relative permeability equation-of-state: the role of …

233
The Pennsylvania State University The Graduate School RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF PHASE CONNECTIVITY, WETTABILITY, AND CAPILLARY NUMBER A Dissertation in Energy and Mineral Engineering by Prakash Purswani © 2021 Prakash Purswani Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2021

Upload: others

Post on 21-Feb-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

The Pennsylvania State University

The Graduate School

RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF PHASE

CONNECTIVITY, WETTABILITY, AND CAPILLARY NUMBER

A Dissertation in

Energy and Mineral Engineering

by

Prakash Purswani

© 2021 Prakash Purswani

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

May 2021

Page 2: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

ii

The dissertation of Prakash Purswani was reviewed and approved by the following:

Zuleima T. Karpyn

Dissertation Co-Adviser

Associate Dean for Graduate Education and Research, College of Earth and Mineral Sciences

Professor of Petroleum and Natural Gas Engineering, Department of Energy and Mineral Engineering

Chair of Committee

Russell T. Johns

Dissertation Co-Adviser

George E. Trimble Chair in Energy and Mineral Sciences

Professor of Petroleum and Natural Gas Engineering, Department of Energy and Mineral Engineering

Amin Mehrabian

Assistant Professor of Petroleum and Natural Gas Engineering, Department of Energy and Mineral

Engineering

Xiaolei Huang

Associate Professor of Data Science, Penn State College of Information Sciences and Technology

Turgay Ertekin

Professor Emeritus of Petroleum and Natural Gas Engineering, Department of Energy and Mineral

Engineering

Special Committee Member

Mort D. Webster

Professor of Energy Engineering

Co-Director Initiative for Sustainable Electric Power Systems

Associate Department Head for Graduate Education, Department of Energy and Mineral Engineering

Page 3: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

iii

ABSTRACT

Relative permeability (kr) is a transport property used for characterizing the flow of multiple phases through

a porous medium. Inputs of kr are integral for reservoir simulations. Multiple parameters such as phase

saturation, wettability of the medium, fluid properties, flow characteristics, pore topology, fluid phase

topology, and fluid/fluid interfacial areas are known to affect relative permeabilities. Current kr models are

functions of phase saturation that are matched for specific flow/experimental conditions. The other

parameters affecting relative permeabilities are inherently captured through these saturation functions.

Representation of relative permeabilities only in the saturation space causes non-uniqueness and path

dependency in relative permeabilities which often cause simulations to fail because they lack generality

and are not physically based. As a result, hysteresis in relative permeabilities arises, which is a major

modeling issue for reservoir simulations.

In this dissertation, models for relative permeabilities are presented by considering functional forms

that include the effects of the key controlling parameters on relative permeabilities. The purpose of this

dissertation is twofold, to (a) understand how different parameters, specifically, phase saturation, phase

connectivity, capillary number, and wettability affect relative permeabilities; (b) propose physically-based

kr models by including the effects of these parameters.

Relative permeabilities are modeled using an equation-of-state (EOS) approach where the exact

differential for relative permeability is written in phase connectivity and saturation ( ˆ ).S − A quadratic

response-based EOS for relative permeability is modeled in the ˆ S − space. Physical limiting conditions on

the state parameters are considered to constrain the EOS model. This model is tested for different capillary

numbers ranging from one to 10-6. In addition, we calculated the partial derivatives of relative permeabilities

in the state parameters using numerical data sets generated with pore-network modeling. A response for

relative permeability is derived in the ˆ S − space following the state function approach. The locus bounded

by residual nonwetting phase connectivity and residual nonwetting phase saturation is presented for two

contact angles in the water-wet regime. Finally, we investigated the role of wettability on phase trapping

Page 4: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

iv

also using pore-network modeling. An extended Land-based hysteresis trapping model is presented and

compared against models from the literature. In addition, models are presented to capture the trends of

residual loci for different contact angles.

Results show that a simple quadratic response for relative permeability in the ˆ S − space captures trends

across different capillary numbers. The model tuned for a capillary number in the capillary dominated

regime can show predictive capability for other capillary numbers within the same regime. The linear kr-S

paths for high capillary numbers (small Corey exponents) and nonlinear kr-S paths for low capillary

numbers (high Corey exponents) are found to occur due to fast and slow changes in phase connectivity,

respectively. Limiting constraints help in the identification of the physical region in the ˆ S − state space.

Results also show that the response derived for relative permeability from relative permeability partial

derivatives using the state function approach can predict relative permeabilities over the entire numerical

data sets, regardless of the direction of flow, thus mitigating hysteresis. Further, the analysis of the effect

of wettability shows that both phase trapping as well as the locus of residual saturation and residual phase

connectivity are sensitive to contact angle changes. For low receding phase contact angles, the residual

locus remains fairly constant, but at higher contact angles, the shape of the residual locus resembles a closed

loop. Pore structure constraint at negligible saturation is found to control the shape of the residual locus.

Phase trapping was found to reduce significantly for high contact angles owing to pore-scale mechanisms

of layer flow of the receding phase and piston-like advance of the invading phase. A newly proposed

extended Land-based model is able to capture residual saturation trends for all contact angles.

Overall, through this research endeavor, we gain insight into the different intrinsic parameters that

affect relative permeability. Through the application of pore-scale measures, these insights are further

manifested into practical models that helps describe relative permeabilities physically.

Page 5: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

v

TABLE OF CONTENTS

LIST OF FIGURES ................................................................................................................................... viii

LIST OF TABLES ..................................................................................................................................... xiii

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

ACKNOWLEDGEMENT .......................................................................................................................... xx

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

1.1. Background ........................................................................................................................................ 1

1.2. Relative permeability ......................................................................................................................... 4

1.2.1. Parameters affecting relative permeabilities ............................................................................... 8

1.2.2. Models for relative permeability ............................................................................................... 21

1.3. Research objectives .......................................................................................................................... 33

1.4. Dissertation layout ........................................................................................................................... 34

1.5. Publication list ................................................................................................................................. 36

CHAPTER 2. IMAGING AND PORE-SCALE MEASUREMENTS ....................................................... 37

2.1. Introduction ...................................................................................................................................... 39

2.2. Methodology .................................................................................................................................... 46

2.2.1. Implementation of the segmentation techniques ....................................................................... 48

2.2.2. Benchmark case ........................................................................................................................ 48

2.2.3. Test case .................................................................................................................................... 48

2.2.4. Supervised machine learning (ML)-based on Fast Random Forest algorithm .......................... 49

2.2.5. Unsupervised machine learning based on k-means and fuzzy c-means clustering ................... 50

2.3. Results and discussion ..................................................................................................................... 52

2.3.1. Bulk measurements ................................................................................................................... 56

2.3.2. Pore-scale measurements .......................................................................................................... 61

2.4. Concluding remarks ......................................................................................................................... 63

Page 6: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

vi

CHAPTER 3. EQUATION-OF-STATE AND CAPILLARY NUMBER ................................................. 65

3.1. Introduction ...................................................................................................................................... 67

3.2. Methodology .................................................................................................................................... 71

3.2.1. Development of a state function ............................................................................................... 71

3.2.2. Phase connectivity..................................................................................................................... 73

3.2.3. Development of relative permeability EOS .............................................................................. 74

3.2.4. Comparison to the development in Khorsandi et al. (2017) ..................................................... 77

3.2.5. Estimation of the coefficients of the EOS ................................................................................. 80

3.3. Results and discussion ..................................................................................................................... 82

3.3.1. Quadratic response for relative permeability ............................................................................ 82

3.3.2. Quadratic response prediction at neighboring conditions ......................................................... 86

3.3.3. Effect of capillary number ........................................................................................................ 86

3.4. Concluding remarks ......................................................................................................................... 90

CHAPTER 4. IMPACT OF WETTABILITY ON PHASE TRAPPING ................................................... 91

4.1. Introduction ...................................................................................................................................... 93

4.2. Methodology .................................................................................................................................. 100

4.2.1. Pore-network simulations ....................................................................................................... 100

4.3. Results and discussion ................................................................................................................... 103

4.3.1. Iso-quality curves for saturation and connectivity – interpreting PNM data .......................... 103

4.3.2. Initial-residual (IR) saturation trapping curves ....................................................................... 106

4.3.3. Modeling IR saturation trapping curves .................................................................................. 108

4.3.4. Phase connectivity – phase saturation ( χ – S) paths and trapping locus ................................. 113

4.3.5. Modeling residual phase connectivity and trapping locus ...................................................... 117

4.8. Concluding remarks ....................................................................................................................... 121

Page 7: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

vii

CHAPTER 5. DEVELOPMENT OF EQUATION-OF-STATE USING PORE-NETWORK MODELING

.................................................................................................................................................................. 123

5.1. Introduction .................................................................................................................................... 125

5.2. Methodology .................................................................................................................................. 132

5.2.1. Example implementation of the EOS ...................................................................................... 136

5.2.2. Two-phase simulations using pore-network modeling ........................................................... 138

5.3. Results and discussion ................................................................................................................... 141

5.3.1. Evolution of connectivity with saturation ............................................................................... 141

5.3.2. Fitting kr-EOS to literature data .............................................................................................. 147

5.3.3. Relative permeability scanning curves: pore-network simulations ......................................... 149

5.3.4. Estimation of relative permeability partial derivatives ........................................................... 154

5.3.5. Prediction of relative permeability .......................................................................................... 156

5.4. Concluding remarks ....................................................................................................................... 160

CHAPTER 6. CONCLUDING REMARKS AND OUTLOOK FOR FUTURE RESEARCH ................ 162

6.1. Key concluding remarks ................................................................................................................ 165

6.2. Future research ............................................................................................................................... 167

APPENDICIES ......................................................................................................................................... 169

Appendix A. Basic theory of Minkowski functionals ........................................................................... 169

Appendix B. A consistent approach for χmax determination .................................................................. 175

Appendix C. Base regression code created for analysis presented in chapter 3 ................................... 177

Appendix D. Data used in chapter 3 (adapted from Armstrong et al. 2016) ........................................ 183

Appendix E. Procedure for developing iso-quality curves discussed in chapter 4 ............................... 185

Appendix F. Residual curves for different stopping criteria and wettability alteration in PNM .......... 186

Appendix G. Base code for running PNM for generating numerical data sets for chapters 4 and 5 .... 188

REFERENCES ......................................................................................................................................... 196

Page 8: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

viii

LIST OF FIGURES

Figure 1. 3-D visualization of fluid flow in porous media showing the endpoint states of primary drainage

and imbibition processes. (Left), (middle), and (right) show the states of the medium before primary

drainage, after the completion of primary drainage, and after the completion of imbibition, respectively.

Figure adapted from Schlüter et al. (2016). .................................................................................................. 2

Figure 2. Schematic showing water-oil relative permeabilities for a water-wet medium. Hysteresis in

relative permeabilities is also displayed. The black curves represent a primary drainage process, while the

red curves represent a primary water injection process. The solid curves are water relative permeabilities,

while the dashed curves are the oil relative permeabilities. Directions of flow are marked by arrows on the

figure. Endpoint relative permeabilities and saturations are also marked on the figure. Adapted from (Blunt

2017). ............................................................................................................................................................ 5

Figure 3. Schematic representation of a 2-D multiphase porous system showing possible phase/phase

contact lines. Left to right the wetting phase saturation increases. When considered in 3-D, phase/phase

interfacial areas would be estimated along the areal region of contact. Adapted from Dalla et al. (2002). 11

Figure 4. Segmented image of a multiphase system acquired using x-ray imaging showing (a) disconnected

nonwetting phase, and (b) more connected wetting phase. ......................................................................... 12

Figure 5. Schematic showing a water-wet solid in the presence of oil. ..................................................... 16

Figure 6. Illustrations of different types of wettability for oil/water/solid systems. From left to right the

medium’s wettability to oil increases. ......................................................................................................... 17

Figure 7. Schematic showing three types of rock wettabilities characterized qualitatively from the

visualization of two-phase relative permeability curves. Here, the two phases are oil and water and the flow

(direction marked by the arrow) represent water injection. The black, green, and red curves are for an oil-

wet, a water-wet, and a mixed-wet rock, respectively. ............................................................................... 18

Figure 8. Schematic of nonwetting phase relative permeability showing hysteresis after flow reversal from

primary drainage (black curve) to primary water injection (red curve). The initial, trapped, residual, and

flowing nonwetting phase saturations are marked on the figure. Adapted from Carlson (1981). .............. 28

Figure 9. Schematic of the laboratory setup and image acquisition system (x-ray MCT scanner). DO stands

for the distance between the detector and the object, while OS stands for the object to source distance. This

figure shows that the sample (object) is very close to the source for finer resolution, the resolution was

coarsened by moving the sample stage laterally from the source, increasing OS and decreasing DO. ...... 47

Page 9: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

ix

Figure 10. Imaged cross-section of dry (top) and brine saturated (middle and bottom) porous glass frit at

different voxel resolutions. The brine used for saturating the porous medium was 1M NaI solution. Non-

local means was used for filtering the raw images to remove image noise. ............................................... 53

Figure 11. Histograms showing the voxel population of the different grayscale intensity values for the

corresponding scans shown in Figure 10. ................................................................................................... 54

Figure 12. Segmented cross-sectional images showing three phases (solid, brine, and air). (a) thresholding

at a resolution of 6 µm (benchmark case), (b) thresholding at a resolution of 18 µm, (c) supervised machine

learning segmentation at a resolution of 18 µm, and (d) unsupervised machine learning segmentation at a

resolution of 18 µm. Zoomed-in version of the images are displayed on the sides to highlight distinct

features of segmented images. The upper and lower regions of interest (marked inside the segmented

images) correspond to labels 1 and 2, respectively. .................................................................................... 56

Figure 13. Vertical profiles of (a) porosity, (b) brine saturation, (c) air saturation, (d) solid fraction, (e)

brine fraction, and (f) air fraction for different segmentation techniques. .................................................. 58

Figure 14. Surface areas of air, brine, and solid phases on images segmented with different techniques. 61

Figure 15. Illustration showing the EOS state approach on a real path (simulation) taken during two-phase

flow simulation in jS and ˆ

j space. ........................................................................................................... 79

Figure 16. Phase saturation, relative permeability, and normalized Euler connectivity for the nonwetting

phase for different capillary numbers used for fitting the quadratic response for relative permeability as

well as for prediction purposes. Courtesy Dr. Ryan T. Armstrong. Data from Armstrong et al. (2016). This

data is tabulated in appendix D. .................................................................................................................. 81

Figure 17. (a) Quadratic response prediction versus simulation data (b) residual between the predicted and

simulation measurements for relative permeability based on the response surface fit shown in Figure 18.

.................................................................................................................................................................... 83

Figure 18. Contour map of the response surface of relative permeability as a function of phase saturation

and normalized Euler connectivity. The capillary number (~ 10-4), wettability, and pore structure have been

kept constant. Data points shown as black dots were taken from the two-phase flow simulations presented

in Armstrong et al. (2016). Dashed line represents a limiting boundary of plausible values. .................... 84

Figure 19. Partial derivative coefficients (calculated using Eqs. (3.10) and (3.11)) expressed as a function

of (a) phase saturation and (b) normalized Euler connectivity. .................................................................. 85

Figure 20. (a) Prediction of relative permeability and (b) residual error for capillary numbers ~10-3 and

~10-5 based on the response surface fit to capillary number 10-4 described in Figure 18. .......................... 86

Figure 21. R2 error for prediction of data at different capillary numbers using sub data set at capillary

number of ~10-4 as the fitted response surface. ........................................................................................... 87

Figure 22. Quadratic response surface fits to sub data sets at different capillary numbers. ...................... 88

Page 10: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

x

Figure 23. Phase relative permeability plots with corresponding phase connectivity value (shown in blue)

(a) high capillary number ~1 (b) low capillary number ~10-5. The red solid lines represent the fit using

Corey model with ~ exponent value of (a) 0.76 and (b) 1.29. The residual saturation in (a) was set to 0 while

computing the Corey exponent because that data point was not known. .................................................... 89

Figure 24. Schematic representation of capillary trapped CO2 by chase brine. Local heterogeneities in pore

structure, mineral complexity, and wettability are depicted. ...................................................................... 94

Figure 25. Initial-residual characteristic curves for selected literature studies. ......................................... 98

Figure 26. Steps in pore-network modeling. (a) The extracted internal structure (pore space) of the porous

medium used in this study; (b) The extracted pore-network of the porous medium in (a) represented in a

ball and stick form; (c), (d), and (e) show three saturation steps of blue phase injection in the pore-network

completely filled with the red phase. (c) shows no injection, (d) is captured after some injection, and (e) is

captured after longer injection period. ...................................................................................................... 101

Figure 27. Iso-quality trends for four wettability cases (θ1 = 180°, 120°, 60°, 0°). Plots to the left show the

iso-saturation curves, while plots to the right show the iso-connectivity curves. The phase saturation, phase

connectivity, and relative permeability are for the receding phase (phase1) during phase2 injection. The

contour lines are plotted at intervals of 0.03 units. The procedure for developing these curves is displayed

in Figure 51 in appendix E. ....................................................................................................................... 104

Figure 28. Comparison of kr-S (upper row) and – S (lower row) paths for two Si values (~ 1.0 and ~ 0.9)

for two receding contact angles cases (θ1 = 180° and 0°). ........................................................................ 106

Figure 29. Initial-residual saturation curves for different wettability cases observed using pore-network

modeling. .................................................................................................................................................. 107

Figure 30. IR saturation trapping curves for four different contact angle cases (θ1 = 180°, 120°, 60°, 0°).

The dashed lines show the fits for the trapping model. The corresponding goodness measure of the fits for

all wettability cases and R2 values are displayed on the right plots. ......................................................... 110

Figure 31. Matching parameters plotted as a function of wettability for the IR trapping model presented in

this research (Eq. (4.7)). The corresponding fits are available in Figure 30. The shaded regions mark the

95% confidence interval. .......................................................................................................................... 111

Figure 32. – S paths for the receding phase during secondary injection process for different contact angle

cases. The red open circles represent the initial condition whereas, the black open circles represent the

residual condition for each – S path. ..................................................................................................... 113

Figure 33. Relative permeability contour plots for the – S paths shown in Figure 32 ......................... 115

Figure 34. Trapping locus of phase saturation and phase connectivity for different contact angles. ...... 116

Page 11: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xi

Figure 35. IR phase connectivity curves for four different wettability conditions (θ1 = 180°, 120°, 60°, 0°).

The dashed lines show the match for the phase connectivity trapping model (Eq. (4.8)). The corresponding

goodness measure of the fit for all wettability cases and R2 value is displayed on the right plot. ............ 117

Figure 36. Matching parameters plotted as a function of wettability for the IR phase connectivity trapping

model. The corresponding fits are available in Figure 35. The shaded regions mark the 95% confidence

interval. ..................................................................................................................................................... 118

Figure 37. Summary of initial-residual phase saturation and phase connectivity trends for four different

contact angles (θ1 = 180°, 120°, 60°, 0°). (a) Initial connectivity versus initial saturation. Here all four

contact angles collapse to one single data set; (b) initial connectivity versus residual connectivity; (c) initial

saturation versus residual saturation; (d) residual connectivity versus residual saturation. ...................... 119

Figure 38. Ball and stick representations of a porous medium with similar number of pores but increasing

number of throats. The Euler characteristic decreases from left to right as connections (throats) are

increased. .................................................................................................................................................. 134

Figure 39. Waterflooding χ - S paths from the literature (Table 11). All cases are for NCa < 10-4. Solid curves

show the best fits to Eq. (5.14). Saturation in the experiments move from right to left as shown by the

direction of the arrow. ............................................................................................................................... 142

Figure 40. The χ - S paths for different drainage and imbibition scans using PNM for the weakly water-wet

case (o

θ ~ 50 ). All PD curves begin at zero oil saturation but terminate at So ~ 1.0 (green), So ~ 0.9 (blue),

and So ~ 0.8 (red). Next, IMB curves terminate at residual conditions (squares). Finally, all SD curves are

simulated to So ~ 1.0. ................................................................................................................................ 144

Figure 41. Fitting parameters (p and k) after matching Eq. (5.14) to PNM scanning data (see Figure 40 for

three of the scans plotted here at 1.0, 0.9, and 0.8 So). The x-axis is the nonwetting saturation at the

termination of PD. The shaded region shows the error bars calculated for 95% confidence limits using the

nlparci function in Matlab®. The contact angle averages at 50o. ............................................................. 145

Figure 42. Best fits to the literature data in Table 11 using the (a) kr-EOS with constant partial derivatives

and (b) Corey form. ................................................................................................................................... 148

Figure 43. PNM simulations of imbibition (IMB) and secondary drainage (SD). Two contact angles are

used: the left column represents a fixed contact angle of 0o, while the right column is for uniformly

distributed contact angles between 40o and 60o with an average of ~ 50o. The black solid line in (a) and (b)

is for primary drainage (oil flood). Figures (a) and (b) also give ten imbibition curves at 0.1 saturation

intervals on the PD curve, while (e) and (f) show their χ - S paths. Figures (c) and (d) are for secondary

drainage and (g) and (h) show their corresponding χ - S paths. The red open circles represent the starting point

for IMB, while the black open circles represent the residual points for each IMB scan (similar to Figure

Page 12: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xii

40). Arrows show the direction of saturation change (IMB points to the left while SD points to the right).

.................................................................................................................................................................. 151

Figure 44. Relative permeability contours plotted from the 200 PNM imbibition simulations for the two

contact angle cases of 0o and ~ 50o. .......................................................................................................... 152

Figure 45. PNM simulations for 200 imbibition and secondary drainage scanning curves that begin at

different PD termination saturations (spaced by 0.005 saturation units). Two contact angles are shown, 0o

(a) and ~ 50o (b). The PD curves begin at So = 0. ..................................................................................... 152

Figure 46. Locus of residual connectivity and residual saturation generated from the different scanning

curves (Figure 43). The blue data point shows the limiting value of connectivity as saturation approaches

zero (see Eq. (5.6)). ................................................................................................................................... 153

Figure 47. Relative permeability partial derivatives estimated from fitting a cluster of data to a plane.. 155

Figure 48. Contour maps of predicted relative permeability using constant partial approach (a, b) and

quadratic response (c, d) for both wettability cases. Actual versus predicted relative permeability values

for the two approaches are shown in (e) and (f). The R2 for both approaches in (e) was ~0.90, while the R2

was ~0.90 and ~0.96 for the constant partial approach and quadratic response, respectively, in (f). ....... 158

Figure 49. Schematic of capillary tube filled with wetting and nonwetting phases. A zoomed-in view of the

interface across the phases is also shown. ................................................................................................. 170

Figure 50. Schematic of different surfaces with different types of Gaussian curvatures. ........................ 172

Figure 51. Schematic showing the procedure for developing the iso-quality curves discussed in chapter 4

(Figure 27). ............................................................................................................................................... 185

Figure 52. Initial-residual saturation curves generated from pore-network modeling for different contact

angle cases measured through the receding phase. The residual saturations are obtained for different

stopping criteria and are shown by different colors. ................................................................................. 186

Figure 53. Primary drainage and water injection capillary pressure scanning curves generated using pore-

network modeling. The different colors (and the arrows mark) give different cycles of injection. These

scanning curves are generated with the endpoint of primary drainage (or starting point of water injection)

at Si = 0.9. .................................................................................................................................................. 187

Page 13: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xiii

LIST OF TABLES

Table 1. Empirical oil/water relative permeability functions for different rock types and wetting conditions

from the literature. ...................................................................................................................................... 25

Table 2. Characteristics of imaging techniques from various experiments of fluid flow in porous media.

.................................................................................................................................................................... 40

Table 3. Summary of percent errors calculated for the different measured properties and segmentation

techniques relative to the benchmark case. The percent errors for the bulk measurements of porosity, fluid

saturation, and phase fractions are calculated for the respective average values across the sample height.

.................................................................................................................................................................... 59

Table 4. Measurements of air-brine interfacial area and Euler number of the nonwetting (air) phase for the

different segmentation techniques. The average air saturation was 14.4 % as determined from the

benchmark case whereas, thresholding, supervised machine learning, and unsupervised machine learning,

showed average values of 11.5%, 15.0% and 11.4%, respectively. ........................................................... 62

Table 5. Physical constraints imposed on the relative permeability response by considering key limiting

conditions that affect relative permeability as a function of phase saturation and phase connectivity. The

phase is assumed nonwetting, although extensions to other phases are easily possible. ............................ 75

Table 6. Euler characteristic values estimated through 2-D extrapolation for the pore structure used during

simulations in Armstrong et al. (2016). ...................................................................................................... 81

Table 7. Model coefficients and the goodness of quadratic response surface fit to phase saturation and phase

connectivity to the data presented in (Armstrong et al. 2016) at the NCa of ~10-4. ..................................... 82

Table 8. Coefficients for the quadratic response and goodness measures for quadratic response surface fits

to sub data sets at different capillary numbers shown in Figure 22. ........................................................... 88

Table 9. IR trapping models from the literature. iS , rS , and max

rS refer to the initial, residual, and

maximum residual phase saturations, respectively, while, C, a, b, c, α, and β are model parameters. ..... 108

Table 10. Properties of the pore-network extracted from the dry micro-CT image of a Bentheimer sandstone

(Lin et al. 2019). The parameter Z (2nt/np) is the coordination number; np and nt are the number of pores and

throats, respectively. The χmin for the image data is the value V-E+F-O exacted from the pore space of the

image. The χmax for the image data is back calculated from the corresponding χmin and Z values. See

additional details for consistent estimation of χmax in appendix B. ........................................................... 139

Table 11. Summary of information for the literature data shown in Figure 39. ....................................... 142

Table 12. Values of parameters fit to Eq. (5.14) using Table 11 data. The curves from the best fits are

shown in Figure 39. The 95% confidence limits were calculated using the nlparci function in Matlab®.143

Page 14: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xiv

Table 13. Summary remarks for p and k values for the evolution of phase connectivity for different cycles

of injection. ............................................................................................................................................... 147

Table 14. Best fit values for the kr-EOS and Corey form shown in Figure 42. The end-point permeability

is the same for comparison purposes. Thus, only χ

α and k are used as tuning parameters for the kr-EOS and

no for Corey. R2 values are shown. ............................................................................................................ 148

Table 15. The average values of the estimated partial derivatives of relative permeability shown in Figure

47. ............................................................................................................................................................. 156

Table 16. Values of plane fitting parameters in Eq. (5.21) used for the kr response in Eq (5.22). ........... 157

Table 17. Example of different objects and their Euler numbers. ............................................................ 174

Table 18. Data used in chapter 3 for nonwetting phase saturation, connectivity, and relative permeability

(adapted from Armstrong et al. 2016). ...................................................................................................... 183

Page 15: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xv

NOMENCLATURE

Shorthand notation

CCUS Carbon capture utilization and storage

CK Carmen-Kozeny

CFD Computational fluid dynamics

2-D/3-D/4-D Two-dimensional/three-dimensional/4-dimensional

DL Deep learning

DO Object-to-detector

EOR Enhanced oil recovery

EOS Equation-of-state

GHG Greenhouse gas

IFT Interfacial tension, (N/m)

IMB Imbibition

IR Initial-residual

IPCC Intergovernmental panel on climate change

LBM Lattice Boltzmann methods

Micro-CT (or µCT

or MCT) Micro-computed tomography

MIA Medical image analysis

ML Machine learning

MRI Magnetic resonance imaging

OS Object-to-source

PNM Pore-network modeling

PV Pore volume (cc)

PVT Pressure-volume-temperature

PD Primary drainage

REV Representative elementary volume (cc)

SD Secondary drainage

WAG Water alternating gas

WEKA Waikato environment for knowledge analysis

Page 16: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xvi

Symbols

A, B, C Coefficients for planar representation of relative permeability for 2D extrapolation in

chapter 3; coefficients for model fitting in chapter 5

αi Coefficients in the quadratic expression for relative permeability, i = 0, 1, 2, 11, 22, 12

αp (αw, αnw, αs) Phase-specific surface area (specific wetting phase surface area, specific nonwetting phase

surface area, and specific solid surface area, respectively) (1/m)

αw-nw Specific fluid/fluid (wetting-nonwetting) interfacial area (1/m)

β0, β1, β2 Betti numbers 0, 1, and 2 respectively

A Fluid/fluid interfacial area (state parameter)

Ap Surface area of phase p (mm2)

C Integration constant; Land’s trapping coefficient; Model constant Burdine’s equation

 C Model constant Burdine’s

Dz Elevation of depth from the reference datum (m)

dl Line element [L]

ds Surface element [L2]

δE Entire surface of object E [L2]

Δx Finite distance within the porous medium (m)

E Number of edges in a polyhedra

E Object in space

f(σ) Interfacial tension scaling factor

F Faces in a polyhedral

g Gravitational acceleration constant (m/s2)

I (e.g., cosθ) Wettability (state parameter)

I Wetting fraction of the solid

a, b, c Coefficients for plane fitting partial derivatives in chapter 3; coefficients for model fitting

in chapter 5

d, e, f Coefficients for plane fitting partial derivatives in chapter 3

K Gaussian curvature [L-2]

k Tuning parameter in the phase evolution function; base (or absolute) permeability (mD)

κ Average fluid/fluid interfacial curvature [L-1]

kabs Absolute permeability (mD)

gk Geodesic curvature [L-1]

kj Effective permeability to phase j (mD)

kr-S Relative permeability—saturation

D

rnwk Nonwetting phase relative permeability during drainage

( )D

rnwnwfk S Nonwetting phase relative permeability during drainage at flowing nonwetting saturation

Page 17: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xvii

( )D

rnwnwik S Nonwetting phase relative permeability during drainage at initial nonwetting saturation

I

rnwk Nonwetting phase relative permeability during imbibition

o

rk End-point relative permeability

nonwettingr

k (or krnwet) Nonwetting phase relative permeability

krj (or r

k ) Relative permeability to phase j

refr

k Reference state relative permeability

krof Oil relative permeability of fixed data point for partial derivative estimation

wettingr

k (or krwet) Wetting phase relative permeability

Lw, Lnw Lengths of wetting and nonwetting phases in capillary tube (mm)

M0, M1, M2, M3 First [L3], second [L2], third [L], and fourth [-] Minkowski functionals

M0w, M0

nw First Minkowski functionals for wetting and nonwetting phases [L3]

M1(E), M2(E),

M3(E)

Second [L2], third [L], and fourth [-] Minkowski functionals for element E under

consideration

M1w, M1

nw Second Minkowski functionals for wetting and nonwetting phases [L2]

M2nw Third Minkowski functionals for wetting/nonwetting interface [L]

µj Viscosity of phase j (cP)

N Number of state parameters

NE Number of reciprocity relations

NCa Capillary number

NP Number of phases

n1, n2, nw, no Corey exponent for phase1, phase2, water, oil; n1 also used as the exponent for the IFT

scaling factor

np Number of pores

nt Number of throats

occp

n Number of occupied pores

occt

n Number of occupied throats

occht

n Number of occupied half-throats

O Number of objects

p Tuning parameter in the phase evolution function

Pc Capillary pressure (Pa)

Pj Pressure of phase j (Pa)

r Radius of capillary tube (mm)

R (R1, R2) Principal radii of curvatures [L]

ρj Density of phase j (kg/m3)

Page 18: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xviii

σ Interfacial tension (N/m)

σo Original interfacial tension (N/m)

σgw, σgo, σow Interfacial tensions of gas/water, gas/oil, and oil/water systems (N/m)

Se Effective phase saturation

*

gFS Reduced free-flowing gas phase saturation

*

grS IFT scaled residual gas saturation

Sgr Residual gas saturation

Si or Snwi Initial phase saturation (or initial nonwetting phase saturation)

Srj (or Sr or Snwr) Residual phase saturation of phase j; residual nonwetting phase saturation

Sj (or S ) Saturation of phase j

Sm Minimum wetting phase saturation

Snwf Flowing nonwetting phase saturation

Snwi Initial nonwetting phase saturation

Snwt Trapped nonwetting phase saturation

Snwet Nonwetting phase saturation

Sof Oil saturation of fixed data point for partial derivative estimation

Sorg Residual oil saturation to gas

*

orgS IFT scaled residual oil saturation to gas

Sref Reference state saturation

Swet Wetting phase saturation

Swir Initial wetting phase saturation

Swirr Irreducible water saturation

Sw,imb, Sw,dr Wetting phase saturation during imbibition and drainage

*

,w imbS , *

,w drS Reduced wetting phase saturation during imbibition and drainage

yi State parameter i

Z Coordination number

S ,

,

A Partial derivatives in the EOS

λ Pore structure (state parameter); also pore-size distribution parameter

λrwet, λrnwet Wetting and nonwetting phase tortuosity ratios

λwet, λnwet Wetting and nonwetting phase tortuosity factors

θ Contact angle, °

θgw, θgo, θow Contact angles for the gas/water/solid, gas/oil/solid, and oil/water/solid systems, °

θi ith corner angle of a rough surface element, i = 1, 2, 3…n

Average contact angle

Porosity

uj Interstitial velocity of phase j (ft/d)

Page 19: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xix

V Number of vertices in a polyhedra

Vb Bulk volume (cc)

Euler characteristic

(δE) Euler characteristic of entire surface of object E

(E) Euler characteristic of object E

ˆ i Normalized initial phase connectivity

0 Normalized phase connectivity at zero saturation

ˆ r Normalized residual phase connectivity

ˆj

(or ) Normalized connectivity of phase j

ˆ S − Connectivity—saturation

ˆ rr S − Locus of residual phase connectivity and residual saturation

χmax Maximum disconnected Euler characteristic of the pore space in a rock

χmax-image Maximum disconnected Euler characteristic of the pore space in a rock from image data

χmax-PNM Maximum disconnected Euler characteristic of the pore space in a rock from PNM data

χmin Maximum connected Euler characteristic of the pore space in a rock

χmin-image Maximum connected Euler characteristic of the pore space in a rock from image data

χmin-PNM Maximum connected Euler characteristic of the pore space in a rock from PNM data

χnonwetting-phase Euler number for nonwetting phase

χpore-structure Euler number for the pore space

ˆref

Reference state connectivity

ˆof

Normalized oil connectivity of fixed data point for partial derivative estimation

Ф Phase distribution

Фr Residual phase distribution

Фref Reference state phase distribution

Page 20: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

xx

ACKNOWLEDGEMENT

I would first like to express my sincere gratitude toward my advisers, Prof. Zuleima Karpyn and Prof.

Russell Johns for providing their never-ending support, and sharing their valuable time, ideas, and critiques

for the betterment of my project. Prof. Karpyn has been a wonderful teacher and a true mentor. She valued

my success as her own and provided me with opportunities to continually grow as an individual and as a

professional, and my life is enriched because of those experiences. Prof. Johns always pushed me one step

further and I will always be grateful to him for the learnings that happened in those moments. He taught me

that growth is possible in moments of discomfort and that it is important to feel comfortable when

experiencing discomfort and uncertainty to become a valuable researcher.

I would like to thank Prof. Martin Blunt for hosting me at Imperial College London, and for sharing his

time, resources, and expertise that added value to my project.

I would also like to thank my dissertation committee, for their time and their critical suggestions that

brought improvements to my project.

I would like to thank my friends and the extended EME family that has been home for the past several

years. It is with their support that such projects become possible.

Lastly, I do not think I can thank my family enough, especially my parents. My education would not

have been possible without their sacrifices and their unwavering support.

Acknowledgements of financial and resource support

Part of this work was performed in support of the National Energy Technology Laboratory’s ongoing

research under the RSS contract number 89243318CFE000003. The author gratefully acknowledges

financial support from the Energi Simulation Foundation, member companies of the Enhanced Oil

Recovery Joint Industry Project at Penn State, and the John and Willie Leone Family Department of Energy

and Mineral Engineering (EME). is also gratefully acknowledged. Additional funding support from the

Holleran and Bowman scholarship from the EME Department is also kindly acknowledged. Lastly, resource

support from the EME Department and the EMS Energy Institute at Penn State, and the Department of

Earth Science Engineering at Imperial College London are gratefully acknowledged.

Prakash Purswani

March 2021

Page 21: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

1

CHAPTER 1. INTRODUCTION

1.1. Background

Understanding fluid flow in a porous medium is at the heart of engineering disciplines such as petroleum

engineering. Single-phase flow in porous systems is less complex, but the introduction of an additional

immiscible phase creates a system with involved physics. Multiphase flow is found in a variety of technical

challenges facing society today. Some examples include sequestration of CO2 in geological formations to

mitigate greenhouse gas emissions, removal of nonaqueous chemicals for remediation of groundwater, and

secondary and tertiary recovery methods to sustain/improve energy production from hydrocarbon

reservoirs.

Simple waterflooding is a multiphase process that is deployed in the secondary phase of oil production.

The waterflooding process requires reinjection of produced water to sustain the production of oil after the

decline of high reservoir pressure. Enhanced oil recovery (EOR), however, involves the injection of an

agent such as a chemical or a gas to further oil production (Lake et al. 2014). EOR techniques are generally

employed during the tertiary phase of oil recovery when waterflooding is in its most matured stage, and oil

production is declining. One of the most common EOR method is the application of polymers to facilitate

favorable mobility ratio (close to 1) by resolving viscosity imbalances between the injecting phase, water,

and the receding phase, oil. Other common examples include the use of a surfactant to cause miscibility

between oil and water by reducing interfacial tension differences. Similarly, gases such as CO2 are often

used to improve oil recovery by oil swelling to cause the oil’s viscosity to reduce for easier recovery (Johns

and Orr 1996). Although the application of CO2 has been with the petroleum industry for decades, it is now

being increasingly favored as an important method for carbon capture purposes. This technique is referred

to as carbon capture utilization and storage (CCUS).

Page 22: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

2

Apart from utilization, CO2 storage known as carbon sequestration is also being ramped up to reduce

carbon emissions for tackling problems of growing global temperatures. A few pilot scale efforts for this

exist in the North Sea. CO2 sequestration is a multiphase process where CO2, primarily in the supercritical

state, is injected in geological formations such as depleted hydrocarbon reservoir, or aquifers for long-term

storage. In the process of injecting CO2, water (already present in the formation) is displaced. This is known

as the primary drainage process for displacement in a water-wet media. The term ‘primary’ is used to

reflect that the external phase (CO2 in this case) is introduced in the pore space for the first time. On the

trailing end of the CO2 plume, however, the CO2 phase experiences displacement by water and this process

is typical of an imbibition process, once again, given the rock formation is primarily water-wet. A 3-D pore-

scale visualization of the endpoint states of a saturated porous medium after primary drainage and

imbibition is shown in Figure 1.

Figure 1. 3-D visualization of fluid flow in porous media showing the endpoint states of primary drainage and imbibition processes. (Left), (middle), and (right) show the states of the medium before primary drainage, after the completion of primary drainage, and after the completion of imbibition, respectively. Figure adapted from Schlüter et al. (2016).

Solid

Pore

space filled

with 100%

wetting

phase

Nonwetting

phase

Page 23: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

3

Figure 1 is adapted from the x-ray imaging experiment by Schlüter et al. (2016). The porous medium is

a bead pack which is shown with transparency to allow for the visualization of the pore space. Figure 1

(left) represents the state just prior to primary drainage. At this point, the pore space is completely filled

with the wetting phase which is shown by the clear background. Figure 1 (middle) shows the end of the

primary drainage process, where the medium is occupied mostly by the nonwetting phase shown in green.

Empty spaces in the middle figure mark the occupancy by the wetting phase which is quantified as the

irreducible wetting phase saturation. Apart from CO2 injection into aquifers, in nature, the primary drainage

process occurs when the oil first migrates from the source rock to the oil reservoir (Blunt 2017). During the

natural density driven flow of oil, oil displaces brine which originally occupies the pore space. Figure 1

(right) shows the state of the medium at the end of the imbibition process. The majority of the medium is

now occupied by the wetting phase (clear) and the green blobs are the trapped nonwetting phase blobs,

which are quantified together as the residual nonwetting phase saturation.

Multiphase processes occur at different scales. From pore-scale (as shown in Figure 1), to core-scale

(example, laboratory corefloods), to the reservoir-scale. At the reservoir-scale, multiphase flow in

geological formations is modeled by using tools such as reservoir simulations. These are physics-based

models that take information from the field as well as laboratory experiments as inputs to make engineering

decisions. Reservoir simulations are critical as they allow for conducting ‘what-if’ scenarios to test diverse

set of conditions and thus help in mitigating risk and finding optimum engineering decisions. To capture

the physics of flow of multiple phases, simulations require transport properties such as relative

permeabilities and capillary pressures. These inputs are often estimated in the laboratory and calibrated for

use in the simulator for the process being simulated.

In this dissertation, the focus is on relative permeabilities, but discussions for capillary pressures is

provided wherever appropriate.

Page 24: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

4

1.2. Relative permeability

Relative permeability is the transport property that helps quantify the flow of multiple phases in a porous

medium. The relative permeability (krj) to a phase (j) is the ratio of the phase’s effective permeability (kj)

to the base (or absolute) permeability (k) of the medium. The effective permeability of a phase is defined

as the permeability of a phase at less than 100% occupancy in a porous medium saturated with multiple

phases. It estimated from extensions to the Darcy’s law and is calculated as,

( )j j zj

j

j

P gDku

x

−= −

, (1.1)

where uj is the flux of phase j; the ratio kj/µj is the proportionality constant termed as the mobility ratio of

phase j; µj and ρj are the viscosity and density of phase j, respectively; the terms inside the bracket, Pj -

ρjgDz, is the potential difference; Pj is the pressure of phase j; g is gravitational acceleration constant; Δx is

the finite distance within the porous medium; Dz is the elevation or depth from the reference datum.

The effective permeability to a fluid is a complex function of a variety of different factors, such as the

phase saturation, pore structure of the medium, wettability of the medium, the flow conditions such as the

interfacial tension between the involved phases, flow rate, and fluid viscosities, and the topology of the

individual phase. Each of these factors are described in more detail in the subsequent sections of this

chapter.

Figure 2 shows a schematic of relative permeabilities to oil and water typical for primary drainage and

primary imbibition processes. This schematic is representative of a water-wet media and other wettabilities

will be considered later. Primary drainage (shown in black) begins at 100% water saturation as the oil

invades the pore space. As water recedes the porous medium, its saturation and relative permeability drops,

while oil saturation and relative permeability to oil increases. The effect of saturation on relative

permeability is intuitive. The more a phase is present in the medium, that phase’s ability to move in the

Page 25: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

5

pore space increases. Thus, the relationship for relative permeability with saturation remains monotonic.

The curvature, the endpoint values on the relative permeability curve, and the endpoint saturations,

however, are distinct for the set of experimental conditions.

Figure 2. Schematic showing water-oil relative permeabilities for a water-wet medium. Hysteresis in relative permeabilities is also displayed. The black curves represent a primary drainage process, while the red curves represent a primary water injection process. The solid curves are water relative permeabilities, while the dashed curves are the oil relative permeabilities. Directions of flow are marked by arrows on the figure. Endpoint relative permeabilities and saturations are also marked on the figure. Adapted from (Blunt 2017).

Sw

1

0 1

k roo

Swirr

k rwo

1-Sor 551-Sor

krw (Primary drainage)kro (Primary drainage)krw (Primary imbibition)kro (Primary imbibition)

Page 26: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

6

The endpoint of primary drainage is marked by negligible change in the oil/water saturation. At this point,

the water saturation left in the medium is termed as the irreducible water saturation (Swirr), and the endpoint

oil relative permeability (kroo) can be estimated from the flow experiment via stabilized pressure and flow

rate measurements. The reverse cycle is the imbibition process (shown in red) where now water is injected

into the medium. This is essentially waterflooding the core to extract oil. As water is injected into the

medium, water saturation rises and so does the relative permeability to water. Now, oil saturation in the

core drops as more oil is recovered, and so does the relative permeability to oil. The endpoint saturation for

oil is called residual oil saturation (or residual nonwetting phase saturation or simply residual saturation).

On the water saturation axis, this is marked as 1-Sor. At this point, the endpoint relative permeability to

water (krwo) can be measured.

The choice of residual saturation is largely subjective. Experimentalists make their own decision to

terminate the flow experiment based on the reason for the experiment. For example, some may conclude

the experiment in the first few pore volume to get trends of early oil recovery, while others may extend for

hundreds of pore volumes (or with the application of centrifuge) to drive to the limits of residual saturation

possible. The ultimate trapped phase saturation depends on the pore structure together with the wettability

of the medium and not on the stopping criterion or flow conditions. We bring more context on the issue of

phase trapping in chapter 5 of this dissertation.

The endpoint relative permeabilities and the endpoint oil/water saturations are critical inputs for

reservoir simulations. In current modeling practices, these endpoint values, measured in the laboratory, are

used for calibration of the kr-S path, specific to the experimental conditions and fed as inputs to the reservoir

simulator. These models will be described in a later section.

Figure 2 shows that the kr-S drainage paths versus imbibition paths are quite distinct for both oil and

water relative permeabilities which shows that relative permeabilities are path dependent in the saturation

space. This conveys that relative permeability models that are calibrated for the drainage cycle cannot be

used for the imbibition cycle. This path dependency or nonuniqueness in relative permeabilities is termed

hysteresis and can cause numerical problems in reservoir simulations. For example, the use of drainage

Page 27: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

7

relative permeabilities over imbibition relative permeabilities can lead to incorrect estimation of oil

recoveries through EOR processes (Carlson 1981). Similarly, without consideration of hysteresis when

simulating carbon sequestration can lead to inaccurate estimates of CO2 migration and ultimate quantity of

trapped CO2 (Juanes et al. 2006).

One of the physical reasons considered responsible for hysteresis in relative permeabilities is associated

with phase trapping. The remaining phase at the end of one cycle of injection (for example, primary

drainage) is linked to possibilities of different reassociation of this phase when the cycle is reversed. This

results in a different flow path of the phase.

One other physical reason responsible for hysteresis in relative permeabilities is contact angle

hysteresis. Contact angle hysteresis is the difference between the contact angles measured at the three phase

(fluid/fluid/solid) contact point when the denser fluid is invading the medium (advancing contact angle)

versus when this fluid is retracting from the medium (receding contact angle). Experimental evidence shows

that advancing contact angle is greater than receding contact angle, and that both can be related to the

intrinsic contact angle, which is the contact angle measure of the static fluid/fluid/solid system on a clean

solid surface (Morrow 1975). Surface roughness is considered as the primary cause for hysteresis in contact

angle. No porous media, when considered at the microscopic scale would have perfectly clean surfaces.

These microscopic irregularities and the resultant entrapped fluid in the rough surface ridges are deemed as

the reasons for hysteresis in contact angles.

Another reason considered to cause relative permeability hysteresis, which is inherently associated with

contact angle hysteresis, is the different types of flow mechanisms that may occur when a phase is

advancing versus when the phase is receding. Piston-like advance, cooperative pore body filling, snap-off,

layer flow, and flow with bypass are some key types of flow mechanisms identified in the literature for

flow under capillary dominated regime (Lenormand and Zarcone 1984; Valvatne and Blunt 2004). Other

flow mechanisms such as drop traffic flow and ganglion dynamics are found to occur at high capillary

numbers (Avraam and Payatakes 1999; Avraam and Payatakes 1995; Rücker et al. 2015).

Page 28: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

8

Early efforts for resolving hysteresis in multiphase flow began with the treatment of capillary pressures.

It was identified that hysteresis occurs because of representation of transport properties strictly in the

saturation space. The hypothesis was that phase saturation alone cannot represent flow and that inclusion

of other pore-scale parameters is necessary toward resolving hysteresis. This hypothesis was proven in the

works by Hassanizadeh and Gray (1993) and Reeves and Celia (1996) where application of fluid/fluid

interfacial areas was included for addressing hysteresis.

Together with phase saturation, multiple other key controlling parameters such as phase connectivity,

fluid/fluid interfacial areas, wettability, capillary number, and pore structure have been recognized in the

literature to affect relative permeabilities. In the following subsections each of these parameters is described

in some detail.

1.2.1. Parameters affecting relative permeabilities

1.2.1.1. Fluid/fluid interfacial areas1

Integral geometry provides a means to quantify the structures of geometrical entities. Researchers have

adopted this approach to quantify connectivity measures of pore structures, as well as wetting and

nonwetting phases to evaluate multiphase flow in permeable systems. There are four useful measures from

integral geometry known as Minkowski integrals that describe the shape of a 3-D geometrical structure

(Armstrong et al. 2018; Blunt 2017; Wildenschild and Sheppard 2013). The first Minkowski (M0) functional

refers to the volume of the structure. For example, pore volume of a pore structure, or saturation of fluid

phases occupying the pore space. The second Minkowski functional (M1), however, corresponds to the

1 Parts of the text presented in this sub-section are published in Purswani et al. (2020) J. Pet. Sci. Eng.

Page 29: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

9

phase’s surface area (Ap). These, when divided by the bulk volume, result in specific surface area of the

phase (αp) (Landry et al. 2014; Landry et al. 2011),

 p

p

b

A

V = . (1.2)

The use of specific surface areas is recommended over actual surface areas because it removes system

dependence. This is similar to the use of specific solid surface areas for the estimation of base permeability

of a porous medium as observed in the development of the Carmen-Kozeny (CK) equation (Lake et al.

2014). The specific surface areas are then used for estimating the specific fluid/fluid interfacial area as

(Dalla et al. 2002),

1 2

.w nw w nw s

= + − (1.3)

Here, αp represents a phase-specific surface area, ( , ,s w

and nw

corresponding to the specific solid surface

area, specific wetting phase surface area, and specific nonwetting phase surface area, respectively); Vb is

the bulk volume of the medium; and αw-nw represents the fluid/fluid interfacial area.

The third Minkowski functional (M2) corresponds to the average curvature at the boundary between

two objects which is commonly used for quantifying local capillary pressure from two-phase image data

(Armstrong et al. 2018; Blunt 2017). Finally, the fourth Minkowski functional (M3) represents the integral

of the total Gauss curvature of an object which is related to the Euler characteristic (χ) as,

4

M

= . (1.4)

Page 30: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

10

The discussion on Euler characteristic is available in the next subsection and additional details and

examples for Minkowski functionals is available in appendix A.

Fluid/fluid interfacial areas have been an important measure to analyze multiphase flow in porous

media for a variety of applications (Culligan et al. 2004; Reeves and Celia 1996). One such application is

the nonaqueous phase liquid dissolution rates which is critical for evaluating environmental chemical

transport. Experimental attempts at measuring interfacial areas already exist, such as the oil/water and

air/water fluid/fluid interfacial area measurements in a sand pack column (Saripalli et al. 1997b; Saripalli

et al. 1997a). These measurements were made by the application of surface-reactive tracers which

selectively adsorb at the fluid/fluid interface and cause its retardation during the miscible displacement

experiment. Although such independent methods exist, the application of x-ray imaging into fluid flow

research has allowed for superior ways of quantifying fluid surface areas and interfacial areas (Blunt 2017;

Culligan et al. 2005; Culligan et al. 2004; Dalla et al. 2002; Landry et al. 2014; Landry et al. 2011).

Sophisticated visualization of the trapped fluid phases and improved algorithms for estimating surface areas

of voxelated entities have enabled the estimation of the total surface areas of the various phases inside of a

porous medium. These can then be used toward estimating interfacial areas among different phase pairs.

A schematic 2-D visualization of a saturated porous media with trapped wetting and nonwetting phases

at two different saturations is shown in Figure 3. The solid, the wetting, and the nonwetting phases are

shown by the gray, blue, and green colors, respectively. The corresponding phase pair interfacial contact

lines are also displayed. Figure 3 (right) can be understood as a snapshot following a saturation step-change

during a typical imbibition process. In a 3-D representation, these contact lines would represent the surface

areas of contact. As the respective phase saturations change, the respective total surface areas of the fluid

phases change, consequently, the fluid/fluid interfacial areas change (qualitatively expressed in Figure 3)

and thus provide a pore-scale measure of the movement of a fluid during multiphase flow.

Page 31: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

11

Figure 3. Schematic representation of a 2-D multiphase porous system showing possible phase/phase contact lines. Left to right the wetting phase saturation increases. When considered in 3-D, phase/phase interfacial areas would be estimated along the areal region of contact. Adapted from Dalla et al. (2002).

1.2.1.2. Phase connectivity2

Figure 4 shows 3-D representations of the nonwetting (left) and wetting (right) phases inside of a saturated

porous medium. These 3-D renderings were acquired using x-ray imaging at static experimental conditions.

More details on the experiment are available in chapter 2 of this dissertation. The disconnected nature of

the nonwetting phase and the connected nature of the wetting phase can be easily visualized. The

nonwetting phase appears as isolated blobs or clusters, while the wetting phase appears more continuous.

2 Parts of the text presented in this sub-section are published in Purswani et al. (2019), Comput. Geosci.

Solid phase

Non-wetting phase

Wetting phase

Wetting phase/solid surface area

Non-wetting phase/solid surface area

Non-wetting phase/wetting phase surface area

Page 32: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

12

Figure 4. Segmented image of a multiphase system acquired using x-ray imaging showing (a) disconnected nonwetting phase, and (b) more connected wetting phase.

Finding a unique mathematical definition for connectivity in porous media has been an active point of

research (Aydogan and Hyttinen 2013). There are a number of connectivity parameters proposed in the

literature such as the Euler characteristic (Vogel 2002), percolation theory (Hovadik and Larue 2007),

connectivity function (Allard 1993), contour tree connectivity (Aydogan and Hyttinen 2013), coordination

number, and fractal dimension (Blunt 2017). Out of these measures, the Euler characteristic () has been

the simplest and most widely used measure of connectivity in porous media (Allard 1993; Aydogan and

Hyttinen 2014).

The Euler number identifies phase connectivity by considering the number of clusters and the number

of connections for these clusters. The Euler number decreases with an increase in the number of clusters.

Euler characteristic is a topological invariant originally proposed by Leonhard Euler for a polyhedra as the

alternating sum of vertices (V), edges (E), faces (F), and objects (O) and is computed as (Richeson 2008),

 V E F O = − + − . (1.5)

a b

Page 33: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

13

Extending the concept to complex phase structures, the Euler Poincaré formula has been widely used

for quantifying connectivity of microstructures as,

0 1 2  = − + , (1.6)

where the parameters 0 1 2, ,  and  are the zeroth, first, and second Betti numbers, respectively.

0

represents the number of clusters, 1 is the number of holes or redundant loops (the maximum number of

breaks that can be made without having the cluster split into two as explained by Herring et al. 2013), and

2 is the number of enclosed voids. 2 is usually considered to be zero for the calculation of the Euler

characteristic (connectivity) of both the wetting and the nonwetting phases. While this may be true for the

nonwetting phase, since there can be no solid grains or wetting phase globules suspended in a continuous

nonwetting phase in a consolidated porous medium, this may not be true when calculating the Euler

characteristic of the wetting phase, where suspended nonwetting phase globules can occur within a

continuous wetting phase. Euler numbers range from   to − + where a highly connected phase has a large

negative value while a highly disconnected phase has a large positive value.

Recent studies conducted using x-ray micro computed tomography (micro-CT) advocate for the use of

either both fluid/fluid interfacial area as well as the Euler characteristic (Mcclure et al. 2018; Mcclure et al.

2016) or suggest use of just the Euler characteristic (Schlüter et al. 2016). In Mcclure et al. 2018 and

Mcclure et al. (2016) it was shown that by including all Minkowski functional (i.e., both Euler characteristic

and fluid/fluid interfacial areas together with saturation) nearly all of the hysteresis observed in capillary

pressure measurements could be accounted for successfully. However, the authors fail to show capillary

pressure predictions by including just phase saturation and Euler characteristic. It is likely that the majority

of the hysteresis could still be captured without the need for the interfacial area measurements. Further, for

modeling purposes, it is suitable to minimize the total number of variables involved, such that the physics

Page 34: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

14

of the problem is reasonably captured and at the same time the overall computational complexity is

minimized. Therefore, in this work, we use Euler characteristic as the measure of phase connectivity.

Pore-scale fluid properties such as fluid/fluid interfacial areas and fluid connectivity (measured through

the Euler characteristic of a fluid phase) are measures that describe the flow of phases within the porous

medium. The porous media properties such as porosity, permeability, and tortuosity, however, help define

the representative elementary volume (REV) over which the continuum assumption holds valid. Thus, for

consistency, the size of the extracted sub volume for pore-scale analysis and property estimation should be

sufficiently large that it is equal to or greater than the porous media-defined-REV.

1.2.1.3. Capillary number

Capillary number is a dimensionless number that is described as the ratio of viscous forces to interfacial

forces. The use of capillary number allows to account for important factors that affect relative

permeabilities, namely, the interfacial tension, fluid viscosity, and the flood rate. It is usually calculated as

(Lake et al. 2014),

,Ca

uN

= (1.7)

where u is the interstitial velocity; µ is the viscosity of the injecting phase; and σ is the interfacial tension

between the flowing phases. Some researchers also include wettability (with the application of the term

cosθ) or the porosity into the definition of capillary number as multiplication factors in the denominator of

Eq. (1.7).

The importance of capillary numbers can be understood from capillary desaturation curves which show

an inverted S-shaped relationship between capillary number and the residual saturation (Lake et al. 2014).

The curve is specific to the experimental conditions of rock and fluids used. It shows that residual

saturations can be reduced significantly at very high capillary numbers. The goal for EOR processes is thus

Page 35: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

15

to achieve high capillary numbers to attain low residual oil saturations. High capillary numbers are

accomplished through either high flood rates (within the engineering design to maintain reasonable

injectivity) or through the use of high viscosity polymers to increase viscous forces, or with the application

of surfactants to reduce interfacial tensions (and consequently reduce interfacial forces).

Two types of flow regimes, namely, capillary dominated versus viscous dominated flow can be

identified from the capillary desaturation curves. The threshold is marked around capillary numbers of 10-

4. Below this threshold, capillary dominated regime occurs which is attained with low flood rate and/or high

interfacial tension conditions. Flow near the wellbore would experience high flow rates as opposed to flow

far away from the wellbore, where the flow is capillary dominated, and thus relative permeabilities in these

regions will be different. Hence, understanding of the full range of capillary number is important. This is

critical for modeling carbon sequestration where storage in the formation is ensured via the capillary

trapping mechanism.

The kr-S paths for different capillary numbers are inherently different. From the experiments by

(Delshad et al. 1987; Fulcher et al. 1985) it was shown that relative permeability paths are straighter (x-

shaped) for high capillary numbers (low interfacial tensions) while the paths are concave for low capillary

numbers. In the work by Fulcher et al. (1985), the effect of capillary number on two-phase relative

permeabilities was investigated through steady-state experiments by considering viscosity and interfacial

tension effects independently. Both wetting and nonwetting phase relative permeabilities as well as residual

saturations were found to be significantly affected by both interfacial tensions and viscosity changes

(Fulcher et al. 1985). The overall impact of capillary number (as a group) was more significant on wetting

phase flow than on nonwetting phase flow. The impact of capillary number on multiphase flow has also

been demonstrated at the pore-scale with two-phase simulations by Armstrong et al. (2016). The

implications of capillary number on relative permeability and modeling efforts are provided in chapter 3 of

this dissertation.

Page 36: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

16

1.2.1.4. Wettability

Wettability is defined as the ability of a solid surface to have preferential affinity to one phase in the

presence of another phase (Anderson 1986a). It is a property of the porous medium. Contact angle

measurement provide the most accurate measure of wettability. It is conventionally measured through the

denser phase. For example, for an oil/water/rock system the contact angle is measured through the water

phase. Figure 5 shows a schematic of a water-wetting solid in the presence of oil. From the Young’s

equation, the balance of interfacial forces, gives the measure of wettability (θ) as follows,

,-ws os

ow

cos

= (1.8)

where σow, σos, and, σws are the oil/water, oil/solid, and water/solid interfacial tensions, respectively.

Figure 5. Schematic showing a water-wet solid in the presence of oil.

There are different types of wettabilities. With respect to oil/water/rock system, the rock may be

characterized as water-wet (θ = 0°-75°), intermediate-wet (θ = 75°-105°), or oil-wet (θ = 105°-180°). These

rough estimates were provided through experiments by Treiber et al. (1972). Complete water-wetness and

complete oil-wetness would occur at θ = 0° and θ = 180°, respectively. Figure 6 shows a schematic of

different rock wettabilities.

Solid

OIL

WATER

Page 37: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

17

Figure 6. Illustrations of different types of wettability for oil/water/solid systems. From left to right the medium’s wettability to oil increases.

Other important classifications of wettability include fractional-wettability and mixed-wettability. Both

these types include parts of the porous medium that may be oil-wet and other parts that may be water-wet.

The difference between the two lies in the way these wettabilities are developed in a porous medium.

Fractional wettability is often developed in unconsolidated porous media by mixing solid grains of different

wettability types (for example, plastic beads—oil-wet and glass beads—water-wet)(Klise et al. 2016;

Landry et al. 2014; Landry et al. 2011). When these grains are mixed and compressed to form one medium,

there are pores that are completely oil-wet versus pores that are completely water-wet.

On the contrary, mixed-wettability is developed in a porous medium due to the process of aging the

rock sample. As such, most oil reservoirs are naturally mixed-wet. In a laboratory, the aging process is

carried out after primary drainage when an initial oil saturation has been established in the rock. Aging

requires subjecting the rock sample to high temperatures and pressures for a period of time. These

conditions are subjective and depend on the experimentalist and are set based on the test requirements. At

initial oil saturation conditions, parts of the rock are in direct contact with the oil and yet other parts are in

contact with water which is present in irreducible amount. During aging, the portions of the rock in contact

with oil are said to become oil-wet. This leads to the generation of mixed-wettability where within the same

pore, parts can be oil-wet and parts can be water-wet (for example, corners of the pore space occupied by

water will remain water-wet). The degree of wettability alteration depends on the aging conditions. Mixed-

wettability was first coined by Salathiel (1973).

Neutral wettingWetting Nonwetting

OIL

Solid

WATER

Completely water-wet Water-wet Neutral-wet Oil-wet Completely oil-wet

Page 38: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

18

Wettability of a medium is often characterized qualitatively using relative permeability curves. A

schematic for oil/water relative permeabilities during water injection are shown for three types of

wettability in Figure 7. The endpoint water relative permeabilities and residual oil saturations are marked

for the three scenarios.

Figure 7. Schematic showing three types of rock wettabilities characterized qualitatively from the visualization of two-phase relative permeability curves. Here, the two phases are oil and water and the flow (direction marked by the arrow) represent water injection. The black, green, and red curves are for an oil-wet, a water-wet, and a mixed-wet rock, respectively.

1

Sw

0 1

kr

flow

krwo

2

krwo

3

1-Sor1 1-Sor3

1-Sor2

krwo

1

Page 39: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

19

The endpoint values of water relative permeabilities and residual oil saturation together with the cross-

point saturations where the oil and water relative permeability curves intersect are used as qualitative cues

for wettability assessment of the rock from a flow experiment. Typically, for a water-wet rock (green curves

in Figure 7), the cross-point saturation is greater than 0.5, and the endpoint water relative permeability is

low ~ 0.2 and can go lower than 0.05 for extremely water-wet media (Blunt 2017; Lake et al. 2014). This

is because water is wetting the surface and would consequently occupy the smaller regions of the pore space

such as the pore corners or crevices. Thus, the water conductance remains low. For an oil-wet rock (black

curves in Figure 7), however, water would occupy the centers of the pore space which leads to higher water

conductance and consequently high endpoint water relative permeabilities (~0.5). Also, for oil-wet rocks,

the cross-point saturation is typically lower than for water-wet rocks.

For mixed-wettability, endpoint water relative permeability remains higher than the water-wet case.

Interestingly, experiments have shown that the residual oil saturation for the mixed-wet case is often the

lowest (Jadhunandan and Morrow 1995; Salathiel 1973). The reasons for this observation are still in debate

in the literature, but one of the hypotheses is that mixed-wettability provides for continuous pathways for

oil to flow in the medium and if flow experiments are prolonged over long periods, oil trapping can be

minimized significantly. Discussions on wettability and its consequences to phase trapping are presented

in detail in chapter 4 of this dissertation.

1.2.1.5. Pore structure

It has always been a challenge to quantify pore structure information. No single metric exists in the

literature. Pore structure differs not just from one type of formation to another, but also from one medium

to another even from the same formation. This is because of heterogeneity that exists in natural geological

formations (Lake et al. 2014). No two naturally occurring porous media will be exactly alike. Some of the

most used porous media properties that give insight into the pore structure are the permeability of the rock,

Page 40: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

20

which describes the ability of a porous medium to transmit fluids, and the porosity of the medium, which

describes the ability of the medium to store fluids. The ratio of the square root of permeability to porosity

is often used as a quantitative measure for characterizing pore structure information. Other measures often

used for modeling base permeabilities, are the tortuosity and specific surface areas of the solid surface.

Tortuosity, a dimensionless porous medium property, is defined as the square of the ratio of capillary tube

length to the length of the representative elementary volume (Lake et al. 2014), which is essentially the

squared ratio of path length traversed by the fluid in the pore space to the length of the porous medium.

Other quantitative measures most frequently used for characterizing pore structure information are the

distribution of pore and grain sizes. Pore-size distributions are experimentally measured though the use of

Mercury (Hg) intrusion porosimetry where a primary drainage capillary pressure curve is generated for a

Hg/air/rock system. Hg is injected as the nonwetting phase into the medium to increasingly high capillary

pressures to enter smaller sized pores. Through such experimentation, information on the pore sizes (and

average pore sizes) for the medium is extracted in the form of a frequency distribution plot. A pore-size

distribution parameter is often set as the calibrating exponent for capillary pressure and relative

permeability curves (Brooks and Corey 1964; van Genuchten 1980; Land 1968).

Measures for the pore structure such as the pore-size distribution, or the square root of permeability

over porosity are bulk (or average) measures for a porous medium. Techniques like x-ray imaging, flow in

micromodels, and pore-network extraction models provide other quantitative measures for characterizing

pore structures by taking information at the pore-scale (Blunt 2017; Fatt 1956; Lenormand and Zarcone

1984). These include the coordination number, aspect ratio, pore topology, geometric shape factor (or the

distribution of the geometric shape factors).

Coordination number and pore topology give direct information on the connectivity of the pore space.

Coordination number is defined for a pore as the average number of throats that are in direct connection to

a pore (Blunt 2017), whereas the topology of the pore space is estimated as the Euler characteristic of the

pore space which can be represented for the entire region of interest by normalizing with respect to the bulk

or pore volume of the region.

Page 41: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

21

Individual pore element sizes play an important role in characterizing the pore structure information.

For example, the pore-body to pore-throat aspect ratio is found to cause hysteresis in relative permeabilities.

High aspect ratios are linked to increased trapping of the nonwetting phase due to increased snap-off events

during imbibition (Jerauld and Salter 1990). The geometric shape factor, however, provides information of

the shape of the pore/throat elements and is defined as the ratio of the cross-sectional area of an element to

the square of its perimeter. If all pore/throat elements of a porous medium were circular and uniform, there

would be no phase trapping in the pore space since all elements will be drained completely by the invading

phase. Therefore, for simulation techniques such as pore-network modeling the shape factors becomes

critical as it helps in attaining pore structures with noncircular (polygonal) network elements that allow for

trapping of phases in pore corners. Availability of polygonal-shaped network elements also allow layer

flow where the phase may be connected through the corners.

In this dissertation, we keep pore structure information constant for the sets of simulations performed

for numerical data set generation. This is critical for the state function approach of modeling relative

permeability. Simultaneous efforts have been on going in our research group to develop state function-

based models for characterizing the base permeability of a porous medium with the knowledge of the

different pore structure metrics.

1.2.2. Models for relative permeability

1.2.2.1. Corey-type models

Initial efforts for modeling relative permeability were presented by Purcell (1949) using a bundle of

capillary tubes, and Burdine (1953) with the application of capillary pressure curves and the tortuosity

parameter. Burdine’s equations for wetting/nonwetting phase relative permeabilities were expressed as,

Page 42: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

22

( )

2

2 0

12

0

/

/

wetS

c

rwet rwet

c

dS Pk

dS P

=

, (1.9)

( )

12

2

12

0

/

/

wet

cS

rnwet rnwet

c

dS P

k

dS P

=

, (1.10)

where krwet and krnwet are the relative permeabilities to the wetting and nonwetting phases, respectively; λrwet

(=λ/λwet) and λrnwet (=λ/λnwet) are the wetting and nonwetting phase tortuosity ratios, respectively; λ is the

porous medium tortuosity factor; λwet and λnwet are the wetting and nonwetting phase tortuosity factors,

respectively; Sm is the minimum phase saturations; and Pc is the capillary pressure.

Burdine provided simplified saturation-based expressions for the wetting and nonwetting phase

tortuosity factors as,

1 m

wer

t mwet

S S

S

−=

− (1.11)

1 m

nw

r

et nwrrnwet

nw

S S

S S

− −

−= (1.12)

where Swet and Snwet are the wetting and nonwetting phase saturations, respectively; Sm and Snwr are the

minimum wetting and residual nonwetting phase saturation, respectively.

Corey (1954) extended Burdine’s equation by approximating,

( )

2

   1

0   

o or o or

o orc

C S S for S S

for S SP

− =

, (1.13)

Page 43: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

23

where ( )  / 1 ,orCC S= −  C is a constant. The following relative permeability equations were presented by

Corey (1954).

4

1

o or

ro

or

S Sk

S

−=

, (1.14)

2 2

1 11

o or o orrg

m or or

S S S Sk

S S S

− − = − −

− −

. (1.15)

Extension to Corey’s equations were presented by Brooks and Corey (1964) in a more general form to

estimate wetting and nonwetting phase relative permeabilities as follows,

( )2 3

rwet ek S

+

= , (1.16)

( )2

21 1rnwet e ek S S

+ = − −

, (1.17)

where Se, known as the effective phase saturation, is defined as ( ) ( )/ 1r rS S S− − ; Sr is the residual phase

saturation; the parameter, λ, is the pore-size distribution index.

These Corey models were further simplified and generalized as exponential models as follows,

1

1 1

1 12 21

n

p rporp rp

rp rp

S Sk k

S S

−=

− −

, (1.18)

Page 44: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

24

where krp1 and 1orpk are relative permeability and endpoint relative permeability to phase1; Sp1, Srp1, and Srp2

represent the saturation of phase1, endpoint saturation of phase1, and endpoint saturation of phase2,

respectively; n1 is the tuning exponent. This expression for water/oil relative permeabilities during

waterflooding then are,

1

wn

o w wirrrw rw

or wirr

S Sk k

S S

−=

− − , (1.19)

1

on

o o orro ro

or wirr

S Sk k

S S

−=

− − , (1.20)

where krw and kro are the water and oil relative permeabilities during waterflooding. orwk is the endpoint

relative permeability to water and Sor is the residual oil saturation which are determined at the end of

waterflooding, whereas orok is the endpoint relative permeability to oil and Swirr is the irreducible water

saturation which are determined at the end of oilflooding (prior to waterflooding). no and nw are the tuning

exponents. Similar to the exponential model other empirical models are available in the literature for

specific set of operating conditions (Fulcher et al. 1985; Honarpour et al. 1982). See Table 1 for empirical

expressions of oil/water relative permeabilities from the literature. For additional saturation-based models

for relative permeability the reader is referred to Honarpour et al. (1986).

Three-phase relative permeability models are not discussed in detail here, but some of the more

commonly used three-phase models include, Naar and Wygal (1961); Stone I (Stone 1970), Stone II (Stone

1973), and Land’s model (Land 1968). A comprehensive comparison of these models against three-phase

experimental data was summarized by Abder (1981).

Page 45: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

25

Table 1. Empirical oil/water relative permeability functions for different rock types and wetting conditions from the literature.

Reference Rock type Wettability Injection type

Honarpour et al. (1982) Sandstone and conglomerate Water-wet Water injection

( )2.9

3.60.035388 0.010874 0.565561 1

w wi w orrw w w wi

wi or wi or

S S S Sk S S S

S S S S

− −= − + −

− − − − (1.21)

Honarpour et al. (1982) Sandstone and conglomerate Oil/intermediate-wet Water injection

( ) ( )( )1.91

1.5814 0.58617 1.2484 11 1

w wi w orrw w wi wi w wi

wi wi or

S S S Sk S S S S S

S S S

− −= − − − − −

− − − (1.22)

Honarpour et al. (1982) Sandstone and conglomerate All Water injection

( )( )

1.8

21

0.76067 2.6318 11 1

oor

wi o orro or o or

or wi or

SS

S S Sk S S S

S S S

− − = + − − − − −

(1.23)

Honarpour et al. (1982) Limestone and dolomite Water-wet Water injection

( )0.43

2.15

10.0020525 0.051371w wi

rw w wia

S Sk S S

k

−= − −

(1.24)

Honarpour et al. (1982) Limestone and dolomite Oil/intermediate-wet Water injection

( )2 4

0.29986 0.32797 0.4132591 1 1

w wi w or w wirw w wi

wi wi or wi or

S S S S S Sk S S

S S S S S

− − −= − − +

− − − − − (1.25)

Honarpour et al. (1982) Limestone and dolomite All Water injection

2

1.26241 1

o or o orro

or wi or

S S S Sk

S S S

− −=

− − − (1.26)

Page 46: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

26

Fulcher et al. (1985) Berea sandstone Water-wet Oil injection

( )( ) 

D

B C ln woro dr

o

k AS

+ =

(1.27)

( )

 

11

w

o

B D ln

w wirrrw dr

wirr

S Sk A

S

+

−=

− (1.28)

( )

( ) 

21

caB D ln N

w wirrrw dr

wirr

S Sk A

S

+ −=

− (1.29)

Fulcher et al. (1985) Berea sandstone Water-wet Water injection

( )

   

1

w

o

B C ln D ln

o orro im

or

S Sk A

S

+ +

−=

− (1.30)

( )

 

11

w

o

B D ln

w wirrrw im

wirr

S Sk A

S

+

−=

− (1.31)

( )

( ) 

21

caB D ln N

w wirrrw im

wirr

S Sk A

S

+ −=

− (1.32)

Residual saturations, endpoint relative permeabilities, and Corey exponents are used for tuning a

specific kr-S path. These empirical models provide good match in most cases, but the challenge with such

representation of relative permeabilities is that the information about the pore structure, wettability, and

capillary number are all incorporated into the tuning exponents. Each kr-S path would thus be distinct and

not generalizable. As such, hysteresis in relative permeability is not resolved. For this, different researchers

have attempted different modeling solutions. Some of the commonly known hysteresis models are

described next.

Table 1. (Continued)

Page 47: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

27

1.2.2.2. Land-type models

Figure 8 shows a schematic displaying hysteresis in nonwetting phase relative permeabilities. Hysteresis in

relative permeability, as described previously, is the path dependency of relative permeability in the

saturation space. The black curves show nonwetting phase relative permeability during primary drainage

while the red curve shows the nonwetting phase relative permeability during imbibition. The imbibition

process is begun at the initial nonwetting phase saturation (Snwi) and ends at the residual nonwetting phase

saturation (Snwr). This set of primary drainage and imbibition forms one set of scanning relative permeability

curves. Other such sets of scanning curves can be generated experimentally, each with its own starting and

ending nonwetting phase saturation.

Page 48: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

28

Figure 8. Schematic of nonwetting phase relative permeability showing hysteresis after flow reversal from primary drainage (black curve) to primary water injection (red curve). The initial, trapped, residual, and flowing nonwetting phase saturations are marked on the figure. Adapted from Carlson (1981).

Naar and Henderson (1961) proposed a model for imbibition relative permeability by considering the

trapped nonwetting phase saturation during the imbibition process. They developed the following

relationship between the imbibition and drainage saturations for the same value of nonwetting phase relative

permeability.

*, , ,

* * 20.5( ),w imb w dr w drS S S= − (1.33)

Snwt

1

Snw

0 1

krn

w

Snwi

Snwr Snw

Snwf

Page 49: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

29

where *,, ) / ( ),( 1w imbi wi iw wmbS S S S= − − is the reduced wetting phase saturation during an imbibition process;

Swi is the initial wetting phase saturation (or the irreducible wetting phase saturation) prior to the imbibition

process; Sw,imb is the wetting phase saturation during imbibition; *

,, ( 1) / ( ),w dr w dr wi wiS S S S= − − is the reduced

wetting phase saturation during a drainage process; Sw,dr is the wetting phase saturation during

drainage. Further, by using Eq. (1.33), the following model for nonwetting phase imbibition relative

permeabilities ( Irnwk ) could be established by using the saturation information of the prior drainage

process,

( ), * 2,

*

0.5,

0.51I

rnww imb

w imb

Sk S

− = −

(1.34)

Land (1968) observed trends of characteristic initial-residual (IR) saturation curves and proposed a

relationship between Snwr and Snwi as follows,

1 1

nwr nwi

CS S

− = . (1.35)

The nonwetting phase relative permeability during imbibition is estimated by extracting information of

flowing (Snwf) and trapped saturation (Snwt). On any point on the krnw-Snw path (see Figure 8), following

relationship exist among the different saturations,

 nw nwt nwfS S S= + . (1.36)

Using Eq. (1.35) and Eq. (1.36), Snwf is estimated as follows (derivation available in Carlson 1981,

originally equation was presented in Land 1968),

Page 50: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

30

( ) ( ) ( )21 4

 2

nwf nw nwr nw nwr nw nwrS S S S S S SC

= − + − + −

. (1.37)

Land (1968) then followed similar treatment as that of Corey (1954) model to propose the following relation

for gas relative permeability with the use of the flowing gas saturation,

( ) ( )2

2* *1 1g grg F Fk S S

+ = − −

, (1.38)

where *gFS is the free-flowing gas phase saturation which is normalized to the effective pore space,

* / ( )1gF wirrgFS S S−= ; and λ is the pore-size distribution index.

Killough (1976) proposed an interpolation-based approach for estimating imbibition relative

permeabilities based on drainage relative permeabilities as follows,

( ) ( ) I D nw nwrrnw nw rnw nwi

nwi nwr

S Sk S k S

S S

= −

, (1.39)

where ( )Irnw nwk S is the imbibition relative permeability and ( )D

rnw nwik S drainage relative permeability at

initial nonwetting phase saturation. This proposed form satisfies the limiting conditions for imbibition

relative permeabilities, where ( ) ( )I Drnw nwi rnw nwik S k S= and ( ) 0I

rnw nwrk S = . Carlson further simplified

Killough’s approach and proposed that ( )Irnw nwk S can be estimated from the corresponding drainage

nonwetting phase relative permeability (Drnwk ) with the knowledge of the flowing nonwetting phase

saturation,

Page 51: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

31

( ) ( ) I Drnw nw rnw nwfk S k S= . (1.40)

Both Killough’s and Carlson’s model require the knowledge of the Land’s trapping coefficient.

Thus far, most relative permeability models are developed for water-wet media. The trapping model by

Spiteri et al. (2008) was an improvement over Land’s model. It was the first model to be used for different

wettabilities. It was developed using initial-residual (IR) trapping data sets generated using pore-network

modeling. Their model was given as,

2

r i iS S S = − , (1.41)

where α and β are model parameters. Similar to the treatments by Carlson (1981), based on the information

of the trapped versus flowing saturation and the previous primary drainage curves, relative permeabilities

for the waterflooding cycle for different contact angles can be calculated.

1.2.2.3. Limitations of relative permeability models

Some limitations of the models described for relative permeability in the previous subsections are as

follows.

• The major limitation of Land-based models is that the path dependency for relative permeability is

not resolved. Based on the information of the trapped phase, the relative permeabilities for the water

injection cycle are predicted which inherently depends on the particular scanning curve. Each set

of scanning curve will consequently have its own initial and residual saturation and its own path to

be traced. And each path will require tuning of its own Land’s trapping coefficient, for example,

Page 52: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

32

for the set of flow conditions that process is subjected to. This leads to ad hoc combinations of

Corey model with Land’s trapping coefficient for calculating relative permeabilities.

• These models are empirical and lack pore-scale physics, despite the understanding of the different

factors that affect relative permeabilities. These models are therefore less predictive away from the

conditions under which they are developed.

• Most Land-based models are developed for water-wet systems and are primarily used for predicting

imbibition relative permeabilities.

• These models are deficient because they require knowledge of previous drainage curve to calculate

imbibition relative permeabilities. In addition, they require knowledge of different inputs such as

the initial and residual nonwetting phase saturations.

Page 53: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

33

1.3. Research objectives

Current relative permeability (kr) models are functions of phase saturations that are matched for specific

flow/experimental conditions. However, as examined from the literature, together with phase saturation,

multiple parameters affect relative permeabilities such as the wettability of the medium, capillary number,

pore structure, fluid phase topology, and fluid/fluid interfacial areas. These other parameters affecting

relative permeabilities are inherently captured through the empirical saturation functions. Representation

of relative permeabilities only in the saturation space causes non-uniqueness and path dependency in

relative permeabilities which often cause simulations to fail because they lack generality and are not

physically based. As a result, hysteresis in relative permeabilities arises, which is a major modeling issue

for reservoir simulations.

Efforts have been presented in this dissertation to model relative permeabilities by considering

functional forms that include the effects of the key controlling parameters on relative permeabilities. The

purpose of this dissertation is twofold, to

(a) understand how different parameters, specifically, phase saturation, phase connectivity, capillary

number, and wettability affect relative permeabilities;

(b) propose physically-based kr models by including the effects of these parameters.

Page 54: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

34

1.4. Dissertation layout

There are five additional chapters in this dissertation after the introductory chapter. For chapters 2 to 5, an

abstract, relevant literature survey, methodology, results and discussions, and conclusions are presented.

Below a brief summary for each chapter is presented.

In chapter 2, a static x-ray imaging experiment of a multiphase system is discussed to quantify

measurement-based errors due to image segmentation. A high-resolution (6 µm) and a low-resolution (18

µm) x-ray scan of the same system was acquired. The high-resolution scan was used as ground truth while

the low-resolution scan was used to test different image segmentation methods and quantify errors in pore-

scale measurements. It was found that pore-scale measures of phase topology and fluid/fluid interfacial

areas are highly sensitive to image analyses procedures such as that of image segmentation. To mitigate

these errors, images with high-resolution should be acquired and these should be obtained in steps to

improve the accuracy of image segmentation. In addition, supervised machine learning based-algorithm

was found to provide the closest pore-scale measures to the ground truth. From this work, the need to

supplement experimental data sets with numerical data sets was identified.

In chapter 3, a state function-based approach for relative permeabilities is discussed. A relative

permeability equation-of-state (kr-EOS) is forced as a quadratic response for kr in the phase connectivity-

phase saturation space ( ˆ ).S − The EOS is constrained to limiting conditions in the ˆ S − space. Although

the model is built for fixed capillary number conditions, it is tested for different capillary numbers, ranging

from one to 10-6. The dependence of phase connectivity on capillary number is also explored. It was found

that a quadratic response for relative permeabilities work across different capillary numbers. The linear kr-

S paths for high capillary numbers (small Corey exponents) and nonlinear kr-S paths for low capillary

numbers (high Corey exponents) were found to occur due to fast and slow changes in phase connectivity,

respectively. From this work, the need for large numerical data sets to calculate relative permeability partial

derivates for the EOS development was identified.

Page 55: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

35

In chapter 4, numerical data sets of phase saturation and phase connectivity are generated using pore-

network simulations to study the effect of wettability on phase trapping. During primary drainage, the

contact angle was set at zero degrees. However, during secondary injection process, the contact angles were

changed from 0° to 180°. Trends of residual phase saturation and residual phase connectivity are analyzed

for different contact angles. Hysteresis trapping models are presented to capture the residual trends and

comparison is presented against models from the literature. It was found that wettability significantly affects

receding phase trapping and that pore-scale mechanisms of layer flow and piston-like advance of the

invading phase become critical when the receding phase is wetting to the surface.

In chapter 5, the workflow of pore-network simulations from chapter 4 is utilized to generate numerical

data sets of nonwetting phase relative permeability, saturation, and connectivity. Here, capillary number

and pore structures were kept fixed, and two wettability cases were considered both in the water-wet regime.

Through hundreds of simulations, the kr, S, and data sets are analyzed to estimate partial derivates of kr

in the ˆ S − space. These partial derivatives are then utilized for the development of an EOS response for

relative permeability. It is found that the EOS predicts kr for the entire data set, regardless of the direction

of flow, thus resolving hysteresis in relative permeabilities.

In chapter 6, key concluding remarks from this study and outlook for future research efforts that can be

built from this dissertation are presented.

Page 56: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

36

1.5. Publication list

Peer reviewed publications

• Purswani P., Johns R.T., Karpyn Z.T., (2021), Relationship between Residual Saturations and

Wettability using Pore-Network Modeling, SPEJ, (in preparation)

• Purswani P., Johns R.T., Karpyn Z.T., (2021), Impact of Wettability on Capillary Phase Trapping

using Pore-Network Modeling, Water Resour. Res., (in preparation)

• Purswani P., Johns R.T., Karpyn Z.T., Blunt M.J., (2021), Predictive Modeling of Relative

Permeability using a Generalized Equation-of-State, SPEJ, (26), 191-205,

https://doi.org/10.2118/200410-PA

• Purswani P., Karpyn Z.T., Khaled E., Yuan X., Xiaolei H., (2021), Evaluation of Image

Segmentation Techniques for Image-Based Rock Property Estimation, J. Pet. Sc. Eng., (195),

https://doi.org/10.1016/j.petrol.2020.107890

• Purswani P., Tawfik, M.T., Karpyn Z.T., Johns R.T., (2019), On the Development of a Relative

Permeability Equation of State, Comput Geosci., (24), 807-818, https://doi.org/10.1007/s10596-

019-9824-2

Conference and talks

• Purswani P., Johns R.T., Karpyn Z.T., (2021), Relationship between Residual Saturations and

Wettability using Pore-Network Modeling, SPE ATCE, Virtual, (in preparation)

• Purswani P., Johns R.T., Karpyn Z.T., Blunt M.J., (2020), Predictive Modeling of Relative

Permeability using a Generalized Equation-of-State, SPE IOR Conference, 31st Aug-4th Sept,

Virtual

• Purswani P., Tawfik, M.T., Karpyn Z.T., Johns R.T., (2018), On the Development of a Relative

Permeability Equation of State, EME Research Showcase, State College, Pennsylvania

• Purswani P., Tawfik, M.T., Karpyn Z.T., Johns R.T., (2018), On the Development of a Relative

Permeability Equation of State, ECMOR XVI, 16th European Conference on the Mathematics of

Oil Recovery, 3-6 September, Barcelona, Spain

• Purswani P., Karpyn Z.T., Johns R.T., (2018), Correlating Transport Parameters Impacting

Multiphase Flow through Permeable Media, Gordon Research Conference, 8-13th July, Maine,

USA

Page 57: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

37

CHAPTER 2. IMAGING AND PORE-SCALE

MEASUREMENTS

Preface

The contents of this chapter were originally published in the Journal of Petroleum Science and Engineering

and are referenced as,

Purswani P., Karpyn Z.T., Khaled E., Yuan X., Xiaolei H., (2020)

Evaluation of Image Segmentation Techniques for Image-Based Rock Property Estimation, J. Pet. Sc. Eng,

(195), https://doi.org/10.1016/j.petrol.2020.107890

Author contributions: Purswani P. and Karpyn Z.T. conceptualized the experiment. Purswani P. and Enab

K. performed the experiments and wrote the original draft in consultation with Karpyn Z.T. All coauthors

contributed toward analyzing the data and updating the manuscript.

Abstract

Accurate characterization of rock and fluid properties in porous media using x-ray imaging techniques

depends on reliable identification and segmentation of the involved phases. Segmentation is critical for the

estimation of porosity, fluid saturations, fluid and rock topology, and pore connectivity, among other pore-

scale properties. Therefore, the purpose of this study was to compare the effectiveness of different image

segmentation techniques when applied to image data analysis in porous media. Two machine learning based

segmentation techniques – a supervised ML technique called Fast Random Forest, and an unsupervised

method combining k-means and fuzzy c-means clustering algorithms – were compared using an

Page 58: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

38

experimental data set. Comparisons are also presented against traditional thresholding segmentation. In

addition, we discuss the potential and limitations of applying deep learning-based segmentation algorithms.

The performance of the segmentation techniques was compared on estimates of porosity, saturation, and

surface area, as well as pore-scale estimates such as fluid/fluid interfacial areas, and Euler characteristic.

X-ray micro-computed tomography images for a sintered glass frit, saturated with two-phases (air and

brine), were acquired at two different voxel resolutions. The high-resolution images (6 µm) were used as

the benchmark case, while the low-resolution images (18 µm) were segmented by three segmentation

techniques: Fast Random Forest, clustering, and thresholding. The results for porosity and phase saturation

from thresholding and from the supervised ML method (i.e., Fast Random Forest) were found to be close

to the benchmark case. Segmentation results from the unsupervised ML method (i.e., clustering) were

largely unsatisfactory, except for total surface area measurements. The supervised ML segmentation results

provided better measurements for air-brine interfacial areas by capturing three-phase interfacial regions.

Also, all segmentation techniques resulted in similar measurements for air-phase Euler characteristic

confirming poor connectivity of the trapped air phase, although the closest results were obtained by the

supervised ML method. Finally, despite the supervised ML segmentation technique being more

computationally intensive, it was found to require less user intervention and its implementation was more

straightforward. In summary, this work provides insights into different segmentation techniques, their

implementation, as well as advantages and limitations with regards to quantitative analysis of pore-scale

properties in saturated porous media.

Page 59: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

39

2.1. Introduction

High-quality, non-destructive imaging is at the heart of innovative science in a variety of disciplines. It

provides researchers with the ability to examine objects at small length scales, which enables the estimation

of a variety of structural and topological properties. In the geosciences, applications of non-destructive

imaging include characterization of rock heterogeneities, pore-network properties, roughness, fluid

distributions, and transport in porous systems (Blunt 2017; Lai et al. 2015; Noiriel et al. 2004; Wildenschild

and Sheppard 2013). X-ray micro-computed tomography (µCT) is one such imaging technique that

generates a three-dimensional (3-D) mapping of linear attenuation coefficients acquired by a digital x-ray

detector. These attenuation coefficients are distinct for each material phase in the object (Cnudde and Boone

2013). As such, an x-ray image provides both quantitative and qualitative information about the elements

constituting the object scanned. To draw meaningful information from these digital images, a series of

image processing steps are necessary. These steps help improve the visual appearance of digital x-ray

images, as well as prepare them for feature and property analyses.

There are three main steps of image processing, namely, pre-processing, segmentation, and post-

processing. Image pre-processing consists of steps to reduce the impact of image artifacts such as noise,

image blur, beam hardening, ring effects, and bright spots (Huda and Abrahams 2015). This is achieved by

the application of image filters like median (Bernstein 1987), mean, non-local mean (Buades et al. 2005),

and edge detecting filters (Sheppard et al. 2004) that help improve the quality of reconstructed raw images

and prepare them for image segmentation. Image segmentation is the process of categorizing (or labeling)

each voxel to a specified class or phase in the object. This labeling step assigns a characteristic number to

all voxels belonging to the same phase. This assists in quantitative analysis on the images, for example,

voxel counting is used for porosity and phase saturation measurements. Lastly, image post-processing is

the operation of fixing any misrepresentation of phases in the segmented image. All image processing steps

are crucial for consistent and accurate feature measurements. The purpose of this work is to evaluate various

image segmentation techniques for image-based rock property estimation.

Page 60: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

40

Due to the growing access of x-ray µCT scanners to numerous researchers in Earth sciences, there has

been an increase in the number of studies that use this technique for studying fluid flow in porous media.

Segmented images are used to measure phase characteristics to provide the observational basis for

understanding different processes such as multiphase fluid flow, structural morphology, pore connectivity,

fluid/fluid, and fluid/solid interfaces (Blunt 2017). These characteristics are subsequently used to quantify

fluid transport through estimations of flow properties such as relative permeabilities and capillary pressures

(Khorsandi et al. 2017). Carefully segmented x-ray images are often used as a starting point for simulating

fluid flow by using techniques such as Lattice Boltzmann simulations (Armstrong et al. 2016; Landry et al.

2014; Liu et al. 2018; Mcclure et al. 2018) or pore-network modeling (Dong and Blunt 2009; Joekar-Niasar

et al. 2010; Joekar-Niasar et al. 2008; Reeves and Celia 1996; Valvatne and Blunt 2004). Table 2 lists a

few of the experimental studies performed over the past two decades. It can be inferred from Table 2 that a

variety of porous systems spanning natural and synthetic media have been studied. It can also be inferred

that, in general, over time, the scanning resolutions have improved as technology advances. Further, there

is a general acceptance for using non-local means filtering technique for pre-processing purposes.

Table 2. Characteristics of imaging techniques from various experiments of fluid flow in porous media.

Reference Porous

media

Image

filtration

Voxel

resolution

(um)

Segmentation

technique Post processing

(Culligan et al.

2004)

Glass bead

pack Median filter 18 k-means clustering -

(Culligan et al.

2005)

Soda lime

beads Median filter 17 k-means clustering -

(Porter and

Wildenschild 2010) Bead pack

Anisotropic

diffusion filter 5.9; 11.8 k-means clustering -

(Karpyn et al. 2010) Glass bead

pack - ~26 Thresholding -

(Landry et al. 2011) Acrylic bead

pack Median filter ~26 Thresholding

Smoothing of

surfaces for area

measurements

Page 61: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

41

(Herring et al. 2013) Bentheimer

sandstone - 10 Indicator Kriging

Removal of

nonwetting clusters

smaller than 100

voxels

(Celauro et al. 2014) Coated glass

bead packs - ~27

Gauss curve fitting

to gray value

histograms

-

(Harper 2013;

Herring et al. 2013;

Joekar-Niasar et al.

2013; Porter and

Wildenschild 2010)

Crushed tuff Anisotropic

diffusion filter 17.5 k-means clustering -

Sintered

glass bead

pack

Median filter 13 Thresholding -

(Herring et al. 2015) Bentheimer

sandstone Median filter 5.8 Thresholding

Removal of air

clusters smaller

than 125 voxels

(Rücker et al. 2015) Gildehauser

sandstone

Non-local

means 2.2 Watershed -

(Berg et al. 2016) Gildehauser

sandstone

Non-local

means 2.2 Watershed

Segmented phases

were cleaned using

morphological

operations

(Schlüter et al.

2016)

Sintered soda

lime bead

pack

Non-local

means and

total variation

denoising filter

8.4 Markov random

field technique -

Sintered

glass bead

pack

Median filter 2.2 Watershed

Removal of

clusters smaller

than 125 voxels

(Gao et al. 2017) Bentheimer

sandstone

Non-local

means 6 Thresholding -

(Singh et al. 2017) Ketton

limestone

Non-local

means 3.28

Seeded watershed

algorithm and

thresholding

Dilation of rock

phase for curvature

analysis

(Lin et al. 2018) Bentheimer

sandstone

Non-local

means 3.58

Seeded watershed

algorithm and

thresholding

Boundary

smoothing for

curvature analysis

(Rücker et al. 2019) Ketton

limestone

Non-local

means 3

Watershed

algorithm and

thresholding

-

Table 2. (Continued)

Page 62: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

42

(Lin et al. 2019) Bentheimer

sandstone

Non-local

means 3.58

Seeded watershed

algorithm and

thresholding

Boundary

smoothing for

curvature analysis

This work

Sintered

glass bead

pack

Non-local

means 6; 18

Thresholding; k

and c means

clustering; and

supervised machine

learning

Removal of small

nonwetting phase

clusters for Euler

number analysis

Pore-scale measurements such as porosity, phase saturation, fluid topology, and fluid/fluid interfacial

areas can be extremely sensitive to the results of image processing steps, in particular, image segmentation.

Segmentation methods can largely be categorized into two groups, global methods and local adaptive

methods (Iassonov et al. 2009). Global methods, such as intensity-based thresholding, work by identifying

valley points on the voxel population histogram of the filtered images. A threshold gray value is set to

classify the voxels, such that gray values above the threshold are identified as one phase, while the voxels

below the threshold are identified as the other phases. This method worked reasonably well for a multiphase

system with sufficient contrast in the gray-levels of each phase, which makes the identification of the valley

points in the histogram easier. Because of its ease of application, intensity-based thresholding continues to

be a common method of segmentation in the digital rocks community (Prodanovic et al. 2015).

Locally adaptive segmentation refers to the segmentation methods that make segmentation decisions

for each voxel in the image. There have been numerous developments on this type of segmentation to

achieve more refined results. Watershed segmentation (Vincent and Soille 1991), converging active contour

method (Sheppard et al. 2004), Markov random field segmentation (Kulkarni et al. 2012), and indicator

kriging (Oh and Lindquist 1999) are a few examples of locally adaptive methods. A comprehensive review

of the implementation and comparison of these locally adaptive methods is available in Schluter et al.

(2014). Machine learning techniques such as fuzzy c-means (Pham and Prince 1999), a combination of k-

means and fuzzy c-means (Dunmore et al. 2018), and supervised machine learning are other examples of

locally adaptive methods of segmentation.

Table 2. (Continued)

Page 63: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

43

The capability of machine learning (ML) approaches in solving classification problems has enabled the

utilization of such techniques for generating segmentation algorithms. Traditional supervised ML

algorithms work as feedback methods by learning from annotated voxel labels of some part of an image in

order to predict the class distribution of each voxel in the whole image. Such an ML model is learnt in a

training process that extracts a vector of features that influence voxel class labels based on feedback from

annotated labels of some voxels (Kotsiantis 2007). After training, the resulting ML model is used to assign

a class label to each voxel in the entire image based on voxel feature values. Support Vector Machines,

Neural Networks (Multilayer Perceptron), Decision Trees, Random Forest, and Fast Random Forest are

examples of supervised ML algorithms that can be used to generate classification models for image

segmentation purposes.

Unsupervised ML approaches, unlike supervised methods, do not need annotations for part of data, but

operate by grouping voxels based on similarities. Clustering, also known as cluster analysis, is one of the

most common types of unsupervised ML methods and it is widely used for classification purposes in data

analysis and data mining. K-means clustering, and fuzzy clustering (c-means or soft k-means clustering)

are two common methods for clustering. Both k-means and c-means clustering are iterative methods that

operate by identifying the similarity of an element in the population to different groups of elements. The

assignment of an element to a particular group is probabilistic in c-means as opposed to deterministic in k-

means.

In the past decade or so, deep learning (DL) methods based on multi-layer artificial neural networks

have produced state-of-the-art results in many fields including computer vision, speech recognition,

medical image analysis, and material inspection. In the area of image segmentation, DL has also achieved

success, including in the segmentation of µCT images. Through training with a large number of fully

annotated images, DL models can extract meaningful visual features automatically and use them to infer

segmentation maps. For 2-D image segmentation, U-net (Ronneberger et al. 2015), Deeplab series (Chen

et al. 2018; Chen et al. 2017a; Chen et al. 2017b), Mask R-CNN (He et al. 2017) among others (Oktay et

al. 2018; Xue et al. 2018; Zhou et al. 2018) cover a variety of different applications from natural images to

Page 64: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

44

medical images. For 3-D image segmentation, prior works (Çiçek et al. 2016; Milletari et al. 2016; Xue et

al. 2019) mainly focus on medical applications such as 3-D magnetic resonance imaging (MRI) or CT

scans.

For rock image segmentation, Wang et al. (2020) introduced a novel 3-D µCT segmentation method

built on U-net (Ronneberger et al. 2015) and ResNet (He et al. 2016) in their recent work. Niu et al. (2020)

and Karimpouli and Tahmasebi (2019) used Convolutional Neural Network (CNN)-based algorithms for

segmenting sandstone data sets. It was found that CNN algorithms can minimize the need for user-defined

inputs (Niu et al. 2020). Although DL methods can achieve promising segmentation results, their feature

learning capacity heavily relies on the large amount of training images as well as high-quality manual

annotations. Moreover, unlike other types of visual recognition tasks such as image classification which

only require image level annotations, the annotation of 3-D µCT rock images for segmentation purposes

requires labeling at a voxel-by-voxel level for the entire image, which can be very expensive and

impractical. Further, because the mineral composition and structural features differ in porous media, voxel-

label annotations obtained for one rock system may not be useful as training data to train the segmentation

model for other rock systems. Thus, DL methods may not be a suitable choice for image segmentation

unless diverse saturated porous media image data are available as training data sets.

In summary, the literature presents a variety of image segmentation techniques including both global

and local techniques. However, newer ML techniques are less commonly used in the porous media

community. Therefore, the purpose of this study is to compare the effectiveness of two ML-based

segmentation techniques – a supervised ML technique called Fast Random Forest, and an unsupervised

method combining k-means and fuzzy c-means clustering algorithms – for segmenting a saturated

multiphase porous medium. Comparisons are also presented against a thresholding-based segmentation

technique. We investigate the segmentation methods on a small-scale dataset where DL models can easily

overfit to training samples which makes them less generalizable. For this reason, we do not include deep

learning-based methods in our comparison. Our goal is to compare feasible segmentation methods for

identification of fluid and solid phases in saturated porous media. We provide a quantitative analysis of the

Page 65: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

45

segmented phases to estimate physical characteristics of porosity, phase saturations, phase surface areas,

interfacial areas, and phase connectivity to demonstrate and compare the capabilities and limitations of each

segmentation technique with recommendations for each when applied to saturated porous media. In this

way, this research provides the readers with insight into emerging machine learning-based image

segmentation techniques, their implementations, their comparative advantages, as well as limitations with

regards to applications in porous media research.

Page 66: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

46

2.2. Methodology

Trusted segmented data are required as a benchmark for comparing the effectiveness of the different

segmentation techniques examined in this research. For this, a saturated porous medium was prepared for

this study and scanned at two different voxel resolutions. The first, benchmark scan was acquired at a voxel

resolution of 6 µm. The second test scan was acquired at a voxel resolution of 18 µm which was used to

test and compare different segmentation techniques. Although 18 µm represents low resolution in this

research, it is still typical of x-ray micro-tomographic studies (Table 2).

The experimental set-up (Figure 9) used in this research is a static set-up consisting of a sintered glass

frit (pore sizes between 100-160 µm) saturated with brine and air, representing wetting and nonwetting

phase, respectively. The x-ray scanner used was GE v | tome | x L300 system with a 300kV x-ray tube. The

sintered glass frit was a specific type of borosilicate glass filter (Robu) procured from Adam and Chittenden

Scientific Glass, California, USA. It is a glass filter widely used for water filtration purposes, 10 mm in

diameter, 2.8 mm long, with ~18% porosity. The sintering allows for the porous medium to be rigid and

maintain its pore structure during handling. The brine phase used in this experiment was a solution of 1M

sodium iodide (NaI). Doping the brine with 1M NaI helps to attenuate more x-rays such that enough contrast

can be achieved to isolate the three phases. This particular concentration of NaI was found to be optimum

and was achieved after multiple trials to minimize imaging artifacts and maximize contrast.

Page 67: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

47

Figure 9. Schematic of the laboratory setup and image acquisition system (x-ray MCT scanner). DO stands for the distance between the detector and the object, while OS stands for the object to source distance. This figure shows that the sample (object) is very close to the source for finer resolution, the resolution was coarsened by moving the sample stage laterally from the source, increasing OS and decreasing DO.

The porous glass frit was held fixed inside a thin (~ 1mm wall thickness) plastic tubing, open at the top,

closed at the bottom and secured to the scanner’s rotating sample mount. The setup was placed within a

few millimeters of the x-ray source to maximize the image resolution to 6 µm (Figure 9). At this position,

a scan for the dry frit was acquired. Next, a pipette filled with the brine solution was used to drop a couple

of droplets into the porous glass frit. After waiting for 20 minutes for the liquid to saturate the glass frit, the

benchmark x-ray scan was acquired. At this stage, the scanned system consisted of three phases (solid glass,

brine, and trapped air). Upon completion of this scan, the sample mount was moved laterally by increasing

the object-to-source (OS) distance and decreasing object-to-detector (DO) distance from the x-ray source

to acquire the exact same scan at a resolution of 18 µm. This was termed as the test scan.

X-ray source

Detector

Glass frit

X-ray chamber

Plastic tubing

Sample mount

NaI dropletPipette

DO OS

Page 68: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

48

2.2.1. Implementation of the segmentation techniques

In this section, we discuss the image processing framework used in this research. Both dry and brine

saturated raw CT images were processed through the non-local means filter to remove image noise. No

other major image artifacts were observed in the CT images. The non-local means filter was found to be

effective as compared to the median filter. The filtering step was not required for the supervised machine

learning segmentation which directly works on raw CT images. This is discussed in the subsequent section.

2.2.2. Benchmark case

To generate the benchmark segmented images, both dry and saturated images of the high-resolution scan

were used to generate reliable segmented images. First, thresholding was conducted on the filtered dry

sample to segment the solid and the pore space. This was easier to accomplish because of the significant

difference in the gray values between the air and the solid phase. Second, the segmented dry images were

subtracted from the saturated images to eliminate the solid phase. This left only the brine and trapped air

phase which were segmented once again using thresholding.

2.2.3. Test case

The comparative analysis between the supervised and unsupervised machine learning segmentation

techniques was conducted using the scanned images at a voxel resolution of 18 µm. Additional comparison

with thresholding is also presented. The implementation of these segmentation techniques is outlined below.

Page 69: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

49

2.2.4. Supervised machine learning (ML)-based on Fast Random Forest algorithm

The supervised ML-based segmentation technique is a multi-threaded implementation of the Random

Forest algorithm, as provided in the WEKA (Waikato Environment for Knowledge Analysis) trainable

segmentation toolbox (Arganda-Carreras et al. 2017). The WEKA segmentation toolkit is implemented as

a built-in plugin in ImageJ. It works as a bridge to apply machine learning tools for image processing and

has been used in a few recent studies for segmentation purposes (Berg et al. 2018; Garfi et al. 2020). The

Random Forest algorithm (Breiman 2001) is a classification algorithm consisting of many decision trees

that operate as an ensemble. Each individual decision tree provides a vote on class prediction and, the class

with the most votes becomes the forest's class prediction. The Random Forest algorithm uses bagging and

feature randomness when building each individual tree to try to build a forest of largely uncorrelated trees

whose prediction by the "wisdom of crowds" is more accurate than that of any individual tree. Random

Forest is considered a fast classifier, but more recently, parallelized versions such as the WEKA Fast

Random Forest implementation enable one individual tree per processor core to take advantage of multi-

core processors, further reducing forest build time.

Below, we outline the general procedure for applying the Fast Random Forest segmentation method in

this research:

1. To build the training set, sample voxels representing each of the different target categories are

selected and each such sample voxel is labeled with the category to which it belongs.

2. To train the random forest classifier, a vector of image features for each voxel is used as the training

feature. The vector of features contains the CT value of the voxel in the raw image, coupled with

the CT values of the same voxel in different filtered images. The available filters provided in the

WEKA toolkit include Gaussian blur, Hessian, derivatives, structure, edges,

minimum/maximum/mean/variance/median, etc. After a trial-and-error process where different

combinations of image filters were tested to arrive at the combination that provided the best

performance, we finally chose the minimum, maximum, mean, and variance filters. Note that these

Page 70: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

50

filters are applied at a voxel-level: the voxels within a small radius (e.g., within a 3x3x3 small

neighborhood) from the target voxel are subjected to the pertinent operation (min, max, mean, or

variance) and the target voxel is set to that value in the filtered image. Once the training features

for all sample voxels (with ground-truth labels) are computed, the random forest classifier is trained

using the (feature-vector, label) pairs for the sample voxels.

3. After the random forest classifier is trained, it is then used to classify every voxel in the entire

image, thus achieving full segmentation of the whole image.

During training of the classifier (step 2 above), the sample dataset of labeled voxels is automatically

divided into three subsets: training, validation, and testing. The subset for training consists of ~ 80% of the

sample data and is used to build the classifier. The subset for validation consists of ~ 10% of the sample

data, and it is used to adjust parameters of the classifier and choose the combination of image filters to use;

that is, many classifiers can be trained with different parameter values (e.g., number of trees in the forest)

and different combinations of image filters, and then the optimal values and filters are chosen based on

which ones give the best performance on the validation subset. Lastly, the subset for testing consists of the

remaining 10% of the sample data and is used to test and report the accuracy of the final chosen classifier.

2.2.5. Unsupervised machine learning based on k-means and fuzzy c-means clustering

The unsupervised technique selected for this study is the medical image analysis (MIA) – clustering

technique, which is an open-source algorithm useful for multiphase image segmentation (Dunmore et al.

2018; Wollny et al. 2013). This segmentation technique requires denoised images as input, so we apply

Gaussian filtering to a raw image before applying the technique, which combines two unsupervised

clustering methods, k-means and fuzzy c-means. K-means clustering is an iterative scheme that places each

member (e.g., a voxel) in a given data set (e.g., all voxels in a CT image) into different clusters representing

different classes. The iterative process starts by defining the centroids (or means) of each cluster arbitrarily.

Page 71: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

51

Then, each data point is grouped to a data cluster as a function of the Euclidean distance between the data

point and a cluster centroid. Next, the cluster centroid values are updated until the difference between the

previous and the updated centroid values meet a user-defined tolerance. C-means clustering is similar to k-

means clustering with one difference which lies in the flexibility of allowing a data point to probabilistically

belong to more than one cluster (Bezdek, James C; Ehrlich, Robert; Full 1984; Dunn 1973). The

probabilistic nature of this approach is enabled by the inclusion of a membership function (with value

between zero and one) and a term called the fuzzifier (a real number between one and two). The membership

function governs the degree to which a particular data point belongs to a particular cluster, whereas the

fuzzifier determines the fuzziness level of a cluster. Larger values of the fuzzifier lead to smaller values of

the membership function and vice versa.

The MIA-clustering algorithm requires little user intervention. Two main input parameters supplied by

the user are the number of classes that exist in the image and the grid-size used for partitioning the image

into overlapping cubes so that segmentation can be refined locally within the cubes. In the first part of the

algorithm, the K-means algorithm clusters all voxels, based on voxel intensity in the denoised CT image,

into the number of classes specified by the user. Subsequently the fuzzy c-means algorithm is applied to

iteratively estimate all class membership probabilities for each voxel, expressed as a vector. Then, by

assigning each voxel to a class based on its highest membership probability, the whole image is clustered

into distinct classes representing structures. However, this global segmentation may miss some fine details

because of intensity inhomogeneities in the input image. Therefore, in the second part of the MIA-clustering

algorithm, fuzzy c-means is applied locally. The whole image volume is subdivided into overlapping small

cubes based on the grid-size parameter. In a cube, the sum of membership probabilities of all voxels for

each class is calculated; if the sum for a class falls below a threshold, then that class is not considered for

the local, refined c-means clustering in the cube. After the local refinement is done for all cubes, class

probabilities for each voxel in overlapping cubes are merged, and once again, voxels are assigned to the

class for which they have the highest membership probability, producing the whole segmented image. More

details about the MIA-clustering algorithm can be found in Dunmore et al. (2018).

Page 72: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

52

2.3. Results and discussion

In this section, results from the implementation and comparison of the machine learning (supervised and

unsupervised) segmentation techniques are discussed. Additional comparisons are presented against

thresholding segmentation. For quantitative analysis, we present bulk measurements of porosity, fluid

saturations, phase fractions, and phase surface areas, as well as pore-scale measurements of phase

connectivity (measured as the Euler characteristic) and fluid/fluid interfacial areas.

The imaged cross-sections of the porous glass frit (dry and brine saturated) are shown in Figure 10, and

the corresponding grayscale intensity histograms are shown in Figure 11. In Figure 10 (top), the brighter

region corresponds to the solid (sintered glass) whereas, the darker region corresponds to the pore (air)

space. The solid, being denser, attenuates more x-rays and appears bright. Figure 10 (top) and Figure 11

(top) show that the quality of the acquired dry x-ray scan is excellent as evidenced from the histogram of

the raw image which is well resolved between the pore and solid space even before applying any

enhancement filters. Minimal x-ray imaging artifacts are observed. Upon the application of the non-local

means filter, the difference in the voxel populations of the solid phase and the pore space becomes clearer.

This assists in the segmentation of two phases by thresholding. We note here that the parameters set for the

non-local means image filtration were kept uniform across all image data sets, irrespective of the resolution,

or whether the data sets were dry, or brine saturated to maintain a common pre-processing procedure for

all images to be segmented.

In Figure 10 (middle) and Figure 10 (bottom), light-gray regions correspond to the brine phase, middle-

gray regions represent the solid phase, and the darker isolated regions represents the air phase. Notice that

the application of the non-local means filter removes image noise (Figure 10 right column). This is more

prominent for the scan at a voxel resolution of 6 µm as opposed to the lower quality scan.

Page 73: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

53

Figure 10. Imaged cross-section of dry (top) and brine saturated (middle and bottom) porous glass frit at different voxel resolutions. The brine used for saturating the porous medium was 1M NaI solution. Non-local means was used for filtering the raw images to remove image noise.

Filtered scan

Dry

Bri

ne

sat

ura

ted

Raw scan

Vo

xel r

eso

luti

on

= 6

µm

Vo

xel r

eso

luti

on

= 1

8 µ

m

Page 74: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

54

Figure 11. Histograms showing the voxel population of the different grayscale intensity values for the corresponding scans shown in Figure 10.

Figure 12 shows a comparison of segmented top view ortho slices using the machine learning and

thresholding segmentation techniques, against the benchmark case. For thresholding, the average of two

distinct attempts was considered. Each attempt was carried out by manually adjusting the threshold mark

between the phases. We see that the air phase, represented by the darker isolated regions, is easier to

recognize and is consequently successfully segmented by all techniques. Segmentation differences amongst

the various approaches are most evident in the identification of brine and solid phases, as shown in Figure

0

5000

10000

15000

20000

25000

30000

35000

1 815

22

29

36

43

50

57

64

71

78

85

92

99

10

611

312

012

713

414

114

815

516

216

917

618

319

019

720

421

121

822

523

223

924

625

3

Vo

xel

po

pula

tio

n

Measure of CT intensity

CT intensityCT intensity

Raw scan

0

5000

10000

15000

20000

25000

30000

1 815

22

29

36

43

50

57

64

71

78

85

92

99

10

611

312

012

713

414

114

815

516

216

917

618

319

019

720

421

121

822

523

223

924

625

3

Vo

xel

po

pula

tio

n

Measure of CT intensity

0

20000

40000

60000

80000

100000

120000

1 815

22

29

36

43

50

57

64

71

78

85

92

99

10

611

312

012

713

414

114

815

516

216

917

618

319

019

720

421

121

822

523

223

924

625

3

Vo

xel

po

pula

tio

n

Measure of CT intensity

Vox

el P

op

ula

tio

n

0

2000

4000

6000

8000

10000

12000

1 815

22

29

36

43

50

57

64

71

78

85

92

99

10

611

312

012

713

414

114

815

516

216

917

618

319

019

720

421

121

822

523

223

924

625

3

Vo

xel

po

pula

tio

n

Measure of CT intensity

0

2000

4000

6000

8000

10000

12000

14000

16000

1 815

22

29

36

43

50

57

64

71

78

85

92

99

10

611

312

012

713

414

114

815

516

216

917

618

319

019

720

421

121

822

523

223

924

625

3

Vo

xel

po

pula

tio

n

Measure of CT intensity

Vox

el P

op

ula

tio

n

CT intensityCT intensity

CT intensityCT intensity

Vox

el P

op

ula

tio

nV

oxe

l Po

pu

lati

on

Vo

xel P

op

ula

tio

n

Vox

el P

op

ula

tio

nFiltered scan

Dry

Bri

ne

satu

rate

d

Vo

xel r

eso

luti

on

= 6

µm

Vox

el r

eso

luti

on

= 1

8 µ

m

Page 75: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

55

12. The benchmark case (Figure 12a) shows that the brine phase has fragmented clusters making the phase

appear more disconnected in the two-dimensional space, although the phase may be connected in the three-

dimensional pore space. This continues to be seen in the other segmented cases; however, relatively bigger

clusters seem to be apparent for the unsupervised machine learning (Figure 12d). Further, it can be noticed

that the benchmark case shows a clear interfacial contact among the air, brine, and the solid phases. When

comparing the interfacial contact detected by the three segmentation techniques, it is observed that all the

segmentation techniques closely detect the interfacial contact between the air and the solid phases.

However, for the interfacial contact between the air and the brine (direct fluid/fluid contact), supervised

machine learning segmented image showed slightly better performance. Due to the missing three-phase

contact in the unsupervised machine learning and thresholding cases (Figure 12b and Figure 12d), we

observe that the air phase appears to be isolated. Oil-brine interfaces are being misidentified as part of the

solid phase during segmentation, thus potentially leading to the loss of fluid/fluid contact areas. This is

problematic because inaccurate estimations of the three-phase contacts can lead to erroneous contact angle

measurements (Alhammadi et al. 2017; Klise et al. 2016; Scanziani et al. 2017). These preliminary

observations are substantiated quantitatively through bulk and pore-scale measures in the following

sections.

Page 76: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

56

Figure 12. Segmented cross-sectional images showing three phases (solid, brine, and air). (a) thresholding at a resolution of 6 µm (benchmark case), (b) thresholding at a resolution of 18 µm, (c) supervised machine learning segmentation at a resolution of 18 µm, and (d) unsupervised machine learning segmentation at a resolution of 18 µm. Zoomed-in version of the images are displayed on the sides to highlight distinct features of segmented images. The upper and lower regions of interest (marked inside the segmented images) correspond to labels 1 and 2, respectively.

2.3.1. Bulk measurements

Bulk measurements of porosity, phase saturations, and fluid surface areas are critical measures for

understanding fluid flow in porous media. Porosity and saturation measures help quantify the amount of oil

and gas reserves present in an oil reservoir (Lake et al. 2014); whereas, area measurements are often used

by hydrologists to quantify the extent of a chemical (non-aqueous phase liquids) spill for groundwater

remediation purposes (Culligan et al. 2005). These measurements are provided for all segmentation

techniques used in this research.

a b

c dc1

c2

b1

b2

d1

d2

a1

a2

Page 77: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

57

2.3.1.1. Porosity, fluid saturations, and phase fractions

Figure 13 displays the bulk measures of porosity, fluid (air/brine) saturation, and individual phase

(air/brine/solid) fractions as a function of the height of the sample calculated using the segmented images

from each technique, while Table 3 presents the corresponding percent errors calculated against the

benchmark case for each average bulk measure. It is observed that for the benchmark case, porosity ranges

between (~15 to 21%) across the height of the sample and decreases slightly at the top part of the sample.

This could be due to the manufacturing artifact of the glass frit or due to slight wear of the glass frit on the

edges as it was pushed in the plastic tubing for acquiring the x-ray scans. It is also observed that all

segmentation techniques were able to capture the porosity trend. However, thresholding and supervised

machine learning performed better as opposed to the unsupervised machine learning (see Table 3 for

comparing % errors calculated from the benchmark case). The average porosity of the scanned medium for

the benchmark case, thresholding, unsupervised machine learning, and supervised machine learning to be

around 17.8%, 17.3%, 22.3%, and 19.1%, respectively.

Page 78: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

58

Figure 13. Vertical profiles of (a) porosity, (b) brine saturation, (c) air saturation, (d) solid fraction, (e) brine fraction, and (f) air fraction for different segmentation techniques.

0

0.2

0.4

0.6

0.8

1

1.2

0.14 0.19 0.24

Hei

ght

(mm

)

Porosity (-)

a

0

0.2

0.4

0.6

0.8

1

1.2

0.7 0.75 0.8 0.85 0.9 0.95 1

Hei

ght

(mm

)

Brine saturation (-)

b

0

0.2

0.4

0.6

0.8

1

1.2

0 0.05 0.1 0.15 0.2 0.25 0.3

Hei

ght

(mm

)

Air saturation (-)

c

0

0.2

0.4

0.6

0.8

1

1.2

0.71 0.76 0.81 0.86

Hei

ght

(mm

)

Solid fraction (-)

d

0

0.2

0.4

0.6

0.8

1

1.2

0.13 0.18 0.23

Hei

ght

(mm

)

Brine fraction (-)

e

0

0.2

0.4

0.6

0.8

1

1.2

0 0.02 0.04 0.06 0.08

Hei

ght

(mm

)

Air fraction (-)

Supervised Machine Learning

Unsupervised Machine Learning

Thresholding

Benchmark Case

f

Page 79: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

59

Table 3. Summary of percent errors calculated for the different measured properties and segmentation techniques relative to the benchmark case. The percent errors for the bulk measurements of porosity, fluid saturation, and phase fractions are calculated for the respective average values across the sample height.

Measured property Phase Thresholding

(% error)

Unsupervised machine

learning

(% error)

Supervised

machine learning

(% error)

Sample porosity - 0.06 24.87 6.83

Fluid saturations Air 20.45 20.61 4.33

Brine 3.44 3.47 0.73

Phase fractions

Air 20.37 1.62 11.61

Brine 3.54 29.56 5.98

Solid 0.01 5.40 1.48

Phase surface areas

Air 27.27 13.51 6.02

Brine 31.84 14.90 29.01

Solid 26.22 8.45 22.85

Phase saturation is a measure of the relative volumes of the fluids occupying the pore space. Figure 13b

displays the brine phase saturation and Figure 13c displays the air phase saturation as a function of height

for each segmented image. We observe a mirror image plot for the two fluids with the wetting phase being

more abundant. It is noticeable that air saturation increases near the bottom of the sample. This is intuitive

since the droplets of brine were introduced from the top of the sample and less brine percolates to the bottom

of the sample – trapping more air at the bottom.

The average saturation of the wetting phase calculated using the segmented images of the benchmark

case, thresholding, unsupervised machine learning, and supervised machine learning were 85.6%, 88.6%,

88.6%, and 85%, respectively, showing that supervised machine learning provided the closest estimates to

the benchmark case with a percent error of just 0.73 % (see Table 3).

Successful segmentation is most challenging for phases that have relatively close gray-level intensities

(brine and solid in this case). We present results for the individual fractions of the solid, brine, and air

phases in Figure 13d, Figure 13e, and Figure 13f. It is found that the air phase makes a small portion of the

total bulk volume at an average fraction of ~ 0.025 when calculated using the images of the benchmark

case. The solid makes a large portion at an average fraction of ~ 0.8 (Figure 13d), while the brine phase

makes the remaining fraction at 0.175. The overall trends from different segmentation techniques continue

Page 80: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

60

to give similar findings as to the benchmark case. As observed for the case of porosity, we see that

unsupervised machine learning differs the most from the benchmark case. It appears that this technique

estimates a larger brine fraction (consequently, lower solid fraction – Figure 13d and Figure 13e). This is

evident when comparing the benchmark case (Table 3) for the air fraction and air saturation, which shows

that despite capturing air fraction fairly, there is a large error for air saturation owing to misrepresented

brine fraction.

2.3.1.2. Surface areas

The surface areas for the different segmentation techniques are plotted in Figure 14 while the corresponding

percent errors calculated against the benchmark cases are presented in Table 3. The algorithm used to

measure the surface area uses a modified marching cube technique to generate a polygonal surface that

wraps around the three-dimensional object (the phases of interest in this case) (Landry et al. 2011). Each

face of the polygonal surface represents a triangle of fixed dimensions based on voxel size. The total surface

area of these triangles represents the total surface area of that phase. All segmented image sets were

imported into Avizo® Fire 9.43 for surface area measurements. Only the interior phase areas are reported

for meaningful comparisons.

Figure 14 shows that the total surface area of air (~ 35 mm2) is the lowest, due to the very small amount

of the air phase. We notice that the three segmentation techniques provide close measures for the total

surface area for the air phase owing to the convenient segmentation of the air phase. We also see that the

total surface areas for the brine and the solid phases are comparable, which is attributed to the fact that the

majority of the solid surface is in contact with the brine phase. Further, it is observed that the unsupervised

machine learning segmentation approach provides close estimates to the benchmark case, as opposed to the

other two segmentation techniques. This may appear to be counter-intuitive since for the other bulk

3 https://www.fei.com/software/amira-avizo

Page 81: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

61

measurements unsupervised machine learning showed poor results. However, this suggests that it is difficult

to compare the different segmentation techniques consistently only on bulk measurements. Therefore, next

we present pore-scale measures of fluid/fluid interfacial areas and Euler characteristic.

Figure 14. Surface areas of air, brine, and solid phases on images segmented with different techniques.

2.3.2. Pore-scale measurements

The measurements of the fluid/fluid (air/brine) interfacial areas and the Euler characteristic (discussed in

section 3.2.2) of the air phase are shown in Table 4. Both these measurements were made using Avizo®

Fire 9.4. We observe large differences in the interfacial area measurements for the different segmentation

techniques when compared to the benchmark case. One reason for the higher surface area and interfacial

area measurement for the benchmark case could be attributed to the coastline paradox which is inherent as

resolution improves (Mandelbrot 1982). Despite this, the closest results are observed for the supervised

machine learning case, whereas unsupervised machine learning and thresholding present quite different

results (two to three orders of magnitude). High fluid/fluid interfacial area for supervised machine learning

0

50

100

150

200

250

300

Air Brine Solid

Are

a (m

m2)

Supervised Machine Learining

Unsupervised Machine Learining

Thresholding

Benchmark Case

Page 82: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

62

(close to the benchmark case) suggests that the air/solid interfacial area is less, such that the air clusters are

less dispersed in the solid phase but rather form continuous contact with the brine and solid phases. These

results complement observations made in Figure 12. This is an important demonstration of the success of

the supervised machine learning technique.

Table 4. Measurements of air-brine interfacial area and Euler number of the nonwetting (air) phase for the different segmentation techniques. The average air saturation was 14.4 % as determined from the benchmark case whereas, thresholding, supervised machine learning, and unsupervised machine learning, showed average values of 11.5%, 15.0% and 11.4%, respectively.

Technique Air-brine interfacial

area (mm2)

Euler number

(air phase)

Supervised machine learning 3.272 298

Unsupervised machine learning 0.013 274

Thresholding 0.003 209

Benchmark case 11.757 534

Unlike the fluid/fluid interfacial area, the Euler characteristic identifies the connectivity of a phase by

considering the number of clusters of the phase and the connection of those clusters. For the Euler

characteristic of the air phase, it is seen that all segmentation techniques, together with the benchmark case

show positive numbers – suggesting a disconnected phase. This is true since the air phase was trapped and

remained largely disconnected upon introduction of the brine phase. Further, it is observed that all three

segmentation techniques gave relatively close results, which can be attributed to the more consistent

segmentation of the air phase. Lower value for the test case as opposed to the benchmark case could also

be attributed to partial volume effects associated with the low-resolution scan which could blur out tiny,

isolated air clusters. To avoid this artifact, usually post-processing of segmented images is carried out, such

as, smoothing, erosion, dilation for removal of very small-sized clusters. However, for the results shown in

Table 4 no post-processing steps were carried out for a one-to-one comparison of each segmentation

technique with the benchmark case. Nevertheless, the closest results were observed for supervised machine

learning, although differences were slight.

Page 83: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

63

2.4. Concluding remarks

In this research, supervised and unsupervised machine learning segmentation techniques were compared to

evaluate their strengths and limitations in segmenting multiple phases in saturated porous media. Additional

comparisons were made against thresholding segmentation technique. For this, a synthetic porous medium

with three phases (solid, brine, and air) was scanned using x-ray micro-CT at two resolutions, a high voxel

resolution of 6 µm and a low voxel resolution of 18 µm. The segmented high-resolution scan was used as

the benchmark case while the low-resolution images were segmented by the three segmentation techniques

to evaluate their performance in estimating the physical properties of the multiphase system. The properties

compared included both bulk measures such as porosity, phase saturations, and phase surface areas, as well

as pore-scale measures such as fluid/fluid interfacial areas and Euler characteristic of the nonwetting phase.

It was found that segmented images by thresholding and supervised machine learning techniques

provided close estimates of porosity, saturation, and phase fractions as that of the benchmark case.

Supervised machine learning provided better measures of fluid/fluid interfacial areas by more appropriately

capturing the three-phase contact regions. All segmentation techniques provide relatively similar estimates

for the air phase connectivity measured through the Euler characteristic, although the closest estimates were

found for the supervised machine learning case. Even though the performance of thresholding and

supervised machine learning segmentation were found to be close, supervised machine learning delivered

as a better segmentation technique for the following reasons. First, supervised machine learning showed

much better capability in capturing fluid/fluid interfacial areas compared to the other techniques. Second,

supervised machine learning required little user-dependent intervention (no image preprocessing required)

and therefore, the user-dependent time consumed for segmentation was less as compared to thresholding.

This is a significant advantage because avoiding prior denoising steps are known to damage image

structures - particularly when it is a 2-D method applied slice wise, as is the case for the implementation of

the non-local means filter. One limitation when applying supervised machine learning would be the longer

Page 84: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

64

computational time needed to train the classifier and then apply classification voxel-by-voxel in order to

segment the whole image.

The segmentation techniques presented and compared in this research are general and can be transferred

within the scope of segmenting porous media grayscale images. For each image, a separate training set of

labeled sample voxels should be collected for the supervised machine learning technique to achieve good

performance. Parameter values such as the number of classes and grid size should be adjusted for different

images when using the unsupervised machine learning technique. The voxel intensity threshold value needs

to be chosen carefully when applying the thresholding-based segmentation. The analysis presented in this

research showed rock property estimates with varying degree of sensitivity to the choice of segmentation

technique. Therefore, such selection has to be informed by the intended application and analysis of rock

properties. We recommend the utilization of the supervised machine learning segmentation techniques to

unfiltered raw images when the speed of the segmentation process is not a challenge and the accuracy of

both bulk and pore-scale measurements is a priority. Unsupervised machine learning, as well as

thresholding segmentation techniques, are recommended for use only on bulk measurements when the

speed of the segmentation or computational resources are a challenge.

Page 85: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

65

CHAPTER 3. EQUATION-OF-STATE AND CAPILLARY

NUMBER

Preface

The contents of this chapter were originally presented at ECMOR XVI, 16th European Conference on the

Mathematics of Oil Recovery, 3-6 September, Barcelona, Spain, 2018. The manuscript was accepted for

publication in the Computational Geoscience Journal and are referenced as,

Purswani P., Tawfik, M.S., Karpyn Z.T., Johns R.T. (2019)

On the Development of a Relative Permeability Equation of State, Comput. Geosci., (24), 807-818,

https://doi.org/10.1007/s10596-019-9824-2

Author contributions: Johns R.T. conceptualized the state function approach. Purswani P. processed the

data set used in the model. Purswani P. and Tawfik M. S. developed the modeling efforts and wrote the

original draft in consultation with Johns R.T. and Karpyn Z.T. All coauthors contributed toward analyzing

the data and updating the manuscript.

Abstract

Standard compositional simulators use composition-dependent cubic equations-of-state (EOS), but

saturation-dependent relative permeability and capillary pressure. This discrepancy causes discontinuities,

increasing computational time and reduced accuracy. In addition, commonly used relative permeability

models such as the Corey model, are empirical functions of phase saturation, where the effect of other pore-

Page 86: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

66

scale phenomena and rock-fluid interaction is incorporated into the tuning parameters. To rectify this

problem, relative permeability has been recently defined as a state function, so that it becomes

compositional dependent and single valued. Such a form of the relative permeability EOS can significantly

improve the convergence in compositional simulation for both two and three-phase flow.

This chapter revisits the recently developed EOS for relative permeability by defining relevant state

variables and deriving functional forms of the partial derivatives in the state function. The state variables

include phase saturation, phase connectivity, wettability index, capillary number, and pore topology. The

developed EOS is constrained to key physical boundary conditions. The model coefficients are estimated

through linear regression on data collected from a pore-scale simulation study that estimates relative

permeability based on micro-CT image analysis. The results show that a simple quadratic expression with

few calibration coefficients gives an excellent match to two-phase flow simulation measurements from the

literature. The goodness of fit, represented by the coefficient of determination (R2) value is 0.97 for relative

permeability at variable phase saturation and phase connectivity, and constant wettability, pore structure,

and capillary number (~10-4). The quadratic response for relative permeability also shows excellent

predictive capabilities.

Page 87: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

67

3.1. Introduction

Multiphase flow in porous media is of great interest in a wide array of applications including hydrocarbon

recovery, groundwater resource utilization (Nadafpour and Rasaei 2014; Parker 1989), CO2 storage (Bachu

and Bennion 2008), aquifer remediation (Chang et al. 2009; Gerhard and Kueper 2003), and two-phase

flow in proton-exchange membrane fuel cells (Akhgar et al. 2017; Lamanna et al. 2014). Each of these

applications is governed by a multitude of underlying pore-scale phenomena, such as Haines jumps (Berg

et al. 2013; Haines 1930), snap-offs (Roof 1970; Singh et al. 2017), corner flow (Mohanty et al. 1987),

capillary and viscous fingering (Nadafpour and Rasaei 2014), diffusion and dispersion (Sahimi and Imdakm

1988), dissolution/precipitation (Al-Khulaifi et al. 2017; Menke et al. 2017), and wettability alteration

(Purswani et al. 2017; Purswani and Karpyn 2019; Zhang and Austad 2006).

Considering the breadth of multiphase flow literature, studies can be classified into two categories:

macroscopic and microscopic. In macroscopic studies, averaged transport properties, such as relative

permeabilities and capillary pressure are measured on core samples to capture the effect of these flow

properties on macro-scale properties of interest, such as oil recovery. The averaged transport properties at

the core scale act as a proxy to pore-scale processes, which govern the multiphase flow process.

Standard compositional simulation employs the principles of continuum mechanics to model

multiphase flow in complex porous media, using averaged transport properties. The most commonly used

two-phase relative permeability model in commercial compositional simulators are the Corey-type models,

which are saturation-based relations that empirically carry the information of other key controlling

parameters that affect relative permeability. As explained in Khorsandi et al. (2018), this has several

limitations, including the necessity of phase labeling, which causes discontinuities as a phase disappears or

changes to another phase, causing serious convergence and stability problems (Khorsandi et al. 2018). In

addition, in order to include the effect of hysteresis, modification of the Corey model is used (Land 1968).

The modified model, however, assumes that phase saturation follows a particular path. Saturation history

is not sufficient to consider the effect of phase distribution on multiphase flow (Majid Hassanizadeh and

Page 88: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

68

Gray 1993). Khorsandi et al. (2017) proposed a new functional form for compositionally dependent relative

permeability based on the state function concept that eliminates the need for phase labeling (Khorsandi et

al. 2017). They demonstrated excellent predictive capability even for complex hysteretic flow. Khorsandi

et al. (2018) then proposed a new compositional simulation approach that eliminated the inconsistencies

caused by saturation-based transport properties, demonstrating significant improvement compared to

current commercial simulators, resolving stability and convergence issues, as well as increased robustness

and accuracy (Khorsandi et al. 2018).

With the advancement in x-ray micro computed tomography (CT), the technology is utilized in several

microscopic multiphase flow studies, allowing for visualization and quantification of previously theorized

pore-scale processes (Berg et al. 2016; Celauro et al. 2014; Herring et al. 2017; Herring et al. 2016; Karpyn

et al. 2010; Landry et al. 2014; Pak et al. 2015). Fast synchrotron-based x-ray tomography allows for real-

time visualization of phenomena like cooperative pore filling, corner filling, droplet fragmentation, snap-

off and (Armstrong et al. 2016; Avraam and Payatakes 1999; Berg et al. 2016). Avraam and Payatakes

(1999) were among the first to propose four main flow regimes observed during multiphase flow through

porous media via 2D micromodel experiments. These included small and large ganglion dynamics, drop

traffic flow, and connected pathway flow. It was observed that as the flow rate of the wetting phase

increased, the flow regime shifted from large ganglion dynamics to small ganglion dynamics to drop traffic

flow to connected pathway flow. A simultaneous increase in relative permeabilities was observed (Avraam

and Payatakes 1999). These findings have been recently corroborated through experiments and simulation

studies by Armstrong et al. (2016).

In more recent studies by Pak et al. (2015) and Khishvand et al. (2016), the authors conducted micro-

CT experiments to visualize and quantify the trapped nonwetting phase structures at variable capillary

numbers. In addition, it was observed during drainage cycles that at higher flow rates, the number of

individual nonwetting phase clusters increases. The authors also noted the importance of the pore structure

by observing that droplet fragmentation was not severe in homogenous rocks like sandstones compared to

the more complex carbonate pore structures (Pak et al. 2015). In such studies, researchers have adopted

Page 89: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

69

various quantitative approaches for characterizing phase connectivity, including Euler characteristic,

coordination number, percolation threshold, fractal dimension, (Blunt 2017) and specific fluid/fluid and

fluid/solid interfacial areas (Landry et al. 2014; Landry et al. 2011).

Micromodel experiments were performed by Osei-Bonsu et al. (2020) to generate phase connectivity

maps of three-phase flow systems useful for water alternating gas (WAG) enhanced oil recovery (EOR)

processes. Two WAG cycles were compared by considering oils of different viscosities and another WAG

injection was performed with surfactant added to the aqueous phase. Different measures for oil/water/gas

phase connectivity were compared, namely, Euler characteristic, normalized Euler characteristic ( ˆ ) ,

specific surface areas of the fluid phases, and wetted fractions of the solid (I). They found that the majority

of oil was recovered during the water injection step. The gas injection step, instead, assisted in fragmenting

the oil phase, which was then subsequently recovered by the water flow in a double-drainage fashion. The

trapped gas phase also prevented the formation of water channeling. Visualization maps of gas phase

connectivity corresponded well with the numerical measure of normalized Euler characteristic and therefore

this metric was found to be a better measure of phase connectivity. The gas phase showed smaller

saturations but higher connectivity due to larger-sized ganglia than that of oil or water. It was also found

that the wetting fraction of the solid (I) was useful toward describing the state space (S, , I) of the phases

involved which may be useful for resolving phase labeling issues during compositional simulations.

There is enough evidence in the literature that suggests that multiphase flow in porous media is affected

by rock properties: rock mineralogy, surface roughness, pore geometry, pore topology, heterogeneity; fluid

properties: viscosity, density; and rock-fluid and fluid/fluid interactions: wettability, adsorption,

precipitation, chemical dissolution and interfacial tension (Blunt 2017). Avraam and Payatakes (1995),

recognized the importance of the dependence of relative permeability on various parameters, in addition to

fluid saturation, such as capillary number, viscosity ratio between injected and displaced fluid, bond

number, advancing and receding contact angles, coalescence factor, pore geometrical and topological

Page 90: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

70

factor, and the history of flow. Within reason, a number of these parameters were varied to evaluate their

impact on fluid distribution and relative permeabilities (Avraam and Payatakes 1995).

Khorsandi et al., (2017) proposed an equation-of-state (EOS) approach for modeling relative

permeabilities as a state function. The main advantage of this approach is that it is physically based and

ensures a single valued solution for relative permeability. The contributing parameters that affect relative

permeabilities were saturation of the fluid phases, phase connectivity, capillary number, wettability of the

medium, and the pore structure of the medium. The authors evaluated the importance of phase saturation

and connectivity on relative permeabilities and found a good match against experimentally published data

(Khorsandi et al. 2017). However, there was no discussion about the verification of the EOS being a valid

state function nor its validity at limiting boundaries of the state variables.

In this research, we present a structured workflow for the development of an EOS for relative

permeability using a response surface modeling approach. We define relevant boundary conditions to

physically constrain the EOS under limiting conditions and derive functional forms for the partial

derivatives. For this development we implement similar state variables proposed by Khorsandi et al. (2017).

The calibrating parameters in the final form of the EOS are determined through linear regression on the

data presented in the recent literature that presents measurements of phase saturation, phase connectivity,

and relative permeabilities. In the following sections we outline the development of the model, provide the

description of the boundary conditions, and present the results that show the fit of the model to the literature

data.

Page 91: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

71

3.2. Methodology

3.2.1. Development of a state function

A state function, by definition, is a property whose value depends only on the condition or state of the

system irrespective of the path taken to reach that state (Cleveland and Morris 2014; Sandler 1989). This

implies that for relative permeability to be a state function, it must only have one value at a given set of the

variables considered. To satisfy this condition, the developed relative permeability function must be an

exact differential. For an exact differential, the partial derivative coefficients must satisfy the Euler

reciprocity relation (Osborne 1908). In thermodynamics, for a property that is a function of n independent

variables to be an exact differential, it must satisfy EN reciprocity relations, where

EN is the number of

conditions given by:

1

1

  .n

E

i

N i−

=

= (3.1)

For a state function 1 1 2 2 n ndQ f dx f dx f dx= + + , the conditions will be of the form:

,,

 

kk

ji

j i x k ix k j

ff

x x

=

, (3.2)

for i 1: 1n − ; j 1:i n + ; k 1: n .

Another condition contributing to the validity of a state function is whether the state variables

considered are independent. Properties are considered independent of each other when one property can be

Page 92: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

72

varied while all other input properties are held constant. On analyzing microscopic multiphase flow

experimental studies in the literature, it is observed that saturation and phase connectivity, are independent

variables except when saturation is exactly one (e.g., Schlüter et al. (2016)). At the same saturation value,

multiple fluid configurations can exist leading to widely different phase connectivity.

As discussed previously, there are numerous pore-scale variables that may contribute to changes in

relative permeability. Hence, given the complex nature of the problem, there are many degrees of freedom

that can be specified to fully define relative permeability as an EOS. Including all the parameters that

contribute to changes in transport properties in the state function would theoretically result in an exact

match of the literature data. Practically, however, a very complex model would be required to account for

all state variables that affect relative permeability. Therefore, in this research, we use the minimum number

of variables that sufficiently define the state of the system, exhibiting a good match with literature data, and

allowing for reliable relative permeability predictions, with an acceptable degree of accuracy.

Khorsandi et al. (2017) proposed an equation-of-state approach (Eq. (3.3) and Eq. (3.4)) to calculate

the change in relative permeability as a function of five measurable, pore-scale state variables (Khorsandi

et al. 2017).

( )ˆ, , , ,r j j j j cak f S I N = . (3.3)

Expressing Eq. (3.3) in exact differential form,

     ˆ  ˆ

r j r j r j r j r j

r j j j j ca

j j j ca

k k k k kdk dS d dI dN d

S I N

= + + + +

, (3.4)

Page 93: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

73

where Sj represents saturation of phase j; ˆj represents connectivity in terms of the normalized Euler

characteristic of phase j; Ij is the wettability index of phase j; NCa is the capillary number; and is the pore

structure.

At constant wettability (dIj = 0), constant pore structure (dλ = 0), and constant flow rate and fluid

properties (i.e.: dNCa = 0), Eq. (3.4) reduces to,

ˆ

ˆˆ

 

j j

r j r j

r j j j

j j S

k kdk dS d

S

= +

. (3.5)

For the simplified relative permeability state function defined in Eq. (3.5), only one reciprocity condition

(see Eq. (3.1)) must be honored as,

ˆ

ˆ ˆj j

r j r j

j j j j S

k k

S S

=

. (3.6)

By forcing relative permeability to satisfy Eq. (3.5), we ensure that there is only one value of relative

permeability as a function of two variables, while simultaneously capturing the essential physics. The error

introduced by this approach is minimized by tuning to literature data.

3.2.2. Phase connectivity

The Euler characteristic depends on saturation, saturation history, pore topology, and the scale of the

measurement. To allow for better comparison based on phase connectivity only, we must normalize the

Euler characteristic to eliminate such effects. Herring et al. (2013) developed a normalization scheme that

Page 94: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

74

eliminated the effect of pore structure by dividing χnon-wetting phase by χpore-structure, which is equivalent to the

Euler characteristic at 100% phase saturation (Herring et al. 2013). However, this only sets an upper bound

to the value of the normalized Euler characteristic, , where ˆ 1− . Khorsandi et al. (2017) modified

this normalization scheme to eliminate the effect of measurement scale and phase saturation, as well as to

set a lower bound to the value of (Khorsandi et al. 2017). In this chapter, we use a simpler expression

for the normalization of the Euler characteristic (Eq. (3.7)) such that is bounded between zero and one.

,ˆ 

  

max

min max

−=

− (3.7)

where χmax represents the limiting case for a completely dispersed phase which is expected to occur when

all the pores in the porous medium are filled with the phase, but no throats are filled to connect the pores;

and χmin represents the case where a phase is fully connected, occupying 100% of the pore space. χmin can

be easily estimated as the Euler number of the pore space from its micro-CT image, whereas χmax can be

estimated as the number of pores from the extracted pore-network of the pore space, or from the

coordination number of the rock type and its χmin value. These minimum and maximum values of the Euler

characteristic are independent of the fluid type.

Equation (3.7) defines an intrinsic connectivity parameter for a homogeneous medium that is no longer

dependent on the pore volume considered. A value of zero for the dimensionless phase connectivity means

the phase is disconnected completely, while a value of one is perfect connectivity.

3.2.3. Development of relative permeability EOS

For developing the EOS (i) must satisfy the reciprocity relation shown in Eq. (3.6), (ii) must honor physical

boundary conditions, and (iii) must be the simplest functional form possible to minimize overfitting of test

Page 95: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

75

data. Therefore, we consider that the relative permeability state function takes a simple form, such that the

partial differential coefficients are linear in ˆj and Sj. Thus, we make the relative permeability state function

a quadratic response to Sj and ˆj expressed by,

2 2

0 1 2 11 22 12ˆ ˆ   ˆ r j j j j j j jk S S S = + + + + + . (3.8)

Next, we describe the principle limiting conditions presented in Table 5 to constrain the EOS. The first

constraint is that for both saturation and phase connectivity equal to 1.0 the relative permeability must be

1.0. At phase saturation just below 1.0, the phase connectivity can theoretically vary over its entire range,

although physically only a small range is likely for a given set of variables.

Table 5. Physical constraints imposed on the relative permeability response by considering key limiting conditions that affect relative permeability as a function of phase saturation and phase connectivity. The phase is assumed nonwetting, although extensions to other phases are easily possible.

We set the phase connectivity to be 1.0 and saturation to be 0.0 for the second constraint. At low phase

saturation, in general, the phase connectivity should be low, however, for a wetting phase under extreme

wetting conditions the connectivity could be high as well. Also, for a nonwetting phase (say oil), the region

near ˆj = 1 and Sj = 0 is a plausible physical region for cases such as film drainage when two other phases

are present (say gas and water). Otherwise, it is unlikely to achieve flow near this region.

Physical Constraint Remarks

1) 1r jk = at 1 jS = and ˆ 1j = Relative permeability at complete saturation must be 1

2) 0r jk = at 0 jS = and ˆ 1j = Relative permeability at low phase saturation should be

negligible

3) 0

ˆ0

j

r j

j S

k

=

= at ˆ 1j = The change of relative permeability with phase connectivity

should be negligible near full phase connectivity

4) ˆ 0

0

j

r j

j

k

S =

=

at 1jS = The change of relative permeability with phase saturation

should be negligible near full phase saturation

Remarks

Relative permeability at 100% saturation must be 1

Relative permeability at low phase saturation should be

negligible

The change in relative permeability with full phase

connectivity should be negligible near full phase connectivity

The change in relative permeability with full phase saturation

should be negligible near 100% saturation

Page 96: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

76

The third and fourth constraints are set to ensure that the partial derivatives are positive over the entire

ˆj and Sj space. These constraints could be removed if more experimental data is available to improve the

values of relative permeability in regions near these limits. We found it necessary to include these

constraints for the data examined in this chapter.

We did not constrain the relative permeability function at Sj = 1.0 and ˆ 0j = , as it is not physical to

reach this value of connectivity. That is, at exactly a saturation of 1.0 the phase connectivity must be 1.0 in

that it is no longer independent, but at a saturation of 0.99 and ˆ 0j = the relative permeability should be

zero. We omitted this constraint from the fitting procedure based on the recognition that complex porous

media would likely never have values near this region. It is likely that there is a limiting value of ˆj as a

function of saturation based on pore morphology and other state variables.

Upon implementing these physical constraints, we obtained the final form of the model shown below.

( ) ( )ˆ ˆ ˆ ˆ2 2

r j 11 j j 22 j j j j j jk = α 1- 2χ + χ +α -2S +S + χ S + χ S . (3.9)

The coefficients 11 and

22 are determined through linear regression on measured data. Evaluation of

the partial derivatives of Eq. (3.9) gives,

( )11 22 11ˆ 12 2  

ˆj

r j

j j

j S

kS

= + −

+

, (3.10)

( ) 222 22

ˆ

2ˆ 1 2 2  

j

r j

j j

j

kS

S

+= + −

, (3.11)

Page 97: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

77

Euler reciprocity shows that these derivatives define a state function (Eq. (3.6). That is,

ˆ

221ˆ ˆ

j j

r j r j

j j j j S

k k

S S

= =

+

. (3.12)

The exact differential form (Eq. (3.5)) then becomes from Eqs. (3.10) and (3.11),

( ) ( ) 222 211 22 1 2 21ˆ ˆ ˆ12 2  12 2  r j jj j j j jd S d S dk S = + − + + − + +  . (3.13)

3.2.4. Comparison to the development in Khorsandi et al. (2017)

The relative permeability EOS proposed by Khorsandi et al. (2017) is shown below (Khorsandi et al. 2017),

( ) kn

r j k j r jΦk C Φ= − , (3.14)

where jΦ is the phase distribution term defined as

jjS

+ ;

r jΦ is the residual phase distribution of phase

j; Ck, αφ, and nk are tuning parameters. It was assumed in this formulation that the ratio of the two partial

differential coefficients was constant. The form of Eq. (3.14) allowed for direct use of the Corey model,

while also making the equation simple. Although Eq. (3.14) satisfies reciprocity, it does not satisfy all

boundary conditions in Table 5 and likely is not reliable except near the tuned experimental data.

We set the value of nk to 2 in Eq. (3.14) to get,

( )2

2 2 2 2 2 2j j jr j k j r j j j r j r jk S Φ ΦS ΦC S = + + + − − . (3.15)

Page 98: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

78

Comparing our development in Eq. (3.9) to Eq. (3.15) we see,

22kC = , (3.16)

11

22

= , (3.17)

11

22

r jΦ

= . (3.18)

More complicated cubic or higher order polynomial equations could also be used, but the simplest form

that reasonably matches experimental data and boundary conditions is preferred to avoid over-fitting. Most

importantly, the EOS developed in this chapter honors the physical boundary constraints presented in Table

5. The response surface formulation provides justification for Eq. (3.14), although future research could

define a form of Eq. (3.14) that honors physical constraints like those in Table 5.

Page 99: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

79

Figure 15. Illustration showing the EOS state approach on a real path (simulation) taken during two-phase flow simulation in

jS and ˆj space.

In Figure 15, we illustrate that once the EOS is determined, the real path from the initial to the final

state can be separated into a constant saturation path followed by a constant phase connectivity path or vice

versa to arrive at the same final state. In this way, relative permeability can be calculated by integrating the

individual partial differential coefficients to arrive at the final state’s relative permeability. The shortcoming

of such an approach is that it requires phase connectivity values at the initial and the final state, which is

not readily available unless sophisticated techniques such as x-ray micro-CT are implemented. One way of

overcoming this shortcoming is to determine a functionality between phase connectivity and phase

saturation so as to bypass the dependence of relative permeability on phase connectivity. This provides an

avenue for future research. Khorsandi et al. (2017) solved this problem by assuming that the change in

connectivity with saturation is constant for any drainage path, and similarly for any imbibition path. They

used simple but different models for drainage and imbibition and tuned them to available data. From these

fixed tuned models, they could predict hysteretic scans that began at different saturations.

Simulation

path

( )11 22 11ˆ 12 2  

ˆj

r j

j j

j S

kS

= + −

+

( ) 222 22

ˆ

2ˆ 1 2 2  

j

r j

j j

j

kS

S

+= + −

Page 100: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

80

3.2.5. Estimation of the coefficients of the EOS

The data set used for estimating the coefficients of the model in Eq. (3.9) is from (Armstrong et al. 2016).

In their paper, the authors coupled experimental research with simulations to study the effect of phase

topology on macroscopic system behavior during two-phase flow in a porous medium. Micro-flow

experiments were conducted in a sintered glass sample with chemically doped water as the wetting phase

and decane as the nonwetting phase. The two phases were co-injected at different fractions maintaining

steady state conditions, and three different flow rates were tested to represent three different capillary

numbers, namely, 10-4,10-5, and 10-6. From the segmented images acquired during micro-flow experiments,

11 different fluid configurations (each representing a different fluid saturation arrangement) were used as

the initial condition for two-phase flow simulations to determine relative permeabilities for a wider range

of capillary numbers obtained by varying fluid properties during these simulations. A 4-D connected

component algorithm was implemented to track the fluid ganglion during simulations for estimating phase

connectivity.

The full data set is displayed in Figure 16. Figure 16a shows the data for phase relative permeability

while Figure 16b shows the data for phase connectivity, which is measured as normalized Euler

characteristic using Eq. (3.7). The phase shown here is the nonwetting phase because the Euler

characteristic for the wetting phase was not reported.

To calculate the normalized Euler characteristic values ˆj , χmax and χmin need to be known. Because the

number of pores and the Euler number of the pore structure was not reported, a 2D extrapolation was carried

out on the data for NCa = 1, which showed the maximum and minimum values for the Euler number. Large

capillary numbers imply very low interfacial tension, which may explain the largest and smallest

connectivity values observed.

Page 101: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

81

Figure 16. Phase saturation, relative permeability, and normalized Euler connectivity for the nonwetting phase for different capillary numbers used for fitting the quadratic response for relative permeability as well as for prediction purposes. Courtesy Dr. Ryan T. Armstrong. Data from Armstrong et al. (2016). This data is tabulated in appendix D.

We assumed a planar relationship among krj, Sj, and ˆj for the 2D extrapolation,

.ˆjr j jk AS B C= + + (3.19)

Three point-extrapolation was carried out such that the extrapolated values were near the actual data as

opposed to extrapolation on the entire data, which could lead to errors in estimation. The first, three data

points of the Euler number were used for determining χmax, where krj was set to zero, while the last three

data points were used for determining χmin, where krj was set as one. Simultaneously, the same three data

points on either end were used for fitting lines through Sj and ˆj for fixed NCa, which were then intersected

with the plane (Eq. (3.19)) to estimate χmax and χmin values. The extrapolated values for the Euler

characteristic of the pore structure are shown in Table 6.

Table 6. Euler characteristic values estimated through 2-D extrapolation for the pore structure used during simulations in Armstrong et al. (2016).

a b

c

max 5788

min -10,704

Page 102: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

82

3.3. Results and discussion

In this section, we present the results for the fitted quadratic response for relative permeability. The sub

data set of capillary number ~10-4 was used for response surface fitting. This fit was used to predict data

sets at different capillary numbers. The goodness of fit is evaluated using residual error and R2 values.

Further, we present the partial derivative coefficient to the exact differential of relative permeability as a

function of phase saturation and phase connectivity. Finally, we compare response surfaces generated for

different capillary numbers using linear regression on the individual sub data sets to evaluate the impact of

capillary number.

3.3.1. Quadratic response for relative permeability

We used linear regression with Matlab® to find the coefficients in the proposed model described in Eq.

(3.9) that best fits the data set at the fixed capillary number of ~10-4. The base code is provided in appendix

C. Table 7 provides the information for the fitting parameters and the goodness measure of the fit.

Table 7. Model coefficients and the goodness of quadratic response surface fit to phase saturation and phase connectivity to the data presented in (Armstrong et al. 2016) at the NCa of ~10-4.

The contour map of the response surface fit is shown in Figure 18. Contour map of the response surface

of relative permeability as a function of phase saturation and normalized Euler connectivity. The capillary

number (~ 10-4), wettability, and pore structure have been kept constant. Data points shown as black dots

were taken from the two-phase flow simulations presented in Armstrong et al. (2016). Dashed line

11 -0.229

22 -0.589

R2 0.971

Root mean squared error 0.147

Page 103: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

83

represents a limiting boundary of plausible values. The dots represent the data points at the fixed capillary

number of ~10-4 used for estimating the coefficients for the response surface. The corresponding plot for

the residual error is shown in Figure 17. As shown, the quadratic response gives small residual error values

scattered around zero with a mean of -0.009, showing little systematic error.

As shown in Figure 18, the general trends of relative permeability versus saturation and relative

permeability versus phase connectivity are honored. The contour map gives the relative permeability value

for a known value of normalized phase connectivity and its corresponding value of phase saturation,

irrespective of the path/direction a particular experiment/simulation may take. The contour map is also

independent of the phase label (gas, oil, or water, for example). This is valid for the fixed wettability, pore

structure, and capillary number used in the development of this quadratic response.

Figure 17. (a) Quadratic response prediction versus simulation data (b) residual between the predicted and simulation measurements for relative permeability based on the response surface fit shown in Figure 18.

a b

Page 104: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

84

Figure 18. Contour map of the response surface of relative permeability as a function of phase saturation and normalized Euler connectivity. The capillary number (~ 10-4), wettability, and pore structure have been kept constant. Data points shown as black dots were taken from the two-phase flow simulations presented in Armstrong et al. (2016). Dashed line represents a limiting boundary of plausible values.

We present a notional boundary on the contour plot in Figure 18, which represent the limits of possible

physical experimental/simulation conditions. The region below and to the right of the curve are extreme

cases controlled by the topology of the rock structure itself, as well as other variables such as wettability.

This region suggests that even at very high phase saturation, the phase remains extremely disconnected.

Such a case would be highly unusual to occur, especially in real porous media. It may occur in theoretical

porous media with a highly disconnected pore structure, or a pore structure with a very large aspect ratio

between the pore and connecting throats so that very high capillary forces are required for the phase to pass

through. Since there is insufficient data in this region, the prediction from our quadratic response fit may

not lead to conclusive results for this region. A similar “unrealistic” region could be present in the upper

left corner of the contour map, although this is not shown in Figure 18. To achieve low saturations with

high connectivity would require very thin wetting films or spreading of an intermediate wetting phase (in

three-phase systems), where even at very low phase saturation a phase remains highly connected. Although

Page 105: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

85

the phase would remain highly connected, the relative permeability would be small in this region of jS

and ˆj space owing to small saturation. More experimental studies should be conducted under extreme

conditions and varying wetting states and pore topology of the medium to acquire more complete data sets

to enhance predictive capabilities of these cases and to capture the loci of zero relative permeabilities in Sj

and ˆj space.

In Figure 19, the partial differential coefficients expressed in the final form of the exact differential

(Eq. (3.13)) are shown as a function of phase saturation (Figure 19a) and phase connectivity (Figure 19b).

Both partial derivatives with respect to relative permeability are always positive, suggesting that relative

permeability increases with an increase in Sj as well as ˆj . We further observe that the rate of increase in

relative permeability decreases with increasing saturation whereas, the rate of increase of the relative

permeability with increasing phase connectivity increases, although this increase is minor and plateaus near

 ~ 0.55ˆj . This suggests that the effect of an increase in phase saturation on relative permeability slowly

declines, while the effect of phase connectivity grows. This is consistent with our understanding that for a

phase to be sufficiently connected in the porous medium some phase saturation should exist.

Figure 19. Partial derivative coefficients (calculated using Eqs. (3.10) and (3.11)) expressed as a function of (a) phase saturation and (b) normalized Euler connectivity.

a b

Page 106: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

86

3.3.2. Quadratic response prediction at neighboring conditions

We now present the predictive capability of the fitted response surface to neighboring conditions. We use

the same data set by Armstrong et al. (2016) but at capillary numbers of ~10-3 and ~10-5. This ensures that

the pore structure and wetting conditions remain the same between the fitted and the predicted cases. The

plots for the response and the corresponding residual errors are shown in Figure 20.

Figure 20. (a) Prediction of relative permeability and (b) residual error for capillary numbers ~10-3 and ~10-5 based on the response surface fit to capillary number 10-4 described in Figure 18.

Figure 20 shows that the predicted response fits the data well. The residual errors between the predicted

and actual relative permeability values show little systematic errors and R2 values near ~0.94 for both

capillary numbers. It is likely that capillary number impacts these values, as is discussed in the next section

in more detail.

3.3.3. Effect of capillary number

To capture the effect of capillary number on relative permeability we use the surface fits to predict relative

permeability at different capillary numbers ranging from ~ one to ~10-6. The goodness measure of these

prediction cases is shown in Figure 21.

a b

Page 107: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

87

Figure 21. R2 error for prediction of data at different capillary numbers using sub data set at capillary number of ~10-4 as the fitted response surface.

From Figure 21 we see that the R2 value showing the goodness of fit is the maximum for capillary

number ~10-4 marked by the dashed red line. This is because the regression was carried out using this

capillary number sub data set. As stated earlier, the R2 values in the neighborhood of the fitted response are

excellent at about 0.94, however, as we move two to three orders of magnitude away from the original

capillary number, the prediction with NCa ~10-4 leads to erroneous values. This clearly suggests the

importance of capillary number as a parameter that affects relative permeability. This is also shown in

Figure 22 for the fitted quadratic responses to individual sub data sets at different capillary numbers (see

Table 8 for fitting parameters). As capillary number increases the response surface becomes more planar

showing that the dependence of relative permeability on phase connectivity is reduced significantly while

relative permeability becomes more sensitive to the change in saturation. The occurrence of these

observations can be quantitatively observed in Table 8, where the magnitude of α11 term decreases

significantly at higher capillary number. A sharp decrease in the magnitude of this term explains the reduced

effect of phase connectivity (see Eq. (3.9)). Overall, these observations are consistent with those observed

by Armstrong et al. (2016), where connectivity was found to be larger at greater capillary numbers and

Page 108: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

88

therefore relative permeability becomes more strongly dependent on phase saturation (Armstrong et al.

2016).

Figure 22. Quadratic response surface fits to sub data sets at different capillary numbers.

Table 8. Coefficients for the quadratic response and goodness measures for quadratic response surface fits to sub data sets at different capillary numbers shown in Figure 22.

Nca = ~10-6

Nca = ~10-4

Nca = ~100

Coefficients for the

quadratic response and

goodness measure

Capillary number

~100 ~10-4 ~10-6

11 -0.009 -0.229 -0.517

22 -0.806 -0.589 -0.667

R2 0.994 0.971 0.863

Root mean squared error 0.064 0.147 0.249

Page 109: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

89

Figure 23. Phase relative permeability plots with corresponding phase connectivity value (shown in blue) (a) high capillary number ~1 (b) low capillary number ~10-5. The red solid lines represent the fit using Corey model with ~ exponent value of (a) 0.76 and (b) 1.29. The residual saturation in (a) was set to 0 while computing the Corey exponent because that data point was not known.

In Figure 23, we show typical Corey fit to the data at extreme capillary numbers. The fit to the exponent

was < 1 for the high capillary number case and > 1 for the low capillary number case. These exponents

govern the curvature of the relative permeability change with respect to saturation, which causes a slower

or faster increase in relative permeability as saturation changes. These plots reveal the inherent importance

of phase connectivity implicitly assumed in Corey’s approach for fitting relative permeability. The

slower/faster change in relative permeability as seen in Figure 23 is the result of the changes in phase

connectivity. Such relative permeability curves are also observed during microemulsion/excess oil and

microemulsion/excess brine relative permeability measurements conducted by Delshad et al. (1987) where

the microemulsion phase relative permeability were observed to increase sharply (Delshad et al. 1987).

This increase was attributed to the wettability and low interfacial tension of the microemulsion phase, which

in principle improved connectivity to increase relative permeability.

Under limiting conditions of NCa approaching ∞, for example, at negligible interfacial tension between

the phases, the two phases will collapse to a single phase. At this point, S and will approach the limiting

value of one and relative permeability would also approach one.

a b

Page 110: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

90

3.4. Concluding remarks

In this chapter, we present the development of a physically-based quadratic state function for relative

permeability in phase saturation and connectivity. The coefficients of the EOS are determined through

linear regression on two-phase flow simulation data from the literature. The following conclusions can be

drawn under the assumptions in which this study is conducted.

• A simple quadratic response for relative permeability gives an excellent fit to simulation data at

fixed capillary numbers.

• The quadratic response fit acquired from one data set shows excellent predictive capabilities at

similar flow conditions. However, away from original conditions, the predictive capability of the

kr response surface decreases owing to the dependence on capillary number.

• Connectivity increases faster at low saturations than at high saturations for high capillary numbers.

This explains the small Corey exponents obtained for ultra-low interfacial tension corefloods. The

reverse is true for small capillary number where capillary effects dominate.

Although the model presented in this chapter may not be the only solution, our approach was designed

to seek the simplest EOS that honors key limiting physical constraints and provides a reasonable fit to the

data from the literature. The presented approach is mathematically simple, which is advantageous for

potential use in compositional simulation. The assumption of a quadratic response surface may not fit all

relative permeability data, where higher-order polynomials may be needed. Nevertheless, the response

surface approach developed here justifies the relative permeability functionality presented by Khorsandi et

al. (2017) for an exponent of 2.0. In addition, one of the main outcomes of using the developed EOS is being

able to eliminate discontinuities resulting from using saturation-dependent relative permeability

correlations in compositional simulators, which increases computational time and reduces accuracy. This

would be especially useful to account for complex processes such as mass transfer and dynamic changes

that often occur during transport of multiple phases in real porous media.

Page 111: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

91

CHAPTER 4. IMPACT OF WETTABILITY ON PHASE

TRAPPING

Abstract

Capillary trapping is an important carbon capture strategy for mitigating greenhouse gas emissions.

Experimental investigations suggest that multiple factors like pore structure, surface roughness, wettability,

and CO2 dissolution can impact capillary trapping. Among these, reservoir wettability is of great

significance as it impacts phase distribution and consequently phase trapping. We use pore-network

modeling to investigate the effect of wettability on phase mobility and capillary trapping. For this, we

recreate strong wetting and nonwetting conditions during the secondary flooding cycles. Different initial

phase saturations are investigated, and the trends of residual phase saturation are studied for each wettability

case. We also present residual phase connectivity trends to explore the impact of wettability on the locus

of residual phase connectivity and phase saturation. Finally, we propose simple models to capture all

residual trends.

Simulations show that when the receding phase is the wetting phase, there is reduced trapping of that

phase which occurs due to combinations of layer flow of the wetting phase and piston-like advance by the

nonwetting phase. Such flow regimes cause nonlinearity in the characteristic initial-residual (IR) saturation

curve with a maximum trapped (residual) saturation of ~0.5 corresponding to initial saturation ranging

between 0.7-0.95. A new extended Land-based model matches the IR trends for all wettabilities.

Simulations also show that the loci of residual phase saturation and phase connectivity, are a strong function

of wettability. At strongly nonwetting conditions, the locus remains fairly constant in phase connectivity,

while at increasingly wetting conditions, the locus shows trends that resemble a closed-loop and the size of

the locus changes. The limiting values of the residual locus are found to be dependent on the pore topology.

Page 112: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

92

The residual locus of phase connectivity and saturation provide a true limiting condition useful for reservoir

modeling. Through network-modeling, this research discusses the impact of wettability on residual trapping

of a phase, which is critical for evaluating the success of long-term CO2 storage projects.

DISCLAIMER

This work was funded by the Department of Energy, National Energy Technology Laboratory, an agency

of the United States Government, through a support contract with Leidos Research Support Team (Neither

the United States Government nor any agency thereof, nor any of their employees, nor LRST, nor any of

their employees, makes any warranty, expressed or implied, or assumes any legal liability or responsibility

for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed,

or represents that its use would not infringe privately owned rights Reference herein to any specific

commercial product, process, or service by trade name, trademark, manufacturer, or otherwise, does not

necessarily constitute or imply its endorsement, recommendation, or favoring by the United States

Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily

state or reflect those of the United States Government or any agency thereof.

Page 113: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

93

4.1. Introduction

If global warming were to persist at current rates, global temperatures may rise by 1.5 °C by 2035 and 2052.

Further elevated levels of warming may cause irreversible damage to the environment (IPCC 2018).

Greenhouse gas (GHG) emissions, such as those of anthropogenic carbon from industrial and energy

sources contribute to global warming. Therefore, long-term sequestration of carbon dioxide in deep

subsurface formations will be an important mitigation strategy for combating emissions of GHGs (Orr

2009; Schrag 2009) In order to sustain energy demand, implementation of carbon capture strategies as an

integral part of the evolving energy mix can ensure energy security for future generations (Dammel et al.

2011; Szulczewski et al. 2012).

Different mechanisms of CO2 sequestration have been identified (Iglauer et al. 2015; IPCC 2005;

Krevor et al. 2015). These include structural trapping of buoyant CO2 plumes under low permeability

caprocks; residual trapping and local capillary trapping caused as entrapment of blobs/ganglia of CO2

because of capillary forces (Pentland et al. 2010; Pentland et al. 2011); dissolution trapping of CO2 in saline

aquifers (Iglauer et al. 2011a; Iglauer et al. 2011b); or mineral trapping of CO2 caused as carbonate mineral

precipitation. Dissolution and mineral trapping are slow processes and would occur over prolonged time

scales; but residual trapping can be achieved faster and thus will be an effective strategy for decades to

come (IPCC 2005). Residual trapping is also attractive due to the availability of the necessary infrastructure

in the energy industry as well as the storage capacity to scale carbon capture. A 2-D schematic

representation of capillary trapping is shown in Figure 24. The schematic depicts heterogeneity in the

porous medium in terms of local pore/grain size variations, mineral complexities, and local variations in

wettability. The moving front of the flowing chase brine leaves capillary trapped globules of CO2 behind.

Page 114: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

94

Figure 24. Schematic representation of capillary trapped CO2 by chase brine. Local heterogeneities in pore structure, mineral complexity, and wettability are depicted.

Laplace’s equation on capillary pressure across the interface between two immiscible fluids, in a

capillary tube geometry, shows that interfacial forces acting on the fluid/fluid/solid system and the

capillary’s cross-section govern the entry pressure through the opening. This implies that both wettability

and pore structure are critical in dictating pore occupancy by the CO2 phase and thus the amount of trapped

CO2. A variety of other factors have also been recognized to impact phase trapping. These include, the

initial CO2 saturation before brine flooding; pore-throat aspect ratio, pore and grain size distribution

(Jerauld and Salter 1990), heterogeneity in porosity and permeability (Krevor et al. 2011; Ni et al. 2019;

Perrin and Benson 2010) rock and fluid interactions such as rock wettability; chemical interactions such as

wettability alteration (Anderson 1987; Salathiel 1973; Tanino and Blunt 2013), rock dissolution, salt

precipitation, or dissolution of CO2 into the brine; as well as mineral heterogeneity of the rock (Iglauer et

al. 2015). Of these factors, the impact of rock wettability on capillary trapping remains a topic of great

interest in the porous media research community.

One reason for the importance of wettability is that it directly affects phase distribution (Blunt et al.

2020; Sun et al. 2019) which can consequently impact capillary trapping. This can have implications for

capillary trapping

CO2 plume

Chase brineLeading end of the CO2 plumeTrapped CO2

Trailing end of the CO2 plume

1

Page 115: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

95

the evaluation of the ultimate quantity of carbon that can be successfully sequestered. Thus, wettability can

govern the suitability of storage sites like aquifers or depleted hydrocarbon reservoirs (Levine et al. 2014).

It is likely that the more nonwetting the formation is to CO2, the larger the trapping capacity, given that

nonwetting phase occupies the centers of pores (Blunt 2017). In addition, wettability is critical for

preserving caprock integrity and CO2 leakage because if caprock minerals become altered to a CO2 wetting

state, the sealing capacity as measured by the entry capillary pressure will be reduced. For example, Chiquet

et al. (2007) found that the wettability to brine in a CO2/brine/mineral system reduced under typical

reservoir pressures. More substantial wettability alteration was observed on mica as opposed to quartz. In

addition, spontaneous imbibition could give rise to preferential leakage pathways for the CO2 to escape.

Although such leakage is deemed to less concerning in comparison to those that may occur because of over-

pressurized faults or abandoned wells (Iglauer et al. 2015).

Extensive research efforts have shown that wettability of CO2/brine/solid systems depends on different

factors such as CO2 phase pressure which can influence the amount of dissolved CO2 in the brine phase

(Chiquet et al. 2007; Espinoza and Santamarina 2010). Other factors such as the brine phase salinity, the

system temperature, mineral heterogeneity, and surface roughness of the rock can all have significant

impact on the wetting preference of the rock. For detailed reviews on the topic of wettability of

CO2/brine/rock systems, the reader is referred to the works of Arif et al. 2019; Iglauer et al. 2015; and

Yekeen et al. 2020. It is largely concluded that an increase in the phase pressure of CO2 results in weakening

water-wetting for the CO2/brine/rock systems.

Bachu and Bennion (2008), and Espinoza and Santamarina (2010) observed a drop in the CO2/brine

interfacial tension with increasing CO2 pressure. This was ascribed to the increase in CO2 concentration at

the fluid/fluid interface because of an increase in the dissolved CO2. Espinoza and Santamarina (2010) also

found oil-wet surfaces to be either intermediate or nonwetting to brine in the presence of CO2. Saraji et al.

(2013) found wettability alteration to weakly water-wet conditions on quartz surface at pressures

corresponding to supercritical CO2 conditions. High temperature was also found to cause an increase in

water advancing contact angles. Despite an increase of up to 10° in contact angles, all measurements

Page 116: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

96

remained primarily under water-wet conditions likely due to the application of cleaned surfaces (Iglauer et

al. 2015; Saraji et al. 2013). Other researchers, however, have found more significant changes. For example,

Kim et al. (2012) showed through micromodel studies conducted with silica grains that wettability of a

CO2/brine/silica system can be altered to weakly water-wet conditions within few hours of CO2 injection.

They observed contact angles as high as 80° measured through the brine phase. The alteration in wettability

was attributed to the dissolution of CO2 into the brine and the associated pH drop to acidic conditions.

Similarly, weakly water-wet trends were observed for higher brine salinity which the authors ascribed to

the thinning of water films at high salinity (Kim et al. 2012). Still, wettability data of CO2/brine/rock

systems for varying salinity and temperature conditions remains sparse and less conclusive in the literature

(Arif et al. 2019; Iglauer et al. 2015).

Some researchers have qualitatively assessed the wettability of CO2/brine/rock system through relative

permeability measurements. For example, Levine et al. (2014) found the endpoint relative permeability to

CO2 to be clustered between 0.35 and 0.4, which is half that of typical water-wet reservoirs, suggesting for

weakly-water wetting conditions. This may result in decreased CO2 injectivity and consequently affect the

disposal capacity of reservoirs that may be prone to leakage at high pressures (Levine et al. 2014). Yet

others, for example, Akbarabadi and Piri (2013) have found supercritical CO2 (scCO2) as nonwetting with

an endpoint drainage relative permeability of 0.19 and cross-over brine saturation of >0.75. However, when

comparing the trapped CO2 saturation for a given initial CO2 saturation, through unsteady-state flow

experiments, the amount of gaseous CO2 that was trapped was greater than scCO2. This was attributed to

brine being more wetting in the presence of gaseous CO2 (Akbarabadi and Piri 2013).

Trapping behavior is characterized through initial-residual (IR) saturation trapping curves. This was

first proposed by Land (1968) to resolve relative permeability hysteresis by linking imbibition relative

permeability to drainage relative permeability (Carlson 1981; Killough 1976; Land 1968). IR characteristic

curves are extremely useful in reservoir engineering from the perspective of phase trapping for carbon

sequestration, and for the resolution of relative permeability hysteresis (Juanes et al. 2006; Spiteri et al.

2008). The IR curves describe the relationship between the initial phase saturation at the start of the water

Page 117: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

97

injection process (example: waterflooding) to the remaining (and/or residual) saturation of the receding

phase and are considered specific to the rock being flooded (Lake et al. 2014). This is due to the dependence

of IR curves on the pore structure and wettability of the rock. Most studies in the literature with water-wet

behavior find similar IR trends, i.e., as initial saturation increases (~ up to 0.5), the residual saturation

increase because of higher amount of nonwetting phase available for trapping, but at initial nonwetting

phase saturation of >0.5, the residual phase saturation either increase slightly or remain constant (Alyafei

and Blunt 2016; Li et al. 2015; Pentland et al. 2010; Pentland et al. 2011).

Most two-phase flooding experiments, however, are unable to capture the region of very high initial

phase saturation (>0.8) (see Fig. 10 in Krevor et al. 2015) likely due to experimental limitation of achieving

very high capillary pressures to attain such high initial nonwetting phase saturations. Rocks that are

nonwetting to brine, show a broader spread in the IR saturation curve than those of water (brine) wet (see

Figure 2 in Alyafei and Blunt 2016). One observation between water-wet and non-water-wet IR curves is

that the residual saturations following brine injection is lower for the non-water-wet systems (Alyafei and

Blunt 2016). Few researchers have observed a nonlinear trend in the IR saturation curve with an apex near

high initial saturations. For example, see experiment 4 in Pentland et al. (2010), experiment with Mt. Simon

rock in Krevor et al. (2012) and Krevor et al. (2011), water-wet Estiallades in Alyafei and Blunt (2016),

Bentheimer sandstone experiments Herring et al. (2015), pore-network simulations by Spiteri et al. (2008).

Figure 25 shows the IR curves for these experiments.

Page 118: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

98

Figure 25. Initial-residual characteristic curves for selected literature studies.

From the literature discussed here, it is evident that wettability is a critical parameter that impacts phase

trapping. Secondly, experimental works show that a range of wetting conditions are possible for the

CO2/brine/rock system. Thirdly, flow experiments provide a good basis for the trapping trends but are

unable to cover the full initial saturation and wettability ranges. Thus, limiting the ability of currently used

models in capturing trapping trends over the full saturation range. Moreover, multiple factors (CO2 pressure/

temperature/ brine salinity/ surface roughness/ mineral heterogeneity) can impact wettability

simultaneously when investigated through experiments.

Only fewer studies have focused on phase connectivity (measured as the Euler characteristic) toward

phase trapping. Herring et al. (2015) and Herring et al. (2013) identified the importance of initial phase

topology toward the residual saturation of the nonwetting phase. However, the medium studied was only

water-wetting, and also the trends of residual phase connectivity were not discussed. In the works by

Khorsandi et al. (2017), Purswani et al. (2020), Purswani et al. (2019), an equation-of-state model for

relative permeability was discussed by including phase connectivity was developed. Experimental and

numerical data sets using pore-network modeling were used to find the functional form of relative

0

0.2

0.4

0.6

0.8

0 0.2 0.4 0.6 0.8 1

Res

idu

al s

atu

rati

on

(-)

Initial saturation (-)

Krevor et al. (2011) - Mt. SimonHerring et al. (2015) - Nca = 10^-8.6Herring et al. (2015) - Nca = 10^-5.6Pentland et al. (2010) - Exp #4Alyafei and Blunt (2015) - EstialladesSpiteri et al. (2008) - Theta = 60 degSpiteri et al. (2008) - Theta = 100 deg

No productionunit slope

Page 119: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

99

permeability. Purswani et al. (2020) showed through the estimation of partial derivatives of relative

permeability that a response for relative permeability could be developed in the phase connectivity-phase

saturation space. They also discussed the locus of residual phase saturation and phase connectivity.

However, only two wettability cases, both under the water-wet regime were studied. Thus, the full spectrum

of wettability was not covered. Spiteri et al. (2008) covered the full spectrum of wettability through pore-

network modeling; however, the importance of phase connectivity was not discussed. Also, the quadratic

trapping model studied seems less adaptable to the more commonly known Land-based models.

In this chapter, we build on the works of Purswani et al. (2020) and Spiteri et al. (2008) to investigate

the importance of wettability on phase mobility and capillary trapping. We use pore-network modeling to

establish strict control on wettability. We change the contact angles during the water injection cycles to

cover the full range of wettability. Different initial phase saturations are investigated, and the trends of

residual phase saturation of the receding phase are studied for each wettability case. We also present

residual phase connectivity trends to explore the impact of wettability on the locus of residual phase

connectivity and phase saturation. We present appropriate functions for modeling the IR saturation trends

as well as the trends of the residual locus. Comparison with commonly used trapping models is also

provided.

We first provide the description of the numerical data sets generated using pore-network simulations

for studying capillary trapping. Next, we provide a description of the commonly known trapping models of

initial-residual (IR) saturation curves. We then introduce a new trapping model for the IR curves. We also

provide a model for the residual phase connectivity and thus model the locus of residual phase saturation

and phase connectivity for all wettability cases.

Page 120: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

100

4.2. Methodology

4.2.1. Pore-network simulations

We used pore-network modeling (PNM) to generate numerical data sets of phase saturation, phase

connectivity, and relative permeability for different wettability conditions. The details of the pore-network

extraction model and two-phase flow simulation model used can be found in our previous work (Purswani

et al. 2020). We used an x-ray computed tomography scan of Bentheimer sandstone (Lin et al. 2019) for

network extraction using the approach of Dong and Blunt (2009). The porosity and permeability of the

pore-network were 22% and 2500mD, respectively, while the number of pores, number of throats, and

coordination were, 16850, 42637, and 5.061, respectively. The network simulations were carried out under

capillary dominated regime (NCa < 10-4), where a saturation step change is guided by capillary entry pressure

at pore throats which governs each network element occupancy by the invading and receding phases. Figure

26 shows a sequence of steps involved in pore-network modeling.

Scanning curves of primary phase1 injection and primary phase2 injection were conducted similar to

Spiteri et al. (2008). Here, we denote the two immiscible phases as phase1 and phase2, and avoid the terms

such as wetting, nonwetting, drainage, or imbibition since wettability of the medium is variable. We also

refer to phase1 as the receding phase because this phase is removed during secondary injection. All initial

and residual saturation or phase connectivity terms discussed in subsequent sections are in the context of

the receding phase (phase1). Similarly, all contact angles in this work are also reported through phase1 for

consistency. The contact angle during phase1 injection was kept at 180° (0° for the phase2) to mimic

completely water-wetting condition, similar to the injection of CO2 in brine filled reservoirs, or oil

migration that occurs in oil fields over geological times. For the phase2 injection, we change the receding

phase contact angle from 180° to 0° (advancing contact angle from 0° to 180°), (mimicking a range of

wettability alteration scenarios, from no alteration (strongly phase2-wet) to strong alteration (strongly

phase1-wet). Most hydrocarbon reservoirs are mixed-wet because of the prolonged contact of oil with the

Page 121: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

101

rock which renders the rock surface to be oil-wet (Jadhunandan and Morrow 1995; Salathiel 1973). We

input the same physical properties (density of 1000 kg/m3 and viscosity of 1 cP) for both phases during

network simulations, although no change would occur in the simulations given the capillary pressure rule-

based execution of the pore-network model.

Figure 26. Steps in pore-network modeling. (a) The extracted internal structure (pore space) of the porous medium used in this study; (b) The extracted pore-network of the porous medium in (a) represented in a ball and stick form; (c), (d), and (e) show three saturation steps of blue phase injection in the pore-network completely filled with the red phase. (c) shows no injection, (d) is captured after some injection, and (e) is captured after longer injection period.

The advantage of using PNM is that strict control on wettability can be achieved. For receding phase

contact angles (θ1) less than 90°, the pore-network’s fluid occupancy is rearranged such that the nonwetting

phase (if occupying the corners), the arc menisci bulge outward for pore corners. In addition, the nonwetting

ba

c d e

Page 122: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

102

phase also occupies the pore centers, while the wetting phase remains sandwiched as wetting layers

(Valvatne and Blunt 2004).

For each θ1, scanning curve pairs of phase1 and phase2 injection sequences were conducted with varying

initial phase1 saturation (~0.05 to 1.0 in intervals of 0.05). This was achieved by terminating the phase1

injection process at the required saturation step. The residual phase1 saturation was achieved for each initial

phase1 saturation by driving phase2 injection until a high magnitude of capillary pressure (magnitude of 105

Pa). Phase1 connectivity was estimated for each saturation step by calculation of Euler characteristic as

(Purswani et al. 2020),

,occ occ occ

p t htn n n = + − (4.1)

where ,occp

n ,occt

n and occht

n are the number of pores, throats, and half throats occupied at each saturation step.

The phase connectivity ( ˆ ) is defined as the normalized Euler characteristic which is calculated as

(Purswani et al. 2019),

,ˆ   max

min max

−=

− (4.2)

where χmin = np – nt and χmax = np; np and nt refer to the number of pores and number of throats in the pore-

network, respectively. Therefore, in this study, χmin = -25787 and χmin = 16850. The base code to extract

phase connectivity data from PNM is provided in appendix G.

Page 123: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

103

4.3. Results and discussion

4.3.1. Iso-quality curves for saturation and connectivity – interpreting PNM data

Trends of phase saturation, phase connectivity, and relative permeability for the nonwetting phase in water-

wet media have been discussed in Purswani et al. (2020). Here, the discussion is extended over the full

range of wettability. To investigate these trends more effectively, we analyze relative permeability iso-

quality curves in the saturation and connectivity space for the receding phase (phase1) during phase2

injection, similar to how isotherms and iso-volume curves are analyzed in a pressure-temperature-volume

(PVT) phase diagram. For example, the iso-connectivity curves convey the behavior of relative

permeabilities with a change in saturation at fixed phase connectivity. These are shown in Figure 27. To

generate these plots, we perform 2-D linear interpolations on the PNM data to arrive at relative permeability

values at equidistant points (intervals of 0.01 units) in the connectivity-saturation space.

Page 124: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

104

Figure 27. Iso-quality trends for four wettability cases (θ1 = 180°, 120°, 60°, 0°). Plots to the left show the iso-saturation curves, while plots to the right show the iso-connectivity curves. The phase saturation, phase connectivity, and relative permeability are for the receding phase (phase1) during phase2 injection. The contour lines are plotted at intervals of 0.03 units. The procedure for developing these curves is displayed in Figure 51 in appendix E.

θ1

= 1

80

o

θ1

= 1

20

1=

60

1=

0o

Iso-saturation Iso-connectivity

Page 125: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

105

For θ1 = 180° and 120°, mostly flat iso-saturation contours are observed at high saturations (>0.5) (left

column of Figure 27), showing that relative permeability is less dependent on phase connectivity at high

saturation. On the contrary, for these high receding phase contact angle cases, from the iso-connectivity

plots (right plots), the importance of phase connectivity on relative permeability can be observed at

intermediate and low saturation (0.5 and below). Similar trends are discussed in chapter 5 through partial

derivative calculations. For the high receding phase contact angle cases, snap-off of the receding

(nonwetting) phase governs trapping, and at low and intermediate saturation, connectivity becomes

important because a critical snap-off event can restrict flow (see iso-saturation plots for θ1 = 180° and 120°

in Figure 27).

As contact angle decreases, θ1 = 60° and 0°, the receding phase becomes the wetting phase. Now, iso-

saturation curves show that at high saturations (>0.7), the contours are no longer flat. Instead, relative

permeability decreases as connectivity increases for high and intermediate saturations. Saturation, however,

continues to be important toward relative permeability as seen for the high receding phase contact angle

cases. With a small change in saturation, kr changes significantly for fixed connectivity. This is especially

true at high and intermediate saturations.

The unexpected trend of increasing connectivity but decreasing relative permeability is more clearly

visible from Figure 28 (also see Figure 33) where two extreme receding contact angle cases (θ1 = 180° and

θ1 = 0°) are compared. Here, the kr-S and ˆ S − paths are compared for Si ~ 1.0 and Si ~ 0.9. From Figure

28, it is seen that the kr-S path for Si ~ 1.0 is below that of Si ~ 0.9 for high saturations (~0.5 – 0.9) when

the receding phase is wetting (θ1 = 0°). At these high saturations, however, the ˆ S − paths for the same

contact angle, is higher for the Si ~ 1.0 path. This suggests that as one moves vertically (at fixed S) across

the two paths from Si ~ 1.0 to Si ~ 0.9, one will encounter that the phase connectivity decreases, but the

relative permeability increases. This is not observed for θ1 = 180°. The anomaly observed for θ1 = 0° arises

because kr decreases more sharply along the Si ~ 1.0 path because of more efficient displacement of the

Page 126: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

106

initially well-connected phase, while the decrease in connectivity is gradual. This demonstrates the

independent nature of the dependence of relative permeability on phase connectivity and relative

permeability. Finally, along the same path (Si ~ 0.9 or Si ~ 1.0), however, the trends between kr-S and ˆ S −

change monotonically as would be expected.

Figure 28. Comparison of kr-S (upper row) and – S (lower row) paths for two Si values (~ 1.0 and ~

0.9) for two receding contact angles cases (θ1 = 180° and 0°).

4.3.2. Initial-residual (IR) saturation trapping curves

We establish phase trapping when the magnitude of capillary pressure (Pc) reaches a very high value

(magnitude of 105 Pa). Other residual saturations for different kr stopping criteria are discussed in appendix

F. For cases where the receding phase (phase1) is nonwetting, Pc remains positive (showing spontaneous

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

k r

S

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

S

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

S

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

S

θ1 = 0o θ1 = 180o

Si ~ 1.0

Si ~ 0.9

Page 127: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

107

imbibition of phase2), whereas when phase1 is wetting, Pc becomes negative and forced injection of phase2

is required.

Figure 29 shows the IR saturation curves for the different wettability cases. The trends are similar to

those seen previously in Spiteri et al. (2008), in that the typical nature of IR curves holds until very high

saturations (~0.8). With increasing initial saturation, the residual saturation increases because more phase

trapping becomes possible with more initial phase. Trapping occurs because of snap-off, which is more

dominant for the high receding phase contact angle cases. For decreasing contact angles (wetting phase1

conditions), we see a decrease in the residual saturation, due to a decrease in snap-off events. For these

contact angles, flow of the receding phase occurs through piston-like displacement by the advancing phase,

as well as through layer flow. Thus, the residual saturation values are driven to low values. This occurs

more strongly at very high initial saturation, which shows that if the medium is initially completely filled

with a particular phase, its initial connectivity may help in driving the residual saturations to very low

values. This is similar to the trends observed experimentally by Herring et al. (2015). For complete initial

phase1 saturation (Si ~ 1.0) and medium completely wetting to phase1 (receding phase), the residual phase1

saturation goes to ~ zero which occurs due to layer flow of phase1. Although such flow scenario may not

be possible through typical core flooding experiment.

Figure 29. Initial-residual saturation curves for different wettability cases observed using pore-network modeling.

Page 128: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

108

4.3.3. Modeling IR saturation trapping curves

Different trapping models have been proposed in the literature. A list of some of the most commonly used

models is presented in Table 9. We implement these models on the PNM data sets and compare these

against a new proposed model.

Table 9. IR trapping models from the literature. iS , rS , and max

rS refer to the initial, residual, and

maximum residual phase saturations, respectively, while, C, a, b, c, α, and β are model parameters. Reference IR trapping model Remarks

Land (1968)

1

ir

i

SS

CS=

+ (4.3)

First trapping model developed for

mitigating relative permeability hysteresis.

Developed from observations in water-wet

media. C is known as the Land’s trapping

coefficient.

Ma and Youngren

(1994) 1

ir b

i

SS

aS=

+ (4.4)

This model is a modification to the Land’s

model when b → 1.0.

Jerauld (1997) ( )( )1 / 111 1

max max

r r

ir

bS S

imax

r

SS

SS

+ −

=

+ −

(4.5) This model was derived from the Land’s

model. The trapping coefficient is replaced

by a calculation from the maximum residual

saturation.

Spiteri et al. (2008) 2

 r i iS S S = − (4.6)

This quadratic model was presented based

on PNM observations where IR curves were

found to show a maximum.

This work 1

ir c

i

aSS

bS=

+ (4.7)

Our model collapses to the Ma and

Youngren model when a → 1.0; and to the

Land’s model when both a and c → 1.0.

Land’s model and other Land-based models are developed on water-wet media with data sets that are

limited in very high initial phase saturation conditions and thus may not appropriately capture the IR

characteristic curves completely. The quadratic model by Spiteri et al. (2008), however, was a substantial

improvement to the currently used Land-based models. However, this model appears restrictive because it

assumes a quadratic nature for all IR curves, which may not be true for water-wet systems. Thus, the Spiteri

et al’s model fails to collapse into the known Land-based models under limiting conditions. Therefore, we

propose a new IR trapping model with a slight adjustment to the Ma and Youngren model (Table 9).

Page 129: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

109

Figure 30 shows the IR saturation trapping curves for four different contact angle cases as well as the

corresponding matches shown by dashed lines for the models presented in Table 9. The goodness measure

of these models for all wettability cases is shown in the right plots in Figure 30, and the matching parameter

information for the model presented in this study (Eq. (4.7)) is provided in Figure 31.

Page 130: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

110

Figure 30. IR saturation trapping curves for four different contact angle cases (θ1 = 180°, 120°, 60°, 0°). The dashed lines show the fits for the trapping model. The corresponding goodness measure of the fits for all wettability cases and R2 values are displayed on the right plots.

Lan

dM

a an

d Y

ou

ngr

enJe

rau

ldSp

iter

i et

al.

This

wo

rk

θ1 = 180o

θ1 = 120o

θ1 = 60o

θ1 = 0o

Page 131: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

111

Figure 31. Matching parameters plotted as a function of wettability for the IR trapping model presented in this research (Eq. (4.7)). The corresponding fits are available in Figure 30. The shaded regions mark the 95% confidence interval.

From Figure 30 we see that Land and other Land-based models (Ma and Youngren, and Jerauld) match

well for high receding phase contact angles (θ1 = 180° and 120°). However, the performance by these

models suffers for lower receding phase contact angles, showing their limitation for water-wet media.

Spiteri et al’s model, however, performs better (R2 = 0.91) than Land-based models. Still, this model does

not match well for cases when receding phase is wetting (θ1 = 60° and 0°). This may likely be because the

curves do not seem typically quadratic. With the model presented in this work (Eq. (4.7) in Table 9),

substantial improvement (R2 ~ 0.99) is found in capturing the trapping trends for all contact angles. We find

that the additional parameter in the numerator, a, is a strong function of wettability and lies between 0 and

1 (Figure 31). It helps achieve control on the nonlinearity of the IR curves. Similar to water-wet conditions,

for high receding phase contact angles, the value of a is closer to one and decreases as the receding phase

contact angles decrease. The other two parameters also change with wettability, but the change is relatively

less.

Page 132: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

112

The abrupt change in the parameters near contact angle of 80° occurs likely as the flow regimes shift

due to shift in wettability. For θ1 = 70° and for Si = 1.0, there is sharp decrease in the Sr value (see the green

curve in Figure 29) which is likely causing this jump in the fitting parameters because of the nature of this

Si-Sr curve. The reason for this is that both wettability and initial saturation govern the nature of the Sr curve.

At θ1 = 70°, there is now a shift in the wettability and that wettability effect is more pronounced when Si =

1, that is when the medium is completely filled with the receding phase.

Page 133: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

113

4.3.4. Phase connectivity – phase saturation ( χ – S) paths and trapping locus

The – S paths for the receding phase (phase1) for four different contact angle cases are displayed in Figure

32. For high receding phase contact angles (θ1 = 180° and 120°), the ˆ S − paths are linear due to gradual

snap-off of the receding phase. The locus bounded by the residual phase1 saturation and phase1 connectivity

( ˆr – Sr) shown by black open circles, remains fairly constant (~ 0.395) for θ1 = 180°. Here, we find

nonlinearity in the locus because of lower Sr values for very high initial phase1 saturation cases owing to

better displacement which is more evident for θ1 = 180°.

Figure 32. – S paths for the receding phase during secondary injection process for different contact

angle cases. The red open circles represent the initial condition whereas, the black open circles represent the residual condition for each – S path.

θ1 = 180o

θ1 = 60o

θ1 = 120o

θ1 = 0o

Page 134: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

114

For lower contact angles (θ1 = 60° and 0°), we continue to find linear ˆ S − trends for higher saturations,

but for lower saturations, goes down to ~ zero (the most disconnected state) at saturation of ~0.15. This

corresponds to the total number of network elements occupied by the receding phase to be roughly equal

to the total number of pores in the network. Further reduction in saturation, causes an increase in , which

occurs due the pore structure constraints and the definition of (Eq. (4.2)). As smaller number of network

elements get occupied, χ < χmax thus, increases and ultimately reaches the limiting condition of ~ 0.395

[–χmax/(χmin– χmax)] as the Euler number reaches a very small number. Thus, the saturation of ~0.15 which

marks the most disconnected state for the receding phase could be referred to as the percolation threshold

for the wetting phase.

Figure 33 shows the relative permeability contours for the ˆ S − paths shown in Figure 32. The

monotonic behavior of kr versus both and S is visible when the phase is nonwetting to the medium (θ1 =

180° and θ1 = 120°). Relative permeability continues to be monotonic with S for cases where the receding

phase is wetting to the solid. Here, the dependence of kr on S is even stronger with sharper change in kr as

S changes slightly (see high saturation regions for θ1 = 60° and θ1 = 0°). Finally, the inverse trends between

kr and for fixed S, discussed previously in Figure 27 and Figure 28, can also be seen from the kr contours

for θ1 = 60° and θ1 = 0° where the receding phase is wetting to the solid.

Page 135: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

115

Figure 33. Relative permeability contour plots for the – S paths shown in Figure 32

In Figure 32, this is the first time that ˆ S − trends for the wetting phase are discussed using network

modeling. In general, the data for wetting phase connectivity remains limited in the literature due to the

difficulty in resolving the wetting phase and accurately calculating the Euler number through image data.

Therefore, more high-resolution experimental data is needed to corroborate these trends.

The ˆ S − path for Si = 1 for the θ1 = 120° crosses the residual locus. However, the relative

permeabilities on this ˆ S − path (near the residual locus) are close to zero (for example, kr = 0.004 at ,S

= 0.360, 0.300 (point near the residual locus for θ1 = 120° in Figure 32)). This shows that kr =f( ,S) may

lose uniqueness near the extreme regions of kr. Thus, consideration of additional parameters such as the

fluid/fluid interfacial may be needed to get a more exact solution for the kr. Nevertheless, the error with the

θ1 = 180o

θ1 = 60o

θ1 = 120o

θ1 = 0o

Page 136: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

116

consideration of kr =f( ,S) is not very significant. Figure 34 shows a compilation of the trapping loci for

the different contact angles.

Figure 34. Trapping locus of phase saturation and phase connectivity for different contact angles.

Few observations can be made from Figure 34. There is a finite value of the minimum trapped phase

connectivity, despite ˆ S − paths showing a ~ zero value of . This is because increases again due to

layer flow (see Figure 32). Next, the shape of the locus opens as contact angle decreases (180° to 80°) and

shrinks back as contact angle decreases further (70° to 0°), showing a strong influence of wettability. This

occurs because as receding phase contact angle decreases, the residual saturation decreases, leaving a more

disconnected state. But, as contact angle decreases further where the receding phase becomes the wetting

phase, the trapped saturation becomes very low, increasing again. The pore structure limiting condition

influences the locus strongly. The shrinking nature of the locus depends on the reducing residual saturation

which occurs at low contact angles because of efficient displacement of the receding phase.

Page 137: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

117

4.3.5. Modeling residual phase connectivity and trapping locus

We model residual phase connectivity ( ˆ )r as a function of the initial phase connectivity similar to how IR

saturation is modeled. We propose a simple quadratic function as,

2 2

  (ˆ ˆ ˆ ˆ ˆ ˆ( ,) )r i o i o o

A B = + +− − (4.8)

where A and B are model parameters; ˆo

is the limiting phase connectivity at negligible phase saturation

[–χmax/(χmin– χmax)]; ˆi

and ˆr

are the initial and residual phase connectivity, respectively. The combination

of modeling the residual phase saturation and residual phase connectivity allows us to model the residual

locus. Figure 35 shows the IR phase connectivity trapping curve for four different contact angle cases as

well as the corresponding match shown by dashed lines for the model for residual phase connectivity. The

goodness measure of the model for all wettability cases is shown in the right plot in Figure 35, and the

matching parameter information for the model presented is provided in Figure 36.

Figure 35. IR phase connectivity curves for four different wettability conditions (θ1 = 180°, 120°, 60°, 0°). The dashed lines show the match for the phase connectivity trapping model (Eq. (4.8)). The corresponding goodness measure of the fit for all wettability cases and R2 value is displayed on the right plot.

θ1 = 180o

θ1 = 120o

θ1 = 60o

θ1 = 0o

Page 138: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

118

Figure 36. Matching parameters plotted as a function of wettability for the IR phase connectivity trapping model. The corresponding fits are available in Figure 35. The shaded regions mark the 95% confidence interval.

Few observations can be made from Figure 35. As discussed previously, the residual phase connectivity

remains fairly constant for high receding phase contact angles. However, as contact angle decreases, the

trends become parabolic in nature. This occurs because of a couple of reasons. First, the topological

constraint of the pore structure. For example, ˆi

cannot go lower than ~0.395 which constraints the curve

on one end. This is because during the primary (phase1) injection, the receding contact angle is set at 180°.

Thus, phase1 is nonwetting to the surface and layer flow does occur to reduce ˆi

. Second, ˆr

for lower

contact angles reduces below 0.395 due to efficient displacement of the receding phase. Finally, ˆr

becomes once again constrained at very high ˆi

because here the residual saturation almost goes to zero

(see Figure 29 and Figure 30). The proposed function (Eq. (4.8)) captures the trends for residual phase

connectivity across all wettability cases (R2 ~ 0.88).

On the basis of Eq. (4.7) and (4.8), the residual phase saturation and residual phase connectivity trends

can be captured. A summary of the initial-residual phase saturation and phase connectivity trends for

different four different contact angles are presented in Figure 37.

Page 139: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

119

Figure 37. Summary of initial-residual phase saturation and phase connectivity trends for four different contact angles (θ1 = 180°, 120°, 60°, 0°). (a) Initial connectivity versus initial saturation. Here all four contact angles collapse to one single data set; (b) initial connectivity versus residual connectivity; (c) initial saturation versus residual saturation; (d) residual connectivity versus residual saturation.

In Figure 37a, initial phase connectivity is shown as a function of initial phase saturation. All four

contact angles collapse to a single data set (shown by blue open circles). This is because the initial condition

prior to secondary phase2 injection, represent the primary phase1 injection at which point all flow conditions

including contact angle of 180° degree are the same. The dashed line in Figure 37a is a match to the

following power law function,

( ) ,ˆ 'ˆ 'k kp S S − = − (4.9)

θ1 = 180o

θ1 = 120o

θ1 = 60o

θ1 = 0o

a) b)

c) d)

Page 140: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

120

where ˆ ' and 'S are points on the ˆ S − path; and p and k are matching parameters. For the match in

Figure 37a, we constrain the model on the ˆ S − path at the first and the last points, namely, (0.395,0) and

(1,1). Figure 37b and Figure 37c were discussed in Figure 30 and Figure 35, respectively. Here, the axes

on Figure 37b are flipped for seamless representation of initial-residual curves. Combination of Figure 37b

and Figure 37c helps develop the residual locus (Figure 37d).

Page 141: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

121

4.8. Concluding remarks

Pore-network simulations show that very high initial phase saturations can lead to reduced trapping and

that an optimum trapping saturation exist for initial saturations in the range of 0.7-0.95 which depends on

the wettability. This is an important finding which suggests that the primary CO2 injection should be

designed such that this range of initial CO2 saturation is targeted to maximize trapping. The influence of

initial phase saturation becomes more significant for cases where the receding phase is strongly wetting.

Thus, the combination of the higher initially connected CO2 phase if is also wetting to the surface, can

reduce the trapping capacity significantly. The most suitable trapping scenario (trapped saturation ~50%)

is found for intermediate wettability conditions and an initial saturation of between 0.6-0.8.

Initial-residual (IR) saturation trends may be inaccurately captured by the application of models that

are built for water-wet media. If the receding phase is wetting to the surface, flow caused by layer flow and

piston-like advance of the invading phase are important physical displacements regimes that can

significantly affect trapping. This has also been seen in experimental studies. For example, Iglauer et al.

(2011a) discuss the presence of sandwiched oil layers which drive low residual oil saturations. They found

Sor in oil-wet medium to be 0.18 as opposed to 0.35 for water-wet medium, with fewer larger-sized oil

clusters. Larger-sized oil clusters were found for the water-wet case by Landry et al. (2011). The general

trapping model presented in this chapter provides a simple and elegant solution to accurately represent

trapping trends across all wettability conditions of the system. In addition, the model easily collapses to the

traditionally known models.

A few researchers have observed a more complex increasing-decreasing-increasing IR trend at mixed-

wetting conditions (Salathiel 1973; Tanino and Blunt 2013). The quadratic concave-up profile as described

in Tanino and Blunt (2013) was not observed through PNM simulations. The difference likely occurs

because as initial saturation increases, more oil contacts the rock surface and low residual oil saturations

are observed due to layer flow. However, at higher initial oil saturations, layer flow continues to exist in

PNM, but experiments show a slight decrease due to oil being forced into the corners of the larger pores or

Page 142: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

122

into micropores that can’t be displaced as effectively and Sor increases further in those experiments (Tanino

and Blunt 2013). However, these trends have not been observed for a CO2/brine/rock system that most flow

experiments are unable to reach such large initial oil saturation values (>0.9). Thus, more experiments

should be designed to corroborate this trend and future network models should be built to incorporate these

trends.

Recent literature has suggested the importance of phase connectivity and by including this term helps

resolve hysteresis in relative permeability (Khorsandi et al. 2017; Purswani et al. 2020; Purswani et al.

2019; Zhao et al. 2019) and capillary pressure (Li and Johns 2018). However, the focus has remained on

the nonwetting phase. In this chapter, we presented phase connectivity trends for receding phase under

wetting and nonwetting conditions. The residual locus provides the true limiting condition of residual phase

connectivity and residual phase saturation for the given pore structure and wettability. This can provide a

known limiting condition to calibrate multiphase models used in reservoir simulations which will be critical

in scaling carbon capture projects. Design of controlled experiments is required to capture wetting phase

connectivity trends as well as the trends of the locus of residual phase connectivity and saturation. In

addition, the limiting conditions of the pore structure topology should be experimentally confirmed.

One limitation of our work is that we have only considered capillary dominated flow which is reflective

of flow away from the wellbore, but it is well established that flow regimes like ganglion dynamics and

drop traffic flow can significantly affect phase connectivity and capillary trapping which occur at high flow

rate conditions (or high capillary numbers), therefore wettability effects in combination with high capillary

numbers will be important when investigating phase trapping (Armstrong et al. 2016; Schlüter et al. 2016).

Overall, the impact of wettability using capillary dominated pore-network simulations is studied.

Through controlled numerical experiments, trends of residual saturation and phase connectivity are

analyzed for the full range of contact angles and initial saturations. As contact angles change, the flow

regimes change, this causes the phases to distribute differently and affect trapping. The trapping models

discussed here are general and can be used across different wetting conditions. Future experiments for

limiting flow conditions of high initial saturations as well as different wettability are recommended.

Page 143: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

123

CHAPTER 5. DEVELOPMENT OF EQUATION-OF-

STATE USING PORE-NETWORK MODELING

Preface

The contents of this chapter were originally presented at the SPE IOR Conference, 31st Aug-4th Sept.,

Virtual, 2020. The manuscript was accepted for publication in the SPE Journal and is referenced as,

Purswani P., Johns R.T., Karpyn Z.T., Blunt M.J. (2020)

Predictive Modeling of Relative Permeability using a Generalized Equation-of-State, SPEJ, (26), 191-

205, https://doi.org/10.2118/200410-PA

Author contributions: Johns R.T., Karpyn Z.T., and Purswani P. conceptualized the approach. Purswani P.

processed the numerical data set used in the model, developed the modeling efforts, and wrote the original

draft in consultation with Johns R.T., Karpyn Z.T., and Blunt M.J. All coauthors contributed toward

analyzing the data and updating the manuscript.

Abstract

Reliable simulation of enhanced oil recovery processes depends on an accurate description of fluid transport

in the subsurface. Current empirical transport models of rock-fluid interactions are fit to limited

experimental data for specific rock types, fluids, and endpoint values. In this chapter, a general equation-

of-state (EOS) approach is developed for relative permeability (kr) based on a set of geometric state

Page 144: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

124

parameters: normalized Euler characteristic (connectivity) and saturation. Literature data and pore-network

modeling (PNM) simulations are used to examine the functional form of the EOS.

Our results show that the new kr-EOS matches experimental data better than the conventional Corey

form, especially for highly nonlinear relative permeabilities at low saturations. Using hundreds of PNM

simulations, relative permeability scanning curves show a locus of residual saturation and connectivity

which defines an important limit for the physical kr region. The change of this locus is also considered for

two contact angles. PNM data further allows for the estimation of the relative permeability partial

derivatives which are used as inputs in the EOS. Linear functions of these partials in the connectivity-

saturation space renders a quadratic response of kr, which shows excellent predictions. Unlike current

empirical models that are based on only one residual saturation, the state function approach allows for

dynamic residual conditions critical for capturing hysteresis in relative permeability.

Page 145: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

125

5.1. Introduction

The flow of two or more phases in a porous medium is quantified through averaged transport properties

such as relative permeability. Relative permeability is critical for accurate multiphase flow simulations of

oil and gas reservoirs, which are routinely used to estimate petroleum reserves, evaluate the success of

enhanced oil recovery (EOR) projects, and estimate the amount of CO2 that can be trapped for sequestration

purposes. Accurate description and practical modeling solutions for relative permeabilities continue to be

a long-standing concern faced by petroleum engineers and hydrologists.

Traditionally, the relative permeability of a phase is expressed as empirical functions of phase

saturations using simple functional forms (Brooks and Corey 1964; Corey 1954). However, a growing body

of literature demonstrates that there are a variety of other parameters that affect relative permeabilities

(Blunt 2017). These include the wettability of the medium, due to which the residual saturation is lower for

a mixed-wet medium than for a strongly water-wet medium (Jadhunandan and Morrow 1995). Flow

conditions such as the flow rate or the interfacial tension between the injection and displacing fluid

(captured in the capillary number) also strongly affect relative permeabilities (Armstrong et al. 2016;

Avraam and Payatakes 1995; Purswani et al. 2019). In addition, topological descriptions of the porous

medium, such as the pore/throat coordination number and aspect ratio also have an effect (Jerauld and Salter

1990).

Various flow conditions and porous medium attributes manifest themselves in different pore-scale

events that govern the movement of fluid through the pore space, thus affecting relative permeability. Pore-

scale flow experiments performed using x-ray micro-computed tomography (micro-CT) assist in the

visualization and quantification of fluid transport at spatial resolutions close to a micron. In addition to

tracking phase volumes, fluid/fluid interfacial area (Culligan et al. 2005; Culligan et al. 2004), distribution

of contact angles (AlRatrout et al. 2017; Klise et al. 2016; Scanziani et al. 2017), the topology of the

nonwetting phase (Herring et al. 2015; Herring et al. 2013; Schlüter et al. 2016) can also be estimated

accurately through such experiments. Fluid movement can in some cases be captured in real-time at

Page 146: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

126

temporal resolutions close to tens of seconds with the application of dynamic x-ray imaging facilities

(Rücker et al. 2015; Singh et al. 2017). Further, through the coupling of x-ray imaging experiments with

fast direct numerical simulations, multiphase transport can be captured at different flow conditions, such as

those at very high capillary numbers (Armstrong et al. 2016; Mcclure et al. 2018; Mcclure et al. 2016).

Pore-network modeling (PNM) is advantageous because repeat numerical experiments can be made for

different porous media topology and at different wettability while holding some parameters constant (Blunt

2017; Valvatne and Blunt 2004).

Pore-scale studies are valuable to gain a deeper appreciation of the complexity of multiphase transport

through improved visualization of displacement processes. These include piston-like displacement,

cooperative pore-filling, snap-off (choke off), Haines jumps, coalescence, drop-traffic flow, and ganglion

dynamics. These mechanisms observed in controlled micromodel experiments (Avraam and Payatakes

1999; Avraam and Payatakes 1995; Lenormand and Zarcone 1984) have been corroborated through three-

dimensional experiments and simulations (Armstrong et al. 2016; Rücker et al. 2015). These pore-scale

mechanisms often lead to phase trapping when the flow direction is reversed. In addition, surface roughness

effects (Morrow 1975) and wettability alteration (Salathiel 1973) lead to hysteresis in contact angles such

that the advancing contact angle is greater than the receding contact angle (Anderson 1986b; Blunt 2017;

Braun and Holland 1995). Phase trapping and contact angle hysteresis cause path dependence of relative

permeabilities in the saturation space, which is termed hysteresis.

Hysteresis in relative permeability continues to be a problem for accurate reservoir modeling. Large

errors manifest in the estimation of recoveries if hysteretic effects are ignored or improperly accounted for

during reservoir simulations (Carlson 1981). Estimation of relative permeability hysteresis and trapping

efficiency of the nonwetting phase is also critical for CO2 sequestration (Akbarabadi and Piri 2015; Juanes

et al. 2006). In addition, hysteretic effects become pronounced for EOR processes that involve a reversal

of flow, such as the water alternating gas (WAG) process, where gas is trapped during the water injection

cycle (Khorsandi et al. 2018; Spiteri and Juanes 2006).

Page 147: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

127

Current modeling techniques for relative permeability (Brooks and Corey 1964) are deficient as these

consider relative permeability to be empirical function of saturation and different physical effects such as

wettability and capillary number are all captured in a single parameter such as the Corey exponent. This

lack of physics in our models likely explains, in part, why relative permeability is often altered to history

match production in oil reservoirs.

Most common hysteresis models take a similar form to that of Corey. Land pioneered the estimation of

imbibition relative permeabilities by accounting for the trapped nonwetting phase. A simplified relationship

was found between the initial (or maximum) and residual nonwetting saturation during imbibition (Land

1968). A pore-size distribution parameter was required for the estimation of imbibition nonwetting phase

relative permeabilities. This was later corroborated through experiments (Land 1971).

Hysteresis models used in conventional simulators are generally extensions of Land’s model.

Killough’s model interpolates between the initial and residual nonwetting saturations for the estimation of

imbibition relative permeabilities. An estimate of the Land’s trapping coefficient is necessary to determine

the residual nonwetting saturation (Killough 1976). Carlson recognized that unlike Land’s and Killough’s

model, the need for a secondary parameter could be avoided given that one experimental data point on the

imbibition relative permeability curve is known. He suggested that the residual nonwetting phase can be

determined using Land’s formulations and the known experimental data point. More known data points on

the relative permeability curve would lead to a better average of the residual nonwetting saturation (Carlson

1981).

One limitation of these models is that they are strictly empirical and are often based on the observations

found for water-wet systems, which means that these models work under the assumption of no trapping

during secondary drainage cycles. This leads to limited prediction capability by these models, especially

for mixed-wet systems (Fatemi et al. 2012; Spiteri et al. 2008). Furthermore, for the implementation of

these models, information of the initial-residual nonwetting saturations together with the prior drainage

relative permeability is required. Spiteri et al. (2008) observed nonlinear trends in the trapping curves for

the oil phase for different wettabilities of the medium. They proposed a quadratic hysteresis model between

Page 148: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

128

the trapped and initial oil saturation and demonstrated the model’s capability in capturing hysteresis for

different wetting conditions.

Yuan and Pope (2012) proposed a Gibbs energy-based model for relative permeability with applications

to two-phase compositional processes. Relative permeabilities were made proxies of compositions, which

were determined by interpolating from a reference state of known Gibbs energy and relative permeability.

Parameters in a simplistic Corey-type model were interpolated to arrive at relative permeabilities for each

change in composition. Trapping and hysteresis in relative permeabilities were later accounted for in a

similar fashion by Neshat and Pope (2018), where parameters in an extended Land model were interpolated

using the Gibbs energy approach. These models eliminated some of the discontinuities observed in

compositional simulation owing to phase identification and labeling problems. Still, it is not clear that

Gibbs energies of each phase are always distinct and that this approach is predictive. Further, these models

depend on lots of tuning parameters and lack physical insight.

Skauge et al. (2019) presented two models to bridge immiscible and miscible WAG EOR processes. In

the first model, empirical three phase relative permeability functions were considered, and phase trapping

was incorporated using the Land formulation. These kr functions were described as functions of all three

phase saturations, the path of the saturations, and the trapped saturation of all phases. In addition, for the

consideration of the compositional effects, another empirical formulation was considered to incorporate

interfacial tension (IFT) changes. This was established using the Coats (1980) functional forms which

include the IFT effect by scaling the endpoint saturations from the original IFT (σo) as follows,

( )*gr grS f S= (5.1)

( )*org orgS f S= (5.2)

Page 149: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

129

where ( )1

1n

of

=

; and Sgr and Sorg are the residual gas and oil saturations, respectively. Equation

(5.1) and (5.2) allow for *grS and *

orgS to be zero as interfacial tension approach zero near supercritical

conditions. The value of n1 usually remains between 4 and 10.

For the second model described by Skauge et al. (2019), wettability, pore-size distribution, and pore-

scale physics was incorporated into the compositional modeling approach. Pore-scale physics was

incorporated using pore occupancy of the phases. For this, the wetting order was based on pore sizes, for

example, the smallest pores would be occupied by the wetting phase, the largest pores would be occupied

by the nonwetting phase, and the intermediate-sized pores would occupy the phase of intermediate

wettability. They extended this idea further to consider the distribution of contact angles in the pore space

and allow for pores of different radii to have different contact angles. This was accomplished through the

Bartell-Osterhof relationship, which is given as,

    .gw gw ow ow go gocos cos cos = + (5.3)

This relationship is established by considering all three Young’s relationships for the three two-phase sets

(oil/water; gas/water; and oil/gas) of interfacial tension equations. Elimination of the fluid/solid IFTs gives

Eq. (5.3).

Equation (5.3) was used to identify the pore occupancy for each phase as well as to model the transition

toward miscibility. For example, for a weakly oil-wet porous medium, water will be intermediate wetting

for an immiscible flood, but will be nonwetting (and gas will be intermediate wetting) for a near-miscible

flood due to the changes in the interfacial tension at near-miscible conditions. As a consequence, in this

example, the path followed in the three-phase displacement process will be different. For the miscible flood,

the path will be gas to oil to water, whereas for the near-miscible flood, the path will be from gas to both

oil and water.

Page 150: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

130

The drawback of these compositional models is that they still require to track of phase labels

(oil/water/gas), which can create problems with compositional simulations because at supercritical

conditions discontinuities can occur that can cause simulations to fail or underpredict recoveries.

One proposed solution to relative permeability hysteresis is to develop a multi-parameter, physics-

based description for relative permeability, similar to capillary pressure, where hysteresis is modeled with

the inclusion of fluid/fluid interfacial area (Hassanizadeh and Gray 1993; Reeves and Celia 1996; Joekar-

Niasar 2008; 2010). However, errors are significant near high and low saturations using that approach,

because of sudden changes in the fluid/fluid interfacial areas at extreme saturations. Schlüter et al. (2016)

recognized that connectivity as quantified by the Euler characteristic is a key parameter for describing

multiphase flow. This was also found in the study by Mcclure et al. (2018), where the authors suggested

that four state variables, namely, fluid saturation (S), specific fluid/fluid interfacial area (αw-nw), fluid/fluid

average interfacial curvature (κ), and normalized fluid phase Euler characteristic ( ˆ ) describe the state of

a fluid phase in a porous medium. Based on this, κ was described as a function of the other three state

parameters, κ = f (S, αw-nw, ). Other constitutive correlations for κ were also considered, κ = f (S) and κ = f

(S, αw-nw) for comparison. The measures of these constitutive correlations were fit against two-fluid flow

simulation data in different porous media. The authors found that κ = f (S, αw-nw, ) gave the most accurate

description of the capillary pressure as opposed to the other constitutive relations considered since this was

the most complete description of the fluid phase. The most dramatic improvement was found upon the

addition of the phase connectivity term (Euler characteristic).

Johns and coworkers suggested a new conceptual framework for modeling relative permeabilities

(Khorsandi et al. 2017, 2018; Li and Johns 2018; Purswani et al. 2019). Khorsandi et al. (2017) proposed

relative permeabilities as a state function based on capturing the essential state parameters. This framework

makes relative permeability single-valued for a combination of state parameters that are fully described at

that state. This methodology can eliminate phase labeling (wetting/nonwetting, oil, water, gas) issues as

demonstrated by Khorsandi et al. (2018) where such labels cause discontinuities for supercritical fluids.

Page 151: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

131

The EOS model was also extended to three phases and for WAG injection. Li and Johns (2018)

implemented a similar framework for proposing a coupled EOS model for capillary pressure and relative

permeability. Partial changes in capillary pressure due to saturation, connectivity, and wettability were

estimated by fitting a quadratic response of capillary pressure to literature data. Purswani et al. (2019)

similarly showed that the relative permeability EOS could be approximated as a simple response surface in

the saturation and connectivity space. Appropriate boundary conditions were set and the EOS was tested

for different capillary number conditions. Zhao et al. (2019) implemented this response surface to develop

a machine learning framework for modeling relative permeabilities.

Despite recent advances, the exact form of the EOS remains unknown and relies on significant

assumptions regarding the approximation of the partial derivatives. Therefore, in this research, we advance

the development of the state function approach for modeling relative permeability by evaluating the partial

change in relative permeability due to the changes in the state parameters considered. First, we test a

physical relationship for the state parameters involved, namely, connectivity (Euler characteristic) as a

function of saturation. This relationship is then incorporated into the exact differential form of the proposed

EOS to match measured data. We also present comparisons to the simplified Corey form of relative

permeability. Micro-computed tomography and synthetic data from PNM are used to examine these

relationships. Next, residual saturation and connectivity trends are analyzed, and values of the partial

derivatives for relative permeability are estimated using PNM. Finally, predictions of relative permeability

are made using the estimated partial derivatives.

Page 152: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

132

5.2. Methodology

In this section, we present a general development of the EOS for relative permeability. A similar

methodology could be applied to capillary pressure and other fluid-rock interaction functions. The exact

differential for relative permeability of a phase j is given as,

1

,j

j

N

i

i

r

i

r

N i

d dyy

kk

=

=

(5.4)

where yi is any state parameter affecting relative permeability, N is the number of state parameters. For NP

phases, there will be NP equations like Eq. (5.4). Although we will refer to “oil” and “water” in this chapter,

such identification is not necessary. Only the values of the state parameters, such as saturation, must be

known to estimate relative permeability.

The task described by Eq. (5.4) is then to define N state parameters that best, or nearly fully, describe

the state of the system. The degrees of freedom are large in porous media and, therefore, the goal is to find

the least number of state parameters that can capture the relative permeability function accurately. Small

errors in relative permeability estimates are acceptable if the most essential aspects of the physics are

captured with only a few input parameters. Thus, such an approach lends itself as an elegant and practical

solution for modeling relative permeabilities. Once these state parameters are known, Eq. (5.4) ensures that

relative permeability is a continuous and unique function of these state parameters.

There are two types of state parameters that can govern relative permeability. First, internal parameters

that describe the geometric state of fluid inside the porous medium, which include saturation, connectivity,

and fluid/fluid interfacial area (Mcclure et al. 2018). Second, external parameters corresponding to the flow,

which include the topology and wettability of the medium and capillary number.

Page 153: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

133

The exact differential of relative permeability of a phase j (krj) for these six parameters is then given

from Eq. (5.4) by,

, , , ,,  , , , ,  , , ,

, , ,

ˆ

, , , , , , , , ,

ˆ

ˆ ˆ ˆ

ˆˆ

j j j

j

j j Caj Ca j Ca

j j j

j j j j Ca j j Ca

r r r

r j j

j j S I lnNA I lnN S A I lnN

r r r

Ca S I A S lnN A S lnN I

Ca

k k kdk dS d dA

S A

k k kdln dI

lnN IN

= + +

+ + +

 

,

A

d

(5.5)

where Sj is the saturation of phase j; ˆj

is the normalized Euler connectivity of phase j defined below; A is

the fluid/fluid interfacial area; I is the wettability of the porous medium (e.g. I = cosθ); lnNCa is the natural

logarithm of capillary number; λ is the topology of the porous medium which can be described in its most

simple form as the ratio of absolute permeability to porosity (λ = kabs/ ), although a more sophisticated

treatment of the pore topology may include, pore body/pore throat aspect ratio, coordination number, as

well as the Euler connectivity of the pore space. Equation (5.5) could also have been written in terms of the

natural logarithm of relative permeability depending on the range of relative permeability modeled. In this

chapter, we drop the subscript j, remembering that saturation, connectivity, and relative permeability can

be written for any phase, oil/water, wetting/nonwetting. Distinctions, if necessary, will be provided to help

the reader, but whether a phase is wetting or not and to what degree, is one of the input state parameters.

The normalization procedure for connectivity follows from our previous work (Purswani et al. 2019).

This is expressed as,

,ˆ   max

min max

−=

− (5.6)

where χ is the Euler characteristic of the phase; χmin and χmax are the minimum and maximum values of the

Euler characteristic for a phase in a given porous medium. χmin corresponds to the completely connected

Page 154: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

134

state of a phase, which is possible at full saturation of that phase, whereas χmax corresponds to the completely

disconnected state of a phase and therefore corresponds to the state where all pores contain the phase, but

they are not connected via throats. Consequently, χmin and χmax can be expressed by,

min p tn n = − and, ,max pn = (5.7)

where np is the number of pores and nt is the number of throats for a porous medium. A schematic

representation is presented in Figure 38. If the size of some throats were to approach that of connecting

pores, the combination will result a single pore. As long as the same number of pores and throats are

removed from the calculation, the Euler number will not change. As an example, in the second illustration

in Figure 38, if the upper two pores and the in between throat were to be combined as a single pore, the

value of χ will still be 5 (np = 9 and nt = 4).

The value of χ can also be determined from micro-CT images when pores and throats are not readily

identifiable. This can be achieved for any saturation by using the alternating sum of the numbers of vertices,

edges, faces, and objects by approximating the voxelated representation of a fluid object(s) as a regular

polyhedral or, as the alternating sum of the Betti numbers (Herring et al. 2013; Sun et al. 2019). However,

for the accurate determination of χmax an extracted pore-network of the dry image is necessary, alternatively,

information of the average coordination number (Z) of a pore structure can be used for estimating np (χmax).

Figure 38. Ball and stick representations of a porous medium with similar number of pores but increasing number of throats. The Euler characteristic decreases from left to right as connections (throats) are increased.

min

ˆ 

 

j max

j

max

−=

min

ˆ 

 

j max

j

max

−=

min

ˆ 

 

j max

j

max

−=

• Pores• Throats

min

ˆ 

 

j max

j

max

−=

Page 155: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

135

Now, consider a set of constant external conditions, such that for a given flow experiment, wettability,

pore structure, and capillary number do not change. This reduces Eq. (5.5) to:

ˆˆ ,r S Adk dS d dA = + + (5.8)

where S

,

, and A

are the partial derivatives of relative permeability with respect to saturation,

connectivity, and fluid/fluid interfacial area, respectively. Equation (5.8) can be rewritten as,

ˆ

.ˆ A

r S

S S

dk dS d dA

= + +

(5.9)

We define a phase distribution term, Φ (Khorsandi et al. 2017), such that,

ˆ

,Φ   ˆ A

S S

d dS d dA

= + + (5.10)

This simplifies the representation of the kr-EOS:

Φ.r Sdk d= (5.11)

This representation of the kr-EOS is applicable to any number of state parameters with no loss of

generality. Equation (5.11) can always be numerically integrated or in special cases integrated exactly.

Upon integrating from a reference state to the final state, we arrive at the most general form of the EOS,

Page 156: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

136

Φ

Φ

Φ.ref

ref

r r Sk k d− = (5.12)

where in this example ( ) ( )ˆˆ/ /

S A SAS

+ + = . Normally, we take 0

refr

k = when ref r

= (one set of

residual state parameters where relative permeability is zero).

For the case when the partial derivatives are constants, Eq. (5.12) can be re-written as,

( ) ,r S rk = − (5.13)

which looks remarkably close to the Corey form of relative permeability, except that there is no exponent.

The constant coefficients for the flow function ( ) can be determined by matching available multiphase

experimental or simulation data. In theory, any path can be taken to determine the change in relative

permeability from an initial set of state parameters to the final state. However, to determine the flow

function for an actual path requires knowledge of how one state parameter changes with another along a

specific path (i.e., ˆ ( ); ( )Af S f S = = ). These dependencies give rise to nonlinear relative permeabilities as

a function of saturation even though no exponent is present in Eq. (5.13).

Equation (5.13) is the same as shown in Khorsandi et al. (2017) except that they arbitrarily included an

exponent. In addition, Eq. (5.12) generalizes the definition of to allow for non-constant partial

derivatives as well, which can further increase nonlinearities in relative permeability with saturation.

Understanding the functional form of these derivatives is one objective of this chapter.

5.2.1. Example implementation of the EOS

We assume here that changes in connectivity and saturation are sufficient to describe the state of relative

permeability (i.e., changes in other potential state parameters are either unimportant in comparison or are

Page 157: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

137

negligible). This assumption is reasonable to first-order based on the findings by Schlüter et al. (2016), and

in any case, would be a significant improvement over current empirical model. Further, we allow for

nonlinear functionality between connectivity and saturation through a power-law expression with two

constants (p and k),

( ) ,ˆ 'ˆ 'k kp S S − = − (5.14)

where ( )ˆ ', S is a specific point that constrains the ˆ S − path. A similar expression was originally proposed

by Schlüter et al. (2016), but here we consider only two parameters for expressing the evolution of

connectivity for both drainage and imbibition. The value of ( )ˆ ', S = (1,1) for the case of drainage by a

nonwetting phase. That is, the nonwetting path must go through (1,1) for very high capillary pressure, i.e.,

when all of the wetting phase is removed. An imbibition path, however, will begin at the endpoint of the

last drainage cycle ( )ˆ ,i i

S .

Equation (5.13) can be substituted into Eq. (5.14) for the case of constant partial derivatives to yield,

( ) ( )ˆ ,k k

r S r rk S S S S = − + −

(5.15)

where ( )ˆ ˆ / Sp = . Here, it is noted that the area term (A) is not included in the development of Eq.

(5.15). Further, an additional boundary condition,   o

r rk k= , at initial saturation ( )i

S , can be used to eliminate

S as,

( ) ( )'

ˆ

.

o

r

S k k

i r i r

k

S S S S

=− + −

(5.16)

Page 158: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

138

This boundary condition is useful for appropriate comparisons to the Corey form, although one could also

use  1.0r

k = when 1S = and ˆ 1 = . As an alternative to Eq. (5.16), one could simplify Eq. (5.13) as,

.

i

o r

r r

rS

k k −

= −

(5.17)

5.2.2. Two-phase simulations using pore-network modeling

Specific connectivity-saturation ( ˆ )S − paths and other state variables are often not controllable in

experiments. For example, the pore structure of a rock can be different from one core test to another, and

parameters such as contact angle can change during hysteresis. Thus, the partial derivatives, which require

holding some parameters constant, cannot be determined consistently through experiments. Compounding

this problem is that controlled pore-scale experiments are intricate to perform. Therefore, PNM is used here

to estimate the derivatives by making numerous numerical experiments and holding state variables, for

example, contact angle, fixed.

Pore-network simulations are widely used in the prediction of reservoir transport properties based on

an idealized representation of the pore space (pores are the wider spaces connected to the narrower throats)

and a set of prescribed rules which guide the movement of multiple phases in the pore space. A capillary

dominated pore-network model ( )410

CaN

− , such as the one used in this study, works as a percolation or

an invasion percolation mechanism based on the wettability and direction of flow (Blunt 2017). These

mechanisms guide each pore filling event by the invading phase based on the occupancy of the neighboring

throats, which governs the fluid/fluid interfacial curvature and consequently the capillary pressure. When

an entry pressure to a pore is met, the invasion takes place. Saturation changes are tracked for each sequence

of filling events and relative permeability for each phase is calculated as the ratio of the flow rate of that

phase to the total flow rate in a single-phase simulation (Valvatne and Blunt 2004).

Page 159: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

139

We used the pore-network model developed by Blunt and coworkers (Raeini et al 2018; Bultreys et al.

2018). This model is an extension of the network models generated previously (Øren et al. 1998; Øren and

Bakke 2003; Patzek 2001; Valvatne and Blunt 2004). The pore structure model was derived from a micro-

CT image of a dry Bentheimer sandstone (Lin et al. 2019). The properties of the extracted pore-network for

this Bentheimer sandstone are listed in Table 10. The network extraction was carried out using the maximal

ball approach developed by Dong and Blunt (2009) and further improved in the generalized network

extraction approach developed by Raeini et al. (2017).

Table 10. Properties of the pore-network extracted from the dry micro-CT image of a Bentheimer sandstone (Lin et al. 2019). The parameter Z (2nt/np) is the coordination number; np and nt are the number of pores and throats, respectively. The χmin for the image data is the value V-E+F-O exacted from the pore space of the image. The χmax for the image data is back calculated from the corresponding χmin and Z values. See additional details for consistent estimation of χmax in appendix B.

PNM Experimental

χmax 16850 16739

χmin -25787 -25618

np 16850 _

nt 42637 _

Z 5.061 5.061

0.22 0.24

kabs 2.493 D 2.198 D

We ran multiple PNM simulations at fixed water-wetting conditions: for one case, the contact angle

was set to exactly zero, and for the other, contact angles between 40o to 60o were uniformly distributed in

the pore-network using a Weibull distribution. For this distribution, the larger angles were assigned to larger

pores. Data for primary drainage and secondary imbibition were generated such that the nonwetting phase

injection during primary drainage was stopped at intermediate saturations, which were then used to start

the secondary imbibition process. Each imbibition process was then run until a high negative capillary

pressure (-105 Pa) was achieved at a final trapped saturation value. Lastly, all imbibition paths were

reversed as a secondary drainage process to achieve maximum oil saturation. A Matlab® sub-routine was

Page 160: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

140

developed to automate this process which allowed for very fine saturation steps of 0.005 in between each

initial nonwetting saturation. Typically, over 40,000 data points were analyzed for each wettability case.

Normalized connectivity was calculated from the pore element occupancy of the nonwetting phase. For

this, a sub-routine was developed in Matlab®. Two-phase flow visualization files were saved for each

saturation point and information of each node (pore/throat/half-throat) occupied by the wetting/nonwetting

phase was extracted. The Euler characteristic of the nonwetting phase was then calculated using,

,occ occ occ

nw p t htn n n = + − (5.18)

where occ

pn is the number of pore nodes occupied by the nonwetting phase;

occt

n is the number of throat nodes

occupied by the nonwetting phase; and occ

htn is the number of half-throats occupied by the nonwetting phase.

A half-throat is a sub-division of a pore element incorporated in recent pore-network extraction codes to

avoid oversimplification of the pore space (Raeini et al. 2017). This helps in describing the pore-network

at a finer level.

Page 161: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

141

5.3. Results and discussion

This section examines the evolution of connectivity as a function of saturation and hysteresis for both water-

wet and mixed-wet media. We examine literature data and synthetically generated data using pore-network

modeling (PNM). Specifically, functions are developed to describe the observed PNM paths in

connectivity-saturation ( ˆ S − ) space for hundreds of drainage and imbibition cycles and for two different

wettability values. We show the resulting residual oil saturation and connectivity locus in the ˆ S − space.

Lastly, we present estimates of the relative permeability partial derivatives using PNM data and give simple

models to predict kr.

5.3.1. Evolution of connectivity with saturation

The ability of a phase to connect or disconnect during transport in a porous medium directly impacts its

ability to flow, and consequently relative permeability. Although it is expected for the connectivity of a

phase to be proportional to its saturation, this may not always be the case. Possible exceptions include thin

film or layer flow in which wettability impacts connectivity.

We first consider example waterfloods for a variety of mixed-wet and water-wet porous media. The

selected literature data consists of similar pore structures, namely, Gildehauser sandstone (Berg et al. 2016),

Bentheimer sandstone (Gao et al. 2017; Lin et al. 2019), and a sintered glass bead pack (Armstrong et al.

2016). The segmented micro-CT images for these flow experiments (Gildehauser, and both Bentheimer)

were processed using commercial image analysis software to remove phase clusters below 125 voxels to

avoid image noise post segmentation (Herring et al. 2015). These segmented images were then used for the

estimation of the Euler characteristic (connectivity) of the oil phase and dry porous media. Next, pore-

Page 162: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

142

networks for these dry images were extracted for the estimation of the coordination number. All porous

media properties are listed in Table 11.

Table 11. Summary of information for the literature data shown in Figure 39.

Reference Rock (Wettability) Method maxχ

minχ Z

�� at

S = 0 Remarks

Berg et al. (2016) Gildehauser sandstone

(Water-wet)

Unsteady-

state injection

/Direct

numerical

simulations

9348 -10949 4.34 0.461

Berg et al.

(2016) used

direct numerical

simulations for

the estimation

of kr

Gao et al. (2017) Bentheimer sandstone

(Water-wet, WW)

Steady-state

co-injection 11369 -11321 3.99 0.501 -

Lin et al. (2019) Bentheimer sandstone

(Mixed-wet, MW)

Steady-state

co-injection 16739 -25618 5.06 0.395 -

Armstrong et al.

(2016)

Sintered glass beads

(Water-wet)

Direct

numerical

simulations

5788 -10704 5.70 0.351

max and

min are

taken from

Purswani et al.

(2019)

Figure 39. Waterflooding χ - S paths from the literature (Table 11). All cases are for NCa < 10-4. Solid

curves show the best fits to Eq. (5.14). Saturation in the experiments move from right to left as shown by the direction of the arrow.

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

GildehauserBentheimer_WWBentheimer_MWSintered glass bead pack

Page 163: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

143

Figure 39 shows the four waterflood paths in the ˆ S − space. The path is nearly linear for the water-wet

Gildehauser sandstone, which suggests the oil phase becomes increasingly disconnected at a fairly constant

rate with increasing water saturation. The disconnections in the oil phase result from snap-off or bypassing

by the wetting phase (water in this case). The trends for the other three floods, however, flatten at lower

saturations owing to a reduction in snap-off events. It is therefore likely that the Gildehauser flood results

would show flattening if data were available at smaller saturations.

Figure 39 also shows greater nonlinearity in the ˆ S − path for the mixed-wet Bentheimer sandstone of

Lin et al. (2019). The average contact angle for this flood is about ~ 80o, which likely resulted in a more

efficient oil displacement (piston-like) since neither phase prefers the rock surface. The number of

disconnections, therefore, change rapidly at higher oil saturations.

We fit each ˆ S − path in Figure 39 to Eq. (5.14), where the values of the fitting parameters (p and k)

are presented in Table 12. Smaller values of k are found for the water-wet cases (Gildehauser and sintered

glass bead), whereas the k value is larger for the mixed-wet case. The water-wet Bentheimer, however, also

gives a larger value of k. This latter case is less definitive as the saturation range is narrow and the data

shows an abrupt change at high oil saturation.

Table 12. Values of parameters fit to Eq. (5.14) using Table 11 data. The curves from the best fits are shown in Figure 39. The 95% confidence limits were calculated using the nlparci function in Matlab®.

Case p k R2

Gildehauser sandstone 0.58 ± 0.16 0.84 ± 0.44 1.00

Bentheimer (WW) sandstone 1.39 ± 0.12 4.46 ± 0.38 0.99

Bentheimer (MW) sandstone 0.66 ± 0.05 4.54 ± 0.80 0.98

Sintered glass bead pack 0.59 ± 0.04 2.65 ± 0.70 0.98

Although some trends are visible from Figure 39, it is difficult to make strong arguments for

connectivity using experimental data alone for several reasons. First, image resolution can strongly affect

the estimation of the Euler characteristic. Misidentification of fluid clusters during image segmentation can

lead to erroneous estimates of the Euler characteristic, as well as the pore space (used for normalization).

Page 164: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

144

Second, imaging experiments are expensive and often provide limited data points. Third, these example

cases demonstrate the complexity that can occur in understanding experimental measurements of porous

media with many simultaneously varying and less controllable parameters (e.g., wettability and pore

structure). For these reasons, we study connectivity trends using pore-network modeling (PNM) next. PNM

is advantageous because simulated paths are repeatable and input parameters are largely controllable.

Figure 40 shows examples of three hysteretic displacements for the weakly water-wet case. Each cycle

begins with primary drainage (PD) (solid curves) from zero oil saturation to a specified termination

saturation (So ~ 0.8, 0.9, and 1.0), followed by imbibition (IMB) (dashed curves) to residual saturation, and

secondary drainage (SD) (dotted curves) to 100% oil saturation. The PD curve ending at So ~ 1.0 reached a

capillary pressure of 105 Pa. The imbibition scans ended when oil no longer flowed as defined by -105 Pa

capillary pressure. More extensive simulations will be presented in later sections to model kr.

Figure 40. The χ - S paths for different drainage and imbibition scans using PNM for the weakly water-

wet case ( o

θ ~ 50 ). All PD curves begin at zero oil saturation but terminate at So ~ 1.0 (green), So ~ 0.9 (blue), and So ~ 0.8 (red). Next, IMB curves terminate at residual conditions (squares). Finally, all SD curves are simulated to So ~ 1.0.

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

PD

I

SD

IMB

Page 165: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

145

The PNM paths shown in Figure 40 are qualitatively similar to the experimental paths in Figure 39.

That is, the PNM paths are nearly linear for the imbibition paths. The drainage paths, however, are

significantly more nonlinear, likely because the nonwetting phase connects well only at higher saturations.

These results are consistent with the displacement experiments discussed by Schlüter et al. (2016).

The paths in Figure 40 were also fit to Eq. (5.14), along with six other similar scanning curves whose

PD paths terminated at smaller, but equally-spaced saturations (So ~ 0.2 – 0.7). The best fit parameters (p

and k) are presented in Figure 41.

Figure 41. Fitting parameters (p and k) after matching Eq. (5.14) to PNM scanning data (see Figure 40 for three of the scans plotted here at 1.0, 0.9, and 0.8 So). The x-axis is the nonwetting saturation at the termination of PD. The shaded region shows the error bars calculated for 95% confidence limits using the nlparci function in Matlab®. The contact angle averages at 50o.

Figure 41 shows that p is nearly constant for all drainage scans, whether primary or secondary. The

parameter, p, for drainage cycles is equal to 1 − as So approaches zero. The value of connectivity ( ) at

this limiting saturation results from only very few pores being filled by the nonwetting phase (the Euler

characteristic becomes a small positive number, such that the minimum and maximum possible values

determine the limit as So approaches zero, e.g., if only one pore were filled by the nonwetting phase, then,

Eq. (5.6) gives, ( ) ( )ˆ 1 16850 25787 16850 0.395/ = − − − = . For PD, therefore, ~ 1 0.4p − or 0.6. Further, all SD

curves that result from PD termination saturations less than about 0.7 have nearly the same p values, because

their initial ˆ S − values (residual values) lie close to the PD path. Only SD curves associated with high

Page 166: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

146

termination saturations (> 0.7) give somewhat larger values for p because their initial ˆ S − values lie below

the PD path in the ˆ S − space.

Imbibition values for p, however, vary significantly from about 0 to 0.8 and are quite uncertain at small

saturations. This uncertainty stems from the lack of a significant change in connectivity over the imbibition

curve. Values for p do not approach 1 − at zero saturation, but rather terminate at a residual ˆ S − value.

Further, an imbibition curve given by Eq. (5.14) does not begin at (1,1), but instead at a termination value

of the PD curve. Thus, p for IMB depend on the initial and final ˆ S − values.

The parameter, k, controls the linearity of the path in ˆ S − space. Figure 41 shows that k is significantly

greater than 1.0 for drainage and ~ 1.0 for imbibition. Thus, all drainage paths curve upwards with

increasing nonwetting saturation, while all imbibition processes are nearly linear as the nonwetting

saturation decreases. The value of k varies from about 3 to 10 for the drainage cycles. The SD and PD

values vary inversely with each other in Figure 41. The k values for the PD curves at small termination

saturations only represent a small portion of the entire PD curve that goes from saturation of 0.0 to 1.0.

Thus, the variation in these values is not representative of the entire PD curve. A value of around 10 matches

well the full PD curve and any portion of that curve in between. The values of k for the SD curves vary

from about 10 at small termination saturations to about 6 at large nonwetting saturation. This result is also

consistent with the value for the PD curve because small termination saturations give a larger change in

saturation so that k approaches 10. Thus, the SD curve that begins at saturation near the PD curve (i.e., for

small termination saturation) takes a similar path as the PD curve to saturation of 1.0, resulting in the same

values of p and k as the PD curve (the SD curve limit for small termination saturation is the PD curve).

These important results could be used to predict the paths in the ˆ S − space where a value of 10 for k

can be used for all drainage processes (PD or SD), while a value of 1.0 for k is reasonably accurate for all

imbibition curves, independent of their starting ˆ S − values. Further, the value of p can be estimated from

Eq. (5.14) for all drainage curves based on the starting ˆ S − values and for all imbibition curves based on

Page 167: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

147

the initial and residual ˆ S − values. A summary for the of p and k values for the different cycles of injection

is provided in Table 13.

Table 13. Summary remarks for p and k values for the evolution of phase connectivity for different cycles of injection.

Injection

process p k

Primary

drainage

All PD curves (if continued till the end)

appear to terminate at ˆ , S = 1,1. Therefore,

by constraining Eq. (5.14) at this point, we

can eliminate, p as (p = 1- 0 ), where 0 is

the value of at S = 0.

A value of k = 10 matches the full PD

curve regardless of where the PD curve

is terminated. This is because all ˆ , S

points lie on the full PD curve.

Imbibition

Imbibition starts and terminates at different

locations in the ˆ S − space, therefore, p, for

imbibition will be a fitting parameter for the

path under consideration

Give the linear nature of the imbibition

curves, k = 1, can be a reasonable

approximation.

Secondary

drainage

All SD curves terminate at ˆ , S = 1,1.

Therefore, by constraining Eq. (5.14) at this

point, we can eliminate, p as (p = 1- 0 ),

where 0 is the value of at S = 0.

To model the curvature observed in the

drainage curves, a high value of k for

drainage works well. Given the starting

point for SD (end of IMB) is different, k

can be a fitting parameter. However,

from observation, a value of k = 10,

works reasonably well for SD curves

5.3.2. Fitting kr-EOS to literature data

This section shows how to use the simplest form of the kr-EOS based on Eq. (5.13) (e.g., constant partial

derivatives) to fit the four data sets considered previously (see Table 11). The path in the ˆ S − space may

not be known from CT scans, so here we tune on only parameters k and

as might be done in practice

(see Eq. (5.15)). For reference, the parameter

is then calculated using the values of p in Table 12.

Figure 42 shows the best fits of the literature data using the new kr-EOS and the conventional Corey

expression. Because the Corey form uses the end-point value, the parameter S

was determined so that the

measured end-point relative permeability is obtained exactly (see Eq. (5.16)). The end-point saturations

Page 168: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

148

were also fixed and not tuned. Thus, the only tuning parameters were

and k for the EOS, while the

conventional exponent was tuned for Corey. As shown in Table 14, the fits are excellent and are slightly

better for the new EOS (see Bentheimer MW case at low saturations) compared to the Corey curve. The

low values of s

for three of the experiments suggest that the connectivity term dominates the behavior of

relative permeability. Unlike Corey, the parameters in the kr-EOS have physical meaning.

Figure 42. Best fits to the literature data in Table 11 using the (a) kr-EOS with constant partial derivatives and (b) Corey form.

Table 14. Best fit values for the kr-EOS and Corey form shown in Figure 42. The end-point permeability is the same for comparison purposes. Thus, only

χα and k are used as tuning parameters for the kr-EOS and

no for Corey. R2 values are shown. Key parameters from literature data kr-EOS Corey

Case ( )absk   D

caN

oiS

orS o

rk S

α χα k 2R

on 2

R

Gildehauser 1.5 9.1x10-7 0.77 0.30 0.49 7x10-8 1.76 2.24 1.00 1.44 1.00

Bentheimer WW 2.07 3x10-7 0.75 0.36 0.31 1x10-8 0.72 3.83 0.99 1.80 0.99

Bentheimer MW 2.198 9.9x10-7 0.89 0.11 0.93 0.14 4.06 10.50 1.00 7.24 0.99

Sintered glass bead pack 22 ~10-4 0.90 0 0.93 0.78 0.50 2.52 1.00 1.24 0.99

Although both the Corey form and kr-EOS show a reasonable match to literature data, the Corey form

has little or no predictive capability unless coupled with other models, such as an extended Land’s model

to capture hysteresis in the scanning curves. Even then Corey’s predictive capability is limited to fitting the

available data. The EOS, however, can honor expected physical trends in kr as a function of the input

0.0

0.2

0.4

0.6

0.8

1.0

0.0 0.2 0.4 0.6 0.8 1.0

0.0

0.2

0.4

0.6

0.8

1.0

0.0 0.2 0.4 0.6 0.8 1.0

GildehauserBentheimer MWBentheimer WWSintered glass bead pack

a b

Page 169: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

149

parameters, i.e., , S, and θ in the examples presented here. The remainder of the paper focuses on modeling

and predicting kr using the EOS.

5.3.3. Relative permeability scanning curves: pore-network simulations

In the previous section, the partial derivatives for the kr-EOS were assumed to be constant. We use pore-

network modeling (PNM) in this section to determine the values of these partial derivatives more exactly

and also to examine how they vary in the ˆ S − space. Relative permeability can then be estimated more

accurately for a given path from these derivatives.

Figure 43 shows the results for ten scans equally spaced in saturation (terminal saturation from PD)

for two contact angles (0o and ~ 50o). Some of these scans for ~ 50o

are also given in Figure 40. The use

of different starting points in the ˆ S − space allows for increasing the region covered in that space by the

various scans. This allows for derivative calculations.

Figure 43a and Figure 43b show that the values of relative permeability intersect each other (crossover),

as has been observed previously by Spiteri et al. (2008). This is also true in the ˆ S − space (see Figure 43e

and Figure 43f). The value of kr at these intersections is nearly the same even though the processes to obtain

that value of ˆ S − were quite different. This demonstrates that for these PNM scans and S are reasonably

sufficient to represent kr, i.e., the state function (EOS) approach is reasonable. The drainage scans (Figure

43c, Figure 43d, Figure 43g, and Figure 43h) do not have significant intersections.

Figure 43 also shows slight differences in kr during primary drainage between the two wettability cases

near low and intermediate oil saturations (see Figure 43a and Figure 43b). Here, the increase in kr is

somewhat greater for the completely water-wet case. Also, there is less hysteresis in kr for the weakly water-

wet case, although the region traversed in the ˆ S − space is larger. This behavior could be attributed to the

sudden pore-filling events possible in strong water-wet medium (water can move around the oil along the

Page 170: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

150

surfaces) as opposed to smoother more piston-like filling in the less water-wet case. Like the results

observed previously, the imbibition ˆ S − paths are significantly more linear than the drainage paths.

Furthermore, it can be noticed from Figure 43e, Figure 43f, Figure 43g, and Figure 43h that the upper

boundary of the ˆ S − space is controlled by the series of imbibition curves starting at different initial

saturations, while the lower boundary of the ˆ S − space is controlled by the secondary drainage curve

starting at the lowest ˆ S − point (one limit of the residual locus). The ˆ S − boundary which is the physical

space where relative permeability values exit depends on the wettability of the medium as well as the pore

structure constraint at S = 0, ˆ ˆ = o point (S, = 0, 0.395 for the porous medium studied here).

Page 171: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

151

Figure 43. PNM simulations of imbibition (IMB) and secondary drainage (SD). Two contact angles are used: the left column represents a fixed contact angle of 0o, while the right column is for uniformly distributed contact angles between 40o and 60o with an average of ~ 50o. The black solid line in (a) and (b) is for primary drainage (oil flood). Figures (a) and (b) also give ten imbibition curves at 0.1 saturation

intervals on the PD curve, while (e) and (f) show their χ - S paths. Figures (c) and (d) are for secondary

drainage and (g) and (h) show their corresponding χ - S paths. The red open circles represent the starting

point for IMB, while the black open circles represent the residual points for each IMB scan (similar to Figure 40). Arrows show the direction of saturation change (IMB points to the left while SD points to the right).

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

a

c

e

g

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0 0.2 0.4 0.6 0.8 1

b

f

d

h

Rel

ativ

e p

erm

eab

ility

Ph

ase

con

nec

tivi

ty

Page 172: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

152

Figure 44. Relative permeability contours plotted from the 200 PNM imbibition simulations for the two contact angle cases of 0o and ~ 50o.

Figure 45. PNM simulations for 200 imbibition and secondary drainage scanning curves that begin at different PD termination saturations (spaced by 0.005 saturation units). Two contact angles are shown, 0o (a) and ~ 50o (b). The PD curves begin at So = 0.

Figure 44 shows the nonwetting phase relative permeability contours generated using the imbibition

PNM simulations. Here, the upper boundary of the ˆ S − space is visible more clearly. Figure 45 shows

the entire PD-IMB-SD scanning curves and shows how relative permeability changes smoothly in the ˆ S −

space for 200 scans, again confirming the use of the state function concept. Further, in the saturation range

of ~ 0.45 to 0.65, ˆ S − paths cross and it is found that 0.126 0.013r

k = for 0o

= and 0.121 0.024r

k =

for ~ 50o

. This demonstrates that for these PNM simulations a similar ˆ S − value gives approximately

θ = 0o~ 50o

0

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 10.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 10.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

a b

Zero krZero kr

Page 173: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

153

one value of kr, independent of the path taken. This powerful concept may allow for the prediction of

complex relative permeability, characteristic of natural porous media.

Relative permeability is zero everywhere in the lower-left region of the ˆ S − space. The boundary of

this region, termed as the residual locus ( ˆ )rr S − , defines an important limit for the physical kr region. The

residual loci for the two wettability cases are compiled in Figure 46.

All imbibition curves must intersect the residual locus. Traditionally, only one residual saturation is

used as determined from a few experiments and arbitrary interpolation-based schemes are implemented for

predicting relative permeability across different paths. The knowledge of the residual locus can have

important implications for the accuracy of simulations when modeling hysteresis.

Figure 46. Locus of residual connectivity and residual saturation generated from the different scanning curves (Figure 43). The blue data point shows the limiting value of connectivity as saturation approaches zero (see Eq. (5.6)).

0

0.1

0.2

0.3

0.4

0.5

0 0.1 0.2 0.3 0.4 0.5

Page 174: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

154

The shape of the locus (Figure 46) shows two values of residual connectivity over a small range in

residual saturations. This region of two values decreases for the completely water-wet case because of little

change in residual connectivity. However, for the weakly water-wet case, efficient displacements of initially

well-connected oil led to a greater decrease in connectivity (fewer trapped oil blobs). This is also seen in

the literature, for example, more oil recovery occurs for mixed-wet rocks owing to the better displacement

of oil (Jadhunandan and Morrow 1995).

5.3.4. Estimation of relative permeability partial derivatives

We now provide estimates of the relative permeability partial derivatives using the 200 PNM scanning

curves show in Figure 45. For estimation of these partials at each ˆ S − value, we fit the equation of a plane

through a cluster of data points in the vicinity of that point using,

( ) ( ) ( )ˆ

,ˆ ˆˆ

r r

r rof of of

S

k kk k S S

S

− = − + −

(5.19)

where ( ),ˆ,   of of rofS k is a fixed data point. The data points in the cluster were selected based on the closest

Euclidean distance to the fixed point. The coefficients in the fitted plane then gave the partial derivatives

in Eq. (5.19) for that point. Different cluster sizes were examined, but an optimum of 700 neighboring data

points gave the best results to smooth the derivative estimates. The derivative values are shown in Figure

47.

Page 175: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

155

Figure 47. Relative permeability partial derivatives estimated from fitting a cluster of data to a plane.

Some common trends for both wettability cases can be inferred from Figure 47. First, the values are

not constant but vary significantly. Second, the partial derivatives in the ˆ S − space were generally

positive so that kr increases with both saturation and connectivity. The partial derivative, ( )ˆ

/r

k S

, was

found to be greater for high saturation than for low saturation (varied from 0 to 2.5), while the reverse trend

was observed for ( )ˆ/r S

k . The derivative, ( )ˆ/r S

k , however, is slightly negative near the residual locus

at the smaller connectivity values. This results from the locus shape, which means that kr below the residual

curve must increase to obtain kr = 0. The average values of the partials are reported in Table 15 which will

be used in the next section for predicting kr values using a single value of partial derivatives.

The overall trends in partial derivatives convey that changes in connectivity are more important at low

and intermediate saturations, while changes in saturation control relative permeability at high saturation.

This is true for both wettabilities considered.

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

Page 176: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

156

Table 15. The average values of the estimated partial derivatives of relative permeability shown in Figure 47.

Wettability ( )ˆ

/r

k S

( )ˆ/r S

k

0o

= 0.43 1.31

~ 50o

0.70 0.85

5.3.5. Prediction of relative permeability

We are interested in the prediction of kr at any given point in the ˆ S − space. Here, we present two

approaches for kr prediction using the partial derivatives estimated in the previous section.

First, we assume constant partial derivatives where Eq. (5.13) is re-written as,

( ) ( )ˆ

,ˆ ˆ. .ˆref

r r

r r re

me San

r

me

f ef

an

k kk k S S

S

+= +

− −

(5.20)

The mean values of the partial derivatives in Eq. (5.20) were taken from Table 15, while the reference

values were chosen at an intermediate ˆ S − value. As such, ( ˆ ,,ref

ref ref rS k ) were set as ( )0.6, 0.439, 0.245

for 0o

= and ( )0.545, 0.417, 0.126 for ~ 50o

. These were chosen because all ˆ S − paths traversed close

to the intermediate values. Ideally, predictions using an exact EOS do not depend on the chosen reference

state.

Second, we fit planes through the partial derivatives shown in Figure 47 using,

ˆ

ˆ ˆˆ

r r

S

k kc and f

SaS b dS e

= + =+ +

+

(5.21)

A constraint, b = d, was set such that the second order derivatives of the partials are equal, i.e.,

2 2ˆ ˆ/ /r rk S k S = . This honors the condition for the exact differential. Table 16 provides the values

Page 177: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

157

of the fitted parameters shown in Eq. (5.21). No additional constraints were used in matching these plane

fits. These plane fits were then used for generating the kr response.

Table 16. Values of plane fitting parameters in Eq. (5.21) used for the kr response in Eq (5.22).

Parameter

Wettability A B C D E F

0o

= 0.97 3.28 -1.50 3.28 -7.34 2.84

~ 50o

1.78 1.21 -0.95 1.21 -2.65 1.36

Equation (5.21) gives a quadratic response for kr when integrated from some reference state to the desired

state,

( ) ( ) ( ) ( ) ( )2 2 2 2ˆ .ˆ ˆ ˆ ˆ ˆ2 2ref

r r ref ref ref ref ref ref

a ek k S b S S c S SS f − + − + − += −+ − + (5.22)

The reference values were kept the same.

Here, we have only considered simple treatments of the partials. Higher order functions such a as

quadratic response could also be considered for the partials in the ˆ S − space which would lead to the

development of a cubic kr response (as opposed to the quadratic response shown in Eq. (5.22)). Moreover,

the functional forms of the partials could also be constrained in the physical ˆ S − space.

The predicted values of kr using the two approaches are presented in Figure 48. Values of kr < 0 were

set to 0.0 and kr > 1.0 were set to 1.0 in Figure 48. Predictions are solely based on the treatment of the

estimated partial derivatives and no constraints at limiting values were used.

The contour maps in Figure 48 show that the predicted kr response is slightly better for the quadratic

response. This is because the quadratic approach gives realistic values in opposite corners of the ˆ S −

space. The corner regions, although likely inaccessible through experiments, correspond with relative

permeability near zero because either connectivity or saturation remains low. This behavior is elegantly

Page 178: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

158

captured by the quadratic response. Also, the residual locus is captured to a reasonable extent as opposed

to a linear ˆrr S − locus given by the constant partial approach. Furthermore, as seen previously, the weakly

water-wet case shows a wider non-zero kr region owing to efficient displacement of the nonwetting phase.

Figure 48. Contour maps of predicted relative permeability using constant partial approach (a, b) and quadratic response (c, d) for both wettability cases. Actual versus predicted relative permeability values for the two approaches are shown in (e) and (f). The R2 for both approaches in (e) was ~0.90, while the R2 was ~0.90 and ~0.96 for the constant partial approach and quadratic response, respectively, in (f).

Figure 48e and Figure 48f show that excellent predictions are obtained in the region of actual data

(obtained from PNM simulations) for both methods. The likelihood of obtaining ˆ S − values outside of the

scanning region is low as evidenced by the fact that a single value for exists for zero saturation (~ 0.4 for

0

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 10.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 10.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 10.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 10.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

0.2

0.4

0.6

0.8

1

0.2 0.4 0.6 0.8 1

a b

c d

0.0

0.2

0.4

0.6

0.8

1.0

0.0 0.2 0.4 0.6 0.8 1.0

Constant partial

Plane fit

0.0

0.2

0.4

0.6

0.8

1.0

0.0 0.2 0.4 0.6 0.8 1.0

Co

nto

ur

map

s

Co

nstan

t partial

Qu

adratic resp

on

se

Constant partial

Quadratic response

e f

Page 179: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

159

the pore structure used in this study). Furthermore, even a simple EOS-based approach of constant partial

derivatives shows substantial improvement over what is traditionally done (Land-based models), where

relative permeability is fixed by Corey’s model for each kr curve, and arbitrary interpolations are used in

between each curve.

Overall, the EOS approach makes kr continuous in the space of the state parameters considered –

mitigating complex quick fixes to hysteresis as is accomplished from conventional approaches. Further, the

approach presented in this research has the ability to improve compositional simulation by avoiding phase

labeling discontinuities by incorporation of wettability into the EOS model.

Page 180: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

160

5.4. Concluding remarks

A general equation-of-state (EOS) approach is presented for modeling relative permeabilities, where we

considered the state parameters of saturation and connectivity. Two-phase flow data from the literature and

numerical data from pore-network modeling (PNM) were used to generate key parameters for the

application and testing of the EOS. The main conclusions from this research are as follows.

• Paths in connectivity-saturation ( ˆ )S − space are more linear during imbibition compared to

drainage. During imbibition, nonwetting phase disconnects owing to snap-off events which occur

at the same rate with saturation, but during drainage, the reconnection of the nonwetting phase

occurs at high saturations. These paths are adequately described by a simple function.

• Values of relative permeability are nearly identical at the same ˆ S − value, which justifies the

state function approach for the case studied

• The new kr-EOS matches literature data well, similar to the conventional Corey form. For relative

permeability curves with significant curvature, however, the new approach matches experimental

data better than Corey, especially at low saturations.

• A residual locus exists in ˆ S − space, which is a function of wettability. Traditional hysteresis

models either ignore this possibility or offer complex interpolation-based solutions. The approach

here gives an elegant predictive methodology that is consistent with physics.

• Partial derivatives of relative permeability with respect to and S show that relative permeability

depends more strongly on connectivity at small saturations, while saturation is more important at

high saturations.

• A simple first-order approximation of the constant partial derivatives gives an EOS with reasonable

predictive capability over the region of the scanning curves considered. Partial derivatives that are

linear functions of and S give a quadratic response of relative permeabilities which show

improvements over the first-order approximation.

Page 181: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

161

PNM offers an excellent bridge for incorporating pore-scale phenomenon toward the development of a

physics-based relative permeability model. Other simulation possibilities could also be employed such

computation fluid dynamics (CFD), Lattice Boltzmann methods (LBM). CFD is generally limited to only

a few pores and smoother geometries. LBM can handle rough surfaces and could be the study of future

work. These simulation models would, however, take much longer run-times, which would greatly slow-

down the possibility of multiple numerical experiments for building very large datasets.

The generalized EOS framework presented in this research can be extended for modeling capillary

pressures (Pc), as well as for finding functional connections between kr and Pc.

Page 182: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

162

CHAPTER 6. CONCLUDING REMARKS AND

OUTLOOK FOR FUTURE RESEARCH

Current modeling efforts for relative permeability depend on empirical functions of phase saturations.

These functions are nonunique and are therefore specific to the flow/experimental condition. As an

example, imbibition nonwetting phase relative permeabilities differ from drainage nonwetting phase

relative permeabilities. Each path then requires its own saturation function which can cause problems in

reservoir simulations of multiphase processes like CO2 sequestration or water alternating gas enhanced oil

recovery where drainage and imbibition processes often occur simultaneously. To address such issues,

research efforts have been presented in this dissertation to model relative permeabilities based on key

controlling parameters such as phase saturation, phase connectivity, capillary number, and wettability that

are known to affect relative permeabilities.

A quadratic response-based equation-of-state (EOS) for relative permeability was modeled in the phase

connectivity and phase saturation ( ˆ )S − space. Limiting conditions on the state parameters were explored

to constrain the EOS model physically. Different capillary number cases ranging from one to 10-6 were

considered in the model. Additional effort investigated the role of wettability on phase trapping using pore-

network modeling (PNM). An extended Land-based hysteresis trapping model was presented and compared

against models from the literature. In addition, models were presented to capture the trends of the loci

bounded by the residual phase connectivity and residual phase saturation for different contact angles.

Finally, numerical PNM data sets for two contact angles in the water-wet regime were used for calculating

partial derivatives of nonwetting phase relative permeabilities in the ˆ S − state parameters. A response for

relative permeability was derived using the calculated partial derivatives and compared against actual

values of relative permeabilities from PNM.

Page 183: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

163

For the implementation of the proposed EOS-based relative permeability model in reservoir

simulations, it is needed that the state parameters, namely, S, (at the least) are characterized for the flow

condition. As saturations are updated for each gridblock and for each timestep, similarly (and

simultaneously), the phase connectivity term should be updated for each gridblock and for each timestep.

Once the saturation and connectivity values are defined, the response surface can be implemented for

updating the relative permeabilities. The coefficients in the response surface should be tuned to

experimental data for the specific flow conditions (capillary number/wettability/pore structure). For

example, the flow conditions near wellbore will experience a different capillary number (higher) than far

away from the wellbore. Similarly, EOR processes of surfactant floods will have a different capillary

number than for a waterflood. Thus, typical special core analysis (SCAL) experiments should be

supplemented with pore-scale imaging/numerical simulations to extract measures of phase connectivity for

tuning the relative permeability response surfaces as is demonstrated in this work. Once tuned these will

remain fixed for the rest of the simulation sequence.

For cases where pore-scale data may not be readily available, it is advised that functional forms such

as the evolution of phase connectivity (as discussed in this dissertation) in the saturation space can be

utilized for characterizing phase connectivity at each saturation point. Here, however, the flow process such

as drainage (nonwetting phase increase) versus imbibition (nonwetting phase decrease) will be required to

accurately capture the evolution of phase connectivity. As such, these functional forms will depend on the

wettability and capillary number conditions for which they are characterized. Future work is therefore

recommended to fully characterize the functional forms of phase connectivity (and even fluid/fluid

interfacial areas) in the saturation space for different experimental conditions. This will be beneficial to

circumvent expensive (and often time-consuming) pore-scale experiments. In addition to the knowledge of

the evolution function, the characterization of the locus of the residual saturation and residual phase

connectivity which is the true limiting condition for relative permeabilities expressed in the ˆ S − space

will be important. This will enable to define the entire physical ˆ S − space where relative permeability

Page 184: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

164

will exist. A response can then be easily developed. This response will represent all scanning curves and

thus hysteresis will no longer be difficult to model.

For compositional and three-phase simulations, the role of wettability is critical. For example, in the

compositional space, properties of fluid phases can approach one another (near critical regions) which can

impact wettability at the fluid/fluid/solid contact point. To include the effect of wettability into the EOS a

simple procedure as outlined in Khorsandi et al. (2021) can be implemented. They defined the tuning

coefficients for the EOS (example, the kr response surface) individually for the “wetting” and “nonwetting”

phases instead of for each “phase label” (oil/gas/water) and proposed that the value of the tuned coefficient

for each phase label can be estimated by weighting between these wetting and nonwetting values, regardless

of the phase itself. In addition, they proposed the use of a wetting fraction for defining each phase label’s

wettability which is essentially the affinity factor of the that phase label to the solid surface.

Page 185: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

165

6.1. Key concluding remarks

The following conclusions can be made from this work.

• A simple quadratic response for relative permeability in the state parameters of phase connectivity and

phase saturation ( ˆ )S − captures trends across different capillary numbers. Model tuned for a capillary

number in the capillary dominated regime can show predictive capability for other capillary numbers

within the same regime.

• The role of phase connectivity is found to be critical for the model development. The sensitivity of kr-

S paths to capillary number is found to be implicitly caused by changes in phase connectivity. The

linear kr-S paths for high capillary numbers (small Corey exponents) and nonlinear kr-S paths for low

capillary numbers (high Corey exponents) are found to occur due to fast and slow changes in phase

connectivity, respectively.

• Limiting constraints help in the identification of physical region in the -S state space. Pore-network

simulations demonstrate that it may not be physically possible to have the state space with high phase

saturation and low phase connectivity and vice-versa.

• The analysis of the effect of wettability shows that both phase trapping as well as the locus of residual

saturation and residual phase connectivity are sensitive to contact angle changes. For low contact

angles, the residual locus remains fairly constant but, at higher contact angles, the shape of the residual

locus resembles a closed-loop due to pore structure constraints at negligible phase saturation at which

point the pore structure topology (min

and max

) govern the value of ˆ .

• Phase trapping was found to reduce significantly for low receding phase contact angles owing to pore-

scale mechanisms of layer flow of the receding phase and piston-like advance of the invading phase. A

newly developed extended Land-based hysteresis trapping model is found to capture the residual trends

for all contact angles.

Page 186: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

166

• Application of numerical techniques like pore-network simulations can help in the development of

physics-informed transport models. Pore-network simulations allow for the generation of hundreds of

data points in the ˆ S − state space under controlled conditions which facilitates the estimation of

relative permeability partial derivatives.

• The response derived for relative permeability from the estimated partial derivatives demonstrates

predictive capability for relative permeabilities over the entire data sets, regardless of the direction of

flow, thus mitigating hysteresis.

Page 187: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

167

6.2. Future research

The following recommendations are provided as possible extensions of future work from this dissertation.

• The modeling efforts presented in this work highlight the need for good quality and numerous

experimental data sets, especially with pore-scale measurements of Euler characteristic and fluid/fluid

interfacial areas that can help in furthering the development and testing of the models. Further, there is

a need for these data sets for different porous medium wettability, particularly, wettability conditions

representing oil-wet or CO2-wet conditions.

• Micro-flow experimental studies can be conducted for porous media with known pore structures

through the application 3-D printing techniques. This will be critical for modeling the kr-EOS where

the pore structure is constant (as was done using pore-network modeling in this study). Similarly, fixed

porous medium will be useful for repeatability of the flow measurements.

• Physics-informed simulation models like PNM can be improved with the provision of conducting

simulations for different capillary numbers. Also, these simulations can be constructed with inbuilt

options to output pore-scale measures of Euler characteristic and fluid/fluid interfacial areas directly.

• Significant errors can occur during image processing (denoising and phase segmentation) when

estimating pore-scale properties via experimental techniques such as x-ray imaging. Newer machine

learning-based tools such as Convolutional Neural Networks (CNN), U-Net, and Res-Net can be

explored for image segmentation and other image pre-processing steps.

• With the availability of numerical flow data sets such as the ones explored in this study, data driven

models such as Artificial Neural Networks (ANN) can be coupled with physical constraints for the state

parameters to develop physics-informed machine learning algorithms for transport properties such as

relative permeabilities. Such machine learning-based algorithms will depend on both the quality and

quantity of flow data which will be used for training and validation purposes. A two-step process is

envisioned. The first step will require experimental flow measurements to corroborate numerical

techniques such as PNM but for different flow conditions such as variable capillary numbers. Next,

Page 188: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

168

with the help of corroborated numerical techniques, large data sets will need to be generated to develop

and test data driven (and physics-informed) models such as ANN.

• Other polynomial forms for the kr-EOS can be studied. For example, the cubic form of kr response in

the ˆ S − space. This will allow for capturing the relative permeability partial derivatives using a

quadratic response for the partial derivatives in the ˆ S − space. The cubic form may also allow for

better representation of the residual locus in the ˆ S − space.

• Wettability indices such as the Amott index are used more traditionally to describe medium wettability.

Although these measures only provide an average wettability estimate. PNM can be used to bridge the

understanding of wettability from a local perspective (such as contact angle measurement) and

wettability indices to compare their effect on residual saturations.

• Although the focus in this work has been on modeling relative permeabilities, adaptations can be

extended to other flow properties such as modeling capillary pressures using the EOS approach

described in chapter 5. Similarly, adaptations can be developed for modeling base porous media

permeabilities by studying different characteristics that describe the structure of a porous medium such

as pore/grain size distribution, surface area of the pore space, topology of the pore space.

• A future extension to the EOS model can be on the inclusion of fluid/fluid interfacial areas into the

equation-of-state. Although this extension should be considered with implications to the final

deployment of the EOS in reservoir simulators—meaning that the improvement in the current EOS

after including fluid/fluid interfacial area should be substantial to warrant complicating the EOS model.

• Trends of fluid/fluid interfacial areas with saturations for different contact angles can be useful for

groundwater remediation applications. Simple models can also be developed to capture these trends

and predict interfacial areas for other experimental conditions. These models can also be useful for

capillary pressure hysteresis models that use fluid/fluid interfacial areas instead of Euler characteristic

for addressing hysteresis.

Page 189: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

169

APPENDICIES

Appendix A. Basic theory of Minkowski functionals

Minkowski functionals describe the geometric measure of size for any object. For a d-dimensional object,

d+1 Minkowski functionals will be required to fully characterize the geometric shape of the object

(Armstrong et al. 2018). For example, a sphere will require four Minkowski functionals, whereas a surface

element such as a triangle or a line segment will require three Minkowski functionals.

Application of Minkowski functionals is in use in understanding pore-scale fluid flow. The pore space

and the phase (oil/water) structures are arbitrary shapes, and their sizes can thus be characterized with the

understanding of Minkowski functionals. Similarly, fluid/fluid interfaces are arbitrary two-dimensional

objects whose sizes can be characterized from Minkowski functionals.

For a three-dimensional object, there will be four Minkowski functionals. The first Minkowski

functional (M0) represents the volume of the element in consideration. This can be used for characterizing

the volume measure for porous media applications such as the porosity and fluid saturations. An example

of the wetting and nonwetting phase in a capillary tube of radius, r, is shown in Figure 49.

Page 190: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

170

Figure 49. Schematic of capillary tube filled with wetting and nonwetting phases. A zoomed-in view of the interface across the phases is also shown.

The following equations express the Minkowski functionals of the wetting/nonwetting phases shown in

Figure 49.

( )3

20

2    1 ,

3

ww

rM r L sin

cos

= + − (A.1)

( )3

20

2     1 ,

3

nwnw

rM r L sin

cos

= − − (A.2)

where M0w and M0

nw are the wetting and nonwetting phase first Minkowski functionals for the phases shown

in Figure 49; Lw and Lnw are the lengths of the capillary occupied by the wetting and nonwetting phases in

Figure 49; θ is the contact angle measured through the wetting phase.

The second Minkowski functional (M1) represent the surface area of the element under consideration.

( )1   ,

E

EM ds

= (A.3)

90- r

2

Wetting phase injection

Nonwetting phase recedes

Lnw Lw

L

2 90- rr

Page 191: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

171

where M1(E) is the second Minkowski functional of element E; ds and δE are the surface element on and

the entire surface of the object E, respectively. This can be used for characterizing the surface area

properties of importance to porous media applications such as the surface area of the solid (or the specific

surface area of the solid), phase total surface areas of the wetting/nonwetting phases (or the specific fluid

surface areas). The specific surface areas of the fluids can be used for estimating the fluid/fluid interfacial

areas. Examples of second Minkowski functionals for the wetting and nonwetting phase elements in Figure

49 can be estimates as,

1 2 1 ,ww

rM r L sin

cos

= + −

(A.4)

1 2 1 ,nwnw

rM r L sin

cos

= + −

(A.5)

where M1w and M1

nw are the wetting and nonwetting phase second Minkowski functionals for the phases

shown in Figure 49.

The third Minkowski functional (M2) represents the integral of the mean curvature of the element under

consideration and is given as,

( )21 2

1 1,M ds

R R

= +

E

E (A.6)

where R1 and R2 are the principal radii of curvatures of the interfacial element E. It can be used for the

estimation of capillary pressure. For the capillary tube example, with spherical curvature (R1 = R2 = R),

2

1 1  2  2  ,nw cos

MR R R r

= + = =

(A.7)

Page 192: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

172

where M2nw is Minkowski functional representing the wetting/nonwetting interphase for the demonstration

in Figure 49; r is the radius of the capillary tube. From Young-Laplace equation, for a spherical interface,

the capillary pressure across the interface can be linked to the third Minkowski functional as,

2 .1 1  2

σ     nwc

cosP M

R R r

= + = =

(A.8)

The fourth Minkowski functional (M3) is represented as the integral of Gaussian curvature and is given

as follows,

( )31 2

1,

E E

EM Kds dsR R

= =

(A.9)

where M3(E) is the fourth Minkowski functional of element E; ds and δE are the surface element and the

entire surface, respectively; K=[1/R1R2] is the Gaussian curvature of the element in consideration. Three

examples of different surfaces with different Gaussian curvatures are shown in Figure 50. The special case

of the Gauss-Bonet theorem relates the fourth Minkowski functional (Gaussian curvature) to the Euler

characteristic of the element.

Figure 50. Schematic of different surfaces with different types of Gaussian curvatures.

Positive Gaussian curvature Negative Gaussian curvature Zero Gaussian curvature

E EE

Page 193: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

173

The most general form of the Gauss-Bonet theorem states that “The integral of the Gaussian curvature

over a surface E with boundary δE plus the integral of the geodesic curvature (kg) around that boundary

plus all exterior corner angles (θi) is equal to 2π times the Euler characteristic χ(δE) of the surface.”

Mathematically, the Gauss-Bonet theorem is expressed as (Gluck 2012; Watkins),

    2 ( ).

E E

Eg iKds k dl

+ + = (A.10)

For a closed element, the integral of the geodesic curvature becomes zero. As such, the following

simplification to the Gauss-Bonet theorem can be made,

( )  2 .

E

EiKds

+ = (A.11)

A further simplification can be made by consideration of closed surfaces that are smooth. Smooth surfaces

ensure that the sum of the exterior angles goes to zero. This simplifies the Gauss-Bonet theorem as,

( )2 .E

E

Kds

= (A.12)

This simplification helps link the fourth Minkowski functional to the Euler characteristic. From Eqs. (A.9)

and (A.12), we get,

( ) ( )3 2 .E E

E

Kds M

= = (A.13)

Page 194: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

174

Further, the Euler characteristic for the surface element ( )( )E and the Euler characteristic for the solid

object ( )( )E are related as follows (Armstrong et al. 2018),

( ) ( )2 .E E = (A.14)

A few examples of 3-D objects and their Euler characteristics are provided in Table 17.

Table 17. Example of different objects and their Euler numbers.

Object M3

Euler number

(Minkowski definition:

A.13 and A.14)

Euler number (Betti

number definition:

χ = β0 – β1 + β2

Hollow sphere

4π 2 1 – 0 + 1 = 2

Solid sphere

4π 1 1 – 0 – 0 = 1

Hollow torus

0 0 1 – 2 + 1 = 0

Solid torus

0 0 1 – 1 + 0 = 0

χ is the Euler characteristic of the object; β0 is the number of objects; β1 is the number of

handles or loops in the objects; β2 is the number of empty cavities in the objects.

Page 195: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

175

Appendix B. A consistent approach for χmax determination

The values of χmin and χmax for a porous medium are used in this research to normalize the Euler

characteristic as follows,

.ˆ 

  

max

min max

−=

− (B.1)

The values of χmin and χmax are calculated differently depending on whether experimental data are used, or

pore-network models are used. When using image data, software like Avizo is used for Euler characteristic

estimation of a phase. For example, χmin is estimated for the pore space using the value of V-E+F-O for the

Euler number, where V is the number of vertices; E is the number of edges; F is the number of faces; and

O is the number of objects (voxels). The parameter χmax, however, cannot be directly estimated from image

data. It must be extracted from a pore-network of the dry image of the pore space. The pore-network gives

the information of the np (number of pores) and nt (number of throats). The parameter χmax is then the np

value.

Using the pore-network information, χmin can also be calculated as np – nt from PNM, which we refer

to as χmin-PNM. However, χmin when estimated by Avizo (χmin-image) may not be the same as χmin-PNM. This is

because, the pore space can be different before and after generating the pore-network. A further problem is

that all of the intermediate χ values for different saturations is estimated using the value of V-E+F-O when

image data is used.

Overall, the following methods can be used for consistency.

1. Use the pore-network approach on each individual saturation to find χ for each saturation. Thus,

we use χmax-PNM and χmin-PNM from the pore-network over the pore space. This, however, is not the

usual method for estimating χ at any saturation as is done using imaging data.

Page 196: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

176

2. Use the value V-E+F-O for all saturations as well as for the pore space to determine χmin-image. Then,

to find χmax, generate a pore-network of the pore space to find the coordination number (Z = 2nt/np)

instead of directly using χmax-PNM as the number of pores. The value of χmax-image is then determined

from the coordination number as,

  .

1  2

mim image

max image Z

− =

(B.2)

Page 197: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

177

Appendix C. Base regression code created for analysis presented in chapter 3

clc

clear all

close all

% % Created by Prakash Purswani

% %----------- Linear Regression with new equality conditions

% %-------4 Equality constraints

% % 1. kr = 1 at S = 1, xhat = 1

% % 2. kr = 0 at S = 0, xhat = 1

% % 3. dkr/dS = 0 at S = 1, xhat = 0

% % 4. dkr/dxhat = 0 at S = 0, xhat = 1

%----------Armstrong et al. (2016): Ca = 10^0

Snw = [0.08784 0.28827 0.30675 0.43963 0.47871 0.47904 0.66023 0.67106 0.76941 0.80114];

krnw = [0.154493616 0.416947583 0.451528114 0.576595911 0.620941523 0.631093263

0.817376274 0.827979284 0.902661087 0.923611531];

Ca = [1.627329193 1.42152634 1.377300437 1.299488153 2.057539683 1.214774557 1.447147458

1.400074764 1.041695422 0.897429262];

X = [2933 -2286 -2353 -5011 -7221 -5957 -10300 -10544 -10682 -10594];

%---------Planar equation fitting AS+BX+C = kr (lower end)

syms A1 B1 C1

for i = 1:3

eqn1(i) = A1*Snw(i) + B1*X(i) + C1 == krnw(i);

end

%-------Linear equation fitting (aS+b = X)

J1 = [Snw(1:3)' ones(3,1)];

Page 198: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

178

j1 = regress(X(1:3)',J1);

%---------Planar equation fitting AS+BX+C = kr (upper end)

syms A2 B2 C2

for i = 1:3

eqn2(i) = A2*Snw(length(Snw)-3+i) + B2*X(length(Snw)-3+i) + C2 == krnw(length(Snw)-3+i);

end

%-------Linear equation fitting (aS+b = X)

J2 = [Snw(end-2:end)' ones(3,1)];

j2 = regress(X(end-2:end)',J2);

Sol1 = solve([eqn1(1), eqn1(2), eqn1(3)], [A1, B1, C1]);

ansA1 = double(Sol1.A1);

ansB1 = double(Sol1.B1);

ansC1 = double(Sol1.C1);

Sol2 = solve([eqn2(1), eqn2(2), eqn2(3)], [A2, B2, C2]);

ansA2 = double(Sol2.A2);

ansB2 = double(Sol2.B2);

ansC2 = double(Sol2.C2);

%---------For Xmax, kr = 0

syms m n

eq1 = ansA1*m + ansB1*n + ansC1 == 0;

eq2 = j1(1)*m - n + j1(2) == 0;

Sol3 = solve([eq1, eq2], [m, n]);

ansm = double(Sol3.m);

ansn = double(Sol3.n);

Page 199: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

179

%---------For Xmin, kr = 1

syms f g

eq11 = ansA2*f + ansB2*g + ansC2 == 1;

eq22 = j2(1)*f - g + j2(2) == 0;

Sol4 = solve([eq11, eq22], [f, g]);

ansf = double(Sol4.f);

ansg = double(Sol4.g);

%------------ Pore structure Euler connectivity values identified

Xmax = ansn;

Xmin = ansg;

%---------------Tuning against Ca = 10^-4

%----------Armstrong et al. (2016): Ca = 10^-4

Snw = [0.09204 0.29338 0.31264 0.44507 0.48577 0.66348 0.67386 0.77207 0.80465 0.88334

0.900609];

krnw = [0.086075394 0.197302896 0.298669867 0.357956049 0.451574991 0.626190811

0.614427609 0.763573275 0.789863744 0.908177118 0.924545659];

Ca = [0.000492236 0.000400132 0.000380349 0.000347567 0.000371549 0.000974638 0.000835806

0.000555354 0.000523234 0.000336238 0.000286985];

X = [158 -400 -661 -1883 -2449 -3341 -3374 -4553 -5306 -7054 -7539];

xhat = (X-Xmax)/(Xmin-Xmax); % normalized Euler connectivity

%----------Linear Regression

X = [(ones(size(Snw)) - 2*xhat + xhat.^2)' (-2*Snw + Snw.^2 + xhat.*Snw)'];

xx = regress((krnw-xhat.*Snw)',X)

a = xx(1);

b = xx(2);

Page 200: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

180

% --------------- Check the goodness of fit ---------------

vtest = a*(1 - 2*xhat + xhat.^2) + b*(-2*Snw + Snw.^2 + xhat.*Snw) + xhat.*Snw;

for i = 1:length(vtest)

if vtest(i) < 0

vtest(i) = 0;

end

end

Residual = vtest - krnw;

MSE = (sum((vtest - krnw).^2));

M = mean(vtest);

SStotal = (sum((vtest - M).^2));

R2_10_4 = 1 - MSE/SStotal;

Mean_Residual = mean(Residual);

count = 1;

for i = 0:0.1:1

count_i(count) = i;

mean_R(count) = Mean_Residual;

mean_for_plot(count) = 0;

count = count + 1;

end

% --------------- Plotting Surface ---------------

syms x y

h1 = ezsurf(a*(1 - 2*y + y.^2) + b*(-2*x + x.^2 + y.*x) + x.*y, [0, 1, 0, 1]);

h1 = findall(h1,'Type','Surface');

X1 = get(h1,'XData');

Y1 = get(h1,'YData');

Z1 = get(h1,'ZData');

Page 201: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

181

%---------Setting negative data for kr in the surface to zero

for i = 1:length(Z1)

for j = 1:length(Z1)

if Z1(i,j) < 0

Z1(i,j) = 0;

end

end

end

figure

plot3(Snw,xhat,krnw,'k.','markersize',30,'linewidth',2)

hold on

contourf(X1,Y1,Z1)

alpha 0.9

shading interp

colorbar

xlabel('Phase saturation','fontsize',20)

ylabel('Normalized Euler characteristic','fontsize',20)

zlabel('Phase relative permeability','fontsize',20)

%title('Fitting Response Surface-Data-Armstrong et al 2017')

%legend('Experimental data')

set(gca,'fontsize',25,'linewidth',2)

% set(get(gca,'YLabel'),'Rotation',10);

% set(get(gca,'XLabel'),'Rotation',-6);

%zlim([0 1])

view(0,90)

grid on

%------------ Plot Error for each data point

figure

Page 202: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

182

plot(Snw,Residual,'ko','markersize',10,'linewidth',2)

hold on

plot(count_i,mean_for_plot,'r--','markersize',10,'linewidth',2)

xlabel('Phase saturation','fontsize',20)

ylabel('Residual','fontsize',20)

set(gca,'fontsize',25,'linewidth',2)

% End of code %

Page 203: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

183

Appendix D. Data used in chapter 3 (adapted from Armstrong et al. 2016)

Table 18. Data used in chapter 3 for nonwetting phase saturation, connectivity, and relative permeability (adapted from Armstrong et al. 2016).

-(log(NCa)) Snw Sw krnw χnw nw

0

0.09 0.91 0.15 2933 0.17

0.29 0.71 0.42 -2286 0.49

0.31 0.69 0.45 -2353 0.49

0.44 0.56 0.58 -5011 0.65

0.48 0.52 0.62 -7221 0.79

0.48 0.52 0.63 -5957 0.71

0.66 0.34 0.82 -10300 0.98

0.67 0.33 0.83 -10544 0.99

0.77 0.23 0.90 -10682 1.00

0.80 0.20 0.92 -10594 0.99

1

0.09 0.91 0.16 1116 0.28

0.29 0.71 0.41 -1140 0.42

0.31 0.69 0.45 -1426 0.44

0.44 0.56 0.57 -3458 0.56

0.48 0.52 0.61 -4190 0.60

0.67 0.33 0.84 -7084 0.78

0.77 0.23 0.92 -8206 0.85

0.80 0.20 0.93 -8583 0.87

0.88 0.12 0.97 -9461 0.92

0.90 0.10 0.97 -10484 0.99

2

0.09 0.91 0.15 591 0.32

0.29 0.71 0.40 -508 0.38

0.31 0.69 0.45 -1426 0.44

0.31 0.69 0.44 -1371 0.43

0.31 0.69 0.43 -1112 0.42

0.44 0.56 0.56 -2467 0.50

0.48 0.52 0.60 -3176 0.54

0.66 0.34 0.81 -5266 0.67

0.67 0.33 0.82 -5390 0.68

0.77 0.23 0.90 -6716 0.76

0.80 0.20 0.92 -7204 0.79

0.90 0.10 0.98 -9783 0.94

Page 204: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

184

3

0.09 0.91 0.09 142 0.34

0.29 0.71 0.22 -399 0.38

0.31 0.69 0.32 -677 0.39

0.44 0.56 0.38 -1889 0.47

0.49 0.51 0.45 -2443 0.50

0.66 0.34 0.72 -3669 0.57

0.67 0.33 0.71 -3597 0.57

0.77 0.23 0.82 -4880 0.65

0.80 0.20 0.85 -5624 0.69

0.88 0.12 0.93 -7322 0.79

0.90 0.10 0.94 -7752 0.82

4

0.09 0.91 0.09 158 0.34

0.29 0.71 0.20 -400 0.38

0.31 0.69 0.30 -661 0.39

0.45 0.55 0.36 -1883 0.47

0.49 0.51 0.45 -2449 0.50

0.66 0.34 0.63 -3341 0.55

0.67 0.33 0.61 -3374 0.56

0.77 0.23 0.76 -4553 0.63

0.80 0.20 0.79 -5306 0.67

0.88 0.12 0.91 -7054 0.78

0.90 0.10 0.92 -7539 0.81

5

0.09 0.91 0.02 167 0.34

0.29 0.71 0.13 -62 0.35

0.31 0.69 0.23 -498 0.38

0.44 0.56 0.31 -1481 0.44

0.48 0.52 0.41 -2071 0.48

0.66 0.34 0.53 -2910 0.53

0.77 0.23 0.77 -4535 0.63

0.80 0.20 0.78 -5265 0.67

0.88 0.12 0.91 -7029 0.78

0.90 0.10 0.93 -7518 0.81

6

0.29 0.71 0.22 -62 0.35

0.44 0.56 0.31 -1288 0.43

0.48 0.52 0.27 -1839 0.46

0.66 0.34 0.56 -2905 0.53

0.67 0.33 0.78 -2966 0.53

0.77 0.23 0.68 -4097 0.60

0.81 0.19 0.71 -4797 0.64

0.88 0.12 0.87 -6509 0.75

0.90 0.10 0.89 -6970 0.77

Table 16. (Continued)

Page 205: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

185

Appendix E. Procedure for developing iso-quality curves discussed in chapter 4

Figure 51. Schematic showing the procedure for developing the iso-quality curves discussed in chapter 4 (Figure 27).

Iso-connectivityIso-saturation

2-D interpolation to find kr at gridded pointsIso-quality curves

Gridded data at equal -S spacingOriginal PNM data

Page 206: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

186

Appendix F. Residual curves for different stopping criteria and wettability alteration in PNM

Figure 52 shows the initial-residual (IR) saturation curves generated using pore-network modeling for four

different contact angles. The contact angles here are reported through phase1 (receding phase during the

secondary process). Four different stopping criteria are shown for obtaining the residual saturations to

compare residual saturations that may be obtained from typical core flooding experiments where the

endpoint of the experiments is judged based on a very high fractional flow near 0.99. These include, three

kr stopping criteria of 10-2,10-3, and 10-5, while the fourth criterion corresponds to the high capillary pressure

of -105 Pa.

Figure 52. Initial-residual saturation curves generated from pore-network modeling for different contact angle cases measured through the receding phase. The residual saturations are obtained for different stopping criteria and are shown by different colors.

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.2 0.4 0.6 0.8 1

S r

Si

Stop at kr = 10^-2

Stop at kr = 10^-3

Stop at kr = 10^-5

Stop at Pc = -10^5 Pa

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.2 0.4 0.6 0.8 1

S r

Si

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.2 0.4 0.6 0.8 1

S r

Si

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.2 0.4 0.6 0.8 1

S r

Si

θ1 = 180o θ1 = 120o

θ1 = 60o θ1 = 0o

Page 207: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

187

Figure 52 shows that residual curves depend on the criteria adopted for marking the residual state. The

capillary pressure-based criterion shows the lowest values of the residual saturation for all contact angles

and for all initial oil saturations, suggesting the ultimate residual saturations at those respective conditions.

Further, the difference in the residual saturations is significant for lower receding phase contact angles of

60° and 0°. This is because at these contact angles pore-scale flow regimes of layer flow and piston-like

advance by the advancing phase play a crucial role in driving low saturations. All four contact angle cases

in Figure F1 show a deviation (around Si ~ 0.35) in the residual saturation curves for the kr stopping criterion

cases. This is because for such low initial oil saturation the phase is not able to flow and as such the relative

permeability remains negligible.

Figure 53 shows capillary pressure scanning curves for two different contact angles generated using

pore-network modeling. For both sets, the primary phase1 injection was conducted at a receding contact

angle of 180°, while during the water injection step, the receding contact angles were 180° (no alteration)

and 30° (after wettability alteration). The capillary pressure for a no wettability alteration remains positive

while a sudden shift in the capillary pressure for the 30° case is observed after wettability alteration.

Figure 53. Primary drainage and water injection capillary pressure scanning curves generated using pore-network modeling. The different colors (and the arrows mark) give different cycles of injection. These scanning curves are generated with the endpoint of primary drainage (or starting point of water injection) at Si = 0.9.

Primary phase1 injection θ1 = 180o

Primary phase2 injection θ1 = 180o

Primary phase2 injection θ1 = 30o

Si = 0.9

(Pa)

Page 208: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

188

Appendix G. Base code for running PNM for generating numerical data sets for chapters 4 and 5

clc

clear all

close all

fclose('all');

% Created by Prakash Purswani

cd C:\Users\pxp5185 % Location for input files of extracted pore network

% Extract pore network’s Euler characteristic (Xmax and Xmin)

[np, nt] = getNpNt('Bentheimer_Lin/Bentheimer'); % get these from _node1.dat and _link1.dat files

Z = 2*nt/np;

Xmax = np;

Xmin = np - nt;

cd C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0 % Current location

% Subroutine for running PNM code (User can keep/remove secondary drainage as per requirement)

counter = 1;

for ii = 0:0.1:0.9

filecontent = fileread('input_pnflow.dat');

newcontent = regexprep(filecontent,'0.000', sprintf('%.3f',ii));

fid = fopen('input_pnflow_run.dat', 'w');

fwrite(fid, newcontent);

fclose(fid);

system('"pnflow_tom.exe" input_pnflow_run.dat');

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\*.m', 'Theta_0_S_0.1_0.2')

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Bentheimer_out.prt',

'Theta_0_S_0.1_0.2')

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\*.txt', 'Theta_0_S_0.1_0.2')

Page 209: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

189

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Bentheimer_res\*.vtu',

'Theta_0_S_0.1_0.2')

sub_folder = sprintf('S=%.3f',ii);

mkdir(['C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2\',sub_folder]);

cd Theta_0_S_0.1_0.2

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2\*.vtu', sub_folder)

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2\*.txt', sub_folder)

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2\*.m', sub_folder)

movefile ('C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2\*.prt', sub_folder)

files = dir;

directoryNames = {files([files.isdir]).name};

for k = 3:length(directoryNames)

currD = files(k).name;

cd(currD)

% kr and S measurements

Bentheimer_cycle1_drain;

Bentheimer_cycle2_imb;

Bentheimer_cycle3_drain;

[row_cycle1,cols_cycle_1] = size(Res_draincycle_1);

for i = 1:row_cycle1

Sw_PD(counter,i) = Res_draincycle_1(i,1);

Pc_PD(counter,i) = Res_draincycle_1(i,2);

krw_PD(counter,i) = Res_draincycle_1(i,3);

kro_PD(counter,i) = Res_draincycle_1(i,4);

end

[row_cycle2,cols_cycle_2] = size(Res_imb_2);

for i = 1:row_cycle2

Page 210: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

190

Sw_I(counter,i) = Res_imb_2(i,1);

Pc_I(counter,i) = Res_imb_2(i,2);

krw_I(counter,i) = Res_imb_2(i,3);

kro_I(counter,i) = Res_imb_2(i,4);

end

[row_cycle3,cols_cycle_3] = size(Res_draincycle_3);

for i = 1:row_cycle3

Sw_SD(counter,i) = Res_draincycle_3(i,1);

Pc_SD(counter,i) = Res_draincycle_3(i,2);

krw_SD(counter,i) = Res_draincycle_3(i,3);

kro_SD(counter,i) = Res_draincycle_3(i,4);

end

% Subroutine for extracting Euler number information

count = 1;

EulerFiles_PD = dir('*1_OInj*.vtu'); % For Primary Drainage cycle

for n = 1:length(EulerFiles_PD)

filename = EulerFiles_PD(n).name;

filetext = fileread(filename);

addpath('C:\Users\pxp5185')

ffaz_data = int8(getVtuData( filetext, 'PointData', 'ffaz'))-1;

connectivity_data = getVtuData( filetext, 'Cells', 'connectivity');

sumNhtO = 0;

sumNhtO_NW = 0;

npnt2=np+nt+2;

count2 = 1;

for i = 0:nt-1

iPt(count2) = connectivity_data(i*3+3)+1;

iP1(count2) = connectivity_data(i*3+1)+1;

iP2(count2) = connectivity_data(i*3+2)+1;

Page 211: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

191

ffT(count2) = ffaz_data(iPt(count2));

ffP1(count2) = ffaz_data(iP1(count2));

ffP2(count2) = ffaz_data(iP2(count2));

if(ffT(count2) == 0)

sumNhtO = sumNhtO +((ffP1(count2)==0) && (iP1(count2) <= npnt2)) + ((ffP2(count2)==0) &&

(iP2(count2) <= npnt2));

end

if(ffT(count2) == 1)

sumNhtO_NW = sumNhtO_NW +((ffP1(count2)==1) && (iP1(count2) <= npnt2)) +

((ffP2(count2)==1) && (iP2(count2) <= npnt2));

end

count2 = count2 + 1;

end

poresffaz=ffaz_data(3:np+2);

throatffaz=ffaz_data(np+3:np+nt+2);

Euler_NW_PD(counter,count) = sum(poresffaz == 1) + sum(throatffaz == 1) - sumNhtO_NW;

Euler_W_PD(counter,count) = sum(poresffaz == 0) + sum(throatffaz == 0) - sumNhtO;

count = count + 1;

end

count = 1;

EulerFiles_I = dir('*2_WInj*.vtu'); % For imbibition cycle

for n = 1:length(EulerFiles_I)

filename = EulerFiles_I(n).name;

filetext = fileread(filename);

addpath('C:\Users\pxp5185')

ffaz_data = int8(getVtuData( filetext, 'PointData', 'ffaz'))-1;

connectivity_data = getVtuData( filetext, 'Cells', 'connectivity');

sumNhtO = 0;

sumNhtO_NW = 0;

npnt2 = np+nt+2;

count2 = 1;

Page 212: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

192

for i = 0:nt-1

iPt(count2) = connectivity_data(i*3+3)+1;

iP1(count2) = connectivity_data(i*3+1)+1;

iP2(count2) = connectivity_data(i*3+2)+1;

ffT(count2) = ffaz_data(iPt(count2));

ffP1(count2) = ffaz_data(iP1(count2));

ffP2(count2) = ffaz_data(iP2(count2));

if(ffT(count2) == 0)

sumNhtO = sumNhtO +((ffP1(count2)==0) && (iP1(count2) <= npnt2)) + ((ffP2(count2)==0) &&

(iP2(count2) <= npnt2));

end

if(ffT(count2) == 1)

sumNhtO_NW = sumNhtO_NW +((ffP1(count2)==1) && (iP1(count2) <= npnt2)) +

((ffP2(count2)==1) && (iP2(count2) <= npnt2));

end

count2 = count2 + 1;

end

poresffaz = ffaz_data(3:np+2);

throatffaz = ffaz_data(np+3:np+nt+2);

Euler_NW_I(counter,count) = sum(poresffaz == 1) + sum(throatffaz == 1) - sumNhtO_NW;

Euler_W_I(counter,count) = sum(poresffaz == 0) + sum(throatffaz == 0) - sumNhtO;

count = count + 1;

end

count = 1;

EulerFiles_SD = dir('*3_OInj*.vtu'); % For Secondary Drainage cycle

for n = 1:length(EulerFiles_SD)

filename = EulerFiles_SD(n).name;

filetext = fileread(filename);

addpath('C:\Users\pxp5185')

ffaz_data = int8(getVtuData( filetext, 'PointData', 'ffaz'))-1;

connectivity_data = getVtuData( filetext, 'Cells', 'connectivity');

sumNhtO = 0;

Page 213: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

193

sumNhtO_NW = 0;

npnt2=np+nt+2;

count2 = 1;

for i = 0:nt-1

iPt(count2) = connectivity_data(i*3+3)+1;

iP1(count2) = connectivity_data(i*3+1)+1;

iP2(count2) = connectivity_data(i*3+2)+1;

ffT(count2) = ffaz_data(iPt(count2));

ffP1(count2) = ffaz_data(iP1(count2));

ffP2(count2) = ffaz_data(iP2(count2));

if(ffT(count2) == 0)

sumNhtO = sumNhtO +((ffP1(count2)==0) && (iP1(count2) <= npnt2)) + ((ffP2(count2)==0) &&

(iP2(count2) <= npnt2));

end

if(ffT(count2) == 1)

sumNhtO_NW = sumNhtO_NW +((ffP1(count2)==1) && (iP1(count2) <= npnt2)) +

((ffP2(count2)==1) && (iP2(count2) <= npnt2));

end

count2 = count2 + 1;

end

poresffaz=ffaz_data(3:np+2);

throatffaz=ffaz_data(np+3:np+nt+2);

Euler_NW_SD(counter,count) = sum(poresffaz == 1) + sum(throatffaz == 1) - sumNhtO_NW;

Euler_W_SD(counter,count) = sum(poresffaz == 0) + sum(throatffaz == 0) - sumNhtO;

count = count + 1;

end

cd C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2

rmdir(['C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0\Theta_0_S_0.1_0.2\',sub_folder],'s');

end

counter = counter + 1;

cd C:\Users\pxp5185\S_01_02_remaining_Copy_2_PD_0

end

% Calculating final outputs

Page 214: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

194

Snw_PD = 1- Sw_PD;

Xhat_NW_PD = (Euler_NW_PD-Xmax)/(Xmin-Xmax);

Xhat_W_PD = (Euler_W_PD-Xmax)/(Xmin-Xmax);

Snw_I = 1- Sw_I;

Xhat_NW_I = (Euler_NW_I-Xmax)/(Xmin-Xmax);

Xhat_W_I = (Euler_W_I-Xmax)/(Xmin-Xmax);

Snw_SD = 1- Sw_SD;

Xhat_NW_SD = (Euler_NW_SD-Xmax)/(Xmin-Xmax);

Xhat_W_SD = (Euler_W_SD-Xmax)/(Xmin-Xmax);

% For making matrices vertical

Snw_PD = Snw_PD';

Snw_I = Snw_I';

Snw_SD = Snw_SD';

Sw_PD = Sw_PD';

Sw_I = Sw_I';

Sw_SD = Sw_SD';

Euler_NW_PD = Euler_NW_PD';

Euler_NW_I = Euler_NW_I';

Euler_NW_SD = Euler_NW_SD';

Euler_W_PD = Euler_W_PD';

Euler_W_I = Euler_W_I';

Euler_W_SD = Euler_W_SD';

Xhat_NW_PD = Xhat_NW_PD';

Xhat_NW_I = Xhat_NW_I';

Xhat_NW_SD = Xhat_NW_SD';

Xhat_W_PD = Xhat_W_PD';

Xhat_W_I = Xhat_W_I';

Xhat_W_SD = Xhat_W_SD';

kro_PD = kro_PD';

kro_I = kro_I';

kro_SD = kro_SD';

Page 215: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

195

krw_PD = krw_PD';

krw_I = krw_I';

krw_SD = krw_SD';

Pc_PD = Pc_PD';

Pc_I = Pc_I';

Pc_SD = Pc_SD';

% Saving excel files in the same folder

writematrix(Snw_PD,'Snw_PD.csv')

writematrix(Snw_I,'Snw_I.csv')

writematrix(Snw_SD,'Snw_SD.csv')

writematrix(Sw_PD,'Sw_PD.csv')

writematrix(Sw_I,'Sw_I.csv')

writematrix(Sw_SD,'Sw_SD.csv')

writematrix(Xhat_NW_PD,'Xhat_NW_PD.csv')

writematrix(Xhat_NW_I,'Xhat_NW_I.csv')

writematrix(Xhat_NW_SD,'Xhat_NW_SD.csv')

writematrix(Xhat_W_PD,'Xhat_W_PD.csv')

writematrix(Xhat_W_I,'Xhat_W_I.csv')

writematrix(Xhat_W_SD,'Xhat_W_SD.csv')

writematrix(Pc_PD,'Pc_PD.csv')

writematrix(Pc_I,'Pc_I.csv')

writematrix(Pc_SD,'Pc_SD.csv')

writematrix(kro_PD,'kro_PD.csv')

writematrix(kro_I,'kro_I.csv')

writematrix(kro_SD,'kro_SD.csv')

writematrix(krw_PD,'krw_PD.csv')

writematrix(krw_I,'krw_I.csv')

writematrix(krw_SD,'krw_SD.csv')

% End of code %

Page 216: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

196

REFERENCES

Abder, C.B.J.: Comparison of three-phase relative permeability models and experimental results, M.S.

Thesis, The Pennsylvania State University (1981)

Akbarabadi, M., Piri, M.: Relative permeability hysteresis and capillary trapping characteristics of

supercritical CO2/brine systems: An experimental study at reservoir conditions. Adv. Water Resour. 52,

190–206 (2013). doi:10.1016/j.advwatres.2012.06.014

Akbarabadi, M., Piri, M.: Co-sequestration of SO2 with supercritical CO2 in carbonates: An experimental

study of capillary trapping, relative permeability, and capillary pressure. Adv. Water Resour. 77, 44–56

(2015). doi:10.1016/j.advwatres.2014.08.011

Akhgar, A., Khalili, B., Moa, B., Rahnama, M., Djilali, N.: Lattice-Boltzmann simulation of multi-phase

phenomena related to fuel cells. AIP Conf. Proc. 1863, (2017). doi:10.1063/1.4992192

Al-Khulaifi, Y., Lin, Q., Blunt, M.J., Bijeljic, B.: Reaction rates in chemically heterogeneous rock:

coupled impact of structure and flow properties studied by x-ray microtomography. Environ. Sci.

Technol. 51, 4108–4116 (2017). doi:10.1021/acs.est.6b06224

Alhammadi, A.M., Alratrout, A., Singh, K., Bijeljic, B., Blunt, M.J.: In situ characterization of mixed-

wettability in a reservoir rock at subsurface conditions. Sci. Rep. 7, 1–9 (2017). doi:10.1038/s41598-017-

10992-w

Allard, D.: Some Connectivity Characteristics of A Boolean model. In: Geostatistics Troia ’92.

Quantitative Geology and Geostatistics. pp. 467–478. Springer, Dordrecht (1993)

AlRatrout, A., Raeini, A.Q., Bijeljic, B., Blunt, M.J.: Automatic measurement of contact angle in pore-

space images. Adv. Water Resour. 109, 158–169 (2017). doi:10.1016/j.advwatres.2017.07.018

Alyafei, N., Blunt, M.J.: The effect of wettability on capillary trapping in carbonates. Adv. Water

Resour. 90, 36–50 (2016). doi:10.1016/j.advwatres.2016.02.001

Anderson, W.G.: Wettability literature survey- Part 1: Rock/oil/brine interactions and the effects of core

handling on wettability. J. Pet. Technol. 38, 1125–1144 (1986)(a). doi:https://doi.org/10.2118/13932-PA

Anderson, W.G.: Wettability literature survey - Part 2: Wettability measurement. JPT, J. Pet. Technol.

38, 1246–1262 (1986)(b). doi:10.2118/13933-pa

Anderson, W.G.: Wettability literature survey - Part 6: the effects of wettability on waterflooding. J. Pet.

Page 217: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

197

Technol. 39, 1605–1622 (1987). doi:https://doi.org/10.2118/16471-PA

Arganda-Carreras, I., Kaynig, V., Rueden, C., Eliceiri, K.W., Schindelin, J., Cardona, A., Seung, H.S.:

Trainable Weka segmentation: A machine learning tool for microscopy pixel classification.

Bioinformatics. 33, 2424–2426 (2017). doi:10.1093/bioinformatics/btx180

Arif, M., Abu-Khamsin, S.A., Iglauer, S.: Wettability of rock/CO 2 /brine and rock/oil/CO 2 -enriched-

brine systems:Critical parametric analysis and future outlook. Adv. Colloid Interface Sci. 268, 91–113

(2019). doi:10.1016/j.cis.2019.03.009

Armstrong, R.T., McClure, J.E., Berrill, M.A., Rücker, M., Schlüter, S., Berg, S.: Beyond Darcy’s law:

The role of phase topology and ganglion dynamics for two-fluid flow. Phys. Rev. E. 94, 1–10 (2016).

doi:10.1103/PhysRevE.94.043113

Armstrong, R.T., McClure, J.E., Robins, V., Liu, Z., Arns, C.H., Schlüter, S., Berg, S.: Porous media

Characterization using Minkowski functionals: Theories, applications and future directions. Transp.

Porous Media. (2018). doi:10.1007/s11242-018-1201-4

Avraam, D.G., Payatakes, A.C.: Flow regimes and relative permeabilities during steady-state two-phase

flow in porous media. J. Fluid Mech. 293, 207–236 (1995). doi:10.1017/S0022112095001698

Avraam, D.G., Payatakes, A.C.: Flow mechanisms, relative permeabilities, and coupling effects in

steady-state two-phase flow through porous media. The case of strong wettability. Ind. Eng. Chem. Res.

38, 778–786 (1999). doi:10.1021/ie980404o

Aydogan, D.B., Hyttinen, J.: Contour tree connectivity of binary images from algebraic graph theory. In:

20th IEEE International Conference on Image Processing (ICIP). pp. 3054–3058. IEEE, Melbourne,

Australia (2013)

Aydogan, D.B., Hyttinen, J.: Characterization of microstructures using contour tree connectivity for fluid

flow analysis. J. R. Soc. Interface. 11, 20131042–20131042 (2014). doi:10.1098/rsif.2013.1042

Bachu, S., Bennion, B.: Effects of in-situ conditions on relative permeability characteristics of CO2-brine

systems. Environ. Geol. 54, 1707–1722 (2008). doi:10.1007/s00254-007-0946-9

Berg, S., Ott, H., Klapp, S.A., Schwing, A., Neiteler, R., Brussee, N., Makurat, A., Leu, L., Enzmann, F.,

Schwarz, J.-O., Kersten, M., Irvine, S., Stampanoni, M.: Real-time 3D imaging of Haines jumps in porous

media flow. Proc. Natl. Acad. Sci. U. S. A. 110, (2013). doi:10.1073/pnas.1221373110

Berg, S., Rücker, M., Ott, H., Georgiadis, A., van der Linde, H., Enzmann, F., Kersten, M., Armstrong,

Page 218: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

198

R.T., de With, S., Becker, J., Wiegmann, A.: Connected pathway relative permeability from pore-scale

imaging of imbibition. Adv. Water Resour. 90, 24–35 (2016). doi:10.1016/j.advwatres.2016.01.010

Berg, S., Saxena, N., Shaik, M., Pradhan, C.: Generation of ground truth images to validate micro-CT

image-processing pipelines. Lead. Edge. 37, 412–420 (2018). doi:10.1190/tle37060412.1

Bernstein, R.: Adaptive nonlinear filters for simultaneous removal of different kinds of noise in images.

IEEE Trans. Circuits Syst. 34, 1275–1291 (1987). doi:10.1109/TCS.1987.1086066

Bezdek, James C; Ehrlich, Robert; Full, W.: FCM: The fuzzy c-means clustering algorithm. Comput.

Geosci. 10, 191–203 (1984)

Blunt, M.J.: Multiphase flow in permeable media : a pore-scale perspective. Cambridge University Press

(2017)

Blunt, M.J., Akai, T., Bijeljic, B.: Evaluation of methods using topology and integral geometry to assess

wettability. J. Colloid Interface Sci. 576, 99–108 (2020). doi:10.1016/j.jcis.2020.04.118

Braun, E.M., Holland, R.F.: Relative permeability hysteresis: Laboratory measurements and a conceptual

model. SPE Reserv. Eng. (Society Pet. Eng. 10, 222–228 (1995). doi:10.2118/28615-PA

Breiman, L.: Random forests. Mach. Learn. 45, 5–32 (2001).

doi:http://dx.doi.org/10.1023/A:1010933404324

Brooks, R.H., Corey, A.T.: Hydraulic properties of porous media. Hydrol. Pap. (1964)

Buades, A., Coll, B., Morel, J.: A non-local algorithm for image denoising. In: IEEE Computer Society

Conference on Computer Vision and Pattern Recognition (CVPR’05). pp. 0–5 (2005)

Bultreys, T., Lin, Q., Gao, Y., Raeini, A.Q., Alratrout, A., Bijeljic, B., Blunt, M.J.: Validation of model

predictions of pore-scale fluid distributions during two-phase flow. Phys. Rev. E. 97, (2018).

doi:10.1103/PhysRevE.97.053104

Burdine, N.T.: Relative permeability calculations from pore size distribution data. J. Pet. Technol. 5, 71–

78 (1953). doi:https://doi.org/10.2118/225-G

Carlson, F.M.: Simulation of relative permeability hysteresis to the nonwetting phase. Proc. - SPE Annu.

Tech. Conf. Exhib. 1981-Octob, (1981). doi:10.2523/10157-ms

Celauro, J.G., Torrealba, V.A., Karpyn, Z.T., Klise, K.A., Mckenna, S.A.: Pore-scale multiphase flow

experiments in bead packs of variable wettability. Geofluids. 14, 95–105 (2014). doi:10.1111/gfl.12045

Page 219: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

199

Chang, L.C., Chen, H.H., Shan, H.Y., Tsai, J.P.: Effect of connectivity and wettability on the relative

permeability of NAPLs. Environ. Geol. 56, 1437–1447 (2009). doi:10.1007/s00254-008-1238-8

Chen, L.-C., Papandreou, G., Member, S., Kokkinos, I., Murphy, K., Yuille, A.L.: Deeplab: Semantic

image segmentation with deep convolutional nets, atrous convolution, and fully connected CRFs. In:

IEEE Transactions on Pattern Analysis and Machine Intelligence. pp. 834–848 (2017)(a)

Chen, L.-C., Zhu, Y., Papandreou, G., Schroff, F., Adam, H.: Rethinking atrous convolution for semantic

image segmentation. arXiv Prepr. (2017)(b). doi:10.1159/000018039

Chen, L.-C., Zhu, Y., Papandreou, G., Schroff, F., Adam, H.: Encoder-decoder with atrous separable

convolution for semantic image segmentation. In: Proceedings of the European Conference on Computer

Vision (ECCV). pp. 801–818. , Munich, Germany (2018)

Chiquet, P., Broseta, D., Thibeau, S.: Wettability alteration of caprock minerals by carbon dioxide.

Geofluids. 7, 112–122 (2007). doi:10.1111/j.1468-8123.2007.00168.x

Çiçek, Ö., Addulkadir, A., Lienkamp, S.S., Brox, T., Ronneberger, O.: 3D U-Net: Learning dense

volumetric segmentation from sparse annotation. In: International Conference on Medical Image

Computing and Computer-assisted Intervention. pp. 424–432. , Athens, Greece (2016)

Cleveland, C., Morris, C.: Dictionary of energy. Elsevier Science & Technology (2014)

Cnudde, V., Boone, M.N.: High-resolution X-ray computed tomography in geosciences: A review of the

current technology and applications. Earth-Science Rev. 123, 1–17 (2013).

doi:10.1016/j.earscirev.2013.04.003

Coats, K.H.: Equation of State compositional model. Soc. Pet. Eng. J. 363–376 (1980).

doi:https://doi.org/10.2118/8284-PA

Corey, A.T.: The interrelation between gas and oil relative permeability, (1954)

Culligan, K.A., Wildenschild, D., Christensen, B.S.B., Gray, W.G., Rivers, M.L.: Pore-scale

characteristics of multiphase flow in porous media : A comparison of air – water and oil – water

experiments. Adv. Water Resour. 29, 227–238 (2005). doi:10.1016/j.advwatres.2005.03.021

Culligan, K.A., Wildenschild, D., Christensen, B.S.B., Gray, W.G., Rivers, M.L., Tompson, A.F.B.:

Interfacial area measurements for unsaturated flow through a porous medium. Water Resour. Res. 40, 1–

12 (2004). doi:10.1029/2004WR003278

Dalla, E., Hilpert, M., Miller, C.T.: Computation of the interfacial area for two-fluid porous medium

Page 220: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

200

systems. 56, 25–48 (2002)

Dammel, J.A., Bielicki, J.M., Pollak, M.F., Wilson, E.J.: A tale of two technologies: Hydraulic fracturing

and geologic carbon sequestration. Environ. Sci. Technol. 45, 5075–5076 (2011). doi:10.1021/es201403c

Delshad, M., Delshad, M., Pope, G.A., Lake, L.W.: Two- and three-phase relative permeabilities of

micellar fluids. SPE Form. Eval. 2, 327–337 (1987). doi:10.2118/13581-PA

Dong, H., Blunt, M.J.: Pore-network extraction from micro-computerized-tomography images. Phys.

Rev. E - Stat. Nonlinear, Soft Matter Phys. 80, 1–11 (2009). doi:10.1103/PhysRevE.80.036307

Dunmore, C.J., Wollny, G., Skinner, M.M.: MIA-clustering : a novel method for segmentation of

paleontological material. 1–18 (2018). doi:10.7717/peerj.4374

Dunn, J.C.: A fuzzy relative of the isodata process and its use in detecting compact well-separated

clusters, 3, 32–57 (1973). doi:10.1080/01969727308546046

Espinoza, D.N., Santamarina, J.C.: Water-CO2-mineral systems: Interfacial tension, contact angle, and

diffusion: Implications to CO2 geological storage. Water Resour. Res. 46, 1–10 (2010).

doi:10.1029/2009WR008634

Fatemi, S.M., Sohrabi, M., Jamiolahmady, M., Ireland, S.: Experimental and theoretical investigation of

gas/oil relative permeability hysteresis under low oil/gas interfacial tension and mixed-wet conditions.

Energy and Fuels. 26, 4366–4382 (2012). doi:10.1021/ef300386b

Fatt, I.: The Network model of porous media. Trans. AIME. 207, 144–181 (1956).

doi:https://doi.org/10.2118/574-G

Fulcher, R.A., Ertekin, T., Stahl, C.D.: Effect of capillary number and its constituents on two-phase

relative permeability curves. J. Pet. Technol. 37, 249–260 (1985). doi:https://doi.org/10.2118/12170-PA

Gao, Y., Lin, Q., Bijeljic, B., Blunt, M.J.: X-ray Microtomography of intermittency in multiphase flow at

steady state using a differential imaging method. Water Resour. Res. 53, 10274–10292 (2017).

doi:10.1002/2017WR021736

Garfi, G., John, C.M., Berg, S., Krevor, S.: The sensitivity of estimates of multiphase fluid and solid

properties of porous rocks to image processing. Transp. Porous Media. 131, 985–1005 (2020).

doi:10.1007/s11242-019-01374-z

van Genuchten, M.T.: A closed-form equation for predicting the hydraulic conductivity of unsaturated

soils. Soil Sci. Soc. Am. J. 44, 892–898 (1980).

Page 221: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

201

doi:https://doi.org/10.2136/sssaj1980.03615995004400050002x

Gerhard, J.I., Kueper, B.H.: Relative permeability characteristics necessary for simulating DNAPL

infiltration, redistribution, and immobilization in saturated porous media. Water Resour. Res. 39, (2003).

doi:10.1029/2002WR001490

Gluck, H.: Differential geometry, https://www2.math.upenn.edu/~shiydong/Math501X-7-

GaussBonnet.pdf

Haines, W.B.: Studies in the physical properties of soil. Physics (College. Park. Md). 20, (1930)

Harper, E.J.: Optimization of capillary trapping for CO2 sequestration in saline aquifers, (2013)

Hassanizadeh, S.M., Gray, W.G.: Thermodynamic basis of capillary pressure in porous media. Water

Resour. Res. 29, 3389–3405 (1993). doi:10.1029/93WR01495

He, K., Gkioxari, G., Dollar, P., Girshick, R.: Mask R-CNN. In: Proceedings of the IEEE International

Conference on Computer Vision. pp. 2961–2969. , Venice, Italy (2017)

He, K., Zhang, X., Ren, S., Sun, J.: Deep residual learning for image recognition. In: IEEE Conference

on Computer Vision and Pattern Recognition (CVPR). pp. 770–778. , La Vegas, NV (2016)

Herring, A.L., Andersson, L., Schlüter, S., Sheppard, A., Wildenschild, D.: Efficiently engineering pore-

scale processes: The role of force dominance and topology during nonwetting phase trapping in porous

media. Adv. Water Resour. 79, 91–102 (2015). doi:10.1016/j.advwatres.2015.02.005

Herring, A.L., Harper, E.J., Andersson, L., Sheppard, A., Bay, B.K., Wildenschild, D.: Effect of fluid

topology on residual nonwetting phase trapping: Implications for geologic CO2 sequestration. Adv. Water

Resour. 62, 47–58 (2013). doi:10.1016/j.advwatres.2013.09.015

Herring, A.L., Middleton, J., Walsh, R., Kingston, A., Sheppard, A.: Flow rate impacts on capillary

pressure and interface curvature of connected and disconnected fluid phases during multiphase flow in

sandstone. Adv. Water Resour. 107, 460–469 (2017). doi:10.1016/j.advwatres.2017.05.011

Herring, A.L., Sheppard, A., Andersson, L., Wildenschild, D.: Impact of wettability alteration on 3D

nonwetting phase trapping and transport. Int. J. Greenh. Gas Control. 46, 175–186 (2016).

doi:10.1016/j.ijggc.2015.12.026

Honarpour, M.., Koederitz, L.., Harvey, H..: Relative permeability of petroleum reservoirs. CRC Press

(1986)

Honarpour, M., Koederitz, L.F., Harvey, A.H.: Empirical equations for estimating two-phase relative

Page 222: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

202

permeability in consolidated rock. J. Pet. Technol. 34, 2905–2908 (1982).

doi:https://doi.org/10.2118/9966-PA

Hovadik, J.M., Larue, D.K.: Static characterizations of reservoirs: refining the concepts of connectivity

and continuity. Pet. Geosci. 13, 195–211 (2007). doi:10.1144/1354-079305-697

Huda, W., Abrahams, R.B.: Radiographic techniques, contrast, and noise in x-ray imaging. 126–131

(2015). doi:10.2214/AJR.14.13116

Iassonov, P., Gebrenegus, T., Tuller, M.: Segmentation of x-ray computed tomography images of porous

materials: A crucial step for characterization and quantitative analysis of pore structures. Water Resour.

Res. 45, (2009). doi:10.1029/2009WR008087

Iglauer, S.., Pentland, C.H.., Busch, A.: CO2 wettability of seal and reservoir rocks and the implications

for carbon geo‐sequestration. Water Resour. Res. 51, 729–774 (2015). doi:10.1002/2014WR015553

Iglauer, S., Paluszny, A., Pentland, C.H., Blunt, M.J.: Residual CO2 imaged with x-ray micro-

tomography. Geophys. Res. Lett. 38, 1–6 (2011)(a). doi:10.1029/2011GL049680

Iglauer, S., Wülling, W., Pentland, C.H., Al-Mansoori, S.K., Blunt, M.J.: Capillary-trapping capacity of

sandstones and sandpacks. SPE J. 16, 778–783 (2011)(b). doi:10.2118/120960-PA

IPCC: IPCC Special Report: Carbon Dioxide Capture and Storage. (2005)

IPCC: IPCC PRESS RELEASE, October 2018. (2018)

Jadhunandan, P.P., Morrow, N.R.: Effect of wettability on waterflood recovery for crude-oil/brine/rock

systems. SPE Reserv. Eng. 10, 40–46 (1995). doi:10.2118/22597-PA

Jerauld, G.R.: General Three-phase relative permeability model for Prudhoe Bay. SPE Reserv. Eng.

(Society Pet. Eng. 12, 255–263 (1997). doi:10.2118/36178-PA

Jerauld, G.R., Salter, S.J.: The effect of pore-structure on hysteresis in relative permeability and capillary

pressure: Pore-level modeling. Transp. Porous Media. 5, 103–151 (1990). doi:10.1007/BF00144600

Joekar-Niasar, V., Doster, F., Armstrong, R.T., Wildenschild, D., Celia, M.A.: Trapping and hysteresis in

two-phase flow in porous media: A pore-network study. Water Resour. Res. 49, 4244–4256 (2013).

doi:10.1002/wrcr.20313

Joekar-Niasar, V., Hassanizadeh, S.M., Leijnse, A.: Insights into the relationships among capillary

pressure, saturation, interfacial area, and relative permeability using pore-network modeling. Transp.

Porous Media. 74, 201–219 (2008). doi:10.1007/s11242-007-9191-7

Page 223: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

203

Joekar-Niasar, V., Prodanović, M., Wildenschild, D., Hassanizadeh, S.M.: Network model investigation

of interfacial area, capillary pressure, and saturation relationships in granular porous media. Water

Resour. Res. 46, 1–18 (2010). doi:10.1029/2009WR008585

Johns, R.T., Orr, F.M.: Miscible gas displacement of multicomponent oils. SPEJ. 1, 39–50 (1996).

doi:https://doi.org/10.2118/30798-PA

Juanes, R., Spiteri, E.J., Orr, F.M., Blunt, M.J.: Impact of relative permeability hysteresis on geological

CO2 storage. Water Resour. Res. 42, 1–13 (2006). doi:10.1029/2005WR004806

Karimpouli, S., Tahmasebi, P.: Segmentation of digital rock images using deep convolutional

autoencoder networks. Comput. Geosci. 126, 142–150 (2019). doi:10.1016/j.cageo.2019.02.003

Karpyn, Z.T., Piri, M., Singh, G.: Experimental investigation of trapped oil clusters in a water-wet bead

pack using X-ray microtomography. Water Resour. Res. 46, 1–25 (2010). doi:10.1029/2008WR007539

Khishvand, M., Alizadeh, A.H., Piri, M.: In-situ characterization of wettability and pore-scale

displacements during two- and three-phase flow in natural porous media. Adv. Water Resour. 97, 279–

298 (2016). doi:10.1016/j.advwatres.2016.10.009

Khorsandi, S., Li, L., Johns, R.T.: Equation of state for relative permeability, including hysteresis and

wettability Alteration. SPE J. 22, 1915–1928 (2017). doi:10.2118/182655-PA

Khorsandi, S., Li, L., Johns, R.T.: A new way of compositional simulation without phase labeling. In:

SPE Improved Oil Recovery Conference. Tulsa, Oklahoma, USA (2018)

Khorsandi, S., Li, L., Johns, R.T.: A new way of compositional simulation without phase labeling. SPEJ.

1–19 (2021). doi:https://doi.org/10.2118/190269-PA

Killough, J.E.: Reservoir simulation with history-dependent saturation functions. Soc Pet Eng AIME J.

16, 37–48 (1976). doi:10.2118/5106-pa

Kim, Y., Wan, J., Kneafsey, T.J., Tokunaga, T.K.: Dewetting of silica surfaces upon reactions with

supercritical CO2 and brine: Pore-scale studies in micromodels. Environ. Sci. Technol. 46, 4228–4235

(2012). doi:10.1021/es204096w

Klise, K.A., Moriarty, D., Yoon, H., Karpyn, Z.: Automated contact angle estimation for three-

dimensional x-ray microtomography data. Adv. Water Resour. 95, 152–160 (2016).

doi:10.1016/j.advwatres.2015.11.006

Kotsiantis, S.B.: Supervised Machine Learning: A review of classification techniques. Informatica. 31,

Page 224: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

204

249–269 (2007)

Krevor, S., Blunt, M.J., Benson, S.M., Pentland, C.H., Reynolds, C., Al-Menhali, A., Niu, B.: Capillary

trapping for geologic carbon dioxide storage - From pore scale physics to field scale implications. Int. J.

Greenh. Gas Control. 40, 221–237 (2015). doi:10.1016/j.ijggc.2015.04.006

Krevor, S.C.M., Pini, R., Li, B., Benson, S.M.: Capillary heterogeneity trapping of CO2 in a sandstone

rock at reservoir conditions. Geophys. Res. Lett. 38, 1–5 (2011). doi:10.1029/2011GL048239

Krevor, S.C.M., Pini, R., Zuo, L., Benson, S.M.: Relative permeability and trapping of CO2 and water in

sandstone rocks at reservoir conditions. Water Resour. Res. 48, 1–16 (2012).

doi:10.1029/2011WR010859

Kulkarni, R., Tuller, M., Fink, W., Wildenschild, D.: Three-dimensional multiphase segmentation of x-

ray CT data of porous materials using a Bayesian Markov random field framework. Vadose Zo. J. 11,

(2012). doi:10.2136/vzj2011.0082

Lai, P., Moulton, K., Krevor, S.: Pore-scale heterogeneity in the mineral distribution and reactive surface

area of porous rocks. Chem. Geol. 411, 260–273 (2015). doi:10.1016/j.chemgeo.2015.07.010

Lake, L.W., Johns, R.T., Rossen, W.R., Gary, P.A.: Fundamentals of enhanced oil recovery. Society of

Petroleum Engineers (2014)

Lamanna, J.M., Bothe, J. V., Zhang, F.Y., Mench, M.M.: Measurement of capillary pressure in fuel cell

diffusion media, micro-porous layers, catalyst layers, and interfaces. J. Power Sources. 271, 180–186

(2014). doi:10.1016/j.jpowsour.2014.07.163

Land, C.S.: Calculation of imbibition relative permeability for two- and three-phase flow from rock

properties. Soc. Pet. Eng. J. 149–156 (1968). doi:10.2118/1942-PA

Land, C.S.: Comparison of calculated with experimental imbibition relative permeability. Soc Pet. Eng J.

11, 419–425 (1971). doi:10.2118/3360-pa

Landry, C.J., Karpyn, Z.T., Ayala, O.: Relative permeability of homogenous-wet and mixed-wet porous

media as determined by pore-scale lattice Boltzmann modeling. Water. 50, 3672–3689 (2014).

doi:10.1002/2013WR015148.Received

Landry, C.J., Karpyn, Z.T., Piri, M.: Pore-scale analysis of trapped immiscible fluid structures and fluid

interfacial areas in oil-wet and water-wet bead packs. Geofluids. 11, 209–227 (2011). doi:10.1111/j.1468-

8123.2011.00333.x

Page 225: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

205

Lenormand, R., Zarcone, C.: Role of roughness and edges during imbibition in square capillaries. In:

SPE, AIME. , Houston, TX (1984)

Levine, J.S., Goldberg, D.S., Lackner, K.S., Matter, J.M., Supp, M.G., Ramakrishnan, T.S.: Relative

permeability experiments of carbon dioxide displacing brine and their implications for carbon

sequestration. Environ. Sci. Technol. 48, 811–818 (2014). doi:10.1021/es401549e

Li, L., Johns, R.T.: Coupled capillary pressure and relative permeability using an equation-of-state

approach. In: SPE Improved Oil Recovery Conference. , Tulsa, OK, USA (2018)

Li, X., Akbarabadi, M., Karpyn, Z.T., Piri, M., Bazilevskaya, E.: Experimental investigation of carbon

dioxide trapping due to capillary retention in saline aquifers. Geofluids. 15, 563–576 (2015).

doi:10.1111/gfl.12127

Lin, Q., Bijeljic, B., Berg, S., Pini, R., Blunt, M.J., Krevor, S.: Minimal surfaces in porous media: Pore-

scale imaging of multiphase flow in an altered-wettability Bentheimer sandstone. Phys. Rev. E. 99, 1–13

(2019). doi:10.1103/PhysRevE.99.063105

Lin, Q., Bijeljic, B., Pini, R., Blunt, M.J., Krevor, S.: Imaging and measurement of pore-scale interfacial

curvature to determine capillary pressure simultaneously with relative permeability. Water Resour. Res.

54, 7046–7060 (2018). doi:10.1029/2018WR023214

Liu, Z., McClure, J.E., Armstrong, R.T.: Influence of wettability on phase connectivity and electrical

resistivity. Phys. Rev. E. 98, (2018). doi:10.1103/PhysRevE.98.043102

Ma, T.D., Youngren, G.K.: Performance of immiscible water-alternating-gas (IWAG) injection at

Kuparuk river unit, North Slope, Alaska. In: Proceedings - SPE Annual Technical Conference and

Exhibition. pp. 415–423. , New Orleans, Lousiana, USA (1994)

Majid Hassanizadeh, S., Gray, W.G.: Toward an improved description of the physics of two-phase flow.

Adv. Water Resour. 16, 53–67 (1993). doi:10.1016/0309-1708(93)90029-F

Mandelbrot, B.B.: The fractal geometry of nature. W. H. Freeman and Co. (1982)

Mcclure, J.E., Armstrong, R.T., Berrill, M.A., Berg, S., Gray, W.G., Miller, C.T.: Geometric state

function for two-fluid flow in porous media. Phys. Rev. Fluids. 3, 1–16 (2018).

doi:10.1103/PhysRevFluids.3.084306

Mcclure, J.E., Berrill, M.A., Gray, W.G., Miller, C.T.: Influence of phase connectivity on the relationship

among capillary pressure, fluid saturation, and interfacial area in two-fluid-phase porous medium

Page 226: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

206

systems. 033102, 1–12 (2016). doi:10.1103/PhysRevE.94.033102

Menke, H.P., Bijeljic, B., Blunt, M.J.: Dynamic reservoir-condition microtomography of reactive

transport in complex carbonates: Effect of initial pore structure and initial brine pH. Geochim.

Cosmochim. Acta. 204, 267–285 (2017). doi:10.1016/j.gca.2017.01.053

Milletari, F., Navab, N., Ahmadi, S.: V-Net : Fully convolutional neural networks for volumetric medical

image segmentation. In: Fourth International Conference on 3D Vision (3DV). pp. 565–571. , Stanford,

California, USA (2016)

Mohanty, K.K., Davis, H.T., Scriven, L.E.: Physics of oil entrapment in water-wet rock. SPE Reserv.

Eng. 2, (1987). doi:10.2118/9406-PA

Morrow, N.R.: The effects of surface roughness on contact angle with special reference to petroleum

recovery. J. Can. Pet. 14, (1975). doi:https://doi.org/10.2118/75-04-04

Naar, J., Henderson, J.H.: An imbibition model - its application to flow behavior and the prediction of oil

recovery. Soc. Pet. Eng. J. 1, 61–70 (1961). doi:https://doi.org/10.2118/1550-G

Naar, J., Wygal, R.J.: Three-phase imbibition relative permeability. Soc. Pet. Eng. J. 1, 254–258 (1961).

doi:10.2118/90-pa

Nadafpour, M., Rasaei, M.R.: Investigating drainage rate effects on fractal patterns and capillary

fingering in a realistic glass micromodel. Teh. Vjesn. 21, 1263–1271 (2014)

Neshat, S.S., Pope, G.A.: Three-phase relative permeability and capillary pressure models with hysteresis

and compositional consistency. SPE J. 23, 2394–2408 (2018). doi:10.2118/191384-PA

Ni, H., Boon, M., Garing, C., Benson, S.M.: Predicting CO2 residual trapping ability based on

experimental petrophysical properties for different sandstone types. Int. J. Greenh. Gas Control. 86, 158–

176 (2019). doi:10.1016/j.ijggc.2019.04.024

Niu, Y., Mostaghimi, P., Shabaninejad, M., Swietojanski, P., Armstrong, R.T.: Digital rock segmentation

for petrophysical analysis with reduced user bias using convolutional neural networks. Water Resour.

Res. 56, 1–11 (2020). doi:10.1029/2019WR026597

Noiriel, C., Gouze, P., Bernard, D.: Investigation of porosity and permeability effects from

microstructure changes during limestone dissolution. Geophys. Res. Lett. 31, 1–4 (2004).

doi:10.1029/2004GL021572

Oh, W., Lindquist, W.B.: Image thresholding by indicator kriging. IEEE Trans. Pattern Anal. Mach.

Page 227: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

207

Intell. 21, 590–602 (1999). doi:10.1109/34.777370

Oktay, O., Schlemper, J., Folgoc, L. Le, Lee, M., Heinrich, M., Misawa, K., Mori, K., McDonagh, S.,

Hammerla, N.Y., Kainz, B., Glocker, B., Rueckert, D.: Attention U-Net: Learning where to look for the

pancreas. arXiv Prepr. (2018)

Øren, P.E., Bakke, S.: Reconstruction of berea sandstone and pore-scale modelling of wettability effects.

J. Pet. Sci. Eng. 39, 177–199 (2003). doi:10.1016/S0920-4105(03)00062-7

Øren, P.E., Bakke, S., Arntzen, O.J.: Extending predictive capabilities to network models. SPE J. 3, 324–

335 (1998). doi:10.2118/52052-pa

Orr, F.M.: Onshore geologic storage of CO2. Science (80-. ). 325, 1656–1658 (2009).

doi:10.1126/science.1175677

Osborne, G.A.: Differential and integral calculus: with examples and applications. D. C. Heath & Co.

(1908)

Osei-Bonsu, K., Khorsandi, S., Piri, M.: Quantitative analysis of phase topology evolution during three-

phase displacements in porous media. Lab Chip. 20, 2495–2509 (2020).

doi:https://doi.org/10.1039/D0LC00232A

Pak, T., Butler, I.B., Geiger, S., van Dijke, M.I.J., Sorbie, K.S.: Droplet fragmentation: 3D imaging of a

previously unidentified pore-scale process during multiphase flow in porous media. Proc. Natl. Acad. Sci.

112, 1947–1952 (2015). doi:10.1073/pnas.1420202112

Parker, J.C.: Multiphase flow and transport in porous media. Rev. Geophys. 3, 311–328 (1989)

Patzek, T.W.: Verification of a complete pore network simulator of drainage and imbibition. SPE J. 6,

251 (2001). doi:10.2118/59312-ms

Pentland, C.H.., Itsekiri, E.., Al Mansoori, S.K.., Iglauer, S.., Bijeljic, B.., Blunt, M.J.: Measurement of

nonwetting-phase trapping in sandpacks. SPEJ. 115, 163–171 (2010). doi:https://doi.org/10.2118/115697-

PA

Pentland, C.H., El-Maghraby, R., Iglauer, S., Blunt, M.J.: Measurements of the capillary trapping of

super-critical carbon dioxide in Berea sandstone. Geophys. Res. Lett. 38, 1–4 (2011).

doi:10.1029/2011GL046683

Perrin, J.C., Benson, S.: An experimental study on the influence of sub-core scale heterogeneities on CO2

distribution in reservoir rocks. Transp. Porous Media. 82, 93–109 (2010). doi:10.1007/s11242-009-9426-

Page 228: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

208

x

Pham, D.L., Prince, J.L.: Adaptive fuzzy segmentation of magnetic resonance images. IEEE Trans. Med.

Imaging. 18, 737–752 (1999)

Porter, M.L., Wildenschild, D.: Image analysis algorithms for estimating porous media multiphase flow

variables from computed microtomography data: A validation study. Comput. Geosci. 14, 15–30 (2010).

doi:10.1007/s10596-009-9130-5

Prodanovic, M., Esteva, M., Hanlon, M., Nanda, G., Agarwal, P.: Digital rocks portal: a repository for

porous media images, https://www.digitalrocksportal.org/

Purcell, W.R.: Capillary pressures - their measurement using mercury and the calculation of permeability

therefrom. J. Pet. Technol. 1, 39–48 (1949). doi:https://doi.org/10.2118/949039-G

Purswani, P., Johns, R.T., Karpyn, Z.T., Blunt, M.: Predictive modeling of relative permeability using a

generalized equation of state. SPE J. (2020). doi:10.2118/200410-PA

Purswani, P., Karpyn, Z.T.: Laboratory investigation of chemical mechanisms driving oil recovery from

oil-wet carbonate rocks. Fuel. 235, 406–415 (2019). doi:10.1016/j.fuel.2018.07.078

Purswani, P., Tawfik, M.S., Karpyn, Z.: Factors and mechanisms governing wettability alteration by

chemically tuned waterflooding: A Review. Energy & Fuels. (2017).

doi:10.1021/acs.energyfuels.7b01067

Purswani, P., Tawfik, M.S., Karpyn, Z.T., Johns, R.T.: On the development of a relative permeability

equation of state. Comput. Geosci. (2019). doi:10.1007/s10596-019-9824-2

Raeini, A.Q., Bijeljic, B., Blunt, M.J.: Generalized network modeling: Network extraction as a coarse-

scale discretization of the void space of porous media. Phys. Rev. E. 96, 1–17 (2017).

doi:10.1103/PhysRevE.96.013312

Raeini, A.Q., Bijeljic, B., Blunt, M.J.: Generalized network modeling of capillary-dominated two-phase

flow. Phys. Rev. E. 97, (2018). doi:10.1103/PhysRevE.97.023308

Reeves, P.C., Celia, A.: A functional relationship between capillary pressure, saturation, and interfacial

area as revealed by a pore-scale network model. Water Resour. Res. 32, 2345–2358 (1996).

doi:10.1029/96WR01105

Richeson, D.S.: Euler’s Gem: The polyhedron formula and the birth of topology. Princeton University

Press (2008)

Page 229: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

209

Ronneberger, O., Fischer, P., Brox, T.: U-Net: convolutional networks for biomedical image

segmentation. In: International conference on medical image computing and computer-assisted

intervention. pp. 234–241., Munich, Germany (2015)

Roof, J.G.: Snap-off of oil droplets in water-wet pores. Soc. Pet. Eng. J. 10, 85–90 (1970).

doi:10.2118/2504-PA

Rücker, M., Bartels, W.B., Singh, K., Brussee, N., Coorn, A., van der Linde, H.A., Bonnin, A., Ott, H.,

Hassanizadeh, S.M., Blunt, M.J., Mahani, H., Georgiadis, A., Berg, S.: The effect of mixed wettability on

pore-scale flow regimes based on a flooding experiment in ketton limestone. Geophys. Res. Lett. 46,

3225–3234 (2019). doi:10.1029/2018GL081784

Rücker, M., Berg, S., Armstrong, R.T., Georgiadis, A., Ott, H., Schwing, A., Neiteler, R., Brussee, N.,

Makurat, A., Leu, L., Wolf, M., Khan, F., Enzmann, F., Kersten, M.: From connected pathway flow to

ganglion dynamics. Geophys. Res. Lett. 42, 3888–3894 (2015). doi:10.1002/2015GL064007

Sahimi, M., Imdakm, A.O.: The effect of morphological disorder on hydrodynamic dispersion in flow

through porous media. J. Phys. A. Math. Gen. 21, 3833–3870 (1988). doi:10.1088/0305-4470/21/19/019

Salathiel, R.A.: Oil recovery by surface film drainage in mixed-wettability rocks. J. Pet. Technol. 25,

1216–1224 (1973). doi:10.2118/4104-PA

Sandler, S.I.: Chemical and engineering thermodynamics. John Wiley and Sons (1989)

Saraji, S., Goual, L., Piri, M., Plancher, H.: Wettability of supercritical carbon dioxide/water/quartz

systems: Simultaneous measurement of contact angle and interfacial tension at reservoir conditions.

Langmuir. 29, 6856–6866 (2013). doi:10.1021/la3050863

Saripalli, K.P., Annable, M.D., Rao, P.S.C.: Estimation of nonaqueous phase liquid - water interfacial

areas in porous media following mobilization by chemical flooding. 31, 3384–3388 (1997)(a).

doi:10.1021/es961082n

Saripalli, K.P., Kim, H., Rao, P.S.C., Annable, M.D.: Measurement of specific fluid/fluid interfacial

areas of immiscible fluids in poreus media. Environ. Sci. Technol. 31, 932–936 (1997)(b).

doi:10.1021/es960652g

Scanziani, A., Singh, K., Blunt, M.J., Guadagnini, A.: Automatic method for estimation of in situ

effective contact angle from X-ray micro tomography images of two-phase flow in porous media. J.

Colloid Interface Sci. 496, 51–59 (2017). doi:10.1016/j.jcis.2017.02.005

Page 230: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

210

Schlüter, S., Berg, S., Rücker, M., Armstrong, R.T., Vogel, H.J., Hilfer, R., Wildenschild, D.: Pore-scale

displacement mechanisms as a source of hysteresis for two-phase flow in porous media. Water Resour.

Res. 52, 2194–2205 (2016). doi:10.1002/2015WR018254

Schluter, S., Sheppard, A., Brown, K., Wildenschild, D.: Image processing of multiphase images

obtained via x-ray microtomography: A review. Water Resour. Res. 50, 3615–3639 (2014).

doi:10.1002/2014WR015256.

Schrag, D.P.: Storage of carbon dioxide in offshore sediments. Science. 5948. 325, 1658–1659 (2009).

doi:10.1126/science.1175750

Sheppard, A.P., Sok, R.M., Averdunk, H.: Techniques for image enhancement and segmentation of

tomographic images of porous materials. Phys. A Stat. Mech. its Appl. 339, 145–151 (2004).

doi:10.1016/j.physa.2004.03.057

Singh, K., Menke, H., Andrew, M., Lin, Q., Rau, C., Blunt, M.J., Bijeljic, B.: Dynamics of snap-off and

pore-filling events during two-phase fluid flow in permeable media. Sci. Rep. 7, 5192 (2017).

doi:10.1038/s41598-017-05204-4

Skauge, A., Sorbie, K.S., VanDijke, M.I.J.: Modelling the transition between immiscible and miscible

WAG. In: IOR 2019 – 20th European Symposium on Improved Oil Recovery. pp. 1–12. European

Association of Geoscientists & Engineers, Pau, France (2019)

Spiteri, E.J., Juanes, R.: Impact of relative permeability hysteresis on the numerical simulation of WAG

injection. J. Pet. Sci. Eng. 50, 115–139 (2006). doi:10.1016/j.petrol.2005.09.004

Spiteri, E.J., Juanes, R., Blunt, M.J., Orr, F.M.: A new model of trapping and relative permeability

hysteresis for all wettability characteristics. SPE J. 13, 277–288 (2008). doi:10.2118/96448-PA

Stone, H.L.: Probability model for estimating three-phase relative perveability. JPT, J. Pet. Technol. 22,

214–218 (1970). doi:10.2118/2116-pa

Stone, H.L.: Estimation of three-phase relative permeability and residual oil data. J. Can. Pet. Technol.

12, 53–61 (1973). doi:10.2118/73-04-06

Sun, C., McClure, J.E., Mostaghimi, P., Herring, A.L., Shabaninejad, M., Berg, S., Armstrong, R.T.:

Linking continuum-scale state of wetting to pore-scale contact angles in porous media. 1–8 (2019)

Szulczewski, M.L., MacMinn, C.W., Herzog, H.J., Juanes, R.: Lifetime of carbon capture and storage as

a climate-change mitigation technology. Proc. Natl. Acad. Sci. U. S. A. 109, 5185–5189 (2012).

Page 231: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

211

doi:10.1073/pnas.1115347109

Tanino, Y., Blunt, M.J.: Laboratory investigation of capillary trapping under mixed-wet conditions.

Water Resour. Res. 49, 4311–4319 (2013). doi:10.1002/wrcr.20344

Treiber, L.E., Archer, D.L., Owens, W.W.: A laboratory evaluation of the wettability of fifty oil-

producing reservoirs. Soc. Pet. Eng. J. 6, 531–540 (1972). doi:10.2118/3526-PA

Valvatne, P.H., Blunt, M.J.: Predictive pore-scale modeling of two-phase flow in mixed wet media.

Water Resour. Res. 40, 1–21 (2004). doi:10.1029/2003WR002627

Vincent, L., Soille, P.: Watersheds in Digital Spaces: An efficient algorithm based on immersion

simulations. IEEE Trans. Pattern Anal. Mach. Intell. 13, 583–598 (1991)

Vogel, H.-J.: Topological characterization of porous media. Morphol. Condens. Matter, Ed. by K.

Mecke, D. Stoyan, Lect. Notes Phys. 600, 75–92 (2002). doi:10.1007/3-540-45782-8_3

Wang, Y. Da, Shabaninejad, M., Armstrong, R.T., Mostaghimi, P.: Physical accuracy of deep neural

networks for 2D and 3D multi-mineral segmentation of rock micro-CT images. arXiv Prepr. (2020)

Watkins, T.: The Gauss-Bonnet theorem and its generalization,

https://www.sjsu.edu/faculty/watkins/gaussbonet.htm

Wildenschild, D., Sheppard, A.P.: X-ray imaging and analysis techniques for quantifying pore-scale

structure and processes in subsurface porous medium systems. Adv. Water Resour. 51, 217–246 (2013).

doi:10.1016/j.advwatres.2012.07.018

Wollny, G., Kellman, P., Skinner, M.M.: Open access MIA - a free and open source software for gray

scale medical image analysis. (2013). doi:10.1186/1751-0473-8-20

Xue, Y., Tang, H., Qiao, Z., Gong, G., Yin, Y., Qian, Z., Huang, C., Fan, W., Huang, X.: Shape-aware

organ segmentation by predicting signed distance maps. arXiv Prepr. (2019)

Xue, Y., Xu, T., Long, L.R., Huang, X.: SegAN: Adversarial network with multi-scale L1 loss for

medical image segmentation. Neuroinformatics. 16, 383–392 (2018). doi:10.1016/B978-012264841-

0/50037-8

Yekeen, N., Padmanabhan, E., Sevoo, T.A., Kanesen, K.A., Okunade, O.A.: Wettability of

rock/CO2/brine systems: A critical review of influencing parameters and recent advances. J. Ind. Eng.

Chem. 88, 1–28 (2020). doi:10.1016/j.jiec.2020.03.021

Yuan, C., Pope, G.A.: A new method to model relative permeability in compositional simulators to avoid

Page 232: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

212

discontinuous changes caused by phase-identification problems. SPE J. 17, 1221–1230 (2012).

doi:10.2118/142093-PA

Zhang, P., Austad, T.: Wettability and oil recovery from carbonates : Effects of temperature and potential

determining ions. 279, 179–187 (2006). doi:10.1016/j.colsurfa.2006.01.009

Zhao, B., Ratnakar, R., Dindoruk, B., Mohanty, K.: A Hybrid approach for the prediction of relative

permeability using machine learning of experimental and numerical SCAL data. 1–24 (2019).

doi:10.2118/196022-ms

Zhou, Z., Rahman Siddiquee, M.M., Tajbakhsh, N., Liang, J.: Unet++: A Nested U-Net architecture for

medical image segmentation. In: Deep Learning in Medical Image Analysis and Multimodal Learning for

Clinical Decision Support. pp. 3–11. , Québec City, QC, Canada (2018)

Page 233: RELATIVE PERMEABILITY EQUATION-OF-STATE: THE ROLE OF …

Curriculum Vita

Prakash Purswani

Education

Ph.D., Energy and Mineral Engineering: Petroleum & Natural Gas Engineering, 2021

Department of Energy and Mineral Engineering, The Pennsylvania State University, University Park, USA

M.S., Energy and Mineral Engineering: Petroleum & Natural Gas Engineering, 2017

Department of Energy and Mineral Engineering, The Pennsylvania State University, University Park, USA

B.E., Chemical Engineering, 2015

Department of Chemical Engineering, BITS-Pilani, Hyderabad, India

Work experience

- Petroleum Engineering Intern, Chevron Corporation (Summer, 2020)

- Course Instructor (PNGE 405), The Pennsylvania State University (Fall, 2019)

- Petroleum Engineering Intern, Chevron Corporation (Summer, 2019)

- Study Abroad Research Scholar, Imperial College London (Spring, 2019)

Awards

- Graduate Student Award, EME Department, Penn State (April, 2019)

- Holleran and Bowman Academic Excellence Award, EME Department, Penn State (February, 2019)

Contact: [email protected]

Google Scholar Purswani Prakash

LinkedIn Purswani Prakash