figures D.6 <strong>and</strong> D.7. For the second set of acoustic velocities, the maximum<br />

pressure reached was 55 barg <strong>and</strong> was located 10 m from the s<strong>in</strong>gle path<br />

coupler. For the third set of acoustic velocities, the maximum pressure reached<br />

is 65 barg <strong>and</strong> is located 170 m from the s<strong>in</strong>gle path coupler. For both cases,<br />

the maximum pressure reported was obta<strong>in</strong>ed <strong>in</strong> case 2.<br />

For all three sets of acoustic velocities, case 2 was always the most severe case<br />

<strong>and</strong> case 1 always the least severe case. The first set of acoustic velocities that<br />

246


<strong>Pressure</strong> [bara]<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Case 1<br />

Case 2<br />

Case 3<br />

0<br />

0 500 1000 1500<br />

Distance [m]<br />

2000 2500<br />

Figure D.7: Maximum <strong>Pressure</strong> Achieved <strong>in</strong> the Oil Off-Load<strong>in</strong>g<br />

System <strong>in</strong> 3 seconds with asteel = 1100 ms <strong>and</strong><br />

aflexible = 450 ms<br />

Appendix D<br />

correspond to the calculated speed of sounds, generated the greater maximum<br />

pressure. The Statfjord A off-load<strong>in</strong>g system actually ruptured 37 m from the<br />

s<strong>in</strong>gle path coupler.<br />

Differences between predicted maximum pressure values reached <strong>and</strong> values<br />

reported <strong>in</strong> (PSA, 2008) are due to uncerta<strong>in</strong>ties. Density <strong>and</strong> compressibility<br />

values were obta<strong>in</strong>ed assum<strong>in</strong>g an hydrocarbon composition. Similarly, the<br />

off-load<strong>in</strong>g system geometry <strong>and</strong> material properties were assumed. Thus,<br />

calculated acoustic impedances are the ma<strong>in</strong> uncerta<strong>in</strong>ties.<br />

Acoustic impedances impact both the amplitude <strong>and</strong> location of the maximum<br />

pressure reachable. Therefore, accurate descriptions of the hydrocarbon<br />

composition <strong>and</strong> state, as well as an accurate description of the off-load<strong>in</strong>g<br />

geometry <strong>and</strong> material properties would help decreas<strong>in</strong>g the uncerta<strong>in</strong>ties.<br />

Conclusions<br />

The sequence of pipe sections has been analyzed based on the exist<strong>in</strong>g Statfjord<br />

A off-load<strong>in</strong>g system. Rupture of the Statfjord A off-load<strong>in</strong>g system<br />

due to rapid pressure transient has been responsible for the second largest oil<br />

spillage on the Norwegian cont<strong>in</strong>ental shelf. Three different sequences have<br />

247


Appendix D<br />

been modelled for assess<strong>in</strong>g the importance of the sections sequence <strong>and</strong> the<br />

lengths of the elements.<br />

Acoustic <strong>in</strong>terfaces are responsible for pressure peaks due to the superposition<br />

of pressure wave reflections. The more sections <strong>and</strong> components <strong>in</strong> the sequence,<br />

the greater the maximum pressure reached. The shorter the elements,<br />

the greater the maximum pressure reached.<br />

Flexible hoses low acoustic impedances decrease the maximum pressure reached;<br />

while rigid sections great acoustic impedances <strong>in</strong>crease the maximum pressure<br />

reached. Another way to express these results is that the maximum pressure<br />

achieved decreases with the stiffness of the sections mak<strong>in</strong>g the off-load<strong>in</strong>g<br />

system sequence.<br />

Quick-act<strong>in</strong>g valve closure functions also <strong>in</strong>fluence the maximum pressure<br />

reached. The greater the fluid velocity gradient dur<strong>in</strong>g the last 20 % of the closure<br />

process, the sharper the pressure wave front created. Practically, waterhammer<br />

surges are generated mostly <strong>in</strong> the last 20 % of the valve closure<br />

process.<br />

Acknowledgement<br />

The authors want to thank the Norwegian Research Council for provid<strong>in</strong>g f<strong>in</strong>ancial<br />

support to Ph.D. student Benjam<strong>in</strong> Pierre under contract 196054/S60.<br />

References<br />

Cheremis<strong>in</strong>off N.P. (1996): Materials Selection Deskbook, Noyes Publications.<br />

Ghidaoui M.S., Zhao M., McInnis D.A. <strong>and</strong> Axworthy D.H. (2005): A Review<br />

of Water Hammer Theory <strong>and</strong> Practice, ASME Applied Mechanics Reviews,<br />

vol. 58(1), pp 49-76.<br />

Horn B.A. <strong>and</strong> Kuipers M. (1988): Strength <strong>and</strong> Stiffness of a Re<strong>in</strong>forced<br />

Flexible Hose, Flow, Turbulence <strong>and</strong> Combustion, vol. 45(3), pp 251-281.<br />

Johnsen E.E. (2008): Private Communication.<br />

Laney C.B. (1998): Computational Gasdynamics, Cambridge University Press.<br />

LeVeque R.J. (2002): F<strong>in</strong>ite Volume Methods for Hyperbolic Problems, Cambridge<br />

University Press.<br />

248


Appendix D<br />

Misiunas D., Vitkovsky J.P., Olsson G., Simpson A. <strong>and</strong> Lambert M.F. (2005):<br />

Pipel<strong>in</strong>e Break Detection Us<strong>in</strong>g <strong>Pressure</strong> Transient Monitor<strong>in</strong>g, Journal of Water<br />

Resources <strong>and</strong> Plann<strong>in</strong>g Management, vol. 131(4), pp 316-321.<br />

Moura C.H.W., Sampaio D.S., de Lacerda I.M. <strong>and</strong> Selli M.F. (2004): Monitor<strong>in</strong>g<br />

Leakages on Oil Production Offload<strong>in</strong>g at Open Seas Us<strong>in</strong>g Statistics Associated<br />

With Mass Balance Methods, ASME Conference Proceed<strong>in</strong>gs 41766,<br />

pp 2243-2247.<br />

Norwegian Petroleum Directorate NPD (2008): Facts, The Norwegian Petroleum<br />

Sector.<br />

Pa<strong>in</strong> H.J. (1993): The Physics of Vibrations <strong>and</strong> <strong>Waves</strong>, Jonn Wiley & Sons.<br />

Petroleum Safety Authority PSA (2008): Oljeutslipp Statfjord OLS-A 12.12.2007.<br />

Pierre B. <strong>and</strong> Gudmundsson J.S. (2008): Water Hammer Simulation of <strong>Pressure</strong><br />

<strong>in</strong> the Failed Flexible Hose for Off-Load<strong>in</strong>g Oil at Statfjord A, Department<br />

of Petroleum Eng<strong>in</strong>eer<strong>in</strong>g <strong>and</strong> Applied Geophysics, <strong>NTNU</strong>, 7491 Trondheim.<br />

Pierre B. <strong>and</strong> Gudmundsson J.S. (2009): Modell<strong>in</strong>g of <strong>Pressure</strong> Wave Propagation<br />

<strong>in</strong> Pipel<strong>in</strong>es: Steady-State <strong>and</strong> Transient Friction, Fifth National Conference<br />

on Computational Mechanics, Trondheim 26-27 May.<br />

Sangesl<strong>and</strong> S. (2008): Private Communication.<br />

Thorley A.R.D. (1987): Check Valve Behaviour Under Transient Flow Conditions.<br />

A State of the Art Review , Fluid Transients <strong>in</strong> Fluid-Structure<br />

Interaction (ASME).<br />

Wang L., Li J., Peng K., J<strong>in</strong> S. <strong>and</strong> Li Z. (2004): Petroleum Pipe Leakage<br />

Detection <strong>and</strong> Location Embeded <strong>in</strong> S.C.A.D.A., ASME Conference Proceed<strong>in</strong>gs<br />

41766, pp 2249-2253.<br />

Wang X.J., Lambert M.F., Ligget J.A. <strong>and</strong> Vitkovsky J.P. (2002): Leak Detection<br />

<strong>in</strong> Pipel<strong>in</strong>es Us<strong>in</strong>g the Damp<strong>in</strong>g of Fluid Transients, Journal of Hydraulic<br />

Eng<strong>in</strong>eer<strong>in</strong>g, vol. 128(7), pp 697-711.<br />

Wood D.J. <strong>and</strong> Jones S.E. (1973): Water-Hammer Charts for Various Types<br />

of Valves, Journal of Hydraulic Division, vol. 99(1), pp 167-178.<br />

Wylie E.B., Streeter V.L <strong>and</strong> Suo L. (1993): Fluid Transients <strong>in</strong> Systems,<br />

249


Appendix D<br />

Prentice Hall.<br />

Nomenclature<br />

a Acoustic Velocity [ms −1 ]<br />

Cr Courant Number<br />

D Diameter [m]<br />

e Pipe Wall Thickness [m]<br />

E Material Young Modulus [Pa]<br />

F Flux Vector<br />

KT Isothermal Compressibility [Pa −1 ]<br />

p <strong>Pressure</strong> [Pa]<br />

Φ Flux Limiter<br />

Q Vector Variable<br />

ρ Density [kg m −3 ]<br />

S Source Term<br />

u Velocity [ms −1 ]<br />

z Acoustic Impedance [kg m −2 s −1 ]<br />

Appendix: Cases Geometry<br />

Figure D.8: Illustration of the Three Cases Considered for the<br />

Modell<strong>in</strong>g<br />

250


Table D.2: Geometry of Case 1<br />

Appendix D<br />

Element Length [m] Property Element Length [m] Property<br />

0 0.75 Rigid 9 11.25 Flexible<br />

1 10.00 Flexible 10 1.00 Rigid<br />

2 11.25 Flexible 11 80.00 Flexible<br />

3 11.25 Flexible 12 10.00 Flexible<br />

4 11.25 Flexible 13 1.00 Rigid<br />

5 2.00 Rigid 14 80.00 Flexible<br />

6 11.25 Flexible 15 28.25 Flexible<br />

7 11.25 Flexible 16 2516.00 Rigid<br />

8 11.25 Flexible<br />

Table D.3: Geometry of Case 2<br />

Element Length [m] Property Element Length [m] Property<br />

0 0.75 Rigid 20 0.75 Rigid<br />

1 0.75 Rigid 21 9.75 Flexible<br />

2 8.50 Flexible 22 0.75 Rigid<br />

3 0.75 Rigid 23 0.75 Flexible<br />

4 0.75 Rigid 24 9.75 Flexible<br />

5 9.75 Flexible 25 0.75 Rigid<br />

6 0.75 Rigid 26 1.00 Rigid<br />

7 0.75 Rigid 27 0.75 Rigid<br />

8 9.75 Flexible 28 78.50 Flexible<br />

9 0.75 Rigid 29 0.75 Rigid<br />

10 0.75 Rigid 30 10.00 Rigid<br />

11 9.75 Flexible 31 1.00 Rigid<br />

12 0.75 Rigid 32 0.75 Rigid<br />

13 2.00 Rigid 33 78.50 Flexible<br />

14 0.75 Rigid 34 0.75 Rigid<br />

15 9.75 Flexible 35 0.75 Rigid<br />

16 0.75 Rigid 36 26.75 Flexible<br />

17 0.75 Rigid 37 0.75 Rigid<br />

18 9.75 Flexible 38 2516.00 Rigid<br />

19 0.75 Rigid<br />

251


Appendix D<br />

Table D.4: Geometry of Case 3<br />

Element Length [m] Property Element Length [m] Property<br />

0 0.75 Rigid 16 11.25 Flexible<br />

1 0.75 Rigid 17 0.75 Rigid<br />

2 10.00 Flexible 18 1.00 Rigid<br />

3 9.75 Flexible 19 0.75 Rigid<br />

4 0.75 Rigid 20 78.50 Flexible<br />

5 0.75 Rigid 21 0.75 Rigid<br />

6 11.25 Flexible 22 10.00 Rigid<br />

7 9.75 Flexible 23 1.00 Rigid<br />

8 0.75 Rigid 24 0.75 Rigid<br />

9 2.00 Rigid 25 78.50 Flexible<br />

10 0.75 Rigid 26 0.75 Rigid<br />

11 9.75 Flexible 27 0.75 Rigid<br />

12 11.25 Flexible 28 26.75 Flexible<br />

13 0.75 Rigid 29 0.75 Rigid<br />

14 0.75 Rigid 30 2516.00 Rigid<br />

15 9.75 Flexible<br />

252

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!