copyright by xiangyu liu 2017

171
Copyright by Xiangyu Liu 2017

Upload: others

Post on 21-Nov-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Copyright by Xiangyu Liu 2017

Copyright

by

Xiangyu Liu

2017

Page 2: Copyright by Xiangyu Liu 2017

The Dissertation Committee for Xiangyu Liu Certifies that this is the approved

version of the following dissertation:

Mud-to-Cement Conversion of Synthetic-Based Drilling Muds using

Geopolymers

Committee:

Eric van Oort, Supervisor

Paul M. Bommer

Hugh C. Daigle

David N. Espinoza

Maria G. Juenger

Sriramya D. Nair

Page 3: Copyright by Xiangyu Liu 2017

Mud-to-Cement Conversion of Synthetic-Based Drilling Muds using

Geopolymers

by

Xiangyu Liu

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin

August 2017

Page 4: Copyright by Xiangyu Liu 2017

Dedication

Dedicated to my wonderful parents, Haimi Liu and Min Wang,

and my dear husband, Jun Lu,

for their endless love, support and encouragement.

Page 5: Copyright by Xiangyu Liu 2017

v

Acknowledgements

First of all I would like to thank my supervisor, Prof. Eric van Oort for the

opportunity to work on this project and for introducing me to the world of well

cementing. Thank you for your invaluable guidance, support and encouragement as I

pursue my degree. Your invaluable suggestions and guidance has inspired and motivated

me to keep exploring research ideas with both fundamental and practical values.

I would like to extend my gratitude to my committee members: Dr. Paul Bommer,

Dr. Hugh Daigle, Dr. Maria Juenger, Dr. Nicolas Espinoza and Dr. Sriramya Nair. Thank

you for your friendly interactions, and for the insightful suggestions and comments on my

work.

To the lab ladies, Sriramya Nair, Qian Wu, Katherine Aughenbaugh, Michelle

Shuck and Hanna Lee - I would not have finished without your support. You all made the

lab full of fun and encouragement. I truly enjoyed the memorable and productive time we

have spent together. It would be a difficult journey without all your support.

I thank Matthew Ramos for helping me run the triaxial compressive strength test

and Besmir Bez Buranaj and Bence Tóth for running the pressure transmission test. The

results were very important for characterizing the materials. I would also like to thank Dr.

Raissa Ferron from the Civil, Architechtural, and Environmental Engineering Department

for allowing me the use of the particle size analyzer in her laboratory.

Special thanks go to Tesse Smitherman, Frankie Hart, Amy Stewart, Glen Baum,

Gary Miscoe and Daryl Nygaard for their technical and administrative support.

Page 6: Copyright by Xiangyu Liu 2017

vi

I would like to acknowledge the contribution of undergraduate assistants who

have aided in the experimental work and data analysis. A special recognition for Hanna

Lee, Marjorie Dininger, and Mateo Valencia.

I would like to thank ConocoPhillips for providing research funding towards this

project. Thank you to PQ Corporation and SEFA Group for providing materials and

technical support.

Finally, my heartfelt gratitude goes to my parents, Haimi Liu and Min Wang, and

my brother, Xiangyi Liu, for their unconditional love and support. To my beloved

husband, Jun Lu, thank you for being so patient, caring and supportive throughout this

process, and for making my graduate life so memorable. You have been - and will always

be - a source of strength and inspiration for me.

Page 7: Copyright by Xiangyu Liu 2017

vii

Mud-to-Cement Conversion of Synthetic-Based Drilling Muds using

Geopolymers

Xiangyu Liu, Ph.D.

The University of Texas at Austin, 2017

Supervisor: Eric van Oort

When constructing wells ranging from simple land wells to complex deepwater

wells, incompatibility between oil-based and synthetic-based muds (OBM / SBM) and

Portland cements can lead to poor cementation and loss of cement integrity, which in turn

may compromise zonal isolation. An alternative cementitious material based on

geopolymers has been developed with improved OBM / SBM compatibility for primary

cementing and lost circulation control as well as well abandonment. Benefits of using

geopolymers go beyond mere OBM / SBM compatibility: it is in fact possible to solidify

non-aqueous drilling fluids (NAF) such as SBM and OBM using geopolymer

formulations. This also means that such NAFs can be disposed of in a more cost-effective

way, which presents a viable option for environmentally acceptable on-site or off-site

disposal of drilling muds and cuttings. In the following, focus will be primarily on the

compatibility between SBM and geopolymers, with the understanding that the results

obtained for SBM can generally be extrapolated to OBM as well.

Geopolymer is a type of alkali-activated material that forms when an

aluminosilicate precursor powder (such as fly ash) is mixed with an alkaline-activating

solution (such as sodium hydroxide). A novel SBM solidification method was developed

Page 8: Copyright by Xiangyu Liu 2017

viii

by blending varied amounts of geopolymer and SBM. The consolidated mud was named

a “geopolymer hybrid cement”.

In an effort to develop the geopolymer hybrid system as a novel well cementing

material, the solidification method was comprehensively studied with various sources of

precursor powders, activators, as well as SBM and OBM formulations. Fresh state

properties, such as slurry rheology and thickening time, and hardened state mechanical

properties, such as compressive strength (under both uniaxial and triaxial confinement

conditions), as well as the self-healing capabilities of the geopolymer hybrid cement were

evaluated.

Strength testing results showed that geopolymer cement can solidify up to a 60/40

geopolymer/SBM ratio by volume. The incorporation of SBM greatly improved the

rheological properties of the geopolymer hybrid, allowing for the otherwise non-

pumpable slurry to become pumpable for well cementation and lost circulation control

purposes. The laboratory evaluations showed that the geopolymer hybrid cement could

meet typical requirements as a well cementing slurry. By changing the amount of

geopolymer and SBM in the slurry, the geopolymer hybrid can be deliberately designed

with high compressive strength for primary cementation, or with lower compressive

strength for lost circulation control. Moreover, geopolymer and geopolymer hybrid

cements reveal true self-healing capability, which means that they can recover and even

increase their strength after prior yielding. This ability would possibly allow such

cements to better adapt to subsurface stress changes acting on abandoned wells, making

them better suited for use in permanent barriers in plug and abandonment operations.

Page 9: Copyright by Xiangyu Liu 2017

ix

Table of Contents

List of Figures ...................................................................................................... xiii

List of Tables .........................................................................................................xx

Chapter 1: Introduction ...................................................................................1

Motivation .................................................................................................1 1.1

Objectives .................................................................................................3 1.2

Dissertation Organization .........................................................................4 1.3

Chapter 2: Background ...................................................................................6

Geopolymer: Definition and Terminology ...............................................6 2.1

Synthesis of Alkali-Activated Fly Ash-based Geopolymer ......................7 2.2

Aluminosilicate source..................................................................7 2.2.1

Activating solution ......................................................................11 2.2.2

Hydroxide Activation..................................................................11

Silicate Activation .......................................................................12

Proportioning ..............................................................................12 2.2.3

Curing regime .............................................................................13 2.2.4

Polymerization Mechanism and Microstructure .....................................14 2.3

Properties and Admixture .......................................................................16 2.4

Compressive strength ..................................................................17 2.4.1

Rheological properties ................................................................17 2.4.2

Setting time control .....................................................................18 2.4.3

Self-healing capability ................................................................20 2.4.4

Other properties ..........................................................................21 2.4.5

Geopolymer in Oilwell Cementing .........................................................22 2.5

Primary cementation ...................................................................22 2.5.1

Well abandonment ......................................................................26 2.5.2

Lost circulation control ...............................................................29 2.5.3

Mud Solidification Technique ................................................................31 2.6

Page 10: Copyright by Xiangyu Liu 2017

x

BFS-based drilling mud solidification ........................................31 2.6.1

Geopolymer-based solidification for organic waste disposal .....33 2.6.2

Summary .................................................................................................34 2.7

Chapter 3: Materials and Methods ................................................................35

Raw Materials .........................................................................................35 3.1

Fly ash composition ....................................................................35 3.1.1

Particle size distribution (PSD) of fly ashes ...............................36 3.1.2

Activator .....................................................................................37 3.1.3

Admixtures ..................................................................................39 3.1.4

Portland cement ..........................................................................40 3.1.5

Alkali-activated slag ...................................................................40 3.1.6

SBM / OBM ................................................................................41 3.1.7

Methods...................................................................................................43 3.2

Compressive strength ..................................................................43 3.2.1

Ultrasonic compressive strength .................................................44 3.2.2

Rheology .....................................................................................44 3.2.3

Thickening time ..........................................................................45 3.2.4

Confined compressive strength ...................................................45 3.2.5

Unconfined self-healing test .......................................................47 3.2.6

Confined self-healing test ...........................................................47 3.2.7

Pipe-in-pipe shear bond strength test ..........................................48 3.2.8

Pressure transmission test ...........................................................49 3.2.9

Porosity and pore size distribution ............................................53 3.2.10

Chapter 4: Hydroxide Activation..................................................................55

Contamination Resistance .......................................................................55 4.1

Properties of Geopolymer Hybrid Cement .............................................61 4.2

Compressive Strength .................................................................61 4.2.1

Downhole and Surface Rheology ...............................................62 4.2.2

Thickening Time .........................................................................64 4.2.3

Validation ................................................................................................65 4.3

Page 11: Copyright by Xiangyu Liu 2017

xi

Effect of Changing Activator Molarity .......................................65 4.3.1

Effect of Changing Aluminosilicate Source ...............................67 4.3.2

Effect of Seawater .......................................................................70 4.3.3

Effect of SBM Composition .......................................................72 4.3.4

Effect of Pressure ........................................................................76 4.3.5

Stability Control ......................................................................................78 4.4

Solidification of Non-Aqueous Drilling Muds .......................................82 4.5

Summary .................................................................................................84 4.6

Chapter 5: Silicate Activation .......................................................................86

Rheological Properties ............................................................................86 5.1

Compressive Strength .............................................................................93 5.2

Thickening Time .....................................................................................95 5.3

Summary .................................................................................................99 5.4

Chapter 6: Mechanical Properties and Self-Healing Capability .................100

Confined Compressive Strength ...........................................................100 6.1

Mechanical Properties ...........................................................................103 6.2

Self-Healing Properties .........................................................................105 6.3

Cement-to-Pipe Bond Strength .............................................................113 6.4

Hydraulic Conductivity .........................................................................116 6.5

Porosity .................................................................................................119 6.6

Ultrasonic Cement Strength ..................................................................122 6.7

Summary ...............................................................................................125 6.8

Chapter 7: Conclusions and future work ....................................................128

Conclusions ...........................................................................................129 7.1

Future Work ..........................................................................................134 7.2

Page 12: Copyright by Xiangyu Liu 2017

xii

List of Abbreviations ...........................................................................................137

List of Key Symbols ............................................................................................139

List of Publications ..............................................................................................140

Bibliography ........................................................................................................141

Page 13: Copyright by Xiangyu Liu 2017

xiii

List of Figures

Figure 2.1 - Ternary phase diagram showing the composition of OPC, blast furnace

slag (BFS), fly ash (FA), silica fume and metakaolin.........................9

Figure 2.2 - Schematic representation of the alkali activation reaction process

(Juenger et al., 2011) .........................................................................15

Figure 2.3 - Effects of calcium and magnesium compounds (at the molar dosage of

0.09 mol) on the setting time of three different alkali-activated fly

ash/kaolinite blends ((Lee and van Deventer, 2002a) .......................20

Figure 2.4 - Primary cementation illustration ........................................................23

Figure 2.5 - Cement plugs in an abandonment well (image adopted from Global CCS

Institute) ............................................................................................27

Figure 2.6 - Lost circulation scenario with (a) partial loss, and (b) total loss (figure

adopted from petrowiki.org) .............................................................30

Figure 3.1 - Volume weighted particle size distribution of all fly ash particles. For the

x-axis, 40 size intervals were generated logarithmically between 0.1 and

1000. Y-axis shows the volume fraction of particles between those sizes.

...........................................................................................................37

Figure 3.2 - Volume weighted particle size distribution of limestone dust (admixture

A) ......................................................................................................40

Figure 3.3 - Schematic diagram of the pipe-in-pipe shear bond strength set-up ...48

Figure 3.4 - Schematic of the pressure transmission test set-up ............................50

Figure 3.5 - Picture of sample assembly in pressure transmission test ..................51

Page 14: Copyright by Xiangyu Liu 2017

xiv

Figure 4.1 - Compressive strength values of hardened Portland cement slurry (P1) and

geopolymer slurry (G1) with SBM (S1) contamination (replacement by

volume) at 24 hours, 170 °F and 3,000 psi. ......................................56

Figure 4.2 - Normalized compressive strength of hardened Portland cement slurry

(P1) and geopolymer slurry (G1) with SBM (S1) contamination

(replacement by volume) at 24 hours, 170 °F and 3,000 psi. ...........57

Figure 4.3 - Rheological properties of (a) Portland cement (P1) and (b) geopolymer

(G1) slurries replaced with various dosages of SBM (S1) by volume at

70 °F ..................................................................................................59

Figure 4.4 – Numerical simulation of drilling mud displacement with cement slurry.

The color bar shows the volume fraction of the cement slurry. The color

gradient at the interface indicates mixing of the two fluids (Enayatpour

and van Oort, 2017) ..........................................................................60

Figure 4.5 - Compressive strength of neat geopolymer (G1) and geopolymer hybrids

(G1S1) at 170 °F and 3,000 psi .........................................................62

Figure 4.6 - Rheological properties of neat geopolymer (G1), geopolymer hybrid

(G1S1-20) and Portland cement (P1-R) at 70 °F and 125 °F ...........63

Figure 4.7 - Thickening time of geopolymer hybrids at 125 °F ............................64

Figure 4.8 - 1-day compressive strength of geopolymer hybrids activated by 6M or

8M NaOH activator...........................................................................65

Figure 4.9 - Rheological properties of geopolymer hybrids activated by 6M or 8M

NaOH activator at 70 °F ...................................................................66

Figure 4.10 - (a) 1-day and (b) 3-day compressive strength of geopolymer hybrids

with three different types of fly ashes ...............................................68

Figure 4.11 - Rheological properties of G2S1 hybrids at 70 °F ............................69

Page 15: Copyright by Xiangyu Liu 2017

xv

Figure 4.12 - Effect of changing aluminosilicate source on the thickening time of

G1S1-20 and G2S1-20 slurries at 125 °F and 3,000 psi ...................70

Figure 4.13 - Effect of using seawater vs. DI water on the compressive strength of G1

and G1S1 hybrids at 70 °F and 3,000 psi .........................................71

Figure 4.14 - Effect of using seawater vs. Di water on the rheological properties of

G1 and G1S1 hybrids at 70 °F ..........................................................71

Figure 4.15 - Rheology of the original SBM (75/25 SWR, 23% CaCl2) as well as

modified SBMs at (a) 80° F, (b) 120° F, and (c) 150° F ..................73

Figure 4.16 - Effect of changing SWR and internal brine CaCl2 concentration of SBM

on (a) rheology and (b) gel strength of G2S1-30-2A at 125 °F and 3,000

psi ......................................................................................................74

Figure 4.17 - Effect of changing SWR and internal brine CaCl2 concentration of SBM

on thickening time of G2S1-30-2A. (BHCT of 125 °F and BHP of 3,000

psi).....................................................................................................75

Figure 4.18 - Effect of changing SWR and internal brine CaCl2 concentration of SBM

on 1-day compressive strength of G2S1-30-2A at 170 °F and 3,000 psi

...........................................................................................................76

Figure 4.19 - Effect of pressure on thickening time of G2S1-30-2A at BHCT of 125

°F .......................................................................................................77

Figure 4.20 - (a) 1-day and (b) 3-day compressive strength of G2S1 hybrids with

varying dosages of stability modifier (A) .........................................79

Figure 4.21 - Effect of adding 1.5% stability modifier (A) on the rheological

properties of (a) G2 and G2S1-20 hybrid, (b) G2S1-30 and G2S1-40

hybrids at 70 °F .................................................................................80

Page 16: Copyright by Xiangyu Liu 2017

xvi

Figure 4.22 - Effect of different dosages of stability modifier (A) on the rheological

properties of G2S1-40 hybrids at 125 °F ..........................................81

Figure 4.23 - Effect of adding 0.75% of stability modifier (A) on thickening time for

G1S1-30 hybrid .................................................................................82

Figure 4.24 - Normalized compressive strength of hardened Portland cement slurries

(P1) and geopolymer slurries (G1) with 20% mud replacement (by

volume) .............................................................................................83

Figure 4.25 - Picture of a geopolymer/mud sample containing 20% (a) S1 and (b) O2

...........................................................................................................84

Figure 5.1 - (a) rheology and (b) gel strength of liquid-form sodium silicate (LSS)

activated geopolymer (0% SBM) at room temperature ....................88

Figure 5.2 - (a) rheology and (b) gel strength of liquid-form sodium silicate (LSS)

activated geopolymer blended with 20% SBM at room temperature90

Figure 5.3 - (a) rheology and (b) gel strength of solid form sodium silicate (SSS)

activated geopolymer (0% SBM) at room temperature ....................91

Figure 5.4 - (a) rheology and (b) gel strength of solid form sodium silicate (SSS)

activated geopolymer blended with 20% SBM at room temperature92

Figure 5.5 - Compressive strength of liquid sodium silicate (LSS) activated

geopolymer for (a) fly ash 1 and (b) fly ash 2 in comparison to liquid

sodium hydroxide (LSH) activated geopolymer without any SBM at 170

°F and 3,000 psi ................................................................................94

Figure 5.6 - Compressive strength of solid sodium silicate (SSS) activated

geopolymer for FA 2 in comparison to liquid sodium hydroxide (LSH)

activated geopolymer without any SBM at 170 °F and 3,000 psi ....95

Page 17: Copyright by Xiangyu Liu 2017

xvii

Figure 5.7 - Effect of various activating solutions on the thickening time of neat

geopolymer slurry (G2) at 125 °F and 3,000 psi ..............................96

Figure 5.8 - Effect of adding 20% original SBM on thickening time of (a) LSH-8M

and (b) SSS-0.24 geopolymer slurries (G2) at 125 °F and 3,000 psi98

Figure 6.1 - Differential stress vs. axial strain for two samples each for geopolymer

subjected to a confining pressure of 100 psi and 500 psi at 7 days 102

Figure 6.2 - Differential stress vs. axial strain for two samples each for geopolymer

hybrid subjected to a confining pressure of 100 psi and 500 psi at 7 days

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

Figure 6.3 - Differential stress vs. axial strain for two samples each for Portland

cement subjected to a confining pressure of 100 psi and 500 psi at 7 days

.........................................................................................................103

Figure 6.4 - Self-healing properties of geopolymer and Portland cement pre-loaded at

(a) 7 days and (b) 28 days under atmospheric conditions. Cylindrical

samples were prepared at 170 °F ....................................................107

Figure 6.5 - Self-healing capability of geopolymer (G) under 500 psi confining stress,

for two samples (a) and (b). The G-7 sample was loaded beyond its yield

point at 7 days, and the same sample was re-tested to failure at 28 days

(G-7-28). Note the evident increase in peak stress observed for the

samples at 28 days ..........................................................................109

Figure 6.6 - Self-healing capability of geopolymer hybrid (GH) under 500 psi

confining stress, for two samples (a) and (b). The GH-7 sample was

loaded beyond its yield point at 7 days and the same sample was re-

tested to failure at 28 days (GH-7-28). Note the evident increase in peak

stress observed for the samples at 28 days ......................................110

Page 18: Copyright by Xiangyu Liu 2017

xviii

Figure 6.7 - Self-healing capability of Portland cement (P) under 500 psi confining

stress, for two samples (a) and (b). The P-7 sample was loaded beyond

its yield point at 7 days and the same sample was re-tested to failure at

28 days (P-7-28). Note the evident reduction in peak stress observed for

the samples at 28 days.....................................................................111

Figure 6.8 - Cement-to-pipe shear bond strength for two samples of Portland cement

and geopolymer with clean steel pipe at 170 °F on day-7 ..............114

Figure 6.9 - Cement-to-pipe shear bond strength for two samples of Portland cement

and geopolymer with SBM coated steel pipe at 170 °F on day-7 ...115

Figure 6.10 - Cross-section view from bottom of the cement-to-pipe shear bond test

samples for (a) Portland cement and (b) geopolymer .....................115

Figure 6.11 - Pressure transmission curves, linear fit and the delay factor for (a)

Portland cement, (b) geopolymer, and (c) slag at 28 days ..............117

Figure 6.12 - Hydraulic conductivity (HC) and linear fit trend lines of Portland

cement, geopolymer, and slag at 28 days .......................................118

Figure 6.13 - Pore size distribution of Portland cement, geopolymer and slag obtained

from MIP test, samples were cured at 170 °F for 28 days ..............121

Figure 6.14 - Geopolymer sonic compressive strength and transit time obtained from

UCA test with built-in empirical correlations developed for Portland

cement, test was conducted at 189 °F and 5,000 psi .......................123

Figure 6.15 - Portland cement sonic compressive strength and transit time obtained

from UCA test with built-in empirical correlations developed for

Portland cement, test was conducted at 170 °F and 3,000 psi ........124

Page 19: Copyright by Xiangyu Liu 2017

xix

Figure 6.16 - Geopolymer sonic compressive strength and transit time obtained from

UCA test with empirical correlations developed for geopolymer, test

was conducted at 189 °F and 5,000 psi ...........................................125

Figure 7.1 - Circular flow diagram showing the design philosophy of

geopolymer/mud hybrid cement .....................................................133

Page 20: Copyright by Xiangyu Liu 2017

xx

List of Tables

Table 2.1 - Empirical correlations for geopolymer specimens prepared with various

NaOH concentrations at two different temperatures. Specimens were

cured under a pressure of 5,000 psi (Khalifeh et al. 2014) ...............29

Table 3.1 - Composition of fly ashes (weight %) ..................................................36

Table 3.2 - Geopolymer mix design ......................................................................39

Table 3.3 - Properties of the SBMs (S1 & S2) ......................................................42

Table 3.4 - Properties of the OBMs (O1 & O2) ....................................................43

Table 3.5 - Parameters used in the MIP measurement ...........................................54

Table 6.1 - Confined compressive strength at 7 days ..........................................103

Table 6.2 - Mechanical properties of hardened slurries at 500 psi confining pressure

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

Table 6.3 - Peak stress values at 7-days and after 21-day waiting period (at 500 psi

confining stress) ..............................................................................112

Table 6.4 - Porosity and pore size evaluation of Portland, geopolymer and slag from

MIP measurement at 7 days ............................................................120

Page 21: Copyright by Xiangyu Liu 2017

1

Chapter 1: Introduction

MOTIVATION 1.1

Field experience and laboratory investigations have established that a major cause

of oilwell cementing failure is contamination of the cement slurry by drilling mud

(Aughenbaugh et al., 2014; Beach and Goins, 1957; Miranda et al., 2007; Morgan and

Dumbauld, 1952). Intermixing of cement slurry and drilling mud will almost inevitably

occur even under nearly ideal displacement conditions. Numerical simulation has shown

that high levels of cement contamination occur when interface instability occurs during

mud displacement (Enayatpour and van Oort, 2017). With advancements in drilling in

more challenging environments, including deepwater, high-pressure / high temperature

(HPHT), unconventional plays accessed by directional / horizontal drilling etc.,

progressively more oil / gas reservoirs are becoming accessible to exploration and

production. This in turn leads to the more difficult cementing operations, and it is a fair

assessment to state that cementing technology has not kept pace with the advancement of

drilling technology. Highly deviated wellbores, wellbore enlargement stemming from

hole instability, poor casing centralization, improper displacement, etc. all increase the

potential of cement contamination (Nair et al., 2015; Nelson and Guillot, 2006), in turn

resulting in poor / insufficient hydraulic zonal isolation by cement. This can lead to a

variety of “knock-on” problems, including sustained annular casing pressure, intermixing

of reservoir fluids, potential contamination of shallow aquifers, increased well control

incident risks, etc. Cement contamination may also play a prominent role in well

abandonment operations, when leftover drilling mud in the well can mix with - and

contaminate - the cement slurry used for abandonment plugs.

Page 22: Copyright by Xiangyu Liu 2017

2

Targeting the contamination issue, a system named “universal fluid (UF)” was

developed over 25 years ago. This system was based on a water-based mud (WBM)

which could be converted into a well cement (Hale and Cowan, 1991; Nahm et al., 1994).

A UF, when used as a drilling mud, delivers all of the necessary characteristics of a

drilling mud. As an added dimension, however, it contains a hydraulic material (blast

furnace slag), which can be activated to set up like a cementitious material when an

alkaline activator is added. The main advantage of a UF in comparison to an ordinary

Portland cement (OPC) system is its improved inherent compatibility with water-based

drilling muds. With the use of a UF, any undisplaced drilling mud or mud filter cake may

be incorporated into the cement, thereby largely eliminating the effect of contamination.

The UF system, however, never found widespread industry acceptance after some

promising early trials and field applications.

To date, a similar mud-to-cement conversion technique has not been successfully

developed and applied for NAFs. Contamination of cement slurry by NAFs is well-

known to be detrimental to important cement slurry properties such as pumping time and

compressive strength. Studies have shown that Portland cement slurries are particularly

sensitive to NAF contamination. With even a relatively small quantity of SBM

contamination, Portland cement slurries can completely lose their compressive strength

(Aughenbaugh et al., 2014; Miranda et al., 2007; Patel et al., 1999). Changing the

industry’s dependence on Portland cement, however, will require a paradigm shift that is

supported by the development of suitable and competent non-Portland based alternatives

that address the shortcomings of Portland cement.

This study will focus on the development of high performance geopolymers for

mud-to-cement conversion of SBMs. Development of geopolymers to be the next

generation of oilwell cements is a promising approach to meet current well cementing

Page 23: Copyright by Xiangyu Liu 2017

3

challenges. Geopolymer cement is a type of alkali-activated material (AAM) that can be

formed by blending an alkaline solution with an aluminosilicate powder such as fly ash

(Duxson et al., 2006). Geopolymers are receiving increasing attention as a replacement

for Portland cement in various industries (e.g. building and road construction) due to their

versatile chemistry, low environmental impact (e.g. the manufacturing of AAMs usually

leads to no or low CO2 emissions, see Provis and van Deventer, 2014), and the fact that

AAMs such as fly-ash are typically waste streams and side-products of other industries

(e.g. energy production from coal burning), allowing relatively inexpensive waste to be

(re-)employed for a useful purpose. The different chemical makeup makes geopolymers

appropriate for use in cases where OPC is less suitable. Geopolymers have largely been

focused on construction applications over the last few decades. A point of special interest

is that researchers have, until now, not been successful in identifying a geopolymer slurry

formulation that is readily pumpable (note that the high viscosity of most geopolymer

slurries allows them to be poured, e.g. for road bases and construction templates, but not

pumped without excessive pumping pressures), which is essential for oilwell cementing

(Khalifeh et al., 2014). Moreover, open literature regarding formulation of these slurries

for oil / gas well applications remains very limited.

OBJECTIVES 1.2

The main objective of this research is to explore the feasibility of using alkali-

activated fly ash (also known as geopolymer) to convert SBM into cementitious

materials. The specific R&D objectives and questions can be broken down as follows:

1. Can geopolymer solidify SBM at various slurry-to-mud mixing ratios?

Page 24: Copyright by Xiangyu Liu 2017

4

2. Can the compressive strength of the geopolymer/mud hybrid system be tailored to

meet various applications such as primary cementing, remedial cementing, or lost

circulation control?

3. Will geopolymer/mud hybrid systems have acceptable pumpability and

rheological properties at both surface and downhole conditions for efficient mud

displacement and cement placement?

4. Will the geopolymer/mud hybrid system be stable at various pressure and

temperature conditions?

5. What are critical hardened state properties of the geopolymer/mud hybrid system

including mechanical properties, porosity and hydraulic conductivity?

6. Can geopolymer and geopolymer/mud hybrid systems be developed for

application in well abandonment and decommissioning? In particular, do these

systems exhibits “self-healing” characteristics, i.e. strength recovery after

yielding, which will contribute to long-term integrity of well abandonment plugs?

DISSERTATION ORGANIZATION 1.3

This dissertation is organized into several sections. Chapter 2 provides a succinct

literature review with introduction to the definition, composition and synthesis of

geopolymers. The geopolymer characteristics that are of importance to oilwell cementing

applications are also reviewed in detail. Chapter 3 gives detailed description of the

materials used in this study and the experimental methods used to characterize the new

cement/mud hybrid system. Chapter 4 presents the results and discussion on hydroxide-

activated geopolymers and their effectiveness in solidifying SBM. This chapter also gives

details about the influence of activator dosage, aluminosilicate source, different types and

composition of mud, and effect of downhole pressure. Chapter 5 evaluates the

Page 25: Copyright by Xiangyu Liu 2017

5

effectiveness of using silicate activation in comparison with hydroxide activation in the

mud solidification method. Chapter 6 focuses on the hardened state properties of the

geopolymer/mud hybrid cement with investigation of the mechanical properties, self-

healing capability, porosity and hydraulic conductivity of the geopolymer system.

Finally, chapter 7 gives an overall summary of the novel mud solidification method

presented in this dissertation, and suggests ideas for future work.

Page 26: Copyright by Xiangyu Liu 2017

6

Chapter 2: Background

The first part of this chapter clarifies geopolymer definition and terminology. The

general rules of geopolymer synthesis are introduced. The reaction process and

microstructure of geopolymers are also discussed, as they influence their behavior and

properties. Next, critical properties of geopolymer including compressive strength,

rheological properties, setting time control and self-healing behavior, which are all of

great importance to oilwell cementing, are reviewed in detail. Subsequently, the current

state of geopolymer research aimed at oilwell cementing applications is presented, with

introduction of basic and essential concepts regarding primary cementing, well

abandonment plug cementation and remedial cementing. Finally, a brief discussion of the

mud solidification technique is provided, showing the potential of using geopolymers to

solidify non-aqueous drilling muds.

GEOPOLYMER: DEFINITION AND TERMINOLOGY 2.1

The blending of an alkali solution with an alumina- and silica-containing powder

leads to a binder phase that can harden and develop compressive strength like an OPC.

This reaction was first patented by Kühl in 1908, and was later on studied in more detail

by Purdon during the development of an alkali-activated blast furnace slag (Kühl, 1908;

Purdon, 1940). Since the 1990s, research into binders with alkali activation has expanded

rapidly all over the world, primarily driven by the potential for substantial reduction in

greenhouse gas emissions that appears to be possible by moving away from using OPC,

the manufacturing of which globally produces a large amount of CO2 (Duxson et al.,

2007b; Juenger et al., 2011).

Page 27: Copyright by Xiangyu Liu 2017

7

A broad classification of “alkali-activated materials” (AAM) is applied to any

binder system formed from an alkali activator and a silicate precursor powder. By this

definition, an AAM can be an alkali-activated calcium silicate such as alkali-activated

slag (AAS) or a more aluminosilicate-rich material such as alkali-activated fly ash. The

name “geopolymer” was first applied by Davidovits in 1978 to a group of mineral

polymer resins derived from the reaction of metakaolin with soluble silicate (Davidovits,

2015). In most recent publications, “geopolymer” is considered to refer to a more specific

type of AAM; although it is still used to describe a wide range of alkali-activated binders,

it mostly refers to alkali-activated low-calcium or calcium-free aluminosilicates (Juenger

et al., 2011; Provis and van Deventer, 2014). Note that the use of the term “geopolymer”

to characterize those alkali-activated materials is still a topic of considerable controversy.

Further discussion on the definition of geopolymer is, however, beyond the scope of this

work, which will use the generally accepted definition as outlined in the 2014 RILEM

report (Provis and van Deventer, 2014). The current study will focus on developing low-

calcium fly ash-based geopolymer cements for oil / gas applications. Background details

on structure, properties, synthesis and applications of fly-ash based geopolymer will be

reviewed in detail in the following sections.

SYNTHESIS OF ALKALI-ACTIVATED FLY ASH-BASED GEOPOLYMER 2.2

Aluminosilicate source 2.2.1

To create a geopolymer slurry, an aluminosilicate precursor powder is mixed with

an alkaline activating solution. Common sources of aluminosilicate include metakaolin

and fly ash. Metakaolin is a dehydroxylated form of kaolinite clay. Alkaline-activated

metakaolin forms a strong and durable binder material (Davidovits, 2015). Fly ash,

Page 28: Copyright by Xiangyu Liu 2017

8

although having more variability in its composition, delivers a more favorable rheology

than metakaolin-based binders at a much lower cost (Kong et al., 2007). It was therefore

selected as the preferred aluminosilicate source for this study.

Fly ash is a by-product generated from coal burning power plants. The

composition of any fly ash greatly depends on the source of coal and the operating

conditions of the boiler. ASTM Standard C618-18 (2015) defines two main classes of fly

ash. The sum of SiO2, Al2O3, and FeO must be above 50% for a class C fly ash and must

be above 70% for a class F fly ash. Class C fly ashes typically have higher total CaO

content than class F fly ashes (Figure 2.1). Due to the rapid cooling in the flue, fly ash

particles typically contain 50-90% of amorphous phases (Ward and French, 2006). The

amorphous phases in the fly ashes are mainly aluminosilicate glasses with a tetrahedral

structure (Hemmings and Berry, 1987; Williams et al., 2005; Williams and van Riessen,

2010). The remaining parts of the fly ash are crystalline phases including quartz,

hematite, mullite, magnetite, etc. (Hemmings and Berry, 1987).

Unlike Portland cement, which is manufactured specifically as a construction

material according to industrial standards, fly ash is a by-product with more variability in

the raw materials, therefore can vary appreciably in reactivity. It has been commonly

accepted that the main characteristics of a fly ash for optimal geopolymer formulation

include a high vitreous content, more specifically high reactive aluminate and high

reactive silicate, since these are the main building blocks of the geopolymer reaction

product (Duxson et al., 2006; Provis and van Deventer, 2014). The quantification of the

crystalline phases in fly ash material can be obtained from X-ray diffraction (XRD). The

determination of glassy phases, however, is difficult. Different elements can occur in

different ways within the ash. For example, silicon, aluminum and iron elements may

exist as crystalline phases such as quartz or magnetite, as aluminosilicate minerals such

Page 29: Copyright by Xiangyu Liu 2017

9

as mullite, or as amorphous aluminosilicate glasses. The amorphous glassy phases do not

have regular arrays of atoms that can be characterized by the XRD measurements;

instead, they show as a “hump” in the X-ray diffraction patterns.

Figure 2.1 - Ternary phase diagram showing the composition of OPC, blast furnace slag

(BFS), fly ash (FA), silica fume and metakaolin

A considerable amount of prior research has been focused on the quantitative

analysis of the glassy constituents and the chemical reactivity of fly ash particles. One

way to quantify the total fraction of amorphous phase is by subtracting the total

crystalline fraction obtained from XRD from the bulk oxide content of the fly ash

(Aughenbaugh, 2013; Ward and French, 2006). This method, however, usually fails to

distinguish between different amorphous phases. Selective dissolution is a chemical

Al2O

3

SiO2

CaO

Silica Fume

OPC

Metakaolin

Page 30: Copyright by Xiangyu Liu 2017

10

method that can also provide information on the glassy content. When fly ash particles

are dissolved in acids (such as hydrofluoric acid, acetic acid or oxalic acid) or caustic

solutions (such as sodium hydroxide), the glassy phases are dissolved and the unreactive

crystalline phases are assumed to be unaffected (Aughenbaugh, 2013; Hemmings and

Berry, 1987; Williams and van Riessen, 2010). The bulk glassy composition can be

estimated by measuring the change in weight during dissolution, i.e., by subtracting the

unreactive crystalline content from the total mass prior to dissolution, the amount of

glassy phases can be determined.

One way to identify the chemical composition of glassy phases in fly ash particles

is to use Energy Dispersive Spectroscopy (EDS) along with Scanning Electron

Microscope (SEM) (Aughenbaugh, 2013; Durdziński et al., 2015). Using EDS, a

spectrum for every pixel can be plotted into a set of element intensity maps. Multi-

spectral image analysis software can be used to assist the analysis and quantification of

groups that has similar chemical compositions. Insight into the microstructure and

location of phases can thus be obtained.

The characterization of fly ash particles can provide very useful information in

geopolymer design. More research is ongoing to analyze the individual glassy phases in

fly ash. Based on literature review, it can be summarized that a suitable fly ash precursor

has to contain sufficient reactive aluminum, as it is the main component in the

aluminosilicate gel (Fernández-Jiménez et al., 2006b; Fernández-Jiménez and Palomo,

2005a). The total reactive silica in the fly ash precursor should be between 40-50% in the

fly ash and the reactive Si/Al ratio should be below 2 (Fernández-Jiménez and Palomo,

2003, Fernández-Jiménez et al., 2006). Other properties that may influence the reactivity

of the fly ash include (1) the morphology and size of the fly ash particles; (2) the calcium

content; (3) the concentration of unburned coal (which preferably should be lower than

Page 31: Copyright by Xiangyu Liu 2017

11

5%); (4) the iron content (which preferably should be no higher than 10%) (Fernández-

Jiménez and Palomo, 2003).

Activating solution 2.2.2

Typical alkaline activators include alkali metal hydroxides, silicates, or a blend of

the two. Carbonate and sulfate activators have also been tested by researchers but are less

effective in low-calcium AAMs (Fernández-Jiménez and Palomo, 2005b; Shi and Day,

1995). The alkali cations in the activator can be sodium, potassium and lithium, with

sodium being the most studied. The choice of cation and the concentration of the alkaline

solution will result in varying properties in the hardened geopolymer (Lizcano et al.,

2012).

Hydroxide Activation

Alkali hydroxides are usually prepared in the form of an aqueous solution.

Typical hydroxide solutions for geopolymer activation range from 8 g/mol to 15 g/mol

(Provis and van Deventer, 2014). Upon mixing with the precursor, the hydroxide ion

(OH-) works as a catalyst in dissolving the solids. The alkali will be incorporated into the

geopolymer structure and acts to balance the negative charges. The structure of the

geopolymer gel will be discussed in more detail later in this review. Hydroxide solutions

have a very high pH, thus they require extreme caution during handling. Once the

geopolymer gels start to form and the hydroxides become chemically bound in the

structure, the resulting hardened material does not pose a hazard to living beings or the

environment.

Page 32: Copyright by Xiangyu Liu 2017

12

Silicate Activation

Alkali silicate solution, or “waterglass”, is also commonly used for geopolymer

activation. Commercially available sodium silicate solutions can be used, or alternatively

they can be prepared in custom form by dissolving amorphous silica (silica fume) in

appropriate alkali hydroxide solutions. The alkali silicate solution provides additional

silicon that is immediately available to interact with other dissolved ions. Studies showed

that the use of sodium silicate leads to a denser geopolymer structure compared to a

geopolymer formed using sodium hydroxide solution (Ma et al., 2012).

Proportioning 2.2.3

Geopolymer can be proportioned based on the water-to-solid ratio or solution-to-

powder ratio in the same way as Portland cement concretes are proportioned (Al Bakri et

al., 2011; Bakharev, 2005; Fernández-Jiménez et al., 2008, 2006b; Oh et al., 2010; Ruiz-

Santaquiteria et al., 2012). Another way to proportion geopolymer is based on molar ratio

of the aluminum and silicon constituents. Khale and Chaudhary (Khale and Chaudhary,

2007) recommend the following oxide ratio for proper polymerization and strength

development for a metakaolin-based geopolymer: M2O/SiO2 = 0.2-0.48, SiO2/Al2O3 =

3.3-4.5, H2O/M2O = 10-25, and M2O/Al2O3 = 0.8-1.6, where M is the alkali metal (such

as sodium or potassium). Formulations with compositional ranges of SiO2/Al2O3 < 1 or

SiO2/Al2O3 > 5 may result in interesting properties for certain specific applications, but

will not be discussed in this literature review. In addition, the alkali composition also

affects the development of geopolymer structure (Duxson et al., 2007a, 2005a).

Davidovits (Davidovits, 1982) recommends (Na2O, K2O)/SiO2 ratio in the range of 0.2-

0.28, and (Na2O, K2O)/Al2O3 ratio in the range of 0.8-1.2. A wide range of recommended

molar ratios can be found in literature (Aughenbaugh et al., 2014; Barbosa et al., 2000;

Page 33: Copyright by Xiangyu Liu 2017

13

Chindaprasirt et al., 2012; Duxson et al., 2007a; Fletcher et al., 2005; Rowles and

O’Connor, 2003). The proportioning of geopolymer solution and precursor must be based

on the composition of the specific aluminosilicate, augmented by trial and error testing.

Compared to metakaolin, proportioning for fly ash-based geopolymer is more

complicated due to greater compositional variability in the latter. Only the reactive

portion of the fly ash should be used in deciding the proportions. As mentioned in the

previous section, finding the amount of reactive phases in a fly ash is very challenging.

Considerable effort has been made on this topic, which still remains an active area of

research (Aughenbaugh, 2013; Chancey et al., 2010; Durdziński et al., 2015). For

simplicity, the non-crystalline part of the fly ash can be considered as the reactive portion

regardless of the varying level of reactivity of different glassy phases.

Curing regime 2.2.4

Curing conditions have a large effect on hardened geopolymer properties. Most

sources report that the synthesis of low-calcium alkali-activated binders requires elevated

temperature. The curing temperature regimes are typically between 40 °C and 100 °C

(Davidovits, 1989; Duxson et al., 2007a; Fernández-Jiménez et al., 2008; Rowles and

O’Connor, 2003; Singh et al., 2015). Room temperature curing is possible, but may

require longer than desirable time to develop measureable strength (Somna et al., 2011).

In literature, both short-term and long-term heat curing conditions have been

studied. Short-term heat curing means that the samples were cured at elevated

temperature for up to 24 hours, followed by room temperature curing until the samples

were tested. Long-term heat curing requires the samples to be cured at elevated

temperature until tested. Palomo et al. (Palomo et al., 1999) studied the effect of short-

Page 34: Copyright by Xiangyu Liu 2017

14

term heat curing at 65 °C and 85 °C using a fly ash based geopolymer. They found that

the higher temperature resulted in higher compressive strength when the samples were

heat cured for 2 or 5 hours. Longer curing time did not result in greater strength. Najafi

Kani and Allahverdi (2009) found that longer pre-curing at room temperature before the

application of heat is beneficial for strength development of the geopolymer.

POLYMERIZATION MECHANISM AND MICROSTRUCTURE 2.3

The chemistry and reaction mechanism of alkali activation is currently a hot topic

in non-Portland cement research. It is widely agreed that the geopolymer reaction is a

dissolution-precipitation process. Upon mixing, both silicate and aluminate species from

the aluminosilicate powder dissolve into the basic solution. Silicon exists in the solution

in the form of HSiO43-

until a critical concentration is reached. Precipitation happens

concurrently forming an aluminum-rich gel in the first stage, and a more silicon-rich gel

in the second stage (Duxson et al., 2006; Duxson and Provis, 2008; Fernández-Jiménez

and Palomo, 2005b).

The micro-structure of the hardened geopolymer binder is a highly cross-linked

aluminosilicate network structure that is very similar to zeolite frameworks (Bell et al.,

2008a, 2008b; Fernández-Jiménez et al., 2008; Juenger et al., 2011; Provis and van

Deventer, 2014). This structure is X-ray amorphous and is oftentimes described as N-A-

S-H gel, where N is Na2O, A is Al2O3, S is SiO2 and H represents water (García-Lodeiro

et al., 2007). The tetrahedral framework of silicon and aluminum is linked by oxygen and

charge-balanced by sodium or other cations (Davidovits, 1989). This aluminosilicate gel

contains a very low level of chemically bound water. It therefore is often characterized as

N-A-S-(H) (Duxson et al., 2005a; Rahier et al., 1996).

Page 35: Copyright by Xiangyu Liu 2017

15

When the precursor solid contains higher amounts of calcium, for instance when

Class C fly ash or blast furnace slag is used in combination with a low-calcium

geopolymer system, a binder phase of C-(A)-S-H gel that is partially crystalline and

partially amorphous will coexist with the N-A-S-(H) gel (Richardson et al., 1994; Wang

and Scrivener, 1995). Figure 2.2 shows a conceptual geopolymer reaction process for

both high-calcium and low-calcium fly ashes in a most general sense. The two stages of

gel evolution are represented by “solidification and hardening” and “ongoing gel

evolution” respectively.

Figure 2.2 - Schematic representation of the alkali activation reaction process (Juenger et

al., 2011)

Dissolution of solid

aluminosilicate source

Silicate species in

activating solution

Rearrangement and exchange

among dissolved species

Gel nucleation

C-(A)-S-H gel N-A-S-(H) gel

Solidification and hardening

Ongoing gel evolution with progression

towards crystallization

Page 36: Copyright by Xiangyu Liu 2017

16

PROPERTIES AND ADMIXTURE 2.4

Here, the compressive strength, rheological properties, setting time control and

self-healing capability of geopolymers are reviewed in more detail, since these are of

great importance to the development of slurries suitable for oilwell cementing. The

admixtures that modify each of the relevant geopolymer properties are also discussed.

In Portland cement systems, “admixture” refers to the additional components that

are added to give special properties to the fresh or hardened cement slurry. Admixtures

may enhance the workability, strength or durability of a given cement mixture. In the

context of an alkali-activated system, the alkaline activators are not considered to be an

admixture since they are a part of the binder chemistry. An admixture will be defined as

the additives that are purposely added to alter the properties of the AAM.

Unlike the Portland cement admixture chemistries, which are well-researched and

documented, admixtures to AAM or geopolymer have not been widely addressed in the

literature. Moreover, the laboratory results can be largely different or even contradictory

among different researchers. This can be attributed to the complexity and variation in the

precursor, the activator, and the type and dosage of the admixture. It is also clear that the

admixtures that are developed for Portland cement chemistries behave very differently in

an AAM, due to the distinct hardening mechanism. For low-calcium alkaline activated

material especially, the majority of the well-known OPC admixtures have been found

ineffective for property modification. In the next section, the admixture chemistries and

their effectiveness on altering geopolymer performance will be discussed in detail for

properties that are critical to oilwell cementing applications.

Page 37: Copyright by Xiangyu Liu 2017

17

Compressive strength 2.4.1

To successfully utilize geopolymers on a significant scale in the construction of

buildings and other infrastructure, their compressive strength was developed to be close

to that of a hardened Portland cement. In literature, studies have reported compressive

strength values higher than 70 MPa (more than 10,000 psi) for geopolymer concrete

cured at temperatures in the range of 50-80 °C (van Deventer et al., 2012). Note that the

compressive strength is expected to increase as the curing temperature increases. Room

temperature curing leads to lower strength in the range of 20-23 MPa (around 3,000 psi)

(Somna et al., 2011). All of these reported values meet the minimum required

compressive strength for oil/gas wells. With the confining pressure present in a downhole

condition, the failure strength of geopolymer will be even higher than the unconfined

compressive strength values, providing sufficient strength for supporting casing and

isolating different zones.

Nasvi et al. (2012a) observed some level of strength reduction at curing

temperatures that are higher than 100 °C (212 °F). They believed that is due to the

breaking up of the intergranular structure of geopolymer at those high temperatures.

Portland cement also suffered from strength retrogression when temperature exceeds 60

°C based on their laboratory evaluations. Nevertheless, the strength reduction rate of

geopolymer is less than that of a Portland cement. The issue of strength retrogression

needs to be specifically addressed when applications for high temperature / high pressure

(HPHT) oil / gas wells are considered.

Rheological properties 2.4.2

Flowability is another important characteristic for oil / gas applications.

Previously reported geopolymer slurries are in general too viscous to be pumped in a

Page 38: Copyright by Xiangyu Liu 2017

18

typical oil / gas well. Adding additional water or a superplasticising admixture are the

two main ways to improve the pumpability of a cement slurry. Similar to Portland cement

slurries, adding more water in geopolymer mixture will improve the workability.

However, excess water will negatively influence the strength and porosity of the

hardened material. Therefore the amount of water requires regulation and optimization

(Barbosa et al., 2000). Effective admixtures for reducing viscosity in Portland cement

slurries, including lignosulfonate-, polynaphthalene- and polycarboxylate-based

superplasticizers, have been tested in AAM slurries. Very distinct, if not contradictory,

effects were observed when using superplasticizers in alkali-activated slags (Criado et al.,

2009; Palacios et al., 2009; Palacios and Puertas, 2005; Puertas et al., 2003). The

contrasting behavior of geopolymers with superplasticizers can be explained by the

chemical instability of superplasticizers at pH greater than 13. Some metakaolin or

bottom ash based slurries showed improvement in workability or slump with

superplastisizer. However, none of those slurries achieved a thin enough viscosity for

pumping the geopolymer slurry down oil / gas wells (Hardjito and Rangan, 2005; Kong

and Sanjayan, 2010; Nematollahi and Sanjayan, 2014). Therefore, designing a

geopolymer slurry with appropriate rheological properties and finding effective

admixtures for the high alkaline environment has remained a big challenge. The topic is

discussed in more detail in the remainder of this dissertation.

Setting time control 2.4.3

Setting time control of geopolymers has been studied by various researchers. It

has been well-documented that borate species act as effective retarders for Portland

cement. Nicholson et al. (2005) studied the effect of using borate in silicate-activated

Page 39: Copyright by Xiangyu Liu 2017

19

class C fly ash and discovered that borates can significantly delay the onset of setting at

7% by weight of fly ash or higher. At these high dosages, however, the binder suffered a

significant drop in compressive strength. The borate is not expected to work as

effectively in lower calcium AAMs, as the borate only resides in the tetrahedral BO4

sides of the geopolymer structure when there is limited amount of calcium in the system.

Other admixtures including potassium salt, phosphoric acid and K2HPO4 have been

identified as effective retarders for alkali-activated high calcium AAMs (Chang, 2003;

Lee and van Deventer, 2002a).

Salt is another class of admixtures that alters the setting time of Portland cement

slurries. For silicate-activated fly ash mixtures, Lee and van Deventer (Lee and van

Deventer, 2002a, 2002b) tested a range of salts and summarized the results in a chart

(Figure 2.3). As shown in the figure, calcium salts generally showed an accelerating

effect, whereas magnesium compounds did not significantly affect the setting time for all

three systems tested.

Other salts have been evaluated for high-Ca alkali-activated fly ash and alkali-

activated blast furnace slag (BFS) cements. Brough et al. (2000) observed that NaCl at

levels up to 4% has an accelerating effect on silicate-activated BFS cements, whereas

dosages higher than 4% of NaCl retards the reaction. In another study, little effect of

NaCl on the setting behavior was observed on a BFS system up to 20% addition

(Sakulich et al., 2009). In contrast, in OPC slurries, NaCl acts as an accelerator at

concentrations up to 15% and as a retarder at concentrations above 20%. Between

dosages of 15% - 20%, NaCl is essentially neutral (Nelson and Guillot, 2006).

Page 40: Copyright by Xiangyu Liu 2017

20

Figure 2.3 - Effects of calcium and magnesium compounds (at the molar dosage of 0.09

mol) on the setting time of three different alkali-activated fly ash/kaolinite blends ((Lee

and van Deventer, 2002a)

Self-healing capability 2.4.4

Ahn and Kishi (2010) reported that geopolymer has a self-healing capability,

where the geopolymer matrix can heal when a micro-crack is present. If a geopolymer

indeed self-heals, it will be a great candidate for well abandonment plugs where

maintaining long-term cement sheath integrity is critical. However, literature about

geopolymer re-healing capability is quite sparse and warrants the need for a thorough

Page 41: Copyright by Xiangyu Liu 2017

21

study. Since remediation of cement in an oil / gas well is costly, difficult, and rarely

successful, it would be beneficial to use a cementitious material that can re-heal in-situ. A

good understanding of the re-healing process and mechanism is imperative for providing

adequate zonal isolation throughout the life of the well.

Other properties 2.4.5

To evaluate the potential of using geopolymer in carbon sequestration and in

offshore operations, Nasvi et al. (2012b) Giasuddin et al. (2013) and Duran (Duran,

2015) studied the effects of saline water curing on geopolymer cement strength and

durability. The results showed that the strength of geopolymer increases with increasing

salinity, possibly because the NaCl content provided more resistance against alkali

leaching from the geopolymer matrix. By comparison, hardened Portland cement slurries

can lose 50% strength when cured in saline water (Nasvi et al., 2012b).

Nasvi et al. (2014a, 2014b, 2013) studied the effect of temperature on the

permeability of class F fly ash-based geopolymers for carbon capture and storage. The

permeability increases with increasing curing temperature during setting and hardening.

The maximum apparent CO2 permeability at any temperature was 0.04 µD, lower than

the CO2 permeability of a typical OPC used in oil / gas wells (0.1-200 µD, see Bachu and

Bennion, 2009; Laudet et al., 2011; Le-Minous et al., 2017), which is well below the

permeability limit (200 µD) recommended by API (Kutchko et al., 2009). These findings

indicate that geopolymer is a viable alternate to ordinary Portland cement for long-term

well integrity as it delivers good durability and low permeability.

Overall, geopolymer has significant potential to be used as an alternative to

Portland cement for oil well cementing. Other than the properties and qualities that have

Page 42: Copyright by Xiangyu Liu 2017

22

been discussed above, geopolymer has also been reported to provide additional benefits:

(1) good durability under harsh conditions (Bakharev, 2005; Lloyd, 2008); (2) resistance

to acid and chemical attack (Bakharev, 2005; Uehara, 2010); (3) low permeability

(Chung et al. 2010); (4) resistance to high temperature (Duxson et al., 2007b; Nasvi et al.,

2012a); (5) good volumetric stability after hardening (Papakonstantinou et al., 2001); (6)

adhesion and binding to multiple surfaces including metallic substrates (Bell et al., 2005);

(7) low cost; (8) low environmental impact. Of course, all of these great properties were

obtained for different mixes as well as a variety of curing conditions. Application-

specific mix design and associated optimization is required to attain the desired

properties.

GEOPOLYMER IN OILWELL CEMENTING 2.5

Geopolymers have been researched during the past few decades for potential as a

replacement for OPC in a wide range of applications. Geopolymer, for instance, can be

designed for building and repairing infrastructure, for making fire resistant concrete, for

hazardous and radioactive waste encapsulation, etc. The application in oilwell cementing,

however, is much less studied. In the following, the basic concepts of oilwell cementing

operations are introduced, along with a critical review of the current state of geopolymer

research in well cementing applications.

Primary cementation 2.5.1

Primary cementation is a technique for placing cement slurries in the annulus

between casing and formation rock or between casing strings (Figure 2.4). In principle,

the primary cementing techniques are similar regardless of the type and size of casing

Page 43: Copyright by Xiangyu Liu 2017

23

string. Cement slurry is pumped down through the string to be cemented, displaces

drilling mud as it moves back up the annulus, and is left to set in the annulus. The

foremost goal of the hardened cement sheath is to provide complete and durable zonal

isolation in the wellbore for the lifetime of the well, meaning that the hydraulic bond

formed by the cement will prevent the migration of reservoir fluids (brine, oil, gas)

between strata and up to the surface.

Figure 2.4 - Primary cementation illustration

Successful primary cementing requires accurate knowledge and preparation of the

well (creating a “cementable” borehole), meticulous planning and testing of the cement

slurry, and proper job execution according to plan. There are many facets to the design of

Casing

Production Liner

Drilling Liner

Tie-back Liner

Cement Sheath

Page 44: Copyright by Xiangyu Liu 2017

24

a well-specific cement slurry. Some of the critical characteristics of a cement slurry

include:

Proper rheological characteristics for mixability and pumpability.

Optimized mud removal and displacement by the cement slurry.

Sufficient pumping time for circulation and placement.

Right-angle setting to prevent formation fluid/gas invasion into the cement.

Sufficient compressive strength and flexural strength.

Good hydraulic bonding capability to seal the casing and formation interfaces.

Appropriate mechanical properties for long-term durability.

To achieve these desired properties, Portland cement systems have been

rigorously studied and additive chemistry has been well developed to make it possible to

satisfy such a wide range of requirements. Geopolymer, on the other hand, is much less

studied, especially under downhole well conditions. Research regarding geopolymer-

based well cementing has been mostly in-house within service companies, and has just

started to receive more attention in academia in recent years.

Salehi et al. (2016a, 2016b) conducted a series of laboratory tests dedicated to the

application of geopolymer for primary cementing. They evaluated the pumping time of a

class F fly ash-based geopolymer at various temperatures and saw strong accelerating

effect when cured at an elevated temperature. Enhancement of thickening time of the

geopolymer slurries were achieved which allowed good pumpability for temperatures up

to 250 °F. A polycarboxymethyl superplasticizer and retarder was also tested in the same

paper and showed a significant retardation effect on thickening time. In the same study,

compressive strength, shear bond strength and durability of hardened geopolymer were

compared with Class H cement mixtures. Laboratory investigation showed that the

geopolymer developed more than 6,000 psi compressive strength by two weeks. The

Page 45: Copyright by Xiangyu Liu 2017

25

compressive strength of geopolymer mixtures increased with temperature, whereas

Portland cement slurries showed strength retrogression. Additionally, geopolymer

mixtures had comparable shear bond strength and better durability compared to Portland

cement at elevated temperatures.

Sugumaran (2015) also conducted a study that looked into the properties of class

F fly ash based geopolymer with different compositions and their acid resistance at

different acid concentrations for oilwell cementing. According to their results, at 60 °C

curing temperature, the optimum compressive strength was achieved when the water-to-

fly ash ratio was kept at 0.3 and when 12 M sodium hydroxide solution was used at an

alkaline-to-fly ash ratio of 0.4. The fly ash geopolymer cement was found to be more

resistant to acid attack when exposed to up to 10% sulfuric acid solution at room

temperature.

Pershikova et al. (Pershikova et al., 2012, 2011; Porcherie et al., 2011; Porcherie

and Pershikova, 2010) patented a pumpable geopolymer formulation for oilfield

application. The composition of the geopolymer was comprised of an aluminosilicate

source, an alkali activator, a rheology modifier containing aluminum compounds (such as

bauxite, aluminum salts aluminum oxide), a strength reinforcing agent (such as fiber or

magnesium silicate), and a filler material for density control. The example slurries

disclosed in the patents showed good pumpability, good mechanical strength, and low

water permeability (< 80 μD) at temperatures up to 90 °C and at a pressure of 3,000 psi.

The thickening time was manipulated by altering the silicon versus aluminum ratio or by

adding lithium compounds.

Overall, these preliminary studies evaluated the strength, rheological properties,

thickening time and acid resistance of geopolymer at downhole conditions. The

performance of geopolymer indicated by these preliminary studies certainly demonstrates

Page 46: Copyright by Xiangyu Liu 2017

26

great potential for using geopolymer as an alternative to current oilwell cement, as it

addresses many of the requirements for cementing a typical oil / gas well. Improved

performance is to be expected once the formulation of geopolymer is optimized and when

the admixture chemistry is more developed for geopolymer mixtures.

Well abandonment 2.5.2

When a well is decommissioned at the end of its life-cycle, usually a set of

cement plugs are placed in the well to maintain and guarantee isolation between

geological layers (Figure 2.5). The primary importance of cement plugs in an abandoned

well is to prevent contamination of groundwater aquifers by hydrocarbon seepage or

formation brine coming from below the aquifer. Contamination during plug cementing

has long been recognized as a serious problem for well abandonment (Nelson and

Guillot, 2006). In abandonment plug operations, a considerable amount of drilling mud

can exist and mix with cement slurry. Consequently, the setting time and mechanical

properties of the cement plug could vary significantly from the designed values, leading

to a highly compromised plug cementation.

Seeking a better well abandonment strategy, previous work has evaluated the use

of Class C fly ash slurry for plugging abandoned wells using coiled tubing (Shah and

Cho, 2001; Shah and Jeong, 2003). Since Class C fly ash has sufficient calcium and

aluminosilicate content, the slurry was activated with pure water instead of an alkaline

solution. Five fly ashes were tested for their thickening time, rheology, shear bond

strength, hydraulic bond strength and gas permeability. The experimental results revealed

that Class C fly ash slurries had sufficient thickening time (longer than 2 hours) and

could be pumped through coiled tubing for well abandonment. The viscosity of the

Page 47: Copyright by Xiangyu Liu 2017

27

slurries decreases as the temperature increases. Shear bond strength values ranged from

100 psi to 1,130 psi among different fly ash sources. The magnesium oxide content in the

fly ash provided good swelling characteristics and increased the shear bond strength. The

hydraulic bond strength values ranged from 1,500 psi to 2,100 psi, comparable to a

typical Class H cement sample under similar curing conditions. The gas permeability of

the geopolymer sample were of the same order of magnitude as Class H cement, ranging

from 0.062 md to 0.197 md, indicating minimal gas migration through the hardened plug.

Figure 2.5 - Cement plugs in an abandonment well (image adopted from Global CCS

Institute)

In another study, Khalifeh et al. (2016, 2015, 2014) investigated Class C fly ash-

based geopolymer slurries that were activated by silicate/hydroxide mixed solutions for

Page 48: Copyright by Xiangyu Liu 2017

28

plug and abandonment (P&A) operation. Thickening time test, rheological properties,

bulk shrinkage, uniaxial compressive strength and sonic compressive strength buildup

were evaluated. In the thickening time test, the geopolymer slurry showed high plateau

consistency around 50 Bc before the consistency finally reached 100 Bc in 40 minutes.

The slurries exhibited a viscous nature with Newtonian-like viscosity and fast gelation. A

maximum bulk shrinkage of 2% was observed for the samples that were activated by

10M NaOH and cured at 125 °C and 5000 psi.

Most importantly, empirical correlations were developed for the first time for

testing geopolymer slurries in an Ultrasonic Cement Analyzer (UCA). An UCA measures

the travel time of ultrasonic waves through a cement sample while it cures under

temperature and pressure. The transit time is correlated to compressive strength using

empirical relationships that were originally established for OPC. An UCA test protocol

has been included in the API Recommended Practice (API RP 10B-2, 2010) as

nondestructive testing method to evaluate cement strength. Such empirical correlations,

however, cannot be applied indiscriminately to other cementitious materials, such as

geopolymer, due to distinct structural differences that affect the ultrasonic travel times.

Khalifeh et al. (2014, 2015) developed a set of equations for geopolymers in an UCA test

by correlating data from unconfined compressive strength (UCS) for specimens that were

prepared with different concentrations of NaOH activator at 87 °C and 125 °C. Table 2.1

shows the acquired empirical correlations.

Page 49: Copyright by Xiangyu Liu 2017

29

Table 2.1 - Empirical correlations for geopolymer specimens prepared with various

NaOH concentrations at two different temperatures. Specimens were cured under a

pressure of 5,000 psi (Khalifeh et al. 2014)

Concentration

of NaOH

(M)

Temperature

(°C)

Empirical Correlations

x: transit time (μsec/in); y: Compressive strength (psi)

6 125 y = -769.81x2 + 13,151x - 48,085

8 125 y = -2,654.8x2 + 50,357x - 229,847

10 125 y = -10,062x2

+ 191,272x – 893,329

6 87 y = -2,154.8x2 + 43,332x - 211,942

8 87 y = -1,144.8x2 + 18,924x - 68,972

10 87 y = 490.19x2 - 13,529x + 88,881

Lost circulation control 2.5.3

Lost circulation, in particular, is a serious well construction problem, and

geopolymer offers great potential to treat such scenarios. If a highly permeable formation,

cavernous formation or fractures are encountered while drilling or cementing, drilling

fluids or cement slurries may totally or partially be lost into these zones instead of

returning back to the surface (Figure 2.6), i.e. circulation is “lost”. Such scenarios will

first be treated with lost circulation materials (LCM) such as particulates and fibers to

temporarily plug the lost zones. If unsuccessful, cement plugs can be placed and slightly

squeezed to seal and consolidate the formation. Note that the probability of a successful

Page 50: Copyright by Xiangyu Liu 2017

30

cement squeeze job for lost circulation control is usually low, often because of

incompatibility between OPC and OBM/SBM used for drilling.

(a) Partial loss (b) Total loss

Figure 2.6 - Lost circulation scenario with (a) partial loss, and (b) total loss (figure

adopted from petrowiki.org)

Miller et al. (2013) developed a geopolymer-based pill as an engineered solution

to treat lost circulation problems. In this approach, an aqueous alkali alumino silicate

(AAAS) was formulated to serve as a chemical sealant. AAAS is a pre-primed silicate

that is in liquid form. When AAAS is subjected to high shear or a reduced pH

environment, it can be triggered to polymerize and form a solid crystalline phase in a

controllable manner. In the study, AAAS was mixed with OBM to form a highly viscous

Flow Flow

Page 51: Copyright by Xiangyu Liu 2017

31

fluid prior to solidification for controlling circulation losses. Setting time was controlled

by adding propylene carbonate into the OBM. When higher compressive strength is

desirable to seal and block a fracture, the AAAS can be formulated into a cementitious

material by adding slag or fly ash. Experimental results showed that a cement plug

containing AAAS and class F fly ash developed good early compressive strength in 3

hours at temperatures higher than 25 °C.

MUD SOLIDIFICATION TECHNIQUE 2.6

Techniques for mud solidification were first introduced to the industry in the late

1980s / early 1990s. As mentioned in the introduction chapter, the conversion of drilling

mud into a cement offers a lot of advantages compared to a conventional cementing

process (Hale and Cowan, 1991):

Better drilling mud removal.

Improved compatibility between drilling mud and cement slurries, with less

negative effect from any mud contamination.

Ability to convert any undisplaced drilling mud and mud filtercake into cement.

Re-cycle drilling mud in-situ.

BFS-based drilling mud solidification 2.6.1

There are multiple approaches to converting a drilling mud into a cementitious

slurry. For instance, Wilson et al. (1989) proposed to add special copolymer dispersants

and accelerators into drilling mud and then add cement to achieve a settable

characteristic. A more developed method that was extensively studied and even field

tested was the slag-based mud solidification method, also called as the “universal fluid”

Page 52: Copyright by Xiangyu Liu 2017

32

(UF) system, in which a WBM was diluted and treated with a chemical activator. BFS

was mixed into the pretreated drilling mud as the hydraulic material to consolidate the

mud (Cowan et al., 1992; Hale et al., 1995; Hale and Cowan, 1994, 1991, Nahm et al.,

1995, 1994, 1993). This method has been deployed for use in several cementing

operations including primary cementing, temporary abandonment plugs and sidetracking

plugs in the last two decades. Laboratory evaluations and field tests showed that the BFS-

based WBM-to-cement conversion technique not only provided good-quality cement, but

also effectively solved gas-migration and lost-circulation problems (Leimkuhler et al.,

1994; Pessier et al., 1994; Song et al., 2000; Wu et al., 1998).

There are several limitations associated with the BFS-based mud solidification

method. In laboratory testing, micro-cracking has been observed in the set cement and the

cause of such cracking remains unclear. Benge and Webster (1994) believed that the set

cement containing BFS and drilling mud have no fibers or crystals connecting the grains,

making the set material less ductile and giving it a tendency to crack. Mueller and

Dickerson (1994) investigated the physical and mechanical properties of the set drilling

mud and observed that the brittleness and stress-cracking of the slag-converted drilling

muds are heavily influenced by factors such as the type and concentration of the

activator. Over-activation of the system could result in detrimental effects such as

abnormal thickening time, excessive heat buildup, high viscosities and diminished

compressive strength.

Another complex factor in the design of a BFS-based mud solidification system is

the varied composition of the drilling mud. The cuttings and chemicals that are added to

the drilling mud continually changes when the bit penetrates different formations to

maintain desired mud properties. Such compositional changes may strongly affect the

cement properties once the mud is converted to a cement slurry. To properly design a

Page 53: Copyright by Xiangyu Liu 2017

33

specific job, a portion of the mud samples has to be isolated and shipped to a laboratory

long before the cementing. This is a limitation that has to be overcome with any mud-

solidification technique.

Despite the progress made on utilizing BFS-based material for WBM

solidification, the same technique has not yet been developed and applied on SBMs or

OBMs. The industry’s dependence on SBM has gone up significantly in the past two

decades. In deepwater drilling especially, SBMs are preferred over WBM due to the

superior performance properties in borehole stabilization, lubrication and improved rate

of penetration. SBMs and OBMs are expensive and are typically recycled whenever

possible to the point where the mud can no longer be reused. The disposal of these NAFs

is governed by strict environmental regulations. Offshore operations, in particular, have

to follow regulations including the OSPAR Commission in the northeastern Atlantic and

the Environmental Protection Agency (EPA) in the Gulf of Mexico. According to EPA,

NAFs as whole mud cannot be discharged in the ocean and NAF (SBM only) left on

cuttings has to be reduced to specific levels before cuttings can be discharged. If a mud

solidification technique becomes feasible, then NAFs that can no longer be recycled may

be processed and consolidated on-site with no need to transport it back to land.

Geopolymer-based solidification for organic waste disposal 2.6.2

The solidification / stabilization (S / S) of industrial wastes using cement-based

materials is a widely applied technique (Wiles, 1987). Traditional cements, however,

have proven to be ineffective for immobilizing organic wastes, as they may inhibit

cement hydration and in general do not chemically bond with the binder (Pollard et al.,

1991; Trussell and Spence, 1994). In a study that investigated the disposal of radioactive

Page 54: Copyright by Xiangyu Liu 2017

34

waste in the nuclear industry, Cantarel et al. (2015) discovered that geopolymer acts as a

strong candidate to encapsulate liquid oil waste effectively. This technique involves the

mixing of oil in an alkali silicate solution to form an emulsion, followed by mixing with

an alumino-silicate source (e.g. metakaolin) to allow for the setting of a geopolymer. The

experimental results showed that metakaolin-based geopolymers can successfully

encapsulate oil waste, with oil droplets uniformly distributed throughout the matrix of the

hardened binder. Good rigidity and mechanical strength were observed. Leaching tests

showed very limited release of oil from the composite material. The geopolymer/oil

composites have been successfully tested at up to 20% by volume. Based on these results,

geopolymer should also have potential to effectively solidify NAFs in a similar manner.

SUMMARY 2.7

Overall, geopolymer cements have significant potential for improved long-term

zonal isolation and well integrity as indicated by previous research studies. Many

geopolymer systems that have been developed require high temperature curing, which

naturally occurs in oil / gas wells. Very limited progress has been achieved in the

following areas limiting its application in oil / gas industry: (1) verifying the performance

of geopolymer paste at various pressure and temperature conditions; (2) study the

compatibility between geopolymer and drilling mud for displacement efficiency; (3)

develop effective methods or identify admixtures to control rheology and workability of

the slurry paste. The current work will focus on addressing these limitations.

Page 55: Copyright by Xiangyu Liu 2017

35

Chapter 3: Materials and Methods

This chapter provides a brief overview about the materials and experimental

procedures. Section 3.1 provides general information about all the materials that were

used in this study. Section 3.2 presents all the experimental methods that were used to

evaluate the geopolymer slurries under various temperature and pressure conditions.

RAW MATERIALS 3.1

Fly ash composition 3.1.1

To create a geopolymer slurry, an aluminosilicate precursor powder is mixed with

an alkaline activating solution. As mentioned in the literature review section, fly ash-

based AAM delivers a more favorable rheology than metakaolin-based binders at a much

lower cost. Thus, in this study, three different class F fly ashes were selected as the

aluminosilicate precursor (ASTM C618-18, 2015). The main composition of each fly ash

(FA) is shown in Table 3.1. Fly ashes FA1 and FA3 were heat treated through a thermal

beneficiation process to increase the reactive component of the ashes. FA3 was passed

through a number 100 sieve to remove large particles. Geopolymer slurries that were

formed from fly ashes FA1, FA2 and FA3 will be referred to in the following as G1, G2

and G3 respectively.

Page 56: Copyright by Xiangyu Liu 2017

36

Table 3.1 - Composition of fly ashes (weight %)

Beneficiated SiO2 Al

2O

3 Fe

2O

3 (SiO

2+Al

2O

3+Fe

2O

3) CaO MgO

Alkalis

(Na2O+0.658K

2O)

SO3

FA1 Yes 49.9 25.3 15.1 90.3 3.0 0.91 0.73 0.44

FA2 No 48.8 20.3 16.33 85.48 6.1 1.1 0.44 2.07

FA3 Yes 53.5 27.0 10.4 90.86 2.7 1.0 0.5 0.24

Particle size distribution (PSD) of fly ashes 3.1.2

The particle size distribution of each fly ash was measured using a Mastersizer

2000 particle size analyzer with a Hydro MU 2000 (Malvern, Worcestershire, United

Kingdom) wet dispersion unit. The specific gravity of fly ash was 2.56 and was obtained

from the manufacturer. Refractive index was picked to be 1.56 based on literature data

(Jewell and Rathbone, 2009). The absorption value was chosen such that the residual

weighted PSD calculated by the Mastersizer software was less than 1. Fly ash particles

were dispersed in isopropyl alcohol (IPA) to prevent hydration throughout the

measurement. The refractive index of IPA was assumed to be 1.39. During testing, the

particles were circulated at a pump speed of 2000 rpm and were sonicated with an

ultrasonic probe for 30 seconds. The particles were then circulated in the particle size

analyzer at a pump speed of 2000 rpm without sonication. Figure 3.1 shows the average

PSD based on five measurements for all the three fly ashes. As shown in the figure, all

the fly ashes showed a d50 between 10-20 µm (meaning 50% of the particles are smaller

than 10-20 µm). For PSD of FA3, a bimodal distribution was observed with peaks at 11.5

µm and 65.0 µm, whereas the data for FA1 and FA2 exhibited a unimodal distribution

with peaks at 14.5 µm and 16.0 µm respectively.

Page 57: Copyright by Xiangyu Liu 2017

37

Figure 3.1 - Volume weighted particle size distribution of all fly ash particles. For the x-

axis, 40 size intervals were generated logarithmically between 0.1 and 1000. Y-axis

shows the volume fraction of particles between those sizes.

Activator 3.1.3

The alkaline activator solution was either hydroxide (sodium hydroxide) or

silicate, where both solid-form (SSS) and liquid-form sodium silicate (LSS) were used.

For hydroxide activation, 8 M NaOH solution was selected as the primary activator. It

has been recurrently reported in literature that 8 M NaOH can successfully activate class

F fly ash-based geopolymer (Aughenbaugh, 2013; Duxson et al., 2006; Fernández-

Jiménez et al., 2006a; Fernández-Jiménez and Palomo, 2005b; Provis and Deventer,

2009). For comparison purposes, 4 M and 6 M NaOH were also tested. NaOH solution

was prepared by weighing reagent grade NaOH pellets and dissolving them in ultrapure

0

1

2

3

4

5

0.1 1 10 100 1000

Volu

me

(%)

Particle Size (μm)

FA1

FA2

FA3

Page 58: Copyright by Xiangyu Liu 2017

38

water (resistivity of 18 MΩ-cm) to the desired concentration. The solution was cooled

down to room temperature prior to use. The activator solution to fly ash ratio was

proportioned to be 0.485 based on preliminary tests for optimum strength and

workability.

For LSS activation, sodium hydroxide pellets and deionized water were added to

a commercially available LSS solution to achieve SiO2/Na2O weight ratios of 0.12, 0.24

or 0.48 (Table 3.2). For SSS activation, the SSS powder was blended with fly ash, and

was activated with a pre-mixed sodium hydroxide solution containing an appropriate

amount of sodium ions to achieve the same SiO2/Na2O ratios of 0.12, 0.24 or 0.48. These

three ratios, although are lower than the values recommended in literature, were selected

because the slurries showed the greatest potential for pumpability based on viscometer

and consistometer measurements. The Na2O/FA ratio was fixed at 0.1 for all the mixes

shown in Table 3.2, leading to a constant Na2O/Al2O3 ratio for any given fly ash. The

geopolymer slurry was hand-mixed with a spatula to incorporate all wet and dry

components, and was then mixed with a paddle stirrer at 480 rpm for 30 seconds.

Page 59: Copyright by Xiangyu Liu 2017

39

Table 3.2 - Geopolymer mix design

Notation Activator SiO2/Na2O water/solid Na2O/fly ash

LSH-8M Liquid Sodium Hydroxide (8 M) - 0.33 0.1

LSS-0.12 Liquid Sodium Silicate 0.12 0.36 0.1

LSS-0.24 Liquid Sodium Silicate 0.24 0.36 0.1

LSS-0.48 Liquid Sodium Silicate 0.48 0.36 0.1

SSS-0.12 Solid Sodium Silicate 0.12 0.36 0.1

SSS-0.24 Solid Sodium Silicate 0.24 0.36 0.1

SSS-0.48 Solid Sodium Silicate 0.48 0.36 0.1

Admixtures 3.1.4

Limestone dust was used as a stability-enhancing additive (A) up to 2% by weight

of fly ash when instability was observed in the slurry. The particle size distribution of

limestone dust was also measured with the Mastersizer 2000 particle size analyzer and

the PSD data is shown in Figure 3.2. The d50 of limestone dust is approximately 3 µm and

d90 is 12 µm, indicating the limestone particles are smaller in comparison with the fly ash

particles (d50 between 10-20 µm).

Page 60: Copyright by Xiangyu Liu 2017

40

Figure 3.2 - Volume weighted particle size distribution of limestone dust (admixture A)

Portland cement 3.1.5

For comparison, a Portland slurry (P1) consisting of Class H Portland cement and

38.5% by weight of cement (bwoc) water was prepared in accordance with API standard

RP 10B-2 (2010). For Portland slurries, sodium glucoheptonate retarder (R) was added at

a dosage of 0.025% bwoc for downhole rheology and thickening time measurements.

Alkali-activated slag 3.1.6

For the sake of comparison, properties of an alkali-activated slag cement were

measured using pressure transmission tester and mercury intrusion porosimetry. The slag

slurry was formed with a grade 120 blast furnace slag and 50% mix water by weight of

slag (bwos). The alkali activator contains 1.5% (bwos) NaOH and 4.5% (bwos) dense

soda ash. When preparing the slurry, NaOH pallets were first dissolved in water in a high

0

1

2

3

4

5

0.1 1 10 100 1000

Vo

lum

e (%

)

Particle Size (µm)

Admixture A

Page 61: Copyright by Xiangyu Liu 2017

41

shear blender. The slag powder and soda ash were premixed and added to the blender.

Once all dry ingredients and liquids were combined together, the slurry was mixed at

12,000 rpm for 35 seconds.

SBM / OBM 3.1.7

The solidification of drilling mud was achieved by directly blending drilling mud

and geopolymer slurry with a paddle stirrer at 480 rpm for 30 seconds and allowing it to

set for various periods of time. In this study, 5-20% of the Portland slurry and 10-40% of

the geopolymer slurry was replaced with SBM / OBM by volume. A SBM (S1) that is

commonly used in the Gulf of Mexico was selected to conduct an in-depth study to

understand various properties of the geopolymer / mud hybrid. Another SBM (S2) and

two spent mineral oil-based muds (O1 and O2) were also solidified using 8 M NaOH

activated fly ash-based geopolymers and their compressive strength was measured. For

the sake of brevity, when 10% of the G1 slurry is replaced with S1, the slurry will be

referred to as G1S1-10. The properties of the four drilling muds used in this study are

shown in Table 3.3 and Table 3.4.

Page 62: Copyright by Xiangyu Liu 2017

42

Table 3.3 - Properties of the SBMs (S1 & S2)

S1 S2

Density (ppg) 9.7 9.7

Water (%) 27% 25%

Oil (%) 63% 59%

Solid (%) 10% 16%

70 °F 120 °F 150 °F 70 °F 120 °F 150 °F

Viscometer Dial

Reading

600 rpm 94 65 48 70 50 36

300 rpm 58 40 28 44 29 21

200 rpm 43 28 20 32 21 15

100 rpm 25 18 15 21 14 10

6 rpm 11 10 9 9 6 4

3 rpm 10 9 8 9 6 4

Plastic Viscosity (cp) 36 25 20 26 21 15

Yield Point (lbf/100sqft) 22 15 8 18 7 6

S: synthetic-based mud

Page 63: Copyright by Xiangyu Liu 2017

43

Table 3.4 - Properties of the OBMs (O1 & O2)

O1 O2

Density (ppg) 12 11.54

Water (%) 24% 18%

Oil (%) 56% 58%

Solid (%) 20% 24%

70 °F 120 °F 150 °F 70 °F 120 °F 150 °F

Viscometer Dial

Reading

600 rpm 349 239 148 349 213 133

300 rpm 317 151 90 246 122 78

200 rpm 241 117 68 178 90 58

100 rpm 156 79 42 104 56 36

6 rpm 43 20 7 23 14 10

3 rpm 34 16 6 18 12 9

Plastic Viscosity (cp) 32 88 58 103 91 55

Yield Point (lbf/100sqft) 285 63 32 143 31 23

O: mineral oil-based mud

METHODS 3.2

Compressive strength 3.2.1

The unconfined compressive strength (UCS) of hardened OPC slurries and

geopolymer slurries was obtained by crushing 2” × 2” cubes that were cured in a

pressurized curing chamber (API RP 10B-2, 2010). The curing temperature was

increased from room temperature to bottom hole circulating temperature (BHCT) of 125

°F in 2 hours and to bottom hole state temperature (BHST) of 170 °F in 10 hours. The

Page 64: Copyright by Xiangyu Liu 2017

44

curing pressure was maintained at bottom hole pressure (BHP) of 3,000 psi. All

compressive strength values reported here are based on the average of 8 cubes. The error

bars in the plots represent one standard deviation. Please note that in order to apply a

confining pressure, a different sample size and sample shape was used for the self-healing

tests as mentioned in a later section.

Ultrasonic compressive strength 3.2.2

The ultrasonic compressive strength of geopolymers was obtained with an UCA

that is commercially manufactured for testing Portland cement. As mentioned in the

background chapter, strength of cement can be estimated ultrasonically with UCA. With

this method, the travel time of ultrasonic energy through a cement sample was measured

and empirical correlations were used to estimate the strength values. In the present study,

a geopolymer sample was tested in UCA with both built-in correlations that were

developed for Portland cement, and with correlations that were developed by Khalifeh et

al. (2014). To test a geopolymer sample in UCA, the geopolymer slurry was formed with

FA1 and 8 M NaOH solution. The slurry was placed in UCA upon mixing. Temperature

was ramped from room temperature to 189 °F (87 °C) in two hours and pressure was

maintained at 5,000 psi (34.5 MPa).

Rheology 3.2.3

Rheological properties of slurries were measured at room temperature with a

typical viscometer with F1 spring and R1B1 rotor configuration following the API

standard API RP 10B-2 (2010). When measuring the downhole rheological properties,

the slurry was pre-conditioned in a HPHT consistometer where the temperature and

Page 65: Copyright by Xiangyu Liu 2017

45

pressure values were ramped to 125 °F and 3,000 psi, respectively, in 80 minutes and

stirred for an additional 30 minutes. At the end of the pre-conditioning period, the slurry

was taken out of the consistometer and poured into a pre-heated viscometer cup to

measure the rheological properties.

Thickening time 3.2.4

Thickening time, or pumping time, of the slurry was measured with a HPHT

consistometer following the API standard RP 10B-2 (2010). The ramping schedule was

80 minutes to BHCT of 125 °F and BHP of 3,000 psi.

Confined compressive strength 3.2.5

The confined compressive strength of hardened slurries was measured using a

triaxial load frame and a confinement vessel. Cylindrical samples(2 in. length × 1 in.

diameter) were prepared and cured in a water bath at 170 °F for 7 days. Prior to testing,

all samples were cut and a surface grinder was used to ensure the top and bottom surfaces

of the samples were smooth and parallel to each other as well as perpendicular to the

circumference of the cylinder. This was done to ensure the load was being applied along

the axis of the cylinder. Each sample was placed between two stainless steel endcaps and

wrapped with heat shrink tubing, which served as an impermeable barrier to the confining

fluid. Axial and radial strain gauges were placed to measure the triaxial displacement.

The confining stress (σ3), or confining pressure (Pc), acts isotropically on the sample.

Hence, the total axial stress (σ1) is a summation of the axial stress applied by the piston

and the confining stress. The difference between the axial stress and the confining stress

is defined as the differential stress (σ1- σ3).

Page 66: Copyright by Xiangyu Liu 2017

46

At the start of each triaxial test, the sample was first loaded with a differential

stress of 30 psi. The sample was then loaded hydrostatically to the target confining

pressure of either 100 psi or 500 psi. The confining pressure was held constant for 10

minutes to ensure the change in axial strain due to creep was negligible compared to the

strain rate during measurement. After the system reached equilibrium, the differential

stress was increased at a constant axial strain rate of 10-5

per second. Loading was

continued until the post-peak regime was adequately defined. Two samples were

measured for each slurry type at both confining pressures. The stress-strain curves were

plotted showing axial and radial strains as a function of differential stress. Elastic

constants were determined over the linear sections of the stress-strain curves using the

following equations:

E =∆σa

∆εa (3.1)

ν =∆εr

∆εa (3.2)

where,

E = Young’s Modulus (psi),

ν = Poisson’s ratio (dimensionless),

σa = Axial stress (psi),

εa = Axial strain (inch/inch), and,

εr = Radial strain (inch/inch).

Page 67: Copyright by Xiangyu Liu 2017

47

Unconfined self-healing test 3.2.6

Self-healing tests were conducted to study the re-healing capability of geopolymer

samples that were subjected to pre-loading. Cylindrical samples (4 in. length × 2 in

diameter) were cured in a water bath at 170 °F. On day 7, four samples were loaded to

failure to measure the UCS. Four samples each were loaded to either 30%, 50% or 70%

of the 7-day UCS, which likely causes varying levels of damage to the internal

microstructure. These partially damaged samples were placed in the water bath at 170 °F

for an additional 21 days. At the end of the re-healing period, the samples were loaded to

failure. By comparing the ultimate strength of the pre-damaged samples to the 28-day

UCS of an undamaged sample, the self-healing capability could be evaluated. Similar

measurements were conducted on samples that were pre-damaged to either 30%, 50% or

70% of the 28-day UCS and re-healed for an additional 28 days. The error bars in the

plots represent one standard deviation. The self-healing capability of the geopolymer

slurry was compared with that of Portland slurry.

Confined self-healing test 3.2.7

The confined self-healing test was performed with the triaxial load frame and a

confinement vessel. Cylindrical samples (4 in. length × 2 in. diameter) were cured for 7

days in a water bath at 170 °F. The samples were cooled down and loaded under 500 psi

confining pressure at a constant axial strain rate of 10-5

per second. The samples were

unloaded after 2% axial strain was reached, and were allowed to cure at 170 °F in a water

bath for an additional 21 days. After this re-healing period, the samples were loaded at

the same axial strain rate until failure. For each sample, the peak stress at 28 days was

compared to the peak stress at 7 days to evaluate the self-healing capability of Portland as

well as geopolymer slurries.

Page 68: Copyright by Xiangyu Liu 2017

48

Pipe-in-pipe shear bond strength test 3.2.8

The cement-to-pipe shear bond strength was measured with a customized pipe-in-

pipe experimental set-up, as shown in Figure 3.3. A steel bar (8 in. length × 1 in.

diameter) was first polished with a medium grade emery cloth and then placed in a 3 in.

inner diameter plastic pipe. Cement slurry was poured into the plastic pipe to a depth of 6

inches around the steel bar. The sample was placed in a humidity controlled environment

at 170 °F for 7 days. After curing, the bottom plate was removed. The sample was placed

on a hollow base and the steel bar was pushed out of the cement sheath. The peak force

divided by the contact surface area of the steel bar yields the shear bond strength of the

cement-to-pipe interface.

Figure 3.3 - Schematic diagram of the pipe-in-pipe shear bond strength set-up

8” 7” 6”

1”

3”

Steel Bar

Cement

Steel Base

Plate

Apply vertical load at

constant rate

Page 69: Copyright by Xiangyu Liu 2017

49

Pressure transmission test 3.2.9

The hydraulic conductivity (HC) of hardened geopolymer was indirectly

measured via pressure transmission testing. For the purpose of comparison, a Portland

cement slurry and an alkali-activated slag cement were also tested.

Pressure transmission test (PTT) was introduced to the oil / gas industry for

testing the tendency of a fluid filtrate, applied at overbalance (differential) pressure, to

invade the sample matrix and elevate the near wellbore pore pressure (Bol et al., 1994;

Darley, 1969; Stowe et al., 2001; Tare and Mody, 2000; van Oort, 1994). The schematic

of the pressure transmission test set-up is shown in Figure 3.4. A detailed test protocol

has been described in literature (van Oort et al., 1996, 2016). Simply explained, in a

pressure transmission test, a testing fluid is injected at a differential pressure across the

sample, allowing the fluid to transmit and penetrate through the sample. The downstream

pressure buildup due to pressure transmission is recorded and the data is processed to

obtain the hydraulic conductivity.

In this study, one sample each was tested for geopolymer, Portland cement, and

slag. The samples (2 in. length × 1 in. diameter) were cured at 170 °F for 28 days. All the

samples were submerged in water throughout curing and sample preparation period until

testing. Figure 3.5 shows a picture of the sample assembly. Porous aluminium frits were

placed on each end of the hardened cement sample to ensure uniform fluid distribution.

The sample stack was then placed between two flow heads, and was wrapped with heat

shrink tube and a viton sleeve. The sample was placed in an iso-static coreholder and

subjected to confining pressure of 1000 psi in a temperature controlled oven operated at

100 °F (35 °C). Tests were performed with two distinct cycles: in the first cycle 3%

artificial seawater (ASW) was used to characterize the conductivity of the artificial pore

fluid, and in a second cycle (after re-equilibrating the sample to initial conditions) light

Page 70: Copyright by Xiangyu Liu 2017

50

mineral oil (LMO) was used. In both cycles, the upstream pressure was maintained at 300

psi and the downstream pressure was set to 50 psi to create a differential pressure of 250

psi.

Figure 3.4 - Schematic of the pressure transmission test set-up

Hardened cement

Confining fluid

Downstream (50 spi)

Viton Sleeve

Upstream (300 psi)

Porous top frit

Porous bottom frit

Top flow head Iso-static coreholder

Bottom flow head

Page 71: Copyright by Xiangyu Liu 2017

51

Figure 3.5 - Picture of sample assembly in pressure transmission test

Simply explained, the downstream pressure build-up behavior due to pressure

transmission through the sample is similar to the charging of a capacitor in a resistor-

capacitor (RC) circuit. The pressure transmission is essentially described as follows (van

Oort et al., 2016):

P(l, t) − Po

Pm − Po= 1 − exp [−

Akt

μβVl] (3.3)

Where

Po = initial pore pressure (Pa),

Pm = upstream fluid pressure (Pa),

P(l, t) = downstream pressure transmission at sample end as a function of time

(Pa),

l = sample length (m),

A = sample cross-sectional area (m2),

V = volume of downstream reservoir (m3),

β = fluid compressibility (Pa-1

),

μ = fluid viscosity (Pa·s),

Page 72: Copyright by Xiangyu Liu 2017

52

k = relative permeability of sample (m2).

The viscosity µ and compressibility β of the test fluid are generally unknown (or

provided to be indeterminate via calculations). The hydraulic conductivity (HC = 𝑘/𝜇𝛽

(m2/s)) is calculated for each of the cycles with ASW and LMO. Rearranging equation

(3.3):

𝐻𝐶 =k

μβ=

Vl

Atln

Pm − Po

Pm − P(l, t) (3.4)

The left side of this equation is hydraulic conductivity, which measures the

diffusivity of the fluids through the samples. Both the hydraulic conductivity

measurements are then compared to yield a “delay factor”:

Delay Factor =Hydraulic Conductivity (LMO)

Hydraulic Conductivity (ASW) (3.5)

The data obtained from PTT is processed in accordance with equation (3.4) and

fitted with a least-squares linear fit. From the slope of the fitted lines obtained for the

pore fluid (ASW) cycle and the test fluid (LMO) cycle, a delay factor can be calculated

using equation (3.5). The PTT data can be used to compare pressure transmission delays

(pore fluid vs. test fluid) and thus characterize the ability of fluid pressure to transmit

through a sample. The delay factor value indirectly reflects by how much the dynamic

pressure transmission can be slowed down. Overall, the pressure transmission test result

demonstrates the sample’s diffusive capability and thus exhibits its permeability

characteristics.

Page 73: Copyright by Xiangyu Liu 2017

53

Porosity and pore size distribution 3.2.10

The porosity and pore size distribution of neat hardened geopolymer, Portland

cement, and alkali-activated slag were measured with mercury intrusion porosimetry

(MIP) test using AutoPore IV Automated Mercury Porosimeter (from Micromeritics

Instrument Corporation). A mercury porosimeter characterizes a material’s porosity (ϕ)

by applying various levels of pressure to a sample immersed in mercury. The pressure

required to intrude mercury into the pores of a sample is inversely proportional to the size

of the pores, and thus the pore size distribution of the porous material can be

characterized.

For each slurry formulation, a cylindrical sample (0.5 in. length × 1 in. diameter)

was cured at 170 °F for 28 days, and was then dried in an oven at 200 °F for at least 24

hours to ensure the weight reaches an equilibrium. The sample was then immersed in

mercury in a pressure-sealed chamber that was connected to a capillary stem with a

capacitor. The pressure in the chamber was increased incrementally from 0 to 60,000 psi.

This pressure range corresponds to pore sizes from a few nanometers (high pressure) to

hundreds of micrometers (low pressure). Each pressure step was maintained until the

volume equilibrium was reached. The bulk sample volume can be determined by the

capacitance when immersed in mercury. The pore volume and porosity can be computed

based on the capacitance measurements. The volume of mercury injected and the pore

volume at each incremental pressure can be used to construct pore size distribution.

Detailed MIP experimental protocols and data interpretation methods can be found in

Bear (1972), Peters (2012), and Purcell (1949). The pore throat radius (rp) can be

obtained from mercury injection pressure (PHg) using the following equation:

rp =2σHg cos θHg

PHg

(3.6)

Page 74: Copyright by Xiangyu Liu 2017

54

where σHg is the air-mercury interfacial tension, θHg is the mercury contact angle.

The values for each parameter used in the pore size distribution and capillary pressure

calculation were summarized in Table 3.5.

In this study, the estimated porosity, pore area and average pore diameter of the

three samples were reported. The pore size distribution curves for each slurry were

plotted and compared.

Table 3.5 - Parameters used in the MIP measurement

Hg advancing angle 130 °

Hg receding angle 130 °

Air-Hg interfacial tension (N/cm) 0.485

Hg density (g/ml) 13.5335

Page 75: Copyright by Xiangyu Liu 2017

55

Chapter 4: Hydroxide Activation

This chapter primarily focuses on evaluating the ability of geopolymer to convert

SBM into a cementitious material. In order to develop the mud solidification method, it is

imperative to first evaluate the compatibility between geopolymer and SBM. The

geopolymer slurry was formed by blending class F fly ash with 8 M sodium hydroxide

activator. Once the enhanced compatibility between geopolymer and SBM was

confirmed, a mud solidification method was achieved by simply incorporating various

quantities of SBM into the geopolymer slurry. The critical path for development of the

geopolymer hybrid cement included the evaluation of both the fresh state properties (such

as rheology and thickening time) and the hardened state properties (such as compressive

strength). To validate the applicability of the proposed solidification method, the effect of

changing the aluminosilicate powder (fly ash) source and the molarity of the hydroxide

activator was studied. The effect of varying the synthetic/water ratio (SWR) and internal

brine composition of the SBM was also analyzed. Seawater was tested as the mixing

water to assess the applicability of this method in offshore deepwater operations. The

thickening behavior of the geopolymer hybrid was evaluated under various BHPs.

Finally, the newly proposed solidification method was tested on various types of SBMs

and OBMs, verifying the compatibility between geopolymer and NAFs.

CONTAMINATION RESISTANCE 4.1

As described in the introduction chapter, OPC in general is incompatible with

NAFs. Contamination of Portland cement slurries with SBM leads to a significant

decrease in compressive strength. Figure 4.1 shows the substantial drop in compressive

strength values of Portland cement by almost 3500 psi with 10% SBM contamination,

Page 76: Copyright by Xiangyu Liu 2017

56

whereas, geopolymer only had a strength reduction of 400 psi at the same contamination

level. Although neat Portland cement had much higher compressive strength than neat

geopolymer, at 20% SBM contamination the two curves intersected, showing equal

compressive strength. Further contamination displayed higher compressive strength for

geopolymer. In addition, the geopolymer can be designed with comparable compressive

strength values to OPC as mentioned in the background chapter. Therefore it is more

appropriate to compare the strength reduction in percentage rather than comparing the

actual strength value.

Figure 4.1 - Compressive strength values of hardened Portland cement slurry (P1) and

geopolymer slurry (G1) with SBM (S1) contamination (replacement by volume) at 24

hours, 170 °F and 3,000 psi.

0

1000

2000

3000

4000

5000

6000

7000

0% 10% 20% 30% 40% 50%

Com

pre

ssiv

e S

tren

gth

(psi

)

% of SBM

P1S1

G1S1

Page 77: Copyright by Xiangyu Liu 2017

57

For both OPC and geopolymer, the change in compressive strength from their

neat state is displayed by plotting the percentage of strength reduction. When neat

Portland cement slurry was contaminated with 5% SBM (replacement by volume), the

compressive strength decreased by 65% and 10% contamination resulted in a 70%

strength reduction (Figure 4.2). Conversely, the geopolymer slurry only had a 30%

strength reduction with 10% SBM contamination. The geopolymer sample retained

measureable compressive strength up to the maximum contamination level tested of 40%

SBM while the Portland slurries had no measurable strength from contamination level of

30% SBM.

Figure 4.2 - Normalized compressive strength of hardened Portland cement slurry (P1)

and geopolymer slurry (G1) with SBM (S1) contamination (replacement by volume) at

24 hours, 170 °F and 3,000 psi.

0%

20%

40%

60%

80%

100%

0 10 20 30 40 50

% o

f U

nco

nta

min

ated

Sam

ple

Com

pre

ssiv

e

Str

ength

% of SBM

P1S1

G1S1

Page 78: Copyright by Xiangyu Liu 2017

58

In addition to the issues with compressive strength, Portland cement slurries also

suffer from loss of pumpability (increase in viscosity) when contaminated with SBM. For

the sake of brevity, when 10% of the G1 slurry is replaced with S1, the slurry will be

referred to as G1S1-10. As shown in Figure 4.3 (a), Portland cement slurries became

progressively more viscous as the level of contamination increased. Above 20%

contamination with S1, Portland cement slurry became too viscous to be measured with a

typical viscometer using a F1 spring and a R1B1 rotor configuration. As shown in Figure

4.4, numerical simulation results have shown that high levels of cement contamination

occur at the leading edge of the cement slurry during displacement when interface

instability occurs (Enayatpour and van Oort, 2017). In a real life scenario, the increase in

viscosity caused by contamination could lead to serious problems including high

pumping pressure and unwanted fracturing of the well.

Geopolymer, on the contrary, showed a reverse rheological trend. As shown in

Figure 4.3 (b), neat geopolymer was too viscous and the upper limit of the viscometer

was reached above 100 s-1

shear rate. Typical neat geopolymer slurries can be placed into

formwork for civil engineering purposes, but cannot be pumped for petroleum

engineering purposes due to their viscous nature. The present work showed that by

adding S1 into geopolymer slurries, at room temperature the rheological properties were

greatly improved. At contamination levels of 30% and 40% the geopolymer slurries

showed rheological profiles that approached that of the neat Portland slurry, allowing a

pumpable geopolymer for well cementation. In addition, it was found that the rheological

properties of geopolymer could further be improved at higher temperatures. These

experimental results will be shown in the following section.

Both compressive strength and rheology measurements proved that geopolymer

slurries exhibit better contamination resistance compared to Portland cement slurries. In

Page 79: Copyright by Xiangyu Liu 2017

59

fact, because a large volume of SBM can be incorporated into geopolymer to form a

pumpable slurry with the ability to harden, it demonstrates the possibility of a mud-to-

cement conversion technique using geopolymer.

Figure 4.3 - Rheological properties of (a) Portland cement (P1) and (b) geopolymer (G1)

slurries replaced with various dosages of SBM (S1) by volume at 70 °F

0

50

100

150

200

250

0 100 200 300 400 500 600

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

P1S1-15

P1S1-10

P1S1-5

P1

(a)

0

50

100

150

200

250

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

G1 G1S1-10G1S1-20 G1S1-30G1S1-40 P1

(b)

Page 80: Copyright by Xiangyu Liu 2017

60

Figure 4.4 – Numerical simulation of drilling mud displacement with cement slurry. The

color bar shows the volume fraction of the cement slurry. The color gradient at the

interface indicates mixing of the two fluids (Enayatpour and van Oort, 2017)

In the remainder of this dissertation, SBM (or OBM) was purposely added to

geopolymer at various quantities. The resulting mixture will be referred to as geopolymer

hybrid. In a real life scenario, this geopolymer hybrid will be the cement slurry that will

be injected to a well. Since geopolymer demonstrates the ability to solidify drilling mud

at various levels of contamination, even if this geopolymer hybrid is further contaminated

at the leading edge, where non-uniform drilling mud contamination is most likely to

occur, the mixture should still have the ability to harden and develop compressive

strength.

In the following sections, properties including strength, rheology and thickening

time of the geopolymer hybrid will be discussed in detail, evaluating the capability of

geopolymer to solidify NAFs for both primary and remedial cementing applications.

Page 81: Copyright by Xiangyu Liu 2017

61

PROPERTIES OF GEOPOLYMER HYBRID CEMENT 4.2

Compressive Strength 4.2.1

The compressive strength development from 16 hours to 14 days is presented in

Figure 4.5 for hardened geopolymer (G1) and geopolymer/SBM hybrids (G1S1) at up to

40% replacement by volume. For samples G1, G1S1-20, and G1S1-30, compressive

strength was measured as early as 16 hours after loading the cubes in the curing chamber.

At 16 hours, G1S1-40 sample was omitted from data collection due to the low strength

measured for the G1S1-30 sample. The strengths were also measured at 24 hours, 48

hours, 72 hours, 7 days, and 14 days. For G1 and G1S1-20 samples, the compressive

strength developed at a fairly rapid rate during the first three days. The G1S1-20 sample

had a 14-day compressive strength slightly above 2000 psi, which is close to half of that

of the neat G1 sample. These values indicated that the G1S1-20 slurry could provide

sufficient compressive strength for most primary cementing applications. In chapter 6, it

will be further proved that the confined compressive strength values are even higher than

the UCS values when the samples are subjected to confining stresses.

On the other hand, the G1S1-30 sample had rapid strength development over the

first 2 days, while the G1S1-40 sample increased in strength very gradually past day-1.

Both slurries reached a plateau after 2 days. The ultimate strength for G1S1-30 slurry was

close to 1000 psi, while for G1S1-40, this value was approximately 400 psi. The G1S1-

30 and G1S1-40 samples can, for instance, be developed into a lost circulation treatment

with a constantly low compressive strength, a desirable quality for such a treatment.

Based on the distinct strength development profiles, the blending of geopolymer slurry

with SBM presents a primary benefit: by deliberately changing the amount of SBM in the

slurry, the compressive strength of the final mix can be fine-tuned to meet the strength

requirement for a specific application.

Page 82: Copyright by Xiangyu Liu 2017

62

Figure 4.5 - Compressive strength of neat geopolymer (G1) and geopolymer hybrids

(G1S1) at 170 °F and 3,000 psi

Downhole and Surface Rheology 4.2.2

The rheological properties of geopolymer hybrids at 70 °F were previously shown

in Figure 4.3 (b). The same measurements were also conducted at elevated temperature.

The neat geopolymer slurry (G1) was too viscous to be conditioned in a consistometer,

therefore was conditioned in a roller oven at 125 °F for 110 minutes. All geopolymer

hybrids were conditioned in a HPHT consistometer as described in experimental

methods. As shown in Figure 4.6, in comparison with data at room temperature (70 °F),

the geopolymer hybrid G1S1-20 had lower shear stresses at each shear rate after

conditioning at temperature and pressure. Neat geopolymer slurry also exhibited

0

1000

2000

3000

4000

5000

0 2 4 6 8 10 12 14

Co

mp

ress

ive

Str

ength

(p

si)

Time (days)

G1

G1S1-20

G1S1-30

G1S1-40

Page 83: Copyright by Xiangyu Liu 2017

63

improved rheological properties after conditioning at 125 °F compared to the room

temperature sample, and although it reached the maximum reading for the F1 spring

(R1B1 configuration), it reached the maximum at a higher shear rate than it did at room

temperature. These results are in contrast with the results for the Portland slurry (see

Figure 4.6), which had increased shear stresses at each shear rate at elevated temperature.

The data showed that improved rheological properties of the geopolymer and geopolymer

hybrids can be expected at elevated temperatures.

Figure 4.6 - Rheological properties of neat geopolymer (G1), geopolymer hybrid (G1S1-

20) and Portland cement (P1-R) at 70 °F and 125 °F

0

50

100

150

200

250

0 100 200 300 400 500 600

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

G1 (70 °F) G1S1-20 (70 °F) P1-R (70 °F)

G1 (125 °F) G1S1-20 (125 °F) P1-R (125 °F)

Page 84: Copyright by Xiangyu Liu 2017

64

Thickening Time 4.2.3

The thickening time of neat geopolymer could not be tested due to its high

viscosity. However, two geopolymer hybrids, G1 with 20% S1 (G1S1-20) and 30% S1

(G1S1-30) slurries, were tested and the results are shown in Figure 4.7. For both slurries,

around 6-18 hours the liquid state consistency maintained at approximately 25 Bc. After

this time period the consistency increased at a steady rate. The results showed that the

thickening time for geopolymer hybrids was very long, with the G1S1-20 sample

reaching 70 Bc at 21 hours. The test for the G1S1-30 sample was terminated at 23 hours,

at which point the consistency was 60 Bc. These results showed that incorporation of

higher dosages of SBM into geopolymers prolongs the thickening time. This behavior is

expected since a higher volume of SBM was added to replace the reactive geopolymer

component. Note that geopolymer hybrids will require acceleration for actual field use,

given these long thickening time values.

Figure 4.7 - Thickening time of geopolymer hybrids at 125 °F

0

10

20

30

40

50

60

70

0:00 4:00 8:00 12:00 16:00 20:00 24:00

Consi

sten

cy (

Bc)

Time (hh:mm)

G1S1-20

G1S1-30

Page 85: Copyright by Xiangyu Liu 2017

65

VALIDATION 4.3

Effect of Changing Activator Molarity 4.3.1

The concentration of alkali activator is a factor that influences the reaction of

geopolymer, and consequently the microstructure of hardened geopolymer (Salehi et al.,

2016a; Somna et al., 2011). In general for NaOH activation, the optimum molarity is

between 6 M and 10 M, while higher molarity results in slightly higher compressive

strength but less favorable rheology. In the present study, 4 M, 6 M and 8 M NaOH

solutions were used to compare their effectiveness in activating fly ash, and in creating

practical geopolymer hybrid cements. As shown in Figure 4.8, geopolymer activated by 6

M or 8 M developed similar 1-day compressive strength values in presence or absence of

SBM. 4 M NaOH activation was also attempted, however, the specimen showed

instability with a large amount of free water separating from the mixture right after

mixing.

Figure 4.8 - 1-day compressive strength of geopolymer hybrids activated by 6M or 8M

NaOH activator

0

500

1000

1500

2000

0% 20%

1-d

ay C

om

pre

ssiv

e S

tren

gth

(psi

)

% of SBM

8M NaOH

6M NaOH

Page 86: Copyright by Xiangyu Liu 2017

66

The effect of NaOH molarity on rheological properties of geopolymer hybrid was

evaluated, and the results are shown in Figure 4.9. As shown, the neat G1 slurry activated

by 6 M NaOH solution, G1(6 M), showed slightly lower rheological readings at lower

shear rates in comparison to the G1(8 M) slurry. When 20% SBM was added, the G1S1-

20(6 M) slurry showed much higher rheological readings when compared with 8 M

NaOH activation. The 4 M NaOH activated slurries were tested and showed significant

separation, thus are not shown on the plot. Based on the rheological measurements it can

be concluded that the 8 M NaOH activation provided the best strength and pumpability.

Therefore, the remaining work described in this dissertation was all conducted with 8 M

NaOH solution.

Figure 4.9 - Rheological properties of geopolymer hybrids activated by 6M or 8M NaOH

activator at 70 °F

0

50

100

150

200

250

0 100 200 300 400 500 600

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

G1(8M) G1S1-20(8M)

G1(6M) G1S1-20(6M)

Page 87: Copyright by Xiangyu Liu 2017

67

Effect of Changing Aluminosilicate Source 4.3.2

In order to show that there are many aluminosilicate sources that are appropriate

raw materials for the proposed mud-to-cement conversion technique, three different

sources of class F fly ash were tested. As described in the materials section, two fly ashes

(FA1 and FA3) were heat treated by the manufacturer to improve their reactivity, while

the other (FA2) was used as produced from the power plant. The results in Figure 4.10

showed that the slurries made from the three different sources did not significantly vary

in strength when cured for 1-day for any amount of SBM. By day-3, G2 was significantly

stronger by itself, but it had similar strengths to the other geopolymers when blended

with SBM. These results lead to the conclusion that even if the composition of the fly ash

varies between sources, the compressive strength of the geopolymer hybrid would not

change significantly. This however has to be verified for more sources of fly ash.

The rheological properties of the G1S1 hybrid (from FA1) were tested at 70 °F

and presented in Figure 4.3. Geopolymer slurry G3 showed comparable rheological

profiles to slurry G1 when no SBM was added. However, G3S1 hybrids (from FA3) had

an instability issue when more than 30% S1 was added, therefore the data are not shown

here. Results for G2S1 hybrids are shown in Figure 4.11. Compared with G1 slurries

shown in Figure 4.3 (b), neat G2 slurry had a significantly lower rheological profile.

Furthermore, the rheological properties of G2S1-30 and G1S1-40 slurries were nearly

identical to those of the neat Portland slurry, P1. These results show that the source

material can have a big impact on the viscosity as well as stability of the hybrid slurry.

Critical evaluation of the precursor is essential during the development and routine

testing (e.g. for field applications) of a geopolymer hybrid.

Page 88: Copyright by Xiangyu Liu 2017

68

Figure 4.10 - (a) 1-day and (b) 3-day compressive strength of geopolymer hybrids with

three different types of fly ashes

0

1000

2000

3000

4000

5000

0% 20% 30% 40%

1-d

ay C

om

pre

ssiv

e S

tren

gth

(p

si)

% of SBM

G1S1

G2S1

G3S1

(a)

0

1000

2000

3000

4000

5000

0% 20% 30% 40%

3-d

ay C

om

pre

ssiv

e S

tren

gth

(psi

)

% of SBM

G1S1

G2S1

G3S1

(b)

Page 89: Copyright by Xiangyu Liu 2017

69

Figure 4.11 - Rheological properties of G2S1 hybrids at 70 °F

The thickening time of G1S1-20 and G2S1-20 slurries are shown in Figure 4.12.

The liquid state consistencies of the two slurries were similar and were both around 20

Bc. After 17 hours, the G1S1-20 slurry started to thicken and reached 100 Bc by 23

hours. The G2S1-20 slurry exhibited a much longer thickening time. As shown in the

materials and methods chapter, the PSD of FA1 and FA2 are very similar to each other.

The differences in the rheological properties and thickening time behavior between G1

and G2 slurries could be attributed to the compositional differences between the fly

ashes. Notice that FA2 has a lower aluminum content (20.3%) while the other two fly

ashes have aluminum content values exceeding 25%. The thinner rheological behavior

and longer thickening time behavior of G2S1-20 in comparison to G1S1-20 could be a

result of fewer aluminum species present in the liquid state.

0

50

100

150

200

250

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

G2

G2S1-20

G2S1-30

G2S1-40

P1

Page 90: Copyright by Xiangyu Liu 2017

70

Figure 4.12 - Effect of changing aluminosilicate source on the thickening time of G1S1-

20 and G2S1-20 slurries at 125 °F and 3,000 psi

Effect of Seawater 4.3.3

For offshore deepwater cementing operations, it is not uncommon to use seawater

instead of fresh water to mix cement slurries since seawater is more readily available in

such situation than fresh water. For such applications, it is important to study the effect of

using seawater as mixing fluid on the compressive strength and rheological properties of

geopolymer slurries. As shown in Figure 4.13 and Figure 4.14, even when the mix water

was changed from deionized (DI) water to seawater, both compressive strength and

rheological properties remained unchanged for neat geopolymer and geopolymer hybrids.

Therefore, it appears to be feasible to use seawater as the mixing water for geopolymer

and geopolymer hybrids while still achieving reliable slurry properties. This helps reduce

cost of the expensive deepwater operation.

0

20

40

60

80

100

120

140

0:00 12:00 24:00 36:00 48:00

Co

nsi

sten

cy (

Bc)

Time (hh:mm)

G2S1-20

G1S1-20

Page 91: Copyright by Xiangyu Liu 2017

71

Figure 4.13 - Effect of using seawater vs. DI water on the compressive strength of G1 and

G1S1 hybrids at 70 °F and 3,000 psi

Figure 4.14 - Effect of using seawater vs. Di water on the rheological properties of G1

and G1S1 hybrids at 70 °F

0

500

1000

1500

2000

0% 20%

Co

mp

ress

ive

Str

ength

(p

si)

% of SBM

G1 (DI)

G1 (seawater)

0

50

100

150

200

250

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

G1 (DI) G1S1-20 (DI) G1S1-30 (DI)

G1 (seawater) G1S1-20 (seawater) G1S1-30 (seawater)

Page 92: Copyright by Xiangyu Liu 2017

72

Effect of SBM Composition 4.3.4

Previous results were all conducted with a field SBM that had a mud weight of

9.7 ppg, 23% CaCl2 and 75/25 SWR. To study the effect of SWR on the properties of the

geopolymer hybrid, additional brine was added to the field SBM to achieve a 60/40 SWR

while maintaining constant CaCl2 concentration in the internal phase. To study the effect

of internal brine phase composition on the properties of the geopolymer hybrid, the field

mud was diluted with water such that the CaCl2 content in the SBM was lowered from

23% to 13%.

Figure 4.15 reports the rheology readings for the three variations of the SBM,

which showed that SBM with higher SWR was less viscous at all shear rates at 80 °F,

120 °F and 150 °F. Changing the brine concentration in the internal phase had no impact

on the rheological behavior.

Figure 4.16 shows the rheological responses of geopolymer hybrids at 125 °F for

the three variations of SBM. As shown in the figure, the average shear stress of

geopolymer hybrid containing 30% original field SBM (75/25 SWR, 23% CaCl2) varied

from 12 Pa (at 5.1 s-1

) to 70 Pa (at 511 s-1

) with 10 sec gel strength of 15 Pa and 10 min

gel strength of 27 Pa. Altering the SWR ratio of the SBM from 75/25 to 60/40 while

maintaining the internal brine composition had no significant change in the rheological

profiles nor the gel strengths within the experimental error. Similar behavior was

observed by altering the internal brine composition, i.e., changing the CaCl2 content from

23% to 13%. In general, drilling muds with lower SWR usually exhibited higher

rheological readings (Figure 4.15). The geopolymer hybrid, on the other hand, was not

affected by the compositional changes in the drilling mud. Thus, the mud solidification

method is expected to be applicable to drilling muds with various SWR’s without

jeopardizing pumpability.

Page 93: Copyright by Xiangyu Liu 2017

73

Figure 4.15 - Rheology of the original SBM (75/25 SWR, 23% CaCl2) as well as

modified SBMs at (a) 80° F, (b) 120° F, and (c) 150° F

0

50

100

150

200

250

0 300 600 900 1200

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

(a) S1 at 80 °F 75/25, CaCl2=23%

60/40, CaCl2=23%

60/40, CaCl2=13%

0

10

20

30

40

50

60

0 300 600 900 1200

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

(b) S1 at 120 °F 75/25, CaCl2=23%

60/40, CaCl2=23%

60/40, CaCl2=13%

0

10

20

30

40

50

60

0 300 600 900 1200

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

(c) S1 at 150 °F 75/25, CaCl2=23%

60/40, CaCl2=23%

60/40, CaCl2=13%

Page 94: Copyright by Xiangyu Liu 2017

74

Figure 4.16 - Effect of changing SWR and internal brine CaCl2 concentration of SBM on

(a) rheology and (b) gel strength of G2S1-30-2A at 125 °F and 3,000 psi

0

50

100

150

200

250

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

75/25 CaCl2=23%

60/40 CaCl2=23%

60/40 CaCl2=13%

(a) G2S1-30-2A

Rheology

0

50

100

150

200

250

75/25 CaCl2=23% 60/40 CaCl2=23% 60/40 CaCl2=13%

Shea

r S

tres

s (P

a)

10 sec 10 min(b) G2S1-30-2A

Gel Strength

Page 95: Copyright by Xiangyu Liu 2017

75

Thickening time of geopolymer hybrids containing 30% field SBM (original as

well as modified) was measured at BHCT of 125 °F and BHP of 3,000 psi. It can be seen

that at the beginning, the liquid state consistencies of all three slurries were

approximately 20-30 Bc (Figure 4.17) and during the first part of the test the samples

maintained low consistency values. Beyond 1 hour 30 minutes, the consistency of the

sample with the original field SBM (75/25 SWR with 23% CaCl2) increased slowly and

reached 70 Bc at approximately 4 hour 50 minutes. When the SWR was changed to

60/40, the time to reach 70 Bc shortened by only 20 minutes, which is within the

experimental error. However, when the internal brine composition was reduced to 13%,

the pumping time further decreased by an additional 50 minutes. Thus, it can be

concluded that changing the SWR of the SBM does not impact pumping time

significantly, but lowering the internal brine content lowers the pumping time.

Figure 4.17 - Effect of changing SWR and internal brine CaCl2 concentration of SBM on

thickening time of G2S1-30-2A. (BHCT of 125 °F and BHP of 3,000 psi)

0

10

20

30

40

50

60

70

0:00 1:00 2:00 3:00 4:00 5:00

Consi

sten

cy (

Bc)

Time (h:min)

75/25 CaCl2=23%

60/40 CaCl2=23%

60/40 CaCl2=13%

Page 96: Copyright by Xiangyu Liu 2017

76

The 1-day compressive strength of geopolymers blended with 20% and 30% of

SBM with varying compositions was measured (Figure 4.18). When blended with 20%

original SBM, the compressive strength decreased from 2240 psi to 1210 psi and with

30% replacement, further decreased to 630 psi. For the other two variations of SBM, a

similar reduction in strength was observed.

Figure 4.18 - Effect of changing SWR and internal brine CaCl2 concentration of SBM on

1-day compressive strength of G2S1-30-2A at 170 °F and 3,000 psi

Effect of Pressure 4.3.5

Having a good understanding of the thickening time behavior of geopolymer

slurries at elevated temperature and pressure is of great importance in the development of

geopolymers for oilwell cementing applications. Salehi et al. (2016) reported that

0

500

1000

1500

2000

2500

3000

75/25 CaCl2=23% 60/40 CaCl2=23% 60/40 CaCl2=13%

Co

mp

ress

ive

Str

ength

(p

si)

0% SBM 20% SBM 30% SBM

Page 97: Copyright by Xiangyu Liu 2017

77

temperature has a big impact on the pumpability of geopolymer slurries, where the

consistency plateaued at higher values at higher temperature within a given time period.

In the present work, pumpability of geopolymer slurries at various pressures was

evaluated. Thickening time of G2S1-30 was measured at different BHPs. Figure 4.19

shows the trend for one representative slurry of three replicates at each pressure. It can be

seen that at the beginning of the test, the liquid state consistency at all three BHPs was

approximately 25-30 Bc. All the samples exhibited similar profiles until reaching a point

of departure around 1 hour 30 minutes. Beyond this point the consistency of the sample

with BHP of 3,000 psi increased slowly and reached 70 Bc at approximately 4 hour 40

minutes. In contrast, the sample subjected to 12,000 psi reached 70 Bc in 3 hours. Thus, it

can be concluded that the pumping time of geopolymer hybrid decreases with increasing

pressure. This acceleration in thickening time with pressure is also observed in Portland

cement slurries, but the degree to which it is affected is based on the specific ramp rate to

reach BHP (Nelson and Guillot, 2006).

Figure 4.19 - Effect of pressure on thickening time of G2S1-30-2A at BHCT of 125 °F

0

10

20

30

40

50

60

70

0:00 1:00 2:00 3:00 4:00 5:00

Co

nsi

sten

cy (

Bc)

Time (h:min)

3000 psi

6000 psi

12000 psi

Page 98: Copyright by Xiangyu Liu 2017

78

STABILITY CONTROL 4.4

A powder form stability modifier (A) was mixed into the slurries to aid with

stability of the blended slurry, which tended to separate at the highest SBM amounts,

particularly in stirred tests such as the thickening time test. The results presented in

Figure 4.20 showed that for G2 and G2S1-20 hybrid, the strengths with addition of 0%,

1.5% or 5% of the stability modifier (A) were nearly the same when considering the error

in the test. For slurries containing G2S1-30 and G2S1-40 samples, with addition of 1.5%

or 5% of the modifier the compressive strength increased in value at both day-1 and day-

3. These results show that not only does the powder work as a stability enhancer, but it

also has the added benefit of increased strength for the slurries with higher levels of

SBM.

The room temperature rheologies of G2 slurries mixed with 1.5% of the stability

modifier were measured, and the results are shown in Figure 4.21. The data for the

geopolymer hybrids with and without the additive are shown for ease of comparison. G2

was selected for this purpose, since it was thinner than G1 and its properties could be

measured without reaching the maximum for the viscometer with a F1 spring. In general,

the shear stresses at each shear rate increased with inclusion of the modifier, which was

expected due to the fineness of the powder additive (refer to PSD shown in Figure 3.2).

The downhole rheological profiles were also measured for the G2S1-40 hybrids with 1%,

1.5%, or 3% of the stability modifier and are presented in Figure 4.22. It can be seen that

at elevated temperatures, there was a significant improvement in the rheologies, and the

profiles were nearly similar to that of the Portland slurry with retarder (P1-R). These

results showed that the stability-enhancing modifier did affect the surface rheology when

added to slurries with high amounts of S1. However, the influence of the modifier on the

slurries was minimum with excellent rheological properties at elevated temperatures.

Page 99: Copyright by Xiangyu Liu 2017

79

Figure 4.20 - (a) 1-day and (b) 3-day compressive strength of G2S1 hybrids with varying

dosages of stability modifier (A)

0

1000

2000

3000

4000

5000

6000

0% 20% 30% 40%

1-d

ay C

om

pre

ssiv

e S

tren

gth

(p

si)

% of SBM

(a) G2S1-0A

G2S1-1.5A

G2S1-5A

0

1000

2000

3000

4000

5000

6000

0% 20% 30% 40%

3-d

ay C

om

pre

ssiv

e S

tren

gth

(psi

)

% of SBM

(b) G2S1-0A

G2S1-1.5A

G2S1-5A

Page 100: Copyright by Xiangyu Liu 2017

80

Figure 4.21 - Effect of adding 1.5% stability modifier (A) on the rheological properties of

(a) G2 and G2S1-20 hybrid, (b) G2S1-30 and G2S1-40 hybrids at 70 °F

0

50

100

150

200

250

0 100 200 300 400 500 600

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

(a) G2-1.5A G2

G2S1-20-1.5A G2S1-20

P1-R

0

50

100

150

200

250

0 100 200 300 400 500 600

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

(b) G2S1-30-1.5A G2S1-30

G2S1-40-1.5A G2S1-40

P1-R

Page 101: Copyright by Xiangyu Liu 2017

81

Figure 4.22 - Effect of different dosages of stability modifier (A) on the rheological

properties of G2S1-40 hybrids at 125 °F

The stability modifier was used in G1 slurries, as well, and thickening time curves

with and without the stability modifier of the G1S1-30 hybrid are shown in Figure 4.23.

The consistency of the slurry containing the modifier was slightly higher at all ages. The

thickening time for the slurry was shortened with addition of 0.75% of the modifier, with

the slurry reaching 70 Bc at 19 hours. The shorter thickening time of the slurry measured

with the stability modifier is a beneficial result from the use of the modifier, however, the

higher consistency values are not.

0

50

100

150

200

250

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

G2S1-40-3A

G2S1-40-1.5A

G2S1-40-1A

P1-R

Page 102: Copyright by Xiangyu Liu 2017

82

Figure 4.23 - Effect of adding 0.75% of stability modifier (A) on thickening time for

G1S1-30 hybrid

SOLIDIFICATION OF NON-AQUEOUS DRILLING MUDS 4.5

The previous sections were all focused on the solidification of one type of SBM

(S1). In this section, compressive strength of geopolymer hybrid was also measured for

another SBM (S2) and two OBMs (O1, O2) and the values were compared with that of

Portland cement (P1). From Figure 4.24 it can be seen geopolymer performed better than

Portland cement with all four muds. At 20% replacement volume, Portland cement

retained less than 30% compressive strength, while geopolymer retained more than 50%

of its compressive strength with all four muds. The S2 and O2 muds required addition of

3% stability modifier. Based on these data, geopolymer demonstrates the possibility of

converting different types of NAFs into cementitious material that may be used for a

0

20

40

60

80

100

0:00 4:00 8:00 12:00 16:00 20:00 24:00

Co

nsi

sten

cy (

Bc)

Time (hh:mm)

G1S1-30

G1S1-30-0.75A

Page 103: Copyright by Xiangyu Liu 2017

83

variety of applications. If desired, the geopolymer formulation can be further optimized

for improved compressive strength. Figure 4.25 shows pictures of two hardened

geopolymer hybrid samples that contained 20% of S1 and O2 muds. No layering or

separation was observed, demonstrating a good sample integrity and uniformity.

Figure 4.24 - Normalized compressive strength of hardened Portland cement slurries (P1)

and geopolymer slurries (G1) with 20% mud replacement (by volume)

0%

20%

40%

60%

80%

100%

S1 S2 O1 O2

% o

f U

nco

nta

min

ated

Com

pre

ssiv

e S

tren

gth

G1-20

P1-200% Mud

Page 104: Copyright by Xiangyu Liu 2017

84

(a) (b)

Figure 4.25 - Picture of a geopolymer/mud sample containing 20% (a) S1 and (b) O2

SUMMARY 4.6

This work presents a new versatile technique to solidify NAFs into cementitious

materials using sodium hydroxide-activated geopolymer. Although the concept of drilling

mud solidification is not new, this is the first time this technology has become applicable

to SBM or OBM systems. The key findings include:

1. Geopolymers showed much better compatibility with NAFs compared to Portland

cement slurries in terms of both UCS and rheological properties.

2. The geopolymer hybrid cement provided sufficient compressive strength and

acceptable rheological properties for well cementing applications.

3. The short term and long term compressive strength of geopolymer can be tailored

to the application with the use of different dosages of SBM. More specifically,

lower dose hybrids may be suitable for primary cementing applications and higher

dose hybrids may find application in lost circulation treatments.

Page 105: Copyright by Xiangyu Liu 2017

85

4. This mud-to-cement conversion technique is not limited to a single source

material. All three types of fly ashes tested were found to be viable candidates for

the solidification method.

5. The three fly ashes with varying compositions did not affect compressive strength

but showed a big influence on the viscosity as well as stability of the hybrid

slurry. Therefore, critical evaluation of the precursor is essential during the

development of a geopolymer hybrid.

6. For offshore deepwater cementing operations, seawater can be used instead of

fresh water without any negative effect on strength or rheology.

7. Changing SBM SWR or calcium chloride concentration showed no significant

effect on the compressive strength or rheology of geopolymer hybrid.

8. Changing the SBM SWR from 70/30 to 60/40 did not alter the thickening time,

but lowering calcium concentration slightly accelerated the thickening behavior.

9. Similar to Portland cement slurries, increased pressure shortened the thickening

time of geopolymer hybrids.

The primary benefits of this solidification method include but are not limited to:

1. Improved compatibility of the slurry with NAFs, and thus enhanced zonal

isolation in wells that have been drilled or abandoned with NAFs.

2. Reduced risk of poor cementation when NAF-based drilling mud displacement is

incomplete and/or when filter cake is present.

3. Inexpensive source material.

4. Environmentally friendly on-site/in-situ or off-site disposal of non-recyclable

NAFs.

Page 106: Copyright by Xiangyu Liu 2017

86

Chapter 5: Silicate Activation

In this chapter, geopolymers activated with silicate alkaline solutions were

evaluated for their applicability in the mud solidification process. To achieve favorable

strength and mechanical properties, sodium silicate was combined with sodium

hydroxide solution in geopolymer synthesis as described in chapter 3. Studies have

shown that silicate-activated geopolymers could lead to a higher-strength and lower-

porosity binder, as the alkali silicate solution provides additional silicate species that

results in the formation of a larger volume of aluminosilicate gel (Duxson et al., 2005;

Criado et al., 2007). For silicate-activated geopolymers, the optimum Na2O/Al2O3 ratio is

reported to be around 1, and SiO2/Na2O ratio is between 0.1 and 2 (Chindaprasirt et al.,

2012; Criado et al., 2008, 2007). The silicate species in the solution control the rate of

structural reorganization and densification during polymerization as well as the degree of

reaction, especially at early age. Thus, engineering the right amount of silicate in the

alkaline solution is a critical task to optimize the slurry design for cementing purposes.

Geopolymer slurries activated by sodium silicate in both liquid form and solid

form were tested for their applicability in the mud solidification method. The rheological

properties, thickening time, and compressive strength of silicate-activated slurries were

compared with hydroxide-activated slurries. Results for neat silicate-activated

geopolymers are first presented, followed by results for geopolymer hybrids formulated

with the original field SBM.

RHEOLOGICAL PROPERTIES 5.1

Figure 5.1 shows the rheological readings of a neat geopolymer slurry activated

with liquid-form sodium silicate (LSS). As can be seen in Figure 5.1 (a), the silicate mix

Page 107: Copyright by Xiangyu Liu 2017

87

with 0.12 SiO2/Na2O ratio (LSS-0.12) had very similar shear stress readings at low shear

rates and gel strength values compared with the sodium hydroxide mix (LSH). The LSS-

0.12 slurry became more viscous beyond 100 s-1

shear rate. The silicate mix with 0.24

SiO2/Na2O ratio (LSS-0.24) had similar rheological readings at higher shear rates

compared to the mix with 0.12 SiO2/Na2O ratio (LSS-0.12). However, the LSS-0.24 mix

was more viscous at lower shear rates and had much higher 10 sec and 10 min gel

strength values (Figure 5.1 b), which raises concerns about pumpability. As the

SiO2/Na2O ratio was increased to 0.48, the slurry became unpumpable and the gel

strength was significantly higher at 10 minutes. These results concur with findings

previously reported in the literature, which indicate that liquid-form sodium silicate tends

to generate more viscous geopolymer mixtures with a tendency to stick to mixing

equipment rather than flowing easily (Lloyd, 2009).

Page 108: Copyright by Xiangyu Liu 2017

88

Figure 5.1 - (a) rheology and (b) gel strength of liquid-form sodium silicate (LSS)

activated geopolymer (0% SBM) at room temperature

0

50

100

150

200

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

LSS-0.12

LSS-0.24

LSS-0.48

LSH-8M

(a)

0

25

50

75

100

125

150

LSS-0.12 LSS-0.24 LSS-0.48 LSH-8M

Shea

r S

tres

s (P

a)

10 sec

10 min

(b)

Page 109: Copyright by Xiangyu Liu 2017

89

Figure 5.2 shows the effect of SBM on liquid silicate-activated geopolymer.

Unlike sodium hydroxide-activated geopolymer, which showed improvement in rheology

when blended with SBM, all three LSS-activated geopolymer slurries showed higher

rheological readings. The calcium ions from the internal phase of SBM likely interacted

with the silicate and formed a gel leading to the higher readings. For LSS-0.24 mix, the

addition of SBM reduced the “stickiness” of the mixture and showed slippage against the

viscometer rotor; however, this mixture was somewhat unstable.

Since the geopolymer activated by LSS was too viscous to be pumped, solid

sodium silicate (SSS) was tested. It exhibited significantly improved rheological

properties. As shown in Figure 5.3, all three SSS-activated geopolymer slurries showed

rheological readings comparable to LSH. The high gelation tendency was also inhibited,

allowing for a pumpable slurry. When SBM was added to SSS-activated geopolymer,

both rheological readings and gel strength values remained unaffected, regardless of the

SiO2/Na2O ratio in the activator (Figure 5.4). Based on these observations, it can be

concluded that LSS-activated geopolymer in general showed poor pumpability. However,

when solid-form silicate was used, the dissolution of silicate species from the solid

silicate activator was delayed, resulting in good pumpability and delayed gelation.

Page 110: Copyright by Xiangyu Liu 2017

90

Figure 5.2 - (a) rheology and (b) gel strength of liquid-form sodium silicate (LSS)

activated geopolymer blended with 20% SBM at room temperature

0

50

100

150

200

0 100 200 300 400 500 600

Shea

r S

tres

s (P

a)

Shear Rate (s-1)

LSS-0.12

LSS-0.24

LSS-0.48

LSH-8M

(a)

0

25

50

75

100

125

150

LSS-0.12 LSS-0.24 LSS-0.48 LSH-8M

Shea

r S

tres

s (P

a)

10 sec

10 min

(b)

Page 111: Copyright by Xiangyu Liu 2017

91

Figure 5.3 - (a) rheology and (b) gel strength of solid form sodium silicate (SSS)

activated geopolymer (0% SBM) at room temperature

0

50

100

150

200

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

SSS-0.12

SSS-0.24

SSS-0.48

LSH-8M

(a)

0

25

50

75

100

125

150

SSS-0.12 SSS-0.24 SSS-0.48 LSH-8M

Shea

r S

tres

s (P

a)

10 sec

10 min

(b)

Page 112: Copyright by Xiangyu Liu 2017

92

Figure 5.4 - (a) rheology and (b) gel strength of solid form sodium silicate (SSS)

activated geopolymer blended with 20% SBM at room temperature

0

50

100

150

200

0 100 200 300 400 500 600

Sh

ear

Str

ess

(Pa)

Shear Rate (s-1)

SSS-0.12

SSS-0.24

SSS-0.48

LSH-8M

(a)

0

25

50

75

100

125

150

SSS-0.12 SSS-0.24 SSS-0.48 LSH-8M

Shea

r S

tres

s (P

a)

10 sec

10 min

(b)

Page 113: Copyright by Xiangyu Liu 2017

93

COMPRESSIVE STRENGTH 5.2

The effect of silicate activation on compressive strength has been well

documented in the open literature (Duxson et al., 2005; Criado et al., 2007). In general,

moving from hydroxide activation to silicate activation tends to increase the compressive

strength. Here the same trend was observed with fly ash FA1 (Figure 5.5 a), with the

LSS-0.24 mix almost doubling in compressive strength compared to the LSH-activated

slurry. The fly ash FA 2, on the other hand, had lower strength value with either liquid-

form (Figure 5.5 b) or solid-form (Figure 5.6) silicate activation. Therefore, the effect of

activator on strength highly depends on the composition of the aluminosilicate precursor.

Factors including glassy content, structure, and particle size distribution could all play a

role and influence the reactivity of fly ash. In addition, for FA1, increasing the amount of

silicate had no further improvements in the compressive strength, which indicates that

there is an optimum SiO2/Na2O ratio in the activator for desired compressive strength.

These results show that analogous to Portland slurry design, a full suite of tests is

necessary using the specific aluminosilicate powder during the geopolymer slurry design

process.

Page 114: Copyright by Xiangyu Liu 2017

94

Figure 5.5 - Compressive strength of liquid sodium silicate (LSS) activated geopolymer

for (a) fly ash 1 and (b) fly ash 2 in comparison to liquid sodium hydroxide (LSH)

activated geopolymer without any SBM at 170 °F and 3,000 psi

0

500

1000

1500

2000

2500

3000

LSH-8M LSS-0.24 LSS-0.72

1-d

ay C

om

pre

ssiv

e S

tren

gth

(p

si) (a) FA 1

0

500

1000

1500

2000

2500

3000

LSH-8M LSS-0.12 LSS-0.24

1-d

ay C

om

pre

ssiv

e S

tren

gth

(psi

) (b) FA 2

Page 115: Copyright by Xiangyu Liu 2017

95

Figure 5.6 - Compressive strength of solid sodium silicate (SSS) activated geopolymer

for FA 2 in comparison to liquid sodium hydroxide (LSH) activated geopolymer without

any SBM at 170 °F and 3,000 psi

THICKENING TIME 5.3

Another important parameter in the development of any cement slurry is

thickening time. For this, silicate activation can be successfully utilized as a means to

accelerate / retard the thickening time. As shown in Figure 5.7, in the absence of SBM,

neat LSH-activated geopolymer reached 70 Bc in 53 hours. For sample LSS-0.12 with

SiO2/Na2O ratio of 0.12, the thickening time was shortened to 22 hours. The thickening

time test for LSS-0.24 was attempted, but the slurry was too viscous to be tested in a

consistometer. Clearly, the presence of dissolved silicate in the solution greatly

accelerated the reaction rate. The rapid gelation and setting behavior of LSS-activated

geopolymer has also been documented by other researchers (Antoni et al., 2016). The

0

500

1000

1500

2000

2500

3000

LSH-8M SSS-0.12 SSS-0.24 SSS-0.48

1-d

ay C

om

pre

ssiv

e S

tren

gth

(p

si)

FA2

Page 116: Copyright by Xiangyu Liu 2017

96

SSS-0.24 sample, on the other hand, had a considerably longer thickening time of 69

hours, indicating the delayed gelation of the aluminate and silicate species. Solely based

on the thickening time data it might appear that the SSS-0.24 sample does not set for a

long period of time. Note that under static conditions all the slurries developed more than

1,000 psi compressive strength by 24 hours (Figure 5.5, Figure 5.6). The dynamic

conditioning process in a consistometer breaks down the gel, thus significantly delaying

the restructuring of reaction products. Once pumping operations have concluded and the

slurry is in place, the setting time will become significantly shortened.

Figure 5.7 - Effect of various activating solutions on the thickening time of neat

geopolymer slurry (G2) at 125 °F and 3,000 psi

Figure 5.8 shows the effect of adding SBM on thickening time of both LSH- and

SSS-activated geopolymer. A very pronounced accelerating effect was observed for both

0

10

20

30

40

50

60

70

0:00 12:00 24:00 36:00 48:00 60:00 72:00

Co

nsi

sten

cy (

Bc)

Time (h:min)

SSS-0.24 LSS-0.12 LSH-8M

Page 117: Copyright by Xiangyu Liu 2017

97

activation chemistries. The thickening time of LSH-activated G2 slurry with 20% SBM

was 3.5 hours, which is 50 hours shorter than for the slurry without SBM (Figure 5.8 a).

The accelerating effect can be attributed to the interaction between calcium ions in the

internal phase of SBM and silicate ions from the activating solution. The SSS-0.24 slurry

showed an initial peak of 65 Bc at 3.5 hours, but continued stirring broke down the initial

gel (Figure 5.8 b). The slurry plateaued after about 14 hours, and the liquid state

consistency lasted for about 5 hours before the slurry began to thicken. The final

thickening time to 70 Bc was 39 hours. As shown in Figure 5.6, the SSS-0.24 slurry

developed a compressive strength of 980 psi by 24 hours, well before the thickening time

reached 70 Bc. This unique behavior of SSS-activated geopolymer is promising for

treating lost circulation in particular. By changing the SiO2/Na2O ratio in the activator,

the initial gelation peak of the slurry can be manipulated to reach the values needed to

plug fractures and stop losses during a lost circulation event. Once the cement plug is in

place, the slurry will start to develop compressive strength. The thickening time of LSS-

activated geopolymer could not be measured due to the high viscosity of the

geopolymer/mud hybrid upon mixing.

Page 118: Copyright by Xiangyu Liu 2017

98

Figure 5.8 - Effect of adding 20% original SBM on thickening time of (a) LSH-8M and

(b) SSS-0.24 geopolymer slurries (G2) at 125 °F and 3,000 psi

0

10

20

30

40

50

60

70

0:00 12:00 24:00 36:00 48:00 60:00 72:00

Co

nsi

sten

cy (

Bc)

Time (h:min)

Neat

20% SBM

(a)

0

10

20

30

40

50

60

70

0:00 12:00 24:00 36:00 48:00 60:00 72:00

Consi

sten

cy (

Bc)

Time (h:min)

Neat

20% SBM

(b)

Page 119: Copyright by Xiangyu Liu 2017

99

SUMMARY 5.4

This chapter explored the feasibility of using silicate-based activator for SBM

solidification. Both liquid form and solid form sodium silicate were tested. The properties

examined include rheology, thickening time and compressive strength. The main variable

in designing silicate-activated geopolymer was the SiO2/Na2O ratio. The following

conclusions are drawn from the experimental findings:

1. LSS activation resulted in unfavorable rheology readings and gel strength. The

presence of SBM further increased the viscosity of LSS-activated geopolymer,

leading to an unpumpable slurry with a rapid setting behavior. In contrast, SSS

activation did not affect the rheological behavior of geopolymer in the presence or

absence of SBM.

2. For neat geopolymer slurries, LSS activation accelerated the pumping time. SSS

activation, by contrast, retarded the pumping time in comparison to hydroxide

activation. The presence of SBM significantly accelerated the pumping time for

both silicate and hydroxide activation. The SSS activation showed an initial high

consistency peak that was broken down with continued shearing.

3. Silicate activation greatly improved the compressive strength compared with pure

hydroxide activation, confirming the findings in existing literature. However, the

effectiveness of strength enhancement highly depended on the aluminosilicate

source selected.

Overall, silicate activation affects both the fresh state properties as well as

hardened state properties of geopolymer hybrid cement in comparison to hydroxide

activation. By adjusting the type and dosage of silicate in the activator, the pumping time

and compressive strength of the geopolymer hybrid can be altered.

Page 120: Copyright by Xiangyu Liu 2017

100

Chapter 6: Mechanical Properties and Self-Healing Capability

To ensure a robust cement design and long term well integrity, this chapter further

studies the mechanical properties of the geopolymer hybrid cement. Specifically, the self-

healing capability of the geopolymer hybrids was evaluated under both uniaxial and

triaxial confinement conditions. The self-healing behavior exhibited from these tests

gives insight into the capability of the geopolymer hybrids to accommodate subsurface

stress variations and regain strength after deformation and failure. It is important to note

here that the term “self-healing” relates to healing of the matrix material itself, not to the

incorporation of e.g. swellable materials to recover hydraulic isolation after the

cementitious material is compromised and invaded by hydrocarbons (Cavanagh et al.,

2007; Reddy et al., 2010; Taoutaou et al., 2011). A true self-healing material is

particularly well-suited for wells that will undergo a lot of geomechanical / tectonic

loads, and can be adopted as a versatile material for temporary / permanent abandonment

of wells.

Furthermore, the cement-to-pipe bond strength of geopolymer hybrids was

measured for the first time in the presence and absence of a layer of drilling mud on the

pipe surface. The hydraulic conductivity, porosity and pore size distribution of

geopolymers were compared with hardened Portland cement and an alkali-activated slag.

CONFINED COMPRESSIVE STRENGTH 6.1

The confined compressive strength tests on geopolymer (G1), geopolymer hybrid

(containing 20% SBM and 80% geopolymer by volume) and uncontaminated Portland

cement slurries were conducted at 100 psi and 500 psi confining pressures. Two samples

were tested for each slurry at each pressure. Figure 6.1, Figure 6.2 and Figure 6.3 showed

Page 121: Copyright by Xiangyu Liu 2017

101

the differential stress with respect to axial strain for each of these samples. The maximum

differential stress is equal to two times of the maximum shear stress in the sample. As

shown in the figures, the differential stress increased with an increase in the confining

stress. The neat geopolymer sample (Figure 6.1) showed a brittle failure pattern at 100 psi

confining pressure with the post peak differential stress dipping by more than 50% of the

peak stress. When a sample exhibits brittle behavior, the stress decreases with increased

strain past the yield point. At 500 psi confining pressure, the geopolymer started to show

brittle-ductile transition behavior wherein beyond post peak the differential stress

gradually decreased as the axial strain was increased. The geopolymer hybrid (Figure 6.2)

showed ductile behavior at both 100 psi and 500 psi confining pressures. Ductile

behavior is characterized by the ability of the material to deform without losing

toughness; in other terms, the stress continuously increases with increasing strain. Ductile

behavior is advantageous for cement sheath integrity in a well because the set cement will

deform or flow instead of cracking. In comparison (Figure 6.3), hardened Portland

cement slurries were more brittle than the geopolymer hybrid at both confining pressures.

The confined compressive strength is defined as the peak stress the sample was

able to withstand. The values are calculated as the summation of differential stress and

the confining stress. In Table 6.1, as expected, the confined compressive strength was

higher at a higher confining pressure for all three slurries. Note that for the geopolymer

hybrid in particular, the confined compressive strength at 500 psi confining stress was in

excess of 2,800 psi, which will be more than sufficient for most types of primary

cementation applications.

Page 122: Copyright by Xiangyu Liu 2017

102

Figure 6.1 - Differential stress vs. axial strain for two samples each for geopolymer

subjected to a confining pressure of 100 psi and 500 psi at 7 days

Figure 6.2 - Differential stress vs. axial strain for two samples each for geopolymer

hybrid subjected to a confining pressure of 100 psi and 500 psi at 7 days

0

2000

4000

6000

8000

0.0% 0.5% 1.0% 1.5% 2.0% 2.5% 3.0%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

100 psi (#1)100 psi (#2)500 psi (#1)500 psi (#2)

100 psi

500 psi

0

2000

4000

6000

8000

0.0% 0.5% 1.0% 1.5% 2.0%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

100 psi (#1)

100 psi (#2)

500 psi (#1)

500 psi (#2)

100 psi

500 psi

Page 123: Copyright by Xiangyu Liu 2017

103

Figure 6.3 - Differential stress vs. axial strain for two samples each for Portland cement

subjected to a confining pressure of 100 psi and 500 psi at 7 days

Table 6.1 - Confined compressive strength at 7 days

Pc = 100 psi Pc = 500 psi

Geopolymer 3330 psi 5000 psi

Geopolymer Hybrid 2000 psi 2870 psi

Portland Cement 5600 psi 7850 psi

MECHANICAL PROPERTIES 6.2

As one of the goals of this study was to investigate the potential use of

geopolymers in P&A operation, it would be beneficial to study the mechanical properties

0

2000

4000

6000

8000

0.0% 0.5% 1.0% 1.5% 2.0%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

100 psi (#1)

100 psi (#2)

500 psi (#1)

500 psi (#2)

100 psi

500 psi

Page 124: Copyright by Xiangyu Liu 2017

104

of the hardened geopolymer slurries under triaxial stress conditions. Properties such as

Young’s modulus and Poisson’s ratio are important performance parameters as they

reflect the flexibility of a cement system and its ability to withstand the varying

geomechanical / tectonic loads that act on cement and cemented casing string during the

lifetime of an abandoned well. Here, these elastic constants were determined from the

linear portion of the stress vs. strain curve of samples that were cured for 7 days and 28

days.

Young’s modulus characterizes the ductility of a material, with a lower Young’s

modulus value corresponding to a more ductile material. As shown in Table 6.2, the

Young’s modulus of neat geopolymer sample was 0.39 Mpsi by 7 days. The sample

became more brittle by 28 days and the Young’s modulus increased to 0.54 Mpsi. Both

the neat geopolymer and geopolymer hybrid showed lower Young’s modulus values

compared with Portland slurry at 7 days and 28 days, indicating that geopolymer was

more ductile than Portland cement. Similar conclusion was observed by Khalifeh et al. in

their study where they compared the properties of potassium silicate-activated class C fly

ash-based geopolymer and Portland cement (2015).

Measurements also showed higher Poisson’s ratio () for both geopolymer and

geopolymer hybrid cements in comparison to the values of Portland cement (Table 6.2).

However, the Poisson’s ratio values of Portland cement obtained are lower than the

values reported by Iverson et al. (2008). The reported Poisson’s ratio was 0.2 for neat

Portland cement, although the samples were cured and measured at different downhole

temperature and pressure conditions. The lower Poisson’s ratio values could be attributed

to the creeping effect. In other words, the sample deformed viscously in a time-dependent

manner and the measurement was conducted before the creep strain became more

appreciable.

Page 125: Copyright by Xiangyu Liu 2017

105

Table 6.2 - Mechanical properties of hardened slurries at 500 psi confining pressure

7 days 28 days

E (Mpsi) ν E ( Mpsi) ν

Geopolymer 0.39 0.12 0.54 0.10

Geopolymer Hybrid 0.37 0.09 0.43 0.14

Portland 0.65 0.06 0.66 0.09

SELF-HEALING PROPERTIES 6.3

The self-healing behavior of geopolymer was studied in order to evaluate the

capability of the cement sheath to adapt its structure to the influence surrounding

environment. Of particular interest is the ability of the cement sheath to regain integrity

and hydraulic isolation after it has been compromised by cracking, e.g. under the

influence of a changing geomechanical load in an abandoned well. Self-healing capability

was evaluated both with and without confining pressure. The results for geopolymer and

Portland cement samples without confinement are shown in Figure 6.4. The vertical axis

shows the compressive strength of the samples that were pre-damaged with respect to the

compressive strength of the samples that were not pre-damaged. As shown in Figure 6.4

(a), geopolymer samples that were pre-loaded at 7 days of curing and re-healed for an

additional 21 days consistently retained a higher percentage of compressive strength

compared to Portland cement at all three pre-loading levels. In addition, geopolymer

samples that were pre-loaded developed higher compressive strength than those without

any pre-load, reflected by the data bars that exceed the baseline (100%). This could be

Page 126: Copyright by Xiangyu Liu 2017

106

attributed to the pre-loading stresses that cracked unreacted particles. More reaction

products could form when these newly cracked surfaces come in contact with pore fluid.

This behavior revealed that geopolymer slurry has intrinsic self-healing capability when

subjected to certain level of damage.

The same test was repeated for samples that were pre-loaded at 28 days of curing

and re-healed for an additional 28 days. As shown in Figure 6.4 (b), for the samples pre-

loaded to 70%, geopolymer retained more than 120% compressive strength, which was

significantly higher than the values for Portland cement. The samples loaded to 30% and

50% of their compressive strength did not show a significant difference between the two

types of slurries. The samples that were pre-loaded to 70% of their compressive strength

were very likely to have microstructure damage, which the geopolymer was able to re-

heal but Portland cement was not. At 30% and 50% pre-loading levels, only little damage

was introduced to the sample structure, which explains why no significant difference

between the two slurries was observed.

Page 127: Copyright by Xiangyu Liu 2017

107

Figure 6.4 - Self-healing properties of geopolymer and Portland cement pre-loaded at (a)

7 days and (b) 28 days under atmospheric conditions. Cylindrical samples were prepared

at 170 °F

40%

60%

80%

100%

120%

140%

30% 50% 70%

28

-day

Co

mp

ress

ive

Str

ength

Rec

over

y

Pre-loading Level at 7-day (% of 7-day Compressive Strength)

(a)

Geopolymer

Portland cement

Baseline compressive

strength of 28-day samples

without pre-damage

40%

60%

80%

100%

120%

140%

30% 50% 70%

56-d

ay C

om

pre

ssiv

e S

tren

gth

Rec

over

y

Pre-loading Level at 28-day (% of 28-day Compressive Strength)

(b)

Geopolymer

Portland cement

Baseline compressive

strength of 56-day samples

without pre-damage

Page 128: Copyright by Xiangyu Liu 2017

108

Since the samples were only loaded up to a maximum of 70% of the compressive

strength in the unconfined tests, these tests can only partially reflect the self-healing

capability of the material. In particular, the unconfined test method could not provide

information regarding the self-healing capability past peak stress due to catastrophic

failure of the samples, a situation not representative of material behavior in the downhole,

confined environment. Therefore, a confined self-healing test was designed such that

materials were loaded beyond their yield point to create significant internal damage to

their matrices. The ability of the hardened cement to fully recover - or to develop even

higher - compressive strength would truly reflect any self-healing capability.

The confined self-healing tests were carried out on geopolymer, geopolymer

hybrid and hardened Portland cement slurries. As described in the experimental method

chapter (Chapter 3), a cylindrical sample was first loaded beyond yield point at 7 days

and was allowed to re-heal for an additional 21 days. Two samples were measured for

each slurry. As shown in Figure 6.5 and Figure 6.6, both geopolymer and geopolymer

hybrid samples showed significant self-healing capability with the 28-day peak stress

values largely exceeding the 7-day peak stress values. In comparison, hardened Portland

cement could no longer support the 7-day stress once the sample was yielded (Figure

6.7). Table 6.3 summarized the changes in peak stresses after re-healing period. As can

be seen, the peak stresses of two neat geopolymer samples increased by 35% and 30%

respectively. The geopolymer hybrid samples also showed re-healing and peak stresses

increased by approximately 18%. The peak stresses of Portland cement samples dropped

by more than 20% for both samples.

Page 129: Copyright by Xiangyu Liu 2017

109

Figure 6.5 - Self-healing capability of geopolymer (G) under 500 psi confining stress, for

two samples (a) and (b). The G-7 sample was loaded beyond its yield point at 7 days, and

the same sample was re-tested to failure at 28 days (G-7-28). Note the evident increase in

peak stress observed for the samples at 28 days

0

1,000

2,000

3,000

4,000

5,000

6,000

0.0% 0.5% 1.0% 1.5% 2.0% 2.5%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

(a)

G-7

G-7-28

0

1,000

2,000

3,000

4,000

5,000

6,000

0.0% 0.5% 1.0% 1.5% 2.0% 2.5%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

(b)

G-7

G-7-28

Page 130: Copyright by Xiangyu Liu 2017

110

Figure 6.6 - Self-healing capability of geopolymer hybrid (GH) under 500 psi confining

stress, for two samples (a) and (b). The GH-7 sample was loaded beyond its yield point at

7 days and the same sample was re-tested to failure at 28 days (GH-7-28). Note the

evident increase in peak stress observed for the samples at 28 days

0

1,000

2,000

3,000

4,000

5,000

0.0% 0.5% 1.0% 1.5% 2.0% 2.5% 3.0%

Dif

fere

nti

all S

tres

s (p

si)

Axial Strain

(a)

GH-7

GH-7-28

0

1,000

2,000

3,000

4,000

5,000

0.0% 0.5% 1.0% 1.5% 2.0% 2.5% 3.0%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

(b)

GH-7

GH-7-28

Page 131: Copyright by Xiangyu Liu 2017

111

Figure 6.7 - Self-healing capability of Portland cement (P) under 500 psi confining stress,

for two samples (a) and (b). The P-7 sample was loaded beyond its yield point at 7 days

and the same sample was re-tested to failure at 28 days (P-7-28). Note the evident

reduction in peak stress observed for the samples at 28 days

0

2,000

4,000

6,000

8,000

0.0% 0.5% 1.0% 1.5% 2.0% 2.5% 3.0%

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

(a)

P-7

P-7-28

0

2,000

4,000

6,000

8,000

0.00 0.01 0.01 0.02 0.02 0.03 0.03

Dif

fere

nti

al S

tres

s (p

si)

Axial Strain

(b)

P-7

P-7-28

Page 132: Copyright by Xiangyu Liu 2017

112

Table 6.3 - Peak stress values at 7-days and after 21-day waiting period (at 500 psi

confining stress)

Pre-loading

Peak Stress

(7-day) psi

Re-loading

Peak Stress

(28-day) psi

% Change

Geopolymer

Sample (a) 4,010 5,420 35%

Sample (b) 3,960 5,120 30%

Geopolymer

Hybrid

Sample (a) 3,420 4,030 18%

Sample (b) 3,520 4,110 17%

Portland

Cement

Sample (a) 6,630 3,960 -40%

Sample (b) 6,890 5,330 -23%

Both unconfined and confined self-healing measurements revealed that

geopolymer-based cementing materials have intrinsic self-healing property that is well-

suited for supporting long-term zonal isolation, as e.g. required by wells that are

abandoned or decommissioned. This self-healing capability of geopolymer can be

attributed to the formation of extra reaction product when the unreacted particles in its

microstructure get exposure to pore fluids.

The loss of zonal isolation after the cement has set can be attributed to mechanical

failure of the hardened cement or the de-bonding of cement sheath from the casing and/or

formation interfaces. Several previous approaches used to mitigate cement failures from

long-term exposures to stresses and temperatures include: (1) adding fibrous materials

into the cement slurry for improved toughness of the cement matrix (LaPrade and Low,

2003); (2) designing cement slurries with the capability to withstand the physical stresses

Page 133: Copyright by Xiangyu Liu 2017

113

that might be encountered during the lifetime of the well (Bybee, 2000; Stiles and

Hollies, 2002); (3) using “self-healing” sealant materials to re-seal the leak path if one is

present (Cavanagh et al., 2007; Reddy et al., 2010; Taoutaou et al., 2011). Note that the

first two approaches target improved failure resistance but provide no self-healing

benefits. The sealant method is usually composed of swelling additives that seal the

cracks when exposed to wellbore fluids or heat, in an attempt to re-establish hydraulic

isolation. It is currently unknown if this approach provides a durable solution for the

longer term. Moreover, unlike the latter method, which only provides sealing of cracks,

geopolymer actually closes any micro-cracks by intrinsically developing more reaction

product, thus providing a superior solution for zonal isolation. A material with self-

healing capability is particularly well suited for wells that will undergo a lot of in-situ

stress variation caused by depletion of the reservoir, hydraulic fracturing operation, or

other geomechanical loading. For well abandonment applications, where long-term

cement integrity is critical, the self-healing geopolymer cement presents a viable option

as a permanent sealant that can re-heal if it becomes compromised. Thus the study of

long-term self-healing properties of geopolymer is encouraged in order to access the

capability of geopolymer to provide effective isolation over a long period of time, even

decades.

CEMENT-TO-PIPE BOND STRENGTH 6.4

When the cement-to-pipe shear bond strength was tested for geopolymer samples

on a clean steel bar at 7 days, the peak bond strength was above 200 psi for both

replicates (Figure 6.8). Beyond this, the load dropped at a decreasing rate due to the

sliding stress (or skin friction) against the pipe (Nahm et al., 1995). In comparison,

Page 134: Copyright by Xiangyu Liu 2017

114

Portland cement samples had a peak value less than 100 psi. The 7-day unconfined

compressive strength for geopolymer and Portland cement is 1,640 psi and 4,100 psi

respectively. The shear bond to compressive strength ratio of geopolymer is therefore

13%. This ratio for hardened Portland cement slurry, however, is only 2%.

Figure 6.8 - Cement-to-pipe shear bond strength for two samples of Portland cement and

geopolymer with clean steel pipe at 170 °F on day-7

The cement-to-pipe shear bond strength was also evaluated when the steel bar was

pre-coated with SBM. This simulates the less-than-optimal conditions when the SBM is

not completely displaced from the surface of the casing. As shown in Figure 6.9, the

presence of SBM significantly lowered the bonding capability of both slurries. The pipe

bond strength of geopolymer is around 30 psi, and is slightly higher than the values for

Portland cement (10 psi). The effect of SBM can also be visually observed as shown in

0

50

100

150

200

250

0 50 100 150 200

Cem

ent-

to-P

ipe

Shea

r B

on

d S

tren

gth

(psi

)

Time (s)

Portland (1)

Portland (2)

Geopolymer (1)

Geopolymer (2)

Page 135: Copyright by Xiangyu Liu 2017

115

Figure 6.10. The color of the hardened slurries close to the pipe surface is darker than the

rest of the samples, indicating the presence of SBM.

Figure 6.9 - Cement-to-pipe shear bond strength for two samples of Portland cement and

geopolymer with SBM coated steel pipe at 170 °F on day-7

(a) (b)

Figure 6.10 - Cross-section view from bottom of the cement-to-pipe shear bond test

samples for (a) Portland cement and (b) geopolymer

0

10

20

30

40

50

0 20 40 60 80 100 120 140 160 180

Cem

ent-

to-P

ipe

Sh

ear

Bo

nd

Str

ength

(psi

)

Time (s)

Portland (1)

Portland (2)

Geopolymer (1)

Geopolymer (2)

Page 136: Copyright by Xiangyu Liu 2017

116

HYDRAULIC CONDUCTIVITY 6.5

As mentioned in chapter 3, the hydraulic conductivity of Portland cement,

geopolymer and alkali-activated slag samples were measured using the pressure

transmission test.

The pressure transmission curves of hardened Portland cement when injecting

artificial seawater (ASW) in first cycle and light mineral oil (LMO) in the second cycle is

shown in Figure 6.11 (a). As can be seen in the figure, the pressure transmission was

much slower when injecting LMO compared to ASW due to the water-wet nature of the

cement sample. The linear trend lines for both pressure transmission curves were also

plotted. By comparing the slope of the linear fit lines, a delay factor around 10 was

obtained, indicating that the pressure invasion of the oil phase will be 10 times slower

when Portland cement was exposed to LMO. The pressure transmission curves for

hardened geopolymer and alkali-activated slag are shown in Figure 6.11 (b) and (c)

respectively. Compared to Portland cement, both geopolymer and slag showed a much

slower pressure transmission rate when injecting LMO. The delay factors for these two

alkali-activated materials were 35 and 45 respectively, much higher than the value of

Portland cement, indicating that it takes longer time for oil phase pressure to transmit

through geopolymer and slag.

Page 137: Copyright by Xiangyu Liu 2017

117

Figure 6.11 - Pressure transmission curves, linear fit and the delay factor for (a) Portland

cement, (b) geopolymer, and (c) slag at 28 days

y = 2.0E-03x

y = 2.2E-04x

0

1

2

3

4

5

0 2000 4000 6000 8000 10000 12000 14000 16000

ln((

Pm

-Po

)/(P

m-P

t))

Time (s)

(a) Portland_ASW

Portland_LMO

Delay factor ~ 10

y = 1.6E-03x

y = 4.8E-05x

0

1

2

3

4

5

0 2000 4000 6000 8000 10000 12000 14000 16000

ln((

Pm

-Po)/

(Pm

-Pt)

)

Time (s)

(b) Geopolymer_ASW

Geopolymer_LMO

Delay factor ~ 35

y = 4.0E-03x

y = 8.7E-05x

0

1

2

3

4

5

0 2000 4000 6000 8000 10000 12000 14000 16000

ln((

Pm

-Po

)/(P

m-P

t))

Time (s)

(c) Slag_ASW

Slag_LMO

Delay factor ~ 45

Page 138: Copyright by Xiangyu Liu 2017

118

To better compare the hydraulic conductivity of these three cementitious

materials, the linear fit trend lines of the ASW pressure transmission curves were

replotted and are shown in Figure 6.12. The hydraulic conductivity values were

calculated based on the slope of the linear fit lines and are indicated in the figure. As can

be seen, the conductivities of geopolymer and Portland cement were comparable and

were both lower than the conductivity of slag. Based on the definition of hydraulic

conductivity (equation 3.4), the permeability is proportional to the hydraulic conductivity

for a specific fluid. Cement with lower permeability is less vulnerable to external fluid

invasion, which is a step in the right direction for ensuring good zonal isolation. In this

study, assuming the LMO has a viscosity of 30 mPa·s and a compressibility of 4.3E-10

Pa-1

at the test pressure and temperature, all three hardened materials had oil permeability

values in the range of 300-1800 μD. It can be speculated that within this permeability

range, hydrocarbon would not migrate through the cement matrix at any detectable rate.

Figure 6.12 - Hydraulic conductivity (HC) and linear fit trend lines of Portland cement,

geopolymer, and slag at 28 days

0

1

2

3

4

5

6

0 500 1000 1500 2000 2500 3000

Ln

((P

m-P

o)/

(Pm

-Pt)

)

Time (s)

Linear (Portland_ASW)

Linear (Geopolymer_ASW)

Linear (Slag_ASW)HC = 2.5E-06 m2/s

HC = 1.2E-06 m2/s

HC = 1.0E-06 m2/s

Page 139: Copyright by Xiangyu Liu 2017

119

POROSITY 6.6

Porosity is another factor that influences the ability of cement to prevent gas

migration and to provide zonal isolation. Space for gas or hydrocarbon to enter the

annulus includes the inherent porosity of the cement, any cracks or fractures in the

cement sheath, and any gap at the cement bonding interfaces. In the previous sections the

ability of geopolymer to resist cracking under triaxial stresses and to bond towards casing

has been discussed. Here the porosity of geopolymer is studied and compared with

conventional Portland cement.

In literature, multiple approaches have been adapted to characterize the porosity

and the pore size distribution of the hardened cement matrix, however, an exact pore

structure characterization remains difficult to achieve. A hardened cement sample

contains pores that have various pore sizes, shapes, interconnectivity, accessibility and

surface roughness. In general, the pore structure of cementitious materials contains air

voids, capillary pores, and gel pores. The pores in hardened cement slurries can be

classified into two categories, gel pores (< 10 nm) that are formed within hydration

products, and capillary pores (10 nm – 10000 nm) that are inter-connected and dominates

diffusivity (Mindess et al., 2003).

The porosity of geopolymer characterized with transmission electron microscopy

(TEM) and gas adsorption porosimetry has been reported in the literature (Kriven et al.,

2006; Metroke et al., 2010). The porosity of geopolymer forms when considerable

amount of water is forced out during the polymerization process and resides between the

precipitates. The pores have a length scale of 5-20 nm. Studies showed that these pores

are filled with free water (Bell et al., 2010, 2009).

The porosity and total pore area obtained from the MIP measurement were

reported in Table 6.4. As shown in the table, the porosity values of all three hardened

Page 140: Copyright by Xiangyu Liu 2017

120

cement slurries were similar at approximately 15%. It can also be seen that the slag

sample showed smaller pore diameter and larger pore area, indicating the slag had

smaller pores compared to the other two cement slurries. The pore size distribution plot

also confirmed this observation. As shown in Figure 6.13, the pore size distribution of

Portland cement and geopolymer were very similar with a normal distribution-like curve

with a single peak at 30 nm on a log scale. The slag sample, on the other hand, was a

bimodal distribution with peak values at 9 nm and 14 nm.

Table 6.4 - Porosity and pore size evaluation of Portland, geopolymer and slag from MIP

measurement at 7 days

MIP-Porosity

(%)

Total Pore Area

(m2/g)

Avg. Pore

Diameter (µm)

Portland Cement 17 8.5 0.048

Geopolymer 14 8.3 0.044

Slag 16 24 0.017

Page 141: Copyright by Xiangyu Liu 2017

121

Figure 6.13 - Pore size distribution of Portland cement, geopolymer and slag obtained

from MIP test, samples were cured at 170 °F for 28 days

It is important to mention that mercury intrusion porosimetry (MIP) test has its

own limitations. MIP cannot provide a true pore size distribution because mercury cannot

pass through the narrowest pores or the isolated pores in the pore network. The total

porosity estimated from MIP test will also differ from the values obtained by other

techniques. The MIP porosity will usually be smaller than the true porosity values due to

inaccessible pores. On the other hand, there are studies that believe the MIP porosity can

be closer to the true value where mercury pressures can collapse small pores so that the

isolated pores become accessible (Beaudoin, 1979; Diamond, 1971).

0

0.005

0.01

0.015

0.02

0.025

0.03

1 10 100 1,000

Incr

emen

tal

Intr

usi

on

(m

L/g

)

Pore Size (nm )

Portland cement

Geopolymer

Slag

Page 142: Copyright by Xiangyu Liu 2017

122

ULTRASONIC CEMENT STRENGTH 6.7

As mentioned in the background chapter, cement strength can be tested via

destructive method, i.e., hardened cement cubes or cylinders of a particular age can be

crushed for its strength. Non-destructive strength testing with ultrasonic cement analyzer

(UCA) is a method that can continuously measure the compressive strength over a period

of time. The strength values are obtained based on empirical correlations between sonic

wave transit time and UCS that were pre-programmed in the UCA. The main drawback

of UCA is the use of a single set of empirical correlations for different formulations.

Regardless of the limitations with UCA test, it is still a popular method to evaluate the

compressive strength of Portland cement as it allows a continuous measurement of the

strength development profile. API recommended practice (API RP 10B-2, 2010) has

included both destructive and non-destructive test methods as means to quantify the

strength values of hardened cement.

In an effort to develop geopolymer for well cementing purposes, the present study

attempted to run geopolymer in UCA following API standard (API RP 10B-2, 2010).

Geopolymer sample was formed with 8 M NaOH activation and fly ash FA1. The

pressure was stepped up to 5,000 psi and the temperature was ramped from room

temperature to 189 °F in 2 hours. Figure 6.14 shows the compressive strength and transit

time data of geopolymer directly obtained from built-in UCA correlations that were

developed for Portland cement. As can be seen in the figure, for geopolymer sample the

transit time varied between 9.5 and 11.5 sec/in and the value reached a peak at

approximately 2 hours before starting to ramp down. For a typical Portland cement

slurry, on the other hand, the transit time started with a higher value around 17 sec/in

and ramped down once the cement started to develop compressive strength (Figure 6.15).

The main issue can be noticed by observing the compressive strength curve. The built-in

Page 143: Copyright by Xiangyu Liu 2017

123

correlations from UCA resulted in negative compressive strength values for the

geopolymer sample for the first 19 hours. The 24 hour compressive strength eventually

reached 235 psi, far from the actual UCS value measured by crushing geopolymer cubes

cured for 24 hours (1,280 psi). Clearly the existing correlations that were developed for

Portland cement do not apply to geopolymer chemistries. This could be attributed to the

fact that fly ash particles are spherical and a lot of times hollow. The reflective and

refractive indices for fly ash differ significantly from Portland cement particles which

leads to a change in the sonic wave propagation path and the transit time.

Figure 6.14 - Geopolymer sonic compressive strength and transit time obtained from

UCA test with built-in empirical correlations developed for Portland cement, test was

conducted at 189 °F and 5,000 psi

9

9.5

10

10.5

11

11.5

12

-600

-400

-200

0

200

400

600

0:00:00 6:00:00 12:00:00 18:00:00 24:00:00

Tra

nsi

t T

ime

(use

c/in

)

Com

pre

ssiv

e S

tren

gth

(psi

)

Time (hh:mm:ss)

Compressive Strength Transit Time

Page 144: Copyright by Xiangyu Liu 2017

124

Figure 6.15 - Portland cement sonic compressive strength and transit time obtained from

UCA test with built-in empirical correlations developed for Portland cement, test was

conducted at 170 °F and 3,000 psi

As discussed in the literature review section, Khalifeh et al. (Khalifeh et al., 2014)

developed a set of UCA correlations for geopolymer (Table 2.1). Here the applicability of

these correlations was tested for their effectiveness in evaluating geopolymer

compressive strength. Notice that geopolymer slurry was formed with class F fly ash as

opposed to the class C fly ash in their study. The same testing temperature and pressure

conditions were used in this study for a direct comparison. Figure 6.16 shows the

compressive strength curve when the published correlations were applied. As shown in

the figure, the negative compressive strength problem was fixed after 2.5 hours.

However, the compressive strength values were above 500 psi for the first two hours,

which does not match the UCS values. In addition, 24 hour cube strength was 1,280 psi,

0

4

8

12

16

20

24

0

1000

2000

3000

4000

5000

6000

0:00:00 12:00:00 24:00:00 36:00:00 48:00:00 60:00:00 72:00:00

Tra

nsi

t T

ime

(use

c/in

)

Co

mp

ress

ive

Str

ength

(p

si)

Time (hh:mm:ss)

Compressive Strength Transit Time

Page 145: Copyright by Xiangyu Liu 2017

125

whereas the 24-hour strength value based on the correlations was more than 6,000 psi.

This result indicated that the published correlations for class C fly ash do not apply to

class F fly ash. More factors including differences in source materials, types of activator

(hydroxide vs. silicate activation), and activator-to-fly ash ratios could all influence the

ultrasonic wave transmission, thus should be considered when developing UCA

correlations for geopolymers.

Figure 6.16 - Geopolymer sonic compressive strength and transit time obtained from

UCA test with empirical correlations developed for geopolymer, test was conducted at

189 °F and 5,000 psi

SUMMARY 6.8

In this chapter, several critical hardened state properties of geopolymer were

evaluated in detail, including mechanical properties, self-healing properties, cement-to-

9

9.5

10

10.5

11

11.5

12

-2000

0

2000

4000

6000

8000

10000

0:00:00 6:00:00 12:00:00 18:00:00 24:00:00

Tra

nsi

t T

ime

(use

c/in

)

Com

pre

ssiv

e S

tren

gth

(psi

)

Time (hh:mm:ss)

Compressive Strength Transit Time

Page 146: Copyright by Xiangyu Liu 2017

126

pipe bond strength, hydraulic conductivity and porosity. The experimental results

indicated that geopolymers can serve as a viable alternative to Portland cement for

primary cementation and as potentially superior candidates for improving barrier integrity

in abandoned wells. The key findings of this work include:

1. This work shows that geopolymers and geopolymer hybrid cements have more

than sufficient compressive strength when subjected to confining stress for most

cementing purposes.

2. Under triaxial loading, both geopolymer and geopolymer hybrid samples

exhibited more ductile behavior in comparison to Portland cement, showcasing

the ability of geopolymer materials to better deal with surrounding stress

fluctuations.

3. The self-healing study demonstrated that geopolymers have great potential to be

used as superior alternatives to Portland cement in well abandonment /

decommissioning applications. The post-failure self-healing capability of

geopolymer was verified under conditions of uniaxial and triaxial loading to

failure. Geopolymers exhibited excellent self-healing of their matrix, with the

material being able to withstand higher stress after they were pre-loaded. Such

behavior was not observed for Portland cement.

4. The cement-to-pipe bonding properties of geopolymer were also studied. When

the steel bar was either clean or was coated with SBM, geopolymers showed

consistently higher shear bond strength values compared with Portland cement.

5. The hydraulic conductivity of geopolymer was compared with hardened Portland

cement slurry and an alkali-activated slag. When injecting ASW, the conductivity

values of geopolymer and Portland were comparable, and they were both lower

than the conductivity of slag. When injecting LMO, the geopolymer and slag

Page 147: Copyright by Xiangyu Liu 2017

127

samples showed higher delay factor than Portland cement, indicating a lower

diffusivity to oil phase.

6. By conducting MIP tests, the porosity of geopolymer was evaluated. The porosity

and pore size distribution of hardened geopolymer and Portland cement were

comparable, whereas the alkali-activated slag showed similar total porosity but

smaller pore sizes.

7. Existing UCA correlations that were developed for Portland cement cannot be

applied to geopolymer slurries because the equations generated negative

compressive strength values. The correlations developed for class C fly ash

geopolymer from a recently published paper yielded positive compressive

strength values, however, the values failed to match the UCS values obtained by

crushing cement cubes made from class F fly ash.

Page 148: Copyright by Xiangyu Liu 2017

128

Chapter 7: Conclusions and future work

The main goal of this research was to investigate and develop a versatile mud

solidification method that is applicable to any type of NAF. In this dissertation a

solidification method was successfully developed by blending SBM / OBM with an

appropriate volume of geopolymer. The resulting hybrid cement delivered many of the

necessary characteristics of an oilwell cement slurry. Through detailed evaluation and

verification, it was shown that the geopolymer hybrid cement could be potentially used

for various applications including primary cementing, lost circulation control, well

abandonment and decommissioning, etc.

In this dissertation, background information about the geopolymer design and

mud solidification method was provided in Chapter 2. Chapter 3 described the materials

and experimental methods that were used in this research. Chapter 4 thoroughly

evaluated the effectiveness of geopolymer to solidify SBM, and characterized the

compressive strength, rheological properties and thickening time behavior of the

geopolymer hybrid cement system. Factors including aluminosilicate source, molarity of

alkaline activator, presence of seawater, composition of SBM were all considered.

Chapter 5 investigated the effectiveness of silicate activation on solidifying SBM.

Chapter 6 investigated the hardened state properties such as mechanical properties,

confined compressive strength, self-healing capability, porosity and hydraulic

conductivity that are crucial for well integrity. Finally this chapter provides key

conclusions from this study along with suggestions for future work.

Page 149: Copyright by Xiangyu Liu 2017

129

CONCLUSIONS 7.1

It has been shown that geopolymers have superior compatibility with NAFs

compared to traditional Portland cement. Geopolymers not only develop sufficient

compressive strength at high levels of mud contamination but also, by incorporating

SBM into geopolymer, their rheological properties can be improved remarkably, allowing

for a pumpable slurry for well cementing operations. By varying the volume ratio of

geopolymer and mud, the properties of the geopolymer hybrid cement can be custom

tailored to the target application. The key findings and conclusions of this research study

are as follows:

Geopolymers showed much better compatibility with NAF-based drilling muds

compared to Portland cement slurries. It was found that geopolymer could solidify

up to 40% of SBM and still develop measurable compressive strength, while

Portland cement slurry lost the ability to harden at the same level of mud

contamination. This observation was the foundation for the development of the

mud solidification technique presented here.

The incorporation of SBM into geopolymer slurries significantly improved the

rheological properties of geopolymer at both room temperature and elevated

temperature, allowing for a pumpable slurry for well cementing applications. In

fact, unlike Portland cement slurries, which become more viscous at elevated

temperatures, geopolymer and geopolymer hybrids showed lower rheological

readings with increasing temperature.

Geopolymer successfully solidified two types of SBMs and two types of OBMs.

The composition of SBM showed limited effects on the strength and rheological

properties of the geopolymer hybrids, meaning that the mud-to-cement conversion

is a versatile method that can be adapted to work with various drilling muds. This

Page 150: Copyright by Xiangyu Liu 2017

130

study further proved that changing synthetic/water ratio or calcium chloride

concentration of SBM had no significant effect on the compressive strength or

rheology of geopolymer hybrid. Changing the synthetic/water ratio of SBM from

70/30 to 60/40 did not alter the thickening time, but lowering the calcium

concentration of the invert phase slightly accelerated the thickening time.

By comparing three different types of fly ashes, it was found that this mud-to-

cement conversion technique is not limited to a single source material, an

important consideration for practical field application which cannot rely on just a

single source. All three types of fly ashes tested in this study showed potential for

formulating geopolymers that are suitable for mud-to-cement conversion. Varying

compositions of fly ashes could lead to varying rheological properties and

pumping time. Therefore, standard API tests have to be conducted for each source

material, e.g. during routine testing in preparation for field application, similar to

Portland cement slurry design.

Triaxial compressive strength testing showed that geopolymer and geopolymer

hybrid cements had more than sufficient compressive strength when subjected to

confinement for most cementing operations. Neat geopolymer and geopolymer

hybrids exhibited more ductile behavior in comparison to Portland cement,

indicating the ability of geopolymer-based material to better accommodate and

deal with surrounding stress fluctuations.

Both unconfined and confined self-healing tests revealed the self-healing

capability of geopolymer and geopolymer hybrids. This was confirmed by the re-

healed peak stress of the cement matrix exceeding the maximum pre-loading

stress. The intrinsic self-healing characteristic is distinctly different from - and is

superior to -current mechanisms for creating self-healing Portland cements, which

Page 151: Copyright by Xiangyu Liu 2017

131

rely on the re-sealing of micro fractures with chemical sealants or with the use of

fibrous materials to reduce crack growth.

The rheological properties, short term and long term compressive strength of the

geopolymer hybrid can be tailored to the target application with the use of an

appropriate volume of SBM, or by altering the alkaline activator composition.

1. Increasing the amount of SBM in the geopolymer hybrids resulted in

lower rheological readings, lower compressive strength values and longer

thickening time.

2. Activation with liquid sodium silicate (LSS) resulted in unfavorable

rheology readings and gel strength values for both geopolymer and

geopolymer hybrids. By contrast, activation with solid sodium silicate

(SSS) did not affect the rheological behavior of geopolymer in the

presence or absence of SBM.

3. For neat geopolymer slurries, LSS activation accelerated the pumping

time. By comparison, SSS activation retarded the pumping time in

comparison to hydroxide activation. The presence of SBM significantly

accelerated the pumping time for both silicate and hydroxide activation.

The SSS activation showed an initial high consistency peak that broke

down with continued shearing.

4. Silicate activation greatly improved the compressive strength compared

with pure hydroxide activation. However, the effectiveness of strength

enhancement highly depended on the aluminosilicate source selected.

Other properties that are critical to oil / gas well cement slurry design were also

investigated. It was found out that:

Page 152: Copyright by Xiangyu Liu 2017

132

1. Similar to Portland cement slurries, increased pressure shortened the

thickening time of geopolymer hybrids.

2. The hydraulic conductivity and porosity values obtained for geopolymer

were comparable to Portland cement values.

3. It was found that using seawater did not affect the rheological properties

or compressive strength of geopolymer hybrid slurries.

Based on the key findings listed above, Figure 7.1 illustrates the slurry design

philosophy that can be adapted to formulate geopolymer hybrid cements. In simple terms,

a geopolymer hybrid cement can be formed when an aluminosilicate source, an alkaline

activator, and SBM or OBM are combined. The fresh state properties including surface

and downhole rheological properties and mixture stability should be first evaluated,

followed by the thickening time test, in order to confirm the pumpability of the slurry.

Next, the hardened state properties including compressive strength, mechanical properties

and bonding characteristics should be evaluated. When any of the properties fail to meet

the requirements of the target application, the slurry design parameters should be

reconsidered. Either property-modifying additives need to be added, or the main

components have to be adjusted. It is very important that key properties are evaluated at

representative downhole conditions, particularly using appropriate pressure and

temperature and their ramp-up history representative of the downhole circulation of the

cement. By performing this iteration, a geopolymer hybrid slurry with desired properties

can be achieved.

Page 153: Copyright by Xiangyu Liu 2017

133

Figure 7.1 - Circular flow diagram showing the design philosophy of geopolymer/mud

hybrid cement

In conclusion, with this new geopolymer-based mud solidification technology, a

wide range of well construction applications open up, including the use of SBM/OBM

compatible lost circulation treatments and plug cementations, the use of geopolymers in

spacer and scavenger fluids prior to primary cementation, the use of geopolymer slurries

in primary cementation, all the way to well abandonment and decommissioning

Fresh-state

Properties

Hardened-state Properties

Slurry Design

• Source material

• Alkali activator

• Admixtures

• Geopolymer/mud ratio

• Mud composition

• Rheological properties

• Thickening time

• Dynamic stability

• Compressive strength

• Compatibility with

various mud

• Mechanical properties

• Bonding properties

• Self-healing properties

• Porosity

• Permeability

Page 154: Copyright by Xiangyu Liu 2017

134

operations. The enhanced compatibility with SBM and OBM may lead to step-change

improvements in deepwater, narrow-margin cementations in particular.

FUTURE WORK 7.2

One of the main limitations with the proposed mud solidification method is the

variation in performance with different source materials. This has been observed with

compressive strength test, thickening time test and rheological property measurements. It

remains unclear which component(s) of the aluminosilicate source affects each one of the

cement properties. It would be valuable to correlate the type and quantity of reactive

contents in the source material to the performance of the resulting slurry.

The other main limitation associated with the current mud solidification method is

that rheology control is mainly achieved by changing the proportion of mud in the

geopolymer hybrid, at the cost of compressive strength reduction. Future work has to

focus on identifying effective superplastisizers / rheology modifying additives that do not

negatively affect other key properties of the slurry.

In terms of thickening time control, the current study was focused on

understanding the changes in the thickening time of geopolymer hybrids with sodium

hydroxide and sodium silicate activation. It has been shown that the fresh state properties

of geopolymer hybrids were significantly affected by the types and dosage of the

activator and by the presence of drilling mud. Future work should focus on further

investigation of the interaction between the mud and geopolymers formed with different

activators and mixed activators. Moreover, further laboratory investigation is necessary to

identify other accelerators and retarders for thickening time control.

Page 155: Copyright by Xiangyu Liu 2017

135

In the present study, the ability of geopolymer hybrid cement to resist excessive

loads triggering yielding in a well has been investigated by conducting confined

compressive strength test and self-healing test. Future research should also look at the

tensile strength and toughness of the hardened geopolymer hybrids, which are all critical

indicators of a cement system’s ability to provide zonal isolation throughout the lifetime

of a well and after abandonment. The present study has attempted to measure the tensile

strength of geopolymer by conducting Brazilian splitting tensile strength tests. However,

the experimental error of this measurement was too large to obtain a good average

strength value. Future work can be directed towards bending or flexural tests. When the

goal of the cement slurry design is for well abandonment and decommissioning, other

properties that need to be evaluated for geopolymer-based cement include:

(1) Volumetric and bulk shrinkage and expansion.

(2) Permeability recovery after yielding / cracking and re-healing, with the ability

to withstand pressure loads associated with hydrocarbon fluid and gas columns.

(3) Long term durability and chemical resistance to wellbore and reservoir fluids

(brine or hydrocarbons).

(4) Static and dynamic stirred fluid loss under differential pressure.

(5) Static gel strength during setting for gas migration control.

(6) Stability control additives for use in blended geopolymer slurries without

significantly affecting the consistency.

(7) Free fluid control.

Low-calcium fly ash-based geopolymers generally requires elevated temperature

curing. The present study only focused on downhole temperature regimes above 125 °F.

In case of deepwater drilling or arctic drilling, where the temperature approaches freezing

point, the slurry compositions presented here would not harden at a practical rate. Future

Page 156: Copyright by Xiangyu Liu 2017

136

work should develop slurry compositions for applications at temperatures below 125 °F.

Several approaches that could enable effective geopolymer activation for low temperature

curing include: (1) utilization of ground fly ash or ultra-fine fly ash with larger surface

area for higher reaction rate; (2) using high-calcium fly ash or blending in calcium

silicate (BFS or Portland cement) to incorporate C-S-H gel in the aluminosilicate

structure.

Page 157: Copyright by Xiangyu Liu 2017

137

List of Abbreviations

AAM Alkali-activated material

AAAS Aqueous alkali alumino silicate

AAS Alkali-activated slag

API American Petroleum Institute

ASTM American Society for Testing and Materials

ASW Artificial seawater

BHCT Bottom hole circulating temperature

BHP Bottom hole pressure

Bwoc By wait of cement

Bwos By wait of slag

BFS Blast furnace slag

DI Deionized

EDS Energy dispersive spectroscopy

EPA Environmental Protection Agency

FA Fly ash

GH Geopolymer hybrid

HC Hydraulic conductivity

HPHT High pressure high temperature

IPA Isopropyl alcohol

LCM Lost circulation material

LMO Light mineral oil

LSS Liquid-form sodium silicate

MIP Mercury intrusion porosimetry

Page 158: Copyright by Xiangyu Liu 2017

138

NAF Non-aqueous drilling fluid

OBM Oil-based mud

OPC Ordinary Portland cement

P&A Plug and abandonment

PSD Particle size distribution

PTT Pressure transmission test

RC Resistor-capacitor

RP Recommended practice

SBM Synthetic-based mud

SEM Scanning electron microscope

S / S Solidification / stabilization

SSS Solid-form sodium silicate

SWR Synthetic / water ratio

TEM Transmission electron microscopy

UCA Ultrasonic cement analyzer

UCS Unconfined compressive strength

UF Universal fluid

WBM Water-based mud

XRD X-ray diffraction

Page 159: Copyright by Xiangyu Liu 2017

139

List of Key Symbols

σ1 Total axial stress

σ3 Confining stress

σ1- σ3 Differential stress

σa Axial stress

σHg Air-mercury interfacial tension

A Sample cross-sectional area

E Young’s Modulus

Pc Confining pressure

PHg Mercury injection pressure

Pm Upstream fluid pressure

P(l,t) Downstream pressure transmission at sample end as a function of time

Po Initial pore pressure

V Volume of downstream reservoir

l sample length

rp Pore throat radius

β Fluid compressibility

μ Fluid viscosity

ν Poisson’s ratio

k Relative permeability

ϕ porosity

θHg Mercury contact angle

εa Axial strain

εr Radial strain

Page 160: Copyright by Xiangyu Liu 2017

140

List of Publications

1. Liu, X., Aughenbaugh, K., Lee, H., Nair, S., & Oort, E. van., 2017. Geopolymer -

Synthetic Based Mud Hybrid Cements for Primary Cementing and Lost

Circulation Control. SPE International Conference on Oilfield Chemistry.

doi:10.2118/184558-MS

2. Liu, X., Ramos, M. J., Nair, S. D., Lee, H., Espinoza, D. N., & van Oort, E., 2017.

True Self-Healing Geopolymer Cements for Improved Zonal Isolation and Well

Abandonment. SPE/IADC Drilling Conference and Exhibition.

doi:10.2118/184675-MS

3. Liu, X., Aughenbaugh, K., Nair, S., Shuck, M., & van Oort, E., 2016.

Solidification of Synthetic-Based Drilling Mud Using Geopolymers. SPE

Deepwater Drilling and Completions Conference. doi:10.2118/180325-MS

4. van Oort, E., Aughenbaugh, K., Nair, S.D. & Liu, X., 2016. “Cementitious

Compositions Comprising a non-aqueous fluid and an alkali-activated material”,

Application No. 15/355,586

5. Liu, X., Nair, S. D., Cowan, M., & van Oort, E., 2015. A Novel Method to

Evaluate Cement-Shale Bond Strength. SPE International Symposium on Oilfield

Chemistry. doi:10.2118/173802-MS

Page 161: Copyright by Xiangyu Liu 2017

141

Bibliography

Ahn, T.-H., Kishi, T., 2010. Crack Self-healing Behavior of Cementitious Composites

Incorporating Various Mineral Admixtures. J. Adv. Concr. Technol. 8, 171–186.

doi:10.3151/jact.8.171

Al Bakri, A.M.M., Kamarudin, H., Nizar, K., et al., 2011. Correlation between

Na2SiO3/NaOH Ratio and Fly Ash/Alkaline Activator Ratio to the Strength of

Geopolymer. Adv. Mater. Res. 341–342, 189–193.

doi:10.4028/www.scientific.net/AMR.341-342.189

Antoni, A., Wijaya, S.W., Hardjito, D., 2016. Factors Affecting the Setting Time of Fly

Ash-Based Geopolymer. Mater. Sci. Forum 841, 90–97.

doi:10.4028/www.scientific.net/MSF.841.90

API RP 10B-2, 2010. Recommended Practice for Testing Well Cements.

ASTM C618-18, 2015. Standard Specification for Coal Fly Ash and Raw or Calcined

Natural Pozzolan for Use in Concrete.

Aughenbaugh, K.L., 2013. Fly ash-based Geopolymers : Identifying Reactive Glassy

Phases in Potential Raw Materials.

Aughenbaugh, K.L., Nair, S.D., Cowan, K., et al., 2014. Contamination of Deepwater

Well Cementations by Synthetic-Based Drilling Fluids, in: SPE Deepwater

Drilling and Completions Conference. Galveston, TX. doi:10.2118/170325-MS

Aughenbaugh, K.L., Williamson, T., Juenger, M.C.G., 2014. Critical Evaluation of

Strength Prediction Methods for Alkali-Activated Fly Ash. Mater. Struct. 48,

607–620. doi:10.1617/s11527-014-0496-z

Bachu, S., Bennion, D.B., 2009. Experimental Assessment of Brine and/or CO2 Leakage

through Well Cements at Reservoir Conditions. Int. J. Greenh. Gass Control 3,

494–501.

Bakharev, T., 2005. Geopolymeric Materials Prepared using Class F Fly Ash and

Elevated Temperature Curing. Cem. Concr. Res. 35, 1224–1232.

doi:10.1016/j.cemconres.2004.06.031

Barbosa, V.F.F., MacKenzie, K.J.D., Thaumaturgo, C., 2000. Synthesis and

Characterisation of Materials Based on Inorganic Polymers of Alumina and

Silica: Sodium Polysialate Polymers. Int. J. Inorg. Mater. 2, 309–317.

doi:10.1016/S1466-6049(00)00041-6

Beach, H.J., Goins, W.C.J., 1957. A Method of Protecting Cements Against the Harmful

Effects Of Mud Contamination.

Bear, J., 1972. Dynamics of Fluids in Porous Media, Environmental science series (New

York, 1972- ). American Elsevier Pub. Co, New York.

Beaudoin, J.J., 1979. Porosity Measurement of Some Hydrated Cementitious Systems by

High Pressure Mercury Intrusion-Microstructural Limitations. Cem. Concr. Res.

9, 771–781. doi:10.1016/0008-8846(79)90073-5

Bell, J., Gordon, M., Kriven, W., 2005. Use of Geopolymeric Cements as a Refractory

Adhesive for Metal and Ceramic Joins, in: Zhu, D., Plucknett, K. (Eds.),

Page 162: Copyright by Xiangyu Liu 2017

142

Advances in Ceramic Coatings and Ceramic-Metal Systems: Ceramic

Engineering and Science Proceedings. John Wiley & Sons, Inc., pp. 407–413.

Bell, J.L., Driemeyer, P.E., Kriven, W.M., 2009. Formation of Ceramics from

Metakaolin-Based Geopolymers: Part I: Cs-Based Geopolymer. J. Am. Ceram.

Soc. 92, 1–8. doi:10.1111/j.1551-2916.2008.02790.x

Bell, J.L., Driemeyer, Kriven, W.M., 2010. Formation of Ceramics from Metakaolin-

Based Geopolymers. Part II: K-Based Geopolymer. J. Am. Ceram. Soc.

Bell, J.L., Sarin, P., Driemeyer, P.E., et al., 2008a. X-Ray Pair Distribution Function

Analysis of a Metakaolin-based, KAlSi2O6·5.5H2O Inorganic Polymer

(Geopolymer). J. Mater. Chem. 18, 5974–5981. doi:10.1039/B808157C

Bell, J.L., Sarin, P., Provis, J.L., et al., 2008b. Atomic Structure of a Cesium

Aluminosilicate Geopolymer: A Pair Distribution Function Study. Chem. Mater.

20, 4768–4776. doi:10.1021/cm703369s

Benge, O.G., Webster, W.W., 1994. Evaluation of Blast Furnace Slag Slurries for

Oilfield Application. Presented at the SPE/IADC Drilling Conference, Society of

Petroleum Engineers. doi:10.2118/27449-MS

Bol, G.M., Wong, S.-W., Davidson, C.J., et al., 1994. Borehole Stability in Shales. SPE

Drill. Complet. 9, 87–94. doi:10.2118/24975-PA

Brough, A.R., Holloway, M., Sykes, J., et al., 2000. Sodium Silicate-Based Alkali-

Activated Slag Mortars: Part II. The Retarding Effect of Additions of Sodium

Chloride or Malic Acid. Cem. Concr. Res. 30, 1375–1379. doi:10.1016/S0008-

8846(00)00356-2

Bybee, K., 2000. New Cement Systems for Durable Zonal Isolation. J. Pet. Technol. 52,

32–33. doi:10.2118/0800-0032-JPT

Cantarel, V., Nouaille, F., Rooses, A., et al., 2015. Solidification / Stabilisation of Liquid

Oil Waste in Metakaolin-Based Geopolymer. J. Nucl. Mater. 464, 16–19.

doi:10.1016/j.jnucmat.2015.04.036

Cavanagh, P.H., Johnson, C.R., Le Roy-Delage, S., et al., 2007. Self-Healing Cement -

Novel Technology to Achieve Leak-Free Wells. Society of Petroleum Engineers.

doi:10.2118/105781-MS

Chancey, R.T., Stutzman, P., Juenger, M.C.G., et al., 2010. Comprehensive Phase

Characterization of Crystalline and Amorphous Phases of a Class F Fly Ash.

Cem. Concr. Res. 40, 146–156. doi:10.1016/j.cemconres.2009.08.029

Chang, J.J., 2003. A study on the Setting Characteristics of Sodium Silicate-Activated

Slag Pastes. Cem. Concr. Res. 33, 1005–1011. doi:10.1016/S0008-

8846(02)01096-7

Chindaprasirt, P., Silva, P.D., Sagoe-Crentsil, K., et al., 2012. Effect of SiO2 and Al2O3

on the Setting and Hardening of High Calcium Fly Ash-Based Geopolymer

Systems. J. Mater. Sci. 47, 4876–4883. doi:10.1007/s10853-012-6353-y

Cowan, K.M., Hale, A.H., Nahm, J.J., 1992. Conversion of Drilling Fluids to Cements

With Blast Furnace Slag: Performance Properties and Applications for Well

Cementing. Presented at the SPE Annual Technical Conference and Exhibition,

Society of Petroleum Engineers. doi:10.2118/24575-MS

Page 163: Copyright by Xiangyu Liu 2017

143

Criado, M., Fernández-Jiménez, A., de la Torre, A.G., et al., 2007. An XRD study of the

effect of the SiO2/Na2O ratio on the alkali activation of fly ash. Cem. Concr. Res.

37, 671–679. doi:10.1016/j.cemconres.2007.01.013

Criado, M., Fernández-Jiménez, A., Palomo, A., et al., 2008. Effect of the SiO2/Na2O

Ratio on the Alkali Activation of Fly Ash. Part II: 29Si MAS-NMR Survey.

Microporous Mesoporous Mater. 109, 525–534.

doi:10.1016/j.micromeso.2007.05.062

Criado, M., Palomo, A., Fernández-Jiménez, A., et al., 2009. Alkali Activated Fly Ash:

Effect of Admixtures on Paste Rheology. Rheol. Acta 48, 447–455.

doi:10.1007/s00397-008-0345-5

Darley, H.C.H., 1969. A Laboratory Investigation of Borehole Stability. J. Pet. Technol.

21, 883–892. doi:10.2118/2400-PA

Davidovits, J., 2015. Geopolymer Chemistry and Applications, 4th ed.

Davidovits, J., 1989. Geopolymers: Inorganic Polymeric New Materials., in: No. 8.

Presented at the Symposium on Chemical Thermodynamics, Calorimetry and

Thermal Analysis 1989, pp. 1633–1656.

Davidovits, J., 1982. Mineral Polymers and Methods of Making Them. US 4349386 A.

Diamond, S., 1971. Critical Comparison of Mercury Porosimetry and Capillary

Condensation Pore Size Distributions of Portland Cement Pastes. Cem. Concr.

Res.

Duran, L.E., 2015. The Effect of Immediate Saline Water Cruing on the Strength,

Ccomposition, and Microstracture of Geopolymer Cement. University of

Massachusetts Amherst.

Durdziński, P., Dunant, C., Haha, M., et al., 2015. A new Quantification Method Based

on SEM-EDS to Assess Fly Ash Composition and Study the Reaction of its

Individual Components in Hydrating Cement Paste. Cem. Concr. Res. 73, 111–

122. doi:10.1016/j.cemconres.2015.02.008

Duxson, P., Fernández-Jiménez, A., Provis, J.L., et al., 2006. Geopolymer Technology:

the Current State of the art. J. Mater. Sci. 42, 2917–2933. doi:10.1007/s10853-

006-0637-z

Duxson, P., Lukey, G.C., Separovic, F., et al., 2005a. Effect of Alkali Cations on

Aluminum Incorporation in Geopolymeric Gels. Ind. Eng. Chem. Res. 44, 832–

839. doi:10.1021/ie0494216

Duxson, P., Mallicoat, S.W., Lukey, G.C., et al., 2007a. The Effect of Alkali and Si/Al

Ratio on the Development of Mechanical Properties of Metakaolin-based

Geopolymers. Colloids Surf. Physicochem. Eng. Asp. 292, 8–20.

doi:10.1016/j.colsurfa.2006.05.044

Duxson, P., Provis, J.L., 2008. Designing Precursors for Geopolymer Cements. J. Am.

Ceram. Soc. 91, 3864–3869. doi:10.1111/j.1551-2916.2008.02787.x

Duxson, P., Provis, J.L., Lukey, G.C., et al., 2007b. The Role of Inorganic Polymer

Technology in the Development of “Green Concrete.” Cem. Concr. Res. 37,

1590–1597. doi:10.1016/j.cemconres.2007.08.018

Page 164: Copyright by Xiangyu Liu 2017

144

Duxson, P., Provis, J.L., Lukey, G.C., et al., 2005b. Understanding the Relationship

Between Geopolymer Composition, Microstructure and Mechanical Properties.

Colloids Surf. Physicochem. Eng. Asp. 269, 47–58.

doi:10.1016/j.colsurfa.2005.06.060

Enayatpour, S., van Oort, E., 2017. Advanced Modeling of Cement Displacement

Complexities. Presented at the SPE/IADC Drilling Conference and Exhibition,

Society of Petroleum Engineers. doi:10.2118/184702-MS

Fernández-Jiménez, A., de la Torre, A.G., Palomo, A., et al., 2006a. Quantitative

Determination of Phases in the Alkaline Activation of Fly Ash. Part II: Degree of

Reaction. Fuel 85, 1960–1969. doi:10.1016/j.fuel.2006.04.006

Fernández-Jiménez, A., Monzo, M., Vicent, M., et al., 2008. Alkaline Activation of

Metakaolin-Fly Ash Mixtures: Obtain of Zeoceramics and Zeocements.

Microporous Mesoporous Mater. 108, 41–49.

doi:10.1016/j.micromeso.2007.03.024

Fernández-Jiménez, A., Palomo, A., 2005a. Mid-infrared Spectroscopic Studies of

Alkali-Activated Fly Ash Structure. Microporous Mesoporous Mater. 86, 207–

214. doi:10.1016/j.micromeso.2005.05.057

Fernández-Jiménez, A., Palomo, A., 2005b. Composition and Microstructure of Alkali

Activated Fly Ash Binder: Effect of the Activator. Cem. Concr. Res. 35, 1984–

1992. doi:10.1016/j.cemconres.2005.03.003

Fernández-Jiménez, A., Palomo, A., 2003. Characterisation of Fly Ashes. Potential

Reactivity as Alkaline Cements. Fuel 82, 2259–2265. doi:10.1016/S0016-

2361(03)00194-7

Fernández-Jiménez, A., Palomo, A., Sobrados, I., et al., 2006b. The Role Played by the

Reactive Alumina Content in the Alkaline Activation of Fly Ashes. Microporous

Mesoporous Mater. 91, 111–119. doi:10.1016/j.micromeso.2005.11.015

Fletcher, R.A., MacKenzie, K.J.D., Nicholson, C.L., et al., 2005. The Composition Range

of Aluminosilicate Geopolymers. J. Eur. Ceram. Soc. 25, 1471–1477.

doi:10.1016/j.jeurceramsoc.2004.06.001

García-Lodeiro, I., Fernández-Jiménez, A., Blanco, M.T., et al., 2007. FTIR Study of the

sol–gel Synthesis of Cementitious Gels: C–S–H and N–A–S–H. J. Sol-Gel Sci.

Technol. 45, 63–72. doi:10.1007/s10971-007-1643-6

Giasuddin, H.M., Sanjayan, J.G., Ranjith, P.G., 2013. Strength of Geopolymer Cured in

Saline Water in Ambient Conditions. Fuel 107, 34–39.

doi:10.1016/j.fuel.2013.01.035

Hale, A.H., Cowan, K.M., 1994. Drilling and Cementing with Blast Furnace Slag/Silicate

Fluid.

Hale, A.H., Cowan, K.M., 1991. Solidification of Water Based Muds. US 505867A.

Hale, A.H., Cowan, K.M., Nahm, J.J., 1995. Universal Fluids for Drilling and Cementing

Wells. US 5464060 A.

Hardjito, D., Rangan, B.V., 2005. Development and Properties of Low-Calcium Fly Ash-

Based Geopolymer Concrete (University of Technology Research Report). Curtin

University, Perth, Australia.

Page 165: Copyright by Xiangyu Liu 2017

145

Hemmings, R.T., Berry, E.E., 1987. On the Glass in Coal Fly Ashes: Recent Advances,

in: Symposium Q – Fly Ash and Coal Conversion By-Products IV, MRS Online

Proceedings Library Archive. p. 3 (36 pages). doi:10.1557/PROC-113-3

Iverson, B., Darbe, R., McMechan, D., 2008. Evaluation of Mechanical Properties of

Cements. Presented at the The 42nd U.S. Rock Mechanics Symposium (USRMS),

American Rock Mechanics Association.

Jewell, B.B., Rathbone, R.F., 2009. Optical Properties of Coal Combustion Byproducts

for Particle-Size Analysis by Laser Diffraction. Coal Combust. Gasif. Prod.

Juenger, M.C.G., Winnefeld, F., Provis, J.L., et al., 2011. Advances in Aternative

Cementitious Binders. Cem. Concr. Res., Conferences Special: Cement Hydration

Kinetics and Modeling, Quebec City, 2009 & CONMOD10, Lausanne, 2010 41,

1232–1243. doi:10.1016/j.cemconres.2010.11.012

Khale, D., Chaudhary, R., 2007. Mechanism of Geopolymerization and Factors

Influencing its Development: a Review. J. Mater. Sci. 42, 729–746.

doi:10.1007/s10853-006-0401-4

Khalifeh, M., Hodne, H., Korsnes, R.I., et al., 2015. Cap Rock Restoration in Plug and

Abandonment Operations; Possible Utilization of Rock-based Geopolymers for

Permanent Zonal Isolation and Well Plugging. International Petroleum

Technology Conference. doi:10.2523/18454-MS

Khalifeh, M., Saasen, A., Vralstad, T., et al., 2014. Potential Utilization of Geopolymers

in Plug and Abandonment Operations. Society of Petroleum Engineers.

doi:10.2118/169231-MS

Kong, D.L.Y., Sanjayan, J.G., 2010. Effect of Elevated Temperatures on Geopolymer

paste, Mortar and Concrete. Cem. Concr. Res. 40, 334–339.

doi:10.1016/j.cemconres.2009.10.017

Kong, D.L.Y., Sanjayan, J.G., Sagoe-Crentsil, K., 2007. Comparative performance of

geopolymers made with metakaolin and fly ash after exposure to elevated

temperatures. Cem. Concr. Res. 37, 1583–1589.

doi:10.1016/j.cemconres.2007.08.021

Kriven, W.M., Bell, J.L., Gordon, M., 2006. Microstructure and Microchemistry of

Fully-Reacted Geopolymers and Geopolymer Matrix Composites, in: Bansal,

N.P., Singh, J.P., Kriven, W.M., et al. (Eds.), Advances in Ceramic Matrix

Composites IX. John Wiley & Sons, Inc., pp. 227–250.

Kühl, H., 1908. Slag Cement and Process of Making the Same. US 900939.

Kutchko, B.G., Strazisar, B.R., Huerta, N., et al., 2009. CO2 Reaction with Hydrated

Class H Well Cement under Geologic Sequestration Conditions: Effects of Flyash

Admixtures. Environ. Sci. Technol. 43, 3947–3952. doi:10.1021/es803007e

LaPrade, D., Low, N., 2003. New Cement System Replaces Mechanical Whipstock Plug.

Society of Petroleum Engineers. doi:10.2118/83466-MS

Laudet, J., Garnier, A., Neuville, N., et al., 2011. The Behavior of Oil Well Cement at

Downhole CO2 Storage Conditions: Static and Dynamic Laboratory Experiments.

Energy Procedia, 10th International Conference on Greenhouse Gas Control

Technologies 4, 5251–5258. doi:10.1016/j.egypro.2011.02.504

Page 166: Copyright by Xiangyu Liu 2017

146

Lee, W.K.W., van Deventer, J.S.J., 2002a. The Effect of Ionic Contaminants on the

Early-age Properties of Alkali-Activated Fly Ash-Based Cements. Cem. Concr.

Res. 32, 577–584. doi:10.1016/S0008-8846(01)00724-4

Lee, W.K.W., van Deventer, J.S.J., 2002b. Effects of Anions on the Formation of

Aluminosilicate Gel in Geopolymers. Ind. Eng. Chem. Res. 41, 4550–4558.

doi:10.1021/ie0109410

Leimkuhler, J.M., Rainbow, F.H.K., Warren, P.B., et al., 1994. Downhole Performance

Evaluation of Blast Furnace Slag-Based Cements: Onshore and Offshore Field

Applications. Presented at the SPE Annual Technical Conference and Exhibition,

Society of Petroleum Engineers. doi:10.2118/28474-MS

Le-Minous, J.C., Mutti, D., Bouvet, A., et al., 2017. Permeability Study of API Class G

and B Cements Considering Seawater and WBM Contamination. Presented at the

SPE/IADC Drilling Conference and Exhibition, Society of Petroleum Engineers.

doi:10.2118/184613-MS

Lizcano, M., Gonzalez, A., Basu, S., et al., 2012. Effects of Water Content and Chemical

Composition on Structural Properties of Alkaline Activated Metakaolin-Based

Geopolymers. J. Am. Ceram. Soc. 95, 2169–2177. doi:10.1111/j.1551-

2916.2012.05184.x

Lloyd, R.R., 2009. Accelerated Ageing of Geopolymers, in: Provis, J.L., van Deventer,

J.S.J. (Eds.) Geopolymers: Structure, Processing, Properties and Industrial

Applications. Woodhead, Cambridge, pp. 139–166.

Lloyd, R.R., 2008. The Durability of Inorganic Polymer Cements. University of

Melbourne.

Ma, Y., Hu, J., Ye, G., 2012. The Effect of Activating Solution on the Mechanical

Strength, Reaction Rate, Mineralogy, and Microstructure of Alkali-Activated Fly

Ash. J. Mater. Sci. 47, 4568–4578. doi:10.1007/s10853-012-6316-3

Metroke, T.L., Henley, M.V., Hammons, M.I., 2010. Effect of Curing Conditions on the

Porosity Characteristics of Metakaolin–Fly Ash Geopolymers, in: Kriven, W.M.,

Zhou, Y., Radovic, M., et al. (Eds.), Strategic Materials and Computational

Design. John Wiley & Sons, Inc., pp. 11–15.

Miller, M., Scorsone, J., Whitfill, D.L., et al., 2013. The Development of a Geopolymer-

Based Pill as an Engineered Solution to Lost Circulation. Society of Petroleum

Engineers. doi:10.2118/166123-MS

Mindess, S., Young, J.F., Darwin, D., 2003. Concrete, Second. ed. Prentice Hall, USA.

Miranda, C.R., Carvalho, K.T., Vargas, A.A., et al., 2007. Minimizing Fluid

Contamination During Oilwell Cementing Operations. Presented at the Offshore

Mediterranean Conference and Exhibition, Offshore Mediterranean Conference.

Morgan, B.E., Dumbauld, G.K., 1952. Use of Activated Charcoal in Cement To Combat

Effects of Contamination by Drilling Muds. J. Pet. Technol. 4, 225–232.

doi:10.2118/952225-G

Mueller, D.T., Dickerson, J.P., 1994. Blast Furnace Slag Technology: Features,

Limitations, and Practical Applications. Presented at the SPE Annual Technical

Page 167: Copyright by Xiangyu Liu 2017

147

Conference and Exhibition, Society of Petroleum Engineers. doi:10.2118/28475-

MS

Nahm, J.J., Javanmardi, K., Cowan, K.M., et al., 1993. Slag Mix Mud Conversion

Cementing Technology: Reduction of Mud Disposal Volumes and Management

of Rig-Site Drilling Wastes. Presented at the SPE/EPA Exploration and

Production Environmental Conference, Society of Petroleum Engineers.

doi:10.2118/25988-MS

Nahm, J.J., Romero, R.N., Javanmardi, K., et al., 1995. Interfacial Sealing Properties of

Slag Mix (Mud-to-Cement Conversion Technology): Laboratory and Field

Evaluation. Society of Petroleum Engineers. doi:10.2118/29407-MS

Nahm, J.J., Romero, R.N., Wyant, R.E., et al., 1994. Universal Fluid: A Drilling Fluid To

Reduce Lost Circulation and Improve Cementing. Society of Petroleum

Engineers. doi:10.2118/27448-MS

Nair, S.D., Wu, Q., Cowan, M., et al., 2015. Cement Displacement and Pressure Control

Using Magneto-Rheological Fluids. Society of Petroleum Engineers.

doi:10.2118/173124-MS

Najafi Kani, E., Allahverdi, A., 2009. Effects of Curing Time and Temperature on

Strength Development of Inorganic Polymeric Binder based on Natural Pozzolan.

J. Mater. Sci. 44, 3088–3097. doi:10.1007/s10853-009-3411-1

Nasvi, M.C.M., Ranjith, P.G., Sanjayan, J., et al., 2014a. Effect of Temperature on

Permeability of Geopolymer: A Primary Well Sealant for Carbon Capture and

Storage Wells. Fuel 117, 354–363. doi:10.1016/j.fuel.2013.09.007

Nasvi, M.C.M., Ranjith, P.G., Sanjayan, J., 2014b. Effect of Different Mix Compositions

on Apparent Carbon Dioxide (CO2) Permeability of Geopolymer: Suitability as

Well cement for CO2 Sequestration Wells. Appl. Energy 114, 939–948.

doi:10.1016/j.apenergy.2013.05.050

Nasvi, M.C.M., Ranjith, P.G., Sanjayan, J., 2013. The Permeability of Geopolymer at

Down-hole Stress Conditions: Application for Carbon Dioxide Sequestration

Wells. Appl. Energy, Special Issue on Advances in sustainable biofuel production

and use - XIX International Symposium on Alcohol Fuels - ISAF 102, 1391–

1398. doi:10.1016/j.apenergy.2012.09.004

Nasvi, M.C.M., Ranjith, P.G., Sanjayan, J., 2012a. Comparison of Mechanical Behaviors

of Geopolymer And Class G Cement As Well Cement At Different Curing

Temperatures For Geological Sequestration of Carbon Dioxide. Presented at the

46th U.S. Rock Mechanics/Geomechanics Symposium, American Rock

Mechanics Association.

Nasvi, M.C.M., Ranjith, P.G., Sanjayan, J., 2012b. Mechanical Properties of Geopolymer

Cement in Brine: Its Suitability as Well Cement for Geological Sequestration of

Carbon Dioxide. Presented at the ISRM Regional Symposium - 7th Asian Rock

Mechanics Symposium, International Society for Rock Mechanics.

Nelson, E.B., Guillot, D., 2006. Well cementing, Second. ed. Schlumberger, Texas, USA.

Page 168: Copyright by Xiangyu Liu 2017

148

Nematollahi, B., Sanjayan, J., 2014. Effect of Different Superplasticizers and Activator

Combinations on Workability and Strength of Fly Ash Based Geopolymer. Mater.

Des. 57, 667–672. doi:10.1016/j.matdes.2014.01.064

Nicholson, C., Murray, B., Fletcher, R., et al., 2005. Novel Geopolymer Materials

Containing Borate and Phosphate Structural Units. Presented at the World

Congress Geopolymer 2005, Saint Quentin, France, 01.07.2005.

Oh, J.E., Monteiro, P.J.M., Jun, S.S., et al., 2010. The Evolution of Strength and

Crystalline Phases for Alkali-Activated Ground Blast Furnace Slag and Fly Ash-

Based Geopolymers. Cem. Concr. Res. 40, 189–196.

doi:10.1016/j.cemconres.2009.10.010

Palacios, M., Houst, Y.F., Bowen, P., et al., 2009. Adsorption of Superplasticizer

Admixtures on Alkali-Activated Slag Pastes. Cem. Concr. Res. 39, 670–677.

doi:10.1016/j.cemconres.2009.05.005

Palacios, M., Puertas, F., 2005. Effect of Superplasticizer and Shrinkage-Reducing

Admixtures on Alkali-Activated Slag Pastes and Mortars. Cem. Concr. Res. 35,

1358–1367. doi:10.1016/j.cemconres.2004.10.014

Palomo, A., Grutzeck, M.W., Blanco, M.T., 1999. Alkali-activated Fly Ashes: A Cement

for the Future. Cement and Concrete Research 29, 1323–1329.

Papakonstantinou, C.G., Balaguru, P., Lyon, R.E., 2001. Comparative Study of High

Temperature Composites. Compos. Part B Eng. 32, 637–649. doi:10.1016/S1359-

8368(01)00042-7

Patel, A.D., Wilson, J.M., Loughridge, B.W., 1999. Impact of Synthetic-Based Drilling

Fluids on Oilwell Cementing Operations. Presented at the SPE INternational

Symposium on Oilfield Chemistry, Society of Petroleum Engineers, Houston,

Texas. doi:10.2118/50726-MS

Pershikova, E., Porcherie, O., Kefi, S., et al., 2012. Pumpable Geopolymers Comprising

A Setting Accelerator. US 20120267106.

Pershikova, E., Porcherie, O., Kefi, S., et al., 2011. Pumpable Geopolymers Comprising a

Fluid-Loss Agent. EP 2338948 A1.

Pessier, R.C., Javanmardi, Nahm, J.J., et al., 1994. Evaluating Slag Solidified Mud

Drillability and Bit Performance. World Oil 215:10.

Peters, E.J., 2012. Advanced Petrophysics: Dispersion, Interfacial Phenomena. Greenleaf

Book Group.

Pollard, S.J.T., Montgomery, D.M., Sollars, C.J., et al., 1991. Organic Compounds in the

Cement-Based Stabilisation/ Solidification of Hazardous Mixed Wastes -

Mechanistic and Process Considerations. J. Hazard. Mater. 28, 313–327.

doi:10.1016/0304-3894(91)87082-D

Porcherie, O., Pershikova, E., 2010. Pumpable Geopolymer Formulation for Oilfield

Application. US 20110073311 A1.

Porcherie, O., Pershikova, E., Rimmele, G., et al., 2011. Pumpable Geopolymers

Comprising A Mixing Aid and Dispersing Agent. EP 2338949 A1.

Provis, J.L., Deventer, J.S.J. van, 2009. Geopolymers: Structures, Processing, Properties

and Industrial Applications. Elsevier.

Page 169: Copyright by Xiangyu Liu 2017

149

Provis, J.L., van Deventer, J.S.J. (Eds.), 2014. Alkali Activated Materials, RILEM State-

of-the-Art Reports. Springer Netherlands, Dordrecht.

Puertas, F., Palomo, A., Fernández-Jiménez, A., et al., 2003. Effect of Superplasticisers

on the Behaviour and Properties of Alkaline Cements. Adv. Cem. Res. 15, 23–28.

doi:10.1680/adcr.2003.15.1.23

Purcell, W.R., 1949. Capillary Pressures - Their Measurement Using Mercury and the

Calculation of Permeability Therefrom. J. Pet. Technol. 1, 39–48.

doi:10.2118/949039-G

Purdon, A., 1940. The Action of Alkalis on Blast-furnace Slag. Journal of the Society of

Chemical Industry - Transactions and Communications 59, 191–202.

Rahier, H., van Mele, B., Wastiels, J., 1996. Low-Temperature Synthesized

Aluminosilicate Glasses. J. Mater. Sci. 31, 80–85. doi:10.1007/BF00355129

Reddy, B.R., Liang, F., Fitzgerald, R., 2010. Self-Healing Cements That Heal Without

Dependence on Fluid Contact: A Laboratory Study. SPE Drill. Complet. 25, 309–

313. doi:10.2118/121555-PA

Richardson, I.G., Brough, A.R., Groves, G.W., et al., 1994. The Characterization of

Hardened Alkali-Activated Blast-Furnace Slag Pastes and the Nature of the

Calcium Silicate Hydrate (c-S-H) Phase. Cem. Concr. Res. U. S. 24:5.

doi:10.1016/0008-8846(94)90002-7

Rowles, M., O’Connor, B., 2003. Chemical Optimization of the Compressive Strength of

Aluminosilicate Geopolymers Synthesised by Sodium Silicate Activation of

Metakaolinite. J. Mater. Chem. 13, 1161–1165. doi:10.1039/B212629J

Ruiz-Santaquiteria, C., Skibsted, J., Fernández-Jiménez, A., et al., 2012. Alkaline

Solution/Binder Ratio as a Determining Factor in the Alkaline Activation of

Aluminosilicates. Cem. Concr. Res. 42, 1242–1251.

doi:10.1016/j.cemconres.2012.05.019

Sakulich, A.R., Anderson, E., Schauer, C., et al., 2009. Mechanical and Microstructural

Characterization of an Alkali-Activated Slag / Limestone Fine Aggregate

Concrete. Constr. Build. Mater. 23, 2951–2957.

doi:10.1016/j.conbuildmat.2009.02.022

Salehi, S., Khattak, M.J., Ali, N., et al., 2016a. Laboratory Investigation of High

Performance Geopolymer Based Slurries. Presented at the AADE Fluids

Technical Conference and Exhibition, Houston, Texas.

Salehi, S., Khattak, M.J., Ali, N., et al., 2016b. Development of Geopolymer-based

Cement Slurries with Enhanced Thickening Time, Compressive and Shear Bond

Strength and Durability. Presented at the IADC/SPE Drilling Conference and

Exhibition, Society of Petroleum Engineers. doi:10.2118/178793-MS

Shah, S.N., Cho, H., 2001. Development of an Environmentally Friendly and Economical

Material for Plugging Abandoned Wells. Presented at the 8th International

Petroleum Environmental Conference, Houston, Texas.

Shah, S.N., Jeong, Y., 2003. The Flow Characteristics of Fly Ash Slurry for Plugging

Abandoned Wells Using Coiled Tubing. Presented at the International Ash

Utilization Symposium, University of Kentucky.

Page 170: Copyright by Xiangyu Liu 2017

150

Shi, C., Day, R.L., 1995. Acceleration of the Reactivity of Fly Ash by Chemical

Activation. Cem. Concr. Res. 25, 15–21. doi:10.1016/0008-8846(94)00107-A

Singh, B., Ishwarya, G., Gupta, M., et al., 2015. Geopolymer Concrete: A Review of

Some Recent Developments. Constr. Build. Mater. 85, 78–90.

doi:10.1016/j.conbuildmat.2015.03.036

Somna, K., Jaturapitakkul, C., Kajitvichyanukul, P., et al., 2011. NaOH-Activated

Ground Fly Ash Geopolymer Cured at Ambient Temperature. Fuel 90, 2118–

2124. doi:10.1016/j.fuel.2011.01.018

Song, M., Wang, W., Ma, K., 2000. Slag MTC Techniques Slove Cementing Problems in

Complex Wells. Presented at the International Oil and Gas Conference and

Exhibition in China, Society of Petroleum Engineers. doi:10.2118/64758-MS

Stiles, D., Hollies, D., 2002. Implementation of Advanced Cementing Techniques to

Improve Long Term Zonal Isolation in Steam Assisted Gravity Drainage Wells.

Society of Petroleum Engineers. doi:10.2118/78950-MS

Stowe, C., Halliday, W., Xiang, T., et al., 2001. Laboratory Pore Pressure Transmission

Testing of Shale. Presented at the National Drilling Technical Conference,

Houston, Texas.

Sugumaran, M., 2015. Study on Effect of Low Calcium Fly Ash on Geopolymer Cement

for Oil Well Cementing. Presented at the SPE/IATMI Asia Pacific Oil & Gas

Conference and Exhibition, Society of Petroleum Engineers. doi:10.2118/176454-

MS

Taoutaou, S., Bermea, V., Andres, J., et al., 2011. Avoiding Sustained Casing Pressure in

Gas wells using Self Healing Cement. Presented at the International Petroleum

Technology Conference, International Petroleum Technology Conference.

doi:10.2523/IPTC-14279-MS

Tare, U., Mody, F., 2000. Stabilizing Boreholes while Drilling Reactive Shales with

Silicate-Base Drilling Fluids. Drill. Contract.

Trussell, S., Spence, R.D., 1994. A Review of Solidification / Stabilization Interferences.

Waste Manag. 14, 507–519. doi:10.1016/0956-053X(94)90134-1

Uehara, M., 2010. New Concrete with Low Environmental Load Using the Geopolymer

Method. Q. Rep. RTRI 51, 1–7. doi:10.2219/rtriqr.51.1

van Deventer, J.S.J., Provis, J.L., Duxson, P., 2012. Technical and Commercial Progress

in the Ddoption of Geopolymer Cement. Miner. Eng., Sustainability through

Resource Conservation and Recycling 29, 89–104.

doi:10.1016/j.mineng.2011.09.009

van Oort, E., 1994. A Novel Technique for the Investigation of Drilling Fluid Induced

Borehole Instability in Shales. Presented at the Rock Mechanics in Petroleum

Engineering, Society of Petroleum Engineers. doi:10.2118/28064-MS

van Oort, E., Hale, A.H., Mody, F.K., et al., 1996. Transport in Shales and the Design of

Improved Water-Based Shale Drilling Fluids. SPE Drill. Complet. 11, 137–146.

doi:10.2118/28309-PA

Page 171: Copyright by Xiangyu Liu 2017

151

van Oort, E., Hoxha, B.B., Hale, A., et al., 2016. How to Test Fluids for Shale

Compatibility. Presented at the AADE Fluids Technical Conference and

Exhibition, Houston, Texas.

Wang, S., Scrivener, K.L., 1995. Hydration Products of Alkali Activated Slag Cement.

Cem. Concr. Res. 25, 561–571. doi:10.1016/0008-8846(95)00045-E

Ward, C.R., French, D., 2006. Determination of Glass Content and Estimation of Glass

Composition in Fly Ash using Quantitative X-ray Diffractometry. Fuel, Special

Issue: The 2005 World of Coal Ash Conference 85, 2268–2277.

doi:10.1016/j.fuel.2005.12.026

Wiles, C.C., 1987. A Review of Solidification / Stabilization Technology. J. Hazard.

Mater., Approaches for dealing with problems at hazardous waste sites 14, 5–21.

doi:10.1016/0304-3894(87)87002-4

Williams, P.J., Biernacki, J.J., Rawn, C.J., et al., 2005. Microanalytical and

Computational Analysis of Class F Fly Ash. Aci Mater. J. 102, 330–337.

Williams, R.P., van Riessen, A., 2010. Determination of the Reactive Component of Fly

Ashes for Geopolymer Production using XRF and XRD. Fuel 89, 3683–3692.

doi:10.1016/j.fuel.2010.07.031

Wilson, W.N., Miles, L.H., Boyd, B.H., et al., 1989. Cementing Oil and Gas Wells using

Converted Drilling Fluid. US 4883125 A.

Wu, D., Sun, F., Peiyan, _, et al., 1998. Low Density Slag / Mud Mix Slurries Applied in

Changqing Oilfield for Sealing Long Annuli. Presented at the SPE International

Oil and Gas Conference and Exhibition in China, Society of Petroleum Engineers.

doi:10.2118/50891-MS