123
University of Calgary PRISM: University of Calgary's Digital Repository Graduate Studies The Vault: Electronic Theses and Dissertations 2016 Thermodynamic Formation Conditions for Propane Hydrates in Equilibrium with Liquid Water Adeniyi, Kayode Israel Adeniyi, K. I. (2016). Thermodynamic Formation Conditions for Propane Hydrates in Equilibrium with Liquid Water (Unpublished master's thesis). University of Calgary, Calgary, AB. doi:10.11575/PRISM/28310 http://hdl.handle.net/11023/3244 master thesis University of Calgary graduate students retain copyright ownership and moral rights for their thesis. You may use this material in any way that is permitted by the Copyright Act or through licensing that has been assigned to the document. For uses that are not allowable under copyright legislation or licensing, you are required to seek permission. Downloaded from PRISM: https://prism.ucalgary.ca

Thermodynamic Formation Conditions for Propane Hydrates in

  • Upload
    others

  • View
    11

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Thermodynamic Formation Conditions for Propane Hydrates in

University of Calgary

PRISM: University of Calgary's Digital Repository

Graduate Studies The Vault: Electronic Theses and Dissertations

2016

Thermodynamic Formation Conditions for Propane

Hydrates in Equilibrium with Liquid Water

Adeniyi, Kayode Israel

Adeniyi, K. I. (2016). Thermodynamic Formation Conditions for Propane Hydrates in Equilibrium

with Liquid Water (Unpublished master's thesis). University of Calgary, Calgary, AB.

doi:10.11575/PRISM/28310

http://hdl.handle.net/11023/3244

master thesis

University of Calgary graduate students retain copyright ownership and moral rights for their

thesis. You may use this material in any way that is permitted by the Copyright Act or through

licensing that has been assigned to the document. For uses that are not allowable under

copyright legislation or licensing, you are required to seek permission.

Downloaded from PRISM: https://prism.ucalgary.ca

Page 2: Thermodynamic Formation Conditions for Propane Hydrates in

UNIVERSITY OF CALGARY

Thermodynamic Formation Conditions for Propane Hydrates in Equilibrium with Liquid Water

by

Kayode Israel Adeniyi

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF SCIENCE

GRADUATE PROGRAM IN CHEMISTRY

CALGARY, ALBERTA

AUGUST, 2016

© Kayode Israel Adeniyi 2016

Page 3: Thermodynamic Formation Conditions for Propane Hydrates in

ii

Abstract

Accurate knowledge for hydrate formation conditions of pure propane in the presence of liquid

water is important to avoid flow assurance issues in processing, storage and transportation of

liquefied petroleum gas, as well as modeling hydrate based separation processes involving type

II hydrates. Experimental dissociation conditions were measured using the phase boundary

dissociation method, which has the advantages of reduced experimental time and generation of

more equilibrium dissociation data when compared to other hydrate measurement techniques. In

this study, phase equilibra measurements are reported for 99.5 mol % and 99.999 mol % propane

at the phase boundary for the liquid water (Lw)-hydrate (H)-propane vapour (C3H8 (g)) and Lw -

H-liquid propane (C3H8 (l)) loci. The results were modeled using van der Waal and Platteeuw

model for the hydrate phase and reduced Helmholtz energy equation for the fluid phases. Results

are compared with highly variant literature data, where large deviations observed for the Lw-H-

C3H8 (l) phase boundary can be mainly attributed to purity and experimental techniques used in

the literature.

Page 4: Thermodynamic Formation Conditions for Propane Hydrates in

iii

Acknowledgements

I am deeply grateful to my Supervisor, Dr. Robert Marriott for the opportunity to work on this

project, patience, support, understanding and supervision of this thesis. Much of my

understanding on phase behavior of hydrate formers is credited to his mentorship.

Thank you to Dr. Viola Birss and Dr. Peter Kusalik for agreeing to serve on my committee and

their support.

I acknowledge the financial support of University of Calgary, Dr Marriott’s NSERC ASRL

Industrial Research Chair in Applied Sulfur Chemistry, Department of Chemistry Graduate

Student Award (2014, 2015 and 2016) and Faculty of Graduate Studies Travel Grant (2015 and

2016).

My sincere appreciation goes to Alberta Sulphur Research Ltd. (ASRL) staff for providing a

friendly working atmosphere. I am grateful to Connor Deering for his help with modeling,

experimental expertise, reviewing my writing and more importantly, his friendship. Thanks to

Zachary Ward for always answering my unending questions about the experimental setup. I also

wish to thank the postdoctoral researchers, Dr. Payman Pirzadeh and Dr. Fadi Alkhateeb for their

words of encouragement and constructive criticism of my writings.

Page 5: Thermodynamic Formation Conditions for Propane Hydrates in

iv

My appreciations also go to the graduate coordinator, Janice Crawford, Patricia Alegre of ASRL

and other administrative staff of the department for their advice and guidance on university

regulations. I would also like to thank my colleagues in the research group for their support.

Omowumi, thank you for your understanding, support, sacrifices and giving me the most

valuable gift of life that I cherished. To my beloved parents, I say thank you for your support and

guidance through the years.

Page 6: Thermodynamic Formation Conditions for Propane Hydrates in

v

Dedication

This thesis is dedicated to God Almighty, the GREAT architect of life and all positive

inspiration.

Page 7: Thermodynamic Formation Conditions for Propane Hydrates in

vi

Table of Contents

Abstract ........................................................................................................................................ ii

Acknowledgements ..................................................................................................................... iii

Dedication.......................................................................................................................................v

Table of Contents ........................................................................................................................vi

List of Tables ................................................................................................................................ix

List of Figures ............................................................................................................................. xi

List of Symbols, Abbreviations and Nomenclature................................................................xiii

Chapter One: Introduction………………………………………………………………….…..1

1.1 Outline ………………………………………………………………………………….……1

1.2 Motivation for study…………………………………………………………………….…...1

1.3 History of gas clathrate hydrates………………………………………………………….....3

1.4 Structure and formation of clathrate gas hydrate…………………………………………....4

1.4.1 Structure I…………………………………………………………………….…….....7

1.4.2 Structure II ………………………………………………………………………........7

1.4.3 Structure H…………………………………………………………..…………….…..7

1.5 Applications of gas hydrate………………………………………………………………….8

1.5.1 Hydrogen (H2) storage………………………………………..…………………….....8

1.5.2 Separation processes…………………………………………………………………..9

1.5.3 Desalination of seawater……………………………………………………………..10

1.5.4 Potential source of energy……………………………………………………………10

1.5.5 Natural gas hydrate in flow assurance………………………………………………..11

1.5.5.1 Gas hydrate prevention and control……………………………..…………...…..12

1.6 Importance of propane hydrate formation conditions studies………………………..........12

1.7 Phase behavior and avoiding hydrate formation………….……………………………….13

Page 8: Thermodynamic Formation Conditions for Propane Hydrates in

vii

1.7.1 Phase behavior of a hydrocarbon hydrate former………….…………..……….......14

1.7.2 Phase behavior of C3H8 + H2O system…………………………………..……........15

1.7.3 Semi empirical model for hydrate dissociation correlation……….……….............17

1.8 Experimental dissociation data for C3H8 hydrate……………………………………..…...17

1.9 Review of gas hydrate thermodynamic models………………………………………...….20

Chapter two: Review of Literature Techniques, Experimental Procedure

and Calibration……………………………………………………………………………...…24

2.1 Outline………………………………………………………………………………..........24

2.2 Methods of studying gas hydrate equilibra…………………………………...………...….24

2.2.1 Dynamic method………………………………………...………………...……......24

2.2.2 Static method…………………………………………………………..……......….25

2.2.2.1 Isothermal method…………………………………………..………...…..25

2.2.2.2 I sobaric method……………………………………………..………...…..26

2.2.2.3 Isochoric method………………………………………………...…....…..26

2.2.2.4 Phase boundary dissociation method………………………………...…..27

2.3 Review of selected experimental apparatus for hydrate dissociation studies……………...29

2.4 Apparatus used for this study………………………………………………………….......32

2.4.1 Interfacing experimental setup with Laboratory Virtual Instruments

Engineering Workbench (LABVIEW) for data acquisition…………………..…....34

2.5 Materials…………………………………………………………………………………...36

2.6 Experimental procedure……………………………………………………………….......36

Chapter three: Experimental Results and Modelling…………………………………….....38

3.1 Outline……………………………………………………………………………………..38

3.2 Thermodynamic modeling………………………………………………………………....38

3.2.1 Fluid phase modeling……………………………………………………………….....39

3.2.2 Description of the hydrate phase……………………………………………….….…..46

3.2.2.1 The Vander Waal and Platteeuw hydrate model…………………………….…....46

3.2.2.2 Calculation of hydrate phase fugacity……………………………………….…...48

Page 9: Thermodynamic Formation Conditions for Propane Hydrates in

viii

3.2.2.3 Hydrate cage occupancy……………………………………………………….......51

3.2.2.3.1 Optimization of Kihara potential paramaters……………………………........53

3.3 Algorithm for calculating equilibrium hydrate formation condition………………………..54

3.4 Experimental results and discussion…………………………………………………...........56

3.4.1 Model comparisons to experimental and literature data along the Lw-H-C3H8(g)…......61

3.4.2 Model comparisons to experimental and literature data along the Lw-H-C3H8(l)….......67

3.4.3 Comparisons of upper quadruple points of this study and literature………………........72

Chapter four: Conclusion, recommendation and future work……………………………....74

4.1 Conclusion……………………………………………………………………………..…....74

4.2 Recommendations……………………………………………………………………….….75

4.3 Future work………………………………………………………………………………....75

Appendix A: Calibrations and results…………………..………………………………………77

A.1.1 Pressure calibration………………………………………………………………………………77

A.1.1.1 Primary transducer calibration through the use of Deadweight Testers……………..77

A.1.1.2 Secondary transducer calibration………………………………………………….……….78

A.1.1.3 Results and discussions……………………………………………………………………….79

A.1.2 Temperature calibration………………………………………………………………………….81

A.1.2.1 The International Temperature Scale……………………………………………………….81

A.1.2.2 Calibration procedure………………………………………………………………….........82

A.1.2.3 Result and discussion……………………………………………………………………….…83

A.1.3 Volume calibration…………………………………………………………………….…85

Appendix B: Pressure versus temperature plots of the experimental run for the dissociation

points along Lw-H-C3H8(g) and Lw-H-C3H8(l) phase boundaries reported in this study………86

Appendix C: Parameters and coefficients used in the reduced energy Helmholtz EOS for

calculation of thermodynamic properties of C3H8 in equation 3.13……………………….…....91

Appendix D: First derivative of and the reducing function r and rT with respect to in …...92

Appendix E: Copyright Permissions……………………………………………………………93

Page 10: Thermodynamic Formation Conditions for Propane Hydrates in

ix

References……………………………………………………………………………………....95

Page 11: Thermodynamic Formation Conditions for Propane Hydrates in

x

List of Tables

Table 1.1. Comparison of structure I, structure II and structure H hydrates…………………….6

Table 1.2. Summary of experimental dissociation conditions along the Lw–H–C3H8 (g) phase

boundary…………………………………………………………………………………………18

Table 1.3. Summary of the experimental dissociation conditions along the Lw–H–C3H8 (l) phase

boundary…………………………………………………………………………………………19

Table 1.4. Summary of C3H8 hydrate upper quadruple points reported in literature…………...20

Table 2.1. Methods used for the study of C3H8 hydrate dissociation conditions….……………32

Table 2.2. Measured gas impurities (in moles) in C3H8 used for this work……….……………36

Table 3.1. Binary parameters of the reducing functions for density and temperature used in

equations 3.18 and 3.19………………………………………………………………………….44

Table 3.2. Thermodynamic reference properties for structure II used in this study…………….50

Table 3.3. Optimised Kihara potential paramaters used for this study………………………….54

Table 3.4. Experimental dissociation conditions for C3H8 hydrates along the Lw–H–C3H8(g)

phase boundary…………………………………………………………………………………..58

Table 3.5. Experimental dissociation conditions for 99.999 mol % C3H8 hydrates along the Lw–

H–C3H8(l) phase boundary………………………………………………………….…………...59

Table 3.6. Model comparison to the experimental data along the Lw–H–C3H8(g) locus……....61

Table 3.7. Summary of literature data along the Lw-H-C3H8(g) phase boundary compared.…..63

Table 3.8. Model comparison to the experimental data along the Lw–H–C3H8(l) phase

boundary…………………………………………………………………………………………68

Table 3.9. Summary of literature data and corresponding purities along the Lw–H–C3H8(l) phase

boundary…………………………………………………………………………………………70

Table 3.10. Quadruple points conditions from this study and literature….…………………...73

Table A.1. Comparison of the pressures measured by the calibrated primary Paroscientific

Transducer, calp , and the uncalibrated Paroscientific Pressure Transducer, measp ………….…80

Table A.2. Comparison of the pressures measured by the calibrated primary Paroscientific

Transducer, pcal, and the uncalibrated Keller Pressure Transducer, measp ……………………...80

Table A.3. The experimentally measured melting points of H2O with the corresponding

deviations………………………………………………………………………………………..84

Page 12: Thermodynamic Formation Conditions for Propane Hydrates in

xi

Table A.4. Comparison of the measured temperatures from the calibrated, Tcal, and uncalibrated

PRT, Tmeas (used inside the water bath)……………………………………...…………….…….84

Page 13: Thermodynamic Formation Conditions for Propane Hydrates in

xii

List of Figures

Figure 1.1. Classification of clathrate gas hydrates……………………………………………..5

Figure 1.2. Typical pressure − temperature diagram for a hydrocarbon hydrate former………15

Figure 1.3. Pressure − temperature diagram of propane + water system……………………….16

Figure 2.1. A Typical pressure-temperature curve of a gas hydrate formation and dissociation

using the isochoric method………………………………………………………………………27

Figure 2.2. Experimental schematic of Deaton and Frost apparatus for phase equilibra studies of

gas hydrates……………………………………………………………………………………...30

Figure 2.3. Diagram of the dynamic apparatus designed and constructed by Hammerschdmit for

studying hydrate formation conditions………………………………………………………….31

Figure 2.4. Schematic diagram of the setup used for the measurement of hydrate dissociation

points…………………………………………………………………………………………….34

Figure 2.5. LabVIEW front panel for the experimental propane hydrate dissociation setup used

in this study……………………………………………………………………………………...35

Figure 3.1. Correlation plot comparing the experimental mole fraction of water in propane to

those calculated using REPFROP 9.1……………………………………………………..……..46

Figure 3.2. Simplified flowchart for the calculation of dissociation temperature used for the

thermodynamic modeling in this study…………………………………………………………..55

Figure 3.3. Pressure versus temperature plot of this study experimental, literature data and this

thermodynamic model along the Lw–H–C3H8 (g) and Lw–H–C3H8 (l) phase boundaries………60

Figure 3.4. Pressure versus temperature plot of this study experimental result, model, empirical

correlations and literature data along the Lw–H–C3H8 (g) locus………………………………...64

Figure 3.5. Temperature deviations between model and this study experimental data, literature

data and correlations along the Lw-H-C3H8(g) locus……………………………………………65

Figure 3.6. Relationship between propane purities and variance of the literature data along the

Lw-H-C3H8(l) locus to the model presented in this study…………………………………….....67

Figure 3.7. The pressures versus temperatures plot of this study dissociation conditions, model

and literature data along the Lw–H–C3H8 (l) locus……………………………………………....69

Figure 3.8. Temperatures deviation of the model presented in this study to the literature data

along the Lw-H-C3H8 (l) locus…………………………………………………………………..71

Page 14: Thermodynamic Formation Conditions for Propane Hydrates in

xiii

Figure 3.9. A graphical representation of upper quadruple point determination from the point of

intersection of the Lw-H-C3H8 (g) locus and Lw-H-C3H8 (l) loci……………………………….72

Figure A.1. A representative temperature-time plot showing the water freezing points for the

PRT probe calibration……………………………………………………………………………83

Figure B.1. Pressure versus temperature profile for 99.999 mol % C3H8 + H2O showing the

cooling, hydrate formation and heating stages along the Lw-H-C3H8(g) locus…………………86

Figure B.2. Pressure versus temperature profile for 99.5 mol % C3H8 + H2O showing the

cooling, hydrate formation and heating stages along the Lw-H-C3H8(g) locus…………………86

Figure B.3. Pressure versus temperature profiles for C3H8 + H2O showing the cooling, hydrate

formation and heating stages along the Lw-H-C3H8(l) locus……………………………………87

Page 15: Thermodynamic Formation Conditions for Propane Hydrates in

xiv

List of Symbols, Abbreviations and Nomenclature

Abbreviations and Symbols

A

AD

a

corrA

Definition

Helmholtz energy

Average deviations

Spherical molecular core radius

Cross sectional area

mixidA . Ideal gas mixture contribution to the

Helmholtz energy

EA Departure function or excess contribution to

the Helmholtz energy

rA Pure fluid residual Helmholtz energy

οA Ideal gas Helmholtz energy

mjA and mjB Langmuir constant fitting parameter related to

guest molecule j in cavity m

Ap Buoyancy corrected factor

b Temperature correction coefficient

CCS Carbon capture and sequestration

mjC Langmuir constant of gas j in cavity type m

ο

pc Ideal gas heat capacity

ο

pwc Reference standard difference in heat capacity

between ice and liquid water

EOS Equation-of-state

if Fugacity of gas component i in a mixture

Page 16: Thermodynamic Formation Conditions for Propane Hydrates in

xv

83,2 HCH Of Fugacity of water in propane phase

OHHCf283 , Fugacity of propane in water phase

H

wf Fugacity of water in hydrate

L

wf Fugacity of pure water

wf Fugacity of water in empty hydrate lattice

F Degree of freedom in a thermodynamic system

ijF Binary interaction parameter for mixture i and j

ITS The International Temperature Scale

g Gravitational constant

H Enthalpy

HBGS Hydrate based gas separation

οh Ideal gas enthalpy

ο

οh Ideal gas enthalpy at arbitrary reference state

I Ice water

BK Boltzmann constant

Lw Liquid water

LHC Liquid hydrocarbon

LPG Liquefied petroleum gas

mi Number of face of type i hydrate former

N Number of moles

ni Number of edges in the hydrate cage of type i

hydrate former

Page 17: Thermodynamic Formation Conditions for Propane Hydrates in

xvi

N Number of components in a thermodynamic

system

NP Number of experimental data points

P Number of phase in a thermodynamic system

PBD Phase boundary dissociation method

p Pressure

PT Pressure transducer

POI Polynomial term in the dimensionless residual

Helmholtz energy equation

Tp Pressure-density-temperature

PRT Platinum resistance temperature

measp Pressure measured from uncalibrated

transducer

calp Pressure measured from a calibrated transducer

R Ideal gas constant

sI Structure I

sII Structure II

sH Structure H

οs Ideal gas entropy

ο

οs Ideal gas entropy at arbitrary reference state

Tt Triple point

T Temperature

οT Arbitrary reference temperature

cT Critical temperature

Page 18: Thermodynamic Formation Conditions for Propane Hydrates in

xvii

rT Reduced temperature

TtTc Vapour pressure of a hydrocarbon

calT Temperature measured from a calibrated probe

calcT Calculated temperature

expT Experimentally measured temperature

R Distance between encaged gas molecule from

the center of the cavity

r Cavity radius

u & v Coefficient for speed of sound derived from

Einstein equation

VA1 Inlet feed valve

VA2 Outlet valve

w(r) Cell potential function for the interaction

between guest molecules

V Vapour phase

V Volume

VBA Microscoft visual basic for applications

vdWP Van der Waal and Platteeuw model

VLE Vapour liquid equilibrium

mv Number of type m cavities per water molecule

Q Quadruple point

Q1 Lower quadruple point

Q2 Upper quadruple point

X Structure I or II hydrate former

Page 19: Thermodynamic Formation Conditions for Propane Hydrates in

xviii

ix Mole fraction of component i

Y Structure H hydrate former

832, HCOHy Mole fraction of saturated water in propane

z Coordination number

ο

wh Reference enthalpy difference between the

empty hydrate lattice and ice phase.

wh Enthalpy change between empty hydrate lattice

and liquid water

L

w

Difference in chemical potential between

empty hydrate cage and pure water phase

H

w

Difference in chemical potential between

empty hydrate cage and filled hydrate cage

ο

w Reference chemical potential difference

between empty hydrate lattice and pure water

∆s Specific entropy change

∆V Specific volume change

mv Reference volume difference between empty

hydrate lattice and pure ice water phase

Fugacity coefficient

w Activity coefficient of water

)(r Potential energy for interaction between

molecule within the cavities

Characteristic minimum energy

Collision diameter

jm Fractional cage occupancy of gas molecule i

within the hydrate cavity m.

Chemical potential of empty hydrate lattice

Page 20: Thermodynamic Formation Conditions for Propane Hydrates in

xix

L Chemical potential of liquid water

H Chemical potential of filled hydrate lattice

Dimensionless Helmholtz energy

ο Dimensionless ideal gas Helmholtz energy

r Dimensionless residual Helmholtz energy

Density

ο Ideal gas density

c Critical density

fluid Density of deadweight hydraulic fluid

r Reduced density

Reduced density

Reduced temperature

ο Reduced temperature at reference state

512

Pentagonal dodecahedron

512

62 Tetrakaidecahedron

512

64 Hexakaidodecahedron

435

66

3 Irregular dodecahedron

512

68 Icosahedron

Page 21: Thermodynamic Formation Conditions for Propane Hydrates in

1

CHAPTER ONE: Introduction

1.1 Outline

The aim of this study was to measure the dissociation conditions for propane hydrates and to

develop a thermodynamic model for accurate calculation for hydrate formation conditions. At

equilibrium, formation conditions are equivalent to the dissociation conditions. In this chapter,

the history, fundamental background and formation conditions of clathrate hydrate is presented

in order to provide the motivation for this work. The different structures of gas hydrates as well

as their applications are discussed. Because propane is the hydrate of interest in this study,

threats posed by its occurrence in flow assurance and application as a potential replacement for

the energy intensive amine processes for removal of acid gas in sour natural gas production are

discussed. A literature review of available formation conditions also is presented. Finally, a

review of the thermodynamic models used for predicting clathrate hydrate formation conditions

is briefly discussed in the final section of this chapter.

1.2 Motivation for study

Propane is the most prevalent liquefied petroleum gas (LPG),1-2

where LPG consumption is

steadily increasing because of its applications as low carbon energy for transportation, farming,

power generation, cooking and heating purposes.1–4

LPG is primarily propane (C3H8) and butane

(C4H10). C3H8 is formed naturally and is found in association with reserves of oil and natural

gas.5 LPG is normally separated from crude oil or natural gas as a by-product, where natural gas

purification produces approximately 55 % of all LPG while crude oil refining accounts for the

remaining 45 %.1,6

Depending on the source of the LPG and production history of the reservoir,

Page 22: Thermodynamic Formation Conditions for Propane Hydrates in

2

non-hydrocarbon impurities also may be present such as water (H2O), hydrogen sulfide (H2S)

and carbon dioxide (CO2) that must be removed before the LPG can be transported in pipelines

and trucks as a salable product.1,4,6

Gas clathrate hydrates are crystalline solid compounds formed from water and suitably small

molecules at appropriate conditions, typically at high-pressure and low-temperature. These

molecules can be hydrocarbons such as methane (CH4), ethane (C2H6), ethene (C2H4) and C3H8

or non-hydrocarbons, such as like H2S and CO2.7-8

Water naturally coexists with oil and gas

inside all subsurface reservoirs. During production, operating conditions inside producing wells,

subsea transfer lines, risers and pipelines can fall within the conditions that favour hydrate

formation.7,9- 10

Gas hydrates can also form in a single phase fluid with dissolved water such as

in transportation pipelines, transport trucks and LPGs storage facilities.11

Both hydrate

formations can lead to serious safety problems and operational shutdowns which can result in

large economic losses. Because of the potential threats presented by hydrate formation in those

aforementioned areas and C3H8 being a main component of LPG, accurate calculation of C3H8

hydrate formation conditions is important, so as to avoid and control its occurrence during LPG

production.

Although gas hydrates are undesirable during production in the oil and gas industry, they have

been considered for the separation of gas impurities from sour natural gas (i.e., natural gas that

contains an appreciable quantity of H2S), coal and CO2 streams for carbon capture and

sequestration (CCS) technologies.9,12–15

C3H8 hydrates have small unoccupied cages; its hydrate

can be formed so that the small sized gas impurities like H2S and CO2 are captured inside the

Page 23: Thermodynamic Formation Conditions for Propane Hydrates in

3

unoccupied hydrate cages and then dissociated to release the captured impurities. This process

offers an alternative to the energy intensive amine treatment that is currently used for removing

CO2 and H2S from sour natural gas. The design of such a process, otherwise known as hydrate

based gas separation (HBGS), requires accurate knowledge for the formation conditions of C3H8

hydrate and mixed hydrates with other impurities.

1.3 History of gas clathrate hydrates

The discovery of gas clathrate hydrates is often attributed to Sir Humphrey Davy in 1810.7–8,16

However in 1778, John Priestley first observed that sulphur dioxide (SO2) would impregnate

water and cause it to repeatedly freeze, whereas hydrochloric acid (HCl) and silicon tetrafluoride

(SiF4) would not induce this effect when he left the window of his laboratory open overnight in

winter.7,17

Unlike Davy’s experiments, Priestley’s temperature (265 K) of the gas mixture was

well below the ice point and it was not clear whether the structure formed was a hydrate.7-8

Davy

reported the hydrate of chlorine, in which he noted that, the ice-like solid formed at temperatures

greater than the freezing point of water, and the solid was composed of more than just water,

thus, a compound structure must have been formed.16

Villard (1888) first discovered and reported the natural gas hydrates of CH4, C2H6, C2H4 and

C3H8.7,18–19

Clathrate hydrates were referred to as a scientific curiosity up until 1934 when

Hammerschmidt discovered that they were the main culprit for plugging oil and gas pipelines in

Canada.20

Natural gas has suitably sized components such as CH4, C2H6, and C3H8 that are

capable of forming clathrate hydrates in pipelines or other facilities under favourable

Page 24: Thermodynamic Formation Conditions for Propane Hydrates in

4

thermodynamics conditions.7,20

Hammerschmidt discovered the presence of natural gas hydrates

in pipelines at relatively high-pressures and low-temperatures, where it was then believed that it

was impossible for ice to exist.20

However, he noted that removal of the liquid water phase

completely eliminated the possibility of hydrate formation in a pipeline, as the hydrate could not

form until the liquid water dew point was reached. 9,20

The understanding has been revised with

recent work, where it is now well understood that hydrates can form without dense-phase water.8

1.4 Structure and formation of clathrate gas hydrate

Gas clathrate hydrates are structures in which suitably sized small molecules are enclosed or

enclathrated in cages formed by water molecules.7-9

The small molecules are commonly referred

to as a “guest” or “former” while the water forming the hydrate cages are called “host”

molecules.8 Water makes up ca. 85 % of the composition of a hydrate lattice while the guest

molecules constitute ca. 15 %.7,9

Within the cavity, small guest molecules can freely rotate and

vibrate, but have limited translational motion.7,13

The cages are composed of hydrogen-bonded

water molecules mainly in the form of five and six–membered rings.7-8

Von Stackelberg and coworkers (1949) studied and identified the crystal structure of hydrates

using X-ray diffraction techniques.21

They classified the structure into cubic structure I (sI) and

cubic structure II (sII) based on the type of cages found in the crystal and guest molecule

sizes.7,21

Hexagonal structure H (sH) was later discovered by Ripmeester et al. in 1987 through

the use of solid state nuclear magnetic resonance and X-ray diffraction techniques.22

The general

classifications and nomenclature for clathrate gas hydrates are shown in Figure 1.1.

Page 25: Thermodynamic Formation Conditions for Propane Hydrates in

5

Jeffrey et al.(1984) proposed a nomenclature in the form im

in for hydrate structures where ni

represent the number of edges in a face of type i and superscript mi is the number of faces.23

The

structure of these rings defines the types of hydrates formed by a guest molecule.9

The three

structures discussed above are composed of small dodecahedral cages (512

) as a building block.

The 512

cage is composed of twelve pentagonal faces, formed by water molecules that are

bonded to each other by hydrogen bonding, with the oxygen atoms at each vertex.7–9

The 512

62

cage is formed from twelve pentagonal and two hexagonal faces because 512

cages alone will

experience strain on the hydrogen bonds.9,21

The 512

64 cage is made up of 5

12 cages and four

hexagonal faces that further relieve the hydrogen bond strain on the 512

cages when they are

connected to each other through the faces.7,9,22

The irregular dodecahedron (435

66

3) cage consists

of six pentagonal, three square and three hexagonal faces that have a considerable amount of

bond strain.9,22,24

The 435

66

3 cage is slightly larger and less spherical than the 5

12 cage, but both

cages can accommodate small guest molecules like CH4 with the 512

68 cage being slightly larger

in size.24

Table 1.1 compares the hydrate structures at different level of cage occupancy.

Figure 1.1. Classification of clathrate gas hydrates. 512

, 512

62, 5

126

4, 4

35

66

3 and 5

126

8 represent

pentagonal dodecahedron, tetrakaidecahedron, hexakaidecahedron, irregular dodecahedron and

icosahedron cages, respectively.25

(reproduced with permission)

Page 26: Thermodynamic Formation Conditions for Propane Hydrates in

6

Table 1.1. Comparison of structure I, structure II and structure H hydrates.7-8,13

Structure I Structure II Structure H

Crystal system

Lattice description

Water Molecules per unit cell

Cubic

Primitive

46

Cubic

Face centered

136

Hexagonal

Hexagonal

34

Coordination number of

cages in different structure*

512

20 20 20

512

62 24 - -

512

64 - 28 -

512

68

435

66

3

-

-

-

-

36

20

Theoretical Formula†

All cages filled

X+5 ¾ H2O

X+5 ⅔H2O

5X+Y+34 H2O

Mole fraction hydrate former

0.1481

0.1500

0.1500

Structure I Structure II Structure H

Only large cages filled X+7 ⅔ H2O X+17 H2O -

Mole fraction hydrate former 0.1154 0.056 -

Volume of unit cell (m³) 1.728 × 10-27

5.178 × 10-27

-

† Where X is the structure I and II hydrate formers while Y is a structure H former.

*Number of oxygens at the periphery of each cavity.

Page 27: Thermodynamic Formation Conditions for Propane Hydrates in

7

1.4.1 Structure I

In a sI hydrate, each unit cell consists of 46 water molecules that form two small dodecahedral

(512

) and six large tetradecahedral (5

126

2) cages.

7,21 Both cages in the sI gas hydrate can

accommodate only small guest gas molecules with molecular diameters up to 6.0 Å such as CH4,

C2H6, CO2, and H2S.7-8,21

The 512

cages have a internal free diameter of ca.5.1 Å, thus, they can

accommodate the smaller guest molecules less than the size of their diameter. Guest molecules

like CH4 (4.36 Å diameter) can effectively occupy the small cages while larger molecules, such

as C2H6 (5.5 Å diameter), occupy the 512

62 cages which have a diameter of 5.86 Å.

7-9

1.4.2 Structure II

The unit cell of sII gas hydrate consists of sixteen small dodecahedral (512

) cages and eight large

hexakaidecahedral (512

64) cages formed by 136 water molecules.

The 5

126

4 cages have a internal

free diameter of 6.66 Å that allows them to accommodate gases with molecular diameters in the

range of 5.9 to 7.0 Å, such as C3H8 and C4H10.7,9

As is the case for sI, unoccupied 512

cages in sII

can potentially accommodate small molecules like CO2, H2S and CH4.9 This is the principle used

in hydrate based gas separation processes (HBGS) such as flue gas removal and CCS

technology.26–32

sII hydrates are the most common form of gas hydrates encountered in natural

gas production.9

1.4.3 Structure H

The unit cells of structure H (sH) are each made up of three small dodecahedral (512

) cages, two

medium irregular dodecahedral (435

66

3) cages, and one large icosahedral cage (5

126

8) formed

from 34 water molecules.7-8,22

Examples of sH hydrate formers are 2,2-dimethylbutane, 2,3-

dimethylbutane, cyclopentane (C5H10). sH hydrates are always double hydrates, meaning small

Page 28: Thermodynamic Formation Conditions for Propane Hydrates in

8

guest molecules, commonly referred to as “help gas” such as CH4 are required to occupy and

stabilize the small (512

) and medium (435

66

3) cages of the structure, while large molecules with

sizes ranging from 7.5 to 8.6 Å, such as the ones listed above, occupy the large 512

68 cages.

7-8,24

The presence of guest molecules inside the hydrate cages causes a stabilization, so when the

majority of the cages are unoccupied, the hydrate structure collapses.7-9

This stabilization is

postulated to be due to van der Waal forces because there are usually no other bonds available

between the host water cages and a guest molecules.7-9,13

Aside from presence of a stabilizing

molecule, the formation of gas hydrate also depends on two other main conditions: (a) the

right

combination of pressure and temperature (i.e., typically high-pressure and low-temperature), and

(b) a sufficient amount of water.7-8,16

Gas clathrate also can be non-stoichiometric, i.e., their cage

occupancy is a function of the pressure and temperature conditions and not the number of cages

available.8

Hydrate formation is commonly favoured in locations such as gas valves where

narrowing within the valves causes Joule-Thompson temperature reduction.8,20

Factors that can

enhance kinetic hydrate formation are nucleation sites such as imperfections in pipelines, weld

spots, or pipeline fittings and high-velocity turbulence.

1.5 Applications of gas hydrate

1.5.1 Hydrogen (H2) storage

Dyadin discovered the clathrate hydrate of hydrogen (H2) in 1991 which has since been

extensively studied for hydrogen storage.33–38

With a diameter of 2.72 Å, H2 typically forms a

type II structure at a pressure of 300 MPa with an H2 to H2O ratio as high as 1:2 due to the 512

Page 29: Thermodynamic Formation Conditions for Propane Hydrates in

9

and the 512

64 cages holding up to 2 and 4 molecules of H2 respectively.

33,35-36 Because of the

relatively high pressures required for pure H2 hydrates, binary hydrates (i.e., hydrate cages

containing two different guest molecules) are easier to form and study, to reduce the pressure for

H2 storage in the clathrate hydrate cages.38

Binary hydrates have been reported for H2 and

tetrahydrofuran (THF) at relatively lower pressures.34,36,38

Although, THF reduces the hydrate

formation pressure and increases the dissociation temperature, it also occupies some of the

hydrate cages that can potentially be used for hydrogen storage.26,39

C5H10 has been reported as an alternative to THF because it also reduces H2 hydrate formation

pressures but it also occupies the hydrate cages.39-40

Skiba et al. (2009) reported that at pressures

between p = 200 – 250 MPa, the double hydrate of the H2–C3H8–H2O system decomposes at a

temperature 20 K higher than pure C3H8 hydrate and H2 occupancy inside the hydrate cages also

increases.41

This makes C3H8 a better potential alternative to both C5H10 and THF for hydrogen

storage.

1.5.2 Separation processes

The World’s proven conventional natural gas reserves are abundant with a large number of the

reserves considered to be sour.7,13,42

Separation of CO2 and H2S from natural gas in the vapour

phases has been done on an industrial scale for at least the last 70 years and has been typically

achieved using aqueous amines.43-44

HGBS processes are being developed for separating CO2

and H2S from sour natural gases as an alternative to this energy intensive amine processes

currently used.26,39,44–45

The captured gas is released at high-pressure and low-temperature,

reducing costs for liquefaction of the CO2 and H2S products.26,39

Another advantage of an HBGS

Page 30: Thermodynamic Formation Conditions for Propane Hydrates in

10

process are that they operate near room temperature during the dissociation process, thereby

reducing heating costs.39

Most HGBS processes are found to be kinetically limited but

thermodynamically feasible.

1.5.3 Desalination of seawater

Seawater is an important source of potable water in many countries with shortages of fresh

water. Desalination technologies such as multi-stage distillation, reverse osmosis and electro-

dialysis are energy intensitve.46-48

Hydrate based separation processes provides a viable

alternative that, again, are estimated to consume less energy.39

The separation of salts and other

gaseous impurities can be achieved by sequential clathrate hydrate formation and dissociation.

When the hydrates form, the cages omit the salts so that when the structure dissociates, only the

gas and pure water is released.48-49

1.5.4 Potential source of energy

At T = 273 K and 1 atmosphere, a cubic meter (m3) of completely filled natural gas hydrate

contains ca. 96 kg of CH4, giving it the potential to be a future natural gas source.13

Large

quantities of natural gas exist as hydrates in the arctic and permafrost regions of the earth, as well

as in the ocean bottoms.7,50–52

The available energy from gas hydrate sources alone is estimated

to be twice that of all other fossil fuels combined.50,53

Although more studies need to be done to

prevent uncontrollable sand production and rapid depressurization of the hydrate bed during

exploration.54

Page 31: Thermodynamic Formation Conditions for Propane Hydrates in

11

1.5.5 Natural gas hydrate in flow assurance

Despite the beneficial uses of hydrate already discussed above, their occurrence during natural

gas production is undesirable because they can plug pipelines and block process facilities during

transportation. Hydrate plugs can form when conditions are favourable such as (i) during a

production start-up, (ii) during a restart following an emergency shut-down and operational shut-

in due to temperature gradient and because reservoir heat is not available, (iii) having uninhibited

water of condensation, or (iv) when local cooling occurs due to flow across a valve or

restriction.7– 9

If not appropriately predicted and avoided, normal operating conditions inside

producing wells, flow lines, valves and pipelines can fall within the hydrate stability zones where

hydrates can lead to the complete shutdown of operations and results in loss of billions of dollars

in revenue.7,9,13

Beside the economic impacts, there can be serious safety hazards associated with gas hydrates.

During the dissociation of hydrate plugs in the pipeline, there can be a substantial pressure drop

across a plug and before the plug detaches from the pipe wall. Upon releasing from the pipeline

wall, the pressure difference can cause the hydrate to reach speeds greater than 300 km hr-1

within the pipeline.7 Not only can the plug itself break through the pipeline, but this phenomenon

also facilitates the compression of the downstream gas which can result in pipelines blowouts or

ruptures possibly leading to the injury or death of any nearby workers.7,13,55

In 2013, two oil

workers working at a ConocoPhillips site in Alberta were struck by a hydrate plug resulting in

the death of one and serious injuries to the other.56

Page 32: Thermodynamic Formation Conditions for Propane Hydrates in

12

1.5.5.1 Gas hydrate prevention and control

Gas hydrate formation in flow assurance can be prevented through one of the following

processes: (a) maintaining the system temperature above the hydrate formation temperature

through the use of heat or insulation, (b) dehydrating the hydrocarbon fluid to below a specified

level, (c) operating at lower pressure than the hydrate formation pressure, or (d) injection of a

chemical inhibitors to prevent or mitigate hydrate formation: (i) thermodynamic inhibitors such

as methanol or glycol to decrease the hydrate formation temperature and prevent crystal

formation, (ii) kinetic inhibitors such as poly N-vinyl pyrrolidone to decrease the rate of

formation and growth of hydrate crystals and (iii) anti-agglomerates, such as quaternary

ammonium salts, allow hydrates to form but to a controlled crystal size so that they can still be

transported through pipelines.7-8,55,57-58

Regardless of the hydrate prevention methods, accurate

calculation of formation conditions are very important in order to determine the best prevention

strategy.

1.6 Importance of C3H8 hydrate formation conditions studies

C3H8 with a diameter of 6.3Å typically forms a sII clathrate hydrate in the presence of water;

however, C3H8 is too large to occupy the 512

cages; therefore, it occupies the large cages of 512

64

(6.66 Å) leaving the 512

cages empty.7,9,11,59-60

The 512

cages in the sII hydrate can potentially

accommodate the molecules with small diameters such as H2S, CO2 and CH4 at appropriate

temperature and pressure conditions. In fact, these smaller formers often stabilize the sII hydrates

even more than just the primary former such as C3H8. The components of gas mixtures are

Page 33: Thermodynamic Formation Conditions for Propane Hydrates in

13

partitioned through forming hydrates according to their relative difference in occupation

thermodynamics and kinetics within hydrate cavities.9,13,61

As discussed earlier, hydrates can affect the processing, storage and transportation of LPGs. If

the conditions are accurately known or predictable, operators are able to properly design schemes

to deal with hydrate issues in flow assurance and develop HBGS processes. Furthermore, C3H8 is

the reference material for which most sII calculations are based on, i.e., it is thought to be better

studied than other sII hydrates.

Thus, the specific aims and objectives of this study are to:

1. Measure the hydrate formation conditions (i.e., pressure and temperature) of C3H8 in the

presences of liquid water by using high purity C3H8 (99.999 mol %).

2. Develop a semi–empirical model based on the Clausius–Clapeyron relation for the rapid

estimation of hydrate formation condition of C3H8.

3. Develop a robust thermodynamic model based on reduced Helmholtz equation–of–state

(EOS) and the van der Waals and Platteuw (vdWP) model for the accurate prediction of

C3H8 hydrate formation conditions.

1.7 Phase behaviour and avoiding hydrate formation

Fundamental thermodynamic properties of natural gas components and mixtures are required for

calculating fluid behaviour over natural gas production conditions.62-64

When considering the

phase behaviour of a hydrate (mixture), the Gibbs Phase Rule is a useful tool in recognizing

changes in the temperature-pressure behavior when changing thermodynamic degrees of

Page 34: Thermodynamic Formation Conditions for Propane Hydrates in

14

freedom. Thermodynamic degrees of freedom (F) or variance of a chemical system is given

as7,65-68

F = C + 2 – P. (1.1)

Where C and P represent the number of components and phases respectively in a chemical

system. For example, a system with pure hydrate former in the gas phase, liquid water and

hydrate phases (C = 2 and P = 3) is univariant at any given temperature, i.e., either pressure or

temperature can be changed independently without changing the state of the system while the

same system with four different phases is invariant (neither pressure nor temperature can be

changed).

1.7.1 Phase behaviour of a hydrocarbon hydrate former

The typical phase diagram for a hydrocarbon former in the presence of water is shown in Figure

1.2. The broken blue line “TtTc” represents the vapour pressure of a pure hydrocarbon hydrate

former. Tt and Tc represent the triple and critical points of the hydrate former, respectively. The

triple point is the point of coexistence of solid, vapour (V) and liquid hydrocarbon (LHC) phases

of the hydrocarbon former while at or beyond the critical point there are no phase boundaries.

For most hydrocarbon formers, four different phases of vapour, hydrate (H), liquid water (Lw)

and ice water (I) can exist together in equilibrium at a relatively lower pressure and temperature

referred to as the lower quadruple point, Q1. LHC, V, Lw and H phases also coexist at a relatively

higher pressure and temperature at the upper quadruple point Q2. The quadruple points (Q) are

invariant for any particular hydrocarbon / hydrate former according to equation 1.1.7-8,66

The

curve 1–1´– 1´´ represents the limit of the hydrate stability region where hydrates formed at

pressures above the curve are thermodynamically stable.66

At higher temperatures, on the right

Page 35: Thermodynamic Formation Conditions for Propane Hydrates in

15

side of line Q2Tc, hydrate phase cannot be formed, but at lower temperature hydrate is formed as

shown at the left hand side of the line Q1Q2. At even lower-temperature and pressure conditions,

an ice phase coexisting with hydrocarbon vapour prevails over hydrate phase formation.7-8,66

Three different phases of I-H-V form along the line labelled 1, usually below T = 273.15 K and

relatively low pressure, while Lw-H-V phases exists in equilibrium with each other at the line

labelled 1´ greater than T = 273.15 K. At relatively high temperatures and pressures, LHC-H-Lw

phases also coexist at the line labelled 1´´.

Figure 1.2. Typical p−T diagram for a hydrocarbon hydrate former. Q1and Q2 represents the

lower and upper quadruple points respectively. Line TtTc represents the vapour pressure line of

hydrocarbon while 1–1´–1´´ represents the hydrate stability region consisting the I–H–V, Lw–

H–V and LHC–H–Lw phase boundaries.

1.7.2. Phase behaviour of C3H8 + H2O system

C3H8 hydrates can exist at temperatures above the normal melting point of either ice or

C3H8.11,59- 60

The phase diagram of the C3H8 + H2O system is shown in Figure 1.3. The red line

and blue curve represents the vapour pressure and hydrate stability region of C3H8 respectively.

lnp

T

LHC–H I–LHC

273.15 K

Tc

Tt

LHC-Lw

I–V

V–LW

Q2

Q11

1

Page 36: Thermodynamic Formation Conditions for Propane Hydrates in

16

C3H8 has two quadruple points in the presence of liquid water: Q1 and Q2.7-8,11,59-60

Beyond Q2,

above temperature T = 279 K, C3H8 hydrate will not form at any pressure but at temperatures

lower than T ≈ 279 K and pressures greater than p ≈ 0.16 MPa, hydrate phase can coexist with

other phases present. However, at temperatures and pressure lower than T = 273.15 K and

p ≈ 0.16 MPa respectively, up to Q1, hydrate phase can coexists with other phases, notably ice

water and C3H8 (g) as shown in the Figure 1.3. Between Q1 and Q2 three different phases of

C3H8 (g), Lw and H coexist in equilibrium while the steep blue line from Q2 represents the

condition for coexistence of C3H8(l)–H–Lw phases typically at relatively high pressure

p ≈ 0.4 MPa and temperatures between T = 273.16 - 279 K.7-8

C3H8 hydrate will begin to

dissociate when the pressure is decreased or temperature is increased while along the hydrate

stability curve.7-8,11,60

Figure 1.3. p – T diagram of C3H8 + H2O system. Q1 is the lower quadruple point and Q2

represent upper quadruple point of C3H8 hydrate. Blue and red lines represent the hydrate

stability region and vapour pressure of C3H8 respectively.

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

1.60

1.80

265 267 269 271 273 275 277 279 281 283 285

Q2

Lw + H

C3H8(l) + Lw

C3H8(g) + LwQ1

I + H

p/

MP

a

T / K

Page 37: Thermodynamic Formation Conditions for Propane Hydrates in

17

Different values of Q1 and Q2 have been reported by various authors. In addition, there is a large

variance in reported dissociation pressure and temperature for the hydrate stability zones. The

literature review for the dissociation conditions and the quadruple points will be discussed in the

next section.

1.7.3 Semi – empirical model for hydrate dissociation correlation

The Clausis-Clapeyron equation provides a relation for estimating the phase transition between

two different phases of a pure component. The slope of tangent (dp/dT) of the line separating two

phases is mathematically expressed as7-8,67-68

VT

H

V

s

dT

dp

, (1.2)

where ∆H, T, ∆V and ∆s represent the enthalpy change, temperature, specific volume change and

the specific entropy change of the phase transition, respectively. Experimental data at the

Lw-H-V and Lw–H–LHC phase boundaries can be correlated using this relation for the rapid

estimation of dissociation pressures and temperatures. This relation also enables the estimation

of H which is difficult to measure experimentally.67-68

1.8 Experimental dissociation data for C3H8 hydrate

Hydrate formation conditions for C3H8 are more commonly reported in the literature along the

Lw–H–C3H8(g) phase boundary because it is easier (experimentally, empirically and

theoretically) to work with C3H8 in the gas phase versus the liquid phase.69-84

This can be

attributed to difficulties of measurement in the liquid phase, where the effect of impurities such

as nitrogen and small leaks in the experimental setup can be more pronounced. Also, there are no

Page 38: Thermodynamic Formation Conditions for Propane Hydrates in

18

empirical correlations for the Lw–H–C3H8(l) phase boundary, because the available data is

variable in this region. Thus, the locus above Q2 is estimated at constant temperature. Table 1.2

shows the summary of the experimental data covering the Lw–H–C3H8(g) equilibrium

conditions. Carroll and Kamath fit C3H8 hydrate dissociation data on the Lw–H–C3H8(g) locus

from T = 273.15 - 278.75 K with a Clausius – Clapeyron type relation.8,69

Carrol gives the relation p (MPa) = exp (–259.5822 + 0.58 * T + 27150.7 / T) while Kamath

correlations gives the relation p (KPa) = exp (67.13 + (−16921.84) / T), where T is in Kelvin in

both correlations.

Table 1.2. Summary of experimental dissociation conditions along the Lw–H–C3H8(g) phase

boundary.

Source No. of data

% Purity

T range / K p range / MPa

Miller and Strong82

8 - 273.20 - 277.13 7.844 - 18.682

Reamer et al.70

Tumba et al.71

6

3

> 99.00

99.50

274.3 - 277.2

274.6 - 278.1

0.2401 - 0.4140

0.250 - 0.540

Engelos and Ngan73

6 99.50 274.2 - 278.3 0.2296 - 0.5353

Robinson and Mehta74

5 99.50 274.20 - 278.87 0.2068 - 0.5516

Patil75

5 99.50 273.60 - 278.00 0.207 - 0.248

Verma et al.72

9 > 99.5 273.9 - 278.4 0.188 - 0.562

Kubota et al.76

15 > 99.5 273.25 - 278.45 0.712 - 0.552

Deaton and Frost77

5 99.80 273.70 - 277.04 0.1827 - 0.3861

Thakore and Holder78

5 99.90 274.00 - 278.15 0.2170 - 0.5099

Page 39: Thermodynamic Formation Conditions for Propane Hydrates in

19

Source No. of data % Purity

T range / K p range / MPa

Den Huevel et al.79

11

99.95

276.77 - 278.55

0.368 - 0.547

Nixdorff80

10 >99.995 273.55 - 278.52 0.18824 - 0.5490

Maekawa84

10 99.999 274.2 - 278.1 0.211 - 0.509

In contrast to the vapour phase, there are relatively few experimental data along the Lw–H–

C3H8(l) phase boundary due to the difficulties previously mentioned. Table 1.3 shows the

literature summary of the dissociation conditions in the Lw–H–C3H8(l) phase boundary.

Table 1.3. Summary of the experimental dissociation conditions along the Lw–H–C3H8(l) phase

boundary.

Reamer et al.1952 reported dissociation data using the lowest C3H8 purity of < 99.5 mol % while

Nixdorff, 2007 used the highest purity of 99.995 mol % C3H8.70,80

There are discrepancies in the

experimental data reported for C3H8 hydrate in the vapour and liquid regions, potentially caused

by the various purities and or the buoyancy of the hydrate used which will be discussed in detail

Source

No. of data

% Purity

T range / K

p range / MPa

Wilcox et al.81

7 - 278.6 - 279.2 0.0807 - 0.6115

Makogon83

9 99.95 278.05 - 278.28 0.555 - 34.999

Den Heuvel et al.79

17 99.95 278.75 - 278.86 0.893 - 9.893

Verma et al.72

4 > 99.50 278.4 - 278.6 0.562 - 11.2999

Reamer et al.70

3 > 99.00 278.6 - 278.8 0.0684 - 0.2046

Page 40: Thermodynamic Formation Conditions for Propane Hydrates in

20

in chapter three. Note that unspecified impurities can increase or decrease the relative stability of

the hydrate phase with respect to the fluid phases.

The reported quadruple points for C3H8 in the presence of liquid water are also sparse. Carroll

(2003) reported T = 273.05 K and p = 0.172 MPa while Harmen and Sloan (2009) measured a

similar pressure to Carroll but a slightly higher temperature of T = 273.10 K for Q1.7,60

Q2 are

more often reported than Q1; a summary of Q2 values reported in the literature are shown in

Table 1.4.

Table 1.4. Summary of C3H8 hydrate upper quadruple points reported in literature.

Source

% Purity

p / MPa

T / K

Makogon83

99.95 0.555 278.3

Robinson and Mehta74

99.5 0.5516 278.872

Den Heuvel et al.79

99.95 0.6a 278.62

Carroll8 - 0.556 278.75

Verma et al.72

99.5 0.562 278.4

Average a

0.566 278.588

(Standard deviation)

0.004 0.2378 aAverage pressure does not include the value of Den Heuvel et al.

79 due to lower

reported precision.

1.9 Review of gas hydrate thermodynamic models

In 1959 Van der Waals and Platteuw proposed the solid state and fluid solution theory for

modeling the hydrate phase based on the equality of water chemical potential between the

Page 41: Thermodynamic Formation Conditions for Propane Hydrates in

21

hydrate phase and the co-existing water phases, i.e., ice, liquid or vapour water.85

The model was

developed based on the assumption that hydrate formation is similar to gas adsorption with the

following conditions: (i) each hydrate cavity is a spherical cage which can hold one gas molecule

at a time, (ii) there is no interaction between the guest molecules and London forces are the only

force present between the guest-host interactions; all other polar forces are assumed to be

integrated in the hydrogen-bonded hydrate lattice, (iii) the host molecule’s contribution to the

free energy is independent of the mode of occupancy by the guest molecules, so the guest

molecule inside the hydrate cage does not distort the hydrate cage, (iv) the enclathrated guest

molecule can only undergo rotational and vibrational motion within the cavities but no

translational motion and, (v) classical statistic mechanics is valid under all conditions. 7,85-86

The Langmuir constant C(T) is an important parameter used to define the interaction behavior

between the gas and water molecules within the cavities and it can be calculated from either the

Lennard-Jones 6-12 potential, Lennard–Jones–Devonshire model or the Kihara potential

model.85

McKoy and Sinanoglu (1963) suggested that the Kihara potential model with a

spherical core was more suitable for estimating the gas-water interactions within the cavities.86-87

In 1972, Parrish and Prausnitz presented a modification to the vdWP model for calculating gas

hydrate equilibra in multi-component systems by introducing the Kihara potential model for

estimating the interaction between the guest and host molecules.7,86

They also presented a

detailed algorithm for calculating dissociation pressure and temperature in the I-H-V and Lw-H-

V loci as well as reference hydrates for different lattice structures.86

Page 42: Thermodynamic Formation Conditions for Propane Hydrates in

22

Holder et al. in 1980 introduced reference thermodynamic properties for gases and hydrates to

replace the reference hydrates initially used for the different cages of hydrate structures in

Parrish and Praunnitz model.88

Their lattice distortion and guest-guest interactions depend on the

properties of the guest and are not accounted for in previous models.89

Chen and Guo (1996)

introduced the concept of equality of fugacities in coexisting phases at equilibrium in gas-water

mixture to hydrate modeling and used the Lennard-Jones 6-12 potential for calculating the gas-

water interactions.90

Klauda and Sandler’s (2000) fugacity model attempted to correct some of

the assumptions made by van der Waal and Platteeuw, primarily by taking into account different

degrees of lattice distortion caused by each guest and therefore subsequent changes in the empty

lattice fugacity or Gibbs free energy. Their proposed model removes the reference parameters

widely used in the previous vdWP type models for hydrate structure and instead introduced some

guest specific parameters.85,91

Klauda and Sandler (2000) also included the energy contributions

from the surrounding second and third shells for calculating the Kihara potential parameter.91

Later in 2003, Klauda and Sandler accounted for the guest-guest interactions and dual occupancy

of guest molecules in the cavities in their model.92

Unlike Klauda and Sandler, Ballard (2004)

accounted for the hydrate lattice distortion due to presence of guest molecules by suggesting

empty hydrates lattice of CH4, C3H8 and CH4 + neohexane as the reference standard hydrates for

sI, sII and sH respectively. Perturbation from the standard states is then accounted for by using

the activity coefficient in his model.7,89

Thus, C3H8 hydrate condition are important, because the

C3H8 hydrate is considered the reference material for type II hydrates in general.

Cubic EOSs such as the Peng-Robinson, Soave-Redlich-Kwong, Valderrama-Patel-Teja

equations and the statistical associating fluid theory are commonly used to model the fluid

Page 43: Thermodynamic Formation Conditions for Propane Hydrates in

23

phases in hydrate dissociation modeling. These are used for computational speed, whereas, for

reference quality results, more accurate EOS such as the reduced Helmholtz energy equation can

be used. Further descriptions of the vdWP model and modeling will be discussed in chapter

three.

Page 44: Thermodynamic Formation Conditions for Propane Hydrates in

24

CHAPTER TWO: Review of Literature Techniques, Experimental Procedure and

Calibration

2.1 Outline

The description of the methods and selected earlier apparatus used for studying the dissociation

conditions of gas hydrate are presented in this chapter. The experimental setup and procedures

employed in this study along the Lw–H–C3H8(g) and Lw–H–C3H8(l) phase boundaries also are

presented. The calibration procedures and results for the pressure transducers, platinum

resistance thermometer and autoclave volume are discussed in Appendix A.

2.2 Methods of studying gas hydrate phase equilibra

There are two primary methods for studying gas hydrate phase equilibra in the laboratory:

dynamic and static methods.93

2.2.1 Dynamic method

In this method, a gas is continuously flowed through a chamber or loop, which is maintained at

condition that favours hydrate formation, usually at low temperature before adding water to the

gas.20,93

This method is more suited for the studies of hydrates formation kinetics, or the effect of

electrolytes and inhibitors on hydrate formation, but it can be used to measure dissociation

conditions.93

In this case, continuous flow of gas through the loop or chamber can be stopped to

allow the system to equilibrate at the desired pressure for hydrates to form. Once hydrates are

formed, the temperature or pressure of the system can then be increased and decreased,

Page 45: Thermodynamic Formation Conditions for Propane Hydrates in

25

respectively, until the hydrates begins to melt so that the dissociation conditions can be measured

by a change in effluent composition.20,93

2.2.2 Static method

This method involves the growth of hydrate crystals in a static high-pressure vessel or autoclave

vessel followed by the subsequent dissociation of the formed hydrate while measuring the

temperature and pressure (with or without a visual window).7,93

For phase equilibra studies of

gas hydrate, this method is preferred because of the ease by which intensive properties such as

the dissociation pressure and temperature can be measured.93

Generally, this method can be

subdivided into either: isothermal, isobaric, or isochoric techniques. 93-94

2.2.2.1 Isothermal Method

In this method, the pressure of a gas-water system is increased above an estimated hydrate

formation pressure at a constant temperature until hydrate begins to form.7,73,77,84,93

The pressure

is usually controlled by withdrawal or addition of gas or aqueous liquid. Hydrate formation

creates a temporary increase in temperature and rapid reduction in the pressure until the hydrate

crystal is formed completely. The temperature increases temporarily because the formation is

exothermic and the pressure reduction is due to the enclathration of the gas molecules inside the

hydrate cages.7,77

Provided there is excess H2O, this gas molecule enclathration causes a pressure

reduction to either Lw–H–V or Lw–H–LHC phase boundary condition.7,70,79

After complete

hydrate formation, the system temperature remains constant and the equilibrium dissociation

pressure is determined by decreasing the pressure and taken as the point of dissolution of the

hydrate crystal phase.7,77,93-94

Page 46: Thermodynamic Formation Conditions for Propane Hydrates in

26

2.2.2.2 Isobaric method

In an isobaric system, the pressure of a gas-water system is maintained constant by suitable

sources such as a positive displacement pump or gas exchange through an external reservoir.7,94

The temperature is then lowered and the initiation of hydrate formation is indicated by a

significant reduction in volume of the gas from the source. The formed hydrates crystals are

heated continuously or stepwise to measure the equilibrium hydrate dissociation temperature by

(i) visually observing through a sight glass, the temperature of complete disappearance of the

hydrate phase, or (ii) point of intersection between the cooling and heating pressure curves or

(iii) heat released using a calorimeter.7,74,94

2.2.2.3 Isochoric method

This method is similar to the isobaric method, but in the isochoric method the system is

examined in a constant volume vessel or cylinder where the gas-liquid water system is cooled to

form hydrates after which the temperature is increased to dissociate the hydrates.7,84,94

Figure 2.1

shows the typical p–T curve of a gas–water system for the cooling, formation and dissociation

processes of hydrates in an isochoric system. The system is first allowed to equilibrate at

conditions above the hydrate formation condition at point “A” before cooling it down to a lower

temperature at point “B” where the hydrate crystals begins to form. The hydrate formation is

characterised by a drastic reduction in pressure, as shown in the curve from “B” to “C”. The

formed hydrate is slowly heated from “C” to “A”. The equilibrium point “D” (intersection

between the cooling and hydrate heating curves) represents the condition where hydrates are

completely melted. At the inflection point, point “D”, the thermodynamic degrees of freedom

increased from 1 to 2 upon heating from “C” to “A”. Increasing the temperature outside the

Page 47: Thermodynamic Formation Conditions for Propane Hydrates in

27

hydrate stabilization regions beyond “D” will results in smaller pressure changes that are

associated with vapour-liquid-equilibra (VLE).

Figure 2.1. A Typical p-T curve of a gas hydrate formation and dissociation using the isochoric

method. A-B; gas cooling, B-C; hydrate formation, C-D; hydrate dissociation.

The isochoric method can provide more pressure and temperature information near a hydrate

forming system in less time when compared to the isobaric and isothermal systems because the

procedure can be automated to be carried out continuously. Also, because the method is a non-

visual technique in most cases, it can be less subjective than the other methods described above

because there is no human error associated with visual observation of hydrate dissociation. 80,94

2.2.2.4 Phase boundary dissociation method

In 2012, Loh et al., presented the phase boundary dissociation (PBD) method for measuring

dissociation conditions of methane hydrates in fresh and sea water in a porous media for

Pre

ssu

re

Temperature

A

B

C

D

Hy

drate

form

ation

Hydrate begin

to form

Complete hydrate

formation

Equilibrium

T and p

C3H8 + H2O loaded

into autoclave

Page 48: Thermodynamic Formation Conditions for Propane Hydrates in

28

pressures ranging from p = 2.30 - 17.00 MPa and varying water compositions.95

This approach is

a modification to the isochoric method in which the hydrates are formed by reducing the

temperature of gas-water system inside a constant volume cylinder or autoclave followed by a

controlled dissociation of the formed hydrates along a phase boundary (e.g Lw–H–V).95-96

According to the Gibbs phase rule (F = C – P + 2), for a pure hydrate former and liquid water

system, i.e., C = 2, P = 3 and F = 1, either the temperature or pressure can be changed without

affecting the number of phases along the phase boundaries/loci. As a result, the system would

dissociate along the triple loci for as long as the three phases coexist and the point of intersection

between the cooling and heating curves is taken as the equilibrium hydrate condition similar to

the isochoric method.94-96

The PBD method has the advantage of generating more equilibrium

data in a short period of time compared to all other previously discussed methods. Ward et al.

2015, reported ca. 4.8 hours per data point when using the PBD method compared to 40 – 45

hours per data point for the isochoric method for the hydrate dissociation condition of H2S along

the Lw-H-V phase boundary.96

Irrespective of the method used for studying gas hydrate phase equilibra, agitation is important

for the design of any apparatus.7-8,20,93

In 1896, Villard first observed that an increase in agitation

caused a decrease in the quantity of liquid water phase and or increase in hydrates formation.

Likewise, Hammerschmidt (1934) also noted that some form of agitation such as gas bubbling

through water, increased velocity (turbulence) or flow fluctuations, initiated and accelerated the

hydrate formation.18,20

The apparatus developed by Deaton and Frost (1946) for studying hydrate

dissociation conditions also was oscillated about a horizontal axis to create a form of agitation

and facilitate good hydrate formation rates.77

It is well known that any type of mild agitation can

Page 49: Thermodynamic Formation Conditions for Propane Hydrates in

29

enhance nucleation in supernatant fluids.97-98

Studies also show that higher stirring rate increase

the nucleation rate and surface area of gas hydrates.97-101

Generally, in a gas-liquid system,

agitation helps to facilitate the mass transfer from one phase to another thereby promoting the

rate of hydrate formation.7-8,100-101

2.3 Review of selected experimental apparatus for hydrate dissociation studies

John Cailletet developed an apparatus in 1887 commonly referred to as the Cailletet apparatus

for the original purpose of studying the liquefaction of oxygen, but it was used later on to study

the dissociation conditions of some mixed hydrates such as CO2 + PH3.79,102-104

The apparatus mainly consists of a thick-walled pyrex capillary tube about 50 mm in length, with

an internal and external diameter of 3 and 10 mm respectively which enabled the visual

observation of phase transition. One end of the tube is closed while the other end is open with a

conical thickening to allow an autoclave vessel made of stainless steel to be mounted on it. The

capillary tube is filled with gas-water mixture constrained over mercury which serves as a

pressure transmitting fluid and prevents the mixture from being contaminated with the silicon oil

used in an adjacent hydraulic device for generating pressure. A glass coated iron rod on top of

the vessel enables mechanical stirring of the mixture and the glass tube is kept at the desired

temperature by a thermostat with circulating oil.102-103

The temperature and pressure were

measured by using a platinum resistance thermometer and a deadweight pressure gauge,

respectively.102

Page 50: Thermodynamic Formation Conditions for Propane Hydrates in

30

Later, in 1937 Deaton and Frost constructed a static hydrate equilibrium apparatus which was

used as a model for many other apparatuses that are use today.7,77

Figure 2.2 shows the schematic

diagram of Deaton and Frost experimental setup main parts. It consists of a high-pressure cell

stainless steel cell containing a quartz or sapphire window. The cell is thermally regulated

through a cooler or heater and is connected to a rocking motor to agitate the system. Gases are

flowed above the liquid water through the valve into the cell and are allowed to exit through an

outlet valve connected to a vacuum or pressure gauge. The isothermal method was used for their

study of hydrate dissociation; the pressure was reduced by letting out some gases inside the cell

to cause hydrate dissociation which was visually observed for the estimation of equilibrium

hydrate pressure.7,77

Figure 2.2. Experimental schematic of Deaton and Frost’s apparatus for phase equilibra studies

of gas hydrates.

Hammerschmidt (1946) designed and used the first dynamic apparatus to investigate natural gas

hydrate kinetic and thermodynamic formation conditions (see Figure 2.3).20,80

The setup consists

of a temperature controlled water bath marked “6” and Pyrex glass tube, “5” which contains the

Rocking

motor

CoolerHeater

Rocking cell

Water bath

Gas + water

Page 51: Thermodynamic Formation Conditions for Propane Hydrates in

31

compressed gas. The gas is flowed through the inlet, “1”, and follows through the loop, “2”,

made of copper tubing and immersed inside the water bath. Water is added from the reservoir

marked “3” by gravity flow to the gas when passing through the internal copper tube, “4”.

Temperature was measured with an iron–constantan thermocouple, “7”, and the gas exit the

setup at “11” through a gas meter, “12” which enables the measurement of the amount of gas

exiting the apparatus. The pressure was measured with a Bourdon tube gauge, “10”, which was

calibrated by a piston gauge. The precooled water bath “6” allows the gas to pass through the

apparatus at various velocities, pressures and temperatures. This apparatus was mainly used to

investigate different kinetic factors that facilitate hydrate formation.20

Figure 2.3. Diagram of the dynamic apparatus designed and constructed by Hammerschdmit for

studying hydrate formation conditions. 1 – gas inlet, 2 – copper precooling coil, 3 – water supply

reservoir, 4 – copper tube, 5 – pyrex glass, 6 – constant temperature bath, 7 – thermocouple

junction , 8 – millivolt meter, 9 – drip, 10 – pressure gauge, 11 – pressure reducing valve, 12 –

gas meter. 20

(reproduced with permission)

Page 52: Thermodynamic Formation Conditions for Propane Hydrates in

32

Different methods are used in the literature for the studies of C3H8 hydrate dissociation

conditions. These methods are summarized in Table 2.1, where the isothermal and isobaric

methods are more common.

Table 2.1. Methods used for the study of C3H8 hydrate dissociation conditions.

Study Method

Reamer et al.

70 Isothermal and Isobaric

Tumba et al.

71 Isochoric

Verma et al.

72 Isothermal

Engelos and Ngan

73 Isothermal

Robinson and Mehta

74 Isobaric

Patil

75 Isobaric

Kubota et al.

76 Isobaric

Deaton and Frost

77 Isothermal

Thakore and Holder

78 Isothermal

Den Heuvel, et al.

79 Isothermal (Cailletet apparatus)

Nixdorff80

Isochoric

Wilcox et al.

81 Isothermal

Miller and Strong

82 Isothermal

Makogon

83 Static and dynamic

Maekawa

84 Isothermal and Isochoric

2.4 Apparatus used for this study.

The setup used in this study was initially assembled by a previous graduate student, Zachary

Ward, for the study of dissociation conditions for mixed sour gas hydrates; a description of the

Page 53: Thermodynamic Formation Conditions for Propane Hydrates in

33

commissioning can be found in Zach’s thesis.105

Repairs were made to the setup for the study of

dissociation conditions of C3H8 hydrate i.e., the PRT and pressure transducer were replaced

because of leaks and electronic malfunction, respectively. Figure 2.4 shows a schematic of the

experimental setup. The autoclave vessel is constructed of Hastelloy–C276 coupled with

magnetic stirrer to enhance the mass transfer in the C3H8-H2O mixture. The autoclave has a

volume of about 45.00 cm3 and has a maximum working pressure p = 20.68 MPa (3000 psia) and

working temperature of T = 263.15 - 308.15 K which are within this study experimental

conditions. The apparatus was originally commissioned with a Paroscientific Inc. Digiquartz

410KR-HT-101 Pressure Transducer and a four-wire 100 Ω platinum resistance thermometer

with a PT-104 temperature data logger (Pico Technologies) which have measurement precisions

of δp = 41045.3 MPa and δT = ± 0.001K, respectively. The Paroscientific Inc. Digiquartz

410KR-HT-101 Pressure Transducer was later replaced with a Keller Druckmesstechnik PA-33X

transducer with a precision of δp = ± 0.001 MPa for measurements in the Lw-H-C3H8(l) locus.

The calibration procedure and results for the PRTs and the pressure transducers are discussed in

Appendix A.106-111

The autoclave vessel was placed inside a PolyScience PP07R-40 refrigerated

circulating bath controlling the temperature to within ± 0.004 K. The stirring assembly was

controlled by an in-house assembled voltage regulation controller and a Hall Effect speed sensor.

C3H8 fluid was injected into the autoclave cell through an inlet high pressure valve, VA1. For

data acquisition, the setup was interfaced with Laboratory Virtual Instrument Engineering

Workbench (LabVIEW) which records the pressure and temperature of the system continuously

and averages every 30 seconds. The PRT and pressure transducer were calibrated by the supplier

and checked by comparison to the ice melting point and comparisons to a previously calibrated

Page 54: Thermodynamic Formation Conditions for Propane Hydrates in

34

transducer at different pressures as well as under a vacuum of 7105.2 MPa at different

temperatures.

Figure 2.4. Schematic diagram of the setup used for the measurement of C3H8 hydrate

dissociation points. VA1, VA2 and PT represent the inlet feed valve, outlet valve and pressure

transducer, respectively.

2.4.1 Interfacing experimental setup with LabVIEW for data acquisition.

The LabVIEW interface was used to merge and communicate with all electronic components of

the setup on a single operating window and thus enabled the automation of experiments while

controlling the temperature and pressure of the system for hydrate formation or dissociation. The

front panel of the experimental setup is shown in Figure 2.5. The indicators labeled “A” and “B”

are used to monitor the pressure and temperature plots versus time inside the autoclave and

throughout the experimental run. The button labelled “C” is use to execute the code for shutting

PolyScience water chiller bath

Vent

Pro

pan

e ta

nk

Autoclave

Impeller

Propane + degassed water

Glycol + water bath

Data logging

computer

VA2VA1

Platinum resistance

thermometer

PT

Page 55: Thermodynamic Formation Conditions for Propane Hydrates in

35

down the system automatically. The elapse interval for saving the time averaged data from the

experimental run is entered into “D” (normally set to 30 s) while the indicator “E” shows the

current temperature set-point during the experiment. The PolyScience circulating water bath is

restarted or shut down with the control knob “G” for temperature regulation of the experiment.

Each desired temperature set-point is entered alongside the duration of time required for each

temperature step. The pressure transducer, PRT and the circulator water bath are switch on or off

by the code executed by the control in box “I”. The temperature indicator or readout “J”

represents temperatures measured by the PRTs, i.e., temperatures measured at room condition

(not shown in Figure 2.4), inside the autoclave cell and water bath (shown in Figure 2.4). This

LabVIEW code allows the user to precisely repeat temperature and pressure conditions with ease

for checking reproducibility.

Figure 2.5. LabVIEW front panel for the experimental C3H8 hydrate dissociation setup used in

this study: A & B, graphic indicator showing the current temperature and pressure measured by

the PRT and pressure transducer respectively inside the autoclave cell; C, control to stop the

experiment run; D, time interval to record the averaged pressures and temperatures for each

experimental run; E, indicator showing the current set temperature of the experimental run; F,

experimental runs name; G, control knob used for stopping or restarting the chiller on the

PolyScience circulator water bath; H, automated temperature set-points program control; I,

control knobs for stopping or restarting the pressure transducer, PRT and the circulating water

bath; J, temperature readout indicator that measures the autoclave cell, room and circulating

water bath temperatures.

D

B

C

E

I

F

GH

A

J

Page 56: Thermodynamic Formation Conditions for Propane Hydrates in

36

2.5 Materials.

C3H8 with listed purities of 99.999% and 99.5% was supplied by Linde Canada Ltd. and Praxair

Inc., respectively. The purity and compositions of C3H8 gases were analyzed with a Bruker 450-

gas chromatograph (GC) equipped with a thermal conductivity detector (TCD) and a flame

ionization detector (FID).

Table 2.2. Measured gas impurities (mol %) in C3H8 used for this work.

Supplied by N2 CO2 CH4 C3H8 i-C4H10

Praxiar Inc 0.4094 0.002 0.006818 99.425 0.1573

Linde Ltd. 0.0000248 ND* 0.00138 99.999 ND*

ND* refers to as not detectable

All water used was taken from EMD Millipore model Milli-Q Type 1 water purification system

polished to a resistivity of 18 MΩ·cm and degassed under vacuum for at least 12 hours.

2.6 Experimental procedure

The PBD method was used to measure the hydrate dissociation conditions of C3H8. The

autoclave was place under a vacuum of 2.5 × 10-7

MPa for a period of 24 hours before each

experiment to flush out impurities from the system. Prior to an experiment, the apparatus was

leak tested by pressurizing the autoclave cell with C3H8 and waiting for 6 hours for pressure

stabilization. C3H8 gas was then flowed through the feed valve, VA1, as shown in Figure 2.4,

Page 57: Thermodynamic Formation Conditions for Propane Hydrates in

37

into the autoclave cell to purge any impurities that may still be trapped in the feed lines or cell

before loading with C3H8 to the desired pressure above the hydrate stability region. For the study

in the Lw–H–C3H8(g) phase boundary, ca. 10 cm3 of polished and degassed water was injected

into the evacuated autoclave by suction. The amount of water corresponds to a mole ratio of 77:1

for water to C3H8 so that liquid water phase was always in excess throughout the experiment. For

measurements in the Lw–H–C3H8(l) region, varying quantity of water was delivered to the

autoclave through a syringe pump after loading the autoclave to a desired pressure.

The C3H8-H2O mixture was then mixed for 8 hours until pressure was stable to within ± 0.005

MPa. Once the system had reached equilibrium, it was cooled and held at 273.35 K for 18 hours

to form hydrates. Figure 2.1, previously shown in sub-section 2.2.2.3, shows a typical curve for

gas cooling, hydrate formation and dissociation stages. A large pressure drop in the autoclave

vessel signifies the hydrate formation.7,94

The formed hydrate was then heated slowly in steps of

0.2 K along the Lw–H–C3H8(g) and 0.05 K along the Lw–H–C3H8(l) phase boundaries,

respectively, to obtain the dissociation points. The system was allowed to equilibrate for

approximately 4 hours at every step before recording the pressure and temperature. Typically,

increasing the temperature in the hydrate stability region causes a sharp increase of pressure due

to the release of enclathrated gas molecules into the gas phase, however, increasing the

temperature outside the hydrate stability region results in a smaller increase in pressure which is

as a result of gas expansion in the autoclave.7,77,94

The pressure versus temperature profile of the

experimental run for each of the data points along the Lw–H–C3H8(g) and Lw–H–C3H8(l) loci

reported in chapter three is shown in Appendix B.

Page 58: Thermodynamic Formation Conditions for Propane Hydrates in

38

CHAPTER THREE: Experimental Results and Modeling

3.1 Outline

C3H8 hydrate dissociation conditions were studied by using the phase boundary dissociation

method described in chapter two. Two purities (99.5 and 99.999 mol %) of C3H8 were used for

the study along the Lw–H–C3H8(g) phase boundary in order to investigate the effect of

impurities on dissociation pressures and temperatures. C3H8 with a listed purity of 99.999 mol %

was used for measurements in the Lw–H–C3H8(l) region. The results obtained from the

experiments were used to fit a Clausius–Clapeyron semi–empirical correlation and to calibrate

more rigorous equations used for thermodynamic modeling of dissociation pressure and

temperature. A mathematical description of the van der Waal and Platteuw model and the

reduced Helmholtz energy EOS used in this study for modeling the hydrate and the fluid phases

respectively are discussed. The algorithm for calculating the dissociation temperature is

presented as well. Finally, the model results were compared to the available literature data and

the deviations are discussed.

3.2 Thermodynamic modeling

In order to define the equilibrium between coexisting species, the partial molar free energy

(chemical potential or fugacity) of each individual species in each phase needs to be well defined

at relevant temperatures, pressures and molar compositions. By definition, equilibrium is

obtained when the free energy of each species in each phase is equal. For example, the

equilibrium condition of three different phases of L, V and H co-existing with each other can be

represented by equation 3.1,64

Page 59: Thermodynamic Formation Conditions for Propane Hydrates in

39

HiiLi ,V,, or HiiLi fff ,V,,

, (3.1)

where µ and f are the chemical potential and fugacity of component i, respectively. By solving

equation 3.1 iteratively, one can mathematically find the equilibrium conditions, i.e., pressure

and temperature, between H2O in a C3H8 gas and H2O which has been incorporated into a C3H8

gas hydrate. The hydrocarbon fluid (vapour and / or liquid C3H8) and hydrate phase fugacities for

this study were calculated by using the reduced Helmholtz energy EOS and the modified van der

Waal and Platteeuw model proposed by Chen and Guo (1996) because of the sound physical

background and high accuracy of these equations.85,90,112-114

3.2.1 Fluid phase modeling

The reduced Helmholtz energy EOS of Lemmon et al., (2009) was used for the calculation of

pure C3H8 thermodynamic parameters such as pressure, density, fugacity and saturation

properties in the fluid phase by using the Reference Fluid Thermodynamic and Transport

Properties (REFPROP 9.1) software.112,115

The equation can be applied from the triple point

temperature of C3H8, T = 85.525 - 650 K and for pressures up to 1000 MPa.112

The reduced

Helmholtz energy equations are composed of ideal and real gas contributions of the fluid. The

ideal gas terms consist of the ideal gas equation and a relation which is used to account for the

isobaric heat capacity at zero pressure, while the real gas contribution describes the residual

behaviour of the fluid.112,116-117

The equation formulated with the Helmholtz energy expressed as

a fundamental properties of density and temperature can be expressed as 112-114,116-117

RT

TATA

RT

TA r ,,, ο .),(),( ο r (3.2)

Page 60: Thermodynamic Formation Conditions for Propane Hydrates in

40

Where A, οA , Ar, ο and α

r represent the Helmholtz energy, ideal gas Helmholtz energy, pure

fluid residual Helmholtz energy, dimensionless form of ideal and real (residual Helmholtz

energy) gas contributions to the Helmholtz energy, respectively. The reduced temperature and

density are defined as )(T

Tc andc

)( . The ideal gas contribution to the Helmholtz energy

is given as

οοο TsRThA , (3.3)

where οs is the ideal gas entropy and οh represents the ideal gas enthalpy which is expressed as

dTchh

T

T

p

ο

οο

ο

ο . (3.4)

In equation 3.4, ο

οh and ο

pc denotes the ideal gas enthalpy and heat capacity at an arbitrary

reference temperature ( οT = 273.15 K), respectively. The ideal gas entropy is given as 112

T

T

p

T

TnRdT

T

css

0οο

ο

ο

ο

ο 1

, (3.5)

where ο represents the ideal gas density at an reference arbitrary pressure οp 0.001 MPa

and temperature οT = 273.15 K:

RT

p

ο

ο

ο . (3.6)

The ideal gas contribution to the Helmholtz energy can be expressed as equation 3.7 by

Page 61: Thermodynamic Formation Conditions for Propane Hydrates in

41

substituting equations 3.4 and 3.5 into equation 3.3

T

T

pT

T

pT

TnRdT

T

csTRTdTchA

ο

ο

οοο

ο

οοο 1

. (3.7)

Alternatively, equation 3.7 can be rewritten in a dimensionless and simplified form:

οο

ο

2

ο

ο

ο

ο

ο

ο

οο 111 d

c

Rd

c

Rn

R

s

RT

h pp

c

. (3.8)

For C3H8, the correlation used for calculating ο

pc in equation 3.8 was developed by fitting the

heat capacity experimental data of Trusler and Zarari (1996) which gives the relationship112,118

2

26

3

ο

1exp

)exp(4

k

kk

k

k

p

u

uuv

R

c, (3.9)

where u and v are coefficients derived from Einstein’s vibrational frequencies equation, which

are given as v3 = 3.043, v4 = 5.874, v5 = 9.337, v6 = 7.922, u3 = 393 K / T, u4 = 1237 K / T, u5 =

1984 K / T, u6 = 4351 K / T and R = 8.3144 J / mol / K.

The functional form of the ideal gas Helmholtz energy can be obtained by substituting equation

3.9 into equation 3.8, 112,114,117

k

i

i bnvaann

exp113116

3

21

ο, (3.10)

where

c

kk

T

ub . (3.11)

The general real gas contribution (residual Helmholtz energy) to the Helmholtz energy is

Page 62: Thermodynamic Formation Conditions for Propane Hydrates in

42

described with an empirical model which is expressed as a sum of the polynomial and

exponential terms: 112,116-117

)exp(,11

ExpPOI

POI

kkkkk

POIII

Ik

ldt

k

dtI

k

k

r NN . (3.12)

The ,r term for C3H8 contains an additional Gaussian term which helps to improve the

prediction of properties in the critical region and is expressed in equation 3.13 as 112

))((exp)exp(, 2218

12

11

6

5

1

kkkk

td

k

k

ldt

k

dt

k

k

r kkkkkkk NNN

.

(3.13)

The values of the parameters and coefficients kN , kt , kl , k , and k were obtained from the

nonlinear regression of the available experimental data for C3H8 vapour liquid equilibrium

(VLE) and pressure-density-temperature (p-ρ-T) by National Institute of Science and Technology

(NIST) which are shown in Appendix C.112

The thermodynamic properties for C3H8 + H2O mixtures are calculated by accounting for the

mixing of the two components which uses a generic mixing equation based on the corresponding

state principle. The Helmholtz energy of the mixture, A, is the sum of the ideal gas, real gas and

mixing or excess contributions which is expressed in the form113,116-117

Emixid AAA .

, (3.14)

Page 63: Thermodynamic Formation Conditions for Propane Hydrates in

43

where mixidA .

and EA denote the ideal mixture and mixing contributions to the Helmholtz energy

respectively. mixidA .

is the sum of the ideal gas Helmholtz energy (ο

iA ) and pure fluid residual

Helmholtz energy (r

iA ) of the component i, in the mixture which can be expressed in the form:

i

r

ii

n

i

i

mixid xRTATAxxTA n1,,) , ,( ο

1

.

, (3.15)

where n and ix represent total number of components and mole fractions of component, i, in the

mixture at temperature T and density ρ, respectively. The functionalized form of the ideal gas,

,ο and residual energy of pure fluid, ,r contributions to the Helmholtz energy are expressed

by equations 3.16 and 3.17 as

ο (ρ, T, x )

i

in

i

i xRT

TAx n1

,ο

1

, (3.16)

and

xxx En

i

r

ii

r ,,,,,1

. (3.17)

Where r

i is the residual term of the reduced Helmholtz free energy of component i which can

be calculated from equation 3.13 for C3H8 and the Wagner and Pruß (2002) equation for H2O.114

The excess contribution to the Helmholtz energy or departure function EA is not required for the

C3H8 + H2O system. The reduced density for a mixture, =)(xr

, and reduced temperature of

a mixture, =T

xTr )(, require mixing function by corresponding states:

114,116-117

Page 64: Thermodynamic Formation Conditions for Propane Hydrates in

44

3

3/1

,

3/1

,

2

,1 1

,,

11

8

1

)(

1

jcicjiijv

jin

i

n

j

ijvvji

r xx

xxxx

x ij

, (3.18)

and

jcic

jiijT

ji

ijTijTj

n

i

n

j

ir TTxx

xxxxxT ,,2

,

,,

1 1

)(

. (3.19)

In equations 3.18 and 3.19, ic, is the critical pressure for component i and jc, represents the

critical pressure of component j while icT , and jcT , represent critical temperatures of component

i and j respectively. The binary parameters used in the equations 3.18 and 3.19 for C3H8 are

shown in Table 3.1.

Table 3.1. Binary parameters of the reducing functions for density and temperature used in

equations 3.18 and 3.19. 116

Mixture i-j vij vij ijT , ijT ,

C3H8−H2O 1.0 1.011759763 1.0

0.600340961

Different derivative functions are formulated from the reduced Helmholtz energy equations for

calculating thermodynamic properties such as pressure, compressibility factor, speed of sound,

isochoric heat capacity, isobaric heat capacity, Gibb energy, internal energy, enthalpy and

entropy using differentiation with respect to density or temperature.114,116

The results obtained by

using the Helmholtz energy equations for pure C3H8 are accurate to within 0.01 to 0.03 % for

densities from T = 85.525 - 350 K, 0.5 % for heat capacities from T = 85.525 - 650 K.112

Page 65: Thermodynamic Formation Conditions for Propane Hydrates in

45

Accurate modeling of the hydrate phase conditions depend on the correct fugacities of the fluid

phases. If the fluid phases are not modeled correctly, this can lead to errors in the modeling of

the hydrate phase because the optimized hydrate parameters would now be based on the

erroneous fluid fugacities.89

The fugacity of component i, in a binary mixture can be calculated

from the expression: 64,113

),,(),,( npTpxnpTf iii . (3.20)

Where ),,( npT denotes the fugacity coefficient of component i in the mixture which can be

calculated from the relationship between molar derivate of r :113,116

),,( npTi =

jnvTi

r

n

nRT

,,

)(exp

. (3.21)

Equation 3.21 can be substituted into equation 3.20 to give

jnvTi

r

iin

nRTxnpTf

,,

)(exp),,(

(3.22)

where n is the number of moles in the mixture in component i and j. The explicit function of this

derivative and other derivatives are shown in Appendix D. Because fugacities for each

component in a gas mixture cannot be directly measured experimentally, experimentally

determined mole fractions of saturated water in C3H8 reported in the literature were compared to

calculated saturated mole fractions at the same conditions for temperature T = 235.55 - 399.89 K

and pressure p = 0.7720 - 67.3962 MPa in order to verify the accuracy of the mixing rules used

for calculating the fugacities of the components in the fluid phase.119-124

The mole fractions of

the saturated component are iteratively calculated from REFPROP 9.1 by using the solver

Page 66: Thermodynamic Formation Conditions for Propane Hydrates in

46

routine within Microsoft Excel. A correlation plot of the calculated mole fraction of saturated

water in C3H8 versus experimental literature data is shown in Figure 3.1. The results obtained by

using this equation are accurate to within an AAD of < 0.2 % with a coefficient of determination

(R2) value of 0.988.

Figure 3.1. Correlation plot comparing the literature experimental mole fraction of saturated

H2O in C3H8, (exp),832 HCOHy to those calculated using REFROP 9.1,

832, HCOHy (calc).

115 □, Song

and Kobayashi (1994);119

∆, Song et al.(2004); 120

○, Kobayashi and Katz (1953); 121

+, Sloan et

al. 1986; 122

*, Bukacek (1955). 123

3.2.2 Description of the hydrate phase

3.2.2.1 The Van der waal and Platteeuw hydrate model

Van der Waal and Platteeuw (vdWP) in 1959 proposed the first hydrate model based on the

statistical thermodynamic and Langmuir adsorption theory for calculating the chemical potential

of a hydrate phase based on some assumption described in the last section of chapter one.85

0.00

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.10

0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10

yH

2O

,C

3H

8(c

alc

)

yH2O , C3H8(exp)

Page 67: Thermodynamic Formation Conditions for Propane Hydrates in

47

The presence of guest molecules inside the hydrate cavity provides the cavity stability, so when

the majority of the cavities are unoccupied, the hydrate cavities dissociate and collapse. The

stability of the hydrate phase is measure as the chemical potential (µ) of water forming the

cavities.7,24

To develop a hydrate model, the hydrate formation process can be viewed as taking

place in two steps. The first step is to form a hypothetical empty hydrate cage from pure water

and the second stage is to fill the hydrate lattice with the former: 8

pure water (L

w ) → empty hydrate lattice (w ) → filled hydrate lattice(

H

w ).

The change in chemical potential accompanying this process is given as

LH

w

(L

w

H

w ) = (H

ww ) + (

w

L

w ), (3.23)

where H

w , w and

L

w represents the chemical potential of water in the filled hydrate lattice,

empty hydrate lattice, and pure liquid water respectively.

At equilibrium conditions, the chemical potential of water in the hydrate phase is equal to any

other coexisting phases present; H

w =L

w or w (ice).

7-8,24,86 By introducing

w to this

equilibrium condition, equation 3.24 is obtained:

L

w

H

w or w , (3.24)

where H

ww

H

w , w

L

w

L

w and www .

Van der waal and Platteeuw derived the change in chemical potential of water in a hydrate phase

and the hypothetical empty hydrate cage as 84

Page 68: Thermodynamic Formation Conditions for Propane Hydrates in

48

)1ln( j

jm

m

m

H

w vRT , (3.25)

where mv represents the number of cavities of type m per water molecule and jm is the

fractional occupancy of the guests molecules j within the hydrate cavities m.

3.2.2.2 Calculation of hydrate phase fugacity

Many attempts have been made to improve the accuracy of the vdWP model over the years. The

breakthrough work of Prausnitz and Parrish in 1972 simplified the use of the vdWP model and is

employed within several commercial computer programs.86

Other modifications to the vdWP

model have been discussed in section 1.5. The modified vdWP model by Chen and Guo (1996)

was used for modeling the hydrate phase in this stability here.90

The model was based on the

concept of equality of fugacities of water for all the phases present at equilibrium as shown in

equation 3.26 rather than chemical potential because of the relative ease of calculating fugacities

for gas mixture components:

H

wf = L

wf or

wf , (3.26)

where H

wf ,L

wf and

wf represents the fugacity of water in hydrate, liquid water and ice phase

respectively. is related to f by 67

ο

ο 1f

fnRT , (3.27)

where the subscript ○ denotes a reference state which is taken as the ideal solution. It follows

that H

wf can be expressed in term of H

w

from equation 3.27,

90

Page 69: Thermodynamic Formation Conditions for Propane Hydrates in

49

RTff

H

w

w

H

w

exp , (3.28)

where

wf is the reference fugacity of the empty hydrate cavity which is expressed as

RTff

L

wL

ww

exp . (3.29)

L

wf was calculated from the Wagner and Pruß (2002) reduced Helmholtz energy EOS.114

Classical

thermodynamics can be used to derive the expression for L

w

, where the simplified method

of Holder et al. (1980) was used for calculating L

w

by directly integrating over pressure and

temperature while using hexagonal ice (ice Ih) water as a reference point from the relationship88

ww

T

T

p

p

www

L

w xdpRT

vdT

RT

h

RTRT

ln

ο ο

2

ο

ο

, (3.30)

where ο

w is the experimentally determined reference chemical potential difference between

water in the empty hydrate lattice and pure water (L) or the ice (α) phases, at an arbitrary

reference temperature T○ (T○ = 273.15 K) and absolute zero pressure οp . wv represents the

reference volume difference between the empty hydrate cage and pure ice water phase. The

volume of the hydrate lattice does not change at low pressures, p < 20 MPa, and ice Ih can be

used as a reference lattice at this condition for estimating wv .86,89

The last term, ww xln , in

equation 3.30, is use to account for the deviation in the chemical potential of a pure liquid or ice

water relative to a water rich mixture.88,125

The activity coefficient, ,w is normally assumed to be

unity unless an inhibitor or a highly soluble gas is present, xw denotes the mole fraction of water

in the liquid water phase, while wh represents the molar enthalpy difference between the empty

Page 70: Thermodynamic Formation Conditions for Propane Hydrates in

50

hydrate lattice and liquid water phase. The molar enthalpy difference ( wh ) between the empty

hydrate lattice and liquid water is temperature dependant and is expressed as88

dTchh

T

T

pwww

ο

ο , (3.31)

where ο

wh is the enthalpy difference between the empty hydrate lattice and ice, at T = 273.15 K

and zero pressure. The change in heat capacity ( pwc ) between the empty hydrate and pure

water phases also depends on temperature:

οο

ο TTbTcc pwpw , (3.32)

whereο

pwc is the reference standard difference in heat capacity between ice and liquid water at

temperatures above 273 K and b represent the coefficient of temperature correction. Table 3.2

shows the values of constants in equation 3.30, 3.31 and 3.32 used in this study.

Table 3.2. Thermodynamic reference properties for structure II used in this study.

Reference

Parameter

Structure II

Source

ο

wh 1025 J mol-1

Dharmawardhana et al.126

ο

w 883.8 J mol-1

Sloan7

ο

pwc -38.13 J mol-1

K-1

Holder et al.88

mv 3.4 cm3 mol

-1 Parrish and Prausnitz

86

b 0.141 Holder et al.88

Page 71: Thermodynamic Formation Conditions for Propane Hydrates in

51

The equality of coexisting phase fugacities in equation 3.26 is used to solve for the equilibrium

pressure and temperature in this study. H

wf and L

are calculated from equation 3.28 and

the relationship developed by Holder et al. in equation 3.30 respectively.85,88

3.2.2.3 Hydrate cage occupancy

The fractional occupancy of the gas molecule within the cavities is calculated by using the

Langmuir adsorption theory which is expressed as 85-86,88-92

jm j jjm

jjm

fC

fC

1, (3.33)

where jf is the fugacity of hydrate former j in cavity m, which was calculated from the reduced

Helmholtz energy EOS described in section 3.3.1. The Langmuir constant ( jmC ) is used to

measure the attraction between the enclathrated gas and water molecules in the cavity and is

given as

drrTK

rw

TKC

BB

jm

2

0

)(exp

4

, (3.34)

where KB, w(r) and r represents the Boltzmann constant, cell potential function (average

resulting field of the enclathrated gas molecules in all position within the cavity), and the

distance between the centre of the encaged gas and water molecules respectively. Van der waal

and Platteuw calculated the contribution to the potential energy due to the interaction of the guest

molecules within the cavity by using the Lennard-Jones 6–12 potential.85

However, McKoy and

Sinanoglu (1963) suggested that the Kihara potential with a spherical hard core provides a better

estimate for the gas-water interactions within the cavity.87

By first considering a gas-water

Page 72: Thermodynamic Formation Conditions for Propane Hydrates in

52

molecule interaction and assuming that the core diameter of water is zero. The potential energy

for the interaction )(r is given by

,)( r for ar 2 (3.35)

and

)(r = 4

,

612

22

arar

for r ˃ 2a, (3.36)

where a2 is the collision diameter, i.e., the distance at which )(r = 0. and a represents

the characteristic energy and spherical molecular core radius respectively. McKoy and Sinanoglu

summed the interaction between the gas and water molecules within the cavities and give the

relationship86-87,90

511

4

5

610

11

12

2)(

R

a

rRR

a

rRzrw (3.37)

where

NN

N

R

a

R

r

R

a

R

r

N11

1 (3.38)

and N can be 4, 5, 10 or 11, z is the coordination number (number of oxygen atoms at the

periphery of the cavity), R is the cavity radius and r represents the distance between the gas

molecule from the center of the cavity. The Kihara cell potential parameters (a, and ) can be

determined in two ways:127

(i) from gas viscosity and second virial coefficient data for pure

substances (ii) by correlating gas hydrates experimental dissociation data to the Kihara potential

parameters. For gas molecules such as C3H8 that only occupy the large cages of structure II, the

Langmuir constant can also be estimated from the relationship developed by Bazant and Trout

Page 73: Thermodynamic Formation Conditions for Propane Hydrates in

53

(2001) which relates the calculated fluid phase fugacity and the experimentally determined

change in chemical potentials via the relationship7,128-129

jmC (T) = j

B

H

w

f

TK1

1

17exp

(3.39)

where H

w

and jf represent the chemical potential difference between a hydrate phase and

empty hydrate cage and the fugacity of the hydrate former respectively.

Based on recommendations made by Mckoy and Sinanoglu, Parrish and Prausnitz (1972)

reported better estimates for dissociation pressures and temperatures that were close to

experimental dissociation conditions by using the Kihara potential for calculating gas-water

molecule interactions for the hydrate modeling of multi-component gases.86-87

They proposed an

equation for calculating the Langmuir constant by correlating the experimental dissociation data

for pure hydrate formers and gave the relationship86

T

B

T

ATC

mjmj

jm exp)( , (3.40)

where mjA and mjB are the fitting parameters related to guest type j in type m cavity at

temperature T. Their parameters were re-optimised with the experimental data measured here to

give a better correlation of experimental conditions.

3.2.2.3.1 Optimization of Kihara potential parameters

The parameters of Parrish and Prausnitz (610992.999 jmA K / MPa and 48.3794jmB K)

given by Karakatsani and Kontogeorgis (2013) were adjusted by minimizing the difference

Page 74: Thermodynamic Formation Conditions for Propane Hydrates in

54

between the fugacities of H2O in the C3H8 phase and hydrate phase.130

The optimized parameters

were determined by minimization of the following objective function using the least square

regression method expressed as

2

,83

2.

NP

i

H

wHCOH ffFObj . (3.41)

The optimized values for jmA and jmB are shown in Table 3.3 and can be used to iteratively

solve for the formation temperature at any pressure.

Table 3.3. Optimised Kihara potential paramaters used for this study.

Phase boundary

610jmA (K / MPa)

jmB (K)

Lw–H–C3H8(g)

999.992

10719.292

Lw–H–C3H8(l)

999.992

23266.346

3.3. Algorithm for calculating equilibrium hydrate formation temperature

To obtain the temperature that gives equilibrium between the hydrate phase and liquid water

phases at a given pressure, one solves equations 3.28 by iteration until the equality of fugacities

in equation 3.26 is satisfied. The flowchart of the steps followed in Microscoft Excel Visual

Basic Application for the estimation of the equilibrium hydrate formation temperature is shown

in Figure 3.3.

Page 75: Thermodynamic Formation Conditions for Propane Hydrates in

55

Figure 3.2. Simplified flowchart for the calculation of dissociation temperature used for the

thermodynamic modeling in this study.

Pressure p is first input and an initial temperature T is guessed (this is currently done using an

empirical fit as an ancillary equation). The program proceeds to execute the next four steps by

first calculating the mole fractions of the two fluid phases (saturated H2O in C3H8, 83,2 HCOHy and

saturated C3H8 in H2O, OHHCy 2,83 ), along with their respective fugacities i.e.,832

, HCOHf and

OHHCf283

, at the guessed temperature and specified pressure from REFPROP 9.1. The Langmuir

constant, jmC , and fugacity of water in the hydrate phase, H

wf , are calculated by using equations

Input p

Is hydrate formation

possible?No

Return error message

Yes

Calculate

No

Calculate: yH2O, C3H8 , yC3H8, H2O ,

fH2O , C3H8 & fC3H8, H2O at p and guessed T

H

wf

Guess T

Calculation of Cjm

Solve Abs | fH2O,C3H8 –H

wf

Return T & p

Assign new T

Yes

| ≤ 10-10

Page 76: Thermodynamic Formation Conditions for Propane Hydrates in

56

3.40 (with the optimized jmA and jmB in Table 3.4) and 3.28 respectively. Temperature is

optimized using the solver routine within Excel.

3.4 Experimental results and discussion

C3H8 with listed purities of (99.5 and 99.999) mol % were used for the study in the Lw–H–

C3H8(g) locus for temperatures T = 273.63 - 278.63 K and pressure p = 0.1887 - 0.5774 MPa

while 99.999 mol % C3H8 was used for measurements along the Lw–H–C3H8(l) phase boundary

for pressures p = 0.5717 - 18.2622 MPa and temperature T = 278.64 - 278.75 K. Previous

studies conducted using the experimental setup used for this study have demonstrated the

accurate determination of equilibrium conditions of CH4 and H2S gas hydrates.96

The

experimental dissociation conditions of C3H8 hydrates for the two purities (99.5 and 99.999 mol

%) along the Lw–H–C3H8(g) phase boundary measurements are shown in Table 3.4. The

dissociation pressure of 99.5 mol % C3H8 is, on average, 0.015 MPa larger than that of the

99.999 mol % C3H8 for the same range of temperature T = 273.63 - 278.63 K and this can be

attributed to the presences of impurities like N2, CO2 and CH4 (assessed by GC TCD/FID as

shown in repeated Table 2.6 below). These molecules can occupy both the small (512

) and large

(512

64) cages of a sII hydrate but have a higher propensity for occupying the small cages.

C3H8,

on the other hand, only occupies the large cages (512

64) in type II hydrate leaving the small cages

(512

) empty. While a larger fraction of cage occupancy by some impurities can result in further

stabilization of the hydrate crystal structure (lower dissociation pressure), impurities in the fluid

phase also lead to destablitization (higher pressure stabilization of the fluid phase).7-9

In this case,

the freezing-point depression is small but apparent; whereas, if the impurities were species such

Page 77: Thermodynamic Formation Conditions for Propane Hydrates in

57

as H2S, one would expect the opposite effect. The experimental dissociation conditions for

99.999 mol % C3H8 in the Lw–H–C3H8(l) phase boundary are shown in Table 3.5 and Figure 3.3

presents the summary of the measured conditions for this study with literature data and

calculated values predicted by the model along the Lw–H–C3H8(g) and Lw– H– C3H8(l) phase

boundaries.70-84

Table 2.6. Measured gas impurities (mol %) in C3H8 used for this work. (Repeated)

Supplied by N2 CO2 CH4 C3H8 i-C4H10

Praxiar Inc 0.4094 0.002 0.006818 99.425 0.1573

Linde Ltd. 0.0000248 ND* 0.00138 99.999 ND*

ND* refers to as not detectable

Page 78: Thermodynamic Formation Conditions for Propane Hydrates in

58

Table 3.4. Experimental dissociation conditions for C3H8 hydrates along the Lw–H–C3H8(g)

phase boundary.

99.999 mol % C3H8 99.5 mol % C3H8

T / Ka p / MPa

b T / K

a

p / MPab

273.63 0.1887 273.63 0.2052

273.83 0.1957 273.83 0.2130

274.03 0.2038 274.03 0.2212

274.23 0.2131 274.23 0.2298

274.42 0.2223 274.43 0.2398

274.62 0.2318 274.63 0.2489

274.83 0.2420 274.83 0.2589

275.02 0.2524 275.03 0.2698

275.22 0.2637 275.23 0.2799

275.43 0.2758 275.43 0.2929

275.63 0.2882 275.63 0.3048

275.83 0.3005 275.83 0.3175

276.03 0.3135 276.03 0.3305

276.22 0.3280 276.23 0.3441

276.42 0.3434 276.43 0.3582

276.62 0.3594 276.63 0.3731

276.82 0.3754 276.84 0.3887

277.03 0.3927 277.04 0.4062

277.22 0.4109 277.23 0.4226

277.42 0.4302 277.44 0.4404

277.63 0.4501 277.63 0.4637

277.83 0.4709 277.83 0.4840

278.03 0.4939 278.03 0.5045

278.23 0.5167 278.23 0.5262

278.43 0.5408 278.43 0.5501

278.62 0.5654 278.63 0.5774 a

Uncertainty for hydrates temperature measurements using the calibrated PRT

was estimated to be ± 0.1 K. b

Uncertainty for the hydrates pressure measurements was estimated to be

±0.0069 MPa.

Page 79: Thermodynamic Formation Conditions for Propane Hydrates in

59

Table 3.5. Experimental dissociation conditions for 99.999 mol % C3H8 hydrates along the Lw–

H–C3H8(l) phase boundary.

T / Ka p / MPa

c

278.64 18.2622

278.65 15.9072

278.65 13.3852

278.64 13.3478

278.68 12.4553

278.67 11.9547

278.68 11.6402

278.69 10.5883

278.69 9.4884

278.68 7.2316

278.72 5.5831

278.73 4.2907

278.74 2.2420

278.75 2.0535

278.75 1.0952

278.75 0.8096

278.75 0.7855

278.75 0.5717

aUncertainty for hydrates temperature measurements using

the calibrated PRT was estimated to be ± 0.1 K. c

Uncertainty for the hydrates pressure measurements was

estimated to ± 0.001MPa.

The experimental data for the two phase boundaries also were used to fit a semi-empirical

correlation based on the Clausius-Clapeyron relation for the rapid calculation of the hydrate

formation conditions. The Lw-H-C3H8(g) locus can be calculated using:

0014.2597.27150

5778.0ln T

Tp . (3.42)

Page 80: Thermodynamic Formation Conditions for Propane Hydrates in

60

Similarly, the Lw–H–C3H8(l) phase boundary also can be calculated from the relationship:

p = 36704 – 131.668T. (3.43)

where p and T are pressure and temperature in MPa and K respectively.

Figure 3.3. Pressure versus temperature for the Lw–H–C3H8(g) and Lw–H–C3H8(l) phase

boundaries (experimental and model). ____

, model; ----, vapour pressure of pure C3H8 calculated

with the reduced Helmholtz energy equation using REFPROP 9.1,115

, this study (99.5 % C3H8);

, this study (99.999 % C3H8); □, Reamer et al.(1952);70

; +, Tumba et al.(2014);71

*, Verma

(1974);72

♦, Engelos and Ngan (1993);73

●, Robinson and Mehta (1976);74

+, Patil (1987),75

■;

Kubota et al.(2003), 76

; ♦, Deaton and Frost (1946);77

■, Thakore and Holder (1987);78

◊, Den

Heuvel et al. (2001);79

▲, Nixdorff (1997);80

○,Wilcox et al.(1941);81

▬, Miller and Strong

(1946);82

∆, Makogon(2003);83

●, Maekawa (2008).84

p/

MP

a

T / K

0.10

1.00

10.00

100.00

275.0 275.5 276.0 276.5 277.0 277.5 278.0 278.5 279.0

Page 81: Thermodynamic Formation Conditions for Propane Hydrates in

61

3.4.1 Model comparison to experimental and literature data along the Lw–H–C3H8(g) region.

The experimental data for 99.999 mol % C3H8 were used to calibrate and validate the

thermodynamic model along the Lw–H–C3H8(g) phase boundary. The deviations between the

temperatures estimated by the model ( calcT ) and experimental temperature ( expT ) for pressure

p = 0.2524 - 0.5654 MPa are reported in Table 3.6.

Table 3.6. Model comparison to the experimental data along the Lw–H–C3H8(g) locus.

expp Texp Tcalc Texp-Tcalc

0.2524 275.02 275.07 -0.05

0.2637 275.22 275.26 -0.04

0.2758 275.43 275.47 -0.04

0.2882 275.63 275.66 -0.03

0.3005 275.83 275.85 -0.02

0.3135 276.03 276.04 -0.01

0.3280 276.22 276.24 -0.02

0.3434 276.42 276.44 -0.02

0.3594 276.62 276.64 -0.02

0.3754 276.82 276.83 0.01

0.3927 277.03 277.02 0.01

0.4109 277.22 277.22 0.00

0.4302 277.42 277.41 0.01

0.4501 277.63 277.60 -0.03

0.4709 277.83 277.79 0.04

0.4939 278.03 277.99 0.04

0.5167 278.23 278.18 0.05

0.5408 278.43 278.37 -0.06

Page 82: Thermodynamic Formation Conditions for Propane Hydrates in

62

The calcTT exp for each experimental datum compared was found to be within the ± 0.1 K

estimated uncertainty for experimental temperature measurements for pressures on the Lw–H–

C3H8(g) phase boundary. The number of literature data, purities, average deviations (AD),

pressure and temperature ranges compared to the model in this study are presented in Table 3.7.

The pressure versus temperature plot of these experimental results, model, empirical correlations

and literature data for the Lw–H–C3H8(g) locus is shown in Figure 3.4. The visual representation

of the deviation between the model and the other data (literatures and correlations) for pressures

p = 0.2524 - 0.5408 MPa is shown in Figure 3.5.70-84

Page 83: Thermodynamic Formation Conditions for Propane Hydrates in

63

Table 3.7. Summary of literature data along the Lw-H-C3H8(g) phase boundary compared.

T / K

p / MPa

No. of

data

compared

Purity

(mol %)

AD* (K)

Source

275.7 - 278.6 0.305 - 0.414 3 >99 -0.14 Reamer et al.70

274.6 – 278.0 0.2069 - 0.5100 4 99.5 -0.20 Patil75

276.4 - 278.1 0.3415 - 0.5006 2 99.5 0.01 Tumba et al.71

276.37 - 278.87 0.3309 - 0.5516 3 99.5 0.13 Robinson and Mehta74

274.95 - 278.30 0.2656 - 0.5353 5 99.5 -0.20 Engelos and Ngan73

274.63 - 278.63 0.2489 - 0.5501 21 99.5 -0.18 This study

275.1 - 278.4 0.2500 - 0.5620 7 >99.5 -0.01 Verma72

276.15 - 278.45 0.3230 - 0.5520 7 >99.5 -0.02 Kubota et al.76

275.37 - 277.04 0.1827 - 0.3861 3 99.8 0.05 Deaton and Frost77

275.15 - 278.15 0.2169 - 0.5099 3 99.9 0.08 Thakore and Holder78

276.77 - 278.55 0.3840 - 0.5658 10 99.95 -0.08 Den Heuvel et al.79

275.49 - 278.43 0.2774 - 0.5490 7 >99.995 0.02 Nixdorff80

275.3 - 278.1 0.2670 - 0.5090 10 99.999 -0.03 Maekawa84

276.8 – 278.0 0.3650 - 0.4720 5 - 0.13 Miller and Strong82

275.0 - 278.6 0.2513 - 0.5616 19 - -0.14 John Carrol8

275.0 - 278.6 0.2513 - 0.5616 19 - -0.29 Kamath correlation69

275.0 - 278.6 0.2513 - 0.5616 19 99.999 -0.02 Maekawa84

275.0 - 278.6 0.2513 - 0.5616 19 99.999 -0.01 This study correlation

* )(1

exp calc

n

TTn

AD , n is the number of data point compared.

Page 84: Thermodynamic Formation Conditions for Propane Hydrates in

64

Figure 3.4. Pressure versus temperature plot of experimental results, model, empirical

correlations and literature data along the Lw–H–C3H8(g) locus. , this study (99.5 % C3H8); ,

this study (99.999 % C3H8); □, Reamer et al.(1952);70

+, Tumba et al.(2014);71

*, Verma,

(1974);72

♦, Engelos and Ngan (1993);73

●, Robinson and Mehta (1976);74

+, Patil (1987);75

■,

Kubota et al.(2003);76

♦, Deaton and Frost (1946);77

■, Thakore and Holder (1987);78

◊, Den

Heuvel et al.(2001);79

▲, Nixdorff (1997);80

○, Wilcox et al.(1941);81

▬, Miller and Strong

(1946);82

, Maekawa (2008);84

……., Kamath correlation (2008);69

––––, this study model;

-----, Carrol correlation (2003);8 -------, this study Clausius-Clapeyron equation; ------, Maekawa

correlation (2008).84

p/

MP

a

T / K

0.20

0.25

0.30

0.35

0.40

0.45

0.50

0.55

0.60

0.65

275.0 275.5 276.0 276.5 277.0 277.5 278.0 278.5

Page 85: Thermodynamic Formation Conditions for Propane Hydrates in

65

Figure 3.5. Temperature difference between the model and experimental data, literature data and

correlations along the Lw-H-C3H8(g) locus. , this study (99.5 % C3H8); , this study( 99.999 %

C3H8); □, Reamer et al.(1952);70

+ , Tumba et al. (2014);71

*, Verma (1974);72

♦, Engelos and

Ngan (1993);73

●, Robinson and Mehta (1976);74

+, Patil (1987);75

■, Kubota et al.(2003);76

, Deaton and Frost (1946);77

■, Thakore and Holder (1987);78

◊, Den Heuvel et al.(2001);79

▲,

Nixdorff (1997);80

○,Wilcox et al. (1941);81

▬, Miller and Strong (1946);82

, Maekawa

(2008);84

……., Kamath correlation (2008);69

––––, this study model;-------, Carrol correlation

(2003);8 -----, this study Clausius-Clapeyron equation; , Maekawa correlation (2008).

84

Tex

p−

Tca

lc/ K

p / MPa

-0.50

-0.40

-0.30

-0.20

-0.10

0.00

0.10

0.20

0.30

0.26 0.31 0.36 0.41 0.46 0.51 0.56 0.61

Page 86: Thermodynamic Formation Conditions for Propane Hydrates in

66

The relationship between the purities and variance of the literature data to the model is shown in

Figure 3.6.70-84

Generally, the higher the purity of C3H8 reported in the literature for

measurements along the Lw–H–C3H8(g) region, the lower the deviation in pressure and

temperature from this study’s model, the only exception to this trend was observed for the data

reported by Tumba et al. (2014) for the 2 experimental points.71

The temperature deviation

between their two values and the model presented in this study was unusually small and lower

than the deviations observed for other literature data with similar purities, i.e., 99.5 mol % C3H8

and some higher purities. Engelos and Ngan (1993), Robinson and Mehta (1976), Patil (1987)

and this study reported dissociation data using 99.5 % mol C3H8; however, those authors did not

report any analysis of impurities.73-75

The data all show a similar AD of between -0.13 and -0.27

K for pressures, p = 0.2069 - 0.5516 MPa when compared to the model in this study. The highest

purity of C3H8 in Lw–H–C3H8(g) phase boundary reported in literature was 99.999 mol % by

Maekawa (2008), which is the same as that used in this study. As expected, the data were

comparable to this model presented in this study to within AD = 0.02 K.84

Also, similar

deviations were observed for data obtained by Nixdorff and Oellrich (1997), for > 99.995 mol %

C3H8.80

The temperatures predicted by Carroll’s and Kamath’s empirical correlations deviate

from this model as the pressure increases from 0.2513 to 0.5616 MPa; although, the Kamath

correlation tends to predict a higher dissociation average temperature T = ~ 0.25 K than the

Carroll correlation for the same pressure.8,69

The Maekawa correlation compared favourably with

this model to within an average temperature of within ± 0.02 K.84

Page 87: Thermodynamic Formation Conditions for Propane Hydrates in

67

Figure 3.6. Relationship between C3H8 purities and variance of the literature data along the Lw-

H-C3H8(g) locus to the model presented in this study ──, this study model; , this study (99.5

mol % C3H8); , this study (99.999 mol % C3H8); □, Reamer et al. (1952);70

∆; Tumba et al.

(2014);71

■, Verma (1974);72

×, Engelos and Ngan, (1993);73

●, Robinson and Mehta (1976);74

+,

Patil (1987);75

♦, Kubota et al.(2003); 76

♦, Deaton and Frost (1946);77

■, Thakore and Holder

(1987);78

◊, Den Heuvel et al. (2001);79

▲, Nixdorff (1997);80

○, Maekawa (2008).84

3.4.2 Model comparison to experimental and literature data along the Lw–H–C3H8(l) region.

The measured and calculated conditions with their corresponding deviations along the Lw–H–

C3H8(l) phase boundary are presented in Table 3.8. The model predicts the dissociation

temperature to an AD = 0.01 K and to within the uncertainty of the experimental temperature

measurements (± 0.1 K). The pressure versus temperature plot of this study’s dissociation

conditions, including the model and literature data, on the Lw–H–C3H8(l) locus are presented in

Figure 3.7.

-1.0

-0.5

0.0

0.5

1.0

98.5 99.0 99.5 100.0

Tex

p−

Tca

lc/ K

Propane purity / mol %

Page 88: Thermodynamic Formation Conditions for Propane Hydrates in

68

Table 3.8. Model comparison to the experimental data along the Lw–H–C3H8(l) phase boundary.

pexp Texp Tcalc Texp-Tcalc

18.2622 278.64 278.63 0.01

15.9072 278.65 278.64 0.01

13.3852 278.65 278.66 -0.01

13.3478 278.64 278.66 -0.02

12.4553 278.68 278.66 -0.02

11.9547 278.67 278.66 0.01

11.6402 278.68 278.67 0.01

10.5883 278.69 278.67 0.02

9.4884 278.69 278.67 0.02

7.2316 278.68 278.68 0.00

4.2907 278.71 278.69 0.02

2.2420 278.74 278.70 0.04

2.0535 278.75 278.70 0.05

1.0952 278.75 278.69 0.06

0.8096 278.75 278.69 0.06

0.7855 278.75 278.69 0.06

0.5717 278.75 278.69 0.06

Page 89: Thermodynamic Formation Conditions for Propane Hydrates in

69

Figure 3.7. The pressure versus temperature plot of this study’s dissociation conditions, model

and literature data along the Lw–H–C3H8(l) locus. , this study model; , this study (99.999 mol

% C3H8); *, Verma (1974);72

◊, Den Heuvel et al.(2001);79

○, Wilcox et al.(1941);81

∆, Makogon

(2003);83

-----, this study Clausius-Clapeyron equation.

This locus leans towards lower temperature as pressure increases, as opposed to higher

temperatures as is reported in some literature.79,89

This similar pattern also was observed by

Dyadin et al. (2001) and Makogon (2003), although, at lower temperatures than the temperatures

reported for this study.84,131

This behaviour can be attributed to lower density of formed C3H8

hydrates in the liquid water and liquid C3H8 phase which causes the hydrates to remain afloat in

these coexisting phases. This behaviour is similar to the behaviour of hexagonal ice in liquid

0

5

10

15

20

25

30

35

40

277.0 277.5 278.0 278.5 279.0

p/

MP

a

T / K

Page 90: Thermodynamic Formation Conditions for Propane Hydrates in

70

water phase, whereby the ice melting line tends towards lower temperatures due to its lower

density (increase in volume upon solidification).

The number of data points, purities, average deviation, pressure and temperature ranges

compared to the model presented in this study along the Lw–H–C3H8(l) locus are presented in

Table 3.9, while the deviations in temperature of the model presented in this study to the

literature data are presented in Figure 3.8.70,72,79,81,83

Table 3.9. Summary of literature data and corresponding purities along the Lw–H–C3H8(l) phase

boundary.

T / K

p / MPa

No. of

data points

Purity (%)

AD (K)

Source

278.05 - 278.28 0.555 - 35.00 9 99.95 -0.46 Makogon83

278.55 - 278.88 0.643 - 9.893 17 99.95 0.08 Den Huevel et al.79

278.2 - 278.6 0.562 - 11.30 4 > 99.50 -0.20 Verma72

278.6 - 278.8 0.684 - 2.046 3 > 99.00 0.01 Reamer et al.70

278.6 - 279.2 0.807 - 6.115 7 - 0.14 Wilcox et al.81

Page 91: Thermodynamic Formation Conditions for Propane Hydrates in

71

Figure 3.8. Hydrate dissociation temperature difference between the model in this study to the

literature data along the Lw-H-C3H8(l) locus. ____, this study model; , this study (99.999 mol %

C3H8); *, Verma (1974);72

◊, Den Heuvel et al. (2001);79

+, Wilcox et al.(1941);81

∆, Makogon

(2003),83

; ----, this study Clausius-Clapeyron equation.

The Clausius-Clapeyron equation compares favourably to the model presented in this study to

within an AD = 0.01 K for pressure ranges p = 0.5717 – 18.2622 MPa. The literature data in the

Lw–H–C3H8(l) region were all within an AD = 0.2 K compared to this model except for the data

reported by Makogon (2003).70,72,79,81,83

As opposed to the Lw–H–C3H8(g) region, temperature

deviation from this model does not correlate with the C3H8 purity. For example, Makogon (2003)

and den Huevel et al’s (2001) reported using 99.995 mol % C3H8 purity. While den Huevel et

al., data shows a significantly lower deviation, comparable to this model. The Makogon (2003)

data show a very large deviation to the model presented in the study.79,83

Tex

p−

Tca

lc/

K

p / MPa

-0.80

-0.60

-0.40

-0.20

0.00

0.20

0.40

0.60

0.80

0 5 10 15 20 25

Page 92: Thermodynamic Formation Conditions for Propane Hydrates in

72

The Reamer et al. (1952) data show the lowest deviation to this model for > 99 mol % C3H8. The

purity of C3H8 used by Wilcox et al. (1941) was not reported but the data were still comparable

to the model of this study to with AD = 0.2 K.70,81

3.4.3 Comparison of Upper Quadruple points of this study and literature.

The upper quadruple point, Q2, for this study was calculated from the point of intersection of

Lw-H-C3H8(g) and Lw-H-C3H8(l) loci.

Figure 3.9. A graphical representation of upper quadruple point determination from the point

of intersection of the Lw-H-C3H8(g) and Lw-H-C3H8 (l ) loci.

0.40

0.50

0.60

0.70

0.80

0.90

1.00

278 278.2 278.4 278.6 278.8 279

C3H8 vapour pressure

p /

M

Pa

T / K

Upper quadruple point (Q2)

Lw

–H

–C

3H

8(l

) lo

cus

Page 93: Thermodynamic Formation Conditions for Propane Hydrates in

73

The calculated pressure, 2Qp = 0.5591 MPa, and other quadruple pressures are shown in Table

3.10 below.8,72,74,79,83

All the 2Qp in the literature fall within the uncertainty of this study

measurements ( p 6.9 × 10-3

MPa) except the pressure reported by den Heuvel et al.

(2001).79

Similarly, most of the quadruple point temperatures, 2QT , reported in the literature

falls within 95 % confidence interval, 2QT =278.68 ± 0.1 K except that of Robinson and Mehta

(1974).

Table 3.10. Quadruple points conditions from this study and literature.

Source

% purity

p / MPa

T / K

Makogon83

99.95 0.555 278.3

Robinson and Mehta74

99.5 0.5516 278.87

Den Huevel et al.79

99.95 0.6 278.62

Carroll8 - 0.556 278.75

Verma72

- 0.562 278.4

This study

99.999

0.5591 ± 0.0069

278.68 ± 0.10

Page 94: Thermodynamic Formation Conditions for Propane Hydrates in

74

CHAPTER FOUR: Conclusion, Recommendation and Future work

4.1 Conclusion

The measured formation conditions for C3H8 hydrate in equilibrium with gaseous and liquid

C3H8 were reviewed, where it was found that the liquid C3H8 region showed a large variance in

the literature measurements. Due to the industrial importance of C3H8 hydrates and because C3H8

hydrate is a reference material for other sII hydrates, the C3H8 hydrate dissociation conditions

were independently measured using the phase boundary dissociation method for T = 273.63 -

278.75 K and p = 0.1887 - 18.2622 MPa. Along the Lw-H-C3H8(g) phase boundary, two

different purities of 99.5 and 99.999 mol % C3H8 were used. The higher pressure dissociation

data reported for 99.5 mol % C3H8 were attributed to stabilization of the fluid phase with respect

to the hydrate due to the presence of fluid phase impurities. C3H8 with a listed purity of 99.999

mole % was used for study of the Lw-H-C3H8(l) locus. A thermodynamic-based model was

optimized using the high-purity data. The model agrees with the experimental data to within the

estimated uncertainty of T ± 0.1 K.

The optimized thermodynamic model also was compared to the available literature data along

the two phase boundaries.7,69-84

Even small amount of impurities were found to be important

when studying C3H8 hydrate dissociation conditions, where larger deviations from the model

reported in this study were observed for the studies that used lower purity C3H8 for

measurements on the Lw-H-C3H8(g) locus. The curve of the Lw-H-C3H8(l) locus also shows an

inclination towards lower temperatures with increasing pressure as opposed to higher

temperatures which has been reported by other studies.84,132

This inclination is expected to be

towards lower temperatures when the densities of C3H8 hydrates are lower than the densities of

Page 95: Thermodynamic Formation Conditions for Propane Hydrates in

75

the other two coexisting phases (liquid water and liquid C3H8). Similar to the literature data

along the Lw-H-C3H8(g) locus, most of the literature data on the Lw-H-C3H8(l) phase boundary

compare favourably to within ± 0.2 K of this model except for Makogon’s (2003) data.

70,72,79,81,83 This can be attributed to the techniques used for measuring the dissociation point

which relies on visual determination of phase transition from one phase to another and the

floating hydrates which can easily get into the pressure transducer.

4.2 Recommendation

The model presented in this study does not converge at hydrate dissociation conditions at low

pressure (< 0.25 MPa). This can be attributed to the Helmholtz energy equation used for

calculating the fugacity in the fluid phase. Here the equations with the GERG mixing rules do

not converge easily at low pressures. Further code development will improve the calculation of

most thermodynamic parameters involving mixtures, including fugacity, at these conditions. As

discussed in section 3.2.2.1, it is assumed that the volume of ice is not changing under 20 MPa.

Higher pressure studies can be carried out to: (i) check the accuracy this present model and

possibly recalibrate some of the parameters used or (ii) to account for the change in volume of

the hydrate as suggested by Ballard.89

4.3 Future work

Normally, laboratory hydrate solid formation conditions are measured in the presence of three

phases (hydrate phase, non-aqueous fluid phase and liquid water phase). In many ways, sub-

Page 96: Thermodynamic Formation Conditions for Propane Hydrates in

76

saturated hydrate formation (no dense phase water) is more applicable to the transportation of

compressed fluids, because they have been previously partially dehydrated. Experimental

measurements are needed for C3H8 in the two phase regions (C3H8(l)–H, C3H8(g) –H)) to

recalibrate the current model and extend its uses for calculations at these conditions, i.e., water

content measurements above hydrate would make these models much more applicable to

industrial issues associated with flow assurance.

Of all the known hydrate formers, H2S can form hydrates at very low pressures and can remain

stable up to a temperature of 303.15 K. H2S also increases the hydrate formation temperature of

hydrocarbons.62

To the best of my knowledge, there are no equilibrium data in the literature

containing more than 50 mol % H2S with any hydrocarbon in the presence of a liquid water

(saturated) phase or water content data above mixed hydrates. High H2S concentrations with

some C3H8, C4H10 or C2H6 impurities in variable amounts over a range of temperature and

pressure would be of interest models applied to in the oil and gas industry. These data also could

be useful in designing a hydrate based gas separation processes to separate H2S in sour gas

streams, where hydrate separation is energy demanding and has the potential to partially replace

some amine processes currently used.

Page 97: Thermodynamic Formation Conditions for Propane Hydrates in

77

Appendix A

Calibrations and Results

A.1.1 Pressure calibration

The Paroscientific Inc. Digiquartz 410KR-HT-101 Pressure Transducer was used for the study of

the dissociation conditions in the Lw–H–C3H8(g) phase boundary, but for measurements in the

Lw–H–C3H8(l) region a Keller druckmesstechnik PA-33X Pressure Transducer was used. There

are two methods which have been used for calibrating these transducers: (i) primary calibration

against a deadweight tester and (ii) secondary calibration against another well calibrated

transducer.

A.1.1.1 Primary transducer calibration through the use of Deadweight Testers.

Deadweight Testers are the primary standard used for calibrating any pressure measuring

transducers and gauges above ambient conditions. There are three primary components of a

Deadweight Tester device: a weight and piston used to apply the pressure, a clamp to attach the

gauge or transducer and a calibrating fluid (isopropanol) for pressure transmission.106

Weights

are used to apply a known force on an accurately determined area on the piston thereby exerting

a pressure on the fluid; this pressure is transferred to the gauge to be calibrated. The pressure at

the piston face, therefore, is equal to the pressure throughout the calibrating fluid in the tester and

is given as107-108

A

gm

A

Fp i

i (A.1.1)

Page 98: Thermodynamic Formation Conditions for Propane Hydrates in

78

where F, A, i

im and g represents the force of the weight on the piston, cross sectional area of

the occupied weight, sum total of the masses of the applied weight and acceleration due to

gravity respectively. To ensure an accurate calibration, the applied force needs to be corrected

for factors such as the local gravity, the buoyancy of the weight on the fluid, the local

temperature and the thermal expansion of the tester, expansion of the effective area due to the

applied pressure, and any additional static head pressure caused by a height difference between

the transducer and piston.106-109

Applying these corrections, equation A.1.1 becomes:

fl

m

a

p

i

i

corr gA

gm

p

1 (A.1.2)

where pA is the buoyancy correction factor, lg represents the local acceleration due to gravity

which was recorded as 9.8082 m.s-2

while f , m and a represent the densities for the fluid,

weight and air which were 785 kg.m-3

, 7300 kg.m-3

and 1.22 kg.m-3

respectively.109

The

corrected pressure, corrp , measured for the different weights were then plotted against the pressure

obtained from the gauge to obtain a calibration equation.

A.1.1.2 Secondary transducer calibration

This type of calibration was achieved by comparing the measurements from a primarily

calibrated transducer to the uncalibrated transducer in hydraulic communication with each other.

It is easier and faster to calibrate a pressure measuring device using this method versus of going

through the deadweight test procedure for an uncalibrated transducer. The uncalibrated

transducers (Paroscientific Inc. Digiquartz 410KR-HT-101 and Keller Druckmesstechhnik PA-

33X) were compared to a primary calibrated transducer Paroscientific Inc. Digiquartz 410KR-

Page 99: Thermodynamic Formation Conditions for Propane Hydrates in

79

HT-101 which was calibrated using the Pressurements Limited T 3800/4 Deadweight Tester by

Connor Deering.109

The Paroscientific Inc. Digiquartz 410KR-HT-101 Pressure Transducer was

initially calibrated by Zachary Ward through a secondary calibration, another secondary check

for this work was also done to confirm the calibration was still valid.105

Different pressures of

nitrogen ranging from 3.39 to 14.00 MPa and under vacuum were used as reference points for

calibrations, where pressurized nitrogen reduces any hydraulic head difference. The calibrated

and uncalibrated transducers were placed in hydraulic communication, the pressure

measurements from the calibrated gauge were plotted against measurements by the uncalibrated

devices to obtain a linear calibration equation.

A.1.1.3 Results and discussion

The Paroscientific and Keller Pressure Transducers can measure pressure up to 20.84 and 100.00

MPa respectively. The measured pressures by the uncalibrated Paroscientific Pressure

Transducer (measp ) were compared to the pressures (

calp ) from the calibrated primary

transducer in the range p = 3.3881 - 13.9339 MPa and under vacuum with Table A.1 showing the

differences between the transducers. The mean average of the differences in pressure

measurements ( calmeas pp ) between the transducers before calibration was 31099.9 MPa with a

95 % standard error of 0.016 observed at 5.8 MPa. A calibration equation was obtained from the

linear regression of calp versus measp :

)0.00032 ±00002.1(meascal pp (A.1.3)

Page 100: Thermodynamic Formation Conditions for Propane Hydrates in

80

Table A.1 Comparison of the pressures measured by the calibrated primary Paroscientific

Transducer,calp , and the uncalibrated Paroscientific Pressure Transducer, measp .

/calp MPa /measp MPa /calmeas pp MPa

13.9339 13.9460 0.0121

10.3479 10.3577 0.0098

6.8819 6.8909 0.0089

3.3881 3.3964 0.0084

0.0000017 0.0107 0.0107

Average 0.0099 ± 0.0015

Similarly, pressure measurements from the uncalibrated Keller transducer were also compared to

the primary calibrated transducer from pressure p = 5.1173 - 19.0525 MPa and under vacuum.

The mean average for the differences before calibration was -0.0402 MPa (Table A.2). The

linear regression of calp against

measp shows

)0.001 ±0058.1(meascal pp (A.1.4)

Table A.2. Comparison of the pressures measured by the calibrated primary Paroscientific

Transducer, calp , and the uncalibrated Keller Pressure Transducer, measp .

calp / MPa measp / MPa /calmeas pp MPa

19.0362 19.0525 0.0163

7.8011 7.7124 -0.0887

5.1173 5.0287 -0.0886

0.0017 0.0019 0.0002

Average -0.0402 ± 0.0564

Page 101: Thermodynamic Formation Conditions for Propane Hydrates in

81

A.1.2 Temperature calibration

A.1.2.1 The International Temperature Scale

The International Temperature Scale (ITS) was adopted by the seventh Conference Generales

des Poids et Mesures in 1927 to overcome the difficulties and variation in measurement of

thermodynamic temperature by use of gas thermometry.110

The temperature scale was amended

and updated at various points through the years, in 1948, 1960, 1968, 1975, 1976 and finally in

1990. The International Temperature Scale of 1990 (ITS-90) supersedes any previously amended

scales and it is the currently accepted standard. The ITS-90 defines temperature in terms of

Kelvin (T90) and Celsius (t90) with the relationship: 110

15.273/ Cº/ 9090 KTt . The ITS-90

also provides a temperature scale between the 0.65 K to the highest temperature practically

measurable in terms of the Planck radiation law for monochromatic radiation, between defined

fixed points and specified references for different temperature ranges.110-111

These fixed points

and specified references are the primary and secondary standards respectively used for the

accurate calibration of a thermometer by comparing the temperature measured to the

standards.111

The fixed points are usually triple point temperatures for pure substances while the

reference points consist of melting and boiling temperatures of various pure substances. Between

the triple point of hydrogen (13.803 K) to the freezing point of silver (1234.93 K), T90 is

measured by means of a platinum resistance thermometer (PRT) calibrated at specified sets of

defining fixed point and specified references provided by ITS-90.110

The ITS-90 provides a

number of secondary reference points whose temperatures also have been accurately determined

from the primary standards.111

These secondary references point can be used in calibrating a

thermometer in place of the primary standards because, in most cases, they are easily

reproduced.

Page 102: Thermodynamic Formation Conditions for Propane Hydrates in

82

A.1.2.2 Calibration procedure

The autoclave is rated for a large temperature range, although the temperature range for C3H8

hydrate dissociation study was only between 271.15 to 280.15 K. ITS-90 recommends that

within that region, thermometers can be calibrated using triple or the melting point temperatures

of H2O at T = 273.15 K and 273.16 K respectively.110

Here only a single point calibration

assumes a constant offset. The 100 ohm, four-wire PRT used for temperature measurement

inside the autoclave was calibrated by the melting point of ice water for a single point

calibration. The temperature was regulated with the PolyScience circulating bath to an precision

of δT = ± 0.004 K using 50:50 ethylene glycol:water as the circulating fluid. The water used was

purified using a EMD Millipore Milli-Q water treating system to a resistance of 18 MΩ·cm

followed by degassing for several hours under vacuum. Prior to loading with degassed water, the

autoclave was evacuated overnight to a vacuum of 7105.2 MPa. About 15.00 cm3 of the

degassed water was injected into the autoclave by suction. The system was first cooled to

T = 278.15 K rapidly for 2 minutes and then sub-cooled to T = 263.15 K to form ice water for 7

hours. This is to allow the ice to anneal before increasing the temperature back to 278.15 K for

60 minutes to obtain the melting point temperature of ice at 273.15 K. This procedure was

repeated three times and the average of the horizontally leveled region (Figure A.1) was obtained

and recorded as the melting point temperature of ice water.

The other PRT (used inside the circulating water bath) was compared to the primary calibrated

PRT. The PRTs (i.e., already calibrated and the one to be calibrated) were used to measure the

temperature of a thermal equilibrated circulating water bath for a period of 2 minutes at different

temperature from T = 273.15 to 333.15 K.

Page 103: Thermodynamic Formation Conditions for Propane Hydrates in

83

Figure A.1. A representative temperature-time plot showing the water freezing points for the

PRT probe calibration.

A.1.2.3 Result and discussion

Although the melting point of H2O is a secondary reference, it is easier to obtain than the triple

point primary reference. In order to achieve satisfactory isothermal phase transitions, two

requirements were needed: (i) a sharp, identifiable initial inflection in the measured temperature,

and (ii) a stable, constant temperature while the transition is occurring.109

To achieve these

conditions, the heating rate was set to 0.167 K min-1

. The average of all the temperatures

recorded during the inflection (see Figure A.1) was taken and the procedure was repeated three

times. The results of the trials are presented in Table A.3 below.

273.10

273.12

273.14

273.16

273.18

273.20

273.22

273.24

273.26

273.28

273.30

Tem

per

ature

/ K

Time

Page 104: Thermodynamic Formation Conditions for Propane Hydrates in

84

Table A.3. The experimentally measured melting points of H2O with the corresponding

deviations.

Trial

T / K δT / K

1. 273.124 ± 0.003

2. 273.118 ± 0.005

3. 273.127 ± 0.001

Average T / K 273.123 ± 0.002

The results of the other PRT, T2, used inside the water bath are presented in Table A.4 The

average of the differences observed between the measurements was 0.159 ± 0.0483 K.

Table A.4. Comparison of the measured temperatures from the calibrated, Tcal, and uncalibrated

PRT, Tmeas (used inside the water bath).

Step T / K

measT / K

calT / K

calmeas TT / K

273.15

273.231

273.118

0.112

283.15 283.362 238.098 0.264

293.15 293.251 293.108 0.143

303.15 303.278 303.133 0.145

313.15 313.298 313.146 0.152

323.15 323.298 323.148 0.149

333.15 333.438 333.292 0.146

Average 0.159 ± 0.048

Page 105: Thermodynamic Formation Conditions for Propane Hydrates in

85

A.1.3 Volume calibration

The volume (V) of the autoclave was calibrated by volume difference. The autoclave cell was

filled with degassed water using a high-pressure syringe pump (Teledyne ISCO Model 260D) at

a constant pressure of p = 25.00 MPa. Before loading with degassed water, the setup was

evacuated for 12 hours. The volume of degassed water (V1) inside the syringe pump was first

recorded before filling the autoclave cell and feed lines that connects the autoclave to the syringe

pump. The water was pumped at pressure p = 25.00 MPa up to the outlet valve on the autoclave

(see Figure 2.1) while closed. When the volume remained constant as indicated by the pump, the

new volume inside the pump was recorded as V2. The difference in the volume (V1 – V2) before

loading and after loading up to valve, VA2, gives the volume of the feedline. Valve, VA2, was

then opened to allow water from the syringe pump, still at pressure p = 25.00 MPa, into the cell.

The system was then left for a period of 48 hours to ensure that all volume was filled with water

and that the pressure remained constant, after which the new volume (V3) of the pump was then

recorded. The volume of the autoclave cell is given as

V = V3 – V2 = 165.72 – 119.51 = 46.21 cm3

(A.1.5)

This method, though accurate enough for this study, had two disadvantages: (i) setup took a very

long time to completely dry after the procedure and (ii) the precision is limited to the volume

measurement of the pump which is ± 0.01 cm3.

Page 106: Thermodynamic Formation Conditions for Propane Hydrates in

86

Appendix B

Pressure versus temperature plots of the experimental run for the dissociation points along

Lw-H-C3H8(g) and Lw-H-C3H8(l) phase boundaries reported in this study.

Figure B.1. Pressure versus temperature profile for 99.999 mol % C3H8 + H2O showing the

cooling, hydrate formation and heating stages along the Lw-H-C3H8(g) locus.

Figure B.2. Pressure versus temperature profile for 99.5 mol % C3H8 + H2O showing the

cooling, hydrate formation and heating stages along the Lw-H-C3H8(g) locus.

0.0

0.1

0.2

0.3

0.4

0.5

0.6

270 275 280 285 290 295 300

p/

MP

a

T / K

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

270 275 280 285 290 295 300

p/

MP

a

T / K

Page 107: Thermodynamic Formation Conditions for Propane Hydrates in

87

Figure B.3. Pressure versus temperature profiles for C3H8 + H2O showing the cooling, hydrate

formation and heating stages along the Lw-H-C3H8(l) locus.

0

1

2

3

4

5

6

7

8

9

270 275 280 285

p/

MP

a

T / K

12

13

14

15

16

17

18

19

20

21

272 274 276 278 280 282

p/

MP

aT / K

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

272 274 276 278 280 282

p/

MP

a

T / K

12

14

16

18

20

22

24

272 274 276 278 280 282

p/

MP

a

T / K

Page 108: Thermodynamic Formation Conditions for Propane Hydrates in

88

Figure B.3. Pressure versus temperature profiles for C3H8 + H2O showing the cooling, hydrate

formation and heating stages along the Lw-H-C3H8(l) locus cont’d.

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

272 274 276 278 280 282

p/

MP

a

T / K

0.0

0.5

1.0

1.5

2.0

2.5

272 274 276 278 280 282

p/

MP

a

T / K

12

14

16

18

20

272 274 276 278 280 282

p/

MP

a

T / K

4

5

6

7

8

9

10

274 276 278 280 282

p/

MP

a

T / K

Page 109: Thermodynamic Formation Conditions for Propane Hydrates in

89

Figure B.3. Pressure versus temperature profiles for C3H8 + H2O showing the cooling, hydrate

formation and heating stages along the Lw-H-C3H8(l) locus cont’d.

4

5

6

7

8

9

10

274 276 278 280 282

p/

MP

a

T / K

10

11

12

13

14

15

16

274 276 278 280 282

p/

MP

a

T / K

0.2

0.3

0.4

0.5

0.6

0.7

272 274 276 278 280 282

p/

MP

a

T / K

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

270 275 280 285 290

p/

MP

a

T / K

Page 110: Thermodynamic Formation Conditions for Propane Hydrates in

90

Figure B.3. Pressure versus temperature profiles for C3H8 + H2O showing the cooling, hydrate

formation and heating stages along the Lw-H-C3H8(l) locus cont’d.

0

1

2

3

4

5

6

7

8

9

270 275 280 285

p/

MP

a

T / K

12

13

14

15

16

17

18

19

20

21

272 274 276 278 280 282

p/

MP

a

T / K

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

272 274 276 278 280 282

p/

MP

a

T / K

12

14

16

18

20

22

24

272 274 276 278 280 282

p/

MP

a

T / K

Page 111: Thermodynamic Formation Conditions for Propane Hydrates in

91

Appendix C

Parameters and coefficients used in the reduced energy Helmholtz EOS for calculation of

thermodynamic properties of C3H8 in equation 3.13.112

k kN kt kd kl k k k

k

1. 0.042910051 1 4

2. 1.7313671 0.33 1

3. -2.4516524 0.8 1

4. 0.34157466 0.43 2

5. -0.46047898 0.9 2

6. -0.66847295 2.46 1 1

7. 0.20889705 2.09 3 1

8. 0.19421381 0.88 6 1

9. -0.22917851 1.09 6 1

10. -0.60405866 3.25 2 2

11. 0.06668065 4.62 3 2

12. 0.01753462 0.76 1 0.963 2.33 0.684 1.283

13. 0.33874242 2.5 1 1.977 3.47 0.829 0.693

14. 0.22228777 2.75 1 1.917 3.15 1.419 0.788

15. -0.23219062 3.05 2 2.307 3.19 0.817 0.473

16. -0.09220694 2.55 2 2.546 0.92 1.5 0.857

17. -0.47575718 8.4 4 3.28 18.8 1.426 0.271

18. -0.01748682 6.75 1 14.6 547.8 1.093 0.948

Page 112: Thermodynamic Formation Conditions for Propane Hydrates in

92

Appendix D

First derivative of and the reducing function r and

rT with respect to in .113

n

jnvTi

r

n

n

,,

)(

=

N

k

r

xk

r

x

jni

rr

jni

r

r

r

kix

n

Tn

Tnn

1,,

.1

.1

1

,

where

jni

r

n

Tn

,

=

jx

N

k k

rk

i

r

x

Tx

x

T

1

and jni

r

nn

,

=

jx

N

k k

rk

i

r

xx

x

1

.

Page 113: Thermodynamic Formation Conditions for Propane Hydrates in

93

Appendix E

Copyright permissions

Page 114: Thermodynamic Formation Conditions for Propane Hydrates in

94

Page 115: Thermodynamic Formation Conditions for Propane Hydrates in

95

REFERENCES

1. Comb, S. Texas Comptroller of Public Account, The Energy Report. 81–88, May, 2008.

2. National Propane Gas Association (NGPA). Fact about propane, America exceptional

energy, 1992.

3. Havard, D. Oil and gas production handbook: An introduction to oil and gas production,

transport, refining and petrochemical industry. 3rd ed., ABB Oil and Gas, 2013, pp 78-79.

4. Katz, D.K. Handbook of natural gas engineering, McGraw-hill book company, New York,

1959.

5. U.S. Department of Energy, Office of Energy Efficiency and Renewable Energy. Freedom

car and vehicle technologies program. Liquefied petroluem gas, Just the Basics, 1992.

6. Younger, A.H. Natural Gas Processing Principles and Technology - part I. University of

Calgary, Alberta, 2004.

7. Sloan, E.D and Koh, C.A. Clathrate Hydrates of Natural Gases, 3rd ed., CRC Press: New

York, Boca Raton, FL, 2007.

8. Carroll, J. Natural Gas Hydrates, A Guide for Engineers. 2nd ed., Gulf Professional

Publishing: Burlington, MA, 2009.

9. Koh, C.A.; Sum, A.K.; Sloan, E.D. Natural Gas Hydrates in Flow Assurance. Gulf

Professional Publishing, 2011.

10. Olabisi, O. T and Sunday, I. S. LPG Hydrate Formation and Prevention using Ethanol and

Methanol. SPE-178333-MS. 2015, 1–8.

11. Gaudette, J.; Servio, P. Measurement of dissolved propane in water in the presence of gas

hydrate. J. Chem. Eng. Data. 2007, 1449–1451.

12. Akatsu,S.; Tomita,S.; Mori,Y.H.; R. Ohmura, Thermodynamic simulations of hydrate-

based removal of carbon dioxide and hydrogen sulfide from low-quality natural gas. Ind.

Eng. Chem. Res. 2013, 52, 15165–15176.

13. Sloan, E. D. Fundamental principles and applications of natural gas hydrates. Nature.

2003, 426, 353–363.

14. Wong, S and Bioletti, R, Carbon dioxide separation technologies, Alberta Res. Counc.,

2002.

15. Babu, P.; Linga, P.; Kumar, R.; and Englezos, P. A review of the hydrate based gas

separation (HBGS) process for carbon dioxide pre-combustion capture. Energy. 2015, 85,

261–279.

16. Davy, H. On a Combination of oxymuriatic gas and oxygene gas. Phil. Trans. Roy. Soc.

Lond. 1811, 101, 155-162.

Page 116: Thermodynamic Formation Conditions for Propane Hydrates in

96

17. Priestley, J. Experiments and observations on different kinds of air and other branches of

natural philosophy Connected with the Subject. New York, Kraus reprint co., 1970.

18. http://memim.com/paul-ulrich-villard.html. accessed December 10, 2015.

19. Villard. Organic chemistry. J. Chem. Soc., Abstr. 1890, 58, 1386-1450.

20. Hammerschmidt, E. G. Formation of gas hydrates in natural gas transmission lines. J. Ind.

Eng. Chem.1934, Washington, D. C., 26, 851–855.

21. Von Stackelberg, M. Gas hydrate, Naturwissenschaften. 1949, 36(11), 327–362.

22. Ripmeester, J. A.; Tse, J. S.; Ratcliffe, C. I.; Powell, B. M. A new clathrate hydrate

structure. Nature. 1987, 325, 135–136.

23. Jeffrey, G.A, Atwood, J.L.; Davies, J.E.D.; MacNicol, D.D. Hydrate inclusion compounds.

Eds., Inclusion compounds, Academic Press, New York, 1984, 135-190.

24. Mehta, A. P and Sloan,E.D. A thermodynamic model for structure-H hydrates. AIChE J.

1994,40, 312–320.

25. Centre for Gas Hydrate Research, Institute of Petroleum Engineering, Heriot-Watt

University(http://www.pet.hw.ac.uk/research/hydrate/hydrates_what.cfm?hy=what),

accessed November 21, 2015.

26. Babu, P.; Linga,P.; Kumar, R.; Englezos, P. A review of the hydrate based gas separation

(HBGS) process for carbon dioxide pre-combustion capture. Energy. 2015, 85, 261–279.

27. Eslamimanesh,A. Thermodynamic studies on Semi-Clathrate Hydrates of TBAB + gases

containing Carbon Dioxide. École Nationale Supérieure des Mines de Paris, 2012.

28. Kiyono, F. Separation and Recovery of CO2 From Exhausted Gas By Hydrates. 220th

ACS National Meeting, Washington, DC, United States, August 20-24, 2000.

29. Koh, C. A.; Sloan, E. D.; Sum, A. K.; Wu, D. T.; Fundamentals and Applications of Gas

Hydrates. Annu. Rev. Chem. Biomol. Eng. 2011, 2, 237–257.

30. Rufford, T. E.; Smart, S.;. Watson, G. C. Y.; Graham, B. F.; Boxall, J.; Diniz da Costa, J.

C.; May, E. F. “The removal of CO2 and N2 from natural gas: A review of conventional

and emerging process technologies. J. Pet. Sci. Eng. 2012, 95, 123–154.

31. Sabil, K. M.; Azmi, N.; Mukhtar, H. A review on carbon dioxide hydrate potential in

technological applications. J. Applied Sciences. 2011. 11, 3534–3540.

32. Eslamimanesh, A.; Mohammadi, A. H.; Richon, D.; Naidoo, P.; Ramjugernath, D.

Application of gas hydrate formation in separation processes: A review of experimental

studies. J. Chem. Thermodyn. 2012, 46, 62–71.

33. Dyadin, Y. A.; Larionov, E. G.; Aladko, E. Y.; Manakov, A. Y.; Zhurko, F. V.; Mikina,

T. V.; Grachev, E. V. Clathrate formation in water-noble gas (Hydrogen) systems at high

pressures. J. of Structural Chemistry, 1999, 40(5), 790–795.

Page 117: Thermodynamic Formation Conditions for Propane Hydrates in

97

34. Strobel, T. A.; Taylor, C. J.; Hester, K. C.; Dec, S. F.; Koh, C. A.; Miller, K. T.; Sloan, E.

D. Molecular Hydrogen Storage in Binary THF-H2 Clathrate Hydrates. J. Phys. Chem.,

2006, 110, 17121-17125

35. Mao, W. L.; Mao, H.-K.; Goncharov, A. F.; Struzhkin, V. V.; Guo, Q.; Hu, J.; Shu, J.;

Hemley, R. J.; Somayazulu, M.; Zhao, Y. Hydrogen clusters in clathrate hydrate. Science.

2002, 297, 2247–2249.

36. Strobel, T. A.; Hester, K. C.; Koh, C. A.; Sum, A. K.; Sloan, E. D.; Properties of the

clathrates of hydrogen and developments in their applicability for hydrogen storage.

Chem. Phys. Lett. 2009, 478, 97–109.

37 Qian, G.R.; Lyakhov, A. O.; Zhu, Q.; Oganov, A. R.; Dong, X. Novel Hydrogen Hydrate

Structures under Pressure. Sci. Rep. 2014, 4, 1–5.

38. Zhang, J.S and Lee, J.W. Equilibrium of hydrogen + cyclopentane and carbon dioxide +

cyclopentane binary hydrates. J. Chem. Eng. Data. 2009, 54, 659–661.

39. Herslund, P. J., von Solms, N., Abildskov, J., & Thomsen, K.. Thermodynamic and

Process Modelling of Gas Hydrate Systems in CO2 Capture Processes. Proceedings of the

7th International Conference on Gas Hydrates (ICGH 2011), Edinburgh, Scotland, United

Kingdom, July 17-21, 2011.

40. Komatsu, H.; Yoshioka, H.; Ota, M, Sato, Y; Wantanabe, Smith, L.; Peter, J. Phase

Equilibrium Measurements of Hydrogen− Tetrahydrofuran and Hydrogen−Cyclopentane

Binary Clathrate Hydrate Systems. J. Chem. 2010, 55. 2214–2218.

41. Skiba, S. S.; Larionov, E. G.; Manakov, A. Y.; Kolesov, B. A.; Ancharov, A. I.; Aladko,

E. Y. Double clathrate hydrate of propane and hydrogen. J. Incl. Phenom. Macrocycl.

Chem. 2009, 63, 383–386.

42. Alberta oil and gas quarterly update; winter 2015. Reporting on the period: Oct.1, 2014 to

Dec.18 2014.

43. Yu, C.H. A Review of CO2 Capture by Absorption and Adsorption. Aerosol Air Qual. Res.

2012, 12, 745–769.

44. Anderson, S.; Newell, R. Prospects for carbon capture and storage technologies. Annu.

Rev. Environ. Resour. 2003, 29, 109–142.

45. Ho, L. C.; Babu, P.; Kumar, R. Linga, P. Hydrate based gas separation process for carbon

dioxide capture employing an unstirred reactor with cyclopentane. Energy. 2013, 63, 252–

259.

46. Eltawil, M. A.; Zhengming, Z.; Yuan, L. A review of renewable energy technologies

integrated with desalination systems. Renew. Sustain. Energy Rev. 2009, 13, 2245–2262.

47. Wang, P and Chung, T.S. A conceptual demonstration of freeze desalination–membrane

distillation (FD–MD) hybrid desalination process utilizing liquefied natural gas (LNG)

cold energy. Water Res. 2012, 46, 4037–4052.

Page 118: Thermodynamic Formation Conditions for Propane Hydrates in

98

48. Kang, K. C.; Linga, P.; Park, K.; Choi, S.J.; Lee, J. D. Seawater desalination by gas

hydrate process and removal characteristics of dissolved ions (Na+, K

+, Mg

2+, Ca

2+, B

3+,

Cl−, SO4

2−). Desalination. 2014, 353, 84–90.

49. Lee, J. D.; Hong, S. Y.; Park, K.; Kim, J. H.; Yun, J. H. A New Method for Seawater

Desalination Via Gas Hydrate Process and Removal Characteristics of Dissolved

Minerals. Spectroscopy. 2011, 274, 7–10.

50. Collett, T. S. Natural Gas Hydrates : Resource of the 21st Century ?, Pratt II Conference,

Petroleum Province of the 21ist

century, Jan. 12-15, 2000. San Diego, California.

51. Makogon, Y. F.; Holditch, S. A.; & Makogon, T. Y. Natural gas-hydrates - A potential

energy source for the 21st Century. J. of Petroleum Science and Engineering. 2007 56 (1-

3), 14–31.

52. Demirbas, A. Methane hydrates as potential energy resource: Part 1 - Importance, resource

and recovery facilities. Energy Conversion and Management. 2010, 51(7), 1547–1561.

53. Ruppel, C. Methane Hydrates and the Future of Natural Gas, MITEI Nat. gas Report,

Suppl. Pap. Methane Hydrates. 2011, 1–25.

54. Qin, J; Kuhs, W. F. Calibration of Raman Quantification Factors of Guest Molecules in

Gas Hydrates and Their Application to Gas Exchange Processes Involving N2. J. Chem. &

Eng. Data. 2015, 60 (2), 369-375.

55. Carroll, J. J.; Maddocks, J. R. Design considerations for acid gas injection.Proceedings-

Laurance Reid Gas Conditioning Conference (1999), 90-116.

56. http://journalofcommerce.com/OHS/News/2013/3/Worker-dies-and-another-injured-in-

separate-incidents-JOC054667W/ accessed December 12, 2015.

57. Perrin, A.; Musa, O. M.; Steed, J. W. The chemistry of low dosage clathrate hydrate

inhibitors. Chem. Soc. Rev. 2013, 42, 1996–2015.

58. Haniffa, M. A. M.; Hashim, F. M. Review of Hydrate Prevention Methods for Deepwater

Pipelines. IEE Conf. Open Syst. 2011, 159–165.

59. Giavarini, C.; Maccioni, F.; Santarelli, M. L. Formation Kinetics of Propane Hydrates.

Ind. Eng. Chem. Res. 2003, 42, 1517–1521.

60. Harmen, A and Sloan, E. D. The Phase Behaviour of the Propane-Water System . Can. J.

Chem. Eng. 1990, 68(1), 151-158.

61. Sun, C.; Li, W.; Yang, X.; Li, F.; Yuan, Q.; Mu, L.; Chen, J.; Liu, B.; Chen, G. Progress

in Research of Gas Hydrate. Chinese J. Chem. Eng. 2011, 19, 151–162.

62. Carroll, J. An examination of the prediction of hydrate formation conditions in sour

natural gas. GPA Europe. Spring Meeting Dublin. 2004.

63. Mansoori, G.A . A unified perspective on the phase behaviour of petroleum fluids. Int. J.

Oil, Gas and Coal Technology. 2, 2009.

Page 119: Thermodynamic Formation Conditions for Propane Hydrates in

99

64. Carroll, J.J.; Phase diagram reveal acid gas injection subtleties. Oil and gas. J. 1998, 96

(9), 92-96.

65. Beltran, J. G.; Bruusgaard, H.; Servio, P. Gas hydrate phase equilibria measurement

techniques and phase rule considerations. J. Chem. Thermodyn. 2012, 44, 1–4.

66. Giavarini, C.; Hester, K. Gas Hydrates: Immense Energy Potential and Environmental

Challenges. London: Springer-Verlag, 2011.

67. V. P. Carey, The properties of gases & liquids, vol 1, (4). 1988. pp 256-257

68. Atkins, Peter; de Paula, Julio. Atkins Physical Chemistry, 9th ed., 2010, Oxford University

Press, pp 146-148.

69. Kamath, V. A. Study of Heat Transfer Characteristics during Dissociation of Gas

Hydrates. University of Pittsburgh, Pittsburgh, PA, Ph.D. Dissertation., 1984.

70. Reamer, H. H.; Selleck, F. T.; Sage, B. H. Some Properties of Mixed Paraffinic and

Olefinic Hydrates. J. Pet. Technol. 1952, 4, 197–202.

71. Tumba, K.; Babaee, S.; Naidoo, P.; Mohammadi, A. H.; Ramjugernath, D. Phase

Equilibria of Clathrate Hydrates of Ethyne + Propane. J. Chem. Eng. Data. 2014, 59 (9),

2914-2919.

72. Verma, V.K. Gas hydrates from Liquid hydrocarbon-water systems.,University of

Michigan: Ann Arbor, MI, Ph.D. Dissertation, 1974.

73. Englezos, P and Ngan, Y.T. Incipient Equilibrium Data for Propane Hydrate Formation in

Aqueous Solutions of NaCl, KCl, and CaC12. J. Chem. Eng. Data, 1993,38, 250-253.

74. Robinson, D and Metha, B. Hydrates in the propane−carbon dioxide−water system. J.

Can. Pet. Technol. 1971, 10 (1), 642−644.

75. Patil, S. L. Measurement of Multiphase Gas Hydrate Phase Equilibria: Effect of Inhibitors

and Heavier Hydrocarbon Components. M.S Thesis, University of Alaska, Alaska, 1987.

76. Kubota, H. K.; Shimizu, Y. Tanaka and T. J. Makita. Thermodynamic Properties of R13

(CClF3), R23(CHF3), R152a(C2H4F2) and Propane Hydrates for Desalination of Seawater,

Chem. Eng. Japan, 1984, 17(4), 423-429.

77. Deaton, W and Frost, J. E. Gas Hydrates and Their Relation to the Operation of Natural-

Gas Pipe Lines; Monograph, U.S. Bureau of Mines, Vol. 8, American Gas Association:

New York. 1946, p 101.

78. Thakore, J. L and Holder, G. D. Solid vapor azeotropes in hydrate-forming systems. Ind.

Eng. Chem. Res.1987, 26 (3), 462–469.

79. Den Heuvel, M. M. M.; Peters, C. J.; de Swaan Arons, J. Gas hydrate phase equilibria for

propane in the presence of additive components. Fluid Phase Equilib. 2002, 193, 245–259.

Page 120: Thermodynamic Formation Conditions for Propane Hydrates in

100

80. Nixdorf, J and Oellrich, L. R. Experimental determination of hydrate equilibrium

conditions for pure gases, binary and ternary mixtures and natural gases. Fluid Phase

Equilib. 1997, 139, 325–333.

81. Wilcox, W.; Carson, D.; Katz, L. Natural gas hydrates. Ind. Eng. Chem.1941, 33, 662–

665.

82. Miller B and Strong, E.R. Hydrate storage of natural gas. Am. Gas Assoc. Monthly. 1946,

28 (2), 63-67.

83. Makogon, Y. F.; Liquid propane + water phase equilibria at hydrate conditions. J. Chem.

Eng. Data. 2003, 48 (2), 347–350.

84. Maekawa, T. Equilibrium conditions of propane hydrates in aqueous solutions of alcohols,

glycols, and glycerol. J. Chem. Eng. Data. 2008, 53 (12), 2838–2843.

85. Van der Waals, J.H and Platteeuw, J.C. Clathrate Solutions. Advances in Chemical

Physics. 1959.

86. Parrish, W. R and Prausnitz, J. M. Dissociation Pressures of Gas Hydrates Formed by Gas

Mixtures. Ind. Eng. Chem. Process Des. Dev. 1972, 11 (1), 26–35.

87. McKoy, V and Sinanoglu , O. Theory of Dissociation Pressures of Some Gas Hydrates. J.

Chem. Phys. 1963, 38 (12), 2946.

88. Holder, G.D.; Corbin, G.; Papadopoulos, K.D. Thermodynamic and molecular properties

of gas hydrates from mixtures containing methane, argon, and krypton. Ind. Eng. Chem.

Fundam. 1980, 19, 282–286.

89. Ballard, A. L. A Non-ideal Hydrate Solid Solution Model for a Multi-phase Equilibria

Program, PhD thesis, Colorado School of Mines: Golden, CO, USA, 2002.

90. Chen, G.J and Guo, T.M. Thermodynamic modeling of hydrate formation based on new

concepts. Fluid Phase Equilibria. 1996, 122 (1-2), 43–65.

91. Klauda, J. B and Sandler, S. I. A Fugacity Model for Gas Hydrate Phase Equilibria. Ind.

Eng. Chem. Res.2000, 39, 3377–3386.

92. Klauda, J.B and Sandler, S.I. Phase behavior of clathrate hydrates: A model for single and

multiple gas component hydrates. Chem. Eng. Sci. 2003, 58, 27–41.

93. Oellrich, L. R. Natural Gas Hydrates and their Potential for Future Energy Supply. Heat

Mass Transf. 2004, 70–78.

94. Tohidi, B.; Burgass, R. W.; Danesh, A.; Ostergaard, K. K; Todd, A. C. Improving the

accuracy of gas hydrate dissociation point measurements. Gas Hydrates Challenges Fut.

2000, 912, 924–931.

95. Loh, M.; Falser, S.; Babu, P.; Linga, P.; Palmer, A.; Tan, T. S., Dissociation of Fresh and

Seawater Hydrates along the Phase Boundaries between 2.3 and 17 MPa. Energy & Fuels.

Page 121: Thermodynamic Formation Conditions for Propane Hydrates in

101

2012, 26, 6240–6246.

96. Ward, Z.T.; Deering, C.E.; Marriott, R.A.; Sum, A.K. Sloan, E.D. and Koh, C.A. Phase

equilibrium data and model comparisons for H2S hydrates. J. Chem. Eng.Data. 2015, 60

(2), 403-408.

97. Mullin, J. W and Raven, K. D. Influence of Mechanical Agitation on the Nucleation of

some Aqueous Salt Solutions. Nature. 1962, 195, 35-38.

98. Dai, S.; Lee, J. Y.; Santamarina, J. C. Hydrate nucleation in quiescent and dynamic

conditions. Fluid Phase Equilib. 2014, 378, 107–112.

99. Young, S.W. Mechanical stimulus to crystallization in supercooled liquids, J. Am. Chem.

Soc. 1911, 33(2), 148–162.

100. Husowitz, B and Talanquer, V. Nucleation in cylindrical capillaries. J. Chem. Phys. 2004,

121, 8021–8028.

101. Smelik, E. A and King, H. E. Crystal-growth studies of natural gas clathrate hydrates

using a pressurized optical cell. Am. Mineral. 1997, 82, 88–98 .

102. de Loos, T. W.; van der Kool, H. J.; Ott, P. L. Vapor-Liquid Critical Curve of the System

Ethane 2-Methylprapane. J. Chem. Eng. Data. 1986, 31, 166–168.

103. De Groen, M.; Vlugt, T.J.H.; de Loos, T.W. Phase behavior of liquid crystals with CO2

J. Phys. Chem. B. 2012, 116, 9101–9106

104. Apanelopoulou, F.A.P. Louis Paul Cailletet: the liquefaction of oxygen and the Emergence

of low-temperature research. Notes Rec. R. Soc. 2013, 67, 355–373.

105. Ward, Z.T. Phase Equilibria of Gas Hydrates Containing Hydrogen Sulfide and Carbon

Dioxide. Ph.D thesis, 2015. Colorado School of Mines: Golden, CO, USA.

106. http://web.cecs.pdx.edu/~gerry/class/EAS361/lab/pdf/lab2_pressureGages.pdf. (Accessed

February 20, 2016.)

107. Tilford, C.R. Pressure and Vacuum Measurements. In Physical Methods of Chemistry;

Rossiter, B.W., Baetzold, R.C., Eds.; John Wiley & Sons, Inc. Vol. 6; pp 101-173.

108. Pressurements Limited, User Manual for T3800 Deadweight Tester.

109. Deering, C. E. Design, Construction, and Calibration of a Vibrating Tube Densimeter for

Volumetric Measurements of Acid Gas Fluids. M.Sc thesis, University of Calgary, 2015.

110. International Temperature Scale of 1990 (ITS-90); Procès-Verbaux du Comité

International des Poids et Mesures, 78th meeting, 1989.

111. Preston-Thomas, H. The International Temperature Scale of 1990 (ITS-90). Metrologia.

1990, 27, 3-10.

112. Lemmon, E. W.; McLinden, M.O.; Wagner, W. Thermodynamic properties of propane.

Page 122: Thermodynamic Formation Conditions for Propane Hydrates in

102

III. A reference equation of state for temperatures from the melting line to 650 K and

pressures up to 1000 MPa. J. Chem. Eng. Data, 2009, 54, 3141–3180.

113. Lemmon, E. W.; Jacobsen, R. T. A generalized model for the thermodynamic properties

of mixtures. Int. J. Thermophys. 1999, 20, 825-835.

114. Wagner, W and Pruß, A. The IAPWS Formulation 1995 for the Thermodynamic

Properties of Ordinary Water Substance for General and Scientific use. J. Phys. Chem.

Ref. Data. 2002, 31, 387-535.

115. NIST Standard Reference Database 23: Reference Fluid Thermodynamic and Transport

Properties (REFPROP), Version 9.1. National Institute of Standards and Technology,

Standard Reference Data Program: Gaithersburg, MD, 2010.

116. Kunz, O and Wagner, W. The GERG-2008 wide-range equation of state for natural gases

and other mixtures: An expansion of GERG-2004. J. Chem. Eng. Data, 2012, 57, 3032–

3091.

117. Emmerich, W and Letcher, T. M. Volume properties: liquids, solutions and vapours.

Royal society of chemistry. ISBN; 978-1-84973-899-6, 2015..

118. Trussler, J.P.M and Zakari, M.P. The speed of sound in gaseous propane at temperatures

between 225 K and 375 K and at pressures up to 0.8 MPa. J. Chem. Thermodyn. 1996. 28,

329-335.

119. Song, K.Y and Kobayashi, R. The water content of ethane, propane and their mixtures in

equilibrium with liquid water or hydrates. Fluid Phase Equilibria. 1994, 95, 281-298.

120. Song,K.Y.; Yarrison,M.; Chapman,W.G. Experimental low temperature water content in

gaseous methane, liquid ethane, and liquid propane in equilibrium with hydrate at

cryogenic conditions. Fluid Phase Equilibria. 2004, 224, 271-277.

121. Kobayashi R. and Katz D. L. Vapor–liquid equilibria for binary hydrocarbon–water

systems. Ind. Eng. Chem. 1953, 45, 440 – 446.

122. Sloan, E.D.; Bourrie, M.S.; Sparks, K.A.; Johnson, J.J. An experimental method for the

measurement of two-phase liquid hydrocarbon-hydrate equilibrium. Fluid Phase

Equilibria, 1986, 29, 233-240.

123. Bukacek, R. F. Institute of Gas Technology, Research Bulletin, vol. 8, 1955.

124. Lemmon, E.W.; Jacobsen, R.T. A generalized model for the thermodynamic properties of

mixtures. Int.J. Thermophys. 1999, 20, 825-835.

125. Muromachi, S.; Nagashima, H. D.; Herri, J. M.; Ohmura, R.. Thermodynamic modeling

for clathrate hydrates of ozone. J. Chem. Thermodynamics. 2013, 64, 193–197.

126. Dharmawardhana, P. B.; Parrish, W. R.; Sloan, E. D.; Experimental Thermodynamic

Page 123: Thermodynamic Formation Conditions for Propane Hydrates in

103

Parameters for the Prediction of Natural Gas Hydrate Dissociation Conditions. Industrial

& Engineering Chemistry Fundamentals.1980, 19 (4), 410–414.

127. Avlonitis, D. The determination of kihara potential parameters from gas hydrate data.

Chemical Engineering Science. 1994, 49(8), 1161–1173.

128. Bazant, M. Z and Trout, B. L. A method to extract potentials from the temperature

dependence of Langmuir constants for clathrate-hydrates. Physica A: Statistical Mechanics

and Its Applications. 2001, 300 (1-2), 139–173.

129. Anderson, B. J.; Bazant, M. Z.; Tester, J. W.; Trout, B. L. Application of the cell potential

method to predict phase equilibria of multicomponent gas hydrate systems. J. of Physical

Chemistry B. 2005, 109 (16), 8153–8163.

130. Karakatsani, E. K and Kontogeorgis, G.M. Thermodynamic Modeling of Natural Gas

Systems Containing Water. Industrial & Engineering Chemistry Research. 2013, 52, 3499–

3513.

131. Dyadin, Y. A.; Larionov, E. G.; Aladko, E. Y.; Zhurko, F. V. Clathration in Propane-water

and Methane-Propane-water Systems at Pressures up to 15 kbar. Dokl. Akad. Nauk .2001,

376, 497-500.