enhanced solvent vapour extraction processes in …

233
ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN THIN HEAVY OIL RESERVOIRS A Thesis Submitted to the Faculty of Graduate Studies and Research In Partial Fulfillment of the Requirements For the Degree of Doctor of Philosophy in Petroleum Systems Engineering University of Regina By Xinfeng Jia Regina, Saskatchewan January 2014 Copyright 2014: X. Jia

Upload: others

Post on 10-Jun-2022

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES

IN THIN HEAVY OIL RESERVOIRS

A Thesis

Submitted to the Faculty of Graduate Studies and Research

In Partial Fulfillment of the Requirements

For the Degree of

Doctor of Philosophy

in

Petroleum Systems Engineering

University of Regina

By

Xinfeng Jia

Regina, Saskatchewan

January 2014

Copyright 2014: X. Jia

Page 2: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

UNIVERSITY OF REGINA

FACULTY OF GRADUATE STUDIES AND RESEARCH

SUPERVISORY AND EXAMINING COMMITTEE

Xinfeng Jia, candidate for the degree of Doctor of Philosophy in Petroleum Systems Engineering, has presented a thesis titled, Enhanced Solvent Vapour Extraction Processes in Thin Heavy Oil Reservoirs, in an oral examination held on December 18, 2013. The following committee members have found the thesis acceptable in form and content, and that the candidate demonstrated satisfactory knowledge of the subject material. External Examiner: *Dr. Zhangxing Chen, University of Calgary

Co-Supervisor: Dr. Fanhua Zeng, Petroleum Systems Engineering

Co-Supervisor: **Dr. Yongan Gu, Petroleum Systems Engineering

Committee Member: **Dr. Yee-Chung Jin, Environmental Systems Engineering

Committee Member: Dr. Chun-Hua Guo, Department of Mathematics & Statistics

Committee Member: Dr. Ezeddin Shirif, Petroleum Systems Engineering

Committee Member: Dr. Farshid Torabi, Petroleum Systems Engineering

Chair of Defense: Dr. Laurie Clune, Faculty of Nursing *via teleconference **Not present at defense

Page 3: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

i

ABSTRACT

Solvent-based techniques, such as solvent vapour extraction (VAPEX) and cyclic

solvent injection (CSI), have emerged as promising processes to enhance heavy oil

recovery. However, there are still a number of technical issues with these processes, such

as the theoretical modeling and performance enhancement. This thesis aims at addressing

the following major technical topics.

Theoretical modeling of VAPEX. Heavy oil−solvent transition zone is where the

VAPEX heavy oil recovery occurs. Existing analytical VAPEX models can neither fully

characterize the transition zone nor accurately predict its growth. Numerical simulation

models use grid sizes that are much larger than the transition-zone thickness (~1 cm) and

thus cannot capture the characteristics of the transition zone. This study develops a new

two-dimensional (2D) mathematical model for the VAPEX process on the basis of its

major oil recovery mechanisms (i.e., solvent dissolution and gravity drainage) inside the

transition zone. This VAPEX model is able not only to accurately describe the distributions

of solvent concentration, oil drainage velocity, and diffusion coefficient across the

transition zone, but also to predict the evolution of the solvent chamber.

Theoretical modeling of the diffusionconvection mass transfer in CSI. CSI is a

solvent huff-n-puff process. One of the differences between CSI and VAPEX is that the

operating pressure is decreased and increased cyclically in CSI. Hence, in addition to

molecular diffusion, CSI has another mass transfer mechanism, convection, which is

attributed to the bulk motion of solvent caused by the pressure gradient between the solvent

chamber and untouched heavy oil zone. This study develops a convection−diffusion

Page 4: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

ii

mass-transfer model for the heavy oil−solvent mixing process of CSI. The diffusion

coefficient and convection velocity are both considered as variables rather than constants.

Results qualitatively show that pressure gradient can greatly enhance the mixing process.

Enhancement of VAPEX and CSI. This study proposes a new process, namely foamy

oil-assisted vapour extraction (F-VAPEX) to enhance the VAPEX performance.

F-VAPEX combines merits of VAPEX (continuous production) and CSI (strong driving

force) together. It is essentially a VAPEX process during which the operating pressure is

cyclically reduced and restored. It is found that the foamy oil flow during the pressure

reduction period can effectively move the partially diluted heavy oil toward the producer.

Results show that F-VAPEX can increase both the average oil production rate and the

ultimate oil recovery of VAPEX. In comparison with CSI, F-VAPEX has a higher oil

production rate and a lower solvent−oil ratio. This thesis also proposes a new process to

enhance the performance of CSI, namely gasflooding-assisted cyclic solvent injection

(GA-CSI). GA-CSI uses dedicated solvent injector and oil producer to prevent the

‘back-and-forth movement’ of foamy oil inside the solvent chamber during the

conventional CSI process. GA-CSI applies a gasflooding slug immediately after the

pressure depletion process of CSI to produce the partially diluted foamy oil left in the

solvent chamber. It is found that the motionless foamy oil due to pressure depletion and

solvent liberation serves as a buffer zone, which effectively reduces the mobility ratio

between the displacing solvent and the displaced oil and leads to a high sweeping

efficiency. In comparison with the conventional CSI process, the GA-CSI process can

increase the oil production rate by over 3 times and in the meantime decrease the

solvent−oil ratio from ~4 to ~3 g solvent/g oil.

Page 5: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

iii

ACKNOWLEDGMENTS

I want to acknowledge the following individuals or organizations:

Drs. Fanhua Zeng and Yongan Gu, my academic advisors, for their excellent

guidance, valuable advice, strong support, and continuous encouragement

throughout the course of this research work at University of Regina;

My thesis supervisory committee members: Drs. Zhangxing Chen (External

Examiner, University of Calgary), Chun-Hua Guo, Ezeddin Shirif, Farshid

Torabi, Yee-Chung Jin, and for their valuable questions and suggestions;

Natural Sciences and Engineering Research Council (NSERC) of Canada for the

Discovery Grants awarded to Drs. Fanhua Zeng and Yongan Gu;

University of Regina for financial support in the form of Graduate Scholarship

through Faculty of Graduate Studies and Research;

Petroleum Technology Research Centre (PTRC) for the Innovation Funds given

to Drs. Fanhua Zeng and Yongan Gu;

My past and present research group members, Mr. Zuojing Zhu, Ms. Lijuan Zhu,

Ms. Suxin Xu, Mr. Tao Jiang, Ms. Xiaoqi Wang, Mr. Shiyang Zhang, Mr.

Mohammad Derakhshanfar, Mr. Xiang Zhou, Mr. Zhongwei Du, and Ms.

Shanshan Yao, for their helpful technical discussions and suggestions during my

Ph.D. studies; and

My friends (Jim Jacobson, Bettie Jacobson, Graham Beke, Debra Beke, Garry

Engler, and friends from Bethany Gospel Chapel) for their care, concern, and

friendship.

Page 6: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

iv

DEDICATION

To my family and friends

especially my girlfriend, Jianli Li,

and my parents, Zhizhou Jia and Xiling Xi,

for their unconditional love, understanding, and support.

Page 7: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

v

TABLE OF CONTENTS

ABSTRACT…… ............................................................................................................ i

ACKNOWLEDGMENTS ............................................................................................. iii

DEDICATION…. ..........................................................................................................iv

TABLE OF CONTENTS ................................................................................................. v

LIST OF TABLES..........................................................................................................ix

LIST OF FIGURES ......................................................................................................... x

NOMENCLATURE ...................................................................................................... xv

CHAPTER 1 INTRODUCTION ................................................................................ 1

1.1 Heavy Oil Resources ..........................................................................................1

1.2 Heavy Oil Recovery Techniques ........................................................................1

1.3 Technical Challenges in Thin Heavy Oil Reservoirs ...........................................3

1.4 Solvent-Based EOR Techniques .........................................................................3

1.5 Problem Statement and Research Objectives ......................................................5

1.5.1 Theoretical modeling of VAPEX .................................................................5

1.5.2 Modeling of the mass transfer in CSI ...........................................................5

1.5.3 Performance improvement for VAPEX and CSI ..........................................6

1.5.4 Research objectives .....................................................................................6

1.6 Thesis Outline ....................................................................................................7

CHAPTER 2 LITERATURE REVIEW ..................................................................... 8

2.1 Vapour Extraction (VAPEX) ..............................................................................8

2.1.1 Physical modeling of VAPEX ................................................................... 10

2.1.2 Mass transfer modeling of VAPEX............................................................ 24

2.1.3 Theoretical modeling of VAPEX ............................................................... 28

2.1.4 Numerical modeling of VAPEX ................................................................ 34

2.2 Cyclic Solvent Injection (CSI) .......................................................................... 35

2.3 Chapter Summary ............................................................................................ 39

CHAPTER 3 MATHEMATICAL MODELING OF VAPEX ................................... 40

3.1 Mathematical Model and Solution .................................................................... 40

3.1.1 Heavy oil–solvent transition zone .............................................................. 40

Page 8: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

vi

3.1.2 Mass transfer in transition zone ................................................................. 44

3.1.3 Fluid flow in transition zone ...................................................................... 47

3.1.4 Moving boundary of transition zone .......................................................... 48

3.1.5 Solution procedures ................................................................................... 50

3.1.6 Heavy oil production rate .......................................................................... 52

3.2 Results and Discussion ..................................................................................... 54

3.2.1 Solvent chamber evolution and recovery factor ......................................... 54

3.2.2 Number of transition-zone segments .......................................................... 59

3.2.3 Permeability .............................................................................................. 59

3.2.4 This study vs. analytical models ................................................................ 65

3.2.5 This study vs. numerical simulation ........................................................... 65

3.3 Chapter Summary ............................................................................................ 78

CHAPTER 4 MATHEMATICAL MODELING OF THE

CONVECTION−DIFFUSION MASS-TRANSFER PROCESS .......... 79

4.1 CSI Process ...................................................................................................... 79

4.1.1 Convection−diffusion equation .................................................................. 81

4.1.2 Diffusion coefficient and convection velocity ............................................ 81

4.2 Mathematical Models ....................................................................................... 84

4.2.1 Governing equation ................................................................................... 84

4.2.2 Boundary and initial conditions ................................................................. 84

4.3 Semi-Analytical Solutions ................................................................................ 86

4.3.1 Model 1: Convection–diffusion model with constant D and variable V ...... 86

4.3.2 Model 2: Convection–diffusion model with variable D and variable V ....... 91

4.4 Validations ....................................................................................................... 93

4.4.1 Validation with an analytical solution for a special case............................. 93

4.4.2 Validation with the numerical solution ...................................................... 95

4.5 Results and Discussion ..................................................................................... 95

4.5.1 Application of the convection–diffusion mass-transfer model .................... 95

4.5.2 Variable and constant diffusion coefficient and convection velocity .......... 97

4.5.3 Effect of convection velocity ..................................................................... 99

4.5.4 Péclet number .......................................................................................... 104

Page 9: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

vii

4.5.5 Effect of gravity force in natural convection ............................................ 106

4.6 Chapter Summary .......................................................................................... 110

CHAPTER 5 FOAMY OIL-ASSISTED VAPOUR EXTRACTION (F-VAPEX)... 111

5.1 Experimental .................................................................................................. 111

5.1.1 Materials ................................................................................................. 111

5.1.2 Experimental set-up ................................................................................. 112

5.1.3 Experimental preparation ......................................................................... 116

5.1.4 Experimental procedure ........................................................................... 118

5.1.5 Other measurements ................................................................................ 120

5.2 Results and Discussion ................................................................................... 121

5.2.1 Foamy oil flow in F-VAPEX ................................................................... 121

5.2.2 F-VAPEX vs. VAPEX/CSI ..................................................................... 127

5.2.3 Effect of well configuration ..................................................................... 134

5.2.4 Residual oil saturation ............................................................................. 145

5.3 Chapter Summary .......................................................................................... 148

CHAPTER 6 GASFLOODING-ASSISTED CYCLIC SOLVENT INJECTION

(GA-CSI) .......................................................................................... 149

6.1 Experimental .................................................................................................. 149

6.1.1 Materials ................................................................................................. 149

6.1.2 Experimental set-up ................................................................................. 150

6.1.3 Experimental preparation ......................................................................... 152

6.1.4 Experimental procedure ........................................................................... 152

6.2 Results and Discussion ................................................................................... 155

6.2.1 Well configuration ................................................................................... 158

6.2.2 Operating scheme (CSI vs. GA-CSI) ....................................................... 162

6.2.3 GA-CSI ................................................................................................... 164

6.2.4 Solvent injection rate ............................................................................... 169

6.2.5 GA-CSI with cylindrical models .............................................................. 169

6.2.6 GA-CSI with rectangular model .............................................................. 172

6.2.7 Residual oil saturation ............................................................................. 172

6.3 Variations of GA-CSI..................................................................................... 177

Page 10: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

viii

6.3.1 Pressure control scheme .......................................................................... 177

6.3.2 Viscous fingering .................................................................................... 180

6.3.3 Oil production ......................................................................................... 182

6.4 Chapter Summary .......................................................................................... 185

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS ............................. 186

7.1 Conclusions.................................................................................................... 186

7.2 Recommendations .......................................................................................... 189

REFERENCES…. ....................................................................................................... 190

Appendix A……. ........................................................................................................ 206

Appendix B……. ......................................................................................................... 208

Appendix C……. ......................................................................................................... 210

Appendix D……. ........................................................................................................ 211

Page 11: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

ix

LIST OF TABLES

Table 2.1 VAPEX experimental studies by Butler’s group. ..................................... 15

Table 2.2 VAPEX experimental studies by Maini’s group. ..................................... 16

Table 2.2 VAPEX experimental studies by Maini’s group (Contd’). ....................... 16

Table 2.3 VAPEX experimental studies by Gu’s group. .......................................... 18

Table 2.4 VAPEX experimental studies by ARC. ................................................... 19

Table 2.5 VAPEX experimental studies by other research groups. .......................... 20

Table 2.6 Comparison of the measured diffusion coefficients of CO2, CH4, C2H6

and C3H8 in different heavy oil and bitumen samples. ............................. 26

Table 2.7 CSI experimental studies in the literature. ............................................... 38

Table 3.1 Parameters of the base case for the mathematical model. ......................... 55

Table 3.2 Parameters of the base case for the numerical simulation. ..................... 667

Table 3.3 Effect of the grid size and estimation of the numerical dispersion. ......... 722

Table 4.1 Parameters of the base case. .................................................................... 83

Table 5.1 Physical properties of the sand-packed models and experimental

conditions for VAPEX, CSI, and F-VAPEX testss ................................ 115

Table 5.2 Cumulative heavy oil production data. .................................................. 129

Table 6.1 Physical properties of the sand-packed models and experimental

conditions for CSI and GA-CSI tests. .................................................... 153

Table 6.2 Cumulative oil and solvent production data. .......................................... 156

Page 12: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

x

LIST OF FIGURES

Figure 2.1 The VAPEX heavy oil recovery process....................................................9

Figure 2.2 Solvent vapour chamber profiles at the end of (a) Rising phase; (b)

Spreading phase; and (c) Falling phase [Zhang, et al., 2006]. .................. 22

Figure 3.1 Transition zone in the VAPEX process. .................................................. 41

Figure 3.2 Approximation to the transition-zone at the (a) beginning and (b)

middle stages of the VAPEX process. ..................................................... 43

Figure 3.3 Boundary movement of a transition-zone segment. ................................. 49

Figure 3.4 Flowchart of the solution calculation for the VAPEX model. .................. 51

Figure 3.5 Discretization of the space and time domains for the numerical solution

to the mass-transfer model with a moving boundary condition. ............... 53

Figure 3.6 Evolution of the solvent vapour chamber during a VAPEX process. ....... 56

Figure 3.7 Oil recover factor of the VAPEX base case . ........................................... 57

Figure 3.8 Solvent concentration distribution at different locations along the

transition zone at different moments: (a) Solvent chamber profiles; (b)

Solvent concentration distributions at the top; (c) Solvent concentration

distributions in the middle; and (d) Solvent concentration distributions

at the bottom of the transition zone.......................................................... 58

Figure 3.9 Effect of dividing number on the average oil production rate................... 60

Figure 3.10 Effect of permeability on the solvent chamber evolution: (a) k = 25 d;

(b) k = 50 d; (c) k = 100 d; and (d) k = 200 d. .......................................... 62

Figure 3.11 Effect of permeability on the oil production rate...................................... 63

Figure 3.12 (a) Heavy oil production rate vs. Square root of diffusion coefficient;

and (b) Oil recovery factor for variable and constant diffusion

coefficients. ............................................................................................ 64

Figure 3.13 Oil production rate predicted by this study and the existing VAPEX

models. ................................................................................................... 66

Figure 3.14 (a) Numerical simulation model; (b) Relative permeability vs. liquid

saturation; and (c) Capillary pressure vs. liquid saturation. ...................... 68

Page 13: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xi

Figure 3.15 Effect of the timestep size on the cumulative oil production (grid size:

0.01 × 0.01 × 0.01 m3). ........................................................................... 70

Figure 3.16 Effect of the grid size on the cumulative oil production (t = 0.001 d). ... 71

Figure 3.17 Effect of the diffusion coefficient on the heavy oil production rate: (a)

Lab-scale grid size simulation results; and (b) Field-scale grid size

simulation results. ................................................................................... 74

Figure 3.18 Mole fraction of solvent in the lab-scale numerical model with different

grid-sizes at 20 h (t = 0.01 d): (a) 0.02 × 0.02 × 0.02 m3; (b) 0.01 ×

0.01 × 0.01 m3; and (c) 0.005 × 0.005 × 0.005 m3. .................................. 75

Figure 3.19 Comparison of the predicted transition-zone thicknesses of this study

and numerical simulation (grid size: 0.02×0.02×0.02 m3; t = 0.01 d). .... 77

Figure 4.1 Vapour solvent-based ‘huff-n-puff’ process (note: bold white arrows

point to the solvent diffusion direction, whereas narrow black arrows

point to convection direction). ................................................................. 80

Figure 4.2 Concentration-dependent diffusion coefficient and flow velocity: (a)

Concentration; (b) Viscosity; (c) Effective diffusion coefficient; and (d)

Convection velocity. ............................................................................... 85

Figure 4.3 Approximation to the convection velocity with a piecewise linear

profile. .................................................................................................... 90

Figure 4.4 Semi-analytical vs. Analytical cD for a convection–diffusion mass

transfer with a special convection velocity .............................................. 94

Figure 4.5 Semi-analytical vs. Numerical cD............................................................. 96

Figure 4.6 Flowchart of calculating the solvent concentration in the transition

zone (t* denotes the termination time). .................................................... 98

Figure 4.7 Comparison of cD for different cases: (a) Variable d & variable V vs.

constant D & variable V; (b) Variable D & variable V vs. variable D &

constant V; (c) Variable D & variable V vs. constant D & constant V;

and (d) Variable D & variable V and D vs. constant D & variable V vs.

variable D & constant V vs. constant D & constant V (constant d is

equal to 5.8×10−9 m2/s; constant v is equal to 1.8×10−6 m/s; t1 = 300 s; t2

= 600 s). ................................................................................................ 100

Page 14: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xii

Figure 4.8 Effect of the pressure gradient on the solvent concentration distribution.

......................................................................................................... 102

Figure 4.9 Effect of crude oil viscosity on the solvent concentration distribution.... 103

Figure 4.10 Effect of diffusion coefficient on the solvent concentration distribution.

......................................................................................................... 105

Figure 4.11 Effect of péclet number on the solvent concentration distribution. ......... 107

Figure 4.12 Effect of péclet number with different linear shape on the solvent

concentration distribution. ..................................................................... 108

Figure 4.13 Effect of gravity force on the solvent concentration distribution. ........... 109

Figure 5.1 Schematic diagram of the experimental set-up in this study. .................. 113

Figure 5.2 (a) Physical model dimensions; (b) Central well configuration; and (c)

Lateral well configuration. .................................................................... 114

Figure 5.3 Pressure-control scheme for (a) VAPEX; (b) CSI; and (c) F-VAPEX. .. 119

Figure 5.4 Injection and production pressure data during a typical F-VAPEX

cycle. .................................................................................................... 123

Figure 5.5 Foamy oil zone (a) before and (b) after foamy oil flow during a

pressure reduction period of an F-VAPEX process (Test #5.3). ............. 124

Figure 5.6 Foamy oil zone during the (a) early, (b) middle, and (c) late stages of

an F-VAPEX test (Test #5.3). ............................................................... 126

Figure 5.7 Cumulative oil production versus time data for the VAPEX, CSI and

F-VAPEX tests with the central well configuration. .............................. 130

Figure 5.8 Cumulative oil production versus time data for the CSI and F-VAPEX

tests with the lateral well configuration. ................................................ 131

Figure 5.9 Enhancement of the oil production rate of VAPEX by F-VAPEX with

different well configurations. ................................................................ 132

Figure 5.10 Enhancement of the oil production rate of CSI by F-VAPEX with

different well configurations. ................................................................ 133

Figure 5.11 Cumulative solventoil ratio versus time data for the VAPEX, CSI, and

F-VAPEX tests with the central well configuration. .............................. 135

Figure 5.12 Cumulative solventoil ratio versus time data for the VAPEX and

F-VAPEX tests with the lateral well configuration. ............................... 136

Page 15: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xiii

Figure 5.13 Foamy oil zone during the (a) early, (b) middle, and (c) late stages of

an F-VAPEX test with the lateral well configuration (Test #5.7). .......... 137

Figure 5.14 Oil production data from the stable pressure period and pressure

reduction period during Test #5.3. ......................................................... 139

Figure 5.15 Oil production from the stable pressure period and pressure reduction

period during Test #5.7 ......................................................................... 140

Figure 5.16 Total oil production from the stable pressure period and pressure

reduction period during the F-VAPEX tests. ......................................... 141

Figure 5.17 Total solvent production data in the stable pressure period and the

pressure reduction period of the F-VAPEX tests. .................................. 144

Figure 5.18 Residual oil saturation at the end of (a) Test #5.1; (b) Test #5.3; and (c)

Test #5.7. .............................................................................................. 146

Figure 6.1 (a) Schematic diagram of the experimental set-up with a cylindrical

model for GA-CSI tests and a CSI test; (b) Dimensions of the

rectangular sand-packed model; and (c) Schematic diagram of the

physical model for a CSI test. ................................................................ 151

Figure 6.2 Pressure-control schemes of (a) GA-CSI and (b) CSI. ........................... 154

Figure 6.3 (a) Cumulative oil production; and (b) SOR of Tests #6.1−3. ................ 157

Figure 6.4 ‘Back-and-forth movement’ of the solvent-diluted heavy oil in a CSI

test: (a) Solvent dissolution into oil during the injection period of a

cycle; (b) Diluted oil flowing to the producer during the production

period; (c) Some diluted oil remaining in the solvent chamber at the

end of the production period; and (d) Diluted oil flowing back during

the solvent injection period of the next cycle. ........................................ 160

Figure 6.5 ‘Back-and-forth movement’ of the solvent-diluted heavy oil during a

cycle of the CSI test (Cycle #40 of Test #6): (a) Oil flowing to the

producer at the early stage of the production period; (b) Oil remaining

in the solvent chamber at the end of the production period; and (c) Oil

flowing back during the solvent injection period of the next cycle

(Cycle #41). .......................................................................................... 161

Page 16: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xiv

Figure 6.6 Gasflooding process during a GA-CSI test (Test #6.7). (a) End of the

blowdown stage; (b) Early gasflooding stage; and (c) Late gasflooding

stage. .................................................................................................... 163

Figure 6.7 Injection and production pressures and the solvent injection rate during

a typical cycle (Cycle #4) of a GA-CSI test (Test #6.3). ........................ 165

Figure 6.8 Cumulative oil production, oil production rate, and solvent–oil ratio

during a typical cycle (Cycle #4) of a GA-CSI test (Test #6.3). ............. 166

Figure 6.9 (a) Heavy oil production; and (b) Solvent gas production during the

blowdown and gasflooding slugs of the production period of a GA-CSI

test (Test #6.3). ..................................................................................... 167

Figure 6.10 Cumulative oil productions of Tests #6.3 (blowdown slugs only), and

#6.1 and #6.2. ....................................................................................... 168

Figure 6.11 Solvent injection rate at early, middle, and late stages of a GA-CSI test

(Test #6.3). ........................................................................................... 170

Figure 6.12 (a) Recovery factor; and (b) Solventoil ratio of the GA-CSI tests with

cylindrical models of different lengths. ................................................. 171

figure 6.13 Oil recovery factor of GA-CSI and CSI tests with the rectangular

physical model. ..................................................................................... 173

Figure 6.14 Residual oil saturation of (a) CSI (Test #6.2); and (b) GA-CSI tests

(Test #6.3). ........................................................................................... 174

Figure 6.15 Residual oil saturation of (a) CSI (Test #6.6); and (b) GA-CSI (Test

#6.7). .................................................................................................... 176

Figure 6.16 Pressure control scheme of PP-CSI. ...................................................... 178

Figure 6.17 Injection and production pressures data during a PP-CSI test. ............... 179

Figure 6.18 Evolution of the solvent chamber throughout a PP-CSI test: (a) Cycle

#2; (b) Cycle #4; (c) Cycle #8; and (d) Cycle #12. ................................ 181

Figure 6.19 Comparison of the oil recovery factor of PP-CSI and GA-CSI tests. ...... 183

Figure 6.20 Oil production from multiple pulses in different cycles of Test #6.8 ...... 184

Page 17: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xv

NOMENCLATURE

Notations a slope of a linear Pe profile

'a slope of a linear dimensionless diffusion coefficient profile

A, B coefficients

b intercept of a linear Pe profile

'b intercept of a linear dimensionless diffusion coefficient profile

c solvent concentration in the solvent-diluted heavy oil, vol.%

c* solvent concentration under a operating pressure, vol.%

cD dimensionless concentration, dimensionless

cmax maximum solvent concentration in a crude oil, vol.%

cmin minimum solvent concentration in a crude oil, vol.%

C modified dimensionless concentration, dimensionless

D diffusion coefficient, m2/s

Dapp apparent diffusion coefficient in the Das−Butler model, m2/s

DD dimensionless diffusion coefficient, dimensionless

F formation electrical resistivity factor

fo, fs weighted volume fractions of crude oil and solvent in Lederer’s equation, fraction

g gravity acceleration, m/s2

H model height, m

h vertical distance between an arbitrary point and the model bottom, m

k permeability, m2

L length of the horizontal section of a horizontal well running VAPEX, m

l length of a transition-zone segment, m

M Kummer’s function

N number of grids

Ns dimensionless number in the Butler−Morkys VAPEX model

Nτ number of substeps in a time step

P pressure, kPa

Page 18: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xvi

Pe Péclet number, dimensionless

q oil drainage rate, m3/s

qo stabilized oil production rate in the Butler−Mokrys/Yazdani-Maini model, m3/s

qin flowrate of solvent-diluted heavy oil entering into a transition zone segment, m3/s

qout flowrate of solvent-diluted heavy oil leaving from a transition zone segment, m3/s

Q cumulative oil production, m3

Qo cumulative oil production in Moghadam et al. model, m3/s

s Laplacian operator

S oil saturation, vol.%

S’ source/sink in the convection−diffusion equation

Soi initial oil saturation, vol.%

Sor residual oil saturation, vol.%

Swr residual water saturation, vol.%

t time, s

tD dimensionless distance

U transition-zone boundary moving velocity, m/s

V convection velocity, m/s

Vp pore volume, m3

Vs sand volume, m3

Vw water volume, m3

V Darcy flow rate of solvent-diluted heavy oil in the transition zone, m/s

W model width, m

w width, m

x x coordinate, m

xD dimensionless distance

y y coordinate, m

z transformed dimensionless distance

Page 19: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xvii

Greek Symbols

α coefficient of viscosity

β coefficient of viscosity

specific gravity

o, s specific gravities of crude oil and liquid solvent

δ transition-zone thickness, m

θ inclination angle of transition zone, degree

λ weight factor in Shu’s equation

μ viscosity of the solvent-diluted heavy oil, mPas

μo, μs viscosities of crude heavy oil and liquid solvent, mPas

ξ distance from the boundary between the solvent chamber and transition zone to an

arbitrary point in the transition zone, m

ξ0 location of the transition-zone boundary at the beginning of a time step, m

ξmax location of the transition-zone boundary next to the solvent chamber, m.

ξmin location of the transition-zone boundary next to the untouched heavy oil zone, m.

ξmv distance of a transition-zone segment moved over a time step, m.

ρ density of solvent-diluted heavy oil, kg/m3

ρo, ρs densities of crude oil and liquid solvent, kg/m3

τ time, s

ψ arbitrary constant

porosity, vol.%

cementation factor, dimensionless

Page 20: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xviii

Subscripts

Abbreviations

app apparent

D dimensionless

dew dew point

o oil

oi initial oil

or Residual oil

out outflow

i ith time interval

in inflow

max maximum

min minimum

mv movement

p pore

s solvent

τ time

ARC Alberta Research Council

CHOPS Cold Heavy Oil Production with Sands

CMG Computer Modelling Group

CPCSI Cyclic Production with Continuous Solvent Inejction

CSI Cyclic Solvent Injection

CSS Cyclic Steam Stimulation

CT Computer Tomography

DPDVA Dynamic Pendant Drop Volume Analysis

EOR Enhanced Oil Recovery

F-VAPEX Foamy Oil-Assisted Vapour Extraction

GA-CSI Gasflooding-Assisted Cyclic Solvent Injection

GEM Generalized Equation of State Model Reservoir Simulator

Page 21: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

xix

Units

ISC In-Situ Combustion

OOIP Original-Oil-In-Place

RF Recovery Factor

SAGD Steam-Assisted Gravity Drainage

SAS Steam Alternating Solvent

SOR Solvent−Oil Ratio

SRC Saskatchewan Research Council

STARS Steam, Thermal and Advanced Reservoir Simulator

VAPEX Vapour Extraction

API American Petroleum Institute gravity

C Celsius

bbl barrel

cc cubic centimeter

cm centimeter

D Darcy

dm decimeter

g gram

h hour

kg kilogram

kPa kilopascal

m meter

m2/s square meter per second

m/s2 meter per square second

min minute

ml mililiter

MPa megapascal

mPas milipascal-second

s second

Page 22: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

1

CHAPTER 1 INTRODUCTION

1.1 Heavy Oil Resources

Effective and economical recovery of unconventional heavy oil and bitumen

resources has become a key technical challenge due to the depletion of conventional

petroleum resources and the increase of hydrocarbon fuel demands. In comparison with

conventional crude oil, heavy oil and bitumen are much more viscous and heavier, and they

are characterized by high viscosities (i.e., higher than 100 mPas for heavy oil and 10,000

mPas for bitumen) and low API (American Petroleum Institute) gravities (i.e., lower than

20.0API for heavy oil and 10.0API for bitumen) [Speight, 1991].

In the world, the total crude oil resources are approximately 9−11 trillion bbls,

among which more than 2/3 are unconventional heavy oil and bitumen [Dusseault, 2001].

Out of the total eight trillion bbls of heavy oil and bitumen reserve, Canada and Venezuela

each possesses 2–3 trillion barrels. In Canada, heavy oil and bitumen resources are found in

Western Canada, mainly in Alberta and Saskatchewan with an estimated

original-oil-in-place (OOIP) of 2.5 trillion barrels [Petroleum Communication Foundation,

2000; Dusseault, 2001; Farouq Ali, 2003]. Most of western Canadian heavy oil and

bitumen deposits are located in the three major basins in northern Alberta: Athabasca, Cold

Lake, and Peace River.

1.2 Heavy Oil Recovery Techniques

In general, there are two kinds of heavy oil and bitumen recovery methods: open-pit

mining and in-situ methods [Butler and Yee, 2002]. Open-pit mining methods are used to

Page 23: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

2

recover minable bitumen deposits that are less than 100 m deep [Petroleum

Communication Foundation, 2000]. It relies on massive earth-moving equipment and

processing facilities, and has limited future capacity since 80 percent of the oil sand

resources lie deep underground and are not accessible by open-pit mining. The latter

extraction methods include three categories: primary production techniques, thermal-based

techniques, and non-thermal-based techniques. These in-situ heavy oil recovery methods

currently rely on the injection of energy-intensive steam and large volumes of natural gas.

In most cases, only 5–10% of the original-oil-in-place (OOIP) can be recovered from

western Canadian reservoirs after primary and secondary oil recovery processes, such as

cold heavy oil production with sand (CHOPS) and waterflooding. Afterward, these

techniques become uneconomical due to reservoir pressure depletion and/or water

encroachment to the production well [Ivory et al., 2010]. Therefore, the latter two

enhanced oil recovery (EOR) techniques are resorted to produce the heavy oil and bitumen

reserves.

Thermal-based methods, such as steam-assisted gravity drainage (SAGD), cyclic

steam stimulation (CSS), and in-situ combustion (ISC) [Butler et al., 1981; Vittoratos et al.,

1990; Moore et al., 1995] can drastically reduce the crude oil viscosity by means of thermal

energy. Specifically, SAGD and CSS have achieved great success in heavy oil reservoirs

with a thickness larger than 10 m. However, many Canadian heavy oil reservoirs have thin

pay zones, for which the thermal-based methods become uneconomical due to large heat

losses to the overburden and underburden.

Solvent-based methods, such as solvent vapour extraction (VAPEX) [Butler and

Mokrys, 1991; Das, 1998], is not an apparent option because of its inefficient gravity

Page 24: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

3

drainage and extremely low oil production rate. As another type of solvent-based

techniques, cyclic solvent injection (CSI) [Lim et al., 1995, 1996; Ivory et al., 2010; Firouz

et al., 2012] has emerged as a promising follow-up process of CHOPS in recent years.

1.3 Technical Challenges in Thin Heavy Oil Reservoirs

A large number of Canadian heavy oil reserves are located in thin reservoirs,

especially in Saskatchewan. Saskatchewan accounts for almost 62% of Canada’s total

heavy oil resources, including 1.7 billion m3 of proven and 3.7 billion m3 of probable

reserves. According to Reservoir Annual [Saskatchewan Energy and Mines, 2000] of the

province's proven initial heavy oil-in-place, 97% is contained in reservoirs with less than

10 m pay zones, and 55% is in reservoirs less than 5 m thick. Primary and secondary

methods combined recover less than 10% of the OOIP, on average. Hence, there is a strong

incentive for the development of appropriate oil recovery techniques, which will maximize

the recovery potential of these thin heavy oil reservoirs [Dong et al., 2006].

The heavy oil and bitumen recovery from these reservoirs with displacement or

thermal recovery processes is neither economically viable nor environmentally friendly

because of the accompanying losses of displacement fluid or energy to the overburden and

underburden. Besides, these methods require huge amounts of water and solvent gas and

vast surface facilities, and are inefficient in the frequently-encountered thin heavy oil

reservoirs. Therefore, it is necessary to develop proper recovery methods to maximize the

recovery potential of the profitability from these thin heavy oil reservoirs.

1.4 Solvent-Based EOR Techniques

In the literature, extensive studies have been conducted to explore the potential of

Page 25: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

4

solvent-based EOR methods, including VAPEX and CSI. VAPEX is a direct experimental

and theoretical analog of SAGD. In the VAPEX process, gaseous condensable solvents,

such as propane and butane [Butler et al., 1995] in conjunction with non-condensable

carrier gases, such as methane and carbon dioxide [Talbi and Maini, 2003], are used to

extract heavy oil and bitumen from reservoir formations. The major oil recovery

mechanisms in this process consist of viscosity reduction through solvent dissolution and

possible asphaltene precipitation and gravity drainage of the solvent-diluted heavy oil. CSI

is basically a solvent huff-n-puff process. It is considered as a solvent-analog of CSS.

The potential advantages of solvent-based EOR methods over thermal-based EOR

methods are: (1) Cost-effectiveness. Solvent-based techniques do not involve large surface

facilities. Therefore, it saves the cost for steam generation equipment and the consequent

costs for operation and treatment of the produced wastewater. For example, VAPEX

requires only approximately 3% of the energy needed for SAGD for the same production

rate [Singhal, et al, 1997]. (2) Environmental friendliness. Solvent-based techniques

produce much less water than thermal-based techniques. Thereby, they would less possibly

cause the environment pollution by the produced wastewater. Moreover, greenhouse gas

emission would be greatly reduced since over 80% of the produced solvent can be captured

and reused in a solvent-based EOR process [Butler and Mokrys, 1991]. (3) Oil in-situ

upgrading. The heavier component of crude oil might be precipitated in the reservoir in the

process of solvent dissolution into heavy oil, which makes the heavy oil become lighter.

This is beneficial for the subsequent oil transportation and processing.

The major disadvantage of solvent-based EOR methods is its low oil production rate,

especially in some thin heavy oil reservoirs. For VAPEX, it is because of the slow mass

Page 26: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

5

transfer and inadequate gravity drainage. For CSI, it might be due to the unproductive, long

injection and soaking periods and the relatively short production period.

1.5 Problem Statement and Research Objectives

1.5.1 Theoretical modeling of VAPEX

Theoretical modeling of VAPEX has not achieved as much progress as the physical

modeling in the past two decades. Existing analytical models are established on the basis of

some major assumptions for the heavy oilsolvent transition zone, such as constant

boundary moving velocity and steady-state mass transfer. They are unable to describe the

solvent chamber evolution. Numerical simulation models use grid sizes much larger than

the transition zone thickness (~1 cm), which makes it difficult for these models to capture

the heavy oil and solvent properties inside the transition zone.

1.5.2 Modeling of the mass transfer in CSI

CSI is a solvent huff-n-puff process. One of the differences between VAPEX and

CSI is that the operating pressure is cyclically decreased and increased in CSI, whereas it is

maintained at a constant value in VAPEX. Hence, in addition to molecular diffusion, the

heavy oil−solvent mixing process in CSI is influenced also by another mechanism,

convection. Convection describes the mass transfer through a bulk motion of the solvent

due to the pressure gradient in the CSI process. So far, few studies have been done to

describe the mass transfer process in CSI. Existing mass-transfer models for VAPEX are

based on Fick’s 2nd law and do not consider the effect of pressure gradient across the

transition zone. In addition, the diffusion coefficient or the dispersion coefficient in

existing models is usually assumed to be a constant, which is not true in the actual cases.

Page 27: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

6

1.5.3 Performance improvement of VAPEX and CSI

The major limitation of the VAPEX process is its extremely low oil production rate,

especially in thin heavy oil reservoirs. This is caused by: (1) The small diffusion

coefficient; (2) Limited contact area between solvent and heavy oil; (3) ‘Concentration

shock’ [Ninniger and Dunn, 2008]; and (4) Inefficient gravity drainage. The first three are

inherent properties of the VAPEX process, resulting in a low mass transfer rate between

the solvent and crude heavy oil. In particular, the concentration shock makes it difficult for

the solvent to pass through the transition zone to dilute the fresh heavy oil. The inefficient

gravity drainage is due to the small inclination angle, especially in thin reservoirs.

Although CSI has stronger driving forces (solution-gas drive and foamy oil flow) for

heavy oil recovery, its technical limitations are as follows: (1) The solvent injection and

soaking periods are unproductive and long, whereas the oil production period is relatively

short. This leads to a low average oil production rate over the entire process. (2) The oil

production rate declines fast and most of the oil production occurs during the early stage of

the production period. This is because the solvent disengages from the oil due to pressure

depletion during the production period, which leads the oil to regain its high viscosity and

eventually lose its mobility. Hence, a considerable amount of solvent-diluted heavy oil

becomes motionless and remains in the reservoir at the end of the production period.

1.5.4 Research objectives

Aiming at the aforementioned technical issues with VAPEX and CSI, this thesis

wants to achieve the following objectives:

1. To develop a new 2D mathematical model to describe the solvent-chamber

evolution during the VAPEX process;

Page 28: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

7

2. To develop a new mass-transfer model to study the effects of the pressure

gradient on the heavy oil−solvent mixing process during the CSI process;

3. To design new operating schemes to enhance the oil production rate of the

conventional VAPEX process; and

4. To design new operating schemes to enhance the performance of the

conventional CSI process.

1.6 Thesis Outline

This thesis is composed of seven chapters. Specifically, Chapter 1 gives an

introduction to the thesis research topic together with the purpose and scope of this study.

Chapter 2 provides an up-to-date literature review on the solvent-based EOR techniques,

such as VAPEX and CSI. Chapter 3 describes a new 2D mathematical model for the

solvent-chamber evolution during the VAPEX process. The mathematical model,

semi-analytical solution, and data analysis are presented in this chapter. Chapter 4 develops

a new convection−diffusion mass-transfer model to investigate the effect of the pressure

gradient on the heavy oil−solvent mixing process during the CSI process. Chapter 5

proposes a new modified VAPEX technique, namely foamy oil-assisted vapour extraction

(F-VAPEX). The experimental set-up, operating scheme, and data analysis of the

F-VAPEX process are presented in this chapter. Chapter 6 presents another a novel

technique, namely gasflooding-assisted cyclic solvent injection (GA-CSI). The special

operating scheme of GA-CSI and its experimental results are described and discussed.

Chapter 7 summarizes the major scientific findings of this thesis study and provides some

technical recommendations for future studies.

Page 29: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

8

CHAPTER 2 LITERATURE REVIEW

2.1 Vapour Extraction (VAPEX)

VAPEX was first studied as a solvent-analogy of SAGD by Butler and Mokrys in

1989. In a typical VAPEX process, a gaseous solvent (typically a lighter hydrocarbon gas)

is injected into a reservoir formation through an upper horizontal injection well. The heavy

oil is diluted by the solvent in the transition zone, drained downward by gravity, and

produced from a lower horizontal well (Figure 2.1). Three zones are formed during this

process: a solvent chamber, an untouched heavy oil zone, and a transition zone in between.

The VAPEX process occurs in the following way: (1) Dissolution of solvent into oil at the

transition zone; (2) Diffusion of solvent molecules in the bulk heavy oil; (3) Reduction of

heavy oil viscosity as the solvent concentration increases; (4) Above a critical

concentration, asphaltene precipitation takes place, further reducing the oil viscosity; and

(5) Due to the effects of Steps #34 and the difference in density between the liquid oil and

gaseous solvent, solvent-diluted heavy oil drains downward along the transition zone to the

production well by gravity and capillary imbibitions.

Page 30: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

9

[http://www.japex.co.jp]

Figure 2.1 The VAPEX heavy oil recovery process.

Page 31: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

10

2.1.1 Physical modeling of VAPEX

A summary of VAPEX laboratory experiments conducted by several research

groups in the past two decades is presented in Table 2.1–5. It includes several major

research groups in VAPEX in Canada, such as Bulter’s group, Maini’s group, and Gu’s

group. The research topics are mainly the traditional VAPEX process. This section first

introduces the previous research work on VAPEX group by group and then summarizes

some major technical aspects of VAPEX that are of interest in this thesis.

Table 2.1 shows the experimental results achieved by Dr. Butler’s group. Butler and

Mokrys [1989, 1991, 1993] used toluene to extract Athabasca and Suncor bitumen samples

in a Hele−Shaw cell. They found that at low permeabilities, oil rate is a linear function of

the square root of the permeability. At high permeabilities, the drainage rate is nonlinear

and approaches asymptotically a constant value that is independent of the permeability.

Das and Butler [1995] conducted VAPEX tests to examine the effect of bottom water on

the VAPEX performance. It was observed that the water injection in a small quantity along

with the solvent and non-condensable gas enhances the extraction rate at the initial stage.

The solvent-diluted heavy oil is efficiently displaced by the injected water which was the

wetting fluid in most oil reservoirs. Major parameters affecting VAPEX performance were

thoroughly studied by Butler and Jiang [1996, 1997] in order to develop optimum

operating conditions for high oil production rates with economical solvent requirements.

The parameters investigated were temperature, pressure, solvent injection rates, pure

solvent type, mixed solvent, well spacing, and well configurations etc. It was found that

propane works better than butane and a mixture of propane and butane works as well as

propane alone. A wider lateral well spacing allows a higher oil production rates and makes

Page 32: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

11

the process more economical. A high start-up solvent-injection rate followed by a reduced

rate performs better than a constant solvent-injection rate.

Table 2.2 shows the VAPEX experimental results of Maini’s group. Boustani and

Maini [2001] undertook a number of experiments with a Hele−Shaw cell to identify the

main process that governs the interfacial mass transfer of solvent into bitumen. The

apparent diffusion coefficient [Das and Butler, 1995] is based on the correlations

developed by Hayduk et al. [1976] as well as the experimental data. It was found that Das

and Butler’s correlations tend to overestimate the diffusion coefficient and underestimate

the overall mass-transfer dispersion coefficient in porous media. Talbi and Maini [2003]

studied a CO2-based VAPEX process for tar sand reservoirs. It was found from their

experimental results that CO2−propane mixture shows better performance in comparison

with the methane−propane mixture at a high operating pressure. The use of CO2 instead of

methane at lower operating pressures is thus justified.

Karmaker and Maini [2003] evaluated the VAPEX process for a reservoir with a

small gas cap. They found that a small gas cap is helpful for the application of VAPEX for

heavy oil recovery. A 1C increase in temperature could increase oil production by 2%.

The oil production rate almost doubled when the original oil viscosity is lowered by 15

times. A long delay in the start of oil production occurs for the increased lateral distance

between the injector and producer. They reinvestigated the oil drainage rate and examined

the scale-up methods for the VAPEX process with three physical models of different sizes.

It is found the grain size distribution does not make a difference on the oil drainage rate,

whereas the model height significantly increases the convective dispersion and the

consequent oil rate. Also they believed that higher oil rates than predicted were possible on

Page 33: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

12

the basis of the results from Hele−Shaw cell experiments and the available scale-up

method. Yazdani and Maini [2004, 2005] designed a new cylindrical model to overcome

the limitation of the rectangular models at higher pressures. The annular space between

two cylindrical pipes constructs the slice-type sand-packed models. It was found that the

stabilized oil drainage rates from their new cylindrical models agree perfectly with those

from the rectangular ones. The new models could also save laboratory space and

construction costs in comparison with flat models. Etminan and Maini [2007] evaluated the

effect of connate water on the VAPEX performance. They found the presence of connate

water causes faster spreading of the solvent vapour chamber in the lateral direction and

tends to increase the thickness of the mixing zone, which seems to be driven by capillary

fingering. In addition, the mobile water increases the oil production rate in the initial stage

and decreases it in the late stage of VAPEX. Moreover, oil deasphalting was found to be

more significant in the presence of connate water.

Zadeh et al. [2008] used a fixed CO2−propane mixture to produce the Athabasca

bitumen. They found it important to control the composition of the injected gas mixture to

avoid a multiple liquid-phase formation. They mapped experimentally and theoretically the

compositional change by using an EOS model during the VAPEX test. Haghighat and

Maini [2010] studied the effect of asphaltene precipitation on VAPEX performance to

determine whether the beneficial effects of asphaltene precipitation would outweigh the

detrimental ones. They found that at higher pressures, the produced oil was substantially

deasphalted but the viscosity was not drastically reduced as expected, i.e., in-situ

deasphalting did not lead to a higher production rate. In addition, the formation damage

caused by asphaltene precipitation and deposition seems irreparable through the huff and

Page 34: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

13

puff injection of toluene. However, a solvent mixture of propane and toluene was found to

be successful in increasing the oil production rate and upgrading the oil quality.

Table 2.3 shows the experimental results obtained by Dr. Gu’s group. Zhang et al.

[2007] carried out a series of VAPEX experiments with a visual rectangular sand-packed

high-pressure physical model, which can be used to visualize the entire VAPEX process,

throughout the vapour chamber rising, spreading, and falling phases. They predicted the oil

production rate by using the modified Butler−Mokrys analytical model and found a good

match with the experimental data. Moghadam et al. [2008] established a new theoretical

model on the basis of the incline angle of the transition zone, to predict the cumulative oil

production rate. The transition zone was assumed to have two straight-line boundaries with

a constant thickness during the VAPEX process. The adjustable transition-zone thickness

keeps almost constant during each VAPEX test and in general, it increases with the

decrease in the permeability of the VAPEX physical model. Furthermore, their results

showed that the horizontal spreading velocity of the solvent chamber is reduced with time

during the spreading phase, thus the heavy oil production rate during this phase declines

with time as the VAPEX process proceeds. Finally, the theoretical predictions showed that

the falling velocity of the solvent chamber is extremely low during its falling phase and

decreases with time as well.

Table 2.4 shows the experimental results obtained by ARC. Cuthiell et al. [2003]

implemented a series of top-down solvent injection experiments under varying conditions,

and the fluid movement was monitored by a CT scanner. They observed that oil production

rate becomes unstable before solvent breakthrough (BT), after which the displacement

remains steady. They also conducted numerical simulation to predict the oil production

Page 35: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

14

rate, the simulated BT time, post-BT oil production rate, and the general character of the

fingering. They found that the experimental data was matched after a certain amount of

physical dispersion is introduced. Frauenfeld et al. [2006] conducted a series of VAPEX

experiments with bottom water. The experimental oil production rates were negatively

impacted by the continuous low permeability layers and initial gas content. The small

diffusivity requires that the surface area exposed to solvents be increased in order to

achieve a commercial oil recovery rate. The bottom water offers a large oilwater contact

area between the wells provides the contact for solvent. Frauenfeld et al. [2007] studied

thermal VAPEX with live heavy oil. They run three experiments to evaluate the VAPEX

process in which the oil had significant initial methane saturation. After solvent injection,

steam was injected into the production well to reflux the solvent. Test results indicate that

the live oil inhibits solvent absorption and hence oil production rates. However, a properly

designed solvent system could produce oil at a reasonable rate. It was also observed that

solvent can be recycled more easily by heating the production well with either electrical or

steam heat. A fairly local heated zone around the wellbore was formed through direct

wellbore heating. Asphaltene precipitation was not significant in these experiments. Zhao

et al. [2005] attempted to combine the advantages of SAGD and VAPEX together to

minimize the energy input in heavy oil and bitumen recovery. They conducted

steam-alternating-solvent (SAS) experiments and the corresponding numerical simulation.

Their results showed that the energy input in the SAS process was 47% lower than that of

the SAGD process for recovering the same amount of oil. In addition, the post-run analysis

revealed that asphaltene precipitation occurred in the porous media. Table 2.5 shows the

VAPEX test results achieved by various researchers who are not mentioned above.

Page 36: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

15

Table 2.1 VAPEX experimental studies by Butler’s group.

No. AuthorHeavy Oil Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Solvent

T P qs t qo RFkg/m3/°C mPas/°C cmcmcm % D °C kPa cc/h h g/h %

1

Butler and

Mokrys[1991, 1993]

Tangleflags 979/15.6 10000/20 Hele−Shaw square 770.068 — 100 1356 C3+hot H2O 36 1034 1034 — 41 97 2 Tangleflags 979/15.6 10000/20 Hele−Shaw square 770.068 — 100 1356 C3+hot H2O 44 1379 1379 — 49 86 3 Tangleflags 979/15.6 10000/20 Hele−Shaw square 770.068 — 100 1356 C3+hot H2O 45 1241 1241 — 60 89 4 Tangleflags 979/15.6 10000/20 Hele−Shaw square 770.068 — 100 1356 C3+hot H2O 47 1379 1379 — 62.5 90 5 Tangleflags 979/15.6 10000/20 Hele−Shaw square 770.068 — 100 1356 C3+hot H2O 55 1724 1724 — 63 84 6 Tangleflags 979/15.6 10000/20 sandpack rectangle 70223.5 Hele-Shaw 100 81030 C3 20 621 621 9 61.11 — 7 Tangleflags 979/15.6 10000/20 sandpack rectangle 70223.5 Hele-Shaw 100 81030 C3 22.8 689 689 9 82.20 — 8 Tangleflags 979/15.6 10000/20 sandpack rectangle 70223.5 Hele-Shaw 100 81030 C3 23-30 827 827 9 101.11 — 9 Tangleflags 979/15.6 10000/20 sandpack rectangle 70223.5 Hele-Shaw 100 81030 C3 20 896 896 9 64.44 — 10 Tangleflags 979/15.7 10000/21 sandpack rectangle 70223.5 glass beads 30.9 830 C3+steam 31 1108 1108 8.5 33.7 — 1

Das and

Butler [1995]

Peace River — 138300/20 sandpack rectangle70.620.53.23050 sand 31 43.5 C4+N2 21.5 814 25.5+1.2 13.3+8.0 — 2 Lloydminster — 9350/20 sandpack rectangle70.620.53.22030 sand 31 194.4 C4+N2 22 779 19.4+1.6 21 46.9+36 68 3 Lloydminster — 9350/20 sandpack rectangle70.620.53.22030 sand 35 191.5 C4+N2 21.7 779 12.5+3 22 28.2 57 4 Lloydminster — 9350/20 sandpack rectangle70.620.53.22030 sand 33 191.1 C4+N2 21.2 779 11.4+3.4 33 28+18 5 Cold Lake — 65000/20 sandpack rectangle70.620.53.22030 sand 33 186.8 C4+N2 21.6 779 12+3.7 27 9.5 44 6 Lloydminster — 9350/20 sandpack rectangle70.620.53.22030 sand 32 192.4 C4+N2 21.6 434 24.8+2.3 21 40.5 — 7 Lloydminster — 9350/20 sandpack rectangle70.620.53.22030 sand 37 195.8 C4+N2 21.8 779 9.5+1.5 35 22.4 — 8 Lloydminster — 9350/20 sandpack rectangle70.620.53.22030 sand 33 194 C4+N2 21.5 959 15+1 29.5 50+25 — 1

Butler and

Jiang [1996]

Atlee Buffalo 969 1140/27 sandpack rectangle 13410.23.8 2030 sand3335 217 n-C4 21 204 10 — 19.4 70 2 Atlee Buffalo 969 1140/27 sandpack rectangle 13410.23.8 2030 sand3335 217 n-C4 27 204 10 — 34 80 3 Atlee Buffalo 969 1140/27 sandpack rectangle 13410.23.8 2030 sand3335 217 n-C4 27 204 5 — 31.4 80 4 Atlee Buffalo 969 1140/27 sandpack rectangle 13410.23.8 2030 sand3335 217 n-C4 27 204 10 — 25.4 80 5 Lloydminster 978 7000/20 sandpack rectangle 35.622.933 2030 sand3537 217 n-C4 20-21 124 — 17.5 33.1 — 6 Lloydminster 978 7000/20 sandpack rectangle 35.622.933 2030 sand3537 217 n-C4 20-21 124 — 15 23.8 — 7 Peace river — 930000/20 sandpack rectangle 35.622.933 2030 sand3537 43.5 n-C4 20-21 124 8.71 14 10.73 — 8 Peace river — 930000/20 sandpack rectangle 35.622.933 2030 sand3537 43.5 n-C4 20-21 124 — 27 — —

Page 37: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

16

Table 2.2 VAPEX experimental studies by Maini’s group.

No. Author Heavy Oil Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Solvent

T P qs t qo RF kg/m3/C mPas/°C cmcmcm % D °C kPa cc/h h g/h %

1

Boustani and

Maini [2001]

Dover 1027 543800/20 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 20.5 875 — 20.43 — 51.682 Dover 1027 543800/20 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 20.5 875 — 20.79 — 48.893 Dover 1027 543800/20 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 20.5 875 — 18.66 — 40.934 Panny 970

51767/10 8971/20

bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 10.5 875 — — — — 5 Panny 970 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 10.5 875 — 7.36 — 40.356 Panny 970 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 10.5 875 — 8.78 — 43.797 Panny 970 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 23 875 — 7.09 — 48.728 Panny 970 bulk rectangle7.47.60.254 Hele-Shaw 100 5376 C3 19.5 875 — 7.66 — 47.691 Talbi

and Maini [2003]

— 982.6 3300/24 sandpack annular 30.48

(30.73, 27.2)

12-16 glass beads 35 640 C3+CO2 24 1723 20 9 58.5248.542 — 983.6 3300/25 sandpack annular 12-16 glass beads 35 640 C3+CH4 24 1723 20 9 61.1651.213 — 984.6 3300/26 sandpack annular 12-16 glass beads 35 640 C3+CH4 24 1723 20 9 57.1432.454 — 985.6 3300/27 sandpack annular 12-16 glass beads 35 640 C3+CO2 23 17230.04 9 92.0243.451

Karmakerand

Maini [2003]

— 998 40000/10 9000/20

sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 10 104840/110/7

8 16.11 — 2 — 998 sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 19 1048 8 19.08 — 3 — 998 sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 10 930 50/1

10/78 16.04 —

4 — 998 600/20 sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 10 1048 8 37.76 — 5 — 998

40000/10 9000/20

sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 10 104810/8 8 10.29 — 6 — 998 sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 10 104840/1 8 15.71 — 7 — 998 sandpackrectangle15.367.33.2 16-20 US mesh 38 330 C3 10 104845/8 8 9.82 — 8 — 980 18500/15 sandpackrectangle 7.511.32.5 12-16 glass beads 40 640 n-C4 15 69 — 8 7.91 74.679 — 980 18500/15 sandpackrectangle 7.511.32.5 16-20 sands 39 330 n-C4 15 69 — 8 6.91 65.2310 — 980 18500/15 sandpackrectangle 7.511.32.5 20-30 sands 37 220 n-C4 15 69 — 8 4.14 39.0811 — 980 18500/15 sandpackrectangle 1522.52.5 12-16 glass beads 40 640 n-C4 15 69 — 8 27.1564.0712 — 980 18500/15 sandpackrectangle 1522.52.5 16-20 sands 39 330 n-C4 15 69 — 8 18.7044.1313 — 980 18500/15 sandpackrectangle 1522.52.5 20-30 sands 37 220 n-C4 15 69 — 8 16.4538.82

Page 38: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

17

Table 2.2 VAPEX experimental studies by Maini’s group (Contd’)

No. Author Heavy Oil Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Solvent

T P qs t qo RF kg/m3/°C mPas/°C cmcmcm % D °C kPa cc/h h g/h %

1 Yazdani

and Maini [2004, 2005]

Dina 982.6 18650/9 sandpack rectangle 30452.5 1216 beads 37 640 C4 9 68 0.11 8 16.55 35 2 Dina 982.6 18650/9 sandpack rectangle 30452.5 1620 beads 36 330 C4 9 68 0.10 8 12.26 24 3 Dina 982.6 18650/9 sandpack rectangle 30452.5 2030 geads 35 220 C4 9 68 0.11 8 9.69 18 4 Dina 982.6 18650/9 sandpack cylindrical30(42.3, 36.3) 1216 beads 37 640 C4 9 40 0.10 8 17.17 43 5 Dina 982.6 18650/9 sandpack cylindrical30(42.3, 36.3) 1620 beads 36 330 C4 9 40 0.11 8 12.03 33 6 Dina 982.6 18650/9 sandpack cylindrical30(42.3, 36.3) 2030 beads 35 220 C4 9 40 0.11 8 9.43 23 1 Zedah

and Maini [2008]

Athabasca 1007/20 148000/20 sandpack annular 30.48

(30.73, 27.2)

1214 beads 35.7 640 C3+CO2 21.1 1974 2.05121.55 0.005 37.3 2 Athabasca 1007/20 148000/20 sandpack annular 1214 beads 35.7 640 C3+CO2 20.8 2015 — 122.42 0.005 38.7 3 Athabasca 1007/20 148000/20 sandpack annular 1214 beads 35.7 640 C3+CO2 20.7 3406 2.56 157 0.006 53.6 4 Athabasca 1007/20 148000/20 sandpack annular 1214 beads 35.7 640 C3+CO2 20.8 784 6.36 48.24 0.017 43.5 1

Etminan [2007]

Frog lake 987.5/22.518600/22.5 sandpack rectangle 67.515.23.1 120140 sand35.18 10 n-C4 30 100 — 100 15.55 48 2 Frog lake 987.5/22.518600/22.5 sandpack rectangle 67.515.23.1 120140 sand33.29 10 n-C4 42 100 — 100 13 41 3 Frog lake 987.5/22.518600/22.5 sandpack rectangle 67.515.23.1 120140 sand32.78 10 n-C4 61 100 — 100 10.13 32 4 Frog lake 987.5/22.518600/22.5 sandpack rectangle 67.515.23.1 120140 sand32.79 10 n-C4 60 100 — 100 12.55 38 1

Haghighatand Maini

[2010]

Ekl Point 987.5/22 27500/22 cylindrical rectangle

30.48

(30.73, 27.20)

140200 sand36.42 2.7 C3 20 814 — 380 1.23 — 2 Ekl Point 987.5/23 27500/23 cylindrical rectangle 140200 sand34.24 2.7 C3 20 850 — 380 1.09 — 3 Ekl Point 987.5/24 27500/24 cylindrical rectangle 140200 sand34.71 2.7 C3+toluene 20 850 — 370 — — 4 Ekl Point 987.5/25 27500/25 cylindrical rectangle 140200 sand36.44 2.7 C3+toluene 20 variable — 390 — — 5 Ekl Point 987.5/26 27500/26 cylindrical rectangle 140200 sand 37.7 2.7 C3 20 750 — 380 1.07 — 6 Ekl Point 987.5/27 27500/27 cylindrical rectangle 140200 sand39.93 2.7 C4 20 200 — 400 1.38 —

Page 39: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

18

Table 2.3 VAPEX experimental studies by Gu’s group.

No. Author Heavy Oil Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Solvent

T P qs t qo RFkg/m3/°C mPas/°C cmcmcm % D °C kPa cc/h h g/h %

1

Zhang et al.

[2007]

Lloydminster 988 24137/20 sandpackrectangle 40102 2030 Ottwa 36.8 441 C4 21 240 — 49.5 3.89 75 2 Lloydminster 988 24137/20 sandpackrectangle 40102 3050 Ottwa 37.5 132 C3 21 240 — 98 1.77 67 3 Lloydminster 988 24137/20 sandpackrectangle 40102 2030 Ottwa 36.2 438 C3 22.5 500 — 101 1.26 46 4 Lloydminster 988 24137/20 sandpackrectangle 40102 3050 Ottwa 36.8 158 C3 23.3 500 — 179.5 0.87 61 5 Lloydminster 988 24137/20 sandpackrectangle 40102 3050 Ottwa 37.4 418 C3 23.4 600 — 54 2.84 56 6 Lloydminster 988 24137/20 sandpackrectangle 40102 2030 Ottwa 35.4 417 C3 23.5 800 — 25 4.76 64 7 Lloydminster 988 24137/20 sandpackrectangle 40102 3050 Ottwa 35.4 122 C3 23.2 800 — 59 2.85 63 8 Lloydminster 988 24137/20 sandpackrectangle 40102 2030 Ottwa 35.8 424 C3 23 900 — 58.75 3.54 73 9 Lloydminster 988 24137/20 sandpackrectangle 40102 2030 Ottwa 35.3 410 C3 23 900 — 78.5 2.1 62 10 Lloydminster 988 24137/20 sandpackrectangle 40102 3050 Ottwa 36.7 143 C3 23.6 918 — 21.33 8.7 76 11 Lloydminster 988 24137/20 sandpackrectangle 40102 3050 Ottwa 36.1 118 C3 23.41050 — 54.5 1.94 — 12 Lloydminster 988 120000/20 sandpackrectangle 40102 5070 Ottwa 30.15 7 C3 20 300 — 72 0.11 — 13

Moghadamet al.

[2008]

Lloydminster 978 11900/20 sandpackrectangle 40102 2030 Ottwa 32.5 310 C3 20.8 854 — 13.5 16.96 — 14 Lloydminster 978 11900/20 sandpackrectangle 40102 3050 Ottwa 32.9 310 C3 20.8 854 — 14 4.08 — 15 Lloydminster 978 11900/20 sandpackrectangle 40102 30v50 Ottwa 33.1 103 C3 20.8 854 — 53 — — 16 Lloydminster 978 11900/20 sandpackrectangle 40102 4060 Ottwa 35.4 96 C3 20.8 854 — 103 3.31 — 17 Lloydminster 978 11900/20 sandpackrectangle 40102 6080 Ottwa 35.7 49 C3 20.8 854 — 55.5 1.62 — 18 Lloydminster 978 11900/20 sandpackrectangle 40102 80100 Ottwa 36.3 25 C3 20.8 854 — 35.5 1.12 — 19 Lloydminster 978 11900/20 sandpackrectangle 40102 80100 Ottwa 36.2 16 C3 20.8 854 — 52 — — 20 Lloydminster 978 11900/20 sandpackrectangle 40102 80100 Ottwa 35.2 15 C3 20.8 854 — 63 0.56 — 21 Lloydminster 978 11900/20 sandpackrectangle 40102 80100 Ottwa 35.2 4.5 C3+C1 24 307 20 70 0.96 —

Page 40: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

19

Table 2.4 VAPEX experimental studies by ARC.

No. Author Heavy Oil Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Solvent

T P qs t qo RF kg/m3/°C mPas/°C cmcmcm % D °C kPa cc/h PV g/h %

1 Cuthiell

et al. [2003]

Lloydminster — 5500/25 sandpackrectangle 25302.8 2040 silica — 90 toluene 25 — 40 0.54 18.3 — 2 Lloydminster — 5500/25 sandpackrectangle 25302.8 2040 silica — 90 toluene 26 — 10 0.58 4.46 — 3 Lloydminster — 5500/25 blend rectangle 25302.8 2040/5070 blend — 88 toluene 27 — 10 0.98 3.46 — 4 Lloydminster — 5500/25 field rectangle 25302.8 Rash lake sand — 8 toluene 28 — 10 1.08 1.3 — 5 Lloydminster — 5500/25 sandpackrectangle 25302.8 2040 silica — 90 toluene 29 — 100 1.44 17.7 — 1

Frauenfeld et al.

[2006, 2007]

Lloydminster — 23000/20 sandpackrectangle 902010 — 33.8 250 C1+C2+C3 15180 3400 20 26.00 32.2 1 2 Lloydminster — 23000/20 sandpackrectangle 902010 — 33.8 250 C2 15180 3400 36 13.40 148 31.73 Lloydminster — 39000/20 sandpackrectangle 902010 — 33.8 250 C2 20240 3600 30 15.00 150 36.94 Kerrobert 994.1/20 50000/20 sandpackrectangle 303010 — 35 400 n-C4 20 100 174 13.12 26 61.65 Kerrobert 994.1/21 50000/21 sandpackrectangle 303010 — 35 4 n-C4 20 100 560.3 189.0 1.18 47.46 Kerrobert 994.1/22 50000/22 sandpackrectangle 303010 — 35 400/90 n-C4 20 100 157.6 21.30 9.31 37 7 Kerrobert 994.1/23 50000/23 sandpackrectangle 303010 — 35 400/90 n-C4 20 100 87.53 13.50 25.4 62 8 Kerrobert 994.1/24 50000/24 sandpackrectangle 303010 — 35 400 n-C4 20 100 318.9 13.50 16.65 40

1 Zhao et al.

[2005]

Northern Alberta — 10000 sandpackrectangle 248010 Ottwa 33 115 C3+steam 218 2200

1 +

129.4 12 288 58.7

Page 41: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

20

Table 2.5 VAPEX experimental studies by other research groups.

No. Author Heavy Oil Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Solvent

T P qs t qo RF kg/m3/°C mPas/°C cmcmcm % D °C kPa cc/h h g/h %

1 Lim et al.

[1995]

Cold lake — 100000/25 sandpack rectangle 474227 — 32.8 80 C3 45 830-1000 — 9 912.94 47 2 Cold lake — 100000/26 sandpack rectangle 474227 — 35 20 C2 20 2800-4000 — 100 115.66 62 3 Cold lake — 100000/27 sandpack rectangle 474227 — 35 110 C2 20 3000-3900 — 75 119.39 48 1

Hadil [2009]

Athabasca 1001/22 225000/22 sandpack cylinder 25 (45, 35) 3040 38 204 C3 70 690 — 3 39.00 — 2 Athabasca 1001/22 225000/22 sandpack cylinder 25 (45, 35) 4050 38 102 C3 70 690 — 3 31.00 — 3 Athabasca 1001/22 225000/22 sandpack cylinder 25 (45, 35) 5070 38 51 C3 70 690 — 3 21.00 — 1

Rezaei [2011]

— 960/25 9231/25 sandpack circular 295

glass beads

40 500 C4 70 1030 — 4 — 55.6 2 — 960/25 9231/25 sandpack circular 295 40 500 C4 80 1030 — 4 — 52.6 3 — 960/25 9231/25 sandpack circular 185 30 500 C4 98 1400 — 6 — 94.5 4 — 960/25 9231/25 sandpack circular 265 30 500 C4 98 1500 — 12 — 72.1 5 — 960/25 9231/25 sandpack circular 105 30 500 C4 98 1600 — 8 — 62.3 6 — 960/25 9231/25 sandpack circular 175 30 500 C4 112 1500 — 7 — 45 7 — 960/25 9231/25 sandpack circular 175 30 500 C4 108 1600 — 8 — 64.5 8 — 960/25 9231/25 sandpack circular 295 40 500 C3 90 1500 — 4 — 55.3 9 — 960/25 9231/25 sandpack circular 155 30 500 C3 85 1500 — 4 — 53.7

10 — 960/25 9231/25 sandpack circular 175 30 500 C3 67 1500 — 4 — 47.8 14 — 960/25 9231/25 sandpack circular 275 30 500 C3 53 1500 — 10 — 60.3 15 — 960/25 9231/25 sandpack circular 205 30 500 C3 52 1500 — 10 — 65.5 16 — 960/25 9231/25 sandpack circular 175 30 500 C3 54 1650 — 6 — 43.3 17 — 960/25 9231/25 sandpack circular 185 30 500 C3 53 1450 — 8 — 74.6 18 — 960/25 9231/25 sandpack circular 195 30 2400 C3 52 1450 — 8 — 56.9 19 — 960/25 9231/25 sandpack circular 155 30 2400 C3 54 1650 — 7 — 40.4 20 — 960/25 9231/25 sandpack circular 155 Berea

core

23 350 C3 53 1500 — 48 — 27.5 21 — 960/25 9231/25 sandpack circular 155 21 350 C4 98 1350 — 28 — 44.4 22 — 960/25 9231/25 sandpack circular 305 21 350 C3 53 1600 — 360 — 41.2

Page 42: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

21

Solvent chamber evolution

Zhang et al. [2007] conducted a series of laboratory-scale VAPEX tests to study the

solvent chamber evolution under different operating conditions. The evolution of the

solvent vapour chamber was roughly divided into four stages: the initial solvent chamber

formation period and the solvent chamber rising, spreading, and falling phases. Butane and

propane were used as two respective solvents to recover a Lloydminster heavy oil sample

at a room temperature and different pressures. Their physical model was packed with

Ottawa sands of different mesh sizes of 20−30 and 30−50. Some representative digital

images were taken at the end of the three phases. First, the initial solvent chamber was

formed around the injector after the solvent was dissolved into the heavy oil and some

diluted heavy oil was produced. Then the solvent chamber rising phase started from the

initially formed solvent chamber until it reached the top of the physical model. As can be

seen in Figure 2.2a, in the spreading phase, the solvent chamber spreads laterally and

finally reaches the upper left-hand and right-hand corners of the physical model, as shown

in Figure 2.2b. Afterward, the solvent chamber kept falling down until the oil production

rate became extremely low. Figure 2.2c shows the solvent chamber profile at the end of its

falling phase. In terms of time and oil production, the spreading and falling phases take the

longest periods and contribute more than 80% of the oil production. Therefore, the solvent

chamber spreading and falling phases are the focus of the mathematical modeling in the

next chapter.

Page 43: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

22

(a)

(b)

(c)

Figure 2.2 Solvent vapour chamber profiles at the end of (a) Rising phase; (b)

Spreading phase; and (c) Falling phase [Zhang, et al., 2006].

Page 44: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

23

Operating pressure

The operating pressure is a crucial factor that affects the heavy oil production rate in

the VAPEX process because the solvent solubility is strongly dependent on the operating

pressure. As one of the major factors for inducing asphaltene precipitation, the operating

pressure influences oil production rate significantly. It was reported that the optimum

operating pressure was set close to but lower than the solvent vapour/dew point pressure to

obtain a higher oil production rate [Butler and Mokrys, 1991; Das and Butler, 1998; Butler

and Jiang, 2000; Boustani and Maini, 2001].

Das and Butler [1995] investigated the effect of the operating pressure on the heavy

oil production rate in the VAPEX process. They tested two different pressures of Pinj = 779

and 434 kPa by using butane as an extracting solvent and nitrogen as a carrier gas to

increase the operating pressure. It was found the operating pressure does not have a

significant effect on the oil production rate if a gas mixture is used to extract heavy oil.

Butler and Jiang [2000] tested Pinj = 30, 185, and 300 psig at a temperature of 27C

with butane, propane, and the mixture thereof as the extracting solvents, respectively. It

was observed that for the first four hours, the two experiments gave almost the same oil

production rate, which was due to the initial communication between the injector and the

producer. During the solvent chamber spreading phase, the experiment at a lower operating

pressure gave a higher oil production rate than that at a higher operating pressure. The

average oil production rate over the entire experiment was reduced by approximately 8% at

an increased operating pressure. This is probably because the exacting solvent became less

gaseous at the increased operating pressure.

Page 45: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

24

Well configuration

The well configuration represents the spatial placement of the injector and producer

placement during the VAPEX heavy oil production process. Butler and Jiang [2000]

conducted an experimental study to investigate the effect of a well configuration on the oil

production rate. Two well configurations were attempted: (1) The injector is closely

located right above the producer; (2) The injector is located horizontally apart from and

above the producer. It is found that the cumulative oil production in the latter case is higher

than that in the former case, even though the oil production rate for the latter case at the

very beginning of the VAPEX test was lower. It was concluded that a wider well spacing is

more beneficial to enhancing the contact area between the solvent vapour and oil so as to

increase the oil production rate. Field-scale well spacing of the order of 100−200 m is

feasible for the situation considered.

2.1.2 Mass transfer modeling of VAPEX

Diffusion equation

The most important mechanism of VAPEX is the significant oil viscosity reduction

through sufficient solvent dissolution, which is actually a mass-transfer process between

the solvent and heavy oil. Fick’s 2nd Law [Fick, 1955] is applied to describe the dissolution

of solvent into a crude heavy oil:

c cDt x x

, (1.1)

where, c is the solvent concentration in heavy oil, vol.%; D is the diffusion coefficient, m2/s;

x is the space variable, m; t is the time variable, s.

Page 46: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

25

Diffusion coefficient

Diffusion coefficient is a transport property that is required to calculate the

mass-transfer rate between the solvent and heavy oil due to molecular diffusion. There are

two categories of diffusion coefficients in the literature: constant value and variable value.

A diffusion coefficient can be assumed to be a constant value when the solubility of the

solvent in heavy oil is not high under the test conditions. Based on this assumption,

diffusion coefficients of gaseous solvents such as methane, propane, and carbon dioxide

were measured by using the so-called pressure decay method [Schmidt, 1985; Upreti and

Mehrotra, 2000, 2002; Tharanivasan et al., 2006], the dynamic pendant drop volume

analysis (DPDVA) method [Yang and Gu, 2006], as well as the modified pressure decay

method [Etminan et al., 2010]. Their measured values are shown in Table 2.6.

Hayduk and Cheng [1971] conducted extensive experimental studies on the

diffusion coefficients of ethane in normal hexane, heptanes, octane, dodecane, and

hexadecane at 25C, and of carbon dioxide in hexadecane at 25 and 50C. They found that

the diffusion coefficient of a solvent depended on the mixture viscosity, which could be

commonly expressed as:

D , (1.2)

where, α and β are constants depending on the crude oil and solvent properties as well as

the operating conditions; is the viscosity of oil−solvent mixture, mPas. β is less than

unity.

Page 47: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

26

Table 2.6 Comparison of the measured diffusion coefficients of CO2, CH4, C2H6 and

C3H8 in different heavy oil and bitumen samples.

Solvent Crude oil Pressure (MPa)

Temperature (C)

Viscosity (mPas)

Diffusivity (10−9 m2/s)

CO2 Athabasca1 5.0 20 361,700 at 20C 0.28 Athabasca2 3.1−4.1 25 767 at 80C 0.16−0.22 Athabasca3 4.0 25 821,000 at 25C 0.12−0.20 Llyodminster4 3.5−4.2 23.9 20,267 at 23.9C 0.46−0.55 Llyodminster5 2.0−6.0 23.9 23,000 at 23.9C 0.22−0.55 Athabasca6 3.2 75 100,000 at 23.9C 0.5 CH4 Athabasca3 4.0 25 224,500 at 25C 0.08−0.11 Athabasca3 4.0−8.0 25 821,000 at 25C 0.06−0.08 Llyodminster4 4.9−5.0 23.9 20,267 at 23.9C 0.21−0.22 Llyodminster5 6.0−14.0 23.9 23,000 at 23.9C 0.12−0.19 Dodecane6 3.5 45, 65 1.34 at 15.6C 4.22−5.28 C2H6 Athabasca3 4.0 25 821,000 at 25C 0.21−0.38 Llyodminster1 1.5−3.5 23.9 23,000 at 23.9C 0.13−0.77 C3H8 Llyodminster5 0.4−0.8 23.9 20,267 at 23.9C 0.49−0.79 Llyodminster5 0.4−0.9 23.9 23,000 at 23.9C 0.09−0.68

Note: 1— Schmidt [1989] 2— Upreti and Mehrotra [2000] 3— Upreti and Mehrotra [2002] 4— Tharanivasan [2004] 5— Yang and Gu [2006] 6— Etminan et al. [2010]

Page 48: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

27

The diffusion coefficients of propane in hexane, heptanes, octane, hexadecane,

n-butanol, and chorobenzene at P = 100 kPa and T = 25C and in n-butanol and

chlorobenzene at 0 and 50C were measured by using the steady-state capillary cell method

[Hayduk et al., 1973], and their results showed the following correlation:

10 0.5450.591 10D (for propane). (1.3)

Das and Butler [1996] applied the general correlation proposed by Hayduk and

Cheng [1971] to determine the values of α and β in Eq. (1.2). Based on their ten VAPEX

experimental tests conducted in the ranges of P = 820−1,160 kPa and T = 21−35C, the

correlations for propane and butane in Peace River bitumen with a viscosity of 126,500

mPas were back-calculated as:

10 0.4613.06 10D (for propane), (1.4)

10 0.464.13 10D (for butane). (1.5)

It is worthwhile to note that in the actual VAPEX process, the solvent-diluted heavy

oil drains down along the transition zone and thus the upward-moving solvent keeps

contacting the fresh heavy oil. In this case, both the molecular diffusion and convective

dispersion of the solvent in the heavy oil contribute to the mass transfer between a heavy

oil and solvent. The back-calculated effective diffusion coefficient by Das and Butler

contained both of the effects [Boustani and Maini, 2001].

Heavy oil viscosity

After the solvent dissolves into the heavy oil, the high viscosity of the crude heavy oil

decreases dramatically. In the literature, the correlation between the viscosity of the

solvent-diluted heavy oil and the solvent concentration was commonly modeled by using

the Lederer equation [Lederer, 1933]:

Page 49: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

28

ss o

of f , (1.6)

s 1cfc c

, 1o sf f . (1.7)

where, o, s, and are viscosities of the crude oil, liquid solvent, and mixture of the two,

respectively, mPas; fo and fs are the weighted volume fractions, vol.%; λ is a weight factor.

Shu [1984] formulated the following correlation to determine the above-mentioned

weight factor for a heavy oilsolvent mixture:

0.5237 3.2745 1.631617.04ln

o s o s

o s

, (1.8)

where o and s are specific gravities of the crude heavy oil and liquid solvent, respectively.

Heavy oil density

The density of solvent-diluted heavy oil can be determined by using the mixture rule

for an ideal solution [McCain, 1990]:

1

1 o sc c

. (1.9)

where, o and s are the densities of the heavy oil and liquid solvent, respectively, kg/m3;

The above equation is applicable only if the volume change due to the solvent dissolution

into the heavy oil is negligible. In addition, the solvent is assumed to be a liquid once it

dissolves into the heavy oil.

2.1.3 Theoretical modeling of VAPEX

Butler−Mokrys model

Butler and Mokrys [1989] carried out the first VAPEX experiments in a Hele−Shaw

cell as a solvent-analog of the SAGD process. They used liquid toluene as the solvent to

Page 50: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

29

recover two bitumen samples. In addition, they developed an analytical mathematical

model on the basis of the following assumptions:

1. Mass transfer of solvent into the bitumen bulk is under pseudo-steady state

condition: 0ct

;

2. Solute−solvent interface moves at a constant unspecified velocity: constU ;

3. Oil flows along the interface in a thin diffusion boundary layer;

4. Drainage of the undiluted bitumen was considered negligible;

5. Effect of surface tension is ignored because it is not crucial;

6. Change in velocity gradient in the direction normal to the flow surface is

negligible: 2

2 constv

;

7. Viscosity, density, and diffusivity, are all concentration dependent and assumed

to be uniform along the boundary and across the cell thickness but changing

across the transition zone: c , c , and D D c .

Two correlations are applied:

1. Concentration is a function of the distance from an arbitrary point to the

boundary between solvent chamber and transition zone: c c ;

2. Geometric relation between the normal velocity U and horizontal velocity xt

:

sin Uxt

.

Page 51: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

30

Three governing equations constitute the foundation of the theoretical model:

1. Fick’s 1st law: dcD Ucd

;

2. Darcy’s Law: sinkv g

;

3. Mass balance equation: o0

t H

y

qdt S xdy .

where, is the distance from the boundary between the solvent chamber and transition

zone to an arbitrary point in the transition zone, m; θ is the inclination angle of the

transition zone, degree; v is heavy oil drainage velocity, m/s; k is the permeability, D; g is

the gravitational acceleration, m/s2; is the porosity, fraction; ∆So = Soi−Sor is the oil

saturation change in the solvent chamber; Soi is the initial oil saturation and Sor is the

residual oil saturation; x and y are the distances in the horizontal and vertical directions,

respectively.

On the basis of the above assumptions, correlations, and governing equations, Butler

and Mokrys derived the famous analytical model for predicting the heavy oil drainage

volume flow rate, qo, during the solvent spreading phase:

o o s2 2q L kg S N H , (1.10)

where, L is the length of a horizontal production well; H is the height of the sand-packed

physical model; and sN is the dimensionless number:

max

min

s

1c

c

c DN dc

c

. (1.11)

Page 52: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

31

In the integrand, ∆ρ is the density difference of the solvent-diluted heavy oil and the liquid

solvent, kg/m3.

Das−Butler model

Das [1995] investigated the VAPEX performance in a sandpack. It was observed that

the oil extraction rate in porous media was 3 to 5 times higher than that predicted by the

Butler−Mokrys analytical model. He attributed the higher production rate in porous media

to the increased bitumen–solvent contact area, increased solvent solubility, and surface

renewal. Since the previous theoretical model was developed on the basis of bulk flow and

could no longer properly predict the oil production rate, Das modified it to make the

prediction better match the measurements:

1o s2 2 ( )q L kg S N h y . (1.12)

where, is the cementation factor, dimensionless. Cementation factor measures the

consolidation of the matrix due to asphaltene precipitation onto the surface of the reservoir

rock. In the meantime, Das replaced the earlier intrinsic molecular diffusion coefficient D

with an apparent diffusion coefficient Dapp:

DDapp . (1.13)

Earlier study [Perkins and Johnston, 1963] indicated this relationship is rather as follows:

appDDF

. (1.14)

F is the formation electrical resistivity factor. Archie [1942] suggested F is related to the

porosity, , and a constant ‘Λ’ by the following equation:

F

(1.15)

Page 53: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

32

The experimentally measurement of is between 1.3 and 2.2, and it changes with the

rock lithology. The more consolidated the reservoir rock, the smaller would be.

Moghadam et al. model

Moghadam et al. [2008] conducted a number of VAPEX experiments with a visual

rectangular sand-packed high-pressure physical model, to examine the effects of the

solvent chamber evolution, the transition-zone thickness, and the inclination angle on the

VAPEX process. Propane was used as the extracting solvent to recover a Lloydminster

heavy oil sample at a pressure slightly lower than the saturation pressure of the solvent.

They found that the ‘inclination angle’ of the spreading phase and the falling phase was

closely related to the oil production rate during the two phases, respectively. A theoretical

VAPEX model was developed on the basis of two assumptions:

1. Two boundaries of the transition zone between the solvent chamber and the

untouched heavy oil zone are assumed to be straight lines with a constant

transition-zone thickness;

2. Downward oil drainage velocity in the transition zone is assumed to be a linear

function of the transverse distance between the solvent chamber and the

untouched heavy oil zone.

The principal governing equations in their model are:

1. Darcy’s Law: sinkv c gc

.

2. Geometric relations: sin HW

; x

sin ; cot x

H .

where, W is the width of the sand-packed physical model, m; δ is the thickness of the

transition zone, m. Cumulative oil production rates during the solvent chamber spreading

Page 54: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

33

and falling phases were derived as:

2o o cotQ H S (for the spreading phase), (1.16)

o o s2 cot tanQ HW S (for the falling phase). (1.17)

where, Qo is the cumulative oil production rates, m2; s is the inclination angle of the

transition zone at the end of the solvent chamber spreading phase.

This is the first attempt to analytically describe the evolution of the solvent vapour

chamber during the VAPEX process. However, its application is quite limited since this

model did not include the solvent diffusion coefficient of the solvent.

Yazdani−Maini model

Yazdani and Maini [2007] examined the effects of the drainage height and grain size

on the stabilized oil drainage rates. On the basis of several sandpack tests in two

rectangular and cylindrical models with three different heights and three sand sizes, they

generated two empirical scale-up correlations for the VAPEX process:

1.26o 0.0174q H k , (1.18)

1.13o 0.0288q H k . (1.19)

They found that the stabilized heavy oil production rate was a function of the

drainage height to the power of 1.1–1.3 rather than 0.5 as predicted in the previous models.

The constant coefficient in the above equation, 0.0174 in Eq. (1.18) and 0.0288 in Eq. (1.19)

represented the combined effect of the gravity drainage, mass transfer, residual oil

saturation, and original oil viscosity. Therefore, these empirical coefficients need to be

determined for specific solvent−oil−reservoir systems. In addition, it is still uncertain for

whether these empirical models can be applied in other cases.

Page 55: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

34

2.1.4 Numerical modeling of VAPEX

Numerical simulators, such as Steam, Thermal and Advanced Reservoir Simulator

(STARS) of Computer Modelling Group (CMG), have been attempted to model the

VAPEX process by different authors [Yazdani, 2007; Qi and Polikar, 2005]. Cuthiell et al.

[2003] used a semi-compositional model (STARS) to model the VAPEX process and

concluded that the simulation could match the solvent breakthrough time, oil production

rate, and the general character of the viscous fingering phenomenon. Das used a fully

compositional model, Generalized Equation of State Model Reservoir Simulator (GEM),

to simulate laboratory VAPEX tests. They applied a high diffusion coefficient and a thick

transition zone in their simulation model to match the experimental data. Wu et al. [2005]

simulated the asphaltene precipitation during the VAPEX process with STARS and

investigated the effect of operation parameters on the VAPEX performance. Rehnema et al.

[2007] conducted a screening study for practical application of VAPEX by using the GEM

module. Zeng et al. [2008] evaluated the VAPEX performance with a Tee well pattern by

using STARS. They concluded that the Tee well pattern shortened the breakthrough time

and increased oil production rate by 28 times. Cuthiell [2012] simulated a laboratory

VAPEX experiment by using a semi-compositional simulator, Tetrad, which was able to

incorporate the diffusion/dispersion physics with the VAPEX process. It was concluded

that most of the gravity drainage occurred in the capillary transition zone.

In general, VAPEX simulations fall into two categories: one is to study the effects of

the reservoir/fluid properties and operating parameters and the other one is to simulate and

validate the laboratory tests. The limitations with the numerical simulation models are: (1)

Simulation models are unable to apply very fine grids to accurately capture the transition

Page 56: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

35

zone which is estimated to be just 2.8–11.2 mm wide [Das and Butler, 1995; Moghadam et

al., 2008; Yazdani and Maini, 2009]; (2) Numerical simulation wastes a great deal of

computational time on the solvent chamber and the untouched heavy oil zone, both of

which occupy a larger area but contributes less oil production in comparison with the

transition zone; (3) Numerical simulation results are not sensitive to the diffusion

coefficient [Qi and Polikar, 2005], which is probably due to the fact that the numerical

dispersion may affect the simulation result to a larger extent than the physical

diffusion/dispersion.

2.2 Cyclic Solvent Injection (CSI)

Solvent-based process with cyclic pressure increase and decrease was investigated

long before the VAPEX process. Shelton and Morris [1973] applied a rich gas to produce

oil in a huff-n-puff mode, where a single well was used alternately as the solvent injector

and oil producer. Allen [1974] patented a huff-n-puff type process in which propane or

butane was injected in cycles to extract oil from a cell packed with Athabasca tar sands. A

typical cyclic solvent injection (CSI) cycle consists of three periods: solvent injection,

soaking, and oil production periods. Unlike VAPEX which is a constant-pressure process,

the pressure of CSI is cyclically built up during solvent injection period and drawn down

during the oil production period.

Lim et al. [1995, 1996] conducted some CSI tests to enhance the oil production of the

VAPEX process. Ethane was applied to produce Cold Lake bitumen at the supercritical

and sub-critical conditions. They found that supercritical ethane performs better than the

sub-critical ethane in terms of either bitumen production rate or the eventual recovery

factor. It was found that the molecular diffusion is not the major mechanism of a

Page 57: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

36

higher-than-expected production rate. Solvent dispersion or viscous fingering might play a

larger role. It was also found that the full utilization of the horizontal well was not achieved

in the model through the residual oil saturation measurement. They observed that the oil

production during the production period comes from two distinct mechanisms: wellbore

inflow and gravity drainage. The production profile exhibited a declining rate in the early

cycles, which suggested the near wellbore inflow mechanism is more significant before the

solvent chamber is fully developed. Afterward, the oil production rate starts to increase,

which is attributed to a growing solvent chamber size as well as the gravity drainage.

Ivory et al. [2010] investigated the CSI process with a real-scale 3 m long stepped

cone model run at field-scale times. A mixture solvent (28 vol.% propane + 72 vol.%

methane) was cyclically injected into the physical model at a pressure of around 3 MPa.

After nearly 6 cycles and 2 years of test, they achieved an oil recovery factor of 6.8% for

the primary production and 50.4% for the entire test, which showed that CSI has potential

to be a good follow-up process of cold production processes.

Dong et al. [2006] designed a methane pressure-cycling (MPC) process to recover

the residual oil after the termination of either primary or waterflood production in some

heavy oil reservoirs. The essence of this method is to restore the solution-gas drive

mechanism for a ‘primary’ production. They found that the mobile-water saturation greatly

affects the performance of the MPC process.

Jamaloei et al. [2012] studied an enhanced cyclic solvent process (ECSP) by using

two solvent gases: one was more soluble (propane) and the other was more volatile

(methane). They found that by using the two-slug injection strategy, the oil recovery factor

could be as high as 34.4% compared with an oil recovery factor of 4.27% by using the

Page 58: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

37

one-slug pure methane. It was concluded that methane CSI can be greatly enhanced by

introducing a propane slug during the injection period. In the ECSP process, methane

provides expansion and some propane stays in the oil to keep the oil viscosity low during

the pressure reduction process. This indicates that the major mechanisms of ECSP are

viscosity reduction and solvent-gas-drive during the early stage of the production.

Jiang et al. [2013] proposed another process, cyclic production with continuous

injection (CPCSI), to enhance heavy oil recovery. In this process, vapour solvent is

continuously injected into the model to maintain the pressure and also supply an extra gas

drive force to flood the solvent-diluted heavy oil out. They found that the oil recovery

factor could be increased to 85% with the CPCSI method.

Studies on the CSI heavy oil recovery process, especial those for post-CHOPS, are

quite limited in the literature. Table 2.7 lists the up-to-date efforts on physical modeling

and numerical simulations of the CSI processes.

Page 59: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

38

Table 2.7 CSI experimental studies in the literature.

No. Author Heavy Oil/bitumen Model Injection Production

Name ρo μo Type Shape

Size Sand

ϕ k Config. Solvent

T P t RF kg/m3/°C mPas/°C cmcmcm % D °C MPa h %

1 Lim

et al., [1996]

Cold lake — 100,000/25 sandpack rectangle 474227 Quartz

0.330.350.35

80 20 110

line source

C3 C2

25 1 4.2

1−3 h 3/13/15 c 4050

2 Dong et al., [2006]

Senlac Cactus Lake

North Plover Lake

— 1,700−5,400sandpack — 30.55 Quartz

35.532.8

6.589.70 lateral C1 —

4.5 drop to

0.5

5−17 h 8 c

23.3 28.8

3 Ivory et al., [2010]

Rush Lake — 39,320/20 sandpack circular cone

300 (H) (9.7, 1)

(Dtop, Dbtm)

Quartz 38 4.5 point 28% C3+

72% CO2 —

2.23.4drop to 0.51

inj: 63−80 d prd: 17−26 d

6 c

52

4 Firouz et al., [2012]

Saskatchewan heavy oil — 1,420/22 sandpack cylinder

30.84

5.08

Quartz 24 1.8 lateral

C1 C3 C4

CO2

20.51.76.8drop to 0.276

soak: 24 h 7−10 c

58−73

5 Jamaloei

et al., [2012]

South Britnell — 1,080/22 sandpack cylinder

101.3

(4.9, 3.2)

Quartz 38 41.8 point C1

C3 — 0.823

0.1

soak: 22 h prod: 0.5 h

6 cycles 34.3

6 Huerta et al., [2012]

Lloydminster — 35,000/20 sandpack — — Quartz

— — lateral 90% CO2+10% H2S —

3.0 drop to

0.5 Soak: 24 h —

7 Jiang et al., [2013]

Lloydminster — 5,875/20 sandpack rectangular 40102 Glass beads 36 4.7 lateral propane 20.2 0.8 soak: 55 min

prod: 5 min 60

Page 60: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

39

2.3 Chapter Summary

From the literature review in this chapter, it can be seen that extensive laboratory

experiments have been conducted to evaluate the VAPEX process, and both analytical

methods and numerical simulation have been attempted to predict the VAPEX

performance. However, the analytical models are only able to roughly estimate the oil

production rate but unable to describe the solvent chamber evolution. Simulation models

can match the production rate but cannot reasonably describe the oil properties inside the

transition zone. On the other hand, the low oil production rate still exists and affects the

applicability of VAPEX. The CSI process is considered as a promising process for

post-CHOPS. Nevertheless, both theoretical and experimental studies of CSI are rather

limited in the literature. A variety of factors, such as the well configuration and operating

scheme, need to be examined to optimize the productivity of the CSI process.

Page 61: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

40

CHAPTER 3 MATHEMATICAL MODELING OF VAPEX

In this chapter, a new mathematical model is developed to describe the solvent

chamber evolution during the VAPEX heavy oil recovery process. This new model is

based on two physical processes: mass transfer and gravity drainage. The mass transfer

process is modelled as a transient process with a variable diffusion coefficient. The heavy

oil−solvent transition zone in which most of the mass transfer occurs is modelled as a

piecewise linear zone that is updated step by step temporally. The boundary of the

transition zone is considered moving with time and calculated on the basis of the material

balance equation. This VAPEX model is able not only to describe the distributions of

solvent concentration, oil drainage velocity, and diffusion coefficient across the transition

zone, but also to predict the solvent chamber evolution and the heavy oil production rate.

3.1 Mathematical Model and Solution

3.1.1 Heavy oil–solvent transition zone

As shown in Figure 3.1, there are three zones during a typical VAPEX process: a

solvent vapour chamber, an untouched heavy oil zone, and a transition zone in between.

Properties of the heavy oil and solvent in the solvent vapour chamber and untouched heavy

oil zone are unchanged: the solvent chamber is filled with the residual oil and solvent

vapour and the untouched heavy oil zone is full of the original untouched heavy oil. The

solvent chamber has three phases during a VAPEX process: rising, spreading, and falling

phases. Due to the complexity and its relatively minor contribution to the oil production,

the solvent chamber rising phase is disregarded and the spreading and falling

Page 62: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

41

Figure 3.1 Transition zone in the VAPEX process.

Solvent chamber

Transition zone

Injector

Heavy oil

Producer

cmin

cmax

Concentration

Distance

Diluted oil

Solvent

Heavy oil

Diluted oil

Page 63: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

42

phases of the solvent chamber are examined in this study. The solvent-chamber evolution

is caused by the oil drainage from the transition zone. Therefore, modeling of the transition

zone is the key to model the entire VAPEX process.

The heavy oil−solvent transition zone is defined on the basis of the solvent

concentration in the solvent-diluted heavy oil. As shown on the left-hand side of Figure 3.1,

the first boundary of the transition zone (Boundary 1) is at the edge of the solvent chamber

and located at c = cmax, and the second one (Boundary 2) neighbors the untouched heavy oil

zone at c = cmin ≈ 0.01 [Butler and Mokrys, 1989]. cmax is the saturation concentration under

the operating pressure and temperature and cmin is the minimum concentration at which the

solvent-diluted heavy oil starts to flow. The position of Boundary 2 depends on the solvent

concentration profile as well as the position of Boundary 1. This study simplifies the

curved boundary (Boundary 1) as a piecewise linear profile. Figure 3.2 schematically

shows the simplified transition zone at the early and middle stages of the VAPEX process.

It is worthwhile to note that Boundary 1 is defined at the very beginning of VAPEX.

Afterward, its position is calculated automatically by the VAPEX model, which is going to

be formulated in this chapter.

The VAPEX model in Figure 3.2 is a simplified model, in which the solvent is

injected from a line source, and the heavy oil is produced from the left bottom corner of the

model. Followings are the assumptions of the new mathematical model:

1. The porosity and permeability are spatially uniform;

2. The drainage of the solvent-diluted heavy oil in the transition zone, which is

assumed to be in liquid phase and caused only by gravity;

3. The heavy oil–solvent mass transfer along the transition zone is negligible;

Page 64: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

43

(a) (b)

Figure 3.2 Approximation to the transition-zone at the (a) beginning and (b) middle

stages of the VAPEX process.

Solvent chamber

Producer

Transition zone

Heavy oil

Solvent Solvent chamber

Transition zone

Heavy oil

Page 65: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

44

4. The diffusion coefficient takes account of both molecular diffusion and

convective dispersion [Das and Bulter, 1995].

Two spatial coordinate systems are used: a two-dimensional (2D) coordinate system

(x, y) with x in the horizontal direction and y in the vertical direction for the solvent

chamber evolution, and a one-dimensional (1D) coordinate system (ξ) for the mass transfer

in the transition-zone segments. The ξ coordinate starts from and is normal to Boundary 1

of each transition-zone segment. The 2D coordinate system is linked to the 1D coordinate

system through the inclination angle of each transition-zone segment.

A new mathematical model is developed on the basis of the major mechanisms of

VAPEX, such as solvent dissolution and gravity drainage. The new model consists of three

sub-models: a mass transfer model, a fluid flow model, and a boundary movement model,

which will be described one by another in the following sections.

3.1.2 Mass transfer in transition zone

As mentioned above, the mass transfer between solvent and heavy oil is the most

important mechanism of VAPEX. Previous studies assumed it as a steady-state diffusion

process, and used Fick’s 1st law to describe it [Butler and Mokrys, 1989]. However, this

assumption is invalid for the actual cases: solvent chamber grows fast at the top and

spreads slowly at the bottom, which indicates that the heavy oil–solvent mass transfer in

the transition zone changes with both time and space. Therefore, this study considers the

heavy oil–solvent mass transfer as a more realistic transient diffusion process, and Fick’s

2nd law is applied to calculate the solvent concentration distribution in each transition-zone

segment:

Page 66: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

45

c cD

, (3.1)

where, τ is the time, s; ξ is the position normal to the Boundary 1 of a transition-zone

segment, m; D is the diffusion coefficient, m2/s.

For a heavy oil−hydrocarbon solvent system, the diffusion coefficient of solvent into

heavy oil is a function of the viscosity of a heavy oil–solvent mixture, which follows a

general formula [Heyduk and Cheng, 1971]:

D , (3.2)

where, is the viscosity of the solvent-diluted heavy oil, mPas; α and β are both constants

depending on the properties of the heavy oil and solvent as well as the operating conditions.

In this study, the Das and Butler’s correlations [1996] are adopted to calculate the diffusion

coefficient:

10 0.4613.06 10D (for propane), (3.3)

10 0.464.13 10D (for butane). (3.4)

In this study, is computed by using the one-parameter Lederer equation [1933]:

ss o

of f , (3.5)

s 1cfc c

, 1o sf f , (3.6)

where, o and s are viscosities of the crude heavy oil and the liquid solvent, respectively,

mPas; fs and fo are the weighted volume fractions, vol.%; the weight factor is obtained by

using the Shu’s correlation [1984]:

0.5237 3.2745 1.6316o s o s

s

17.04ln o

, (3.7)

Page 67: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

46

where, γo and γs are specific gravities of the crude heavy oil and the liquid solvent,

respectively. The density of the heavy oil–solvent mixture is calculated by using:

1

1 o sc c

. (3.8)

At the left boundary (Boundary 1) of the transition-zone segment, the solvent

concentration is assumed to be the saturation concentration under the operating pressure

and temperature:

max0,c c . (3.9)

The right boundary of the transition zone is treated as a closed boundary:

, 0c L

, (3.10)

Initially, there is no solvent in the oil:

, 0 0c . (3.11)

It is worthwhile to note that the transition zone is updated step by step temporally. In

the first time step, the left boundary condition (BC) and the initial condition (IC) are Eqs.

(3.9) and (3.11), respectively. During the following steps, however, the left boundary is

always moves toward the untouched heavy oil zone and the left boundary condition

becomes a free boundary problem:

max,c s c , (3.12)

where, s is the location of Boundary 1 during a time step, m. Suppose that Boundary 1 of

one transition-zone segment is at ξ0 in the beginning of a time step and it moves at a

velocity of U during that step. Then the location of Boundary 1 during the step becomes:

0s U . (3.13)

Page 68: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

47

Boundary moving velocity U can be determined by using the mass balance equation, which

will be specified in the later section. The IC for the second and the following time steps is

actually the solvent concentration distribution at the end of their previous time steps:

1 , 0t tc c . (3.14)

3.1.3 Fluid flow in transition zone

Given a known concentration profile, the viscosity and density of the solvent-diluted

heavy oil can be calculated by using Eqs. (3.5) and (3.8), respectively. Hence, the gravity

drainage velocity across the transition zone can be determined by using the Darcy’s law:

sinskv c gc

, (3.15)

where, v is the drainage velocity, m/s; k is the permeability, D; ρ(c) and ρs are densities of

the solvent-diluted heavy oil and the liquid solvent, kg/m3; g is the gravitational

acceleration, m/s2; θ is the inclination angle, degree.

The drainage velocity gradually decreases from the maximum value at Boundary 1 to

the minimum value at Boundary 2 of the transition zone. The total amount of the

solvent-diluted heavy oil that drains from one segment into another can be calculated by

integrating Eq. (3.15) across the transition zone:

max

min

dq v

, (3.16)

where, q is the oil drainage flux of one segment, m2/s; max and min are the locations of

Boundary 1 and Boundary 2 of the transition zone, respectively, m.

Page 69: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

48

3.1.4 Moving boundary of transition zone

As shown in Figure 3.3, the solvent-diluted heavy oil flows into and drains out of a

transition-zone segment at fluxes of qin and qout, respectively. Over a short period of time,

τ, Boundary 1 moves by ξ due to the depletion of the oil. Assume the length change of

the outer segment boundary during τ is trivial and the oil saturation change is uniform in

the depleted area, Then ξ can be determined by using the mass balance equation:

out in oi orq q l S S , (3.17)

where, l is the length of the transition-zone segment boundary, m; Soi and Sor are the initial

and residual oil saturations, respectively, vol.%. qin and qout can be obtained from Eq. (3.16).

If the timestep size is small enough, the boundary moving velocity can be approximated as:

out ind q qUd l S

. (3.18)

Eq. (3.18) is substituted into Eq. (3.13) for the concentration calculation.

With a known timestep size ∆t and the boundary moving velocity U, the moving

distance over the time step can be obtained:

mv U t . (3.19)

The movement of the transition-zone segment in the horizontal direction is:

sinmvx

. (3.20)

Similarly, the movement of the transition-zone segment in the vertical direction is:

cosmvy

. (3.21)

Eq. (3.20) is used to estimate the solvent chamber evolution during its spreading phase and

Eq. (3.21) is used to estimate the solvent chamber evolution during its falling phase.

Page 70: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

49

Figure 3.3 Boundary movement of a transition-zone segment.

vin Heavy oil

h

Solvent chamber Transition zone

ξ at t ξ+dξ at t+dt

vout

ξ

Page 71: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

50

3.1.5 Solution procedures

Due to the complexity of the correlation between D and c, it is difficult to analytically

solve the nonlinear governing equation. Therefore, an approximate solution method is

applied: (1) the solvent concentration is calculated step by step; (2) for the first step, D is

considered as a constant since the model is solvent free in the beginning; (3) for the

following steps, D is a function of c at the end of the previous step; thus, D is a known

variable during one step. In this way, the governing equation can be reduced to a linear

partial differential equation (PDE). Specifically, the calculation procedures are described

as below (Figure 3.4):

1. Applying BCs and IC of Eqs. (3.9−3.11) to the governing equation of Eq. (3.1)

for the first time step to obtain a solvent concentration profile for each

transition-zone segment;

2. Computing ρ(c), μ(c), D(c), v(c), q, and U of the solvent-diluted heavy oil by

using Eqs. (3.12, 3.9−3.11, 3.7 or 3.8, 3.19, 3.20, and 3.22), respectively;

3. Calculating the left-boundary movement of each transition-zone segment in the

horizontal or vertical direction by using Eq. (3.24) or (3.25), and then updating

the solvent-chamber profile at the end of the present time step;

4. Entering the next step and updating the solvent diffusion coefficient, left BC, and

IC with those obtained at the end of the previous time step;

5. Calculating a new solvent concentration profile for each transition-zone segment;

and

6. Repeating Steps #2−5 till the end of the process.

Page 72: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

51

Figure 3.4 Flowchart of the solution calculation for the VAPEX mathematical model.

ρ

v

D

U

Governing Eq.

Finish

IC BCs

c

Yes No Terminate

Page 73: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

52

It is worthwhile to emphasize that the mass-transfer model is solved with the finite

difference method (FDM). The key to solving the diffusion equation with a moving

boundary is to discretize of the time and space domains. Figure 3.5 shows the discretization

of ξ and τ axes for a transition-zone segment during a time step. Suppose that there are Nξ

grids with a grid size of ∆ξ across the transition zone, the time needed for the boundary to

pass by ∆ξ at a moving velocity of U is:

U

. (3.22)

Then the number of sub-step during a step ∆t can be estimated as:

tN

. (3.23)

If Nτ < 1, it can be treated as a fixed boundary problem. Otherwise, it will be solved as a

moving boundary problem. The detailed solution to the mathematical model with the

Crank−Nicolson FDM is presented in Appendix A.

3.1.6 Heavy oil production rate

The cumulative heavy oil production can be found by integrating the area of the

solvent vapour chamber in the (x, y) coordinate system:

0

W

Q S H y x dx , (3.24)

where, Q is the cumulative heavy oil production, m3; H is the model height, m; W is the

model width, m. The heavy oil production rate can be obtained by taking the derivative of

Q with respect to t.

Page 74: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

53

Figure 3.5 Discretization of the space and time domains for the numerical solution to

the mass-transfer model with a moving boundary condition.

s = ξ/U

t2=∆ξ/U

t1=0

t3=2∆ξ/U

t4=3∆ξ/U

t5=4∆ξ/U

ξN = L ξ3=2∆ξ ξ2=∆ξ ξ1=0 ξ4=3∆ξ ξ5=4∆ξ

Page 75: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

54

3.2 Results and Discussion

Table 3.1 lists the parameters of a base case. The other cases are discussed below and

their parameters that are different from those in Table 3.1 will be specified.

3.2.1 Solvent chamber evolution and recovery factor

Figure 3.6 shows the solvent chamber evolution with time for the base case. Solvent

chamber grows faster at the top than at the other parts. That is because nothing flows into

the first segment but the solvent-diluted heavy oil keeps draining out, which is different

from the other segments. After around 14 h, the solvent chamber reaches to the right-hand

side of the model, indicating the solvent chamber is near the end of its spreading phase and

starts the falling phase at that moment. Figure 3.7 presents the heavy oil recovery factor

(RF) curve. The RF curve is quite flat in the first a few hours and then increases linearly

during the solvent chamber rising phase, indicating a stabilized oil production rate. This is

consistent with observations in the previous studies. In the last 10 h, the RF curve decreases

slightly as the solvent chamber keeps falling. This is caused by the reduced inclination

angle and the diminishing gravity drainage.

The solvent concentration is quite small due to its fast movement during the first 15 h.

It becomes slightly higher during the solvent chamber falling phase due to the decreased

drainage and accumulation of solvent-diluted heavy oil. For the middle segment, its

thickness is quite stable throughout the VAPEX process, indicating a balance between the

mass transfer and fluid flow. For the bottom segment, because of the smaller gravity effect

and solvent accumulation, the transition zone grows steadily from several millimeters in

the beginning to several centimeters at the end of the VAPEX process. Hence, it can be

seen that transition zone is a dynamic zone and it always changes with time and location.

Page 76: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

55

Table 3.1 Physical properties and operating conditions of the base case for the

VAPEX mathematical model.

Parameters Value Model dimensions, m2 0.1 × 0.1 Porosity, % 35 Permeability, D 50 Relative oil permeability, fraction 1 Solvent type C3H8 Solvent solubility, g solvent/100 oil 26.5 Heavy crude oil viscosity @ 23°C and 800 kPa, mPas 12,000 Heavy crude oil density @ 23°C and 800 kPa, kg/m3 975 Solvent viscosity (liquid) @ 23°C and 800 kPa, mPas 0.106 Solvent density (liquid) @ 23°C and 800 kPa, kg/m3 517 Diffusion coefficient, m2/s 1.306×10-9-0.46 Operating pressure, kPa 800 Operating temperature, C 23 Connate water saturation, % 5 Residual oil saturation, % 15 Number of transition-zone segments 12

Page 77: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

56

Horizontal distance, m

0.00 0.02 0.04 0.06 0.08 0.10

Ver

tical

dis

tanc

e (m

)

0.00

0.02

0.04

0.06

0.08

0.10

Figure 3.6 Evolution of the solvent vapour chamber during a VAPEX process.

2 h

30 h

Page 78: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

57

Time (h)

0 5 10 15 20 25 30

Oil

reco

very

fact

or (%

)

0

20

40

60

Figure 3.7 Oil recover factor of the VAPEX base case.

Spreading phase Falling phase

Page 79: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

58

x (m)

0.00 0.02 0.04 0.06 0.08 0.10

y (m

)

0.00

0.02

0.04

0.06

0.08

0.10

t=0.1 ht=5.0 ht=10.0 ht=15.0 ht=20.0 h

(m)

0.00 0.01 0.02 0.03 0.04 0.05

c/c m

ax

0.0

0.2

0.4

0.6

0.8

1.0t=0.1 ht=5.0 ht=10.0 ht=15.0 ht=20.0 h

(m)0.00 0.01 0.02 0.03 0.04 0.05

c/c m

ax

0.0

0.2

0.4

0.6

0.8

1.0t=0.1 ht=5.0 ht=10.0 ht=15.0 ht=20.0 h

(m)0.00 0.01 0.02 0.03 0.04 0.05

c/c m

ax

0.0

0.2

0.4

0.6

0.8

1.0t=0.1 ht=5.0 ht=10.0 ht=15.0 ht=20.0 h

Figure 3.8 Solvent concentration distribution at different locations along the transition

zone at different moments: (a) Solvent chamber profiles; (b) Solvent concentration

distributions at the top; (c) Solvent concentration distributions in the middle; and (d)

Solvent concentration distributions at the bottom of the transition zone.

(a) (b)

(c) (d)

Page 80: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

59

3.2.2 Dividing number of the transition zone

Figure 3.9 shows the effect of the dividing number of transition-zone segments on

the solvent chamber profile. The transition zone is respectively divided into 5, 10, 15, 20,

and 25 segments and the corresponding average oil production rates are shown in Figure

3.9. It can be seen that the more segments, the higher the average oil production rate.

However, the production rate incremental decreases with the increase of total segment

numbers. In addition, more segments involve much longer computation time due to the

numerical solution method applied to the mass transfer model. Therefore, it is important to

divide the transition zone with a reasonable number of segments. This study applies 10

segments to the transition zone in the following calculations.

3.2.3 Permeability

Permeability is one of the most important model properties. The previous studies

concluded that the stabilized heavy oil production rate is proportional to the square root of

the permeability. This study analyzed four different permeability values: 5, 50, 100, and

200 D. These values are chosen to be consistent with the experimental permeability ranges

in the literature. Figure 3.10 demonstrates the solvent chamber profiles from the very

beginning to 2, 4, 6, 8, and 10 h for the four permeability cases. As expected, a higher

permeability will lead to faster oil recovery. Figure 3.11 compares the stabilized heavy oil

production rate against the square root of the permeability, and a good linear trend was

regressed with R2 = 0.9696, which agrees well with the conclusions in the previous studies.

Page 81: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

60

Dividing number

0 5 10 15 20 25 30

Oil

prod

uctio

n ra

te (c

c/h)

0.10

0.11

0.12

0.13

0.14

Figure 3.9 Effect of dividing number on the average oil production rate.

Page 82: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

61

The new VAPEX model is able to deal with both constant and variable diffusion

coefficients. Figure 3.12a shows a good linear dependence of the stabilized heavy oil

production rate on the square root of the constant diffusion coefficient, which is consistent

with the Butler−Mokrys model. This study further analyzes the effect of variable diffusion

coefficients. Figure 3.12b displays the recovery factor curve for three variable diffusion

coefficients, which adopts the correlations in Eqs. (3.3−3.4) with the same exponent of

−0.46 but different coefficients of α = 1.5, 1.0, and 0.5×10−9. As expected, a larger

diffusion coefficient will make the solvent chamber grow faster than a smaller diffusion

coefficient. Figure 3.12b also compares a constant diffusion coefficient (D = 2.29×10−9

m2/s) with an equivalent variable diffusion coefficient (D = 0.5×10−9μ−0.46). It is found that

the RF curve for the former case is much smaller than that for the latter case. This implies

that the constant diffusion might underestimate the oil production rate of a VAPEX process

because the concentration-dependent variable diffusion coefficient becomes larger and

larger with time, whereas the constant diffusion coefficient does not. Therefore, the

constant diffusion coefficient may be acceptable for a short time but unacceptable for a

longer time of simulation.

Page 83: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

62

x (m)0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

0.00

0.02

0.04

0.06

0.08

0.10

t=0 ht=2 ht=4 ht=6 ht=8 ht=10 h

x (m)0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

0.00

0.02

0.04

0.06

0.08

0.10

t=0 ht=2 ht=4 ht=6 ht=8 ht=10 h

x (m)0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

y (m

)

0.00

0.02

0.04

0.06

0.08

0.10

t=0 ht=2 ht=4 ht=6 ht=8 ht=10 h

x (m)0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14

0.00

0.02

0.04

0.06

0.08

0.10

t=0 ht=2 ht=4 ht=6 ht=8 ht=10 h

k = 5 Darcy k = 50 Darcy

k = 100 Darcy k = 200 Darcy

y (m

)

y (m

)y

(m)

Figure 3.10 Effect of permeability on the solvent chamber evolution: (a) k = 25 D; (b)

k = 50 D; (c) k = 100 D; and (d) k = 200 D.

(a) (b)

(c) (d)

Page 84: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

63

0 2 4 6 8 10 12 14 16

q o (cc

/h)

0

1

2

3

4

5

(Darcy0.5)

Figure 3.11 Effect of permeability on the oil production rate.

20.3394 , 0.9696q K R

K

Page 85: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

64

(m2/s)1/2

0.00000 0.00002 0.00004 0.00006 0.00008 0.00010 0.00012

q o (cc

/h)

0.0

0.5

1.0

1.5

2.0

2.5

(a)

Time (h)

0 2 4 6 8 10 12 14 16

Oil

reco

very

fact

or (%

)

0

2

4

6

8

10

12

14

16

18

20

D=2.25e-9

D=0.5e-9m-0.46

D=1.0e-9m-0.46

D=1.5e-9m-0.46

(b)

Figure 3.12 (a) Heavy oil production rate vs. square root of diffusion coefficient; and

(b) Oil recovery factor for variable and constant diffusion coefficients.

225669 , 0.9275q D R

9

9 0.46

9 0.46

9 0.46

2.25 10

0.5 101.0 101.5 10

DDDD

D

Page 86: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

65

3.2.4 This study vs. analytical models

Existing analytical models including the Bulter−Mokrys model, the Das−Butler

model, and the Yazdani−Maini model [Yazdani and Maini, 2005], and the model in this

thesis are all applied to calculate the heavy oil production rate for the base case. Four

permeability values are considered: k = 25, 50, 100, and 200 D. Figure 3.13 shows the

calculation results. The new model’s prediction is close to that of the Yazdani–Maini

model and much higher than those of the other two models. The reason may be that the first

two models were developed on the basis of the steady-state mass transfer and constant

diffusion coefficients. The coefficients in the Yazdani–Maini model incorporated the

effects of all system variables, such as porous media, grain size, and convective dispersion.

However, these empirical coefficients are regressed for specific laboratory tests only and

their applicability for a general VAPEX process remains questionable.

3.2.5 This study vs. numerical simulation

A numerical simulation model is developed by using the CMG STARS module

[Version 2011, Computer Modelling Group Limited, Canada] in this section to compare

with the new mathematical model in this research. Properties of the simulation model are

listed in Table 3.2 and Figure 3.14.

Page 87: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

66

k (Darcy)

0 20 40 60 80 100 120 140 160 180 200 220

q o (cc

/h)

0.00

0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40This studyYazdani-Maini (2005)Yazdani-Maini, (2005)Das-Butler (1995)Butler-Mokrys, (1989)

Figure 3.13 Oil production rate predicted by this study and the existing VAPEX

models.

Page 88: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

67

Table 3.2 Physical parameters and operation conditions for the base case of the

numerical simulation.

Parameters Value Model dimensions (lab-scale), m3 0.4 × 0.02 × 0.1 Model grid (lab-scale) 40 × 2 × 10 Run time (lab-scale), d 3 Model dimensions (field-scale), m3 1 × 1 × 1 Model grid (field-scale) 100 × 100 × 10 Run time (field-scale), d 3,000 Porosity, vol.% 35 Permeability, D 50 KV1 (k value correlation), kPa 2.31106 KV2 (k value correlation), 1/kPa 0 KV3 (k value correlation) 0 KV4 (k value correlation), C −2,725.4 KV5 (k value correlation), C −273.15 Dispersion coefficients, m2/d 0.000864 Injection pressure, kPa 800 Production pressure, kPa 799 Gas/liquid relative permeability curve Figure 3.14

Page 89: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

68

(a)

(b) (c)

Figure 3.14 (a) Numerical simulation model; (b) Relative permeability vs. liquid

saturation curve; and (c) Capillary pressure vs. liquid saturation curve.

Page 90: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

69

Numerical dispersion

Numerical dispersion is always a major concern in a numerical simulation, since it is

inherent in the finite difference simulation method and arises from time and space

discretization [Smith, 1985; Fanchi, 2006; Chen, 2007]. Numerical dispersion depends on

gridblock size x, timestep size t, as well as numerical formulation [Fanchi, 2006]. Fig.

13(a) displays the effect of the timestep size on the cumulative heavy oil production. Four

timestep sizes (t = 0.1, 0.01, 0.001, and 0.0001 d) are run and it can be seen that the

cumulative heavy oil production for t = 0.001 d behaves strangely near the end of

production. This shows the instability of simulation results caused by the timestep size.

Figure 3.16 presents four scenarios of grid sizes. Comparing Scenarios #2 and #4, it is

found that the heavy oil production rate is quite sensitive to the grid size as well as to the

geometric ratio. In contrast, although the mass-transfer model in this study is numerically

solved, it suffers less numerical dispersion because of the small grid size (ξ ≈ 0.0001 m)

applied to the transition-zone segment discretization. Therefore, the diffusion coefficient

of this study is used to roughly estimate the numerical dispersion in the simulation results.

First, the stabilized heavy oil production rate is calculated for the scenarios in Figure 3.16.

Then the coefficient α in Eq. (3.2) is adjusted to make the heavy oil production rate for each

case equal to that of the numerical simulation. Finally, the equivalent numerical dispersion

is estimated by subtracting the results for the original diffusion coefficient from those

adjusted one. Table 3.3 listed the results, suggesting that the error caused by gridding could

be as high as 60% and a smaller grid size can lead to more reliable stable simulation results.

Page 91: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

70

Time (d)

0 1 2 3

Cum

ulat

ive

oil p

rodu

ctio

n (c

c)

0

20

40

60

80

100

120

140

160

180

200

t = 0.0001 d t = 0.001 d t = 0.01 d t = 0.1 d

Figure 3.15 Effect of the timestep size on the cumulative oil production (grid size: 0.01

× 0.01 × 0.01 m3).

Page 92: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

71

Time (d)

0 1 2 3

Cum

ulat

ive

oil p

rodu

ctio

n (c

c)

0

20

40

60

80

100

120

140

160

180

Scenario #1Scenario #2Scenario #3Scenario #4

Figure 3.16 Effect of the grid size on the cumulative oil production (t = 0.001 d).

Page 93: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

72

Table 3.3 Effect of the grid size and estimation of the numerical dispersion.

Parameters Scenario #1 Scenario #2 Scenario #3 Scenario #4 This study ΔX, m 0.005 0.01 0.01 0.02 ― ΔY, m 0.01 0.02 0.005 0.01 ― ΔX:ΔY 1:2 1:2 2:1 2:1 ― Oil production rate, cc/h 1.729 1.685 1.511 1.589 1.463

Effective diffusion coefficient, m2/s 0.000778 0.000778 0.000778 0.000778 ―

Original α (αor) ― ― ― ― 1.306×10−9 Adjusted α (αad) 2.13×10−9 1.97×10−9 1.42×10−9 1.65×10−9 ― Relative error, % 63.093 50.842 8.729 26.341 ―

Note: The original α is equivalent to the constant diffusion coefficient. The adjusted α

matches the predicted production rates by using this study’s model with the numerical

simulation result. Relative error = (αad−αor)/αor×100%.

Page 94: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

73

Diffusion coefficient

One of the most important mechanisms of VAPEX is the solvent−heavy oil mass

transfer, which is characterized by the diffusion coefficient. This section further analyzes

the sensitivity of variable and constant diffusion coefficients to the modeling results of this

study and the numerical simulation, respectively. Five variable diffusion coefficients (D =

5, 1, 0.5, 0.1, and 0.05 × 10−9μ−0.46 m2/s) and five equivalent constant diffusion coefficients

in oil phase (D = 22.9, 4.58, 2.29, 0.458, and 0.229 × 10−9 m2/s) are applied for one

lab-scale and one field-scale models. Figure 3.17 demonstrates the results. It is found that

for the lab-scale cases, the oil production rates are quite sensitive to diffusion coefficient

for both this study and the numerical simulation. However, for the field-scale cases, the

new model’s results keep the similar trend to that in the lab-scale cases, whereas the

simulation results become insensitive to the diffusion coefficients. For D = 2.29×10−10 to

4.58×10−9 m2/s in the numerical simulation, though the latter diffusion coefficient is 20

times of the former one. Its corresponding heavy oil production rate (7.04 m3/d) is just 1.14

times of that of the former one (6.15 m3/d). This insensitivity is probably caused by the

larger grid size, resulting larger numerical dispersion in the field-scale simulations.

Transition-zone thickness

Transition zone is the most important area for the VAPEX heavy oil recovery process,

since it contributes to the most of the heavy oil production. In order to locate the transition

zone and capture the mass transport phenomena inside it, Yazdani and Maini [2007] stated

that the grid size of simulation model should be smaller than the transition-zone thickness,

which was estimated to be approximately 1 cm in the literature [Das et al., 1995;

Moghadam et al., 2008; Yazdani and Maini, 2009a]. Finer grids enable a more detailed

Page 95: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

74

D (m2/s)

0.0 5.0e-9 1.0e-8 1.5e-8 2.0e-8 2.5e-8

q o (cc

/h)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

2.0

Numerical simulationThis study

(a)

D (m2/s)

0.0 5.0e-9 1.0e-8 1.5e-8 2.0e-8 2.5e-8

q o (m

3 /d)

0

2

4

6

8

10

Numerical simulationThis study

(b)

Figure 3.17 Effect of the diffusion coefficient on the heavy oil production rate: (a)

Lab-scale grid size simulation results; and (b) Field-scale grid size simulation results.

Page 96: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

75

(a)

(b)

(c)

Figure 3.18 Mole fraction of solvent in the lab-scale numerical model with different

grid-sizes at 20 h (t = 0.01 d): (a) 0.02 × 0.02 × 0.02 m3; (b) 0.01 × 0.01 × 0.01 m3; and

(c) 0.005 × 0.005 × 0.005 m3.

Page 97: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

76

description of the component exchanges in the transition zone, yet the small grid size

and the accompanied minor time step will be limited by the computational time. Figure

3.18 shows the mole fraction distribution of solvent in the lab-scale numerical simulation

results with different grid sizes. For a grid size of 0.02 × 0.02 × 0.02 m3, the transition zone

has approximately one grid at the top, two grids in the middle, and more than three grids at

the bottom, corresponding to a thickness of ~2 cm at the top, ~4 cm in the middle, and over

6 cm at the bottom. This happens similarly to the cases of 0.01 × 0.01 × 0.01 m3 and 0.005

× 0.005 × 0.005 m3. Obviously, this estimate is much larger than the experimentally

measured and mathematically calculated ranges (between 0.3 and 1.5 cm) in the literature

[Samane et al., 2008]. In contrast, the assessment on the transition-zone thickness made in

this study (~1 cm at the top and middle and ~2 cm at the bottom) is relatively closer to the

previous conclusion, as clearly displayed in Figure 3.19.

In summary, in comparison with the numerical simulation model, this study’s model

demonstrated more sensitivity to diffusion coefficient. In addition, it is able to more

accurately describe the properties of the heavy oil and solvent inside the transition zone.

Page 98: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

77

Figure 3. 19 Comparison of the predicted transition-zone thickness of this study and the

numerical simulation (grid size: 0.02×0.02×0.02 m3; t = 0.01 d) at 20 h.

Page 99: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

78

3.3 Chapter Summary

This chapter formulates a new mathematical model to predict the VAPEX process.

The following conclusions can be drawn:

1. The new VAPEX mathematical model is developed on the basis of its major

mechanisms, such as mass transfer and gravity drainage;

2. The new model is able to describe the solvent concentration, oil viscosity and

density, diffusion coefficient, and drainage velocity inside the transition zone;

3. The evolution of the solvent chamber during its spreading and falling phases, as

well as the heavy oil production rate can be predicted by using the new model;

4. This new model confirms the linear correlations of qo vs. k and qo vs. D

for a permeability range of 5−200 Darcy and 50−0.5 × 10−10μ0.46, respectively.

5. It is found that the constant diffusion coefficient applied in the existing analytical

models may underestimate the oil production rate because it disregards the

growth of the diffusion coefficient during the VAPEX process; and

6. In comparison with the numerical simulation, the new model presented in this

chapter demonstrates more sensitivity to the diffusion coefficient and has much

less numerical dispersion.

Page 100: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

79

CHAPTER 4 MATHEMATICAL MODELING OF THE

CONVECTION−DIFFUSION MASS-TRANSFER PROCESS

Due to the inherent slow oil production rate of VAPEX, another solvent-based

method, CSI, is studied in this chapter. A convection−diffusion mathematical model is

developed for the mass-transfer process in the CSI process. Convection velocity represents

the effect of pressure gradient between the solvent chamber and untouched heavy oil zone.

In this model, variable diffusion coefficient and convection velocity are considered and a

special approximation method for them is applied to obtain the semi-analytical solution.

Results qualitatively show that the mass-transfer process between solvent and heavy oil

can be significantly enhanced by the bulk motion of the solvent due to the pressure gradient

during the solvent injection period of the CSI process, especially at the early stage.

4.1 CSI Process

As a solvent-based EOR method, CSI showed promising potential to recover heavy

oil and bitumen in thin heavy oil reservoirs. CSI is basically a solvent huff-n-puff process.

Typically, each cycle consists of three periods, as schematically shown in Figure 4.1. First,

a vapour solvent is injected into the reservoir at a high pressure for some time (injection

period). Then the well is shut in for a period of time to let the solvent soak into the crude oil

(soaking period). Finally, the well is opened and its pressure is reduced so that the

solvent-diluted crude oil can be produced from the reservoir (production period). After one

cycle, the well would be converted into an injector again and the entire process will be

repeated for another cycle, until the oil production rate reaches an economic limit.

Page 101: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

80

(a) Injection (b) Soaking (c) Production (d) Flow velocity

Figure 4.1 Vapour solvent-based ‘huff-n-puff’ process (Note: bold white arrows point

to the solvent diffusion direction, whereas narrow black arrows point to convection

direction).

x

V Injector/ producer Solvent

Transition zone

Dead oil

Solvent Transition

zone

Dead oil

Solvent Transition

zone

Dead oil

High Low

Page 102: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

81

This chapter focuses on the mass-transfer process during the solvent injection period,

during which the pressure of the injected solvent is higher than that of the untouched heavy

oil. This causes a pressure gradient between the solvent chamber and the untouched crude

oil. Under the effect of the pressure gradient, solvent would have a bulk motion that could

accelerate the mixing process between solvent and crude oil. This chapter analyzes the

contribution from the pressure gradient to the mass-transfer process.

4.1.1 Convection−diffusion equation

One foremost feature of all solvent-based EOR techniques is oil viscosity reduction

due to the mass transfer between crude oil and solvent: solvent molecules mix with the bulk

heavy oil through Brownian motion (concentration gradient) and/or bulk motion (pressure

gradient). Without the latter bulk motion, the mass transfer between solvent and crude oil is

a diffusion process that can be modelled by using the Fick’s law. With the bulk motion, the

mass-transfer process is modelled by using the convection−diffusion equation:

'c cD cV St x x

, (4.1)

where, V is the convection velocity, m/s; S’ is the source/sink term. Diffusion coefficient

describes the effect of random walk of the diffusing particles, whereas the convection

velocity represents the effect of the bulk motion between the solvent and the crude oil.

4.1.2 Diffusion coefficient and convection velocity

Viscosity

The viscosity of the solvent-diluted heavy oil is calculated by using the Lederer−Shu

correlations [Lederer, 1933; Shu, 1984], as specified in Eqs. (3.5−7).

Page 103: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

82

Diffusion coefficient

The diffusion coefficient of a hydrocarbon solvent into crude oil is determined by

using the Das and Butler correlations which are back-calculated on the basis of their

VAPEX experiments, as specified in Eqs. (3.3−4).

Convection velocity

Convection velocity in a porous medium can be described by using the Darcy’s law

[1933]. Under the effect of a pressure gradient and gravity force, it is

k dPV gdx

. (4.2)

where, dPdx

is the pressure gradient, kPa/m; Δρ is density difference between the

solvent-diluted crude oil and the liquid solvent, kg/m3; The density of solvent-diluted oil

can be determined by using the mixture rule for an ideal solution, as shown in Eq. (1.9).

Convection velocity is commonly treated as a constant mean value [Scott and Jirka,

2002] in various previous studies. This simplification is acceptable for the cases where

fluid is incompressible and the flow velocity is quite uniform, such as tracer flow [Sposito

and Weeks, 1998]. However, in solvent-based EOR processes, this simplification may be

unreasonable. More specifically, the diffusion coefficient and convection velocity are both

functions of viscosity and viscosity is a function of concentration, which is further a

function of time and location, as shown by Eqs. (3.3−5, 4.2). Thereby, at a certain time,

diffusion coefficient and convection velocity are both functions of location inside the

transition zone: V = V[c(x)] and D = D[c(x)].

Using the parameters of a base case in Table 4.1, a concentration profile is calculated

and shown in Figure 4.2a, based on which a viscosity curve is calculated by using Eq. (3.6)

Page 104: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

83

Table 4.1 Physical properties and operating conditions of the base case.

Property Value Length, m 0.01 Permeability, D 10 Oil gravity 0.975 Solvent gravity (liquid) 0.517 Oil viscosity, mPas 6000 Solvent viscosity (liquid), mPas 0.1 Pressure gradient, kPa/m 5 Diffusion coefficient, m2/s 1.306×10−9μ−0.46 Time, s 600 Solubility, g solvent/g oil 0.26 Inlet solvent concentration, fraction 0.329 Inlet diffusion coefficient, m2/s 1.21×10−8

Page 105: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

84

and plotted in Figure 4.2b. The associated diffusion coefficient and convection velocity

curves are computed by using Eqs. (3.4) and (4.1) and plotted in Figures 4.2c–d,

respectively. Calculations in the following sections are all based on the parameters of the

base case. For some special cases, their particular parameters will be noted.

4.2 Mathematical Models

4.2.1 Governing equation

Considering the effects of the concentration gradient and pressure gradient on the

mass transfer process between two materials, one dissolving into another (i.e., crude oil

and solvent), and disregarding the source/sink term, the governing equation would be:

c cD cVt x x

, (4.3)

4.2.2 Boundary and initial conditions

One boundary of the modeling object, transition zone, is assumedly saturated with

solvent at all times, which means a Dirichlet boundary condition (BC). The other boundary

of the transition zone is regarded as a Neumann BC. Initially, the entire model is free of

solvent. The BCs and initial condition (IC) are described by:

*( 0, ) , 0x tc c t

, (4.4)

( , )

0, 0x t

c tx

, (4.5)

( , 0) 0, 0x tc x

, (4.6)

Page 106: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

85

x (m)0.000 0.002 0.004 0.006 0.008 0.010

c

0.0

0.1

0.2

0.3

0.4

x (m)0.000 0.002 0.004 0.006 0.008 0.010

D (m

2 /s)

0.0

2.0e-9

4.0e-9

6.0e-9

8.0e-9

1.0e-8

1.2e-8

1.4e-8

x (m)0.000 0.002 0.004 0.006 0.008 0.010

V (m

/s)

0

1e-6

2e-6

3e-6

4e-6

5e-6

6e-6

7e-6

x (m)0.000 0.002 0.004 0.006 0.008 0.010

c

P

0

1000

2000

3000

4000

5000

6000

7000

Figure 4.2 Concentration-dependent diffusion coefficient and flow velocity: (a)

Concentration; (b) Viscosity; (c) Effective diffusion coefficient; and (d) Convection

velocity.

(a) (b)

(c) (d)

Page 107: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

86

where, c* is maximum concentration under the operating conditions, fraction; is

transition zone thickness, m. It is worthwhile to note that in fact, the boundaries of the

transition zone are moving and the solvent chamber is expanding all the time, which makes

the mass transfer model become a free boundary problem. This study does not consider the

whole mass transfer process, but focuses on a mass-transfer process in a short time interval

and thus uses fixed-boundary conditions to simplify the problem.

4.3 Semi-Analytical Solutions

4.3.1 Model #1: Convection–diffusion model with constant D and variable V

This model considers a convection–diffusion equation with a constant D and a

variable V. With the definitions of following dimensionless variables and number:

Dinlet

ccc

, 2Dt Dt

, Dxx

, VPeD

, (4.7)

where, cinlet is the inlet concentration at x = 0, which is equal to c* under the operating

conditions, fraction; Pe is Péclet number, dimensionless. The governing equation, BCs,

and IC of Model 1 can be normalized as:

2

2

( 0, )

( 1, )

( , 0)

1 0

0 0

0 0 1

D D

D D

D D

D D DD

D D D D

D Dx t

DD

D x t

D Dx t

c c c PePe ct x x x

c t

c tx

c x

. (4.8)

Pe measures the relative importance of convection to diffusion during a mass-transfer

process. It is worthwhile to state that in this thesis, Pe is a variable rather than constant

dimensionless number, and it actually represents the convection velocity.

Page 108: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

87

Solution to Model #1with a constant D and a linear V

In order to derive the solution to a convection–diffusion model with a general Pe

profile, a simple convection–diffusion model with a simple linear Pe profile is studied as

the first step:

DPe ax b , (4.9)

where, a and b are the slope and x-intercept of a linear Pe profile. a and b should meet a<0,

b>0, and |b|>|a| for a positive flow velocity (direction of V is consistent with that of D), and

a>0, b<0, and |b|>|a| for a negative flow velocity (direction of V is opposite to that of D).

Substituting Eq. (4.8) for Pe in Eq. (4.8):

2

2D D D

D DD D D

c c cax b act x x

. (4.10)

Performing the Laplace transformation and integrating IC, Eq. (4.10) can be transformed

into an ordinary difference equation (ODE):

22

21 04

d C z J Cdz

, (4.11)

where

2

4 2D Da bx x

DC c e

, (4.12a)

, 0Dax bz aa

or , 0Dax bz aa

, (4.12b)

1 , 02

sJ aa

or 1 , 02

sJ aa

, (4.12c)

where, C and z are transformed cD and x in the Laplacian domain, respectively; s is the

Laplacian operator. Eq. (4.11) is the canonical form of the parabolic cylindrical function

[Abramowitz and Stegun, 1970], to which the general solution is in the following form:

Page 109: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

88

1 2C AC BC . (4.13)

Here, C1 and C2 are two independent solutions to Eq. (4.11):

21 2

41

1 1, ( , , )2 4 2 2

z J zC J z e M

, (4.14)

21 2

42

3 3, ( , , )2 4 2 2

z J zC J z ze M

, (4.15)

where, M is the Kummer’s function; A and B are two constant coefficients and can be

obtained by applying the two BCs. Appendix A presents the detailed derivation of the

solution to Model #1 with a constant D and a linear V.

Solution to Model #1 with a constant D and a variable V

On the basis of the above foundation, a more general variable Pe profile is studied

here. As shown in Figure 4.3, a special approximation method is applied to solve a

convection–diffusion model with an arbitrary monotonous smoothly curved Pe profile: the

curved V profile is approximated with a piecewise linear (n segments) profile. Since the

governing equation and IC for each segment are the same as those in Model #1 with a

constant D and a linear V, the transformed dimensionless concentration in Laplacian

domain for each segment should have the same form as Eq. (4.13):

1, 2, 1i i i i iC AC B C i n , (4.16)

where, Ai and Bi can be obtained by applying the BCs of the ith segment: D,i

Di

D x

c qx

for

the left boundary and , 1

1

D i

Di

D x

c qx

for the right one (qi and qi+1 denote the mass-transfer

rates at the left and right boundaries of the ith segment, respectively). It is worthwhile to

Page 110: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

89

note the left BC for the first segment is 01

DD x

c , and that the right BC for the last

segment is 1

0D

D

D x

cx

.

Back replacing the determined Ai and Bi into the Eq. (4.16), C is obtained over the

entire spatial domain as functions of qi:

* *1 1 1 2,C J z A B q , (4.17a)

* *1, 1 < <i i i i iC J z A q B q i n , (4.17b)

* *,n n n nC J z A q B . (4.17c)

Considering the continuity condition for the solvent concentration at the interior common

boundaries between any neighboring two segments:

1D,i D,i

D,i D,ix xc c , (4.18)

and in the Laplacian domain, this is

2 21 1

, ,

4 2 4 21

i i i i

D i D i

a b a bx x x x

i ix x

C e C e

. (4.19)

Applying Eq. (4.19) to Eqs. (4.17a−c), n−1 equations can be obtained:

* * * *2 2 1 2 2 2 3 1A B q B q A , (4.20a)

* * * *1 1 1 1 0 2 < <i i i i i i i i iA q A B q B q i n , (4.20b)

* * * *1 1 1n n n n n n n nA q B A q B , (4.20c)

where

21 1

4 2i i i ia a b bx x

i e

. (4.21)

Page 111: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

90

Figure 4.3 Approximation to the convection velocity with a piecewise linear profile.

max( 0, )x tc c

( , )

0x L t

cx

( , 0) 0x tc

( , ) ?x tc

x1 i ii ix x

c c

Pe

Page 112: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

91

Eqs. (4.20a−c) can be coupled altogether to form a linear system:

1 1 1 1n n n n M q F , (4.22)

where, the coefficient matrix [M](n-1)×(n-1) is a tridiagonal matrix; {q}n-1 is the

to-be-determined unknown column matrix; the column matrix {F}n-1 and the coefficient

matrix [M](n-1)×(n-1) can be constructed given a piecewise linear Pe profile. {q}n-1 can be

solved by using the Thomas algorithm [Muller, 2001]. Back replacing it into Eqs.

(4.17a−c), C over the entire space can be obtained. Finally, solvent concentration in

physical domain can be obtained by using the Stehfest Laplace inverse transform [Stehfest,

1970] (loop number in this study is set as 8).

4.3.2 Model #2: Convection–diffusion model with variable D and variable V

This model considers both diffusion coefficient and convection velocity as variables:

D=D(x) and V=V(x). With the definitions of following dimensionless variables and

number:

Dinlet

ccc

, Dinlet

DDD

, Dxx

, 2inlet

Dt Dt

, inlet

VPeD

, (4.23)

where Dinlet is the diffusion coefficient at the left/inlet boundary under the operating

conditions, m2/s. The governing equation, BCs, and IC of Model #2 can be normalized as:

2

2

( 0, )

( 1, )

( , 0)

1 0

0 0

0 0 1

D D

D D

D D

D D D DD D

D D D D D

D Dx t

DD

D x t

D Dx t

c c D c PeD Pe ct x x x x

c t

c tx

c x

. (4.24)

Page 113: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

92

Similar to Model #1, a simple case for Model #2 with a linear DD and a linear Pe is first

studied as a first step for a more general case:

DPe ax b , ' 'D DD a x b . (4.25)

where 'a and 'b are the slope and x-intercept of the linear DD profile. Substituting Eq.

(4.25) for DD and Pe in the governing equation of Model #2:

2

2' ' 'D D DD D D

D D D

c c ca x b a ax b act x x

. (4.26)

Eq. (4.26) can be analytically solved in the same way as that for Model #1 with a constant

D and a linear V. The general solution is:

1, ; 1, 2 ;C AM B M , (4.27)

here, definitions of ϛ, ξ, and ε are provided in Appendix B; coefficients A and B can be

determined by applying the BCs. The solvent concentration in the physical domain can be

obtained by conducting the Stehfest Laplace inverse transformation (loop number in this

study is chosen as 8).

For a more general case where DD and Pe profiles are arbitrary monotonous smooth

curves, DD and Pe can be respectively approximated to have a piecewise linear profile.

Then the model can be semi-analytically solved by using the same approach as described in

the previous section. It is worthwhile to note that DD and Pe profiles must be divided into

the same segments on the x axis when the approximations of them are made.

The above-derived solutions are for the IC of cD(xD,0) = 0. In the case of tD > 0, the

transformed governing equation in the Laplacian domain would be a non-homogeneous

equation whose general solution can be obtained:

1 2 1 2' 'C AC BC A C B C . (4.28)

Page 114: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

93

Here, A and B are the coefficients for the corresponding homogeneous equation and have

the same values as those in Eq. (4.16); A’ and B’ can be determined by using the BCs and

the method of undetermined coefficients [Zill, 2001].

4.4 Validations

4.4.1 Validation with an analytical solution for a special case

Considering a special case of Model #1 where D is a constant and Pe is a hyperbolic

function of xD:

2D

D

Pe xx

, (4.29)

where, ψ is an arbitrary constant. The analytical solution to Eq. (4.29) can be obtained in

the Laplacian domain (the detailed derivations are given in Appendix C) as:

(2 )

2( )

(1 )

D Dsx s x

s

e eC ss e

, (4.30)

where

(1 ) 1(1 ) 1

ss

. (4.31)

Then the hyperbolic Pe profile is approximated with a piecewise linear profile, and the

model is semi-analytically solved by using the aforementioned approach. Figure 4.4

compares the analytical and semi-analytical solutions, suggesting that the semi-analytical

solution is not reliable when the Pe profile is roughly approximated with five segments but

rather accurate with twenty-five segments. The accuracy of the semi-analytical solution

depends on the approximation to the Pe profile—The better approximation is, the more

accurate the semi-analytical result would be. However, too many segments would greatly

Page 115: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

94

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6Semianalytical, N=5Semianalytical, N=15Semianalytical, N=25Analytical

xD0.0 0.2 0.4 0.6 0.8 1.0

Pe

0

2

4

6

8

10

Figure 4.4 Semi-analytical vs. analytical cD for a convection–diffusion mass transfer

with a special convection velocity

20.25D

Pex

Page 116: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

95

increase the computational time but trivially enhance the incremental accuracy. Therefore,

the segment number for obtaining a reliable and precise solution varies with the linearities

of the D and V profiles. It is worthy of stating that in order to improve the proximity of a

piecewise linear profile, the segment-division can be densified where Pe changes

drastically and sparsed where it varies slowly, rather than evenly distributed.

4.4.2 Validation with the numerical solution

The Crank−Nicolson finite difference method (FDM) with a truncation error of

O(∆x4) is applied to acquire the numerical solution to the aforementioned

convection–diffusion model. Two spatial grid sizes (0.00005 and 0.0001 m) and two time

steps (10 and 20 s) are used for the numerical solution. Figure 4.5 compares the

semi-analytical and numerical solutions, showing the numerical solution with a grid size of

0.00005 m and a time step of 10 s gives the best match with the semi-analytical solution.

This suggests that the numerical solution is accurate enough as long as the grid size and

time step are sufficiently small.

4.5 Results and Discussion

4.5.1 Application of the convection–diffusion mass-transfer model

The convection−diffusion mass transfer model is applied to a CSI process (Figure

4.1). The solvent concentration distribution inside the transition zone is calculated by using

the semi-analytical solutions to the above models. Eqs. (3.3−5, 4.2) show that D and V are

both functions of and is a function of c, which make the governing equation a

non-linear partial differential equation (PDE). In this study, this non-linear PDE is solved

in a special way. First, the time domain is divided into a number of steps. Second, at one

Page 117: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

96

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2Numerical-1Numerical-2Numerical-3Semianalytical

Figure 4.5 Semi-analytical vs. numerical cD.

Page 118: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

97

time step, D and V are treated as functions of the solvent concentration at the end of the

previous time step; Therefore, D and V can be explicitly plotted so that the governing

equation becomes a linear PDE that is semi-analytically solved by using the method

provided in this study. Then the solvent concentration can be calculated and its value at the

end of that step will be used as the initial condition for the next time step. Figure 4.6

schematically demonstrates the flow chart for calculating the solvent concentration in the

transition zone. First, a solvent concentration profile can be computed with the governing

equation, BCs, and IC at a certain time point. Then, based on the solvent concentration, μ

and of solvent-diluted crude oil can be calculated by using Eqs. (1.6−8) and Eq. (1.9),

respectively; then D and V can be respectively obtained by using Eqs. (3.5−6) and (4.2).

Finally, the IC can be updated by the present solvent concentration profile; D and V in the

governing equation should also be updated for the calculation in the next time point. Table

4.1 lists the parameters of a base case, and most calculations in the following part are for

the base case. For comparison cases, their particular parameters will be specified.

4.5.2 Variable and constant diffusion coefficient and convection velocity

The effects of constant and variable diffusion coefficient/flow velocity on the solvent

concentration distribution across the transition zone are analyzed in this section. Four cases

are considered:

Case #1: a variable D and a variable V

Case #2: a constant D and a variable V

Case #3: a variable D and a constant V

Case #4: a constant D and a constant V

Page 119: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

98

Figure 4.6 Flowchart of calculating the solvent concentration in the transition zone (t*

denotes the termination time).

Equations 1. Semi-analytical solutions 2. Mixture rule of ideal solution, Eq. 1.9 3. Lederer–Shu correlation, Eq. 1.6−8 4. Das–Butler correlation, Eq. 3.5−6 5. Darcy’s law, Eq. 4.2

Governing Eq. IC BCs

c

ρ

V

D Finish

Yes t > t*

1

2 3

4

5

Page 120: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

99

For comparison purposes, the constant D in Cases #2 and #4 is equivalent to the mean

value of the variable D over a range of [Dinlet, 0.01Dinlet], and the constant V in Case #3 and

#4 is equivalent to the mean value of the variable V over a range of [Vinlet, 0.01Vinlet]. Two

times (300 and 600 s) are calculated. Figure 4.7 compares the results of the four cases.

By Analyzing of the results for the four cases, several conclusions can be made: (1)

compared with Cases #2−4, cD for Case 1 (variable D and variable V) is underestimated

near the left boundary but overestimated at the other locations. (2) The deviation between

Case #1 and the rest cases are smaller at a shorter time t1 but larger at a longer time t2.

Because the constant D and V are equivalent to the variable D and V at the initial time and

are unchanged as time increases; however, the variable D and V become larger and larger

with time. Therefore, the constant D and V would be less than the mean values of the

variable D and V at a later time. (3) Although the integral area of cD curves for the four

cases are closer to each other, their solvent concentration profiles have quite distinct shapes.

This is most evident in Figure 4.7c. The shape of the cD curve is mainly determined by D

and V in the governing equation. (4) The effect of constant V on the cD profile is much

larger than that of D, as can be seen in Figures 4.7a and 4.7b. This implies that the pressure

gradient may play a larger role than the concentration gradient and that the crude

oilsolvent mass transfer can be strengthened by a pressure difference.

4.5.3 Effect of convection velocity

Pressure gradient

During the majority of huff-n-puff process, the pressure over the entire reservoir is

unbalanced: the pressure in the solvent chamber is higher (positive) than that in the

untouched crude oil zone during the huff period whereas lower (negative) than that in the

Page 121: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

100

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2t1: var. D & var. Vt2: var. D & var. Vt1: const. D & var. Vt2: const. D & var. V

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2t1: var. D & var. Vt2: var. D & var. Vt1: var. D & const. Vt2: var. D & const. V

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2t1: var. D & var. Vt2: var. D & var. Vt1: const. D & const. Vt2: const. D & const. V

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2t2: var. D & var. Vt2: const. D & var. Vt2: var. D & const. Vt2: const. D & const. V

Figure 4.7 Comparison of cD for different cases: (a) Variable D & variable V vs.

constant D & variable V; (b) Variable D & variable V vs. variable D & constant V; (c)

Variable D & variable V vs. constant D & constant V; and (d) Variable D & variable V

and D vs. constant D & variable V vs. variable D & constant V vs. constant D & constant

V (Constant D is equal to 5.8×10−9 m2/s; constant V is equal to 1.8×10−6 m/s; t1 = 300 s; t2

= 600 s).

(a) (b)

(c) (d)

Page 122: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

101

untouched crude oil zone during the puff period. This part analyzes the effect of the

pressure gradient on the solvent concentration distribution across the transition zone. It is

worthwhile to mention that in fact, the pressure gradient becomes smaller and smaller with

time. Here, a constant pressure gradient is used to simplify the calculation and qualitatively

illustrate the problem.

Figure 4.8 shows the cD profiles under the effect of positive, zero, and negative

pressure gradients. Compared with the pure diffusion process where the pressure gradient

equals zero, the cD profile is greatly improved by a positive pressure gradient of 5 kPa/m:

the integral area of the cD profile for 5 kPa/m is almost twice of that for 0 kPa/m. In contrast,

the cD profile is slightly shrunk by a negative pressure gradient of −5 kPa/m: the integral

area of the cD profile for −5 kPa/m is ~80% of that for 0 kPa/m. It is found that the positive

pressure gradient could prompt the solvent dissolution into the crude oil while the negative

pressure gradient would hinder the solvent dissolution into the crude oil, implying that the

transition zone expands during the huff period while shrinks during the puff period. This

also demonstrates one of the advantages of the cyclic solvent process (such as CSI) over

the continuous solvent process (such as VAPEX): the crude oilsolvent mixing process can

be more effective in CSI than in VAPEX.

Crude oil viscosity

Figure 4.9 presents the effect of the crude oil viscosity on the cD profile. Four

viscosities are chosen: 600, 6,000, 60,000, and 600,000 mPas. The results are quite

straightforward: (1) The solvent can dissolve further into the crude oil with a lower oil

viscosity, which means that less viscous crude oil can be more easily diluted by the solvent;

Page 123: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

102

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2grad_P=5.0 kPa/mgrad_P=2.5 kPa/mgrad_P=0 kPa/mgrad_P=-2.5 kPa/mgrad_P=-5.0 kPa/m

Figure 4.8 Effect of the pressure gradient on the solvent concentration distribution.

Page 124: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

103

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2mPa.smPa.smPa.smPa.s

Figure 4.9 Effect of crude oil viscosity on the solvent concentration distribution.

Page 125: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

104

(2) The transition-zone thickness increase is not linear: the dimensionless transition-zone

thickness is increased by 0.04 for a viscosity change from 600,000 to 60,000 mPas, by

0.12 of the unit dimensionless distance for a viscosity change from 60,000 to 6,000 mPas,

and by 0.24 for a viscosity change from 6,000 to 600 mPas. This indicates that the

solvent-based CSI process can be more effective for a relatively less viscous crude oil.

Note that the transition-zone thickness covers a distance from xD (cD = 1) to xD (cD = 0.01).

Diffusion coefficient

Figure 4.10 displays the effect of the diffusion coefficient on the solvent

concentration distribution. The diffusion coefficient correlation of propane and Peace

River bitumen [Das and Butler, 1996], Eqs. (3.5−6), is used as a basic formula. Parameter β

is kept constant as −0.46 since it is around −0.5 in most cases as cited in the literature; α

varies from 1.0×10−9 to 3.0×10−9. The results show that the cD profile with a smaller D

would have a sharper front yet a shorter diffusing distance while the cD profile with a larger

D would have a gentler front but a longer diffusing distance. This is because a smaller D

would make solvent molecules aggregate near the inlet, leading to a higher concentration

but sharper decline near the inlet since very little solvent can dissolve into the crude oil.

4.5.4 Péclet number

All the above-mentioned factors (i.e., pressure gradient, viscosity and diffusion

coefficient) can be integrated into a dimensionless number, Péclet number. The above

analyses indicate a general trend: an increased V can accelerate the crude oilsolvent

mixing. In practice, the permeability is easy to measure but the pressure gradient and

diffusion coefficient across the transition zone are difficult to determine. Therefore, some

simple Pe profiles are assumed to qualitatively analyze its effect on the cD distribution.

Page 126: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

105

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2=1.0e-9=1.5e-9=2.0e-9=2.5e-9=3.0e-9

Figure 4.10 Effect of diffusion coefficient on the solvent concentration distribution.

0.46D

Page 127: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

106

Concave, linear, and convex Pe profiles

Figure 4.11 displays the cD profiles for Pe curves of three different shapes: concave,

linear, and convex with the same starting and ending points and mean values. Comparing

the results, it is found that a declining Pe profile tends to make the solvent accumulate at

the inlet boundary, which is the bumping (cD > 1) portion in the cD profile. The sharper the

decline is, the more easily the solvent would aggregate—the amplitude of cD profile for the

concave Pe is the largest among the three scenarios.

Linear Pe profile with different slopes

Figure 4.12 shows the cD profiles for different linear Pe profiles with the same

x-incept but different slopes. It can be seen that a smaller Pe slope would lead to a lower cD

profile, indicating a less efficient crude oilsolvent mixing. In addition, Figure 4.12

demonstrates a proof to the conclusion generated in Figure 4.11: a larger decrease of Pe

could lead to a more notable bumping of the cD profile.

4.5.5 Effect of gravity force in natural convection

Effect of natural convection on the solvent concentration is studied. In some

solvent-based EOR processes, such as the rising phase of VAPEX and upwards leaching,

solvent diffuses upwards into the crude oil that is diluted and drained downward. In this

case, the flow direction is against the diffusing direction and the flow velocity in the

governing equation is negative. Figure 4.13 shows the concentration distributions of

propane and butane for a diffusion process with and without natural convection, suggesting

that butane with gravity force is the least efficient while propane without gravity force is

the most effective one. This means that the natural convection caused by density difference

may hinder the solvent from mixing with the crude oil.

Page 128: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

107

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2

ConvexLinearConcave

xD

0.0 0.2 0.4 0.6 0.8 1.0

Pe

0

2

4

6

8

Figure 4.11 Effect of Péclet number on the solvent concentration distribution.

Page 129: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

108

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2Pe=-8xD+8Pe=-6xD+6Pe=-4xD+4Pe=-2xD+2

xD0.0 0.2 0.4 0.6 0.8 1.0

c D

0

2

4

6

8

Figure 4.12 Effect of Péclet number with different linear shape on the solvent

concentration distribution.

Page 130: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

109

xD

0.0 0.2 0.4 0.6 0.8 1.0

c D

0.0

0.2

0.4

0.6

0.8

1.0

1.2Propane, without gravityPropane, with gravity Butane, without gravityButane, with gravity

Figure 4.13 Effect of gravity force on the solvent concentration distribution.

Solvent

Heavy oil

With gravity force Without gravity force

Solvent

Heavy oil

Page 131: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

110

4.6 Chapter Summary

This chapter develops two 1D convection–diffusion mass transfer models for the CSI

process: one model considers a constant diffusion coefficient and a variable flow velocity

and the other considers both parameters as variables. Semi-analytical solutions are

obtained and applied to analyze the mass transfer process between crude oil and solvent.

The following conclusions can be made:

1. The accuracy of the semi-analytical solution largely depends on the

approximation of the diffusion coefficient and convection velocity profiles—a

better approximation can lead to a more accurate solution.

2. The approximation of the actual variable diffusion coefficient with a constant

value can be inaccurate, since the latter does not consider the change of

diffusion coefficient throughout a CSI process. This is true for the convection

velocity.

3. The convection velocity can play a larger role than the diffusion coefficient in the

crude oilsolvent mixing process during a CSI process.

4. An increased convection velocity can accelerate the dissolution of solvent into

the crude oil during the solvent injection period of CSI.

5. Gravity force may reduce the mass transfer between crude oil and solvent due to

the natural convection.

Page 132: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

111

CHAPTER 5 FOAMY OIL-ASSISTED VAPOUR EXTRACTION

(F-VAPEX)

CSI benefits from a stronger mass transfer (Chapter 4) and a higher oil production

rate during the pressure reduction period but suffers from the unproductive and long

injection and soaking periods and the consequent low average oil production rate. This

chapter proposes a new process, namely foamy oil-assisted vapour extraction (F-VAPEX).

F-VAPEX combines VAPEX and CSI together to take advantage of the continuous

production of VAPEX and the stronger mass transfer and oil production mechanisms of

CSI. It is essentially a VAPEX process during which the operating pressure is cyclically

decreased and increased. Technical details, experimental results, and comparative analyses

of the new technique are presented in this chapter.

5.1 Experimental

5.1.1 Materials

The heavy oil sample was collected from a western Canadian heavy oil reservoir,

with a density of o = 976 kg/m3 and a viscosity of μo = 5,875 cP, both of which were

measured at the atmospheric pressure and a room temperature of 20.2C. The asphaltene

content of the original heavy oil sample was measured by using the standard ASTM D2700

method [2003] with filter papers (No. 5, Whatman, England) with a pore size of 2.5 μm and

found to be 17.69 wt.% (n-pentane insoluble). Propane with a stated purity of 99.5 mol.%

(Praxair, Canada) was used as the extracting solvent. Glass beads with an average size of

90−150 μm were used to pack the cylindrical and rectangular physical models.

Page 133: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

112

5.1.2 Experimental set-up

Figure 5.1 shows the schematic diagram of the experimental set-up, which is

comprised of four major operation units: a solvent injection unit, a physical model, a fluids

production unit, and a data acquisition system. The solvent injection unit consists of a

propane cylinder (Praxair, Canada), a two-stage gas regulator (KCY Series, Swagelok,

USA) installed on the propane cylinder, a solvent injection valve, and an injector.

The major component of the experimental set-up is a visual rectangular sand-packed

high-pressure physical model. This physical model has a rectangular cavity (40 10 2

cm3) grooved in a steel plate to be packed with sand. The front of the model is covered with

an acrylic as a visual window, through which the test process can be visualized and

photographed. A digital camera (Rebel T3, Canon, Japan), in conjunction with a florescent

light sources (Catalina Lighting, USA), is used to take digital images of the solvent

chamber during each VAPEX test. The technical details regarding the rectangular physical

model can be found elsewhere [Moghadam et al., 2008].

Two types of well configurations are adopted for the VAPEX and F-VAPEX tests

(Figure 5.2): (1) Central well configuration. The producer is set at the center bottom part of

the model, whereas the injector is placed 3 cm above the producer; (2) Lateral well

configuration. The injector and producer are positioned at the top right and bottom left

corners of the physical model, respectively. The lateral well configuration is applied to

simulate a pair of horizontal injector and producer with proper vertical and horizontal

separation distances. The well configuration for each test is specified in Table 5.1. It is

worthwhile to mention that one CSI test (Test #5.2) used a single well alternately as the

injector and producer to simulate a conventional solvent huff-n-puff process.

Page 134: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

113

Figure 5.1 Schematic diagram of the experimental set-up in this study.

Injection unit Physical model

Propane Camera

Data acquisition unit Production unit

Back-pressure regulator Digital pressure gauge

Pressure transducer

Notebook computer

Steel tubing Plastic flexible hose Data ware

Sand pack Injector Producer

Scale Surge flask Flow meter

Oil collector

Page 135: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

114

(a)

(b)

(c)

Figure 5.2 (a) Physical model dimensions; (b) Central well configuration; and (c)

Lateral well configuration.

20 cm 7 cm

Height = 10 cm

Width = 40 cm Thickness = 2 cm

Page 136: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

115

Table 5.1 Physical properties of the sand-packed model and experimental conditions

for the VAPEX, CSI and F-VAPEX tests.

Test No. Well Configuration Operating Scheme Cycle Length (h)

PV (cc)

Soi (%)

k (D)

(%)

5.1 VAPEX, Fig. 5.3a — 284.8 92.7 4.9 35.6

5.2 CSI, Fig. 5.3b 1 284.5 92.1 4.7 36.2

5.3 F-VAPEX, Fig. 5.3c 1 286.6 93.5 5.4 35.7

5.4 F-VAPEX, Fig. 5.3c 2 286.1 92.4 5.1 36.7

5.5 F-VAPEX, Fig. 5.3c 4 285.8 93.2 5.8 35.9

5.6 VAPEX, Fig. 5.3a — 287.1 94.5 4.9 35.8

5.7 F-VAPEX, Fig. 5.3c 1 283.2 93.8 4.7 35.9

5.8 F-VAPEX, Fig. 5.3c 2 288.8 94.1 5.2 36.3

Injector/ producer

Page 137: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

116

The production unit is comprised of a production valve, a high-sensitivity

back-pressure regulator (BPR) (LBS4 Series, Swagelok, USA), a produced oil collector, an

electronic scale (ML302E, Mettler Toledo, Switzerland), and a precision drum-type gas

flowmeter (TG05, Ritter, Germany). The produced oil was collected in a flask and weighed

by using the electronic scale to determine the cumulative oil production for all the tests.

The produced gas volume was recorded by using the precision drum-type gas flowmeter.

The data acquisition system includes a high-precision digital pressure transducer

(PPM-2, Heise, USA) and a notebook personal computer (PC) (Hewlett Packard, USA).

The injection pressure and production pressure are recorded in the notebook PC

automatically and continuously. The cumulative oil production and solvent production are

recorded into the notebook PC manually at a fixed time interval of 1 h during each test.

5.1.3 Experimental preparation

Sand-packing

Prior to the sand-packing, a leakage test was carried out by using water at a pressure

of 1,200 kPa. After the water leakage test, the glass beads with a grain size of 90−150 μm

(diameter) were used to pack the physical model. Once the cavity of the physical model

was fully packed with glass beads, it was covered with the polycarbonate plate, acrylic

plate and metal frame in sequence. Then the physical model was positioned vertically or

horizontally, and the sands were dried by using the pressurized air for at least 48 h. The

physical model was shaken with an air-actuated vibrator (BV, Vibco, USA) for at least two

hours. Some void space might be formed at the top of the cavity after the dry sands were

shaken and settled downward. Therefore, the physical model needed to be repacked 2 to 3

times in the same way until no void space was formed at the top of the physical model.

Page 138: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

117

Porosity measurement

The imbibition method was used to measure the porosity of the sand-packed physical

model. More specifically, the physical model was vacuumed and then saturated with water

by imbibitions. With the measured volume of the imbibed water and the known volume of

the cavity of the physical model, its porosity can be calculated. The measured porosity for

the experiments was found to be in a range of 34.5% to 37.5%.

Permeability measurement

During the permeability measurement, a digital pressure transducer was used to

record the pressure difference. Distilled water with the density of 1,000 kg/m3 was used as

the working medium. The permeability of the sandpack is determined by using the Darcy’s

law for a steady-state one-phase flow prior to each test. The pressure drop of the distilled

water at the two ends of the physical model was measured and recorded by using a digital

pressure transducer. The permeability measurements were conducted for three times, and

the average permeability of the sand-packed physical model was found to be k = 4.6−6.9

Darcy.

Initial oil saturation

After the permeability measurement, the wet glass beads were dried by using the

pressurized air for at least 48 hours. The heavy oil was injected into the physical model at a

volume flow rate of 0.1−0.25 cm3/h until it was completely saturated with the heavy oil

sample. The initial oil saturation was measured as the ratio of the injected oil volume to the

pore volume of the physical model, which was found to be in the range of Soi =

90.2−94.5%.

Page 139: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

118

5.1.4 Experimental procedure

In this study, eight laboratory tests were performed with three operating schemes

(VAPEX, CSI, and F-VAPEX). The VAPEX and CSI tests served as base tests for the

F-VAPEX tests. The first five tests were conducted with the central well configuration to

evaluate the F-VAPEX process and the last three tests were undertaken with the lateral well

configuration to validate and further assess the F-VAPEX process. Pressure-control

schemes for the VAPEX, CSI, and F-VAPEX tests are shown in Figure 5.3 and described

in the following sub-sections.

VAPEX

The extracting solvent (propane) is continuously injected into the physical model at

Pinj = 800 kPa (Figure 5.3a) and T = 20.2C, which is close to the propane’s saturation

pressure Pdew = 841 kPa at T = 20.2C. Meanwhile, the BPR is properly controlled so that

the pressure inside the physical model is maintained at P = 800 kPa and no oil is

accumulated above the producer.

CSI

Each CSI cycle lasts for 1 h and consists of two periods: (1) Injection period. Propane

is continuously injected into the sand-packed physical model at Pinj = 800 kPa and T =

20.2C for 55 min; (2) Production period. The production pressure Pprod is reduced from

800 to 200 kPa within 5 min (Figure 5.3b).

F-VAPEX.

F-VAPEX is a cyclic process and each cycle continues over 1 h that comprises two

periods:

Page 140: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

119

Time (hh:mm)

00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10

Pres

sure

(kPa

)

0

200

400

600

800

1000

(a)

Time (hh:mm) 00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10

Pres

sure

(kPa

)

0

200

400

600

800

1000

One cycle

Injection Production

(b)

Time (hh:mm) 00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10

Pre

ssur

e (k

Pa)

0

200

400

600

800

1000

One cycle

Stable pressure period

Pressure reduction period

(c)

Figure 5.3 Pressure-control scheme for (a) VAPEX; (b) CSI; and (c) F-VAPEX.

Page 141: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

120

1. Stable pressure period. This period lasts for 55 min, during which the injection

and production pressures are operated in the VAPEX mode: the model pressure is

maintained at Pinj = 800 kPa and the oil is produced continuously.

2. Pressure reduction period. This period lasts for 5 min, during which the solvent

injector is closed and both the BPR and the producer are quickly opened to

induce a sharp blowdown.

Some F-VAPEX tests with longer cycle lengths (2 and 4 h) prolong the stable

pressure period and pressure reduction period proportionally. For instance, Test #5.4 (cycle

length = 2 h) has a 110 min of stable pressure period and a 10 min of pressure reduction

period in each cycle. It is worthwhile to mention that prior to the formal process of the

VAPEX and F-VAPEX tests, the initial communication between the injector and the

producer was established by keeping the injection pressure at a pre-set pressure (800 kPa)

and the production pressure at the atmospheric pressure until a column of continuous gas

bubbles were observed at the producer. In this study, the communication took 30–45 min to

establish and led to an oil production of 1.1–2.3 g for the tests with the central well

configuration and 2.3–3.5 h to result in an oil production of 2.1–3.6 g for the tests with the

lateral well configuration.

5.1.5 Other measurements

Residual water and oil saturations

The residual water and oil saturations at different representative locations inside the

sand-packed physical models were measured after each test. First, the physical model was

opened and sand samples saturated with the residual water and oil were taken and placed

into beakers of 25 ml. Second, the beakers were placed in an oven and heated at 70C for

Page 142: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

121

24 h so that the residual water in the sand sample can be evaporated. The weight difference

before and after the heating was noted as Ww. Finally, the water-free sand sample was

rinsed with toluene to remove the residual oil, and then heated inside the oven to vaporize

the toluene from the sand sample. The weight change before and after the rinsing and

heating was noted as Wo. The final weight of the dried and cleaned sand samples were

noted as Ws. The volumes of the residual oil, residual water, and sand were computed by

dividing Ww, Wo, and Ws by their respective densities w, o, and s. The pore volume of the

initial sand sample, Vp, was calculated by using the volume of the final dried and cleaned

sand sample and the measured porosity. The pore volume, residual water saturation, and

residual oil saturation were determined by using the following equations:

p s 1V V

, (5.1)

wwr

p

VSV

, (5.2)

oor

p

VSV

. (5.3)

5.2 Results and Discussion

5.2.1 Foamy oil flow in F-VAPEX

Foamy oil zone

Figure 5.4 shows the measured injection and production pressure versus time data for

an F-VAPEX test (Test #5.3). A special phenomenon, namely ‘foamy oil flow’, was

observed during the pressure reduction period of Test #5.3, as shown in Figure 5.5a. Three

zones can be identified in Figure 5.5a: a solvent chamber, an untouched heavy oil zone, and

Page 143: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

122

a ‘foamy oil zone’ in between. The boundary of foamy oil zone on the right-hand side of

the model is roughly marked with the white dash lines. It is found that the foamy oil zone

grew throughout the F-VAPEX test, especially during the pressure reduction period.

During the early stage of a pressure reduction period, the foamy oil zone shrank slightly

due to the production of the solvent-diluted heavy oil from it. Afterward, when the pressure

was decreased to a certain level (i.e., bubble-point pressure), a flow front suddenly

emerged from the boundary between the foamy oil zone and untouched heavy oil zone and

moved quickly toward the solvent chamber and the producer. This speculated foamy oil

flow lasted for a short period of time (typically 5–20 s) and resulted in an expanded foamy

oil zone. Figure 5.5b shows the model only 10 s later than that in Figure 5.5a. It can be seen

that the foamy oil zone became much larger on both sides of the model after the expansion.

In addition, the foamy oil zone became darker near the solvent chamber and lighter near the

untouched heavy oil zone, and its boundaries also became clearer in Figure 5.5b. This is

because the foamy oil flow moved solvent-diluted heavy oil closer to the producer and

redistributed the oil saturation inside the foamy oil zone.

Page 144: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

123

Time (hh:mm)

00:00 00:20 00:40 01:00 01:20 01:40

Pre

ssur

e (k

Pa)

0

200

400

600

800

1000Pinj

Pprod

00:30 00:40 00:50 01:00770

780

790

800

810

Figure 5.4 Injection and production pressure data during a typical F-VAPEX cycle.

Page 145: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

124

(a)

(b)

Figure 5.5 Foamy oil zone (a) before and (b) after foamy oil flow during a pressure

reduction period of an F-VAPEX process (Test #5.3).

Solvent chamber

Foamy oil zone

Untouched heavy oil zone

02/11/2013 08:58:14

02/11/2013 08:58:04

Page 146: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

125

Evolution of foamy oil zone

Figure 5.6 shows the foamy oil zone during the early, intermediate, and late stages of

an F-VAPEX process (Test #5.3). It is found that the foamy oil zone was rather small at

the early stage of Test #5.3 (Figure 5.6a). It grew larger and larger with time (Figure 5.6b)

and occupied almost half of the model at the late stage (Figure 5.6c). Moreover, it can be

seen that the solvent chamber had a funnel shape and the foamy oil zone on each side of

the model had irregular shapes, wider at the bottom and thinner at the top. This is because

the solvent-diluted heavy oil was drained downward to the bottom of the foamy oil zone

by gravity, which led to a stronger foamy oil flow and more significant expansion at the

bottom than at the top of the foamy oil zone during the pressure reduction period.

Effects of foamy oil zone

The foamy oil flow during the pressure reduction period of an F-VAPEX test has two

major effects:

1. Mass transfer enhancement. The foamy oil flow redistributed the solvent-diluted

heavy oil inside the model and greatly alleviated the ‘concentration shock’,

giving the solvent more chance to touch the heavy crude oil. In addition, since the

foamy oil zone was a two-phase zone that had a large solvent−oil contact area, the

partially diluted heavy oil could be more easily and completely diluted by solvent

during the stable pressure period.

2. Production enhancement. In addition to the gravity drainage and pressure

gradient [Knorr and Imran, 2012] in conventional VAPEX, F-VAPEX

introduced two more production mechanisms, i.e., solution-gas drive and foamy

oil flow, to enhance the heavy oil recovery.

Page 147: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

126

(a)

(b)

(c)

Figure 5.6 Foamy oil zone during the (a) early, (b) middle, and (c) late stages of an

F-VAPEX test (Test #5.3).

Page 148: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

127

It is worthwhile to mention that the foamy oil flow also helped to estimate the size of

the foamy oil zone. Because the foamy oil zone boundary, especially the boundary near the

untouched heavy oil zone, became distinguishable after the foamy oil flow (Figure 5.6b).

Therefore, the size of the foamy oil zone can be estimated to optimize the operating

conditions of F-VAPEX.

Oil production mechanisms

Similar to VAPEX, F-VAPEX had oil production throughout the process. During the

stable pressure period, the solvent-diluted heavy oil was continuously drained downward

by gravity and intermittently produced by a small pressure gradient around the producer, as

shown by the close-up of the injection and production pressure profiles in the insert of

Figure 5.4. The small pressure gradients can suck out the solvent-diluted heavy oil around

the producer without causing a serious solvent breakthrough. During the pressure reduction

period, the solvent-diluted heavy oil was produced through the pressure gradient,

solution-gas drive, and foamy oil flow that are similar to the puff period of the

conventional CSI process.

5.2.2 F-VAPEX vs. VAPEX/CSI

Oil production

Table 5.2 summarizes the measured test durations t, cumulative oil production data

Qo, and cumulative solvent production data Qg, and the calculated average oil production

rates qo, oil recovery factors (RFs), and solvent−oil ratios (SOR) for the VAPEX, CSI, and

F-VAPEX tests in this study. Figure 5.7 shows the cumulative oil production versus time

data for the VAPEX, CSI, and F-VAPEX tests with the central well configuration. It can be

seen that the F-VAPEX tests (Tests #5.3−4) had higher average oil production rates and the

Page 149: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

128

ultimate oil recovery factors than the VAPEX (Test #5.1) and CSI (Test #5.2) tests. In

addition, Test #5.3 achieved the best performance among all the three F-VAPEX tests

(Tests #5.3−5.5). Figure 5.8 shows the cumulative oil production versus time data for the

VAPEX and F-VAPEX tests with the lateral well configuration. It verifies the superiority

of the F-VAPEX process over the conventional VAPEX and CSI processes in terms of the

average oil production rate. The average oil production rate of VAPEX was enhanced by

F-VAPEX by approximately 50% with the central well configuration and about 115% with

the lateral well configuration, as shown in Figure 5.9. In addition, the average oil

production rate of CSI was slightly enhanced by F-VAPEX with the central well

configuration and significantly improved (over 100%) by F-VAPEX with the lateral well

configuration (Figure 5.10).

Page 150: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

129

Table 5.2 Cumulative heavy oil and solvent production data.

Test no. t (h)

Qo (g)

Qs (dm3)

qo (g/h)

Oil RF (% OOIP)

SOR (g solvent/g oil)

5.1 60 114.8 38.421 1.91 44.6 0.61 5.2 57 153.6 393.13 2.69 60.1 4.70 5.3 60 179.3 345.52 2.99 68.6 3.54 5.4 60 171.5 312.06 2.86 66.5 3.34 5.5 60 156.1 299.23 2.60 60.1 3.52 5.6 60 137.6 51.075 2.29 52.0 0.68 5.7 28 143 311.38 5.11 55.2 4.00 5.8 28 135.2 252.11 4.83 51.0 3.43

Page 151: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

130

Time (h)

0 10 20 30 40 50 60 70

Cum

ulat

ive

oil p

rodu

ctio

n (g

)

0

20

40

60

80

100

120

140

160

180

200VPAEXCSIF-VAPEX, cycle length = 1 hF-VAPEX, cycle length = 2 hF-VAPEX, cycle length = 4 h

Figure 5.7 Cumulative oil production versus time data for the VAPEX, CSI and

F-VAPEX tests with the central well configuration.

CSI Injector/ producer

VAPEX & F-VAPEX

Page 152: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

131

Time (h)

0 10 20 30 40 50 60 70

Cum

ulat

ive

oil p

rodu

ctio

n (g

)

0

20

40

60

80

100

120

140

160

180

200VPAEXF-VAPEX, cycle length = 1 hF-VAPEX, cycle length = 2 h

Figure 5.8 Cumulative oil production versus time data for the CSI and F-VAPEX tests

with the lateral well configuration.

VAPEX & F-VAPEX

Page 153: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

132

Test No.

3 4 5 6 7 8

Oil

prod

uctio

n ra

te e

nhan

emen

t (%

)

0

20

40

60

80

100

120

140

160

180

Figure 5.9 Enhancement of the oil production rate of VAPEX by F-VAPEX with

different well configurations.

VAPEX / F-VAPEX

VAPEX / F-VAPEX

100%F VAPEX VAPEX

VAPEX

q qenhancementq

Page 154: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

133

Test No.

3 4 5 6 7 8

Oil

prod

uctio

n ra

te e

nhan

emen

t (%

)

0

50

100

150

Figure 5.10 Enhancement of the oil production rate of CSI by F-VAPEX with different

well configurations.

F-VAPEX

F-VAPEX

CSI Injector/ producer

100%F VAPEX CSI

CSI

q qenhancementq

Page 155: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

134

Solvent−oil ratio

Figure 5.11 shows the cumulative solvent−oil ratio versus time data for the VAPEX,

CSI, and F-VAPEX tests with the central well configuration. It can be seen that F-VAPEX

with a cycle length of 2 h had a SOR higher than that of VAPEX but lower than that of CSI.

Figure 5.12 shows the cumulative solvent−oil ratio versus time data for the VAPEX and

F-VAPEX tests with the lateral well configuration. It is found that F-VAPEX also had a

higher SOR, which is consistent with the observation in Figure 5.11. Although F-VAPEX

requires much more solvent than VAPEX for the oil production, most of the injected

solvent can be recovered and reused [McMillen, 1985; Butler and Mokrys, 1991; Singhal

et al., 1997]. Nevertheless, the solvent usage is still a major issue for the solvent-based

methods, and the oil production rate and SOR should be optimized in an F-VAPEX

process.

5.2.3 Effect of well configuration

Foamy oil zone

Figures 5.13 shows the foamy oil zone during the early, intermediate, and late stages

of an F-VAPEX test with the lateral well configuration (Test #5.7). The foamy oil zone

above the untouched heavy oil zone in Test #5.7 had a different shape from that in Test

#5.3 (Figure 5.5). It grew mainly in the vertical direction in Test #5.7 and in the horizontal

direction in Test #5.3. Although with different shapes, the foamy oil zones were both

caused by the foamy oil flow during the pressure reduction period.

Page 156: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

135

Time (h)

0 10 20 30 40 50 60 70

Cum

ulat

ive

solv

ent-o

il ra

tio (g

sol

vent

/g o

il)

0

1

2

3

4

5

6VPAEXCSIF-VAPEX, cycle length = 1 hF-VAPEX, cycle length = 2 hF-VAPEX, cycle length = 4 h

Figure 5.11 Cumulative solvent−oil ratio versus time data for the VAPEX, CSI, and

F-VAPEX tests with the central well configuration.

Page 157: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

136

Time (h)

0 10 20 30 40 50 60 70

Cum

ulat

ive

solv

ent-o

il ra

tio (g

sol

vent

/g o

il)

0

1

2

3

4

5

6VPAEXF-VAPEX, cycle length = 1 hF-VAPEX, cycle length = 2 h

Figure 5.12 Cumulative solvent−oil ratio versus time data for the VAPEX and

F-VAPEX tests with the lateral well configuration.

Page 158: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

137

(a)

(b)

(c)

Figure 5.13 Foamy oil zone during the (a) early, (b) middle, and (c) late stages of an

F-VAPEX test with the lateral well configuration (Test #5.7).

Solvent chamber Foamy oil zone

Untouched heavy oil zone

Page 159: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

138

Oil production

Figure 5.14 shows the oil production from the stable pressure period and pressure

reduction period of the even numbered cycles of Test #5.3 (odd numbered cycles are

hidden for the sake of a clear illustration). Apparently, the stable pressure period produced

much more oil than the pressure reduction period in most cycles of Test #5.3. In total, the

former recovered 121.4 g of oil and the latter produced only 58.9 g. This trend agreed with

the oil production data from Test #5.7 (Figure 5.15) as well as the other F-VAPEX tests

(Tests #5.4−5, and #5.8) (Figure 5.16). The smaller contribution from the pressure

reduction period is caused by the solvent dissociation and the resulting oil viscosity

re-increase and mobility loss due to pressure reduction. However, without the pressure

reduction period, F-VAPEX would become a conventional VAPEX and its oil production

rate would be much lower (Table 5.2). Therefore, the pressure reduction period is an

indispensable part of the F-VAPEX process. Because the foamy oil flow during the

pressure reduction period moved the solvent-diluted heavy oil closer to the producer,

which facilitated the oil production during the subsequent stable pressure period.

Page 160: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

139

Cycle number

0 10 20 30 40 50 60

Oil

prod

uctio

n (g

)

0

1

2

3

4

5

6Stable pressure periodPressure reduction period

Figure 5.14 Oil production from the stable pressure period and pressure reduction

period during Test #5.3.

F-VAPEX

Page 161: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

140

Cycle number

0 5 10 15 20 25

Oil

prod

uctio

n (g

)

0

2

4

6

8

10

12

14Stable pressure periodPressure reduction period

Figure 5.15 Oil production from the stable pressure period and pressure reduction

period during Test #5.7.

F-VAPEX

Page 162: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

141

Test No.

3 4 5 6 7 8

Oil

prod

uctio

n (g

)

0

20

40

60

80

100

120

140

160

180Stable pressure periodPressure reduction period

Test no. Cycle length (h) 3 1 4 2 5 4 7 1 8 2

Figure 5.16 Total oil production from the stable pressure period and pressure reduction

period during the F-VAPEX tests.

Page 163: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

142

Comparison of Tests #5.3 (Figures 5.14) and #5.7 (Figures 5.15) suggests: (1) Test

#5.7 has a higher oil production rate than Test #5.3 in their early stages; (2) The oil

production rate declines with time in both Tests #5.3 and #5.7 and the decrease in Test #5.7

is much faster than that in Test #5.3; (3) Test #5.7 has a lower oil production rate than Test

#5.3 in their late stages. In Test #5.7, oil was produced mainly by pressure gradients during

the stable pressure period rather than gravity drainage due to the small inclination angle. In

the early stage, Test #5.7 had a larger solvent−oil contact area and more solvent-diluted

heavy oil than Test #5.3. Therefore, the oil could be more effectively produced in Test #5.7

than in Test #5.3. In the middle and late stages, the foamy oil zone could not reach to the

upper portion of the model and a high gas saturation band (solvent chamber) formed above

it in Test #5.7. Consequently, the solvent easily broke through from the solvent chamber

during the stable pressure period, which suppressed the oil production. During the pressure

reduction period, the foamy oil mainly flows vertically upward rather than horizontally

toward the producer, which also hindered the oil production. In contrast, Test #5.3 always

had a considerable inclination angle and the foamy oil flow constantly moved the oil

toward the producer, which led to a more stable oil production throughout the test.

Cycle length

Figure 5.17 shows the total solvent production data of the F-VAPEX tests with

different cycle lengths. Obviously, more solvent was required in the pressure reduction

period than that in the stable pressure period in the F-VAPEX tests with a cycle length of 1

h (Tests #5.3 and #5.7), which is contrary to the F-VAPEX tests with longer cycle lengths

(Tests #5.4−5, and #5.8). This suggests that a longer cycle length required a less amount of

solvent meanwhile had a similar oil production rate (Table 5.2). However, this does not

Page 164: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

143

mean a longer cycle length would necessarily lead to a higher oil production rate, as shown

by the results of Tests #5.4 (cycle length: 2 h) and #5.5 (cycle length: 2 h). Test #5.5 saved

4.17% of the total solvent usage but lost 9.10% of the total oil production in comparison

with Test #5.4. Because the increase of cycle length decreased the cycle number, which

further affected the foamy oil flow to mobilize the oil toward the producer. Therefore,

during the stable pressure period, solvent broke through easily once the oil around the

producer was produced, which increased the solvent production. This interpreted why Test

#5.5 used more solvent than Test #5.4 during their stable pressure periods (Figure 5.17) but

produced less oil than Test #5.4 (Table 5.2).

Page 165: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

144

Test No.

3 4 5 6 7 8

Solv

ent p

rodu

ctio

n (d

m3 )

0

50

100

150

200

250

300Stable pressure periodPressure reduction period

Test no. Cycle length (h) 3 1 4 2 5 4 7 1 8 2

Figure 5.17 Total solvent production data in the stable pressure period and the pressure

reduction period of the F-VAPEX tests.

Page 166: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

145

5.2.4 Residual oil saturation

The residual oil and water saturation distributions inside the model were measured at

the end of each test. The residual water saturation was found to be in the range of Swr =

2.3−5.4%. Figure 5.16 shows the sandpack models at the end of Tests #5.1, #5.3, and #5.7.

The rough front surface of the sandpack models was formed because the viscous heavy oil

in the untouched heavy oil zone stuck to the cover plate. Figures 5.18a and 5.18b compare

the residual oil saturation distributions at several representative locations in the models of

Tests #5.1 and #5.3. It can be seen that the residual oil saturation at Location #1 (solvent

chamber) of both tests are about 10%. The residual oil saturation at Location #3 (untouched

heavy oil zone) of both tests are about 90%, which are close to their respective initial oil

saturations (92.7% for Test #5.1 and 93.5% for Test #5.3). The residual oil saturations at

Locations #2 and #4−5 (foamy oil zone) of Test #5.3 was found to be Sor = 37.6−49.9%,

which are much lower than those (Sor = 78.5−85.5%) at the same locations of Test #5.1. In

addition, the residual oil saturation in the foamy oil zone of Test #5.7 is found to be Sor =

51.2% and 31.3%, which is consistent with the measured residual oil saturation in the

foamy oil zone of Test #5.3.

Figure 5.19 shows the cross-sectional views of the Test #5.4 and #5.6, respectively. It

can be seen that oil saturation distributions are quite uniform in the thickness direction.

The asphaltene precipitation is observed in Test #5.6, which is shown as the multiple rigid

dark strips mingled with soft brown strips in Figure 5.19b. This is consistent with the

previous study [Das 1998]. The asphaltene precipitation in other F-VAPEX tests are not as

pronounced as that in Test #5.6.

Page 167: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

146

(a)

(b)

(c)

Figure 5.18 Residual oil saturation at the end of (a) Test #5.1; (b) Test #5.3; and (c)

Test #5.7.

38.5%

23.1%

11.4%

78.5%

85.5% 91.2%

85.4%

63.3%

16.1%

12.5%

47.3%

3.6% 40.6%

48.9% 91.1%

77.7%

5.7%

12.2%

67.1%

78.8% 89.2% 41.2%

17.3%

5

4

49.9% 5

37.6% 4

2

1

3

86.4% 3

2 47.3%

16.3% 1

1 51.2% 2 37.3%

Injector

Producer

Injector

Producer

Injector

Producer

Page 168: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

147

(a)

(b)

Figure 5.19 Cross-sectional views of the post-test sandpack of (a) Test #5.4 and (b)

Test #5.6.

Page 169: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

148

5.3 Chapter Summary

This chapter presents a new solvent-based process, F-VAPEX, to enhance heavy oil

recovery of conventional VAPEX/CSI. F-VAPEX benefits the technical advantages of

both VAPEX and CSI, such as continuous production and strong driving force.

In comparison with VAPEX, F-VAPEX introduces more production mechanisms,

including gravity drainage and intermittent sucking during the stable pressure period and

the solution-gas drive and foamy oil flow during the pressure reduction period. Foamy oil

flow moves the solvent-diluted heavy oil closer to the producer, which not only enhances

the oil production during the pressure reduction period but also facilitates the oil recovery

during the subsequent stable pressure period. The average oil production rate of VAPEX is

increased by 1.15 times with F-VAPEX. In comparison with CSI, F-VAPEX has a higher

average oil production rate and a lower solvent−oil ratio. The oil saturation inside the

foamy oil zone is measured to be in the range of 35−50%.

F-VAPEX with the lateral well configuration produces oil faster in the early stage but

slower in the late stage than that with the central well configuration. A longer cycle length

can lower the solvent gas usage but reduce the oil production. The cycle length has to be

optimized for an F-VAPEX process.

Page 170: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

149

CHAPTER 6 GASFLOODING-ASSISTED CYCLIC SOLVENT INJECTION (GA-CSI)

The CSI process takes advantage of solution-gas drive and foamy oil flow (as shown

in Chapter 5) for the oil production. However, CSI process suffers from the solvent

liberation during its production period. This leads the partially diluted heavy oil to regain

its high viscosity and eventually lose its mobility. How to recover the partially diluted

heavy oil becomes a key challenge for a CSI process. This chapter first experimentally

analyzes the conventional CSI process with different well configurations. It is found that

the ‘back-and-forth movement’ of some partially diluted heavy oil in the solvent chamber

limits the oil productivity of the conventional CSI process. On the basis of this observation,

a new process, namely gasflooding-assisted cyclic solvent injection (GA-CSI), is proposed

to enhance the performance of the CSI process. In the GA-CSI process, two wells are used

respectively as the solvent injector and oil producer, and a gasflooding slug is applied after

the pressure reduction process to produce the partially diluted foamy oil left in the solvent

chamber. The experimental results show that GA-CSI can significantly enhance the CSI

performance in terms of both the average oil production rate and the ultimate oil recovery

factor.

6.1 Experimental

6.1.1 Materials

Heavy oil sample and solvent (propane) material are the same as those specified in

the previous chapter: Crude heavy oil has a viscosity of 5,875 mPas and a density of 975

kg/m3. Propane with a purity of 99.5 mol.% is used as the extracting solvent.

Page 171: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

150

6.1.2 Experimental set-up

Figure 6.1a shows a schematic diagram of the experimental set-up, which was

comprised of four major operation units: a solvent injection unit, a physical model, a fluids

production unit, and a data acquisition system. The solvent injection unit, fluids production

unit, and data acquisition system are quite similar to those specified in Chapter #5. The

only difference is that a digital gas flowmeter (XFM17, Aalborg, USA) was installed in the

solvent injection unit to record the solvent injection rate and the cumulative solvent

injection during the tests, especially during the solvent injection period.

The major characteristic of the experimental set-up in this chapter is that two types of

physical models were used in this chapter to evaluate the performance of the GA-CSI

process: three cylindrical models and a 2D rectangular model. The cylindrical models were

steel pipes with the constant inner diameter (ID) of 3.8 cm and different lengths of 34, 63,

and 93 cm, respectively. The injector and producer were installed in the center of the caps

at two ends. The 2D rectangular model was the same as described in the previous chapter.

It is worthwhile to note that in this chapter, the rectangular physical model was placed

horizontally rather than vertically. The first five tests were conducted with the cylindrical

physical models and the last two tests were undertaken with the rectangular physical model.

Two types of well configurations were adopted during the tests: (1) A single well is

alternately used as the injector or producer (one-well configuration, see Figure 6.1c) and (2)

The injector is horizontally apart from the producer (two-well configuration, see Figure

6.1a). In the experiments, two CSI tests were carried out with the one-well configuration,

while one CSI test and four GA-CSI tests were performed with the two-well configuration.

Page 172: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

151

(a)

(b)

(c)

Figure 6.1 (a) Schematic diagram of the experimental set-up with a cylindrical model

for GA-CSI tests and a CSI test; (b) Dimensions of the rectangular sand-packed model;

and (c) Schematic diagram of the physical model for a CSI test.

Physical model Injector/ Producer

To Solvent injector

To Production unit

Sand pack

P P

H = 2 cm L = 40 cm

W = 10 cm

Solvent injection unit

Data acquisition unit

Physical model

Fluids production unit

Pressure transducer Computer

Producer Injector Sand pack

Gas regulator

Propane cylinder

Gas flow meter

Scale Surge flask Flow meter

Back-pressure regulator Digital pressure gauge

Oil collector

Steel tubing Plastic flexible hose Data ware

Page 173: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

152

6.1.3 Experimental preparation

Experimental preparations, such as sand packing, porosity, permeability, and initial

oil saturation measurements, are the same as those described in Chapter 5. Table 6.1

summarizes the detailed physical properties of the sand-packed physical models.

6.1.4 Experimental procedure

In this chapter, seven laboratory tests were performed with two operating schemes

(CSI and GA-CSI). The first five tests were conducted with the cylindrical physical models

to analyze the CSI and GA-CSI processes. The last two tests were undertaken with the

rectangular physical model to verify the effectiveness of the GA-CSI process. The

pressure-control processes for the CSI and GA-CSI tests are schematically shown in Figure

6.2 and described in the following section.

CSI

Each CSI cycle lasted for 1 h and consisted of two periods (Figure 2a): (1) Injection

period. Propane is continuously injected into the sand-packed physical model at Pinj = 800

kPa and T = 20.2C for 50 min; (2) Production period. The production pressure Pprod is

reduced from 800 to 200 kPa within 10 min.

GA-CSI

Similar to CSI, each GA-CSI cycle lasted for 1 h and also consisted of two periods: a

44−48 min of constant-pressure injection period (Pinj = 800 kPa) and a 12−16 min of

production period that has three stages:

1. Blowdown: The solvent injection valve was closed and the oil production valve

was opened so as to decrease the Pprod from 800 to approximately 200 kPa within

3−5 min;

Page 174: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

153

Table 6.1 Physical properties of the sand-packed physical model and experimental

conditions for CSI and GA-CSI tests.

Test No. Well Configuration Production

Scheme PV (cc)

Soi (%)

k (D)

(%)

6.1 CSI, Fig. 6.2a 140 94.2 5.8 36.5 6.2 CSI, Fig. 6.2a 138 94.9 4.9 35.9 6.3 GA-CSI, Fig. 6.2b 139 94.2 5.3 36.2 6.4 GA-CSI, Fig. 6.2b 271 96.3 4.8 36.4 6.5 GA-CSI, Fig. 6.2b 393 96.1 4.6 35.7 6.6 CSI, Fig. 6.2a 281 93.9 4.8 35.1 6.7 GA-CSI, Fig. 6.2b 284 93.3 5.1 35.5 6.8 PP-CSI, Fig. 6.16 283 93.6 5.2 35.9

L = 34 cm, D = 3.8 cm

L = 34 cm, D = 3.8 cm

L = 93 cm, D = 3.8 cm

L = 34 cm, D = 3.8 cm L = 63 cm, D = 3.8 cm

40 10 2 cm3

40 10 2 cm3

40 10 2 cm3

Page 175: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

154

Time (hh:mm)

00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10 01:20

Pres

sure

(kPa

)

0

200

400

600

800

1000

Pprod

Injection Production

ONE CYCLE

(a)

Time (hh:mm)

00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10 01:20

Pres

sure

(kPa

)

0

200

400

600

800

1000Pinj

Pprod

Injection Production

ONE CYCLE

(b)

Figure 6.2 Pressure-control scheme of (a) GA-CSI and (b) CSI.

Page 176: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

155

2. Reinjection: The injection valve is opened and the production valve is closed.

Propane is re-injected into the physical model to restore the model pressure to the

previous level; and

3. Gasflooding: The producer is reopened and the production pressure is carefully

adjusted to maintain a proper pressure difference between the injector and the

producer. The flooding stage continues over 6, 8, and 10 min for the 34, 63, and

93 cm cylindrical models, respectively. For rectangular model, it lasts for 8 min.

It is worthwhile to mention that prior to the cyclic process of the CSI and GA-CSI

tests with the two-well configuration, an initial communication between the injector and

the producer was established by keeping the injection pressure at a pre-set pressure (800

kPa) and the production pressure at the atmospheric pressure until a column of continuous

gas bubbles were observed at the producer.

6.2 Results and Discussion

Table 6.2 summarizes the measured test durations t, cumulative oil production data

Qo, and cumulative solvent production data Qg, and the calculated average oil production

rates qo, oil recovery factors (RFs), and solvent−oil ratios (SOR) for the CSI and GA-CSI

tests in this study. Figure 6.3 shows the cumulative oil production and solvent−oil ratio

versus time data for two CSI tests and one GA-CSI with the same cylindrical physical

model but different well configurations. It can be seen from this figure that the

performance of the two-well CSI test (Test #6.2) is much better than that of the one-well

CSI test (Test #6.1), whereas the GA-CSI test (Test #6.3) performs the best among the

three tests. Apparently, the performance of a cyclic solvent process is affected more by the

operating scheme than by the well configuration. The enhancement of oil production rate

Page 177: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

156

Table 6.2 Cumulative oil and solvent production data of eight CSI and GA-CSI tests.

Test No. t (h)

Qo (g)

Qs (dm3)

qo (g/h)

Oil RF (% OOIP)

SOR (g solvent/g oil)

6.1 29 64.4 150.7 2.2 48.8 4.3 6.2 22 75.5 154.2 3.4 57.6 3.8 6.3 12 96.9 158.6 8.1 74.0 3.0 6.4 14 176.3 276.7 12.6 67.5 2.9 6.5 16 258.4 362 16.2 68.4 2.6 6.6 47 126.1 283.1 2.7 47.8 4.1 6.7 15 169.9 325.1 11.2 63.6 3.5 6.8 13 178.2 359.2 13.5 67.2 3.6

Page 178: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

157

Time (h)

0 5 10 15 20 25 30

Oil

reco

very

fact

or (%

)

0

20

40

60

80

(a)

Time (h)

0 5 10 15 20 25 30

Cum

lativ

e so

lven

t-oil

ratio

(g s

olve

nt/g

oil)

0

1

2

3

4

5

(b)

Figure 6.3 (a) Cumulative oil production; and (b) SOR of Tests #6.1−3.

Test #6.3

Test #6.2 (CSI)

Test #6.3

Test #6.2 (CSI)

Test #6.1

Test #6.1

Page 179: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

158

due to the operating-scheme change from Test #6.2 to Test #6.3 (qo = 4.7 g/h) is much

larger than that due to the well-configuration change from Test #6.1 to Test #6.2 (qo = 1.2

g/h). The detailed effects of these factors on the performance of a cyclic solvent process

will be analyzed in the following sections.

6.2.1 Well configuration

Tests #6.1−2 have similar physical model properties and pressure-control schemes,

except for the well configuration. Test #6.1 used a single well alternately and cyclically as

the injector or producer, whereas Test #6.2 used two wells as the injector and the producer,

respectively. The well placements resulted in a significant difference on their performance.

Test #6.1 achieved an average oil production rate of 2.2 g/h and Test #6.2 obtained 3.4 g/h.

The reason for the lower average oil production rate in Test #6.1 is the ‘back-and-forth

movement’ of the solvent-diluted heavy oil in the solvent chamber, which is schematically

illustrated in Figure 6.4 and explained below.

Similar to VAPEX, a sand-packed model during a cyclic solvent process also has

three zones: a solvent chamber, an untouched heavy oil zone, and a foamy oil zone in

between. During the solvent injection period, propane is injected into the model and

dissolved into the partially diluted heavy oil inside the foamy oil zone and the dead oil at

the boundary between the foamy oil zone and the untouched heavy oil zone, as shown in

Figure 6.4a. During the production period, solvent in the solvent chamber is first released

and the solvent-diluted heavy oil starts to flow to the producer. Meanwhile, the solvent

dissolved into the heavy oil begins to nucleate into extremely small bubbles due to pressure

reduction [Smith, 1988] and most of these bubbles will keep entrained in the oil and move

toward the producer, causing the so-called foamy oil flow [Sarma and Maini, 1992] (Figure

Page 180: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

159

6.4b). Afterward, the gas bubbles grow larger and larger and finally disengage from the oil

phase to form a continuous gas phase when the pressure is below the ‘pseudo-bubble-point

pressure’ [Kraus et al., 1993]. As a result, the solution gas is quickly released and the

partially diluted heavy oil regains a high viscosity and gradually loses its mobility, and

some foamy oil remains in the model at the end of the production period, as shown in

Figure 6.4c. During the solvent injection period of the subsequent cycle, the injected

solvent re-dissolves into the partially diluted heavy oil in the foamy oil zone and pushes the

oil back to the untouched heavy oil zone (Figure 6.4d). This ‘back-and-forth movement’ of

the foamy oil during the oil production period of one cycle and the solvent injection period

of the next cycle would hinders the oil production, and its influence is expected to become

more and more serious as the solvent chamber grows longer and longer.

This ‘back-and-forth movement’ hypothesis was validated by the digital photographs

of Test #6.6 with the rectangular model. Figure 6.5a shows a clear flowing front in the early

stage of the oil production period of a cycle of Test #6.6. Several foamy oil flowing fronts

can be seen in Figure 6.5b and they became almost immobile at the end of the production

period when Pprod ≈ 200 kPa. During the solvent injection period of the subsequent cycle,

the partially diluted heavy oil in the foamy oil zone was re-diluted and pushed backward by

the injected solvent, and a thick backward flow band of the oil moving away from the

injector was observed (Figure 6.5c).

In contrast, with the two-well lateral well configuration in Tests #6.2−5 and #6.7, the

‘back-and-forth movement’ of the foamy oil did not exist since the oil always flew in one

direction from the injector to the producer.

Page 181: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

160

(a) (b)

(a)

(c) (d)

Figure 6.4 ‘Back-and-forth movement’ of the solvent-diluted heavy oil in a CSI test:

(a) Solvent dissolution into oil during the injection period of a cycle; (b) Diluted oil

flowing to the producer during the production period; (c) Some diluted oil remaining in

the solvent chamber at the end of the production period; and (d) Diluted oil flowing back

during the solvent injection period of the next cycle.

Injector Producer

Injector Producer

Solvent chamber

Untouched heavy oil Diluted oil

Page 182: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

161

(a)

(b)

(c)

Figure 6.5 ‘Back-and-forth movement’ of the solvent-diluted heavy oil during a cycle

of the CSI test (Cycle #40 of Test #6): (a) Oil flowing to the producer at the early stage of

the production period; (b) Oil remaining in the solvent chamber at the end of the

production period; and (c) Oil flowing back during the solvent injection period of the

next cycle (Cycle #41).

Flow

fron

t Fl

ow fr

ont

Bac

kflo

w

Producer

Producer

Injector

Flow

fron

t

Flow

fron

t

Page 183: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

162

6.2.2 Operating scheme (CSI vs. GA-CSI)

The only difference in the pressure-control scheme between Tests #6.2 and #6.3 is

that Test #6.3 has reinjection and flooding processes after the pressure reduction process

during the oil production period of each cycle. However, this small change in the operating

scheme made a large difference in their performance. The average production rate of Test

#6.3 is 2.38 times of that of Test #6.2. This is because although the oil viscosity was

re-increased to some extent at the end of the oil production period, there was still a large

amount of solvent dissolved into the oil. In addition, the oil was relatively uniformly

distributed in the foamy oil zone. Therefore, during the reinjection and flooding processes,

the partially diluted oil in the foamy oil zone near the injector can be quickly diluted by the

solvent and pushed toward the producer to form a flooding front, which was served as a

‘buffer zone’ to greatly control the mobility ratio between the displacing solvent and the

displaced oil.

Figure 6.6 presents a solvent flooding process during a GA-CSI test (Test #6.7). The

brown area in Figure 6.6a shows the foamy oil zone at the end of a blowdown stage. The

white area in Figure 6.6b indicates the swept zone with low residual oil saturation.

Moreover, the advancing front was rather uniform, which was probably attributed to the

‘buffer zone’. At the end of the flooding stage, the advancing front was separated into

several large fingers that resulted in a reduced sweeping efficiency as well as a decreased

oil production rate (Figure 6.6c).

Page 184: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

163

(a)

(b)

(c)

Figure 6.6 Gasflooding process during a GA-CSI test (Test #6.7). (a) End of the

blowdown stage; (b) Early gasflooding stage; and (c) Late gasflooding stage.

Producer Injector

Producer Injector

Producer Injector

Page 185: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

164

6.2.3 GA-CSI

Figure 6.7 shows the injection pressure, production pressure, and solvent injection

rate during a typical cycle of Test #6.3. It is worthwhile to mention that during the oil

production period, the BPR was adjusted to the minimum level so as to induce a larger

pressure gradient and a higher pressure reduction rate, which would result in more bubbles

and greater foamy oil stability [Handy, 1958; Maini et al., 1996; Sheng, 1997]. Figure 6.8

depicts the cumulative oil production, oil production rate, and the corresponding SOR

during the blowdown, reinjection, and flooding stages of the oil production period of a

representative cycle of Test #6.3.

It is found that during the blowdown stage, the oil production rate decreased sharply,

while the SOR increased quickly. During the flooding stage, the cumulative oil production

curve had an ‘S’ shape while the SOR fluctuated around 1.9 g solvent/g oil. Figure 6.8 also

shows that the oil production during the flooding stage was significantly higher than that

during the blowdown stage, which was a general trend in Test #6.3 and the other GA-CSI

tests. Figure 6.9a confirms this trend and also shows that the oil production from the

flooding stage of each cycle first increased and then gradually decreased, while the oil

production from the blowdown stage was 3 g/cycle. The total oil production from the

flooding stages of Test #6.3 was 61.1 g, which was 1.71 times higher than that from the

blowdown stages (35.8 g). Figure 6.9b shows that the solvent gas production during the

blowdown and flooding stages increased steadily with the cycle number throughout Test

#6.3. Figure 6.10 compares the cumulative oil production data due to pressure reduction

in Tests #6.1−3, indicating that the oil production trend due to pressure reduction in the

GA-CSI test was similar to those in the CSI tests.

Page 186: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

165

Time (hh:mm)

00:00 00:20 00:40 01:00 01:20

Pres

sure

(kPa

)

0

200

400

600

800

1000

Solv

ent i

njec

tion

rate

(cc/

min

)

0

200

400

600

800

1000

1200

1400

Pinj

Pprod

qs,inj

Injection

ONE CYCLE

1 2 3

1: Blowdown2: Reinjection3: Gasflooding

Figure 6.7 Injection and production pressures and the solvent injection rate during a

typical cycle (Cycle #4) of a GA-CSI test (Test #6.3).

Page 187: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

166

Time (min)

0 1 2 3 4 5 6 7 8 9 10 11 12

Cum

lativ

e oi

l pro

duct

ion

(g)

0

2

4

6

8

10

12

14

Oil

prod

uctio

n ra

te (g

/min

)

0

1

2

3

4

5

6

Solv

ent-o

il ra

tio (g

sol

vent

/g o

il)

0.0

0.5

1.0

1.5

2.0

2.5

Blowdown Reinjection Gasflooding

Figure 6.8 Cumulative oil production, oil production rate, and solvent–oil ratio during

a typical cycle (Cycle #4) of a GA-CSI test (Test #6.3).

Page 188: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

167

Time (h)

0 1 2 3 4 5 6 7 8 9 10 11 12 13

Oil

prod

uctio

n (g

)

0

2

4

6

8

10

BlowdownGasflooding

(a)

Time (h)

0 1 2 3 4 5 6 7 8 9 10 11 12 13

Solv

ent p

rodu

ctio

n (s

c,L)

0

2

4

6

8

10

12

14 BlowdownGasflooding

(b)

Figure 6.9 (a) Heavy oil production; and (b) Solvent gas production during the

blowdown and gasflooding slugs of the production period of a GA-CSI test (Test #6.3).

Page 189: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

168

Time (h)

0 5 10 15 20 25 30

Cum

lativ

e oi

l pro

duct

ion

(g)

0

20

40

60

80

Figure 6.10 Cumulative oil productions of Tests #6.3 (blowdown slugs only), and #6.1

and #6.2.

Test #6.3

Test #6.2 (CSI)

Test #6.1

Page 190: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

169

6.2.4 Solvent injection rate

Figure 6.11 shows the solvent injection rate during the early (Cycle #2), middle

(Cycle #6), and late (Cycle #11) stages of Test #6.3. All the three curves declined and

reached to a small value within 45 min, indicating that the dissolution of propane into the

heavy oil became rather slow after 45 min of constant pressure injection. This justifies the

selection of one hour as the full cycle length for the CSI and GA-CSI tests in this study. In

addition, it can be seen that less solvent was injected and dissolved into the heavy oil

during the early and late stages than during the middle stage. The lower solvent injection

rate was due to the small solvent chamber size and limited contact area at the early stage

and the high solvent saturation in the heavy oil at the late stage. The higher solvent

injection rate at the middle stage was probably because of the more developed solvent

chamber and the relatively lower solvent saturation. Assuming that the injection period of

each cycle ends when the solvent injection rate is below certain value, the cycle length

must be a function of time and change with the solvent chamber size and solvent saturation

throughout a test. Variable cycle lengths for CSI/GA-CSI need to be studied in future.

6.2.5 GA-CSI with cylindrical models

Figure 6.12 shows the oil RFs and SORs of three GA-CSI tests with three cylindrical

physical models of different lengths. It is found that the three oil RF curves all have an ‘S’

shape and their final values are close to each other and decrease slightly with the increase

of the model length (Figure 6.12a). In addition, longer cylindrical models result in lower

ultimate SOR values in comparison with shorter models (Figure 6.12b). This is because

longer models had smaller pressure gradients for the solvent gas displacement during the

flooding stage. On one hand, a smaller driving force would lead to a lower oil production

Page 191: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

170

Time (min)

00 10 20 30 40

Sol

vent

inje

ctio

n ra

te (s

cm3 /m

in)

0

100

200

300

400

500

600

Figure 6.11 Solvent injection rate at early, middle, and late stages of a GA-CSI test

(Test #6.3).

Test #6.3

Page 192: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

171

Time (h)

0 5 10 15 20

Oil

reco

very

fact

or (%

)

0

20

40

60

80

100

(a)

Time (h)

0 5 10 15 20

Cum

ulat

ive

solv

ent-o

il ra

tio (g

sol

vent

/g o

il)

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

(b)

Figure 6.12 (a) Recovery factor; and (b) Solvent−oil ratio of the GA-CSI tests with

cylindrical models of different lengths.

Test #6.5

Test #6.4

Test #6.3

Test #6.5

Test #6.4 Test #6.3

Page 193: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

172

for longer cylindrical models. On the other hand, a smaller driving force may result in

neither an earlier gas breakthrough nor a larger solvent usage. In short, the pressure

gradient is an important factor for the solvent flooding process and further study is needed

to determine its optimum value for a GA-CSI process.

6.2.6 GA-CSI with rectangular model

Figure 6.13 shows the oil recovery curves for a CSI test and a GA-CSI test with the

rectangular model. It is obvious that Test #6.7 performed much better than Test #6.6. The

average oil production rate of Test #6.7 was 4.48 times of that of Test #6.6, which validated

the superiority and effectiveness of the GA-CSI process over the conventional one-well

CSI process. Comparison of Figure 6.3a and Figure 6.13 shows that the enhancement on

the average oil production rate of the conventional one-well CSI process by the GA-CSI

process with the short cylindrical model (Test #6.3 vs. Test #6.1) is consistent with that

with the rectangular model (Test #6.7 vs. Test #6.6).

Comparison of Tests #6.4 and #6.7 shows that with the same operating scheme and

similar permeabilities and PVs, both tests achieved similar oil production rates (12.6 g/h

for Test #6.4 and 12.1 g/h for Test #6.7) and ultimate oil RF values (67.5% of the OOIP for

Test #6.4 and 64.1% of the OOIP for Test #6.7).

6.2.7 Residual oil saturation

In this study, the residual water and oil saturations at different representative

locations were measured by analyzing the sand samples taken at the end of the CSI and

GA-CSI tests. The residual water saturation was found to be in the range of 26% for all

the measurements and this study focuses on the distributions of the residual oil saturation.

Figure 6.14 shows the digital photographs of the cross sections of the short cylindrical

Page 194: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

173

Time (h)

0 10 20 30 40 50

Oil

reco

very

fact

or (%

)

0

10

20

30

40

50

60

70

Figure 6.13 Oil recovery factor of GA-CSI and CSI tests with the rectangular physical

model.

Test #6.7

Test #6.6

Page 195: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

174

(a)

(b)

Figure 6.14 Residual oil saturation of (a) CSI (Test #6.2); and (b) GA-CSI tests (Test

#6.3).

29.8%

l = 1 cm (Injection/production side)

l = 8 cm

54.1%

l = 24 cm

65.2%

l = 32 cm

72.2%

l = 1 cm (Injection side)

4.1%

14.8%

l = 8 cm

6.9%

l = 32 cm (Production side)

37.3%

l = 24 cm

50.3%

9.8%

Page 196: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

175

sand-packed physical model (L = 34 cm) at the end of Tests #6.1 and #6.3, respectively.

The measured residual oil saturations were in an excellent agreement with the sand

samples’ colours. The whiter the sand samples, the lower the corresponding residual oil

saturation would be. It is found that the colour of the cross sections of Test #6.1 was quite

uniform (Figure 6.14a), which is because: (1) The foamy oil flow uniformly redistributed

the oil during the production period of the CSI process; (2) The effect of gravity force

was negligible. In contrast, it is found that the residual oil saturation in the upper part of

the model of Test #6.3 was much lower than that in the lower part (Figure 6.14b). This is

due to the gravity overriding, which made the solvent-diluted heavy oil move downward

during the test so that the obtained sand sample was lighter in the upper part and darker in

the lower part.

Figure 6.15 shows the front of the rectangular sand-packed physical model after

sampling at the ends of Tests #6.6 and #6.7. Obviously, the model color of Test #6.6 was

much darker than that of Test #6.7, which was consistent with fact that the oil RF of Test

#6.6 was much lower than that of Test #6.7. It is worthwhile to emphasize that the residual

oil saturation was lower at the two ends but higher in the middle part of Test #6.6 (Figure

6.15a), which is due to the aforementioned ‘forth-and-back movement’ of the foamy oil.

The residual oil saturation of Test #6.7 declined from the left-hand side to the right-hand

side (Figure 6.15b). Precipitated asphaltenes were observed as the gray patches on the

right-hand side of the model for Test #6.7. However, It seems that asphaltene precipitation

did neither affect the solvent injection nor the oil production throughout the test.

Page 197: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

176

(a)

(b)

Figure 6.15 Residual oil saturation of (a) CSI (Test #6.6); and (b) GA-CSI (Test #6.7).

45.7%

34.7%

54.1%

43.4%

60.1%

55.7%

46.7%

60.4%

52.3%

54.2%

54.2%

51.9%

63.9%

26.4%

32.6%

34.6%

23.1%

36.3%

39.7%

18.8%

27.8%

25.0%

16.1%

3.1%

16.4%

20.8%

Injector/Producer

Injector Producer

Sampling hole

Page 198: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

177

6.3 Variations of GA-CSI

In the GA-CSI process, foamy oil flow and gasflooding are coupled together to

provide a strong driving force for the heavy oil production. This section presents an

extension of the GA-CSI process, pressure-pulsing cyclic solvent injection (PP-CSI).

6.3.1 Pressure control scheme

PP-CSI is a special form of GA-CSI. The physical properties of a PP-CSI test are

listed in Table 6.1 and its operating scheme is showed in Figure 6.16 and described as

follows.

The pressure control scheme of PP-CSI is similar to that of GA-CSI. Solvent

injector and oil producer are placed horizontally apart. Its pressure is cyclically operated

and each cycle has two periods:

Injection period. Propane is continually injected into the sand-packed physical model

at 800 kPa and 20.2C for ~40 min.

Production period. The production period contains several pressure pulses and each

pulse is a three-step process: blowdown, reinjection, and gasflooding. More specially, in

each pulse, first, decrease the model pressure to induce foamy oil flow; Then, build up the

pressure by injecting solvent for a few minutes; Finally, maintain a certain pressure

difference between injector and producer for a period of gasflooding. Afterward, the

pressure pulse is repeated for another pulse until the oil production rate drops below an

economical limit.

Figure 6.17 shows the typical injection and production pressure measured during the

PP-CSI test (Test #6.8). It is worthwhile to mention that the pulse can be applied as many

times as necessary.

Page 199: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

178

Time (hh:mm)

00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10 01:20

Pres

sure

(kPa

)

0

200

400

600

800

1000

Ping

Pprod

Injection Production

ONE CYCLE

1 2 3

Pulse

1: blowdown2: re-injection3: flooding

Figure 6.16 Pressure control scheme of PP-CSI.

Page 200: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

179

Time (hh:mm)

00:00 00:10 00:20 00:30 00:40 00:50 01:00 01:10 01:20 01:30

Pre

ssur

e, k

Pa

300

400

500

600

700

800

900

Pinj

Pprod

Injection Production

Figure 6.17 Injection and production pressures data during a PP-CSI test.

Page 201: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

180

6.3.2 Viscous fingering

Figure 6.18 displays the solvent chamber evolution throughout the PP-CSI test (Test

#6.8). In the early stage (Cycle #2 and #4), it can be seen that viscous fingers are formed

near the injector. The formation of the solvent fingers is due to the high mobility ratio

between the solvent and heavy oil during the flooding process. Viscous fingering reduces

the sweeping efficiency and caused early solvent breakthrough during the immiscible

flooding processes, such as water flooding, chemical flooding. However, in the PP-CSI

process, viscous fingers play a good role in the following ways: first, the finger growth in

the length was not as fast as anticipated due to the foamy oil flow. Meanwhile, its growth in

the width is much better than expected. Second, solvent fingers greatly increased the

solvent–oil contact area, which significantly enhanced the mass-transfer rate.

Figure 6.18c−d shows the solvent chamber at the middle and late stage of the PP-CSI

process. It can be seen that the solvent fingers at the early stage mingle together to form a

big one, and it did not breakthrough at Cycle #8 when 54.6% of the OOIP was recovered.

From the colour of the model, it can be seen that a sweeping efficiency was achieved in the

PP-CSI test. A noticeable solvent chamber connection to the producer occurred in Cycle

#12 when 61.1% of the OOIP was recovered.

Page 202: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

181

(a)

(b)

(c)

(d)

Figure 6.18 Evolution of the solvent chamber throughout a PP-CSI test: (a) Cycle #2;

(b) Cycle #4; (c) Cycle #8; and (d) Cycle #12.

Injector Producer

Injector Producer

Injector Producer

Injector Producer

Page 203: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

182

6.3.3 Oil production

Figure 6.19 compares the cumulative oil production versus time data for the GA-CSI

and PP-CSI tests (Tests #6.7−8). Obviously, the oil was recovered faster in the PP-CSI test

(Test #6.8) than in the GA-CSI test (Test #6.7). The average oil production rate of the

PP-CSI test was qo = 13.5 g/h, which was 20.1% higher than that of the GA-CSI test (qo =

11.2) g/h. Furthermore, the final recovery factor of the PP-CSI test was 3.6% higher than

that of the GA-CSI test.

Figure 6.20 shows the oil production of multiple pulses (Cycles #2, #7, #11, and #12)

in different cycles of the PP-CSI test. It can be seen that the oil production during the

second and third pulses are similar to that in the first one. In addition, the pulse number was

increased from 2 to 3 in the early stage since the oil production increased with the cycle.

However, in the late stage, the pulse number was reduced (from 3 in Cycle #11 to 2 in

Cycle #12), which is because the oil production decreased with the pulse number and the

decline rate increased in the late stage of the test. It is worthwhile to note that in Test #6.8,

the length of each cycle was set as 1 h and the pulse was set as ~8 min for comparison. In

fact, the operational parameters of the PP-CSI process, such as cycle length, pulse number,

and pulse length, need to be optimized.

Page 204: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

183

Time (h)

0 5 10 15 20

Oil

reco

very

fact

or (%

)

0

20

40

60

80

PP-CSI (Test #6.8)GA-CSI (Test #6.7)

Figure 6.19 Comparison of the oil recovery factor of PP-CSI and GA-CSI tests.

Tests #6.76.8

Page 205: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

184

Cycle number

1 2 3 4

Oil

prod

uctio

n (g

)

0

2

4

6

8

10

12Pressure pulse #1Pressure pulse #2Pressure pulse #3

4 7 11 12

Figure 6.20 Oil production from multiple pulses in different cycles of Test #6.8

Page 206: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

185

6.4 Chapter Summary

In this chapter, a new operating scheme, GA-CSI, is designed and evaluated through

a series of laboratory tests. The following conclusions can be drawn:

1. The performance of conventional one-well CSI is hampered by the

‘back-and-forth movement’ of the foamy oil in the solvent chamber.

2. The major difference between conventional CSI and GA-CSI is the

pressure-control scheme. GA-CSI applies a gasflooding slug between the ‘puff’

and ‘huff’ periods of a conventional CSI process.

3. Aside from solution-gas drive, foamy oil flow, and gravity drainage, GA-CSI

introduces a new and stronger production mechanism, gasflooding, into oil

recovery.

4. The gasflooding slug of the GA-CSI process can effectively produce the partially

diluted heavy oil remained in solvent chamber due to solvent liberation and

mobility loss at the end of the pressure reduction process. A good sweeping

efficiency of gasflooding was observed because of the ‘buffer zone’, which

reduces the mobility ratio between the displacing solvent and the displaced oil.

5. GA-CSI can greatly enhance the performance of conventional CSI in terms of

both the average oil production rate and the ultimate recovery factor. The average

oil production rate enhancement of CSI by GA-CSI is 3.64 times with the

cylindrical model and 4.52 times with the rectangular model.

6. As a variation of GA-CSI, PP-CSI produces oil faster but uses more solvent than

GA-CSI. More research work is needed to further explore and optimize the

variations of GA-CSI.

Page 207: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

186

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS

7.1 Conclusions

This thesis conducts some theoretical modeling of the traditional solvent-based EOR

techniques (VAPEX and CSI) and proposes several new solvent-based EOR processes to

improve their performances. The major conclusions are summarized as follows:

Modeling of VAPEX

1. A new mathematical model of VAPEX is established to predict the evolution of

the solvent vapour chamber during its rising and falling phases. This new model

is developed on the basis of the major mechanisms of VAPEX, such as viscosity

reduction and gravity drainage. The transient heavy oil−solvent mass transfer and

the moving boundary condition are considered in this new model.

2. The new VAPEX mathematical model is able to not only estimate the growth of

the solvent chamber, but also to describe the solvent concentration, oil viscosity

and density, diffusion coefficient, and drainage velocity inside the transition zone.

It is found that the upper-part of the transition zone is thinner and moves faster

than the lower part. It is also found that the constant diffusion coefficient is

acceptable only for a short period of time. It underestimates the oil production

rate for a longer time period since it ignores the increase of the diffusion

coefficient during the VAPEX process.

3. In comparison with the numerical simulation, the new theoretical model

demonstrates more sensitivity to the diffusion coefficient and has less numerical

dispersion.

Page 208: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

187

Modeling of the mass transfer in CSI

4. A 1D convection–diffusion mass-transfer model is developed to describe the

heavy oil−solvent mixing process in CSI. It estimates the effect of pressure

gradient on the mass transfer process. Solvent diffusion coefficient and

convection velocity are both considered as functions of solvent concentration in

this new model.

5. Semi-analytical solutions are obtained through a special approximation of the

variable diffusion coefficient and convection velocity.

6. Modeling results qualitatively suggest that a pressure gradient between the

solvent chamber and untouched heavy oil zone provides a pushing force for the

solvent to mix with heavy oil during the solvent injection period of CSI. The

convection plays a larger role than the diffusion during the CSI process,

especially in the early stage of the solvent injection period.

Enhanced VAPEX (F-VAPEX)

7. F-VAPEX is a combined process of VAPEX and CSI. It is essentially a VAPEX

process during which the operating pressure is cyclically reduced and restored.

8. F-VAPEX is superior to the VAPEX in terms of both the average oil production

rate and the ultimate oil recovery factor. F-VAPEX has a higher oil production

rate and a lower solvent−oil ratio in comparison with CSI.

9. Production mechanisms of the F-VAPEX process include the intermittent

sucking and gravity drainage during the stable pressure period and the

solution-gas drive and foamy oil flow during the pressure reduction period.

Page 209: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

188

10. The foamy oil flow during the pressure reduction period moves the

solvent-diluted heavy oil toward the producer to facilitate the oil production in

the subsequent stable pressure period. As a result, the stable pressure period of

F-VAPEX contributes more oil production than the pressure reduction period.

11. F-VAPEX with the lateral well configuration produces oil faster in the early

stage but slower in the late stage than that with the central well configuration.

12. A longer cycle length for F-VAPEX saves gas usage but may reduce the oil

production.

13. The oil saturation inside the foamy oil zone is measured to be in the range of

35−50%.

Enhanced CSI (GA-CSI)

14. The oil productivity of conventional one-well CSI process is largely limited by

the ‘back-and-forth movement’ of the partially diluted foamy oil in the solvent

chamber. In the GA-CSI process, the solvent injector and oil producer are placed

laterally apart, which effectively eliminates the ‘back-and-forth’ movement of

conventional CSI.

15. GA-CSI applies a gasflooding slug between the ‘puff’ and ‘huff’ periods of the

conventional CSI process. Therefore, in addition to solution-gas drive, foamy oil

flow, and gravity drainage, GA-CSI introduces a stronger production mechanism,

gasflooding. The average oil production rate of CSI is enhanced by GA-CSI by

3.64 times with a cylindrical model and 4.52 times with a rectangular model.

16. Gasflooding slug in GA-CSI results in a good sweeping efficiency due to the

Page 210: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

189

‘buffer zone’. The ‘Buffer zone’ is actually a foamy oil band at the flooding front

that effectively controls the mobility ratio between the displacing solvent and the

displaced oil.

7.2 Recommendations

The following recommendations for future work are made on the basis of the

research in this study.

Modeling of the mass transfer process in CSI

In the present convection–diffusion model, the pressure gradient is assumed as a

constant, which is not reasonable for a practical case. Therefore, variable and dynamic

pressure gradients need to be considered in the future mass-transfer model.

Characterization of the foamy oil flow

Foamy oil flow and the resulting foamy oil zone are observed in this study. The

foamy oil flow and the foamy oil zone in the F-VAPEX and CSI needs to be further

characterized. In addition, the mass transfer of solvent into foamy oil also needs to be

investigated.

Parametric study of F-VAPEX and GA-CSI

More laboratory experiments and numerical simulation need to be conducted to

analyze and optimize these new processes in order to maximize their productivities.

Solvent recovery

Solvent retention is a major concern in the solvent-based EOR processes. Without

high solvent recovery, these processes would be economically unviable. How to

effectively recover the retained solvent needs more research efforts.

Page 211: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

190

REFERENCES

Abramowitz, M. and Stegun, I.A. 1970. Handbook of mathematical functions with

formulas, graphs, and mathematical tables. Dover Publications, New York.

Ali, S.M.F. 2003. Heavy oil–evermore mobile. J Pet Sci Eng 37 (1–2): 5–9. doi:

0920-4150/02.

Alkindi, A.S., Al-Wahaibi, Y.M., and Muggeridge, A.H. 2012. Experimental and

numerical investigation of oil-drainage rate during Vapex process. SPE Journal 16

(2): 343–357. doi: 10.2118/141053-PA.

Allen J.C., Gillespie, R.E. and Burnett, D.B. 1984. Superheated solvent method for

recovering viscous petroleum. US Patent No. 4,450,913.

Allen, J.C. 1973. Method for recovering viscous oils by solvent extraction. Canadian

Patent No. 1,008,361.

Allen, J.C. 1974. Gaseous solvent heavy oil recovery. Canadian Patent No. 1,027,851.

Allen, J.C. and Redford, D.A. 1978. Combination solvent-noncondensible gas injection

method for recovering petroleum from viscous petroleum-containing formations

including tar sand deposits. US Patent No. 4,109,720.

Allen, J.C., Woodward, C.D., and Brown, A. 1976. Multiple solvent heavy oil recovery

method. US Patent No. 3,954,141.

Al-niami, A.N.S. and Rushton, K.R. 1978. Radial dispersion to an abstract well. Journal of

Hydrology 39 (3−4): 287−300. doi: 10.1016/0022-1694(78)90006-9.

Page 212: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

191

Alshmakhy, A.B. and Maini, B.B. 2012. Effects of Gravity, Foaminess, and Pressure

Drawdown on Primary-Depletion Recovery Factor in Heavy-Oil Systems. J Can Pet

Technol 51 (6): 449–456. doi: 10.2118/163067-PA.

Alshmakhy, A.B. and Maini, B.B. 2009. Effect of Foaminess on the Performance of

Solution Gas Drive in Heavy Oil Reservoirs. J Can Pet Technol 48 (3): 27–35. Doi:

10.2118/09-03-36.

Alshmakhy, A.B. and Maini, B.B. 2012. Foamy-Oil-Viscosity Measurement. J Can Pet

Technol 51 (1): 60–65. doi: 10.2118/136665-PA.

Alston, R.B. 1981. Enhanced oil recovery employing blend of carbon dioxide, inert gas and

intermediate hydrocarbons. US Patent No. 4,299,286.

ALvestad, J., Lagisquet, A., and Gilje, E. 2013. Solvent injection recovery process. US

Patent No. 2013/0000894 A1.

B utler, R.M. and Mokrys, I.J. 1993. Recovery of Heavy Oils using Vapourized

Hydrocarbon Solvents: Further Development of the VAPEX Process. J Can Pet

Technol 32 (6): 56–62. doi: 10.2118/93-06-06.

B utler, R.M. and Mokrys, I.J. 1998. Closed-loop Extraction Method for the Recovery of

Heavy Oils and Bitumens Underlain by Aquifers the VAPEX Process. J Can Pet

Technol 37 (4): 41−50. doi: 10.2118/98-04-04.

Boustani, A. and Maini, B.B. 2001. The role of diffusion and convective dispersion in

vapour extraction process. J Can Pet Technol 40 (4): 68–77. doi: 10.2118/01-04-05.

Butler, R.M. 1985. A New Approach To The Modeling Of Steam-Assisted Gravity

Drainage J Can Pet Technol 24 (3): 42–51. JCPT-85-03-01. doi: 10.2118/85-03-01.

Butler, R.M. 1997. Process and apparatus for the recovery of hydrocarbons from a

Page 213: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

192

reservoir of hydrocarbon. US Patent No. 5,607,016.

Butler, R.M. and Jiang, Q. 2000. Improved recovery of heavy oil by Vapex with widely

spaced horizontal injectors and producers. J Can Pet Technol 39 (1): 48–56. doi:

10.2118/00-01-04.

Butler, R.M. and Mokrys I.J. 1995. Process and apparatus for the recovery of hydrocarbons

from a hydrocarbon deposit. US Patent No. 5,407,009

Butler, R.M. and Mokrys, I.J. 1989. Solvent Analogue Model of Steam-assisted Gravity

Drainage. AOSTRA Journal of Research. 5 (1): 17–32.

Butler, R.M. and Mokrys, I.J. 1991. A New Process (VAPEX) for Recovering Heavy Oils

using Hot Water and Hydrocarbon Vapour. J Can Pet Technol 30 (1), 97–106. doi:

10.2118/91-01-09.

Butler, R.M. and Stephens, D.J. 1981. The gravity drainage of steam-heated heavy oil to

parallel horizontal wells. J Can Pet Technol 20 (2): 90–96. JCPT-81-02-07. doi:

10.2118/81-02-07.

Butler, R.M., McNab, G.S., and Lo, H.Y. 1981. Theoretical studies on the gravity drainage

of heavy oil during in-situ steam heating. Can. J. Chem. Eng. 59 (4), 455–460.

doi: 10.1002/cjce.5450590407.

Chakrabarty, T. 2010. Solvent for extracting bitumen from oil sands. US Patent No.

2010/0130386 A1.

Chen, Z. 2007. Reservoir simulation: mathematical techniques in oil recovery. Society of

Industrial and Applied Mathematics, Phyladelphia, PA.

Christiansen, R.L. and Smith, S.R. 1993. Sequentially flooding a subterranean

hydrocarbon-bearing formation with a repeating cycle of immiscible displacement

Page 214: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

193

gases. US Patent No. 5,232,049.

Chuang, B.C., Bose, M., Morton, S.A., Elkow, K.J., Erlendson, E. 2009. Methods of

improving heavy oil production. US Patent No. 7,527,096 B2.

CMG, STARS manual, Version 2009.

Cohen, Y. and Metzner, A.B. 1981. Wall effects in laminar flow of fluids through packed

beds. AIChe J. 27 (5):705–715. doi: 10.1002/aic.690270502.

Cuthiell, D. and Edmunds, N. 2012. Thoughts on simulating the VAPEX process. Paper

SPE-158499 presented at the SPE heavy oil conference, Calgary Canada, 12–14

June.

Cuthiell, D., McCarthy, C, Kissel, G., and Cameron, S. 2006. The role of capillary in

VAPEX. Paper PETSOC 2006-073, presented at the Canadian International

Petroleum Conference, Calgary, Alberta, 1315 June. doi: 10.2118/2006-073.

Cuthiell, D., Kissel, G., Jackson, C., Frauenfeld, T., Fisher, D., and Rispler, K. 2006.

Viscous Fingering Effects in Solvent Displacement of Heavy Oil. J Can Pet Technol

45 (7): 29–39. doi: 10.2118/06-07-02.

Cuthiell, D., Mccarthy, C., Frauenfeld, T., Cameron, S., and Kissel, G. 2003. Investigation

of the VAPEX process using CT scanning and numerical simulation. J Can Pet

Technol 42 (2): doi: 10.2118/03-02-04.

Darvish, M.P., and Firoozabadi, A. 1999. Solution-gas Drive in Heavy Oil Reservoirs. J

Can Pet Technol 38 (4): 54–61. doi: 10.2118/99-04-06.

Das, S.K. 1998. VAPEX: an efficient process for the recovery of heavy oil and bitumen.

SPE Journal 3 (3): 232–237. doi: 10.2118/50941-PA.

Das, S.K. 2005. Diffusion and dispersion in the simulation of VAPEX process. Paper

Page 215: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

194

SPE-97924 presented at the 2005 SPE International Thermal Operation and Heavy

Oil symposium, Calgary, Canada, 1–3 November.

Das, S.K. and Butler, R.M. 1994. Effect of aphalentene deposition on Vapex process a

preliminary investigation using a Hele-Shaw cell. J Can Pet Technol 33 (6): 39–45.

Das, S.K. and Butler, R.M. 1996. Diffusion coefficients of propane and butane in Peace

River bitumen. Can J Chem Eng 74 (6): 985–992.

Das, S.K., Butler, R.M. 1998. Mechanism of the vapour extraction process for heavy oil

and bitumen. J Pet Sci Eng 21 (1): 43–59. doi: 0920-4105/98.

Denbina, E.S., Boberg, T.C., and Rottor, M.B. 1991. Evaluation of Key Reservoir Drive

Mechanisms in the Early Cycles of Steam Stimulation at Cold Lake SPE Reservoir

Engineering 6 (2): 207–211. SPE-16737-PA. doi: 10.2118/16737-PA.

Dong, M., Huang, S., and Hutchence, K. 2006. Methane Pressure-Cycling Process With

Horizontal Wells for Thin Heavy-Oil Reservoirs. SPE Reservoir Evaluation &

Engineering 9 (2): 154–164. doi: 10.2118/88500-PA.

Dunn, S.G., Nenniger, E.H., and Rajan, V.S.V. 1989. A study of bitumen recovery by

gravity drainage using low temperature soluble gas injection. Can J Chem Eng 67 (6):

978-991. doi: 10.1002/cjce.5450670617.

Etminan, S.R., Maini, B.B., Chen, Z., and Hassanzadeh, H.S.R. 2010. Constant-Pressure

Technique for gas diffusivity and solubility measurements in heavy oil and bitumen.

Energy & Fuels 24 (1): 33-549. doi: 10.1021/ef9008955.

Frauenfeld, T., Lillico, D., Jossy, C., Vilcsak, G., Rabeeh, S., and Singh, S. 1998.

Evaluation of Partially Miscible Processes for Alberta Heavy Oil Reservoirs. J Can

Pet Technol 37 (4): 17–24. JCPT 98-04-01. doi: 10.2118/98-04-01.

Page 216: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

195

Fang, W. and Wheeler, T.J. 2012. Method for accelerating start-up for steam-assisted

gravity drainage (SAGD) operations. US Patent No. 2012/0227965 A1.

Fick, A., Poggendorff’s Annalen, 94 (1855) 59-86. (in English) Phil. Mag. 10 (4) 30–39.

Firouz A.Q. and Torabi, F. 2012. Feasibility study of solvent-based huff-n-puff method

(Cyclic solvent injection) to enhance heavy oil recovery. Paper SPE 157853

presented at the SPE Heavy Oil Conference Canada, Calgary, Alberta, 12–14 June.

doi: 10.2118/157853-MS.

Franenfeld, T.J.W. and Lillico, D.A. 1999. Solvent-assisted method for mobilizing viscous

heavy oil. US Patent No. 5,899,274.

Frauenfeld, T., Jossy, C., and Wang, X. 2007. Experimental studies of thermal solvent oil

recovery process for live heavy oil. J Can Pet Technol 46 (11): 41-46. doi:

10.2118/07-11-03.

Frauenfeld, T., Jossy, C., Rispler, K., and Kissel, G. 2006. Evaluation of the bottom water

reservoir VAPEX process. J Can Pet Technol 45 (9): 29–35. doi: 10.2118/06-09-02.

Gates, I.D. and Bunio, G. 2008. In situ process to recover heavy oil and bitumen. US Patent

No. 2008/0017372 A1.

Gutek, A.M.H., Harschnitz, B., Myers, R.D., and Okazawa, T. 2003. Combined steam and

vapour extraction process (SAVEX) for in situ bitumen and heavy oil production.

US Patent No. 6,662,872 B2.

Gutek, H., Harschnitz, B., Myers, R.D., and Okazawa, T. 2004. Combined steam and

vapour extraction process. Canadian Patent No. 2,325,777.

H. Darcy, Les Fontaines Publiques de la Ville de Dijon, Dalmont, Paris, 1856.

Haghighat, P. and Maini, B.B. 2010. Role of asphaltene precipitation in vapex process. J

Page 217: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

196

Can Pet Technol 49 (3) 14–21. doi: 10.2118/134244-PA.

Hayduk, W., Castaneda, R., Bromfield, H., Perras, R.R. 1973. Diffusivities of propane in

normal paraffin, Chlorobenzene, and Butanol solvents. AIChE Journal 19 (4),

859–861. doi: 10.1002/aic.690190432.

Hayduk, W., Minhas, B.S. 1982. Correlations for Prediction of molecular diffusivities in

liquids. Can. J. Chem. Eng., 60 (2): 295–299. doi: 10.1002/cjce.5450600213.

Hoier, L., Alvestad, J., Lagisquet, A., and Gilje, E. 2013. Solvent and gas injection

recovery process. US Patent No. 2013/0025858 A1.

Holm, L.W. and Josendal, V.A. 1974. Mechanisms of oil displacement by carbon dioxide.

J. Pet. Tech. 26 (12): 1427–1436. DOI:10.2118/4736-PA.

Huang, W.S. and Chien, S.F. 1986. Method and apparatus for producing viscous

hydrocarbons from a subterranean formation. US Patent No. 4,577,691.

Ito, Y., Ichikawa, M., and Hirata, T. 2001. The Effect of Gas Injection on Oil Recovery

During SAGD Projects J Can Pet Technol 40 (1): 38–43. JPCT-01-01-03. doi:

10.2118/01-01-03.

Ivory, J, Chang, J, Coates, R, and Forshner, K. 2010. Investigation of cyclic solvent

injection process for heavy oil recovery. J Can Pet Technol 49 (9): 22–33. doi:

10.2118/140662-PA.

Jamaloei, B.Y., Dong, M.Z., Mahinpey, N., and Maini, B.B. 2012. Enhanced cyclic solvent

process (ECSP) for heavy oil and bitumen recovery in thin reservoirs. Energy &

Fuels 26 (5): 2865–2874. doi: 10.1021/ef300152b.

James, L.A., Rezaei, N., and Chatzis, I. 2009. VAPEX, Warm VAPEX and Hybrid

VAPEX - The State of Enhanced Oil Recovery for In Situ Heavy Oils in Canada. J

Page 218: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

197

Can Pet Technol 47 (4): 1–7. doi: 10.2118/08-04-12-TB.

Javaheri, M. and Abedi, J. 2008. Modeling Mass Transfer Boundary Layer Instability in

the CO2-Based VAPEX Process. J Can Pet Technol 48 (8): 42-48. JPCT- 2008-089.

doi: 10.2118/09-08-42.

Jensen, E.M., Uhrich, K.D., and Hassan, D.J. 1998. Single well vapor extraction process.

US Patent No. 5,771,973.

Jia, X., Jiang, T., Zeng, F., and Gu, Y. 2013. Enhanced vapor extraction through foamy oil

flow. Presentated at the SPE Heavy Oil Conference Canada held in Calgary, Alberta,

Canada, 11–13 June 2013.

Jiang, Q. and Butler, R.M. 1996. Experimental Studies on Effects of Reservoir

Heterogeneity on the Vapex Process. J Can Pet Technol 35 (10): 46–54. doi:

10.2118/96-10-04.

Jiang, T., Jia, X., Zeng, F., and Gu, Y. 2013. A novel solvent injection technique for

enhanced heavy oil recovery: cyclic production with continuous solvent injection.

Paper SPE 165455 presented the SPE Heavy Oil Conference Canada, Calgary,

Alberta, 11–13 June. doi: 10.2118/165455-MS.

Kapadia, R.A., Upreti, R.S., Lohi, A., and Chatzis, I. 2006. Determination of gas

dispersion in vapor extraction of heavy oil and bitumen. J Pet Sci Eng 51 (3–4):

214–222. doi: 10.1016/j.petrol.2006.01.001.

Knorr, K.D. and Imran, M. 2011. Extension of Das and Butler Semianalytical Flow Model.

J Can Pet Technol 50 (6): 53–60. doi: 10.2118/148944-PA.

Knorr, K.D. and Imran, M. 2012. Solvent-chamber development in 3D-physical-model

experiments of solvent-vapour extraction (SVX) processes with various

Page 219: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

198

permeabilities and solvent-vapour qualities. J Can Pet Technol 51 (6): 425–436. doi:

10.2118/149190-PA.

Kumar, R. and Mahadevan, J. 2012. Well-Performance Relationships in Heavy-Foamy-Oil

Reservoirs. SPE Production & Operation 27 (1): 94–105. doi: 10.2118/117447-PA.

Kumer, A., Jaiswal, D.K., and Kumer, N. 2009. Analytical solutions of one dimensional

convection−diffusion equation with variable coefficient in a finite domain. J. Earth

Sys. Sci. 118 (5), 539–549. doi: 10.1007/S12040-009-0049-Y.

Lederer, E.L. 1933. Proc. World Pet. Cong. (London) 2, (1933): 526–528.

Lewis, E. and Mohanty, K.K. 2011. Partially miscible VAPEX displacement of a

moderately viscous oil. J Pet Sci Eng 77 (1): 104–110. doi:

10.1016/j.petrol.2011.02.008.

Lim, G.B., Kry, R.P., Harker, B.C., and Jha, K.N. 1995. Cyclic stimulation of Cold Lake

oil sand with supercritical ethane. Paper SPE 30298 presented at the International

Heavy Oil Symposium, Calgary, Alberta, 19–21 June. doi: 10.2118/30298-MS.

Lim, G.B., Kry, R.P., Harker, B.C., and Jha, K.N. 1996. Three-dimensional scaled physical

modeling of solvent vapour extraction of Cold Lake bitumen. J Can Pet Technol 35

(4): 32–40. doi: 10.2118/96-04-03.

Luhning, R.W., Das, S.K., Fisher, L.J., Bakker, J., Grabowski, J., Engleman, J.R.,

Wong, S., Sullivan, L.A., and Boyle, H.A. 2003. Full Scale VAPEX

Process-Climate Change Advantage and Economic Consequences A. J Can Pet

Technol 42 (2): 29–34. doi: 10.2118/03-02-02.

Maini, B.B. 1996. Foamy Oil Flow in Heavy Oil Prodution. J Can Pet Technol 35 (6):

21-24. doi: 10.2118/96-06-01.

Page 220: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

199

Maini, B.B. 2001. Foamy oil flow. Journal of Petroleum Technology 53 (10): 54-64. doi:

10.2118/68885-MS.

Maini, B.B., Sarma, H.K., and George, A.E. 1993. Significance of foam-oil behavior in

primary production of heavy oils. J Can Pet Technol 32 (9): 50–54. doi:

10.2118/93-09-07.

Marino, M.A. 1974. Numerical and analytical solutions of dispersion in a finite, adsorbing

porous medium. Journal of the American Water Resources Association. 10 (1):

81–90. doi: 10.1111/j.1752-1688.1974.tb00542.x.

McMillen, J.M. 1985. Method of solvent stimulation of heavy oil reservoirs. US Patent No.

4,531,586.

Moghadam, S., Nobakht, M., and Gu, Y. 2008. Theoretical and physical modeling of a

solvent vapour extraction (VAPEX) process for heavy oil recovery. J Pet Sci Eng 69

(2009), 93–104. doi: 10.1016/j.petrol.2008.12.029.

Mokrys, I.J. 2001. Vapor extraction of hydrocarbon deposits. US Patent No. 6,318,464 B1.

Mokrys, I.J. and Butler, R.M. 1993. The rise of interfering solvent chambers: solvent

analog model of steam-assisted gravity drainage. J Can Pet Technol 32 (3): 26−36.

doi: 10.2118/93-03-02.

Moore, R.G., Laureshen, C.J., Belgrave, J.D.M., Ursenbach, M.G., and Mehta, S.A. 1995.

In-situ combustion in Canadian heavy oil reservoirs. Fuel 74 (8): 1169–1175. doi:

0016-2361/95.

Muller, K.E. 2001. Computing the confluent hypergeometric function, M(a, b, x).

Numerical Mathematics 90 (1): 179–196. doi: 10.1007/s0022110100285.

Nasr T.N. and Isaacs, E.E. 2001. Process for enhancing hydrocarbon mobility using a

Page 221: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

200

steam additive. US Patent No. 6,230,814 B1.

Nasr, T.N. and Isaacs, E. 2004. Process for enhancing HC on mobility using a steam

additive Canadian Patent No. 2,323,029.

Nenniger, E.H. 1979. Hydrocarbon recovery. Canadian Patent No. 1,059,432.

Nenniger, J. 2007. Method and apparattus for heavy oil production. Canadian Patent No.

2235085.

Nenniger, J. and Dunn, S. 2010. Situ extraction process for the recovery of hydrocarbons.

US Patent No. 2010/0236783 A1.

Nenniger, J. and Nenniger, E. 2005. Method and apparatus for stimulating heavy oil

production. US Patent No. 6,883,607 B2.

Nenniger, J. and Nenniger, E. 2008. Method and apparatus for stimulating heavy oil

production. Canadian Patent No. 2,351,148.

Nenniger, J. and Nenniger, E. 2008. Method and apparatus for stimulating heavy oil

production. US Patent No. 7,363,973 B2.

Nenniger, J.E. and Dunn, S.G. 2008. How fast is solvent based gravity drainage? Paper

PETSOC 2008-139 accepted for the Proceedings of the Canadian International

Petroleum/SPE Gas Technology Symposium 2008 Joint Conference (Petroleum

Society’s 59th Annual Technical Meeting), Calgary, Alberta, 17–19 June. doi:

10.2118/2008-139.

Nghiem, L.X., Kohse, B.F., and Sammon, P.H. 2001. Compositional Simulation of the

VAPEX Process. J Can Pet Technol 40 (8). 54–61. JCPT-01-08-05. doi:

10.2118/01-08-05.

Page 222: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

201

Ostos, A.N. and Maini, B.B. 2005. An Integrated Experimental Study of Foamy Oil Flow

during Solution Gas Drive. J Can Pet Technol 44 (4): 43–50. doi: 10.2118/05-04-05.

Rahnema, H., Kharrat, R. and Tostami, B. 2008. Experimental and numerical study of

vapor extraction process (VAPEX) in heavy oil fractured reservoir. Petroleum

Society Paper 2008-116 presented at Canadian International Petroleum Conference,

Calgary, Alberta, June 17–19.

Rathmell, J.J., Stalkup, F.I., and hassinger, R.C. 1971. A laboratory investigation of

miscible displacement by carbon dioxide. Paper SPE 3483 presented at Fall Meeting

of the Society of Petroleum Engineers of AIME, 3-6 October 1971, New Orleans,

Louisiana. doi: 10.2118/3483-MS.

Redford, D.A. and Luhning, R.W. 1999. In Situ Recovery from the Athabasca Oil

Sands-Past Experience and Future Potential, Part II. J Can Pet Technol 38 (13): 1–13.

doi: 10.2118/99-13-41.

Richardson, W.C. and Kibodeaux, K.R. 2001. Chemically assisted thermal flood process.

US Patent No. 6,305,472 B2.

Sayegh, S. and Maini, B.B. 1984. Laboratory evaluation of the CO2 huff-n-puff process for

heavy oil reservoirs. J. Can. Pet. Technol. 23 (3): 29–36. doi: 10.2118/84-03-02.

Schmidt, T. 1989. Mass transfer by diffusion, in: AOSTRA Technical Handbook on Oil

Sands, Bitumens and Heavy Oils, AOSTRA, Edmonton, Alberta.

Scott, A.S. and Jirka, G.H. 2002. Environmental fluid mechanisms part I: Mass transfer

and diffusion. Karlsruhe, Germany.

Page 223: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

202

Shelton, J.L. and Morris, E.E. 1973. Cyclic Injection of Rich Gas into Producing Wells to

Increase Rates from Viscous-Oil Reservoirs. Journal of Petroleum Technology 25

(8): 890–896. doi: 10.2118/4375-PA.

Sheng, J.J. 1997. Foamy oil flow in porous media. PhD dissertation, University of Alberta,

Edmonton, Canada.

Sheng, J.J., Maini, B.B., Hayes, R.E., and Tortike, W.S. 1997. Experimental Study of

Foamy Oil Stability. J Can Pet Technol 36 (4): 31–37. doi: 10.2118/97-04-02.

Sheng, J.J., Maini, B.B., Hayes, R.E., and Tortike, W.S. 1999. Critical Review of Foamy

Oil Flow. Tranp Porous Med 35 (2). 157–187. doi: 10.1023/A:1006575510872.

Shu, W.R. 1984. A viscosity correlation for mixtures of heavy oil, bitumen, and

petroleum fractions. SPE Journal 24 (3): 227–282. doi: 10.2118/11280-PA.

Shu, W.R. 1984. Viscous oil recovery method. US Patent No. 4,466,485.

Singhal, A.K., Das, S.K., Leggitt, S.M., Kasraie, M., and Ito, Y. 1997. Steam-Assisted

Gravity Drainage and VAPEX Process Reservoir Screening. Journal of Petroleum

Technology 1997: 1122–1124.

Smith, G.E. 1988. Fluid flow and sand production in heavy oil reservoirs under solution

gas drive. SPE Production Engineering 3 (2): 169–180. doi: 10.2118/15094-PA.

Sposito, G. and Weeks, S.W. 1998. Tracer convection by steady groundwater flow in a

stratified aquifer. Water Resources Research 34 (5): 1051–1059.

Stalkup, I.F. 1978. Carbon dioxide miscible flooding: Past, present, and outlook for the

future, Journal of Petroleum Technology 30(8): 1102-1112. doi: 10.2118/7042-PA

Stehfest, H. 1970. Numerical inversion of Laplace transformation. Communications of the

ACM 13 (1): 47–49.

Page 224: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

203

Tharanivasan, A.K. 2004. Measurements of Molecular Diffusion Coefficients of Carbon

Dioxide, Methane and Propane in Heavy Oil Under Reservoir Conditions, Master

Thesis, University of Regina, Regina, Saskatchewan (December 2004).

Tharanivasan, A.K., Yang, C., and Gu, Y. 2006. Measurements of Molecular Diffusion

Coefficients of Carbon Dioxide, Methane and Propane in Heavy Oil under Reservoir

Conditions. Energy & Fuels 20 (6): 2509–2517. Doi: 10.1021/ef060080d.Etminan,

Upreti, S.R. and Mehrotra, A.K. 2000. Experimental Measurement of Gas Diffusivity in

Bitumen Results for Carbon Dioxide. Ind. Eng. Chem. Res. 39 (4): 1080–1087. doi:

10/1021/ie990635a.

Upreti, S.R. and Mehrotra, A.K. 2002. Diffusivity of CO2, CH4, CZH6 and N2 in

Athabasca bitumen. Canadian Journal of Chemical Engineering 80: 116–125.

Upreti, S.R., Lohi, A., Kapadia, R.A., and El-Haj, R. 2007. Vapour extraction of heavy oil

and bitumen: a review. Energy & Fuels 21 (3): 562–1574. doi: 10.1021/ef060341j.

Vittoratos, E., Scott, G.R., and Beattie, C.I. 1990. Cold lake cyclic steam stimulation: A

multi-well process. SPE Reservoir Engineering 5 (1): 19–24. SPE-17422-PA. doi:

10.2118/17422-PA.

Vogel, J.V. 1987. Gravity stabilized thermal miscible displacement process. US Patent No.

4,697,642.

Wheeler, M.F. and Dawson, C.N. 1992. An operator-splitting method for

convection−diffusion−reaction problems. London: Academic Press.

Wong, Kenny and Hall, W.L. 1982. Viscous oil recovery method. US Patent No.

4,324,291.

Page 225: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

204

Wong, R.C.K. 2003. Sand Production in Oil Sand under Heavy Oil Foamy Flow. J Can Pet

Technol 42 (3): 56–61. doi: 10.2118/03-03-06.

Wu, X., Polokar, M., and Cunha, L.B. 2005. A numerical approach to simulate and design

VAPEX experiments. Paper 2005-180 presented at the Petroleum Society 6th

Canadian International Petroleum Conference, Calgary, Canada, June 7–9.

Xu, S., Zeng, F., Gu, Y., and Knorr, K.D. 2012. Upscaling study of Vapour Extraction

through Numerical Study. Transp Porous Med 95 (3): 697–715. doi:

10.1007/s11242-012-0069-y.

Yang, C. and Gu, Y. 2006. Diffusion coefficients and oil swelling factors of carbon dioxide,

methane, ethane, propane, and their mixtures in heavy oil. Fluid Phase Equilibria

243 (2006): 64–73. doi:10.1016/j.fluid.2006.02.020.

Yazdani, A. and Maini, B.B. 2005. Effect of drainage height and grain size on production

rates in the Vapex process: Experimental study. SPE Reservoir Evaluation &

Engineering 8 (3): 205–213. doi: 10.2118/89409-PA.

Yazdani, A. and Maini, B.B. 2008. Modeling of the Vapex process in a very large physical

model. Energy & Fuels 22 (1): 535–544. doi: 10.1021/ef700429h.

Yazdani, A. and Maini, B.B. 2009. Pitfalls and solutions in numerical simulation of Vapex

experiments. Energy & Fuels 23 (8): 3981–3988. doi: 10.1021/ef900200f.

Zhang, H., Luo, P., and Gu, Y. 2006. Physical modeling of heavy oil production rate in a

vapour extraction process. Paper PETSOC 2006-142 presented at the Petroleum

Society’s 7th Canadian International Petroleum Conference (57th Annual Technical

Meeting), Calgary, Alberta, 13–15 June. doi: 10.2118/2006-142.

Zhao, L., Nasr, T.N., Huang, H., Beaulieu, G., Heck, G., and Golbeck, H. 2005. Steam

Page 226: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

205

alternating solvent process lab test and simulation. 44 (9): 37–43. doi:

10.2118/05-09-04.

Zeng, F., Knorr, K.D., and Wilton, R.R. 2008a. Enhancing oil rate in solvent vapour

extraction processes through Tee-well pattern. Paper SPE-117528 presented at the

2008 SPE International Thermal Operations, Calgary, Canada, 20–23, October.

Zill, D.G. 2001. A first course in differential equations with modeling applications. Pacific

Grove: Brooks/Cole Pub Co.

Page 227: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

206

APPENDIX A

Numerical solution to the diffusion equation with a moving boundary

The solvent–heavy oil mass transfer model is composed of the governing equation,

boundary conditions with a moving velocity, and initial conditions:

c cD

, (A1)

1 , 0t tc c , (A2)

max,c s c , (A3)

, 0c L

, (A4)

0s U . (A5)

Discretize the space and time domains, and rewrite the governing equation by using

Crank–Nicolson FDM:

1 1 1 1 11 1 1 1

1 11

2i i i i i i i i i i

i i i ic c c c c c c c c cD D D D

. (A6)

Let

22

, (A7)

then

1 1 1 1 11 1 1 1 1 1

ti i i i i i i i i i i i i ic c D c c D c c D c c D c c

. (A8)

Rearrange Eq. (A8)

1 1 11 1 1 1 1 1 1 1i i i i i i i i i i i i i iD c D D c D c D c D D c D c . (A9)

Page 228: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

207

Apply the FDM to the BCs:

1 maxc c , (A10)

1 0N Nc c . (A11)

For the second step, left boundary point moves to x2; therefore, apply left BC to x2:

12 maxc c . (A12)

For points x3~xN-1:

1 1 11 1 1 1 1 1 1 1i i i i i i i i i i i i i iD c D D c D c D c D D c D c . (A13)

Apply the right BC to the last point xN

1 0N Nc c . (A14)

Hence, a matrix function can be formed as:

n-1 n-1 n-1 n-1c F

M , (A15)

where, the coefficient matrix [M] is a tri-diagonal matrix, {c}n-1 is the to-be-determined

unknown concentration matrix. The column matrix {F}n-1 and the coefficient matrix

[M](n-1)×(n-1) can be constructed given a velocity profile. Eq. (A15) can be solved by using

Thomas method to get the solvent concentration profile in the solvent-diluted heavy oil.

Page 229: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

208

APPENDIX B

Semi-analytical solution to an convection-diffusion equation with a constant d and a

variable v.

Mathematical model in dimensionless form is:

2

2D D D

DD D D D

c c c PePe ct x x x

,

(B1)

( 0, ) 1 0D DD Dx tc t

, (B2)

( 1, )

0 0D D

DD

D x t

c tx

, (B3)

( , 0) 0 0 1D DD Dx tc x

. (B4)

Set cD as 2

* 4 2D Da bx x

Dc c e

, and assume a linear Pe as DPe ax b , then Eq. (B1)

becomes:

* 2 *2 *

2

1( )4 2D

D D

c c aax b ct x

. (B5)

Performing the Laplace transformation and considering IC:

2 *2 *

2

1 ( ) 04 2D

D

c aax b s cx

. (B6)

Allow 1 ( )Dz ax ba

, and substitute xD with z in Eq. (B6) yields:

22

2

1 04

C z J Cz

, (B7)

Page 230: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

209

where, )21( a

sJ . Eq. (B7) is the canonical form of hyperbolic cylindrical equation,

whose general solution is:

1 2C AC BC , (B8)

where, C1 and C2 are the independent odd and even solutions. Coefficients A and B can

be obtained by applying the BCs:

21 24

11 1( , ) ( , , )

2 4 2 2z J zC J z e M

,

21 24

23 3( , ) ( , , )

2 4 2 2z J zC J z ze M

. (B9)

Applying BCs, Eqs. (B2−B3), A and B can be obtained as:

0 01 2

1 1As C C

, 0 01 2

1Bs C C

, (B10)

where,

1111

1122

2

2

C a ba Cz

C a ba Cz

; subscripts ‘1’ and ‘2’ of C stand for the two

independent solutions; superscripts ‘0’ and ‘1’ denote the left and the right boundaries,

respectively; M denotes the Kummer’s function. The concentration in real domain can be

obtained by applying the Stehfest Laplace inverse transformation:

2

4 2

1

ln 2 ( )D Da b nx x

D jjD

c e V C st

, (B11)

where,

min2 2+

2

12

(2 )!( 1)( )!( )!( 1)!( )!(2 )!2

n n, jn j

jjk

k kV n k k k j k k j

, ln 2

D

s jt

. (B12)

Page 231: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

210

APPENDIX C

Definitions of the dimensionless terms

The dimensionless terms in Eq. (4.27) are defined as:

a sa

, (C1)

2

' ''

a a s b aa

, (C2)

2' ( ' )

'Da ax a a b

a . (C3)

Page 232: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

211

APPENDIX D

Analytical solution to an convection-diffusion equation with a special hyperbolic

convection velocity profile.

For a hyperbolic Pe profile, Eq. (4.29), set ( )

D

D

cCx

and substitute it in Eq.

(4.3):

2

2D D

C Ct x

. (D1)

Applying the Laplace transformation and considering the IC, Eq. (D1) becomes:

2

2D

CsCx

. (D2)

Its solution is:

D Ds x sxC Ae Be . (D3)

Coefficients A and B can be obtained by applying the BCs in the Laplacian domain:

D D( 0, )

1x t

Cs

, (D4)

( 1, )

(1 ) 0D D

D x t

C Cx

. (D5)

Back replace A and B into Eq. (D3), the analytical solution becomes:

(2 )

2( )

(1 )

D Dsx s x

D s

e eC xs e

, (D6)

Page 233: ENHANCED SOLVENT VAPOUR EXTRACTION PROCESSES IN …

212

where, (1 ) 1(1 ) 1

ss

. The concentration in real domain can be acquired by applying

the Stehfest Laplace inverse transformation.