a geopolymer cementing system for oil wells subject to

36
Accepted Manuscript A geopolymer cementing system for oil wells subject to steam injection Maria D.M. Paiva, Emílio C.C.M. Silva, Dulce M.A. Melo, Antônio E. Martinelli, José F. Schneider PII: S0920-4105(18)30510-2 DOI: 10.1016/j.petrol.2018.06.022 Reference: PETROL 5029 To appear in: Journal of Petroleum Science and Engineering Received Date: 17 January 2018 Revised Date: 3 June 2018 Accepted Date: 7 June 2018 Please cite this article as: Paiva, M.D.M., Silva, Emí.C.C.M., Melo, D.M.A., Martinelli, Antô.E., Schneider, José.F., A geopolymer cementing system for oil wells subject to steam injection, Journal of Petroleum Science and Engineering (2018), doi: 10.1016/j.petrol.2018.06.022. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Upload: others

Post on 03-Oct-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A geopolymer cementing system for oil wells subject to

Accepted Manuscript

A geopolymer cementing system for oil wells subject to steam injection

Maria D.M. Paiva, Emílio C.C.M. Silva, Dulce M.A. Melo, Antônio E. Martinelli, José F.Schneider

PII: S0920-4105(18)30510-2

DOI: 10.1016/j.petrol.2018.06.022

Reference: PETROL 5029

To appear in: Journal of Petroleum Science and Engineering

Received Date: 17 January 2018

Revised Date: 3 June 2018

Accepted Date: 7 June 2018

Please cite this article as: Paiva, M.D.M., Silva, Emí.C.C.M., Melo, D.M.A., Martinelli, Antô.E.,Schneider, José.F., A geopolymer cementing system for oil wells subject to steam injection, Journal ofPetroleum Science and Engineering (2018), doi: 10.1016/j.petrol.2018.06.022.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service toour customers we are providing this early version of the manuscript. The manuscript will undergocopyediting, typesetting, and review of the resulting proof before it is published in its final form. Pleasenote that during the production process errors may be discovered which could affect the content, and alllegal disclaimers that apply to the journal pertain.

Page 2: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

1

A GEOPOLYMER CEMENTING SYSTEM FOR OIL WELLS SUBJECT TO STEAM 1

INJECTION 2

Maria D.M. Paiva1, a, Emílio C.C.M. Silvab, Dulce M.A. Meloc, Antônio E. Martinellid and José F 3

Schneidere 4

Corresponding author 5 1 UFRJ, POLI/COPPE, NUMATS 6 Av. Athos da Silveira Ramos, 149 (Bloco I-110) 7 Cidade Universitária 8 Rio de Janeiro, RJ - 21941-972 - Brazil 9 [email protected] 10 11 a Universidade Federal do Rio Grande do Norte (UFRN), Programa de Pós-Graduação em Ciência e 12

Engenharia de Materiais. 13 Av Salgado Filho 3000 14 Lagoa Nova 15 Natal, RN - Brazil 59072970 16 17 b Petróleo Brasileiro S.A, CENPES 18

Av. Horácio Macedo, 950, Escritório 8, CENPES/Ampliação 19 Cidade Universitária 20 Rio de Janeiro, RJ - 21941915 – Brazil 21 [email protected] 22

23 c UFRN, Centro de Ciências Exatas, Instituto de Química. 24

Av Salgado Filho 3000 25 Lagoa Nova 26 Natal, RN - 59072970 – Brazil 27 [email protected] 28 29 d UFRN, Centro de Tecnologia, Departamento de Engenharia de Materiais. 30

Av Salgado Filho 3000 31 Lagoa Nova 32 Natal, RN - 59072970 - Brazil 33 [email protected] 34 35 e USP, Instituto de Física de São Carlos, Departamento de Física e Ciência Interdisciplinar. 36

Av.Trabalhador São-carlense, 400 37 Centro 38 Caixa-postal: 369, São Carlos, SP - 13560-970 - Brazil 39 [email protected] 40 41

Page 3: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

2

42

ABSTRACT 43

Oil wells subjected to cyclic steam injection often present cement sheath failures due to thermal 44

loads. Currently, different Portland based systems are employed to mitigate this risk. We developed 45

and tested geopolymer systems that meet oil well cementing requirements and achieve high 46

mechanical performance. A metakaolin-potassium-based formulation, containing microsilica, 47

mineral fiber and retarder, was adjusted to the desired rheology and thickening time. Triaxial 48

compression and indirect tensile tests were performed and the properties were used in finite element 49

simulations of the life of a model onshore well. Compared to the Portland based system currently in 50

use, the geopolymer has superior uniaxial compressive strength and lower stiffness. Tensile strength 51

increased significantly with the addition of mineral fibers and became higher than that of the 52

Portland system. Numerical simulations show that the geopolymer can withstand steam injection 53

temperatures comparable to a conventional system, remaining in the elastic region. These results 54

show that geopolymers are a viable alternative well cementing system for this application, with 55

significant improvement in mechanical performance. 56

Keywords: geopolymer; alkali-activated binder; well cementing integrity; mechanical properties; 57

triaxial testing; computer simulation. 58

1 INTRODUCTION 59

The cement sheath in an oil well must provide structural support to the casing and surface 60

equipment, compose the well barrier against fluid migration, protect the casing from corrosion and 61

isolate production zones (Nelson, 1990). Cement is placed in the annulus between the casing and 62

the borehole wall by pumping slurry through the casing and displacing the previous fluid in the 63

annulus. After setting, the cement will bond to the external wall of the casing and to the borehole 64

wall, which may be a previous casing string. 65

Page 4: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

3

There are two sets of requirements for a cementing system. In the fluid state, the density and 66

rheology of the slurry must be adjusted to keep downhole pressures, due to hydrostatics and to 67

friction losses, within an operational window. Rheological properties must be maintained for the 68

duration of the cementing operation, but setting and strength evolution must be as fast as possible 69

after placement. In the solid state, the requirements include mechanical strength, low permeability 70

and resistance to chemical degradation (Nelson, 1990). 71

Onshore oil fields in northeastern Brazil contain shallow sandstone formations bearing heavy oil, 72

intercalated with potable water aquifers, both of which are very important to the local economy. 73

Current exploration techniques, described by Rosa et al. (2006), employ continuous and huff-and-74

puff steam injection to improve productivity and recovery factors of these heavy oil formations. 75

These techniques expose the cement sheath to large temperature variations during its operating life, 76

which expose the cement to extreme stress states (Bour, 2005; Dean and Torres, 2002). 77

The main mechanism is the thermal expansion of the casing, compressing the cement sheath against 78

the formation. Possible failure modes include radial cracking due to tangential stresses, compressive 79

failure and debonding at the interfaces between the cement sheath and casing and/or rock. 80

Dean and Torres (2002) reported that cement sheath failures in Brazilian onshore fields are mainly 81

due to radial cracking. In some cases, there is an increase in water fraction due to loss of zonal 82

isolation and steam may be observed at the surface. Beside the high costs of remedial jobs, there is a 83

risk of contaminating aquifers used by the local population. Bour (2005) analyzed numerically a 84

model well in Central California in a very compliant formation. Tensile failure was the primary 85

failure mechanism for all well geometries, since the formation did not confine the cement. 86

Bour (2005), as well as Goodwin and Crook (1992) detail requirements for the design of cement 87

slurries for steam injection wells, recommending that slurries must be compliant, resilient, ductile 88

and strong in tension. Finite element analyses demonstrate that conventional Portland-based slurries 89

with silica addition are not suitable for these conditions, having high compressive strength but low 90

tensile strength and significant brittleness. 91

Page 5: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

4

Alternative ordinary Portland cement (OPC)-based slurry formulations proposed in the literature 92

and sold commercially usually include additions such as mineral fibers and/or elastomers. In 93

northeastern Brazil, latex-based slurries are routinely used to improve the mechanical properties, 94

with mixed results (Dean and Torres, 2002). 95

Geopolymers (Davidovits, 1991, 1994a) or alkali-activated binders (Pacheco-Torgal et al., 2008) 96

are inorganic polymers composed of silicate and aluminate tetrahedra linked by shared oxygen 97

atoms and stabilized by cations such as Na+ or K+. The structure is analogous to zeolites; however, 98

their structure is amorphous or semicrystalline. Geopolymers are formed by activating source 99

materials with an alkaline solution, usually a combination of hydroxides and silicates. Fly ash and 100

metakaolin (calcinated kaolin) are common sources of silica and alumina, but several other 101

materials can be used (Olawale, 2013). 102

Compared to Portland-based formulations, the use of geopolymers (Davidovits, 1991) for well 103

cementing is still in early stages. Based on work by Marinho (2004), Nóbrega (2006) and Pinto 104

(2007), Paiva (2008) developed formulations that are equivalent to cement slurries when tested with 105

API RP 10B-2 (2013) methods. However, effective retarders of the geopolymerization reaction 106

were necessary to achieve the thickening time required in cementing operations. Cementing service 107

company Schlumberger published two patents (Barlet-Gouedart et al., 2010a, 2010b) on the use of 108

geopolymers in oil well cementing, followed by Halliburton (Chatterji, 2016), which proposed 109

alternative slag- and pumice-based compositions. These and more recent patents, such as (Porcherie 110

et al., 2016), focus on effective retarders to enable application in diverse well environments. In one 111

of the first field applications, cementing service company Sanjel employed hybrid slurries with up 112

to 60% geopolymer in more than thirty wells in Alberta, Canada, with cost savings and good 113

mechanical properties (Mahmoudkhani et al., 2008). 114

Khalifeh et al. (2014) evaluated fly ash based geopolymers for plug and abandonment using API RP 115

10-B (2013) tests, finding that the materials are suitable for use, while pointing that gelation and 116

rapid setting are major concerns. Salehi et al. (2016) performed a sensitivity analysis, observing that 117

Page 6: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

5

optimization of the mix design is critical for well cementing geopolymers. Salehi et al. (2017, 2018) 118

expanded this work with a comprehensive set of API tests, obtaining a geopolymer mixture with 119

potential to replace OPC well cementing systems. The authors tried different classes of retarders 120

and plasticizers, reporting difficulties in obtaining reliable behavior, as well as shortcomings of the 121

API thickening time procedure when applied to geopolymers. 122

Two major challenges for ordinary Portland cement in current well applications are salt and CO2-123

rich environments, such as the Brazilian pre-salt (Alves et al., 2009). Nasvi et al. (2012a, 2012b, 124

2014c) evaluated pure geopolymers in salt solutions and their applicability as CO2 sequestration 125

well cements. Besides economic and environmental advantages, geopolymers exhibit higher 126

durability (Nasvi et al., 2014b), resistance to acids and chemical attacks (Nasvi et al. 2014a), as well 127

as the higher strength compared to OPC pastes (Nasvi et al., 2014c), better bond to casing (Salehi, 128

2017), less shrinkage (Salehi, 2018), no carbonation and no alkali-aggregate reaction. Nasvi et al. 129

(2012a) studied the mechanical behavior at downhole conditions, identifying changes in 130

deformation capacity and failure mode with temperature. Brine reduces the strength of geopolymers, 131

but this reduction is less pronounced than the analogous effect on OPC-based pastes (Nasvi et al., 132

2012b; Giasuddin et al., 2013). Combined with a very low permeability at downhole conditions 133

(Nasvi at al., 2014b), compared to API requirements and OPC pastes, these properties make 134

geopolymers an advantageous alternative for zonal isolation in aggressive well environments. 135

An undesired property of geopolymers is their brittleness and low tensile strength, requiring 136

reinforcement (Silva, 2000; Thaumaturgo and Silva, 1999). A comprehensive review by Sá Ribeiro 137

et al. (2017) discusses how additions, such as particles and fibers, can enhance compressive, tensile 138

and flexural strength, increase fracture toughness and ductility of geopolymer composites. 139

Challenging applications, such as cementing oil wells subject to steam injection, explore the limits 140

of what is economically feasible to obtain with Portland-based cementing systems. Alternative 141

cementing systems may outperform ordinary materials and simultaneously reduce costs and the 142

carbon footprint of well construction. Application of these systems require demonstration that all 143

Page 7: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

6

operational requirements are satisfied and that mechanical and zonal isolation objectives will be 144

achieved. The new system developed in this work was subject to a full set of industry-specific tests 145

and was proven to maintain its characteristics over the life of the well, by numerical simulation. 146

2 EXPERIMENTAL 147

2.1 Materials and mix design 148

A Portland-based formulation (P2) used in the field was prepared using a local cement provided by 149

CIMESA, similar to API Class G. This formulation is a 1790 kg/m3 [15.0 ppg] slurry with w/c ratio 150

0.63, 40% microsilica by weight of cement (bwoc), supplied by Elkem and 50 mL per kg of cement 151

[0.6 gps] latex. All additives (latex, latex stabilizer, defoamer, dispersant and fluid loss control) 152

were supplied by Halliburton Brazil. Styrene-butadiene latex is supplied as a water-internal, oil-153

external emulsion of latex polymer, which is stabilized by a separate emulsifying additive 154

(ethoxylated nonylphenol). The defoamer is based on polypropylene glycol. The combined 155

dispersant and fluid loss control additive is a sulfonated organic salt condensate. The second fluid-156

loss additive is a copolymer of 2-acrylamido-2-methylpropane sulfonic acid (AMPS) and N,N-157

dimethylacrylamide (NNDMA). 158

Geopolymer formulations P3, P4 and P5 are based on metakaolin from Metacaulim do Brasil, using 159

potassium hydroxide p.a. from Vetec and potassium silicate solution from Diatom do Brasil as 160

alkaline activators. Two batches of metakaolin were used, labeled “Metakaolin RN” and 161

“Metakaolin RJ”, found to be very similar in terms of physical and chemical characterization. Most 162

of the work was done with the first batch, except for formulations P3.B, P4.B and P5.B, described 163

below, which used the second batch. A polypropylene glycol-based defoamer provided by 164

Halliburton Brazil was used in all cases. P5 uses a copolymer of AMPS and acrylic acid as a 165

retarder additive, supplied by Halliburton Brazil. 166

To the base geopolymer formulation P3, P4 adds 10.6% microsilica (by weight of metakaolin, 167

b.w.o.m.) and 43 mL per kg of metakaolin [0.5 gps] of AMPS-based retarder. P5 is also based on 168

Page 8: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

7

P3, adding 5.3% b.w.o.m. microsilica and 2.1 vol. % natural wollastonite, provided by Energiac. 169

Molar ratios were kept fixed at SiO2/Al2O3 = 3.8, K2O/Al2O3 = 1.0 and H2O/K2O = 12. Resulting 170

densities varied between 1690 kg/m3 [14.1 ppg] and 1730 kg/m3 [14.4 ppg]. The design parameters 171

of the geopolymer formulations are listed in Table 1 below. 172

Starting with a fixed weight of pure metakaolin, microsilica (if applicable) is added as a weight 173

fraction of the metakaolin. Next, the amount of potassium silicate solution is calculated to obtain 174

the desired SiO2/Al 2O3 ratio. The K2O/Al2O ratio is reached by adding potassium hydroxide. 175

Finally, water is added until the H2O/K2O ratio is obtained. To this base geopolymer, the retarder 176

and defoamer are added as fixed volumes per unit mass of metakaolin. Finally, the wollastonite 177

mass is calculated to obtain the desired volume fraction of fiber addition. 178

Table 1 – Mix design parameters, additions and additives. 179

Parameter Units P3 P4 P5 Control material*

Microsilica % b.w.o.m. 0 10.6 5.3

SiO2/Al 2O3 - 3.8 3.8 3.8 K2SiO3(aq) K2O/Al2O3 - 1.28 1.0 1.0 KOH(s)

H2O/K2O - 12 12 12 Water

Retarder mL/kg [gps] 0 43 [0.5] 43 [0.5]

Defoamer mL/kg [gps] 1.7 [0.02] 1.7 [0.02] 1.7 [0.02]

Wollastonite vol. % 0 0 2.1

* Material used to obtain the given molar ratio.

The target molar ratios of P3 were based on studies performed by Marinho (2004), Nóbrega (2006) 180

and Pinto (2007). The SiO2/Al2O3 ratio was chosen to maximize the compressive strength (Duxson 181

et al., 2005; Steveson and Sagoe-Crentsil, 2005). The target K2O/Al2O3 ratio was 1.0 to prevent 182

incomplete polymerization of the geopolymer network (Davidovits, 1991). P3 required a higher 183

value (1.28) to keep the target H2O/K2O with the available materials. The H2O/K2O controls 184

directly the alkalinity of the activator and the workability of the slurry. We fixed this parameter at 185

12, which to obtain workable slurries for oil well applications. Decreasing this parameter leads to 186

very short setting times (Paiva, 2008) due to the increase in alkalinity. 187

Microsilica was introduced as an alternative source of silica, to reduce the potassium silicate 188

requirement, therefore decreasing the cost and improving the handling in the field. Fixing the 189

Page 9: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

8

microsilica/metakaolin weight ratio, we kept constant the total amount of solids, since Paiva (2008) 190

determined that more than 38 vol. % of solids resulted in viscosities above the target of 0.150 Pa.s 191

[150 cP]. The volume fraction of mineral fiber was fixed to obtain a constant reinforcement effect. 192

Due to its effect on viscosity, we reduced the amount of microsilica when adding wollastonite. 193

Additives were specified in amounts compatible to those used in OPC-based slurries, using field 194

units (1 gps is 1 gallon [3.7854 L] per 94 lbs [42.64] sack of cement) and replacing metakaolin for 195

cement. 196

Two different curing regimes were employed for geopolymer formulations P3, P4 and P5. 197

Geopolymer samples cured with pressure are labeled with the suffix B (P3.B, P4.B and P5.B) to 198

distinguish them from those cured at atmospheric pressure (P3, P4 and P5). Those pressure-cured 199

samples were prepared using the “Metakaolin RJ” batch, using the design method above to 200

reproduce the same molar ratios and obtain the same geopolymer matrices. 201

2.2 Characterization 202

Physical-chemical analysis of metakaolin, microsilica and wollastonite included X-ray fluorescence 203

(XRF), X-ray diffraction (XRD, thermogravimetry (TG), differential thermal analysis (DTA, SEM, 204

EDX and high-resolution 29Si MAS NMR. The potassium silicate composition was determined by 205

XRF. Details of sample preparation and complete results can be found in Paiva (2008). 206

Geopolymers cured for mechanical tests were also sampled for physical-chemical characterization 207

by XRD, FTIR, TG, DTA, SEM, 29Si MAS NMR, 1H NMR, differential scanning calorimetry 208

(DSC) and thermomechanical analysis (TMA). 209

2.3 Standard well cement tests 210

Conventional and geopolymer slurries were prepared and tested according to API RP 10B-2 (2013) 211

for rheology, free water, fluid loss, and thickening time determination. . The flow curve was 212

measured using a Chandler Model 3506 Viscometer and the results were interpreted using the 213

Bingham model. Fluid loss was obtained in a Fann Fluid Loss Tester. Thickening time was 214

determined using a Chandler Model 8340 HPHT Consistometer. The bottom hole static temperature 215

Page 10: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

9

(BHST) was calculated for a model well 530 m [1740 ft] deep, using a geothermal gradient of 216

4.2°C/100 m [2.3°F/100 ft], obtaining 49°C [120°F]. The calculated circulating temperature 217

(BHCT) is 31°C [86°F]. The temperature and pressure schedule was calculated according to API 218

RP 10B-2 (2013). 219

2.4 Mechanical tests 220

Slurries were placed in cylindrical molds of 5.0 cm [2.0 in] diameter and 10 cm [3.9 in] height. 221

Geopolymers were insulated with PVC film to prevent drying shrinkage (Tavares et al., 2004). All 222

samples were cured in a wet chamber for 14 days at 49ºC [120ºF] and atmospheric pressure, except 223

for those labeled with the suffix B (P3.B, P4.B and P5.B), which were pre-cured at 10 MPa 224

[1450 psi] and 49ºC [120ºF] for 48 h before being moved to the wet chamber for the remaining 12 225

days. 226

For each formulation, three samples were tested in uniaxial compression according to ASTM C39 227

(2016) at 0.1 mm/min [3.9 mils/min]. Compressive tests were instrumented with axial LVDTs. 228

Poisson ratios were determined using axial and lateral strain gauges. Nine samples were tested in 229

triaxial compression (ASTM D2664-04, 2004) at 2, 6 and 10 MPa [290, 870 and 1450 psi] 230

confining pressures, using three samples per pressure. Indirect tensile strength tests (ISRM, 2014) 231

were performed at 0.8 mm/min [31.5 mils/min] in four samples per formulation. 232

3 NUMERICAL 233

Simulations were performed in five selected transverse sections of a 500 m [1640 ft] vertical well, 234

as shown in Fig. 1. The well has a 0.1778 m [7 in] casing of nominal weight 340 N/m [23 ppf], 235

placed on a 0.2223 m [8 3/4 in] circular hole with a 70% stand-off and cemented to the surface with 236

the given cement paste, with industrial water inside the casing. Fig. 2 shows a well section and a 237

detail of the 2D plane strain finite element mesh employed, with 9353 nodes and 3040 elements. 238

Tripling the number of elements in the cement sheath, casing and 1.5 m of surrounding rock 239

changed the failure temperature by less than 0.1°C, so further mesh refinement was not necessary. 240

Page 11: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

10

All elements are coupled temperature-displacement 8-node quadrilaterals, with biquadratic 241

displacement, bilinear temperature and reduced integration. The formation is a circular section with 242

50 m [164 ft] radius, mechanically pinned and kept at constant temperature at the far end. 243

244

Fig. 1 – Well geometry, formations and depths of the simulated sections. At 200 m, the same section 245 is simulated with both formations. 246

247

Fig. 2 – Detail of the geometry and finite element mesh of a 2D well section (eccentric casing, 248 cement sheath and formation). The problem is symmetric around the vertical axis. There are 240 249 elements in the casing and 400 in the cement sheath. 250

Steel and rock are modeled as linear elastic, while the cement or geopolymer is elastoplastic with a 251

Mohr-Coulomb failure criterion and a Rankine (maximum principal stress) criterion in tension. 252

Table 2 summarizes the material properties adopted in the model. Pressures are equilibrated 253

radially with the gradient of water inside the casing and the gradient of slurry at the casing-cement 254

Page 12: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

11

and cement-rock interfaces. The initial stress state of the paste is the hydrostatic pressure after 255

placement. The initial temperature is uniform, calculated from the geothermal gradient. 256

Table 2 – Material properties used in thermomechanical simulations. Mechanical properties 257 measured in this work are listed in Table 6. 258

Property Units Steel Carbonate Sandstone P2 P3, P4, P5

Young modulus GPa [ksi]

200 [2900]

7.0 [1015]

3.43 [497]

Poisson ratio 0.30 0.23 0.28

Specific heat kJ/m3 ºC 4000 2000 2788 2200 1750

Thermal conductivity W/m ºC 15 2.4 1.59 0.7 0.3

Thermal expansion 10-5/ºC [10-5/°F]

1.16 [0.644]

0.80 [0.444]

1.0 [0.555]

1.38 [0.767]

0.40 [0.222]

Density kg/m3 [ppg]

7854 [65.5]

2560 [21.3]

2220 [18.5]

1799 [15.0]

1700 [14.2]

259

Steam injection was idealized by heating to 260°C [500°F] at 110°C/h [198°F/h] and ramping the 260

internal pressure at the inside of the casing to the pressure of saturated steam at this temperature 261

(4.59 MPa [665 psi]) in 10 minutes. 262

During heating, the cement or geopolymer sheath is monitored for plastic strain, caused by violation 263

of one or more failure criteria. If the material fails, the current temperature and failure mode is 264

recorded. 265

4 RESULTS AND DISCUSSION 266

4.1 Characterization 267

Chemical composition of the materials was determined by FRX and crystallinity was assessed by 268

XRD. Metakaolin is 95 wt. % aluminosilicate (49.3 wt. % alumina and 45.7 wt. % silica), with an 269

amorphous hump located between 20° and 30° 2θ and quartz (PDF 00=046-1045) and kaolinite 270

(PDF 00-014-0164) as crystalline contaminants (Fig. 3). Microsilica is 95.9 wt. % amorphous 271

silica (Fig. 4). Wollastonite is 97.8 wt. % crystalline calcium silicate hydrate (PDF 00-043-1460) 272

(Fig. 5). The commercial silicate solution contains 37.8 wt. % potassium silicate (26.7 wt. % K2O 273

and 11.1 wt. % SiO2) in aqueous solution. 274

Page 13: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

12

275 Fig. 3 – XRD patterns of the two batches of metakaolin used in the study. 276

277 Fig. 4 – XRD pattern of microsilica 278

279

Quartz

Kaolinite (C)

RJ

RN

Angle (2θ)

Angle (2θ)

Angle (2θ)

Wollastonite

Page 14: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

13

Fig. 5 – XRD pattern of wollastonite. 280

281

Thermal analysis (Fig. 6) shows that the materials are thermally stable up to 600°C. Metakaolin and 282

microsilica lose about 5% of their mass, gradually over the whole interval analyzed, suggesting 283

dehydration of adsorbed water for microsilica and metakaolin up to 200°C, and dehydroxylation of 284

kaolinite from 400°C to 500°C. 285

286

Fig. 6 – Thermogravimetric analysis of raw materials. 287

288

Scanning electron microscopy images (Fig. 7) shows that metakaolin and microsilica are composed 289

of approximately spherical particles consistent with the size distribution measured by laser 290

granulometry (not shown). Wollastonite has an acicular geometry, with fibers about 100 µm long. 291

EDX analysis confirms the homogeneity and overall composition determined by FRX. 292

Metakaolin Microsilica Wollastonite

Wei

ght

Loss

(T

G, %

)

Temperature (oC) D

TG

(%

/sec

)

Page 15: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

14

1000x/ 10 µm

128 x 128 pixels

(a) Metakaolin (Al ���� red and Si ���� blue). 293

294

1000x/ 10 µm

128 x 128 pixels

(b) Microsilica (Si ���� blue and O ���� green). 295

296

1000x/ 10 µm

128 x 128 pixels

(c) Wollastonite (Si ���� blue and Ca ���� yellow) 297

298

Fig. 7 – SEM and EDX of raw materials. 299

Page 16: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

15

300

Fig. 8 displays the 29Si MAS NMR spectra of the original materials. Metakaolin shows Q4(m Al) 301

silicates, with a peak at -91.6 ppm corresponding to kaolinite (Q3 silicates). Microsilica presents a 302

main peak at -110.5 ppm (Q4 silicates), with contributions from Q3, Q2 and Q4(m Al) groups. 303

Wollastonite displays Q2 silicates. 304

305 Fig. 8 – 29Si MAS NMR spectra of raw materials. 306

4.2 Geopolymer characterization 307

X-ray diffraction patterns of the geopolymer formulations show an amorphous hump centered 308

around 29°, shifted from the humps observed in both metakaolin and microsilica patterns. Weak 309

quartz and kaolinite peaks were observed in some samples. P5 displayed small wollastonite peaks, 310

consistent with the pattern of that material and with the small fraction of fibers in the formulation. 311

This is evidence that a similar microstructure was obtained using microsilica as an alternative 312

source of silica. Pre-curing at 10 MPa [1450 psi] for 48 h did not change the patterns, with 313

exception of spurious peaks at around 40° 2θ for P4 and 45° 2θ for P5.B probably related to 314

Wollastonite

Microsilica

Metakaolin

Chemical deviation (ppm)

Page 17: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

16

samples contamination, showing that pressure levels representative of downhole conditions do not 315

change the geopolymer structure. 316

317

318

Angle (2θ)

Angle (2θ)

Page 18: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

17

319

Fig. 9 – XRD diffractograms of geopolymer pastes. 320

321

FTIR spectra (Fig. 10) are also very similar for all three geopolymer formulations, including those 322

cured under pressure, confirming that the same product was generated. The main absorbance peak is 323

located between 1010 cm-1 and 1030 cm-1, shifted from silica peaks observed in the original 324

materials, an evidence of Si–O–Al bond formation. Smaller peaks corresponding to silicate 325

tetrahedra, Si–O–Al and Si–O–Si stretching are also observed. A wide peak around 3350 cm-1 326

suggests that –OH groups remain in the structure. 327

Angle (2θ)

Page 19: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

18

328

329

Wave number (cm-1)

Tra

nsm

itanc

e (%

) T

rans

mita

nce (

%)

Wave number (cm-1)

Page 20: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

19

330

Fig. 10 – FTIR spectra of geopolymer pastes. 331

332

Thermogravimetric analysis (Fig. 11) identified approximately 7% mass loss before 100°C and 333

progressive loss until 900°C, up to about 15% mass loss. The percentage seems to be related to the 334

amount of water present in the geopolymer systems. DTA and DSC show an exothermal event 335

simultaneous with the initial loss before 100°C. Formulations with retarder lost about 2% of their 336

masses around 310°C, which can be attributed to the decomposition of the retarder, also observed in 337

the DTA and DSC curves (Fig. 11). 338

339

Tra

nsm

itanc

e (%

)

Wave number (cm-1)

Wei

ght

Loss

(T

G, %

)

DT

G (

%/s

ec)

Temperature (oC)

Page 21: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

20

340

341

Fig. 11 – TG, DTA, DSC of geopolymer pastes. 342

343

Thermomechanical analysis (Fig. 12) shows geometrical stability up to 100°C and significant 344

shrinkage (up to 6% at 600°C) above this. Formulations containing wollastonite displayed less 345

shrinkage (4.8% at 600°C). 346

Temperature (oC)

Temperature (oC)

Page 22: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

21

347

Fig. 12 – TMA of geopolymer pastes. 348

349

The authors interpret that most of the mass lost below 100°C is due to water evaporation in 350

macropores, which does not cause significant shrinkage, as suggested by Duxson et al. (2005). 351

Above 100°C, loss of water in micropores causes drying shrinkage due to capillary effects. Above 352

250°C, kaolin dehydroxylation causes further shrinkage. 353

Electron microscopy of fracture surfaces show a compact structure, compatible with a well-formed 354

geopolymer, with nanometer scale voids. This suggests that the Si/Al ratio was adequate and that 355

the defoamer prevented the formation of air bubbles. Formulation P5 displayed a larger number of 356

pores, which may be related to the presence of wollastonite fibers. These fibers were apparently 357

incorporated to the matrix. 358

Temperature (oC)

Len

gth

Var

iatio

n (%

)

Page 23: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

22

10000x / 1 µm

30000x / 500 nm (a) P3 359

10000x / 1 µm

30000x / 500 nm (b) P4 360

10000x / 1 µm

30000x / 500 nm (c) P5 361

Fig. 13 – SEM of geopolymer pastes. The lengths (µm and nm) correspond to the size of scale bar. 362

EDX imaging shows that Si, Al and K are uniformly distributed in the matrix. A few aluminum-rich 363

particles were embedded in the matrix, probably unreacted particles. Calcium oxide-rich fibers can 364

Page 24: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

23

be found in formulation P5. Pastes cured at 10 MPa [1450 psi] did not display qualitative changes 365

in the EDX images. 366

1000x / 10 µm

128 x 128 pixels

(a) P3 (Si � blue, Al � red and K � green). 367

1000x / 10 µm

128 x 128 pixels

(b) P4 (Si � blue, Al � red and K � green). 368

1000x / 10 µm

128 x 128 pixels

(c) P5 (Si � blue, Al � red and K � green) 369

Page 25: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

24

Fig. 14 – SEM and EDX of geopolymer pastes. The lengths (µm and nm) correspond to the size of 370

scale bar. 371

372

Hydrogen (1H-NMR) time domain analysis was performed to determine the relaxation time (T1). As 373

shown in Table 3, more than 80% of the relaxation occurs in the 0.1 to 0.6 ms range, and a 374

significant amount occurs in the 1 to 20 ms range. Samples with wollastonite (P5) displayed 375

additional relaxation in the 30 to 50 ms range. Faure and Rodts (2008) interpreted the higher time 376

range (30 to 150 ms) as corresponding to the interaction between pore water and pore surface. A 377

lower range (2 to 3 ms) was attributed to micro-organized water. These results suggest that most of 378

the water is in the form of –OH groups (lowest range) and micropore water (1 to 20 ms range). The 379

fraction in the 30 to 50 ms range, observed in formulation P5, may be related to the larger pores 380

observed by SEM. Larger relaxation times are absent, probably because the samples were dried 381

before testing. 382

Table 3 – Summary of 1H-NMR results. TCi and Ai are the spin-network relaxation times and 383 the corresponding fraction. Pn.B are geopolymers pre-cured with pressure. 384

Paste TC1 (ms) A1 (%) TC2 (ms) A2 (%) TC3 (ms) A3 (%)

P3 0.249 96.8 4.94 3.2

P3.B 0.533 85.5 1.60 14.5

P4 0.140 86.9 5.24 13.1

P4.B 0.243 92.8 13.4 7.2

P5 0.149 82.7 1.46 11.4 44.0 5.9

P5.B 0.168 82.1 1.16 14.1 30.0 3.8 385

Silicon (29Si MAS NMR) spectra are shown in Fig. 15 and quantified in Table 4. The samples are 386

generally similar: there are two main wide 29Si bands, the most intense of which is centered around 387

-92 ppm, corresponding to geopolymer silicates Q4(m Al) and some Q3. The other is centered 388

around -110 ppm and correspond to Q4(1Al) e Q4 silicates. According to Singh et al. (2005), the 389

latter do not strictly belong to the geopolymer, but Davidovits (1994b) observed a similar peak at -390

115 ppm when microsilica was added to the formulation. The fraction of geopolymer silicates 391

Page 26: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

25

increases when pre-curing with pressure. The base geopolymer P3/P3.B presents a weaker Q4(1Al) 392

band, which is also better resolved from the Q4 line. In these, a band from -60 ppm to -70 ppm 393

evidences Q0 or Q1 silicates, indicating low connectivity groups, which may be an unreacted 394

fraction of the potassium silicate solution. P4/P4.B have the largest fraction of geopolymer silicates. 395

396 Fig. 15 – 29Si MAS NMR spectra of geopolymers. Pn.B samples were pre-cured with pressure. 397

Table 4 - Quantitative 29Si MAS NMR interpretation. 398

Paste Q1 (%)

Q3 + Q4(3Al) + Q4(2Al) (%)

Q4(1Al) + Q4 (%)

Ratio

P3 5.6 63.9 30.5 2.1

P3.B - 76.3 23.7 3.2

P4 - 80.8 19.2 4.2

P4.B - 82.2 17.8 4.6

P5 - 67.8 32.2 2.1

P5.B - 78.0 22.0 3.5

399

4.3 Standard well cement tests 400

Results of API RP 10B-2 (2013) tests are summarized in Table 5. Slurry placement simulations 401

show that the rheology of geopolymer slurries is suitable for well cementing in the model scenario, 402

Chemical deviation (ppm)

Page 27: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

26

with an equivalent circulating density (ECD) compatible with the operational window, maintaining 403

usual displacement rates. Excessive filtrate loss (more than 100 mL/30 min) and free water (1 mL) 404

was observed for pure geopolymer (P3), but the addition of microsilica and consequent reduction of 405

the fluid content was sufficient to reduce the filtrate loss and free water of formulations P3 and P4. 406

If further control were required, Barlet-Gouedart et al. (2010a) suggests that Portland cement 407

additives can be effective in controlling filtrate loss and free water of geopolymers. 408

The base geopolymer P3 has a very short thickening time at the BHCT (31°C, 88°F), compared to 409

the operational requirement of three hours. This time was also very sensitive to temperature, being 410

less than half of the thickening time at room temperature (not reported). Unretarded versions of 411

geopolymers P4 and P5 had very similar results, as well as versions retarded with lignosulfonate-412

based additives (values not reported). However, the addition of the AMPS-based retarder to P4 and 413

P5 resulted in thickening time similar to that of the reference slurry (P2), as shown in Table 5. In all 414

cases, the consistency curve in Bearden units (Bc) shows a stable behavior with a right angle set 415

(less than 30 minutes between 50 Bc and 100 Bc). Based on the results of the standardized tests, we 416

can conclude that formulations P4 and P5 could be successfully pumped and placed in the intended 417

field operations. 418

Table 5 – API RP 10B-2 (2013) test results. 419

Slurry Temperature °C [ºF]

Plastic viscosity Pa.s [cP]

Yield stress Pa [lbf/100ft2]

Fluid loss cm3/30 min

Free water cm3

TT* min

P2 31 [88] 0.0674 [67.4] 5.84 [12.2] 59 0 218

P3 27 [80] 0.103 [103] 4.33 [120] 120 1.0 109

P4 31 [88] 0.140 [140] 17.0 [66] 66 0.5 210

P5 31 [88] 0.153 [153] 25.0 [92] 92 0.5 180 * Thickening time at the BHCT (31°C, 88°F)

420

4.4 Mechanical tests 421

Table 6 summarizes the mechanical properties measured in this work. Reference slurry P2 is 422

designed to be more compliant and deformable than a neat cement paste, but it is weaker in both 423

compression and tension. Latex allows the slurry to deform plastically before rupture, as shown in 424

Page 28: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

27

Fig. 16. The base geopolymer P3 is even more compliant, but its strength is lower than that of P2. 425

Addition of microsilica improves the compressive strength and increases the stiffness, while 426

wollastonite effectively increases the tensile strength, as observed by several authors (Silva, 2000; 427

Thaumaturgo and Silva, 1999; Silva and Thaumaturgo, 2003; Dias and Thaumaturgo, 2004; Zhang 428

et al. 2008). Even with fibers, failure of the geopolymers was always very brittle, as seen in Fig. 17 429

to Fig. 20. Curing at 10 MPa for 48 h had no significant effect on the mechanical properties. 430

Table 6 – Mechanical properties of cement (P2) and geopolymer (P3, P4 and P5) pastes. 431

Property Units P2 P3 P4 P5

Uniaxial compressive strength MPa [psi]

18.1 [2630]

17.4 [2520]

38.6 [5600]

37.4 [5420]

Young modulus GPa [ksi]

6.8 [986]

3.50 [508]

5.03 [730]

5.07 [735]

Poisson ratio 0.20 0.20 0.21 0.19

Friction angle ° 20.73 22.1 20.5 20.8

Cohesion MPa [psi]

5.86 [850]

7.33 [1060]

13.3 [1930]

13.3 [1930]

Tensile strength MPa [psi]

2.06 [299]

1.26 [183]

1.51 [219]

2.30 [336]

432

433

Fig. 16 – Stress-strain curves of triaxial compression tests on the reference paste P2. Strain values 434 are influenced by the compliance of the frame. Conversion: 1 MPa = 145 psi. 435

Strain

Str

ess

(MP

a)

Page 29: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

28

436

Fig. 17 – Stress-axial strain and stress-lateral strain curves of uniaxial compression tests on 437 geopolymer pastes. Conversion: 1 MPa = 145 psi. 438

Fig. 18 – Stress-strain curves of triaxial compression tests on the base geopolymer P3. Strain values are influenced by the compliance of the frame. Conversion: 1 MPa = 145 psi.

Strain

Str

ess

(MP

a)

Str

ess

(MP

a) S

tres

s (M

Pa)

Strain

Strain

Page 30: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

29

Fig. 19 – Stress-strain curves of triaxial compression tests on the geopolymer with microsilica P4. Strain values are influenced by the compliance of the frame. Conversion: 1 MPa = 145 psi.

Fig. 20 – Stress-strain curves of triaxial compression tests on the geopolymer with microsilica and mineral fiber P5. Strain values are influenced by the compliance of the frame. Conversion: 1 MPa = 145 psi.

4.5 Numerical Simulations 439

Steam injection causes the expansion of the casing against the cement or geopolymer and the 440

formation. For stiff formations (sections 1 and 2), the rock confines the material in the annulus, 441

inducing compressive stress states that could cause shear failure. All formulations were able to 442

resist these stresses and reach the final temperature, as shown in Table 7, but the geopolymers have 443

a wider safety margin due to their higher compressive strength. For this scenario, there is no risk of 444

failure with communication to the surface. 445

Compliant formations (sections 3 to 5) allow the cement sheath to expand, creating tensile stresses. 446

In the shallowest section (3), the initial hydrostatic pressure is insufficient to prevent tensile failure 447

of the sheath; therefore, all formulations fail in tension during heating. This translates to a risk of 448

zonal communication at shallow depths. 449

Base geopolymer P3 is the most compliant, thus generating the lowest tensile stress levels. This 450

explains why the formulation resists the highest temperature, even with the lowest tensile strength. 451

Formulations P4 and P5 have similar stiffness; therefore, the fiber addition is the determining factor 452

Str

ess

(MP

a)

Strain

Page 31: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

30

that allows P5 to withstand almost the same temperature as P2. The authors noticed that 453

geopolymers have lower heat conductivity and hold larger temperature gradients, resulting on a 454

secondary failure at section 4. 455

Table 7 – Failure temperature (when applicable) when heating to 260°C [500°F]. All failures 456 were tensile by radial cracking. 457

Section P2 P3 P4 P5

1. Carbonate at 150 m [492 ft] – – – –

2. Carbonate at 200 m [656 ft] – – – –

3. Sandstone at 200 m [656 ft] 170°C [339°F]

217°C [423°F]

150°C [302°F]

167°C [332°F]

4. Sandstone at 350 m [1148 ft] – – 225°C [438°F]

238°C [461°F]

5. Sandstone at 500 m [1640 ft] – – – – 458

5 SUMMARY AND CONCLUSIONS 459

In this work, the authors developed and tested three typical metakaolin-based geopolymer 460

formulations, a base geopolymer (P3), a geopolymer with microsilica (P4) and a geopolymer with 461

microsilica and mineral fiber (P5). Starting with ideal molar ratios and preliminary studies in the 462

literature (P3), we introduced an alternative source of silica, increasing the fraction of solids (P4) 463

and added fiber reinforcement (P5). From P3 to P5, we observed an increase in compressive 464

strength of 120%, in elastic modulus of 45% and in tensile strength of 80%, as well as a 20% 465

reduction in shrinkage. 466

Physical-chemical analysis by XRD, RMN, SEM and FTIR verified the quality of the source 467

materials and showed that the same compact and uniform geopolymer network was obtained, after 468

partial replacement of silicate by microsilica, due to care in preserving molar ratios, demonstrating 469

that microsilica contributed to the geopolymer matrix formation and that the fiber remained intact as 470

a reinforcement. 471

These formulations were subjected to a set of standard well cementing tests, demonstrating that P4 472

and P5 are suitable for field operations in the proposed scenario, replacing Portland-based slurry P2. 473

Page 32: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

31

The thickening time of the base geopolymer P3 is two short for operations, but the AMPS-based 474

setting retarder was effective for P4 and P5 at 31°C [88°F], increasing the thickening time from less 475

than two hours (P3) to three and a half hours (P4 and P5), a time comparable to the reference slurry 476

P2. 477

Mechanical tests show that compared to the reference cement paste P2, the base geopolymer P3 is 478

more compliant, but very brittle and weaker in tension. Addition of microsilica in P4 increased its 479

stiffness, while keeping it more compliant than P2, and improved significantly the compressive 480

strength. Mineral fibers in P5 improved the tensile strength with no detrimental side effects. All 481

geopolymer formulations were very brittle, while P2 was able to sustain significant deformation 482

before failure. This advantage is offset by the fact that P4 and P5 have more than twice the 483

compressive strength of P2, 10% higher tensile strength and 25% smaller stiffness. 484

Numerical simulations demonstrate that the best geopolymer is able to withstand injection 485

temperatures as high as the reference paste. The failure mode is strongly dependent on the initial 486

state of stress and on the stiffness of the formation. Compliant formations and low initial pressures 487

lead to tensile failures. Stiff formations tend to confine the cement sheath, changing the mode to 488

shear failure (microcracking in the case of P2, or brittle fracture for the geopolymers). 489

The main contribution of this work is a design workflow from fluid state properties to solid state 490

properties and, through numerical simulation, to the mechanical integrity over the life of the well. 491

This allows us to recommend these formulations for use in the field for this particular application, 492

grounded on solid experimental and numerical results. We suggest that more applications of 493

geopolymers in well cementing should be analyzed with this comprehensive methodology, 494

effectively integrating these alternative materials to the oil industry portfolio. 495

6 ACKNOWLEDGEMENTS 496

The authors acknowledge the support of Halliburton Services, which provided some of the chemical 497

additives used in the work, and the mechanical testing facilities of LABEST/COPPE/UFRJ. 498

Page 33: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

32

This work was supported by the National Agency for Petroleum, Natural Gas and Biofuels (ANP); 499

and Petrobras S.A. 500

Declarations of interest: none. 501

7 REFERENCES 502

Alves, I., Aragao, A., Bastos, B-L. et al. Pre-Salt Santos Basin C: Well Construction 503 Learning Curve Acceleration. In: Offshore Technology Conference, Paper OTC 20177, 504 Houston, Texas, USA, 2009. 505

API RP 10B-2: Recommended practice for testing well cements, second edition, American 506 Petroleum Institute, 2013. 507

ASTM C39 / C39M-16b: Standard Test Method for Compressive Strength of Cylindrical Concrete 508 Specimens, American Society For Testing And Materials, West Conshohocken, PA, 2016. 509

ASTM D2664-04: Standard Test Method for Triaxial Compressive Strength of Undrained Rock 510 Core Specimens Without Pore Pressure Measurements, American Society For Testing And 511 Materials, Filadélfia, PA, 2004. 512

Barlet-Gouedart, V., Zusatz-Ayache, B., Porcherie, O. Geopolymer composition and application in 513 oilfield industry. U.S. Patent US7794537B2, 2010a. 514

Barlet-Gouedart, V. , Zusatz-Ayache, B., Porcherie, O. Geopolymer composition and application 515 for carbon dioxide storage. U.S. Patent US7846250B2, 2010b. 516

Bour, D. Cyclic steam well design – A new approach to solve an old problem of cement sheath 517 failure in cyclic steam wells, In: SPE Western Regional Meeting, Irvine, CA, SPE paper 93868, 518 2005. 519

Chatterji, J.; Brenneis D. C.; Keys, C. L. Geopolymer cement compositions and methods of use. U.S. 520 Patent US9346711B2, 2016. 521

Davidovits, J. Geopolymers: inorganic polymeric new materials. Journal of Thermal Analysis 37, 522 1636-1656, 1991. 523

Davidovits, J. Method for obtaining a geopolymeric binder allowing to stabilize, solidify and 524 consolidate toxic or waste materials. U.S. Patent 5.349.118, 1994a. 525

Davidovits, J., Properties of Geopolymer Cements. In: Proceedings First International Conference 526 on Alkaline Cements and Concretes, Kiev, Ukraine, p.131, 1994b. 527

Dean, G. D., Torres, R. S. Novel Cement System for Improved Zonal Isolation in Steam Injection 528 Wells, SPE paper 78995, 2002. 529

Dias, D. P.; Thaumaturgo, C., Fracture toughness in geopolymeric concretes reinforced by basalt 530 fibers, Cement and Concrete Composites 27, 49–54, 2004. 531

Duxson, P.; Provis, J.L. et al. Understanding the relationship between geopolymer composition, 532 microstructure and mechanical properties, Colloids and Surfaces A: Physicochem. Eng. Aspects 269, 533 47-58, 2005. 534

Faure, P. F., Rodts, S., Proton NMR relaxation as a probe for setting cement pastes, Magnetic 535 Resonance Imaging 26, 1183–1196, 2008. 536

Giasuddin, H. M.; Sanjayan, J. G.; Ranjith , P. G. Stress versus strain behavior of geopolymer 537 cement under triaxial stress conditions in saline and normal water. Development 2, 12, 2013. 538

Page 34: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

33

Goodwin, K. J.; Crook, R. J. Cement sheath stress failure, SPE Drilling Engineering, SPE paper 539 20453, 1992. 540

International Society for Rock Mechanics, Suggested Method for Determining Tensile Strength of 541 Rock Materials - 1978 [EUR 4], In: Ulusay, R. (ed.), The ISRM Suggested Methods for Rock 542 Characterization, Testing and Monitoring: 2007-2014, Springer, 2014. 543

Khalifeh, M.; Saasen, A. S.; Vralstad, T.; Hodne, H. Potential utilization of Geopolymers in plug 544 and abandonment operations, paper SPE-169231-MS, In: SPE Bergen One Day Seminar, Bergen, 545 Norway, 2014. 546

Mahmoudkhani, A. H., Huynh, D. N. T. et al. New environmental-friendly cement slurries with 547 enhanced mechanical properties, In: CIPC/SPE Gas Technology Symposium, Calgary, Canada, SPE 548 paper 115004, 2008. 549

Marinho, E.P. Desenvolvimento de Pastas Geopoliméricas para Cimentação de Poços de Petróleo 550 [Development of Geopolymeric Slurries for Oil Well Cementing, in Portuguese], Doctoral Thesis, 551 Universidade Federal do Rio Grande do Norte, Natal, Brazil, 2004. 552

Nasvi, M M.C., Gamage, R. P., Jay, S. Geopolymer as well cement and the variation of its 553 mechanical behavior with curing temperature. Greenhouse Gas Sci Technol. 2, 46-58 , 2012a. 554

Nasvi, M M.C., Gamage, R. P., Jay, S. Mechanical properties of geopolymer cement in brine: its 555 suitability as well cement for geological sequestration of carbon dioxide (CO2). In: ARMS 2012 - 556 7th Asian Rock Mechanics Symposium, October 15-19, Seoul, Korea, 2012b. 557

Nasvi, M.C.M; Ranjith, P.G.; Sanjayan, J. A numerical study of CO2 flow through geopolymer 558 under down-hole stress conditions: Application for CO2 sequestration wells. Journal of 559 Unconventional Oil and Gas Resources 7, 62-70, 2014a. 560

Nasvi, M.C.M, Ranjith, P.G., Sanjayan, J., Bui, H. Effect of temperature on permeability of 561 geopolymer: A primary well selant for carbon capture and storage wells. Fuel 117, 354-363, 2014b. 562

563

Nasvi, M.C.M, Ranjith, P.G., Sanjayan, J., Haque, A., Li, X. Mechanical behavior of wellbore 564 materials saturated in brine water with different salinity levels. Energy 66, 239-249, 2014c. 565

Nóbrega, A.K.C., Materiais Cimentantes de Pega Rápida à Base de Metacaulim [Fast Setting 566 Cementing Materials Based on Metakaolin, in Portuguese], Master´s Dissertation, Universidade 567 Federal do Rio Grande do Norte, Natal, Brazil, 2006. 568

Olawale, M. D. Syntheses, Characterization and Binding Strength of Geopolymers: A Review. 569 International Journal of Materials Science and Applications, 2 (6) 185-193, 2013. 570

Pacheco-Torgal, F., Castro-Gomes, J. and Jalali, S. Alkali-activated binders: a review. Part 1. 571 Historical background, terminology, reaction mechanisms and reaction products. Construction and 572 Building Materials 22, 1305-1314, 2008. 573

Paiva, M. D. M., Otimização e análise mecânica de pastas geopoliméricas para uso em poços 574 sujeitos à injeção cíclica de vapor [Optimization and mechanical analysis of geopolymeric slurries 575 for use in wells subject to cyclic steam injection, in Portuguese], Doctoral Thesis, Universidade 576 Federal do Rio Grande do Norte, Natal, 2008. 577

Pinto, E. N. M. G., Influência da adição de tetraborato de sódio e poliuretana não iônica nas 578 propriedades de pastas geopoliméricas para cimentação de poços de petróleo [Influence of sodium 579 tetraborate and nonionic poliurethane on the properties of geopolimeric slurries for oil well 580 cementing, in Portuguese], Master´s Dissertation, Universidade Federal do Rio Grande do Norte, 581 Natal, Brazil, 2007. 582

Page 35: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT

34

Porcherie, O.; Pershikova, E.; Evans, N. et al. Compositions and methods for well cementing. U.S. 583 Patent US9394202B2, 2016. 584

Sá Ribeiro, R. A.; Sá Ribeiro, M. G.; Kriven, W. M. A review of particle- and fiber-reinforced 585 metakaolin-based geopolymer composites, J. Ceram. Sci. Tecnol. 307-322, 2017. 586

Rosa, A. J., Carvalho, R. S., Xavier, J. A. D. Engenharia de Reservatórios de Petróleo [Oil 587 Reservoir Engineering, in Portuguese], Interciência, Rio de Janeiro, Brazil, 2006. 588

Salehi, S.; Khattak, M.J.; Rizvi, H.; Karbalaei, S.F.; Kiran, R. Sensitivity analysis of fly ash 589 geopolymer cement slurries: Implications for oil and gas wells cementing applications, Journal of 590 Natural Gas Science & Engineering 37, 116-125, 2016. 591

Salehi, S.; Khattak, M.J.; Bwala, A. H.; Karbalaei, S.F. Characterization, morphology and shear 592 bond strength analysis of geopolymers: Implications for oil and gas wells cementing applications, 593 Journal of Natural Gas Science & Engineering 38, 323-332, 2017. 594

Salehi, S.; Khattak, M.J.; Ali, N.; et al. Study and use of geopolymer mixtures for oil and gas well 595 cementing applications. Journal of Energy Resources Technology 140 (1), 012908, 2018. 596

Silva F. J.; Thaumaturgo, C., Fiber reinforcement and fracture response in geopolymeric mortars, 597 Fatigue and Fracture of Engineering Materials and Structures 26, 167–172, 2003. 598

Silva, F. J. Reforço e fratura em compósitos de matriz álcali-ativada [Reinforcement and fracture in 599 composites with alkali-activated matrix, in Portuguese], Doctoral Thesis, Instituto Militar de 600 Engenharia, Rio de Janeiro, 2000. 601

Singh, P. S., Trigg, M., et al., Geopolymer formation processes at room temperature studied by 29Si 602 and 27Al MAS-NMR, Mater. Sci. and Eng. A 396, 392–0402, 2005. 603

Steveson, M.; Sagoe-Crentsil, K., Relationships between composition, structure and strenth of 604 inorganic polymers. Part I – metakaolin-derived inorganic polymers, J. Mater. Sci. 40, 2023–2036, 605 2005. 606

Tavares, P.B.; Vieira, E. et al., Análise Microscópica de matrizes geopoliméricas [in Portuguese]. 607 In: GEOPOLÍMEROS: Desenvolvimentos recentes e aplicações na engenharia, Portugal, 2004. 608

Thaumaturgo, C., Silva, F. J. Argamassa de poli(sialato-siloxo) reforçada por microfibras de 609 wollastonita [in Portuguese]. Acta Microscópica 8, 99-100, 1999. 610

Zhang, Y.; Sun, W.; Li, Z. et al., Impact properties of geopolymer based extrudates incorporated 611 with fly ash and PVA short fiber, Construction and Building Materials 22, 370–383, 2008. 612

613

Page 36: A geopolymer cementing system for oil wells subject to

MANUSCRIP

T

ACCEPTED

ACCEPTED MANUSCRIPT• Composite geopolymer formulations satisfy field operation requirements

• AMPS-based setting retarder was effective at the temperature of this application

• New formulations are less stiff than the Portland-based reference

• Simulations show that resistance to steam injection is comparable to the reference