Dissolved CO2 Increases Breakthrough Porosity in ... - ACS Publications

Jun 16, 2017 - simulations with high resolution digital models of North Sea chalk, we show that the breakthrough porosity is an important indicator of...
0 downloads 0 Views 2MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Dissolved CO2 increases breakthrough porosity in natural porous materials Yi Yang, Stefan Bruns, Susan Louise Svane Stipp, and Henning Osholm Sørensen Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b02157 • Publication Date (Web): 16 Jun 2017 Downloaded from http://pubs.acs.org on June 18, 2017

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 33

Environmental Science & Technology

Dissolved CO2 increases breakthrough porosity in natural porous materials Yi Yang*1, Stefan Bruns1, Susan Louise Svane Stipp1 and Henning Osholm Sørensen1 1

Nano-Science Center, Department of Chemistry, University of Copenhagen, Universitetsparken 5, DK-

2100 Copenhagen, Denmark

*[email protected]; Tel: +45 9161 4575; Fax: +45 3532 0219

ACS Paragon Plus Environment

Environmental Science & Technology

1

ABSTRACT

2

When reactive fluids flow through a dissolving porous medium, conductive channels form, leading to

3

fluid breakthrough. This phenomenon is caused by the reactive infiltration instability and is important in

4

geologic carbon storage where the dissolution of CO2 in flowing water increases fluid acidity. Using

5

numerical simulations with high resolution digital models of North Sea chalk, we show that the

6

breakthrough porosity is an important indicator of dissolution pattern. Dissolution patterns reflect the

7

balance between the demand and supply of cumulative surface. The demand is determined by the

8

reactive fluid composition while the supply relies on the flow field and the rock’s microstructure. We

9

tested three model scenarios and found that aqueous CO2 dissolves porous media homogeneously,

10

leading to large breakthrough porosity. In contrast, solutions without CO2 develop elongated convective

11

channels known as wormholes, with low breakthrough porosity. These different patterns are explained

12

by the different apparent solubility of calcite in free drift systems. Our results indicate that CO2 increases

13

the reactive subvolume of porous media and reduces the amount of solid residual before reactive fluid

14

can be fully channelized. Consequently, dissolved CO2 may enhance contaminant mobilisation near

15

injection wellbores, undermine the mechanical sustainability of formation rocks and increase the

16

likelihood of buoyance driven leakage through carbonate rich caprocks.

17

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

Environmental Science & Technology

18

Introduction

19

A great challenge in predicting the consequences of geologic carbon storage (GCS) is to quantify the

20

interaction between flowing reactive fluids and geologic formations as a function of time.1-8 Making

21

progress in this direction requires considering a GCS reservoir as a constantly evolving flow system, that

22

is subject to geochemical and geomechanical instabilities.9, 10 The reactive infiltration instability (RII) is

23

the morphological instability of a migrating dissolution front resulting from heterogeneities in

24

petrophysical properties. RII controls the microstructural evolution of rock in response to an imposed

25

flow field and is the key to the self-organization of fluid flow through natural porous materials.11-17

26

Identifying the trigger of this instability, as well as the chemical reactions leading to the morphological

27

development,18, 19 is especially important for predicting the evolution of the sealing integrity of caprock

28

and the geomechanical deformation susceptibility of reservoir structures as host rocks are eroded away.

29

The morphological evolution of porous media caused by the reactive infiltration instability can be

30

described qualitatively using dissolution patterns (homogeneous, ramified, channelized etc.;20-22 also

31

Figure 1), or quantitatively, using breakthrough porosity, ϕc, the macroscopic porosity of a sample when

32

fluid creates a penetrating channel.20, 23, 24 Given a sample size, the breakthrough porosity is an indicator

33

of dissolution pattern. A large ϕc is typically associated with a uniformly dissolving morphology while a

34

small ϕc indicates the formation of fingering channels. Breakthrough porosity can also be qualitatively

35

related to the mechanical stability of a sample after fluid penetration. A large ϕc means little solid is left

36

behind and the porous microstructure is mechanically weak, and vice versa. These features of

37

breakthrough porosity motivated the use this parameter to describe the dynamic process of

38

microstructure development. Worth mentioning is that the breakthrough porosity of an evolving porous

39

medium is different from the solute breakthrough curve in a steady state flow field. The former

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 33

40

describes microstructural evolution while the latter characterises the transport within a static geometry

41

and can be effectively computed using streamline tracing.25

42

A strong coupling between mineral dissolution rate and rock permeability is essential for a dynamic

43

process involving RII. Coupling can be decomposed into 3 aspects of sensitivity.21, 26 First, the sensitivity

44

of the flow field to rock microstructure determines how the reactants and products of the water-rock

45

interactions are conveyed in a flow field so the chemical system remains far from equilibrium.27-30 This

46

sensitivity can be affected by variables such as the geometry and the texture of the pore surfaces,31-35

47

the properties of the fluids and their interactions during flow,36-38 as well as the body forces exerted by,

48

e.g. gravitational or electrostatic fields.39 Second, the sensitivity of the reaction rates to fluid

49

composition, which determines the spatial variations of chemical conversion (i.e. the extent of chemical

50

reaction, which is characterised by the relative amount of reactant converted to product).40 This

51

sensitivity is often complicated by the intrinsic mineral heterogeneity in natural porous media,41-44 and

52

by the large number of aqueous species involved in the reactions.45, 46 Finally, the sensitivity of the rock

53

microstructure to chemical conversion closes the loop of the positive feedback leading to RII. 47 This

54

sensitivity reflects the density, molar mass and the solubility of the dissolving or precipitating minerals.

55

If any of the three sensitivities becomes zero, the infiltration instability vanishes and the self-

56

organization of the porous structure ceases.13, 48, 49

ACS Paragon Plus Environment

4

Page 5 of 33

Environmental Science & Technology

57 58

Figure 1. A qualitative demonstration of the dissolution front instability in a 2D domain. The instability

59

determines the morphology of the dissolving porous medium. Yellow indicates the dissolving front. (a)

60

Stable dissolution front: mineral dissolution always takes place at the same distance from the injection

61

point. (b) – (d): increasingly unstable dissolution fronts, where the distance between the injection point

62

and the dissolution front varies considerably. The same amount of reactive fluid was injected in each

63

simulation and snapshots were taken after the same reaction time. Greater instability leads to a higher

64 65 66 67

likelihood of channel formation (wormholing).

68

thus affects all three types of sensitivity. For example, buoyant gaseous or supercritical CO2 might

69

induce Rayleigh-Taylor instability,50 influencing the spreading of the reactive aqueous phase and mixing

70

of the fluids.51 Dissolution of CO2 in water changes many aspects of the kinetics of the water-rock

71

interactions. Mineral dissolution rates increase because of lower pH and a stronger contribution from

72

carbonic and bicarbonate species through surface complexation.52, 53 The apparent order of reaction,

73

i.e., the sensitivity of reaction rate to reactant concentration, decreases because of the buffering effect

74

of dissolved CO2 as a weak acid.33 Apparent mineral solubility can also be affected by modified ion

Introducing CO2 into a natural porous formation changes many of the factors that control behaviour,

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 33

75

activity and buffered pH. Increased solubility is a thermodynamic driving force for the dissolution

76

reactions. It can change the relative stability of the secondary phases and as a result, the sensitivity of

77

the solid architecture to the extent of reaction.7, 8, 54, 55 Finally, the impact of CO2 depends on how the

78

gas is introduced.56, 57 Direct injection,58 surface mixing59 and wellbore mixing4, 60 all yield distinct

79

pressures and solution compositions.

80

The dynamics of dissolving microstructures and consequently, the CO2 effect on the breakthrough

81

properties of natural porous materials remain poorly understood for three reasons. First, structural

82

heterogeneities (e.g., spatial variations in porosity and permeability) can be viewed as perturbations for

83

reaction front migration. These perturbations are amplified by infiltration instability. Thus, in a series of

84

experiments, the initial microstructure must be identical for meaningful comparison of results, but this is

85

not possible in experiments with real materials.19, 31, 32 Second, there are few numerical models currently

86

available that can directly account for the pore scale heterogeneities that cause the instabilities that are

87

important for simulations to be valid.61-63 Recent advances in X-ray imaging, however, have made it

88

possible to create models using actual images from natural samples and use them to numerically

89

simulate microstructure evolution in 2.5D64 or even in 3D.65, 66 Such an operation usually requires

90

binarisation (segmentation) of the image data because of the limited applicability of governing

91

equations that are derived from first principles. For example, Navier-Stokes equations cannot be applied

92

directly for a mixed system of solid and fluids.67, 68 Such “hard segmentation” loses a large portion of the

93

information contained in each pixel, which can erase the important heterogeneities that trigger the

94

unstable migration of the dissolution front.69, 70 This loss of detail is especially significant when the

95

imaging resolution is not sufficient to fully resolve the grains and pores in very fine grained materials,

96

such as chalk. Finally, simulating microstructure evolution is a free boundary problem of partial

97

differential equations and often requires prohibitive computational resources, even with binarised

98

geometry.14, 48

ACS Paragon Plus Environment

6

Page 7 of 33

Environmental Science & Technology

99

In this study, we used a model that reflects the submicrometre scale heterogeneities that are present in

100

real samples to study the influence of dissolved CO2 on the microstructure evolution of natural porous

101

materials. We focus on the geochemical aspect of the comparison, i.e. CO2 is considered an aqueous

102

species that modifies the dissolution rate and the apparent solubility of calcite but does not affect the

103

physical properties of fluid. In three model scenarios, we tested the effect of reactive fluids with three

104

different compositions, as they flowed through the simulation domain. We used high resolution X-ray

105

tomography images (25, 50 and 100 nm voxel size) to characterise the microstructure in natural chalk

106

samples. Chalk was chosen as a rapidly dissolving model material. It is the dominant bedrock in

107

northwest Europe, often hosting oil and gas reservoirs that have been considered for GCS. After

108

computing voxel specific porosities from the 32 bit images, the greyscale datasets were implemented

109

directly into numerical simulations. The simulations allowed us to differentiate between the effects of

110

the initial microstructure and solution composition and to investigate their effects on microstructure

111

evolution. The simulated spatial and temporal evolution of rock properties (porosity, permeability and

112

surface area) provided insight into the interplay between breakthrough porosity, wormholing and

113

aqueous CO2 concentrations.

114

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 33

115

Methods and Materials

116

Three chalk pieces from the Hod formation, drilled from different locations in the North Sea Basin, were

117

imaged with X-ray holotomography at 29.49 keV at ID22 at the European Synchrotron Research Facility

118

(ESRF).61 The 3D microstructure was reconstructed from 1999 radiographs at 25, 50 and 100 nm voxel

119

resolution and processed as described by Bruns et al.70 Post reconstruction image processing included

120

ring artefact removal by the method developed by Jha et al.,71 followed by iterative nonlocal means

121

denoising66 and sharpening using the image deconvolution of Wang et al.72 Greyscale intensities of the

122

voxels were then converted to localized porosity values using linear interpolation between the void

123

phase and the carbonate phase, as identified by fitting a Gaussian mixture model.67

124

The greyscale data were then imported into numerical simulations using a previously developed reactor

125

network model.26 The model describes each voxel as a combination of 7 ideal reactors, i.e. each dataset

126

is modelled as a network of continuously stirred tank reactors (CSTRs) interconnected by plug flow

127

reactors (PFRs).40 The volume, permeability and specific surface area of each reactor were assigned

128

according to the voxel size and porosity of neighbouring voxels. Hydraulic pressure drops only in the

129

PFRs. Darcy flow was solved to obtain the flow field. The extent of the dissolution reaction in each

130

reactor was calculated by solving the performance equations (Eqs S3-7) for all reactors in the network

131

simultaneously. The microstructure was evolved in each time step by applying mass balance to update

132

voxel porosity (Eqs S8-9). Equations used in the model are summarised in the SI. A comprehensive

133

sensitivity analysis of the parameters used in the model has been presented in the context of entropy

134

production patterns.73 More information can be found in Yang et al.26, 74 The simulation domains in this

135

study are randomly chosen subvolumes of three tomographic datasets of the chalk pieces. Each

136

subvolume (“sample”) consisted of 1003 32-bit voxels, corresponding to 15.625 µm3, 125.00 µm3 and

137

1000.0 µm3 for the three levels of resolution. A constant flow velocity, 50 µm/s, was imposed at the

ACS Paragon Plus Environment

8

Page 9 of 33

Environmental Science & Technology

138

fluid inlet and outlet. This flowrate is selected so that 1) the residence time of the fluid corresponds to a

139

reactive region comparable to the size of the simulation domain and that 2) Darcy’s law is applicable on

140

the voxel level.

141

For each of the 40 randomly chosen samples (15 samples each from the 25 nm/pixel and 100 nm/pixel

142

datasets and 10 samples from the 50 nm/pixel dataset), we developed 3 model scenarios and simulated

143

the microstructural evolution with and without dissolved CO2. This led to 6 conditions (Table 1) and 240

144

instances of simulation (Table S2). Scenario I represents the ambient conditions where MilliQ water,

145

equilibrated with 1 bar CO2, percolated through the porous material at room temperature. In Scenario

146

II, concentrated CO2 was premixed with seawater at ambient temperature and then injected into a

147

porous material at intermediate CO2 partial pressure. This procedure is similar to the method used in

148

the CarbFix project in Hellisheidi, Iceland.60 Scenario III represents a direct injection into deep

149

formations where the supercritical CO2 mixes with brine under reservoir conditions and then migrates

150

away from the injection wellbore. This scenario is most relevant to the state-of-the art practices for

151

sedimentary basins. The combination of temperature (100 oC) and pressure (250 bar) used in this

152

scenario are representative of that of a North Sea chalk reservoir.75 In the systems without CO2, we used

153

hydrochloric acid to adjust the initial pH of the solutions. Doing so allows the solution pairs in each

154

scenario to have the same pH and thus a comparable calcite dissolution rate at the fluid entrance. The

155

physical differences between the solutions, e.g. the variations of fluid viscosity or solute diffusivity

156

stemming from the elevated aqueous concentration of CO2, were not accounted for in our simulations.

157

We used the rate of entropy production from fluid friction over the entire sample, SP (nJ K-1 s-1), to

158

identify a breakthrough event. SP quantifies the pressure drop in 3D,

159

S P = − ∫∫∫ ( Q ⋅∇ P ) dV T V

,

(1)

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 33

160

where V indicates an integration over the entire sample, Q represents the volumetric flow rate (m3 s-1),

161

P, the pressure (Pa) and T, the temperature (K). When porous media are dissolved by reactive fluids, the

162

evolution of entropy production depends on the initial microstructure, the rate(s) of the dissolution

163

reaction(s) and the flow rate. In addition to the breakthrough porosity, we also recorded the number of

164

pore volumes required before breakthrough (# of PV, i.e., the total volume of percolated fluid divided by

165

the initial pore volume of a sample), whenever an SP inflection could be identified. If a sample dissolved

166

homogeneously, the breakthrough porosity was reported as the difference between its initial porosity

167

and 1. The # of PV is not reported for these cases.

168

Speciation calculations for the three model scenarios (ambient, premixing and direct injection; details in

169

Table 1) were made with PHREEQC (Version 3) using the llnl database.76, 77 For the composition of

170

seawater in Scenario II, we used data from Nordstrom et al.78 The Peng-Robinson equation of state was

171

used to calculate fugacity coefficients.79 For the aqueous species concentrations in Scenarios II and III,

172

we used the B-dot equation, an extension of the Debye-Hückel model.80 We used the rate law for calcite

173

dissolution from Pokrovsky et al.52, 53 to approximate chalk dissolution rates. Surface speciation was

174

determined from aqueous speciation results in a closed, free drift compartment, where the aqueous Ca

175

concentration served as the master variable for determining both pH and the saturation index of calcite.

176

The rates were then determined and compiled in Figure S1.

177

Table 1. The model scenarios.78

ACS Paragon Plus Environment

10

Page 11 of 33

Environmental Science & Technology

178 179

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 33

180

Results and Discussion

181

Figure 2 shows two patterns of SP evolution: a wormhole pattern, where channels form in the porous

182

medium, and a homogeneous pattern, where the whole domain dissolves evenly. During wormhole

183

growth, SP drops rapidly one or more times, representing necking in regions with high porosity, i.e.,

184

small breakthrough of fluid. Necking events are represented by local SP’ minima, the first derivative of SP

185

with respect to time, and can thus be identified by inflection points (Figure 2a). Because of the structural

186

heterogeneity of natural porous media, the critical porosity at which necking occurs, as well as the

187

frequency of its occurrence, differs for each sample even with the same solution composition. In this

188

study, we refer to the porosity at the global minimum of SP’ as the breakthrough porosity (ϕc, green

189

arrow in Fig. 2a). This choice does not imply the uniqueness of the necking event during percolation.

190

Figure 2d shows cross sections of a wormhole sample before and after breakthrough and the

191

corresponding spatial patterns of entropy generation. Formation of flow paths with higher permeability

192

lead to bypassing fluid from the less porous regions. After breakthrough, the macroscopic entropy

193

production is solely determined by the significant pressure drop at the necks of pores (e.g., ϕa = 0.50).

ACS Paragon Plus Environment

12

Page 13 of 33

Environmental Science & Technology

194

195

Figure 2.

196

developing porous structures. Evolution of SP in isothermal (a) wormhole and (b) homogeneously

197

dissolving systems. SP’ and SP’’ are first and second derivatives of SP with respect to percolation time. We

198

used the inflection point at the global minimum of SP’ to represent the occurrence of breakthrough (e.g.,

199

ϕc in a). When ϕc > 1, no inflection exists and the medium dissolves homogeneously. (c to e) Cross

200

sections of porosity (ϕ) and entropy production rate (T⋅SP) from 3D simulations using GeoID 1832 (Table

201

S2). Fluid flows from left to right. (c) The initial geometry. (d) A wormhole system following on from

202

(c), before (ϕa = 0.30) and after (ϕa = 0.50) breakthrough. (e) A homogeneous dissolution system,

203 204 205

following on from (c), at the same reaction progression (ϕb = 0.30 and ϕb = 0.50).

Typical temporal and spatial patterns of entropy generation, Sp, from fluid friction in

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 33

206

Figure 2b shows a temporal pattern of SP where no abrupt pressure drop can be observed before the

207

depletion of solid material. This pattern indicates a homogeneous dissolution where SP’ increases

208

monotonically. The spatial patterns in Figure 2e reveal that the morphology change is not strongly

209

dependent on flow direction, i.e. the entire sample behaves as a homogeneous medium, despite its

210

geometric complexity, and in different regions, porosity decreases at similar rates. Because of the

211

absence of a favoured flow path, no significant fluid focusing occurs.

212

Figure 3 shows results of 240 simulations. In 74% of all the simulations, compared to solutions with

213

hydrochloric acid, dissolved CO2 led to greater breakthrough porosity, but the conditions under which

214

CO2 entered the system had an influence. For atmospheric CO2 pressure (Scenario I), the breakthrough

215

porosity was higher than the HCl control in 83% of the cases. In the premixing scenario (II), 78% of the

216

breakthrough with dissolved CO2 took place at a greater porosity. The percentage decreased to 63% in

217

the direct inject scenario (III), where CO2 mixed with saline water under reservoir conditions. In all

218

systems, fluids with CO2 required fewer pore volumes (# of PV) to breakthrough than the CO2 free

219

systems. This difference in # of PV was between one and two orders of magnitude and depended only

220

weakly on the initial porosity (ϕ0).

221

The homogeneous boundaries (dashed lines in Fig. 3) are lines marking the difference between the

222

initial porosity and a porosity of 1 (complete dissolution). They were determined from mass balance

223

and decreased linearly with ϕ0. Symbols that fall on these lines represent cases where the structures

224

dissolved homogeneously, i.e. where the initial geometric complexity of the simulation domain does not

225

affect the position of the symbols. In these cases, the system evolution was not sensitive to the initial

226

heterogeneities so could be considered more stable and predictable. The # of PV is inversely

227

proportional to the initial porosity because less fluid is needed for breakthrough in a more porous

228

material.

ACS Paragon Plus Environment

14

Page 15 of 33

Environmental Science & Technology

229

In Figure 3, it is interesting to note that no significant resolution dependence is observed. The resolution

230

of tomography typically affects the characterisation of porosity and surface area.61 However, our model

231

uses greyscale tomographic data. The greyscale images preserve the local material density information

232

because the X-ray absorption is averaged in each voxel, even when the resolution is low. Consequently,

233

our simulations are not subject to the resolution dependence of porosity characterisation.70 Imaging

234

resolution does affect surface area characterisation. However, the resolution difference (e.g., 25 nm vs

235

100 nm) does not yield significantly different specific surface area values for chalk,70 especially within

236

our fairly small simulation domains. On a similar note, our simulations have not accounted for either

237

chemical heterogeneity or any variation of effective diffusivity within a domain. These limitations may

238

contribute to the uncertainty of simulation results, but not to a significant extent given the small domain

239

sizes. Another limitation of the current model is that it does not account for mineral precipitation, which

240

may significantly affect breakthrough behaviours. A mathematical scheme considering both mineral

241

dissolution and precipitation in evolving microstructures in imposed flow fields is a focus of our future

242

works.

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 33

243 244

Figure 3. Breakthrough porosity and the corresponding number of pore volumes of fluid (# of PV, red

245

symbols). ∆ϕc (blue) represents the difference between the breakthrough porosity (ϕc) and the initial

246

porosity (ϕ0) of a sample. All samples are cubes and consist of 1 million voxels. The open symbols show

ACS Paragon Plus Environment

16

Page 17 of 33

Environmental Science & Technology

247

results with CO2 and the filled symbols represent CO2 free systems. The shape of the symbols represents

248

the voxel size: square: 100 nm; triangle: 50 nm and circle: 25 nm. The grey dashed line near the top of

249

each figure marks the homogeneous boundary, which corresponds to cases where no inflection of SP is

250

observed before solid depletion (i.e., the sample dissolves homogeneously throughout the percolation).

251

For points on this boundary, the number of PVs is not shown because of the difficulty in defining

252

breakthrough. (a) Ambient conditions, CO2 in ultrapure water (Scenario I). (b) Premixing conditions, CO2

253

dissolved in seawater (Scenario II). (c) Direct injection conditions, CO2 dissolved in saline water under

254

reservoir conditions (Scenario III).

255

256

Both the differences in the breakthrough porosity (∆ϕc) and the number of pore volumes (# of PV) can

257

be explained by the observed dissolution patterns. Given the same initial sample microstructure, the

258

dissolution pattern is determined by distribution of reactants in the medium. Figure 4 shows the

259

decrease of fluid reactivity (measured by reaction rate and pH) as a function of cumulative surface (CS),

260

‫׬‬଴ ೝ೐ೞ ܵܵ‫ݐ݀ ∙ ܣ‬, in the 3 model scenarios. ߬௥௘௦ represents the residence time of the fluid in the porous

261

medium (s), and SSA, the specific surface area (m-1). The physical significance of CS is the overall surface

262

area that the fluid “sees” as it travels through a sample. Therefore, the curves represent the chemical

263

history of an isolated fluid parcel, an imaginary unit of flowing fluid, travelling along a streamline. Here

264

we assume that the parcel only interacts with the solid along the streamline and does not exchange

265

mass with other fluid parcels. As solid dissolves, pH and the saturation index (SI) in the fluid parcel

266

increase. Both effects slow the reaction. The reaction front along the streamline is the position at which

267

the reactivity of the fluid parcel drops to a very small value (e.g. 10-6 mol/m2/s in Figure 4). This

268

definition suggests the equivalent effect of the residence time and that of the surface area in

269

determining the reaction front.

270

In general, a dissolution pattern reflects a demand-supply relation for cumulative surface. When the

271

demand is high and the supply is low, a porous medium dissolves homogeneously. In contrast,



ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 33

272

wormholes form when the demand is low and the supply is high. The demand is determined by the

273

chemical properties of the reactive fluid, while the supply is affected by both the available geometric

274

surface area and the fluid residence time. Although specific surface area (SSA) depends exclusively on

275

the sample microstructure, the residence time (߬௥௘௦ ) depends on both the microstructure and the fluid

276

flow rate. Consequently, different dissolution patterns can coexist in a system. This is because the

277

patterns reflect the competition between the reaction front and the cumulative surface within a region

278

of interest (ROI). Given the same solution conditions, one observes a homogeneous dissolution pattern

279

when the fluid leaves the ROI without depleting reactivity. This is typical for fluids with a high apparent

280

mineral solubility, i.e. a high demand for CS. The reaction front of such a fluid appears only when

281

residence time is sufficiently long or when the specific surface area is large. If, in contrast, the reaction

282

front is reached well before the fluid leaves the ROI, wormholes appear. The residence time can be

283

increased by using a lower flow rate or by making the sample larger (i.e., a large ROI). For example,

284

Figure S2 shows that with an ROI 5 times longer in the flow direction, flow rate change alone can affect

285

the observed dissolution pattern.

286

Given sample size (i.e., ROI) and flow rate, the CS supply of a sample is fixed, and distinct dissolution

287

patterns are most likely to occur when the ROI contains only one of the two reaction fronts. This is

288

demonstrated by the green shaded areas in Figure 4. These areas are left bounded by the dissolution

289

front of systems without CO2 and right bounded by those with dissolved CO2. This difference between

290

the positions of the dissolution fronts, defined by the apparent solubility of the solutions, is the most

291

important geochemical difference between the solutions in each pair. It provides a straightforward

292

explanation for the impacts of different infiltrating fluids on dissolution patterns, i.e. why the same

293

domain dissolves differently with and without CO2. If the CS of a sample falls to the left of the shaded

294

areas, the supply is always less than the demand and the sample dissolves homogeneously. If the CS falls

295

to the right of the shaded areas, the supply is always greater than the demand and therefore wormholes

ACS Paragon Plus Environment

18

Page 19 of 33

Environmental Science & Technology

296

appear regardless of whether CO2 is present. The fluid velocity in this study (50 µm/s) produced

297

distributions of CS that overlapped with these shaded areas. Therefore, the widths of the shaded

298

regimes explain the percentages of patterns observed in the model scenarios. In the ambient scenario

299

(I), the reaction fronts with and without CO2 are farthest from each other (the shaded area is widest in

300

Figure 4a). Without CO2, the ROI is larger than the reactive volume and the wormhole pattern appears.

301

With dissolved CO2, the opposite is true. As a result, in 33 of the 40 comparisons (83%), breakthrough

302

porosity was higher for systems with CO2 than without. In contrast, the two reaction fronts were very

303

similar in the direct injection scenario (III, Figure 4c). Only 63% of the simulated cases (25/40) showed

304

that dissolved CO2 increased the breakthrough porosity compared to the cases without CO2. The

305

distance between the two reaction fronts is intermediate in the premixing scenario (Figure 4b) and so is

306

the percentage (78%) of increased ϕc in the presence of CO2.

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 33

307 308

Figure 4. Decrease of calcite dissolution rate (solid lines) with cumulative surface area (‫׬‬଴ ೝ೐ೞ ܵܵ‫) ݐ݀ ∙ ܣ‬

309

in a free drift system. The integral represents the overall surface area that an isolated parcel of reactive

310

fluid experiences before leaving the sample. If a sample has a greater cumulative surface, it is more likely

311

to show wormhole patterns (and vice versa). Shaded regimes are bounded by the dissolution fronts of the

312

two fluid compositions (with vs. without CO2). Also shown is the evolution of pH (dashed lines). (a)



ACS Paragon Plus Environment

20

Page 21 of 33

Environmental Science & Technology

313

Ambient conditions (Scenario I). (b) Premixing conditions (Scenario II). (c) Direct injection conditions

314 315

(Scenario III).

316

A wormhole pattern lowers the breakthrough porosity because of fluid focusing. Similarly, more

317

effective use of geometric surface on the pore scale for the dissolution reaction explains why systems

318

with dissolved CO2 need less fluid to breakthrough. Both phenomena are demonstrated in Figure 5,

319

where a case study of microstructural evolution in Scenario I is presented. The initial geometry for both

320

simulations is shown in Figure S3. The two systems start with the same percolative entropy production

321

rate because of their identical initial geometry. The isothermal SP for the system with CO2 decreases

322

gradually, indicating a homogeneous dissolution pattern. In contrast, the system without CO2 showed a

323

sharp decrease in SP near ϕ = 0.4, suggesting significant necking of pores, typical of wormhole growth. SP

324

without CO2 dropped below that with CO2 after an overall porosity of 0.45, which suggests bypassing

325

flow through a fully developed wormhole. The pressure drop after this point is determined by the shape

326

of the channel rather than the amount of solid in the pores. Also shown in Figure 5a is the evolution of

327

reactive surface area (RSA). Although the two systems start with exactly the same geometric surface

328

area, their RSA differs by almost one order of magnitude because of the rapid depletion of fluid

329

reactivity in the CO2-free system. The initial increase in RSA is characteristic of a system with infiltration

330

instability. The decrease of RSA is caused by the depletion of solid material. Although RSA in both cases

331

evolve with similar trends, the absolute difference increases with time, contributing to the difference in

332

# of PV at breakthrough.

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 33

333 334

Figure 5. A case study of CO2 effect on microstructure evolution (simulations uid-3813 vs. uid-96443).

335

(a) Evolution of isothermal entropy generation rate (T⋅SP) and reactive surface area (RSA) with overall

336

porosity as an indicator of reaction progress. (b) Distribution of residence time multiplied by specific

337

reactive surface area (SRSA) in the context of calcite dissolution rates in Scenario I. The same shaded

338

regime is shown here, as in Figure 4a. The blue bars show the density function of fluid with CO2 and the

339

red bars, without. Also shown are microstructures, cross sections (10 ൈ 10 µm ) of pH and of reactive

340 341

surface distribution for percolation (c) with and (d) without CO2.

2

ACS Paragon Plus Environment

22

Page 23 of 33

Environmental Science & Technology

342

Figures 5b to d compare the evolution of microstructure in the presence and absence of CO2. In Figure

343

5b the CS for the complete flow field is approximated by the product of specific reactive surface area

344

(SRSA, m2/m3) and residence time distribution (RTD, giving the probability density of fluid parcels with

345

residence time, τ). The same shaded regime in Figure 4a is shown. Both distributions fall entirely into

346

the regime between the two reaction fronts. As a result, the dissolution front with CO2 cannot be

347

observed in the simulation domain because it is beyond the sample size (blue bars). In contrast, the

348

dissolution front without CO2 is fully contained in the domain. This is because the sample provides

349

approximately 10 times more cumulative surface required to reach the front (the maximum distance

350

from the injection point to where the dissolution rate drops to zero, the red bars). The development of

351

the distribution at different overall porosities (ϕ = 0.3, 0.4 and 0.5) can be decomposed into two parts.

352

The morphing of their shapes reflects the redistribution of fluid as the micropore evolves. The lower

353

bound of the distribution represents the flow pathways with minimum flow resistance. As the sample

354

dissolves, the mean residence time shifts to the left, indicating the tendency of the fluid to focus in the

355

more permeable pathways. Fluid focusing during wormhole development is directly reflected in the

356

standard deviations. With CO2, the standard deviation of the cumulative surface increases from 1.4 ×

357

104 s/m for ϕ = 0.3 to 1.6 × 104 s/m for ϕ = 0.5, suggesting more uniform fluid distribution. In contrast,

358

without CO2, cumulative surface area decreased from 1.0 × 104 s/m to 6.5 × 103 s/m for the same

359

overall porosity. Therefore, much less geometric surface in the pores is in contact with reactive solution

360

during wormhole growth (e.g., spatial distribution of RSA in Figures 5c and d). The development of

361

reactive surface determines the horizontal shifting of the distribution. With dissolved CO2, the

362

maximum of RSA appears after an overall porosity of 0.5, thus increasing the mean of the distribution. In

363

contrast, the overall RSA in the CO2 free system decreases after an overall porosity of 0.35, shifting the

364

entire distribution of cumulative surface leftward.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 33

365

The spatial distributions of pH in Figure 5c and 5d provide a pore scale interpretation for the stabilised

366

migration of dissolution front on the macroscale. pH is an indicator of calcite dissolution rate. In this

367

case, both solutions started with the same rate at the fluid entrance (on the left of the cross sections,

368

corresponding to a pH of 3.91). Figure 5d shows that in a CO2-free system this rate drops much faster

369

than the rate in a CO2-buffered system, i.e. the reaction rate in the absence of CO2 is very sensitive to

370

the extent of reaction and can drop from the initial value to zero within a few voxels. This sensitivity is

371

an important in forming the loop of positive feedback leading to infiltration instability. In addition, the

372

buffering effect of dissolved CO2 also increases the area of the geometric surface that has access to

373

reactive fluid (in Fig. 5d most geometric surface is in contact with highly saturated fluid with pH greater

374

than 7). Overall, this combined effect of rate sensitivity and surface area in the presence of CO2 results

375

in breakthrough with significantly fewer PVs in all cases, even though the breakthrough porosity has

376

been increased in 74% of the cases.

377

The CO2 effect on breakthrough porosity has several implications. Mineral dissolution serves as a trigger

378

for many water-rock interactions. In GCS, dissolution provides cations for carbon mineralization and at

379

the same time, it initiates microstructure evolution that determines the mechanical strength of the

380

formation. Greater instability in the dissolution front, often associated with wormhole development, is

381

favoured when geomechanical stability is desirable. This is because lower breakthrough porosity leads

382

to flow structures that dissipate injected fluid effectively, without removing a substantial portion of solid

383

that serves as mechanical support. Our results show that dissolved CO2 has adverse effects as it

384

stabilises the migration of dissolution front. This stabilisation is reflected in an increased demand for the

385

cumulative surface in order to form wormholes. When the size of a sample is fixed, so is the supply of

386

cumulative surface. The increased demand therefore makes fluid penetration through wormholing less

387

likely. Also, the wide spread of fluid reactivity during homogeneous dissolution may increase the chance

388

of contaminant mobilisation near injection wellbores. This likelihood is related to an increased region of

ACS Paragon Plus Environment

24

Page 25 of 33

Environmental Science & Technology

389

disintegrating carbonates during acidic fluid injection. In addition to sedimentary basins, the other

390

geologic settings suitable for GCS may also contain carbonates, especially in the sealing structures. Our

391

results suggest that the fractures cemented by carbonates are subject to geochemical alterations and

392

may serve as active flow pathways for CO2 leakage. The presence of CO2 may enhance the opening of

393

these pathways and dramatically shorten the time for fluid breakthrough. Furthermore, we have shown

394

that the tomography-based greyscale approach is a powerful tool to assemble geochemical and

395

microstructural knowledge in studying evolving systems. Similar approaches may be useful in other GCS

396

scenarios and in subsurface applications.

397

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 33

398

ASSOCIATED CONTENT

399

Supporting Information. Reactor network model details. Calcite dissolution rate based on 3 published

400

rate laws, for the 3 model scenarios. Cross sections of simulations showing the coexistence of different

401

dissolution patterns within one sample at various flow rates. The initial geometry upon which the case

402

study presented in Figure 5 was based. Tabulated results for all simulations. This material is available

403

free of charge via the Internet at http://pubs.acs.org.

404

AUTHOR INFORMATION

405

Corresponding Author

406

* [email protected]

407

Author Contributions

408

YY designed the research and conducted the simulations. SB processed the tomography data.

409

HOS and SLSS advised during the research. All authors contributed to writing the paper.

410

ACKNOWLEDGMENTS

411

We thank Heikki Suhonen at the ID22 beamline at ESRF (The European Synchrotron Radiation Facility)

412

for technical support. We are grateful to F. Engstrøm from Maersk Oil and Gas A/S for providing the

413

sample. We deeply appreciates K. Dideriksen and three anonymous reviewers’ careful review of the

414

manuscript and for the thoughtful comments. Funding for this project was provided by the Innovation

415

Fund Denmark, through the CINEMA project, the Innovation Fund Denmark and Maersk Oil and Gas A/S,

416

through the P3 project as well as the European Commission, Horizon 2020 Research and Innovation

417

Programme under the Marie Sklodowska-Curie Grant Agreement No 653241 for the Postdoctoral

418

Fellowship to YY. We thank the Danish Council for Independent Research for support for synchrotron

419

experiments through DANSCATT.

ACS Paragon Plus Environment

26

Page 27 of 33

420

Environmental Science & Technology

TOC

421 422

ACS Paragon Plus Environment

27

Environmental Science & Technology

Page 28 of 33

423

REFERENCES

424 425

1. Bourg, I. C.; Beckingham, L. E.; DePaolo, D. J., The Nanoscale Basis of CO2 Trapping for Geologic Storage. Environmental Science & Technology 2015, 49, (17), 10265-10284.

426 427

2. DePaolo, D. J.; Cole, D. R., Geochemistry of Geologic Carbon Sequestration: An Overview. Reviews in Mineralogy and Geochemistry 2013, 77, (1), 1-14.

428 429 430

3. Alfredsson, H. A.; Oelkers, E. H.; Hardarsson, B. S.; Franzson, H.; Gunnlaugsson, E.; Gislason, S. R., The geology and water chemistry of the Hellisheidi, SW-Iceland carbon storage site. International Journal of Greenhouse Gas Control 2013, 12, 399-418.

431 432 433 434

4. Gislason, S. R.; Wolff-Boenisch, D.; Stefansson, A.; Oelkers, E. H.; Gunnlaugsson, E.; Sigurdardottir, H.; Sigfusson, B.; Broecker, W. S.; Matter, J. M.; Stute, M.; Axelsson, G.; Fridriksson, T., Mineral sequestration of carbon dioxide in basalt: A pre-injection overview of the CarbFix project. International Journal of Greenhouse Gas Control 2010, 4, (3), 537-545.

435 436

5. Bentham, M.; Kirby, M., CO2 storage in saline aquifers. Oil & Gas Science and Technology 2005, 60, (3), 559-567.

437 438

6. Ellis, B. R.; Peters, C. A., 3D Mapping of calcite and a demonstration of its relevance to permeability evolution in reactive fractures. Advances in Water Resources 2016, 95, 246-253.

439 440 441

7. Ellis, B. R.; Fitts, J. P.; Bromhal, G. S.; McIntyre, D. L.; Tappero, R.; Peters, C. A., Dissolution-driven permeability reduction of a fractured carbonate caprock. Environmental Engineering Science 2013, 30, (4), 187-193.

442 443 444

8. Deng, H.; Ellis, B. R.; Peters, C. A.; Fitts, J. P.; Crandall, D.; Bromhal, G. S., Modifications of Carbonate Fracture Hydrodynamic Properties by CO2-Acidified Brine Flow. Energy & Fuels 2013, 27, (8), 4221-4231.

445 446 447

9. Emmanuel, S.; Anovitz, L. M.; Day-Stirrat, R. J., Effects of Coupled Chemo-Mechanical Processes on the Evolution of Pore-Size Distributions in Geological Media. Reviews in Mineralogy and Geochemistry 2015, 80, (1), 45-60.

448

10.

449 450

11. Szymczak, P.; Ladd, A. J., Reactive-infiltration instabilities in rocks. Part 2. Dissolution of a porous matrix. Journal of Fluid Mechanics 2014, 738, 591-630.

451 452

12. Szymczak, P.; Ladd, A. J. C., Reactive-infiltration instabilities in rocks. Fracture dissolution. Journal of Fluid Mechanics 2012, 702, 239-264.

453 454

13. Ortoleva, P.; Chadam, J.; Merino, E.; Sen, A., Geochemical self-organization II: the reactiveinfiltration instability. Am. J. Sci 1987, 287, 1008-1040.

455 456

14. Chadam, J.; Hoff, D.; Merino, E.; Ortoleva, P.; Sen, A., Reactive Infiltration Instabilities. IMA Journal of Applied Mathematics 1986, 36, (3), 207-221.

Hubbert, M. K.; Willis, D. G., Mechanics of hydraulic fracturing. 1972.

ACS Paragon Plus Environment

28

Page 29 of 33

Environmental Science & Technology

457 458 459

15. Steefel, C. I.; Lasaga, A. C., A coupled model for transport of multiple chemical-species and kinetic precipitation dissolution reactions with application to reactive flow in single-phase hydrothermal systems. American Journal of Science 1994, (294), 529-592.

460 461

16. Steefel, C. I.; Lasaga, A. C. In Evolution of dissolution patterns: Permeability change due to coupled flow and reaction, ACS Symposium Series, 1990; Oxford University Press: 1990; pp 212-225.

462 463

17. Hinch, E.; Bhatt, B., Stability of an acid front moving through porous rock. Journal of Fluid Mechanics 1990, 212, 279-288.

464 465 466

18. Hao, Y.; Smith, M.; Sholokhova, Y.; Carroll, S., CO2-induced dissolution of low permeability carbonates. Part II: Numerical modeling of experiments. Advances in Water Resources 2013, 62, Part C, 388-408.

467 468

19. Cao, P.; Karpyn, Z. T.; Li, L., The role of host rock properties in determining potential CO2 migration pathways. International Journal of Greenhouse Gas Control 2016, 45, 18-26.

469 470

20. Fredd, C. N.; Fogler, H. S., Influence of transport and reaction on wormhole formation in porous media. AIChE Journal 1998, 44, (9), 1933-1949.

471 472 473

21. Golfier, F.; Zarcone, C.; Bazin, B.; Lenormand, R.; Lasseux, D.; Quintard, M., On the ability of a Darcy-scale model to capture wormhole formation during the dissolution of a porous medium. Journal of Fluid Mechanics 2002, 457, 213-254.

474 475 476

22. Glasbergen, G.; Kalia, N.; Talbot, M. S. In The Optimum Injection Rate for Wormhole Propagation: Myth or Reality?, 8th European Formation Damage Conference, 2009; Society of Petroleum Engineers: 2009.

477 478

23. Fredd, C. N.; Scott Fogler, H., The kinetics of calcite dissolution in acetic acid solutions. Chemical Engineering Science 1998, 53, (22), 3863-3874.

479 480

24. Thompson, K. E.; Fogler, H. S., Modeling flow in disordered packed beds from pore‐scale fluid mechanics. AIChE Journal 1997, 43, (6), 1377-1389.

481 482 483

25. Pereira Nunes, J. P.; Bijeljic, B.; Blunt, M. J., Time-of-Flight Distributions and Breakthrough Curves in Heterogeneous Porous Media Using a Pore-Scale Streamline Tracing Algorithm. Transport in Porous Media 2015, 109, (2), 317-336.

484 485 486

26. Yang, Y.; Bruns, S.; Stipp, S.; Sørensen, H., Reactive infiltration instability amplifies the difference between geometric and reactive surface areas in natural porous materials. in review at Environmental Science & Technology (arXiv:1704.01060) 2017.

487 488 489

27. Brunet, J.-P. L.; Li, L.; Karpyn, Z. T.; Kutchko, B. G.; Strazisar, B.; Bromhal, G., Dynamic Evolution of Cement Composition and Transport Properties under Conditions Relevant to Geological Carbon Sequestration. Energy & Fuels 2013, 27, (8), 4208-4220.

490 491

28. Gale, J.; Hendriks, C.; Turkenberg, W.; Huerta, N. J.; Bryant, S. L.; Strazisar, B. R.; Hesse, M., 10th International Conference on Greenhouse Gas Control TechnologiesDynamic alteration along a fractured

ACS Paragon Plus Environment

29

Environmental Science & Technology

Page 30 of 33

492 493

cement/cement interface: Implications for long term leakage risk along a well with an annulus defect. Energy Procedia 2011, 4, 5398-5405.

494 495 496

29. Huerta, N. J.; Hesse, M. A.; Bryant, S. L.; Strazisar, B. R.; Lopano, C. L., Experimental Evidence for Self-Limiting Reactive Flow through a Fractured Cement Core: Implications for Time-Dependent Wellbore Leakage. Environmental Science & Technology 2013, 47, (1), 269-275.

497 498 499

30. Zhang, L.; Dzombak, D. A.; Nakles, D. V.; Brunet, J.-P. L.; Li, L., Reactive Transport Modeling of Interactions between Acid Gas (CO2 + H2S) and Pozzolan-Amended Wellbore Cement under Geologic Carbon Sequestration Conditions. Energy & Fuels 2013, 27, (11), 6921-6937.

500 501 502

31. Menke, H. P.; Bijeljic, B.; Andrew, M. G.; Blunt, M. J., Dynamic Three-Dimensional Pore-Scale Imaging of Reaction in a Carbonate at Reservoir Conditions. Environmental Science & Technology 2015, 49, (7), 4407-4414.

503 504 505

32. Noiriel, C., Resolving Time-dependent Evolution of Pore-Scale Structure, Permeability and Reactivity using X-ray Microtomography. Reviews in Mineralogy and Geochemistry 2015, 80, (1), 247285.

506 507 508

33. Noiriel, C.; Luquot, L.; Madé, B.; Raimbault, L.; Gouze, P.; van der Lee, J., Changes in reactive surface area during limestone dissolution: An experimental and modelling study. Chemical Geology 2009, 265, (1–2), 160-170.

509 510

34. Noiriel, C.; Madé, B.; Gouze, P., Impact of coating development on the hydraulic and transport properties in argillaceous limestone fracture. Water Resources Research 2007, 43, (9).

511 512

35. Noiriel, C.; Gouze, P.; Bernard, D., Investigation of porosity and permeability effects from microstructure changes during limestone dissolution. Geophysical Research Letters 2004, 31, (24).

513 514 515

36. Jha, B.; Juanes, R., Coupled multiphase flow and poromechanics: A computational model of pore pressure effects on fault slip and earthquake triggering. Water Resources Research 2014, 50, (5), 37763808.

516 517 518

37. Nicolaides, C.; Jha, B.; Cueto-Felgueroso, L.; Juanes, R., Impact of viscous fingering and permeability heterogeneity on fluid mixing in porous media. Water Resources Research 2015, 51, (4), 2634-2647.

519 520

38. Matin, R.; Misztal, M. K.; Hernandez-Garcia, A.; Mathiesen, J., Evaluation of the Finite Element Lattice Boltzmann Method for Binary Fluid Flows. arXiv:1608.01906 2016.

521 522 523

39. Carballido-Landeira, J.; Trevelyan, P. M. J.; Almarcha, C.; De Wit, A., Mixed-mode instability of a miscible interface due to coupling between Rayleigh-Taylor and double-diffusive convective modes. Physics of Fluids 2013, 25, (2), 024107.

524

40.

525 526 527

41. Yang, Y.; Min, Y.; Lococo, J.; Jun, Y.-S., Effects of Al/Si ordering on feldspar dissolution: Part I. Crystallographic control on the stoichiometry of dissolution reaction. Geochimica et Cosmochimica Acta 2014, 126, 574-594.

Levenspiel, O., Chemical Reaction Engineering. Wiley & Sons, Inc.: 1999.

ACS Paragon Plus Environment

30

Page 31 of 33

Environmental Science & Technology

528 529

42. Yang, Y.; Min, Y.; Jun, Y.-S., Effects of Al/Si ordering on feldspar dissolution: Part II. The pH dependence of plagioclases’ dissolution rates. Geochimica et Cosmochimica Acta 2014, 126, 595-613.

530 531

43. Gouze, P.; Luquot, L., X-ray microtomography characterization of porosity, permeability and reactive surface changes during dissolution. Journal of Contaminant Hydrology 2011, 120–121, 45-55.

532 533 534

44. Zachara, J.; Brantley, S.; Chorover, J.; Ewing, R.; Kerisit, S.; Liu, C.; Perfect, E.; Rother, G.; Stack, A. G., Internal Domains of Natural Porous Media Revealed: Critical Locations for Transport, Storage, and Chemical Reaction. Environmental Science & Technology 2016, 50, (6), 2811-2829.

535 536 537

45. Lu, J.; Kharaka, Y. K.; Thordsen, J. J.; Horita, J.; Karamalidis, A.; Griffith, C.; Hakala, J. A.; Ambats, G.; Cole, D. R.; Phelps, T. J., CO 2–rock–brine interactions in Lower Tuscaloosa Formation at Cranfield CO 2 sequestration site, Mississippi, USA. Chemical Geology 2012, 291, 269-277.

538 539 540 541

46. Zheng, L.; Apps, J. A.; Spycher, N.; Birkholzer, J. T.; Kharaka, Y. K.; Thordsen, J.; Beers, S. R.; Herkelrath, W. N.; Kakouros, E.; Trautz, R. C., Geochemical modeling of changes in shallow groundwater chemistry observed during the MSU-ZERT CO 2 injection experiment. International Journal of Greenhouse Gas Control 2012, 7, 202-217.

542 543 544

47. Upadhyay, V. K.; Szymczak, P.; Ladd, A. J., Initial conditions or emergence: What determines dissolution patterns in rough fractures? Journal of Geophysical Research: Solid Earth 2015, 120, (9), 6102-6121.

545

48.

546 547

49. Ortoleva, P.; Merino, E.; Moore, C.; Chadam, J., Geochemical self-organization I: reactiontransport feedbacks and modeling approach. Am. J. Sci 1987, 287, (10), 979-1007.

548 549

50. Kalisch, H.; Mitrovic, D.; Nordbotten, J. M., Rayleigh–Taylor instability of immiscible fluids in porous media. Continuum Mechanics and Thermodynamics 2016, 28, (3), 721-731.

550 551 552

51. Thomas, C.; Loodts, V.; Rongy, L.; De Wit, A., Convective dissolution of CO2 in reactive alkaline solutions: Active role of spectator ions. International Journal of Greenhouse Gas Control 2016, 53, 230242.

553 554 555

52. Pokrovsky, O. S.; Golubev, S. V.; Schott, J.; Castillo, A., Calcite, dolomite and magnesite dissolution kinetics in aqueous solutions at acid to circumneutral pH, 25 to 150 C and 1 to 55 atm pCO 2: new constraints on CO 2 sequestration in sedimentary basins. Chemical Geology 2009, 265, (1), 20-32.

556 557

53. Pokrovsky, O. S.; Golubev, S. V.; Schott, J., Dissolution kinetics of calcite, dolomite and magnesite at 25 C and 0 to 50 atm pCO 2. Chemical Geology 2005, 217, (3), 239-255.

558 559 560

54. Deng, H.; Fitts, J. P.; Crandall, D.; McIntyre, D.; Peters, C. A., Alterations of Fractures in Carbonate Rocks by CO2-Acidified Brines. Environmental Science & Technology 2015, 49, (16), 1022610234.

561 562

55. Fitts, J. P.; Peters, C. A., Caprock Fracture Dissolution and CO2 Leakage. Reviews in Mineralogy and Geochemistry 2013, 77, (1), 459-479.

Ortoleva, P. J., Geochemical Self-Organization. Oxford University Press: Clarendon Press: 1994.

ACS Paragon Plus Environment

31

Environmental Science & Technology

Page 32 of 33

563 564

56. Shariatipour, S. M.; Mackay, E. J.; Pickup, G. E., An engineering solution for CO2 injection in saline aquifers. International Journal of Greenhouse Gas Control 2016, 53, 98-105.

565 566 567

57. Emami-Meybodi, H.; Hassanzadeh, H.; Green, C. P.; Ennis-King, J., Convective dissolution of CO2 in saline aquifers: Progress in modeling and experiments. International Journal of Greenhouse Gas Control 2015, 40, 238-266.

568 569

58. Kumar, A.; Noh, M. H.; Ozah, R. C.; Pope, G. A.; Bryant, S. L.; Sepehrnoori, K.; Lake, L. W., Reservoir simulation of CO2 storage in aquifers. SPE Journal 2005, 10, (03), 336-348.

570 571 572

59. Burton, M.; Bryant, S. L., Eliminating buoyant migration of sequestered CO2 through surface dissolution: implementation costs and technical challenges. SPE Reservoir Evaluation & Engineering 2009, 12, (03), 399-407.

573 574 575

60. Matter, J. M.; Stute, M.; Snæbjörnsdottir, S. Ó.; Oelkers, E. H.; Gislason, S. R.; Aradottir, E. S.; Sigfusson, B.; Gunnarsson, I.; Sigurdardottir, H.; Gunnlaugsson, E., Rapid carbon mineralization for permanent disposal of anthropogenic carbon dioxide emissions. Science 2016, 352, (6291), 1312-1314.

576 577 578

61. Muter, D.; Sorensen, H.; Jha, D.; Harti, R.; Dalby, K.; Suhonen, H.; Feidenhans'l, R.; Engstrom, F.; Stipp, S., Resolution dependence of petrophysical parameters derived from X-ray tomography of chalk. Applied Physics Letters 2014, 105, (4), 043108.

579 580 581

62. Gooya, R.; Bruns, S.; Müter, D.; Moaddel, A.; Harti, R.; Stipp, S.; Sørensen, H., Effect of tomography resolution on the calculated microscopic properties of porous materials: Comparison of sandstone and carbonate rocks. Applied Physics Letters 2016, 109, (10), 104102.

582 583 584

63. Müter, D.; Pedersen, S.; Sørensen, H. O.; Feidenhans'l, R.; Stipp, S. L. S., Improved segmentation of X-ray tomography data from porous rocks using a dual filtering approach. Computers & Geosciences 2012, 49, 131-139.

585 586 587

64. Deng, H.; Molins, S.; Steefel, C.; DePaolo, D.; Voltolini, M.; Yang, L.; Ajo-Franklin, J., A 2.5 D Reactive Transport Model for Fracture Alteration Simulation. Environmental Science & Technology 2016, 50, (14), 7564-7571.

588 589

65. Blunt, M. J.; Bijeljic, B.; Dong, H.; Gharbi, O.; Iglauer, S.; Mostaghimi, P.; Paluszny, A.; Pentland, C., Pore-scale imaging and modelling. Advances in Water Resources 2013, 51, 197-216.

590 591

66. Steefel, C. I.; Beckingham, L. E.; Landrot, G., Micro-Continuum Approaches for Modeling PoreScale Geochemical Processes. Reviews in Mineralogy and Geochemistry 2015, 80, (1), 217-246.

592 593

67. Molins, S., Reactive Interfaces in Direct Numerical Simulation of Pore-Scale Processes. Reviews in Mineralogy and Geochemistry 2015, 80, (1), 461-481.

594 595 596

68. Molins, S.; Trebotich, D.; Steefel, C. I.; Shen, C., An investigation of the effect of pore scale flow on average geochemical reaction rates using direct numerical simulation. Water Resources Research 2012, 48, (3).

597 598 599

69. Bruns, S.; Stipp, S. L. S.; Sørensen, H. O., Looking for the Signal: A guide to iterative noise and artefact removal in X-ray tomographic reconstructions of porous geomaterials. Advances in Water Resources 2017, 105, 96-107.

ACS Paragon Plus Environment

32

Page 33 of 33

Environmental Science & Technology

600 601

70. Bruns, S.; Sørensen, H. O.; Stipp, S. L. S., Rock Properties of Compacted North Sea Chalks characterized by Greyscale Analysis. in review at Water Resources Research 2016.

602 603 604

71. Jha, D.; Sørensen, H. O.; Dobberschütz, S.; Feidenhans; apos; l, R.; Stipp, S. L. S., Adaptive center determination for effective suppression of ring artifacts in tomography images. Applied Physics Letters 2014, 105, (14), 143107.

605 606

72. Wang, Y.; Yang, J.; Yin, W.; Zhang, Y., A New Alternating Minimization Algorithm for Total Variation Image Reconstruction. SIAM Journal on Imaging Sciences 2008, 1, (3), 248-272.

607 608

73. Yang, Y.; Bruns, S.; Stipp, S.; Sørensen, H., Temporal pattern of entropy production induced by reactive infiltration instability in natural porous media. arXiv:1704.01059 2017.

609 610

74. Yang, Y.; Hakim, S. S.; Bruns, S.; Uesugi, K.; Dalby, K. N.; Stipp, S. L. S.; Sørensen, H. O., Path selection during wormhole growth. in review at Science Advances (arXiv:1704.01062) 2017.

611 612 613

75. Zhen-Wu, B.; Dideriksen, K.; Olsson, J.; Raahauge, P.; Stipp, S. L. S.; Oelkers, E., Experimental determination of barite dissolution and precipitation rates as a function of temperature and aqueous fluid composition. Geochimica et Cosmochimica Acta 2016, 194, 193-210.

614 615 616

76. Appelo, C. A. J.; Verweij, E.; Schäfer, H., A hydrogeochemical transport model for an oxidation experiment with pyrite/calcite/exchangers/organic matter containing sand. Applied Geochemistry 1998, 13, (2), 257-268.

617 618 619

77. Daveler, S. A.; Wolery, T. J., EQPT, A Data File Preprocessor for the EQ3/6 Software Package: User’s Guide and Related Documentation (Version 7.0). Lawrence Livermore National Laboratory, UCRlMA-110662 PT II 1992, 1-89.

620 621 622 623 624

78. Nordstrom, D. K.; Plummer, L. N.; Wigley, T. M. L.; Wolery, T. J.; Ball, J. W.; Jenne, E. A.; Bassett, R. L.; Crerar, D. A.; Florence, T. M.; Fritz, B.; Hoffman, M.; Holdren, G. R.; Lafon, G. M.; Mattigod, S. V.; McDuff, R. E.; Morel, F.; Reddy, M. M.; Sposito, G.; Thrailkill, J., A Comparison of Computerized Chemical Models for Equilibrium Calculations in Aqueous Systems. In Chemical Modeling in Aqueous Systems, American Chemical Society: 1979; Vol. 93, pp 857-892.

625 626

79. Stryjek, R.; Vera, J., PRSV: An improved Peng—Robinson equation of state for pure compounds and mixtures. The Canadian Journal of Chemical Engineering 1986, 64, (2), 323-333.

627 628

80. Pitzer, K.; Brewer, L., revised edition of Thermodynamics by GN Lewis and M. Randall. In McGraw-Hill, New York, NY: 1961.

629

ACS Paragon Plus Environment

33