Wellbore Cement Porosity Evolution in Response to Mineral Alteration

surface characteristics, complicating permeability and storage security predictions. In this paper,. 11 we report a small/wide angle scattering study ...
0 downloads 0 Views 2MB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

Wellbore Cement Porosity Evolution in Response to Mineral Alteration During CO Flooding 2

Michael C. Cheshire, Andrew G. Stack, J. William (Bill) Carey, Lawrence M. Anovitz, Timothy R Prisk, and Jan Ilavsky Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b03290 • Publication Date (Web): 13 Dec 2016 Downloaded from http://pubs.acs.org on December 16, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 25

Environmental Science & Technology

1

Wellbore Cement Porosity Evolution in Response to

2

Mineral Alteration During CO2 Flooding

3

Michael C. Cheshire1*, Andrew G. Stack1, J. William Carey2, Lawrence M. Anovitz1, Timothy R.

4

Prisk1, Jan Ilavsky3 1

5 2

6

Oak Ridge National Laboratory, Oak Ridge, TN 37831

Los Alamos National Laboratory, Los Alamos, NM 87545 3

7

Argonne National Laboratory, Argonne, IL 60439

8 9

Abstract

10

Mineral reactions during CO2 sequestration will change the pore-size distribution and pore-

11

surface characteristics, complicating permeability and storage security predictions. In this paper,

12

we report a small/wide angle scattering study of wellbore cement that has been exposed to

13

carbon dioxide for three decades. We have constructed detailed contour maps that describe local

14

porosity distributions and mineralogy of the sample, and relate these quantities to the carbon

15

dioxide reaction front on the cement. We find that the initial bimodal distribution of pores in the

16

cement, 1-2 nm and 10-20 nm, are affected differently during the course of carbonation

17

reactions. Initial dissolution of cement phases occurs in the 10-20 nm pores, and leads to the

18

development of new pore-spaces that are eventually sealed by CaCO3 precipitation, leading to a

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 25

19

loss of gel and capillary nanopores, smoother pore surfaces, and reduced porosity. This suggests

20

that during extensive carbonation of wellbore cement, the cement becomes less permeable due to

21

carbonate mineral precipitation within the pore space. Additionally, the loss of gel and capillary

22

nanoporosities will reduce the cement reactivity to CO2 due to reactive surface area loss. This

23

work demonstrates the importance of understanding not just changes in total porosity, but how

24

the distribution of porosity evolves with reaction that affects permeability.

25 26

Introduction

27

Porous media are the principal host for any subsurface fluid and are, therefore, the main sites

28

for geochemical reactions associated with CO2-sequestration. How geochemical reactions and

29

their kinetics affect these media has been the subject of considerable research, particularly with

30

respect to potential mineral trapping of CO2.1 Recent work has considered potential pore-size

31

dependencies of geochemical reactions.2,3 If pore-size dependent geochemical reactions occur,

32

they could have significant impact on long-term security through changes to permeability and

33

sealing capacity in the subsurface CO2 storage complex.4,5,6,7 While it may be expected local

34

changes in porosity due to dissolution and carbonate or other mineral precipitation may

35

increase/reduce permeability4,5, pore-size dependent reactions may affect permeability in

36

unforeseen ways since functions describing permeability have a non-linear dependence on pore

37

size.6,7 A complicating factor is that large permeability variations (up to five orders of

38

magnitude) are observed in rocks with similar porosities.8 This natural variability in the starting

39

material makes reactive transport models difficult to implement, because permeability cannot be

40

determined on the basis of porosity alone.7 Complicating permeability calculations of porous

41

media are the tortuosity and constrictive parameters.9,10,11 Tortuosity is defined as ratio of the

ACS Paragon Plus Environment

2

Page 3 of 25

Environmental Science & Technology

42

flow length without porous media versus path length in porous media.9,10,11 Constrictivity, also

43

known as bottleneck effect, describes the variable openness of the flow path in a natural porous

44

media.9,10,11 Therefore, changes in the interfacial regions within the pore-spaces have direct

45

impact on the tortuosity, constrictivity, and subsequently, permeability. Furthermore, given the

46

large uncertainty in the initial permeability, coupled to pore structure alterations (i.e., changes to

47

porosity, pore-size distribution, tortuosity, and constrictivity), predicting permeability evolution

48

in response to anthropogenic perturbation becomes problematic.

49

In this study, we investigate detailed changes in porosity as a function of pore-size that occurs

50

in a field sample of wellbore cement exposed to CO2 during CO2-enhanced oil recovery.

51

Wellbore cement integrity is a major concern for CO2 sequestration operations as the cement is a

52

key element of the hydrodynamic sealing system, and thus CO2-reacted cement represents a

53

potential leakage pathway.12,13 In addition to their intrinsic importance in establishing zonal

54

isolation in wells, ordinary portlandite cements are an excellent analogue for natural nano-porous

55

geological materials containing phases that with a high degree of reactivity permit more efficient

56

experimental observations. Examples of potential CO2 storage materials with reactive phases

57

include some caprocks14,15, volcanic sediments such as the Nagaoka injection site16, and basalts

58

present at Big Sky17 and Carbfix.18,19

59

The wellbore cement investigated here was recovered from a CO2 injection well located in the

60

SACROC unit in West Texas.20 The well was a producer for 25 years prior to conversion to a

61

CO2/water injection well for 30 years for the purposes of enhanced oil recovery (Figure 1a).

62

Analysis of the wellbore cement showed that it was partially carbonated due to CO2 leakage

63

along the cement/rock interface (Figure 1b).20 The unaltered portion of the wellbore cement is

64

composed of 86 wt. % amorphous materials. The remaining material consists of typical cement

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 25

65

phases, e.g., portlandite, katoite, brownmillerite, and calcite. The carbonated cement had

66

experienced dissolution and precipitation reactions in which cement phases dissolved (i.e.,

67

portlandite and calcium silica hydrate) and calcite and aragonite with lesser amounts of vaterite

68

(total CaCO3 phases 55%) precipitated along with various amorphous phases (i.e., silica). Thus,

69

the SACROC wellbore cement shows a reaction series that may be expected to occur at CO2

70

sequestration sites.21 This process of carbonated fluids interacting with reactive phases with

71

potential porosity and permeability alteration impacts sequestration site performance.10 We,

72

therefore, focused this investigation on measuring cement porosity evolution in response to

73

chemical and mineral alteration during CO2 injection to provide a better understanding of pore-

74

scale processes and how those processes might influence permeability and mechanical

75

properties.

76 77

Experimental Methods

78

The SACROC sample was prepared and analyzed as an ~50 µm thick thin-section mounted on

79

a borosilicate slide with epoxy (Figure 1b). It was analyzed via ultra-small (USAXS), small X-

80

ray scattering (SAXS) (to reveal pore distributions) and wide-angle (WAXS) X-ray scattering (to

81

reveal Bragg peaks of mineral phases) at the Advance Photon Source, Argonne National

82

Laboratory on beamline 9ID.22 Beam size for (U)SAXS measurements was 0.2 mm × 0.2 mm

83

and WAXS measurements were 0.2 mm × 1.5 mm. For (U)SAXS measurements, the sample was

84

rastered across the X-ray beam developing a 4 × 31 point grid covering most of the sample.

85

WAXS measurements were collected across the shale - orange cement (i.e., alteration zone) and

86

orange – grey cement interfaces (Figure 1b). Data reduction and analyses were conducted using

87

the Irena and Nika macro packages yielding Q-Intensity plots [Q = 4π/λ*sin θ, where θ is the

ACS Paragon Plus Environment

4

Page 5 of 25

Environmental Science & Technology

88

scattering angle and λ is the X-ray wavelength (0.6888(8) Å)].23,24 Porosity data was determined

89

using the methods described by Wang et al.25 and Anovitz et al.26, while scattering characteristics

90

were modeled using the Unified Fit model with a phase contrast determined between open

91

porosity (unepoxied pores) and cement solids.27,28 For detailed methods description regarding

92

scattering application to the characterization and analysis of porosity and pore structures were

93

refer the reader to review Anoviz and Cole29. Orange and grey zone cement compositions are

94

provided by Carey et al.20. Contouring and interpolation were done using Igor Pro 6.37 (Lake

95

Oswego, 2015). Manual indexing, phase identification, and peak area calculations were

96

performed on WAXS curves to determine the spatial distribution of mineralogy in the sample.

97

Errors presented with total porosity and pore sizes are one-sigma standard deviation from each

98

population.

99

Electron microscopic analyses were performed using a Hitachi S4800 scanning electron

100

microscope (SEM) at the High Temperature Materials Laboratory at Oak Ridge National

101

Laboratory (ORNL). The sample was carbon coated prior to SEM analysis and analyzed at 15.0

102

kV. Ion microscopy was performed with Zeiss’ Orion NanoFab He-ion microscope at ORNL’s

103

Center for Nanophase Materials Sciences.

104 105

Results and Discussion

106

Bulk mineral analyses shows the alteration zone is primarily (55 wt. %) carbonate minerals

107

(i.e., calcite, aragonite, and vaterite) and residual amorphous materials (44 wt. %).20 The exact

108

nature of the amorphous material is not well known, but may include unreacted or partially

109

altered calcium silica hydrate (C-S-H) phases, amorphous SiO2, iron (oxy)hydroxides, and

110

aluminum (oxy)hydroxides.31,32,33,13 WAXS data from this investigation shows the carbonated

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 25

111

cement is mineralogically zoned with aragonite predominantly along the shale interface with

112

calcite progressively increasing towards the reaction front and spiking at the edge of the reaction

113

boundary (Figure 2a). There is a weak inverse relationship between calcite and aragonite

114

suggesting environmental changes leading to the predominance of either aragonite or calcite.

115

Lighter colored diagonal veins within the orange alteration zone are dominated by calcite,

116

whereas, orange colored areas appear to be aragonite dominated. Vaterite was previously

117

reported20 from bulk XRD data, but is absent in the WAXS data. This absence is most likely due

118

to high detection limits associated with the WAXS setup. Carbonate mineral detection limits for

119

the WAXS system are currently not known, but previous20 bulk XRD analyses on the SACROC

120

sample show vaterite to be present around 7 wt.%.

121

The genetic relationships among calcite and aragonite (and vaterite) in this particular sample

122

are not known. Whether these carbonate minerals are from direct precipitation or a solid-state

123

transformation is essential to developing a detailed model on the microenvironments associated

124

with well-bore cements during CO2 flooding or sequestration. Our observations suggest that

125

calcite does not form as a function of aging (the region adjacent to the shale has been carbonated

126

longer than the interior regions of the cement). The data suggest a hypothesis that aragonite

127

forms at the highest degrees of supersaturation and calcite forms at lower supersaturation, as it is

128

possible that the pH and CO2 content of the cement pore fluid declined as CO2 penetrated the

129

cement.

130

The parent cement contains primarily amorphous phases with heterogeneously distributed

131

portlandite (Figure 2b), both in abundance and particle orientation. There are at least two calcite

132

filled fracture sets in the parent wellbore cement that are perpendicular and subparallel to the

133

reaction interface. The timing associated with the calcite infilling is not documented, so it is

ACS Paragon Plus Environment

6

Page 7 of 25

Environmental Science & Technology

134

difficult to determine its closure is concurrent with the CO2-induced alteration or occurred during

135

a separate event. In either case, calcite-filled fractures suggest that these fractures acted as

136

conduits for CO2/CO32--bearing fluids at some point during the wellbore’s lifetime.16 Fracture

137

self-sealing within cement from calcite and amorphous silica precipitation has been

138

demonstrated experimentally by Huerta et al.34 and Abdoulghafour et al.35. In these particular

139

examples Ca2+, H4SiO4(aq), and CO32- residence times within the fracture system determine

140

whether pH is buffered followed by calcite and silica precipitation.34,35 The nearly complete

141

calcite infilling associated with the SACROC fractures would further suggest that fluid flow

142

during this period was minimal. Interestingly, work from Walsh et al.36,37 adds that hydraulic

143

apertures are further reduced due to changing mechanical properties and fracture roughness

144

deformation. Portlandite within the parent cement indicates that CO2 has not diffused into the

145

parent cement where fractures are not present.20

146

SEM analysis of the parent cement shows textures associated with typical cement phases

147

(Figure 3a). Individual unhydrated clinkers are still present in the cement phase, but the majority

148

of the sample appears to be hydration products filling the voids between the clinkers (Figure 3a).

149

As CO2 or H2CO3 + H2O permeated into the cement, the parent cement textures were completely

150

removed as carbonates and amorphous phases formed (Figure 3b).

151

Plotting X-ray scattering data with respect to analysis position allowed us to produce contour

152

maps of total porosity, porosity ranges, and fractal characteristics in relationship to microscopic

153

and mineralogical data. Figure 4a shows the total porosity contours overlaid on the alteration

154

zone and parent wellbore cement. Scattering results from within the unaltered, parent cement

155

(i.e., gray areas) indicates total porosity ranging from 0.129 to 0.473 and an average porosity of

156

0.27(8) with a bimodal nanopore distribution (Figure 4b; i.e., 1 - 2 nm and 19(2) nm populations)

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 25

157

representing gel nanopores and medium capillary nanopores.38,39 The average porosity

158

determined via scattering data is similar to the 0.335 porosity value reported by Carey et al.20

159

using He-pycnometry. These results are consistent with previous works showing good

160

agreements between low-pressure gas sorption techniques (such as CO2, N2 BET and He-

161

pycnometry) and scattering techniques used for porosity and pore structure determinations.40 The

162

alteration zone (i.e., orange areas) shows a less heterogeneous porosity distribution, ranging from

163

0.075 to 0.252, with an average porosity of 0.14(4) (Figure 4a). Reaction fronts between

164

alteration zone and parent wellbore cements show a gradual porosity decrease as the parent

165

cement is carbonated. High porosity zones occur in the parent cement and are farthest away from

166

the alteration front, while low porosity zones are associated with the alteration zone. Empirical

167

porosity-permeability relationship such as the Kozeny-Carman relation suggests that this change

168

in porosity has the potential to result in permeability changes along the cement/rock interface of

169

approximately seven orders of magnitude (0.075 yielding 3.8×10-11 Darcys to 0.473 yielding

170

2.4×10-4 Darcys.7,41-43 Although the Kozeny-Carman model is a simplified porosity-permeability

171

relationship, empirical results all indicate a strong non-linear dependence of permeability on

172

porosity.7

173

At the reaction interface, pore occlusion occurs primarily within the 5 - 250 nm pores. Pore

174

volumes are approximately cut in half for 100 nm or larger pores, whereas a reduction in volume

175

of only about a third is observed for 10 nm pores, and significant reduction for 1 - 2 nm pores

176

(Figure 4b). CO2-alteration appears to completely modify cement porosity through dissolution

177

and reprecipitation, thereby removing all measurable gel nanopores (i.e., 1 - 2 nm pore sizes)

178

while developing a unimodal, capillary pore distribution centered at 17(2) nm (Figure 4b). Pore

179

sizes greater than ~250 nm appear to have experienced little alteration. If the trend in Figure 4 is

ACS Paragon Plus Environment

8

Page 9 of 25

Environmental Science & Technology

180

taken at face-value, it seems to suggest a sequence of where initial reactions, perhaps

181

precipitation, lead to constriction of 10-20 nm pores (points A-B), followed by concomitant

182

dilation of 10-20 nm pores and disappearance of the 1- 2 nm pores, likely due to dissolution

183

(points B-C), and lastly back-filling of 10-20 nm pores, likely due to precipitation (points C-D).

184

This pore-size dependence will modify permeability differently than if reactions occurred

185

uniformly across pore size.

186

Reaction front porosity characteristics from the SACROC wellbore cement are not consistent

187

to reaction front porosities observed in previous experiments.30,31,44 Experiments where various

188

cements were exposed to supercritical CO2 show sharp leading edge boundaries, characterized

189

by a high-porosity zone due to acid-induced dissolution and by a low porosity zone from

190

carbonate and/or silica precipitation at the front.30,31,44 These differences may suggest the

191

occurrence of multiple alteration events in the SACROC sample due to multiple periods of CO2

192

infiltrations and potential pulses of CO2 upwelling. The experiments, on the other hand, were

193

conducted under steady pCO2 conditions and shorter reaction times (i.e., up to 1 year versus 30

194

years).30,31,44

195

The scattering data yield information on the fractal dimension of the pore-fluid surfaces

196

present in the cement as alteration proceeds. Porous media often display two types of fractal

197

microstructure: (1) mass fractals, whose dimension (Dm; power-law slope = -Dm) is indicated by

198

a power-law slope of ≥ -3 on a plot of log I(Q) as a function of log Q; and (2) surface fractals

199

(Ds; power-law slope = (Ds – 6)), whose slope is between -3 and -4 (Figure 5a).25,26,29,45-48 Mass

200

fractals (Dm) describe the spatial pore distribution of porous solids. Compact structures typically

201

have Dm ~ 3 and Dm decreases as the pore volumes become increasingly open networks.25,45,46

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 25

202

For surface fractals (Ds), smooth surfaces are described with Ds = 2 (i.e., slope = -4), and,

203

approach 3 with roughening (i.e., slope = -3).25,45,46

204

The parent cement in this sample has a rough pore-solid interface indicated by an average

205

surface fractal, Ds, of 2.96(13), and Ds decreases to 2.43(20) due to CO2-alteration (Figure 5b).

206

The decrease in Ds implies an overall smoothing of the surface and possible reduction of surface

207

area. Decreasing Ds has been suggested34 to proceed via preferential solute (e.g., CO32-, Ca2+)

208

sorption onto rough, more energetically favorable sites, effectively annealing rough surfaces and

209

reducing the number of reactive sites.

210

Mass fractal scattering (i.e., Dm ) is observed over length-scales ranging from 22 nm to 630 nm

211

for both the alteration zone and parent wellbore cements. Changes in Dm suggest that less

212

compact, more open structures develop due to carbonation (Figure 5c). The parent cement shows

213

a fairly uniform Dm ranging from 2.81 to 2.99, further suggesting a more compact and tight pore

214

structure compared to the alteration zone.49,50

215

Reasons for the limited infiltration of CO2 and carbonation are difficult to determine without

216

detailed knowledge of the temporal variations of CO2 at the cement-shale interface. However, a

217

possible explanation based on experimental data for limited CO2-brine migration is that coupled

218

geochemical/geomechanical processes reduce permeability36,37, swelling of the amorphous

219

material35, filling of fractures with calcium carbonate34, or due to a fluctuating/limited CO2

220

supply. In any case, limited wellbore cement carbonation is consistent with experimental studies

221

of cement and supercritical CO2.30-32

222

Carbonation reactions result in an increase in volume with a possible change in molar volume

223

up to +53.6 cm3 mol-1 (assuming a C-S-H phase with composition 3.4CaO•2SiO2•3H2O, molar

224

volume = 128 cm3 mol-1) for a dewatered, carbonated cement with an amorphous silica

ACS Paragon Plus Environment

10

Page 11 of 25

Environmental Science & Technology

225

byproduct.39 The observed gel and capillary nanoporosity loss is consistent with these estimates

226

of complete carbonation. Gel nanopore loss coincides with C-S-H dissolution during

227

carbonation. Further, C-S-H and portlandite dissolution develops new porosities that are

228

subsequently closed off during CaCO3 precipitation.

229

Empirical observations have shown a non-linear inverse relationship between cement porosity

230

and compressive/tensile strengths51 and carbonation of cement has been found to increase

231

compressive strength.52,53 However, in three-point bending tests, CO2 attack on cement yielded a

232

~93% decrease in strength and ~84% decrease in elastic modulus.54 An interesting complexity

233

overlying these porosity-mechanical relationships are microfracture developments near the

234

reaction front due to molar volume expansion during carbonate mineral formation coupled with

235

alteration zone (i.e., unaltered cement, portlandite depleted, carbonate mineral, and amorphous

236

aluminosilicate zones) displaying differing mechanical properties.55,56 Carbonate crystallization

237

can induce pore over-pressurization causing damage to the matrix in form of microfractures.

238

Fabbri55 speculates that most of these microfractures will occur at or near the reaction interface.

239

Many instances these microfractures will not be included in the porosity calculations and

240

therefore can make porosity-mechanical behavior relationships difficult to quantify. Additional

241

to the microfracture effects, Hangx56 has shown that alteration zones from carbonation displays

242

24-56% mechanical weakening in the portlandite-depleted and 150-181% mechanical

243

strengthening in the carbonation zones as compared to neat cement.

244 245

Environmental Implications

246

The ultimate environmental implications of this work are if carbonation reactions increase, or

247

decrease, CO2 storage security in cement. A review of cement carbonation behavior in relation

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 25

248

to CO2 sequestration indicates that carbonation of wellbore cement reduces permeability,10

249

consistent with the results on total porosity in this study. This would enhance CO2 storage

250

security. Coupled to this may be variable impacts on mechanical strength of the cement, but the

251

strength requirements for wellbore cement are limited.10 This work shows that reactions during

252

cement carbonation reduced the gel and capillary nanoporosities, but did not significantly alter

253

larger-scale porosity. This raises the importance of determining how the distribution of pore sizes

254

changes in context to the total porosity and how this affects overall permeability. Furthermore,

255

CO2 permeation and alteration was limited to only 8-10 mm (0.3 mm year-1) over the 30 years of

256

CO2 interaction with the wellbore cement. The small penetration depth supports work suggesting

257

that cement matrix leakage is not a primary leakage pathway.33 These observations can be related

258

to reservoir rock and caprock properties as well as cement. Pore-scale dependent reactions

259

coupled with a nonuniform alteration of the pore-sized distribution during reactive transport of

260

CO2 have the potential for compounding effect on the storage security of subsurface CO2.2,3

261 262

Acknowledgment

263

This work was supported as part of the Center for Nanoscale Control of Geologic CO2, an

264

Energy Frontier Research Center funded by the U.S. Department of Energy, Office of Science,

265

Basic Energy Sciences under Award #DE-AC02-05CH11231. This research used resources of

266

the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User

267

Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract

268

#DE-AC02-06CH11357. JWC acknowledges support from the DOE National Energy

269

Technology Laboratory (NETL) under Grant #FE-371-14-FY16, which is managed and

270

administered by Los Alamos National Laboratory and funded by DOE/NETL and cost/sharing

ACS Paragon Plus Environment

12

Page 13 of 25

Environmental Science & Technology

271

partners. He-ion microscopy was conducted at the Center for Nanophase Materials Sciences,

272

which is a DOE Office of Science User Facility. The authors would like to thank Larry Lake for

273

providing production data.

274 275

REFERENCES

276

(1) Gaus, I. Role and impact of CO2-rock interactions during CO2 storage in sedimentary rocks.

277

Int. J. Greenhouse Gas Control. 2010, 4, 73-89.

278

(2) Stack, A.G.; Fernandez-Martinez, A.; Allard, L.F.; Banuelos, J.L.; Rother, G.; Anovitz, L.M.;

279

Cole, D.R.; Waychunas, G.A. Pore-size-dependent calcium carbonate precipitation

280

controlled by surface chemistry. Environ. Sci. Technol. 2014, 48, 6177-6183.

281 282 283

(3) Stack, A.G. Pore-Size Dependent Precipitation Reactions: A Geochemical Frontier. Rev. Mineral. Geochem. 2015, 80, 165-190. (4) Borgia, A.; Pruess, K.; Kneafsey, T.J.; Oldenburg, C.M.; Pan, L. Numerical simulation of salt

284

precipitation in the fractures of a CO2-enhanced geothermal system. Geothermics. 2012,

285

44, 13–22.

286 287

(5) Cohen, Y. and Rothman, D.H. Mechanisms for mechanical trapping of geologically sequestered carbon dioxide. Proc. R. Soc. A. 2015, 471, 20140853.

288

(6) Freeze, R.A. and Cherry, J.A. Groundwater. Prentice Hall: Englewood Cliffs, NJ. 1979.

289

(7) Zhang, S.; DePaolo, D.J.; Xu, T.; Zheng, L. Mineralization of carbon dioxide sequestered in

290

volcanogenic sandstone reservoir rocks. Int. J. Greenhouse Gas Control. 2013, 18, 315-

291

328.

292

(8) Gibson-Poole, C.M.; Svendsen, L.; Underschultz, J.; Watson, M.N.; Ennis-King, J.; van

293

Ruth, P.J.; Nelson, E.J.; Daniel, R.F.; Cinar, Y. Site characterization of a basin-scale CO2

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 25

294

geological storage system: Gippsland Basin, southeast Australia. Environ. Geol. (N. Y.).

295

2008, 54, 1583-1606.

296

(9) Bear, J. Dynamics of Fluids in Porous Media. Dover: New York. 1988.

297

(10) Steefel, C.; Molins, S.; Trebotich, D. Pore scale processes associated with subsurface CO2

298 299 300

injection and sequestration. Rev. Mineral. Geochem. 2013, 77, 259-303. (11) Berg, C.F. Permeability description by characteristic length, tortuosity, constriction, and porosity. Transp. Porous Med. 2014, 103, 381-400.

301

(12) Gasda, S.E.; Bachu, S.; Celia, M.A. Spatial characterization of the location of potentially

302

leaky wells penetrating a deep saline aquifer in a mature sedimentary basin.

303

Environmental Geology. 2004, 46, 707-720.

304 305

(13) Carey, J. W. Geochemistry of wellbore integrity in CO2 sequestration: Portland cementsteel-brine-CO2 interactions. Rev. Mineral. Geochem. 2013, 77, 505-539.

306

(14) Smith, M.M.; Sholokhova, Y.; Hao, Y.; Carroll, S.A. Evaporite caprock integrity: An

307

experimental study of reactive mineralogy and pore-scale heterogeneity during brine-CO2

308

exposure. Environ. Sci. Technol. 2012, 47, 262-268.

309

(15) Kampman, N.; Busch, A.; Bertier, P.; Snippe, J.; Hangx, V.; Di, Z.; Rother, G.; Harrington,

310

J.F.; Evans, J.P.; Maskell, A.; Chapman, H.J.; Bickle, M.J. Obersvational evidence

311

confirms modeling of the long-term integrity of CO2-reservoir caprocks. Nature

312

Communications, 2016, 7, 12268.

313 314

(16) Mito, S.; Xue, Z.; Ohsumi, T. Case study of geochemical reactions at the Nagaoka CO2 injection site, Japan. Int. J. Greenhouse Gas Control. 2008, 2, 309-318.

ACS Paragon Plus Environment

14

Page 15 of 25

Environmental Science & Technology

315

(17) Zakharova, N.V.; Goldberg, D.S.; Sullivan, E.C.; Herron, M.M.; Grau, J.A. Petrophysical

316

and geochemical properties of Columbia River flood basalt: Implications for carbon

317

sequestration. Geochem., Geophys., Geosyst. 2012, 13, Q11001.

318

(18) Gislason, S.R.; Wolff-Boenisch, D.; Stefansson, A.; Oelkers, E.H.; Gunnlaugsson, E.;

319

Sigurdardottir, H.; Sigfusson, B.; Broecker, W.S.; Matter, J.M.; Stute, M.; Axelsson, G.;

320

Fridriksson, T. Mineral sequestration of carbon dioxide in basalt: A pre-injection

321

overview of the CarbFix project. Int. J. Greenhouse Gas Control. 2010, 4, 537-545.

322

(19) Matter, J.M.; Stute, M.; Snæbjörnsdottir, S.O.; Oelkers, E.H.; Gislason, S.R.; Aradottir,

323

E.S.; Sigfusson, B.; Gunnarsson, I.; Sigurdardottir, H.; Gunnarsson, E.; Axelsson, G.;

324

Alfredsson, H.A.; Wolff-Boenisch, D.; Mesfin, K.; Taya, D.F.; Hall, J.; Dideriksen, K.;

325

Broecker, W.S. Rapid carbon mineralization for permanent disposal of anthropogenic

326

carbon dioxide emissions. Science, 2016, 352, 1312-1314.

327

(20) Carey, J.W.; Wigand, M.; Chipera, S.J.; WoldeGabriel, G.; Pawar, R.; Lichtner, P.C.;

328

Wehner, S.C.; Raines, M.A.; Guthrie, G.D. Analysis and performance of oil well cement

329

with 30 years of CO2 exposure from the SACROC Unit, West Texas, USA. Int. J.

330

Greenhouse Gas Control. 2007, 1, 75-85.

331

(21) Johnson, J. W.; Nitao, J. J.; Knauss, K. G. Reactive transport modeling of CO2 storage in

332

saline aquifers to elucidate fundamental processes, trapping mechanisms and

333

sequestration partitioning. Spec. Publ. - Geol. Soc. London. 2004, 233, 107-128.

334

(22) Ilavsky, J.; Jemian, P.R.; Allen, A.J.; Zhang, F.; Levine, L.E.; Long, G.G. Ultra-small-angle

335

X-ray scattering at the Advanced Photon Source. J. Appl. Crystallogr. 2009, 42, 469-479.

336

(23) Ilavsky J. and Jemian, P. R. Irena: tool suite for modeling and analysis of small-angle

337

scattering. J. Appl. Crystallogr. 2009, 42, 347-353.

ACS Paragon Plus Environment

15

Environmental Science & Technology

338 339

Page 16 of 25

(24) Ilavsky, J. Nika: software for two-dimensional data reduction. J. Appl. Crystallogr. 2012, 45, 324-328.

340

(25) Wang, H.-W.; Anovitz, L.; Burg, A.; Cole, D.R.; Allard, L.F.; Jackson, A.J.; Stack, A.G.;

341

Rother, G. Multi-scale characterization of pore evolution in a combustion metamorphic

342

complex, Hatrurim basin, Israel: Combining (ultra) small-angle neutron scattering and

343

image analysis. Geochim. Cosmochim. Acta. 2013, 121, 339-362.

344

(26) Anovitz, L.M.; Cole, D.R.; Rother, G.; Allard, L.F.; Jackson, A.J.; Littrell, K.C. Diagenetic

345

changes in macro- to nano-scale porosity in the St. Peter Sandstone: An (ultra) small

346

angle neutron scattering and backscattered electron imaging analysis. Geochim.

347

Cosmochim. Acta. 2013, 102, 280-305.

348 349 350 351 352 353

(27) Beaucage, G. Approximations leading to a unified exponential/power-law approach to small-angle scattering. J. Appl. Crystallogr. 1995, 28, 717-728. (28) Beaucage, G. Small-angle scattering from polymeric mass fractals of arbitrary mass-fractal dimension. J. Appl. Crystallogr. 1996, 29, 134-146. (29) Anovitz, L.M and Cole, D.R. Characterization and analysis of porosity and pore structures. Rev. Mineral. Geochem. 2015, 80, 61-164.

354

(30) Kutchko, B.G.; Strazisar, B.R.; Dzombak, D.A.; Lowry, G.V.; Thaulow, N. Degradation of

355

well cement by CO2 under geologic sequestration conditions. Environ. Sci. Technol.

356

2007, 41, 4787-4792.

357

(31) Kutchko, B.G.; Strazisar, B.R.; Lowry, G.V.; Dzombak, D.A.; Thaulow, N. Rate of CO2

358

attack on hydrated class H well cement under geologic sequestration conditions. Environ.

359

Sci. Technol. 2008, 42, 6237-6242.

ACS Paragon Plus Environment

16

Page 17 of 25

Environmental Science & Technology

360

(32) Kutchko, B.G.; Strazisar, B.R.; Huerta, N.; Lowry, G.V.; Dzombak, D.A.; Thaulow, N.

361

CO2 reaction with hydrated class H well cement under geologic sequestration conditions:

362

Effects of flyash admixtures. Environ. Sci. Technol. 2009, 43, 3947-3952.

363

(33) Carroll, S.; Carey, J.W.; Dzombak, D.; Huerta, N.J.; Li, L.; Richard, T.; Um, W.; Walsh, S.;

364

Zhang, L. Review: Role of chemistry, mechanics, and transport on well integrity in CO2

365

storage environments. Int. J. Greenhouse Gas Control. 2016, 49, 149-160.

366

(34) Huerta, N.J.; Hesse, M.A.; Bryant, S.L.; Strazisar, B.R.; Lopano, C. Reactive transport of

367

CO2-saturated saturated in a cement fracture: Application to wellbore leakage during

368

geologic CO2 storage. Int. J. Greenhouse Gas Control. 2016, 44, 276-289.

369

(35) Abdoulghafour, H.; Luquot, L.; Gouze, P. Characterization of the mechanisms controlling

370

the permeability changes of fractured cements flowed through by CO2-rich brine.

371

Environ. Sci. Technol. 2013, 47, 10332-10338.

372

(36) Walsh, S.D.C.; Mason, H.E.; Du Frane, W.L.; Carroll, S.A. Experimental calibration of a

373

numerical model describing the alteration of cement/caprock interfaces by carbonated

374

brine. Int. J. Greenhouse Gas Control. 2014, 22, 176-188.

375

(37) Walsh, S.D.C.; Mason, H.E.; Du Frane, W.L.; Carroll, S.A. Mechanical and hydraulic

376

coupling in cement-caprock interfaces exposed to carbonated brine. Int. J. Greenhouse

377

Gas Control. 2014, 25, 109-120.

378 379 380 381

(38) Mindess, S.; Young, J.F.; Darwin, D. Concrete, 2nd Edition. Pearson: London, England, 2002. (39) Jennings, H.M. and Tennis, P.D. Model for the developing microstructure in Portland cement pastes. J. Am. Ceram. Soc. 1994, 77, 3161-3172.

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 25

382

(40) Clarkson, C.R.; Solano, N.; Bustin, A.M.M.; Chalmers, G.R.L.; He, L.; Melnichenko, Y.B.;

383

Radlinski, A.P.; Blach, T.P. Pore structure characterization of North American Shale gas

384

reservoirs using USANS/SANS, gas adsorption, and mercury intrusion. Fuel. 2013, 103,

385

606-616.

386 387 388 389 390 391

(41) Carman, P.C. Fluid flow through granular beds. Transactions of the Institute of Chemical Engineers. 1937, 15, S32-S48. (42) Goto, S.; Roy, D.M. The effect of W/C ratio and curing temperature on the permeability of hardened cement paste. Cem. Concr. Res. 1981, 11, 575-579. (43) Cui, L.; Cahyadi, J.H. Permeability and pore structure of OPC paste. Cem. Concr. Res. 2001, 31, 277-282.

392 393

(44) Rimmele, G.; Barlet-Gouedard, V.; Porcherie, O.; Goffe, B.; Brunet, F. Heterogeneous

394

porosity distribution in Portland cement exposed to CO2-rich fluids. Cem. Concr. Res.

395

2008, 38, 1038-1048.

396

(45) Martin, J.E. and Hurd, A.J. Scattering from fractals. J. Appl. Crystallogr. 1987, 20, 61-78.

397

(46) Allen, A.J.; Baston, A.H.; Wilding, C.R. Small angle neutron scattering studies of pore and

398

gel structures diffusivity, permeability and damage effects. Pore Structure and

399

Permeability of Cementitious Materials. 1989, 137, 119-125

400

(47) Anovitz, L.M.; Lynn, G.W.; Cole, D.R.; Rother, G.; Allard, L.F.; Hamilton, W.A.; Porcar,

401

L.; Kim, M.-H. A new approach to quantification of metamorphism using ultra-small and

402

small angle neutron scattering. Geochim. Cosmochim. Acta. 2009, 73, 7303-7324

403

(48) Anovitz L.M.; Cole D.R.; Jackson A.J.; Rother G.; Littrell K.C.; Allard L.F.; Pollington

404

A.D.; Wesolowski D.J. Effect of Quartz Overgrowth Precipitation on the Multiscale

ACS Paragon Plus Environment

18

Page 19 of 25

Environmental Science & Technology

405

Porosity of Sandstone: A (U)SANS and Imaging Analysis. Geochim. Cosmochim. Acta.

406

2015, 153, 199-222.

407

(49) Beddoe, R.E. and Lang, K. Effect of moisture on fractal dimension and specific surface of

408

hardened cement paste by small-angle x-ray scattering. Cem. Concr. Res. 1994, 24, 605-

409

612.

410 411 412 413

(50) Winslow, D.; Bukowski, J.M.; Young, J.F. The fractal arrangement of hydrated cement paste. Cem. Concr. Res. 1995, 25, 147-156. (51) Chen, X.; Wu, S.; Zhou, J. Influence of porosity on compressive and tensile strength of cement mortar. Construction and Building Materials. 2013, 40, 869-874.

414

(52) Taylor, H. F. W. Cement Chemistry. Academic Press: London, 1990.

415

(53) Takla, I.; Burlion, N.; Shao, J.-F.; Saint-Marc, J.; Garnier, A. Effects of the storage of CO2

416

on multiaxial mechanical and hydraulic behaviors of oil-well cement. J. Mater. Civ. Eng.

417

2011, 23, 741-746.

418

(54) Li, Q.; Lim, Y.M.; Flores, K.M.; Kranjc, K.; Jun, Y.-S. Chemical reactions of portland

419

cement with aqueous CO2 and their impacts on cement’s mechanical properties under

420

geologic CO2 sequestration conditions. Environ. Sci. Technol. 2015, 49, 6335-6343.

421

(55) Fabbri, A.; Corvisier, J.; Schubnel, A.; Brunet, F.; Goffe, B.; Rimmele, G.; Barlet-

422

Gouedard, V. Effect of carbonon the hydro-mechanical properties of Portland cements.

423

Cement and Concrete Research. 2009, 39, 1156-1163.

424

(56) Hangx, S.J.T.; van der Linden, A.; Marcelis, F.; Liteanu, E. Defining the brittle failure

425

envelopes of individual reaction zones observed in CO2-exposed wellbore cement. Env.

426

Sci. Tech. 2016, 50, 1031-1038.

427

ACS Paragon Plus Environment

19

Environmental Science & Technology

428

Page 20 of 25

List of Figures

429 430

Figure 1. A) Production and injection history from well 49-6 within the SACROC unit (L. Lake,

431

personal communication, 2015). B) Image from the sliced sample with parent wellbore cement

432

(grey), CO2 alteration zone (orange), and shale interfaces (modified from Carey et al.16).

433

ACS Paragon Plus Environment

20

Page 21 of 25

Environmental Science & Technology

434

A

p

p

p

p

p

435

B

p

436

Figure 2. A) Diffraction data from the WAXS analyses. Vertical graphs show raw integrated

437

peak intensities from the calcite 104 (red), aragonite 111 (purple), and portlandite 110, 101

438

(blue) indices. For clarity, the spots from the WAXS scans are not presented, but the vertical

439

graphs are present in the analysis locations. B) Helium-ion photomicrograph showing portlandite

440

plates (P, on edge) within the parent cement. Cubic features are probably halite.

441

ACS Paragon Plus Environment

21

Environmental Science & Technology

A

200 µm

B

200 µm

Page 22 of 25

442 443

Figure 3. Backscatter-SEM images showing the different textural and composition zones within

444

the SACROC sample. A) Grey, parent cement shows unhydrated clinkers with calcium silica

445

hydrate between clinkers. B) The carbonated region displays reworked textures with most of the

446

parent cement textures removed.

447

ACS Paragon Plus Environment

22

Page 23 of 25

Environmental Science & Technology

448 449

Figure 4. A) Contours of total porosity distribution associated with the SACROC wellbore

450

cement sample showing changes in porosity across the reaction interface between relatively

451

unaltered gray cement (bottom) and altered, orange zone (top). B) Individual porosity

452

distributions for points along the X=35 mm transect. Porosity of the parent cement (point A) has

453

a bimodal distribution with peak frequencies at 1 - 2 nm and 19(2) nm size distributions.

454

Reaction interface shows similar size distribution to the parent cement with a significant loss of

455

the larger pore sizes. Porosities within the orange alteration zone (points C and D) are markedly

456

altered and show a unimodal distribution with 17(2) nm pore sizes. Error bars represent

457

estimated uncertainty in the collected scattered X-ray intensity values propagated to the pore

458

volume distribution.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 25

459 460

Figure 5. A) (U)SAXS plot of log I(Q) as a function of log Q for a parent cement (black curve;

461

x35, y59) and carbonated cement (green curve; x35, y48). Red curves overlaying each scattering

462

curve are the modeled scattering curve from the Unified Fit modeling package. B) Surface fractal

463

dimension approaching Ds = 2 within the alteration zone. Carbonation is smoothing the pore-

ACS Paragon Plus Environment

24

Page 25 of 25

Environmental Science & Technology

464

solid interface. C) Mass fractal dimensions across the alteration zone and parent cement. The

465

mass fractal is much more variable in the carbonation zone (both higher and lower).

466 467

For Table of Contents Only

468

ACS Paragon Plus Environment

25