• Nie Znaleziono Wyników

Geochemical modelling of hydrogen gas migration in an unsaturated bentonite buffer

N/A
N/A
Protected

Academic year: 2021

Share "Geochemical modelling of hydrogen gas migration in an unsaturated bentonite buffer"

Copied!
23
0
0

Pełen tekst

(1)

1 Geochemical Modelling of Hydrogen Gas Migration in Unsaturated Bentonite

1

Buffer 2

3

Majid Sedighi1*, Hywel Rhys Thomas1, Shakil Al Masum1, Philip James Vardon1,2, 4

Duncan Nicholson3, Alex Chen3 5

6

1 Geoenvironmental Research Centre, Cardiff School of Engineering, Cardiff 7

University, Queen’s Buildings, The Parade, Cardiff, CF24 3AA, United 8

Kingdom. *Corresponding author (SedighiM@cf.ac.uk) 9

2 Now at Geo-Engineering Department, Technical University of Delft, Postbus 5, 10

2600 AA, Delft, The Netherlands. 11

3 Arup, 13 Fitzroy Street, London, W1T 4BQ, United Kingdom. 12

13 14

Abstract 15

This paper presents an investigation of the transport and fate of hydrogen gas through 16

compacted bentonite buffer. Various geochemical reactions that may occur in the multiphase 17

and multicomponent system of the unsaturated bentonite buffer are considered. A reactive 18

gas transport model, developed within an established coupled thermal, hydraulic, chemical 19

and mechanical (THCM) framework is presented. The reactive transport module of the model 20

considers the transport of multicomponent chemicals both in liquid and gas phases, 21

together with an advanced geochemical reaction model. The results of a series of 22

numerical simulations of the reactive transport of hydrogen in unsaturated bentonite are 23

presented, in which hydrogen gas has been injected at a realistic rate due to the corrosion of 24

steel canister into a partially saturated bentonite buffer. Gas pressure development and the 25

fate of the hydrogen gas considering the geochemical reactions are studied. The results show 26

the high buffering capacity of unsaturated bentonite, when considering a steel canister, over a 27

period of 10,000 years. The presence of accessory minerals is shown to have an important 28

role in mitigating excess hydrogen ions, thus increasing the dissolution capacity of the system 29

to gas. Development of various forms of aqueous complexations between the inorganic 30

components and hydrogen ions were also found to be important in buffering the excess 31

hydrogen evolved. Based on the results obtained, it is postulated that the presence of various 32

chemical components in the clay buffer may influence the transport and fate of the hydrogen 33

gas. 34

(2)

2 The generation and migration of gases within a potential nuclear waste repository has been 36

the subject of considerable attention and interest for a number of years. Research has been 37

commissioned into a number of aspects, covering both the gas generation potential of the 38

materials to be used in the repository and the migration of the generated gases through the 39

surrounding engineered barriers and geosphere (e.g. Pusch et al. 1985; Horseman et al. 1999; 40

Graham et al. 2002). To the authors’ knowledge, this research has largely focused on what 41

might be described as the “physics” of the problem, i.e. the impact of gas generation and 42

migration within performance assessment approaches on the physical condition of a potential 43

high level radioactive waste disposal repository (e.g. Horseman et al. 1999; Vardon et al. 44

2013). 45

The fate of gas and its long term behaviour considering gas-geochemical interactions in the 46

multiphase and multicomponent unsaturated compacted clay buffer system has received less 47

attention. Various chemical reactions may take place between the generated gas and the 48

clay/water system affecting the fate of the gas in the near field. Additionally, these 49

interactions may alter the behaviour of the clay buffer (Yong et al. 2010). A number of 50

studies have investigated the fate of hydrogen considering geochemical reaction in clay 51

system and related to high level radioactive waste disposal (e.g. Xu et al. 2008; Lassin et al. 52

2011; De Windt et al. 2014). Lassin et al. (2011) presented an investigation of the hydrogen 53

solubility in pore water of unsaturated clay (Argillite) in the context of the geological 54

disposal of nuclear waste under different temperatures and capillary pressures. Accordingly, 55

an increase of the hydrogen solubility in the pore water-rock-gas system has been reported 56

with the decrease in relative humidity or capillary water pressure. A number of scenarios of 57

hydrogen production and diffusion have been studied in a disposal cell of high level 58

radioactive waste disposal in argillaceous formation using multiphase multicomponent 59

reactive transport modelling by De Windt et al. (2014). The effects of the saturation state on 60

the fate of gases and geochemical reaction are highlighted in the results. 61

The work presented here attempts therefore to extend current knowledge to accommodate 62

some of the above aspects, so that the impact of gas migration on the geochemical condition 63

of the unsaturated compacted bentonite buffer in a potential nuclear waste repository might 64

be explored. The results of a series of numerical simulations of gas transport in unsaturated 65

bentonite buffer are presented. 66

A reactive gas transport model developed within an established coupled thermal, hydraulic, 67

chemical and mechanical (THCM) framework has been applied to conduct the numerical 68

(3)

3 simulations. Recent developments in the gas reactive transport module of the coupled THCM 69

model are presented. 70

The gas reactive transport model has been used under isothermal and constant pore water 71

saturation conditions. Scenarios are then developed considering heterogeneous re-saturation 72

of the buffer which may provide some partially saturated regions for gas flow. In the 73

application presented, the focus has been to examine the importance of the chemical reactions 74

on the buffering the hydrogen gas injected to the system. 75

Interactions of gas with chemical components present in the bentonite-water system through 76

various geochemical reactions are considered in the simulations presented. 77

Compacted MX-80 bentonite has been selected for this study, under three different initial 78

water contents. 79

The results of pressure development and the fate of hydrogen gas are presented. A particular 80

focus of the research is to study of the mechanisms involved in mitigating excess gas 81

generation. The effects of hydrogen reactions on the geochemistry of the bentonite buffer are 82

also discussed. 83

1. Reactive chemical/gas transport model 84

A reactive transport model has been developed based on a coupled thermal, hydraulic, 85

chemical and mechanical (THCM) formulation which includes the transport of multiple 86

chemicals in both liquid and gas phases. The governing equations have been implemented in 87

the numerical coupled THCM model developed at the Geoenvironmental Research Centre, 88

Cardiff University, i.e. COMPASS (e.g. Thomas & He 1998; Thomas et al. 2012). 89

Geochemical reactions that may occur between the components existing in liquid, gas and 90

solid phases have been considered in the model development via coupling a geochemical 91

model, i.e. PHREEQC v2 (Parkhurst & Appelo 1999), with the coupled THCM model 92

(COMPASS) (Sedighi 2011; Thomas et al. 2012) . The COMPASS-PHREEQC platform has 93

been extended by the linkage of the gas transport model with the geochemical model and 94

inclusion of further geochemical features into the coupled platform (Masum 2012). 95

In this paper, an application of the chemical/gas reactive transport model is presented where 96

the recent developments in the chemical and gas reactive transport module (Sedighi 2011; 97

Masum, 2012; Thomas et al. 2012) is presented. These developments in particular include i) 98

the extension of the multicomponent chemical transport model to the gas species (Masum et 99

(4)

4

al. 2012; Vardon et al. 2013) and ii) further integration of the geochemical features to the

100

COMPASS-PHREEQC model by including the gas reactions. The governing equations of gas 101

reactive flow are described in this paper as the simulations do not consider the variations in 102

thermal, hydraulic or mechanical processes inside the buffer. Details of the theoretical and 103

numerical formulations of thermal, hydraulic, chemical and mechanical behaviour have been 104

discussed in details elsewhere (e.g. Thomas and He, 1998; Seetharam et al. 2007, Thomas et 105

al. 2012).

106

The reactive transport module of the model (COMPASS-PHREEQC) has been extensively 107

verified and the accuracy of implementation of the theoretical and numerical formulation in 108

the model has been tested against several benchmarks and alternative analytical and 109

numerical models for both the reactive transports of chemicals in liquid phase (Sedighi 2011; 110

Thomas et al. 2012) and chemicals in gas phase (Masum 2012). The reactive transport of 111

multicomponent chemicals in compacted bentonite under coupled thermal, hydraulic and 112

chemical conditions has been also validated against a series of experimental investigations 113

(Sedighi et al. 2012). 114

115

1.1 Formulations of the reactive transport of chemicals 116

The governing equations for transport of chemicals in liquid and gas are based on 117

conservation of mass. Major transport mechanisms including advection, diffusion and 118

dispersion are considered in the formulations. A sink/source term has also been included in 119

the mass conservation equation which represents the geochemical reactions that can occur 120

in/between chemical components in the three-phase (solid-gas-liquid) system of unsaturated 121

soil. 122

The conservation equation is expressed in terms of molar concentration of each chemical 123

component present in the associated phases (i.e. liquid or gas). Under constant temperature 124

and moisture content, the formulations of the chemicals in liquid and gas phases can be 125

simplified as: 126

∙ (1)

where, represents the volumetric content of liquid ( or gas ( . is the 127

chemical concentration of the ith component in liquid or gas phase. represents the

128

sink/source term which represents the total mass gained or lost due to homogeneous and/or 129

(5)

5 heterogeneous geochemical reactions occurred in or between the three-phase system, i.e. 130

(solid, water and gas system of the unsaturated soil). stands for time. is the total flux of 131

ith chemical component in the associated phase.

132

As previously described, the total chemical flux considered in the formulation, includes 133

advection, diffusion and dispersion of the components. The details of the transport 134

mechanisms and developments related to multicomponent chemicals in aqueous phase were 135

provided by Thomas et al. (2012). Developments on the transport model include an extended 136

formulation of multicomponent chemicals in liquid phase under combined electrochemical 137

and thermal diffusion potentials. 138

The advective flow of gas components is considered to be driven by the pressure gradient. 139

The mechanism is expressed using Darcy’s law, described as: 140

, (2)

where, , is the advective flux of the ith component in gas phase and is the molar

141

concentration of the gas component. represents the effective permeability to the gas 142

mixture. is the total gas pressure. 143

The total gas pressure is obtained considering ideal gas behaviour, given as: 144

(3)

where, is the universal gas constant, is temperature and represents the number of gas 145

components in the system. 146

Equation (2) and (3) describe the advective flow as a function of the total concentrations of 147

the chemicals in the gas phase. Such approach allows considering the effects of pressure 148

development of each component on the advection of the other components. 149

Diffusive flow of gases is considered via the extended Fick’s law of diffusion for 150

multicomponent gas mixtures (Taylor & Krishna 1993). The diffusion of ith component in gas

151

phase is described in a general form of multicomponent system as: 152

, (4)

where, , is the diffusive flux of the ith component. represents the effective diffusion

153

coefficient of the ith component due to the concentration gradient of the jth component.

(6)

6 The governing equation of gas transport, presented in equation (1) can then be presented as: 155

∙ (5)

The component of the geochemical sink/source term in the governing chemical transport 156

equation is calculated using a geochemical model, PHREEQC version 2 (Parkhurst & Appelo 157

1999). The geochemical reactions considered in the coupled model include ion exchange, 158

precipitation and dissolution of minerals, surface complexation, and redox reactions. The 159

transport model (COMPASS) and the geochemical model (PHREEQC) are linked together 160

incorporating a sequential non-iterative approach (Sedighi 2011; Thomas et al. 2012). 161

Detailed information of the COMPASS-PHREEQC development, the coupling scheme and 162

verifications of the model can be found in Sedighi (2011) and Masum (2012). 163

In a general form, the gas transport equation is coupled with all other primary variables of the 164

model which includes: pore water pressure, temperature, chemical concentrations and 165

deformation vector. Such coupling is reflected via the variation of degree of saturation with 166

suction and temperature as well as changes of the porosity related to deformation modelling. 167

Under the conditions of the simulations considered in this work (as described in section 2), no 168

variation of pore water pressure, temperature and porosity is considered and level of coupling 169

with other processes is minimized. The solution to the coupled differential equations is based 170

on finite element for spatial discretisation and finite difference for temporal discretisation 171

(Thomas & He 1998). 172

2.0 Simulation scenarios 173

The problem considered in this study is an unsaturated compacted bentonite exposed to 174

hydrogen gas which is generated due to the corrosion of a steel canister in a high level 175

radioactive disposal repository. The simulation scenarios that are studied are based on a 176

condition that the re-saturation of the clay buffer may occur heterogeneously. In other words, 177

it is assumed that during the hydro-thermal phase, despite the long term exposure of the 178

compacted bentonite re-saturation would not homogenously occur and some areas would 179

remain partially saturated. In fact, the results of a number of mock-up scale experiments and 180

in-situ tests on re-saturation of bentonite have shown that the full saturation has not been 181

observed (e.g. Thomas et al. 2003). Despite the fact that the compacted bentonite buffer may 182

remain for a long time under the effects hydraulic flow from the surrounding rock and an 183

extensive numerical modelling results on re-saturation of bentonite, there is still uncertainty 184

(7)

7 about the extent of the saturation with time as clay microstructural effects and rock 185

heterogeneities can considerably affect the hydraulic behaviour and predictions. This implies 186

that there can be some partially saturated regions where gas flow can preferentially occur 187

through them. The hydrogen can then be the subject of various geochemical buffering 188

processes during the migration in the unsaturated clay buffer. 189

The analysed domain is a two dimensional domain of 500×350 mm of bentonite compacted 190

at dry density of 1600 kg/m3. MX-80 bentonite was selected for this study. The porosity 191

associate with the dry density is approximately 0.4. Simulations were carried out for three 192

initial water contents of the soil, i.e. 10%, 15% and 18% (gravimetric percentage) which 193

corresponds to degrees of saturation of 40%, 60% and 70%, respectively. The suction values 194

associated with the initial water contents of the samples are obtained from the water retention 195

data of compacted MX-80 reported in the literature (Åkesson et al. 2005) and by using the 196

well-established van Genuchten’s relationship for soil-water retention curve (van Genuchten 197

1980). The values of suctions associated with each degree of saturation are approximately 198

70MPa, 35MPa and 25MPa corresponding to 40%, 60% and 70% degree of saturation, 199

respectively. It is noted that the total water content has been considered in the chemical 200

reaction calculations. It is acknowledged that in compacted bentonite, water exists in several 201

porosity scales (e.g. Bradbury & Baeyens 2002) and this can affect the distribution of 202

chemicals and the fate of the pore water composition and geochemical reactions due to the 203

available water in each pore space. 204

The physical, chemical and geochemical properties of the MX-80 bentonite required for the 205

analysis have been obtained from data reported by Bradbury & Baeyens (2002). Table 1 206

presents a selective basic mineralogical and geochemical characteristic of the MX-80. 207

Considering the partial re-saturation conditions described and the fact that the corrosion of 208

steel and consequently the release of hydrogen will occur after the thermal phase, the 209

simulations have been carried out under isothermal conditions with no hydraulic variations 210

(constant moisture content). It is acknowledged that both thermal and hydraulic processes can 211

change the initial conditions of the sample and consequently alter the geochemical 212

conditions. However, in the scope of the current study, the main focus has been to investigate 213

the extent of the gas pressure development in an unsaturated buffer and the effects of 214

geochemical reactions on buffering the gas. Simplified conditions have been considered for 215

this study to facilitate the interpretation of the results in relation to the objectives of the 216

investigation. 217

(8)

8 The domain was discretised into 80 equally sized quadrilateral elements. A variable time step 218

has been used in this work which allows a variation of time-step depending on the 219

convergence criteria. The simulations have been carried out for a period of 10,000 years. 220

2.1 Initial and boundary conditions 221

The initial pore water composition of the bentonite associated with each water content was 222

obtained via a geochemical analysis of the system using the basic mineralogical and 223

geochemical properties of the MX-80, provided in Table 1. A procedure was followed to 224

analyse and obtain the pore water composition which is similar to that described by Bradbury 225

& Baeyens (2003). Table 2 provides the composition of the pore fluid and Table 3 presents 226

the mineral contents and exchangeable composition of compacted MX-80, calculated from 227

the geochemical analysis. 228

The initial water pressure was obtained from the water retention curve of the compacted MX-229

80 (Börgesson & Hernelind 1999), for each initial water content. 230

Constant flux of gas was considered at one boundary whilst the other boundary was assumed 231

to be impermeable to gas flow (water boundary). Different rates of hydrogen gas 232

development due to the steel corrosion have been reported from theoretical and experimental 233

investigations (e.g. Papillon et al. 2003; Taniguchi et al. 2004; Smart et al. 2006). Based on a 234

comprehensive literature review reported by Masum (2012) and Vardon et al. (2013), the 235

presence of bentonite may enhance the corrosion process, although the protective surface of 236

the consisted may lower the long term corrosion rate. Based on the range of values reported 237

for the hydrogen generation reported by Taniguchi et al. (2004) and Smart et al. (2006) for 238

compacted bentonite, a constant hydrogen flux equal to 2.0×10-11 kg/m2/sec has been selected 239

which corresponds to a range of 1 to 3 μm/year rate of steel corrosion. 240

A schematic diagram of the domain, chemical initial conditions and boundary conditions 241

applied for the case of the scenario with the initial water content of 10% is shown in Fig 1. 242

2.2 Material parameters 243

Material parameters required for the simulations include i) the flow properties and ii) 244

equilibrium constants of chemical reactions considered. 245

The relationship used to describe the gas conductivity was adopted from Ho & Webb (2006) 246

given as: 247

(9)

9 (6)

where, is the intrinsic permeability of the clay and is the viscosity of the gas. is the 248

relative gas conductivity: 249

1 1 (7)

where is a constant which is obtained from the water retention curve described by the van 250

Genuchten’s relationship (van Genuchten 1980). This value has been reported to be 0.43 for 251

the MX-80 (Börgesson & Hernelind 1999). , represents the effective saturation, which can 252

be given as: 253

1

⁄ (8)

where, and is the degree of saturation and the residual degree of saturation, 254

respectively. 255

In the absence of the experimental data for the intrinsic gas permeability of compacted MX-256

80 bentonite, an intrinsic permeability value equal to 10-12 m2 has been considered. This value 257

has been obtained from the experimental tests carried on the compacted FEBEX bentonite, 258

indicating a range between 10-12 and 10-16 m2 for specimens with dry densities between 1500 259

and 1700 kg/m3 (Huertas et al. 2000). The dynamic viscosity of the hydrogen gas is assumed 260

to be equal to the air viscosity at standard temperature and pressure. 261

The effective diffusion of hydrogen gas in a single component system has been obtained 262

from: 263

, (9)

where , represents the diffusion of hydrogen in air. is the tortuosity factor for gas 264

transport which was obtained from a relationship proposed by Millington & Quirk (1961): 265

⁄ ⁄ (10)

where, is the porosity of the clay and .is the degree of gas saturation. 266

Similarly, the diffusion coefficients of chemicals in liquid phase were considered via a 267

similar relationship as that described in equation (9): 268

(10)

10 where, , represents the self diffusion coefficients of ions in water. is the tortuosity factor 269

of chemical flows obtained from the relationship proposed by Millington & Quirk (1961): 270

(12)

2.3 Geochemical reactions 271

Various geochemical reactions that may occur due to the exposure of the system to hydrogen 272

gas have been considered. All the reactions included are calculated under equilibrium with 273

gas pressure, considering a “closed” thermodynamic system. 274

Four main categories of geochemical reactions were considered in the simulations: 275

1) Redox reactions including hydrogen phase exchange between gas and liquid phases (i.e. 276

hydrogen gas dissolution in water). 277

2) Precipitation/dissolution of minerals. 278

3) Aqueous complexations between different components in liquid phase. 279

4) Exchange of ions between the components in liquid phase and exchangeable ions of the 280

clay. 281

Table (4) presents the reactions considered and the equilibrium constants associated with 282

each reaction. The equilibrium constants of the reactions described in categories 1 to 3 have 283

been obtained from the values provided in PHREEQC ver 2.15 (Parkhurst & Appelo 1999). 284

The equilibrium constants of ion exchange reactions described under the category 4 were 285

obtained from the values presented by Bradbury & Baeyens (2003) for MX-80 bentonite. 286

It is noted that under the scenarios considered the water pressure is highly negative and the 287

water in under metastable condition (Lassin et al. 2005). The implies that the application of 288

Henry’s low and the equilibrium constant of the hydrogen dissolution may involve a degree 289

of approximation under the high negative pore water as it has been described in the literature 290

(e.g. Lassin et al. 2005). Under high negative water pressure, application of other 291

relationships such as the equation state proposed by Spycher & Reed (1988) has been 292

suggested and tested for hydrogen solubility in hydrogen solubility in pore water of partially 293

saturated argillites (Lassin et al. 2011). It is acknowledged that the application of the Henry’s 294

low used in this study under the high negative pore water conditions of the simulations may 295

underestimate the extent of the hydrogen dissolution in water. 296

(11)

11 An assumption has been considered that all the chemical reactions are assumed to be under 297

equilibrium conditions. However, it is acknowledged that further analysis is required to 298

assess the effects of the kinetically controlled reactions on the fate of hydrogen. In particular, 299

the effects of kinetics on sulphate reduction reaction which play an important role in 300

buffering the excess hydrogen ions (as it is shown in the results) requires further analysis. 301

This is due to the fact that long half-life has been reported for the sulphate reductions under 302

the conditions related to the radioactive waste disposal (Truche et al. 2009). Future work will 303

be required to provide a more comprehensive understanding of the system under mixed 304

equilibrium and kinetically controlled reactions. 305

3.0 Simulation results and discussion 306

The variations of the major gas and chemical variables/components observed from the 307

simulations at three different initial water contents, i.e. 10, 15 and 18 %, are presented. It is 308

noted that no water flow into/from the boundaries have been considered in the simulations. 309

Therefore, the initial moisture content considered remains the same which allows 310

investigating the geochemical effects at three partially saturated samples with the initial water 311

contents described above. 312

Selective results presented and discussed here include the variations of i) concentration of 313

hydrogen gas, ii) pH and pe (pe is defined as the activity of the free electron in water, i.e. 314

pe log e ) and iii) major minerals. 315

The results are only presented for the boundary condition adjacent to the gas injection face of 316

the soil domain. At this location (representing the canister-bentonite buffer interface), the 317

buffers may experience longer physical and chemical interactions due to the hydrogen influx 318

than the other locations of domain. At this location, a higher gas pressure development is 319

more likely as the soil-water system is exposed to a constant hydrogen flux whilst in 320

downstream of the sample, buffering of the gas due to chemical reactions at locations closer 321

to the gas injection boundary may hinder the pressure development. This location is however 322

can be considered as a representative of the domain under the long term simulations. 323

Considering the size of the domain (35 cm) and closed boundary conditions at the 324

downstream of the sample, a uniform distribution of chemicals and gas has been observed in 325

the entire domain in the results of 10,000 years or higher (Masum 2012). Since the focus of 326

the paper has been to study the gas pressure development in long term analysis, the results of 327

(12)

12 the location selected for the results can be considered as the representative of the whole 328

domain in 10,000 years analysis. 329

3.1 Hydrogen gas evolution 330

The result of prediction of hydrogen concentrations at the gas injection face is presented in 331

Fig. 2 for three initial water contents. The maximum gas concentration developed at the 332

boundary is related to the highest water content. A “S” shape type of behaviour is observed 333

for all three scenarios which indicates a gradual accumulation of gas at the boundary, 334

followed by a relatively sharp increase in the concentration. The concentration then follows 335

another period of gradual increase. The available pore space for gas decreases in the 336

simulation scenarios with higher initial water contents, therefore, an increase of the gas 337

concentration is observed with increasing the initial water content. 338

At the end of simulation period, the maximum gas concentration at 10% water content has 339

reached an equivalent pressure of 2.45 bar. This value is significantly lower compared to a 340

similar prediction in which only dissolution of gas into pure water has been considered, i.e. 341

maximum gas pressure 9.0 bar (Masum et al. 2012). This observation suggests that 342

significant involvement of geochemical processes can mitigate the excess gas generated in 343

the system. In terms of the spatial distribution of hydrogen, the results of analysis show a 344

uniform gas concentration/pressure across the domain at 10,000 years. 345

Based on the prediction results obtained, the saturation index of hydrogen after 10,000 years 346

simulation shows a negative value of -6.43. The saturation index (SI) represent the logarithm 347

of ratio of the ion activity product ( to the solubility product constant ( , i.e. 348

. This suggests that the solution is under-saturated and able to accommodate 349

more hydrogen. The results of similar simulation for up to 100,000 years indicated that after 350

60,000 years, the gas pressure starts to build up rapidly in the system to a maximal value of 351

4.95 MPa. A detailed explanation of the possible chemical reactions to buffer the gas in the 352

system is presented later in the discussion. 353

3.2 pH and pe 354

The pH and redox potential (pe) variations of the buffer at the gas injection face are presented 355

in Fig. 3 and 4, respectively. Several processes are involved to control the pH behaviour of 356

the buffer including the hydrogen dissolution in water and mineral precipitation/dissolution. 357

Relatively small variations of the pH are observed. The values obtained at the end of 358

(13)

13 simulations remain closely to the initial pH value of the system (around 8.0) for all three 359

initial water contents. 360

The redox behaviour shows a three -stage evolution pattern with time. The evolution involves 361

a gradual decrease, with a rapid drop in a short period of time, which is followed by relative 362

steady-state for long term. In general, the intrusion of hydrogen to the system reduces the 363

redox potential of the system by increasing the activity of free electron in water, i.e. pe. The 364

first phase of gradual decrease is more highlighted in the case of the scenario with 10% initial 365

water content. 366

Higher reduction in pe is observed for samples with higher water content in Fig. 4. The pe 367

decreases from the initial value of -4.75 to -5.35, -5.57 and -5.64 for 10%, 15% and 18%, 368

respectively. The variations of the redox potential can affect the dissolution and precipitation 369

behaviour of the minerals (Appelo & Postma 2005). 370

Similar time scale can be observed for the sharp rise of the gas concentration (or pressure) 371

from Fig. 2 and the sharp decrease of pe. These observations will be discussed along other 372

geochemical variations in section 3.4. 373

3.3 Minerals (Gypsum and Calcite) 374

The evolution of gypsum and calcite at the gas entrance boundary is presented in Fig. 4 and 375

5, respectively. The initial concentration of gypsum varies from 0.0113 to 0.003 mol/kg of 376

the soil with increasing the water content from 10% to 18%. As shown in Fig. 4, in scenario 377

with 10% initial water content, the total amount of available gypsum in the system is 378

dissolved almost linearly in the pore-water after a period of approximately 6000 years. The 379

dissolution of gypsum has occurred faster in scenarios with higher water content. Dissolution 380

of gypsum increases the amount of calcium (Ca2+) and sulphate (SO42-) ions in the pore-381

water. 382

The amount of calcite shows an increase with time. During the prediction period, calcite 383

mineral has accumulated in the system from an initial value of approximately 0.0705 to 384

0.0822 (mol/kg of soil). 385

It is noticeable from Fig. 6 that the precipitation of calcite continues for a relatively long 386

duration for the scenario with 10% initial water content. The calcite precipitation has also 387

linearly continued for approximately 4700 years until it reaches a steady state in the case of 388

18% water content simulation. The increasing precipitation of calcite is related to the 389

progressive dissolution of gypsum, as shown in Fig. 5. This process involves the reaction of 390

(14)

14 bicarbonate ions with calcium ions added by dissolution of gypsum. In the case of scenario 391

with 18% initial water content, after approximately 4700 years, the saturation index of calcite 392

reaches zero. This suggests that the pore-water has become saturated with precipitated calcite. 393

It is noted that the change in calcite content also affects the concentration of calcium, 394

bicarbonate ions and exchangeable ions. 395

3.4 Discussion of the results 396

Based on the prediction results presented in previous section, a large amount of injected 397

hydrogen gas was found to be buffered in clay-water system. The mitigation of hydrogen gas 398

in the system involved several geochemical reactions as shown in Table 4. These reactions 399

directly or indirectly increase the capacity of the system to adsorb gas and buffer it in the 400

system. 401

Dissolution of hydrogen gas into water is the main phase change process, given as: 402

H ↔ 2H 2e 403

The dissolution of hydrogen into pure water has limited capacity to buffer the excess 404

hydrogen in the system. In the absence of other chemical components or minerals in water, 405

only a small quantity of gas can be dissolved in water according to Henry’s law. However, by 406

considering the whole geochemical system, the pore water demand for hydrogen ion can be 407

increased considerably due the larger number of reactions which involve hydrogen ions. 408

The constant influx of hydrogen increases the amount of (H ) ion and decreases the redox 409

potential of the porewater. At lower pe, the solution has a higher tendency to donate protons 410

or H to the system. As an example, the demand for SO42‐ and H increases when HS is

411

produced through the redox reaction: 412

SO42‐ 9H 8e↔HS4H 2O

It is noted that the sulphates reduction presented in the above reaction is reported to occur 413

only under bacterial activity and considering the long half-life of the reaction reported should 414

be analysed as a kinetically controlled reaction (Truche et al. 2009). In this study all 415

reactions, including the sulphate reaction has been considered under equilibrium. It is 416

acknowledged that further analysis considering kinetics of the reaction will provide further 417

insight into the natural conditions. 418

This process increases the dissolution of gypsum in the water to provide SO42‐ ions. 419

(15)

15 The gypsum dissolution in fact provides excess amount of SO42- and Ca2 ions in the system, 420

required to generate a number of other hydrogen impregnated compounds such as species 421

CaHSO , H S, HCO , CaHCO . The evolution of these species in the solution accelerates the 422

consumption of both hydrogen ions (H ) and (e ), shifting the equilibrium position of the 423

hydrogen dissolution (H ↔ 2H 2e ) from left to right causing more hydrogen to be 424

dissolved in the solution from gas phase. 425

After the completion of the dissolution of all gypsum, the demand for hydrogen in liquid 426

phase reduces and a rapid rise of hydrogen concentration in gas phase can be observed (Fig. 427

1). Other aqueous complexation process has continued to mitigate the excess amount of 428

hydrogen in the system but less effectively than when gypsum was available in the system. 429

During the gypsum dissolution period the increasing amount of H+ ions in the liquid from gas 430

phase causes the pH of the solution to drop (Fig. 3). The pH started to rise gradually 431

following the calcite precipitation process until both calcite precipitation and pH evolution 432

reaches to a steady state. Over the duration of the simulation, the pH of the system has not 433

changed considerably and only varied from an initial value of 8.0 to a final value of 8.12. It is 434

noted that surface complexation reactions can also control the pH buffering. These include 435

the reactions of hydrogen ions with weak sites (SOH) in MX-80 (Bradbury & Baeyens 2002). 436

The latter process was not included in the simulations presented. 437

438

4.0 Conclusions 439

The transport and fate of hydrogen gas in an unsaturated bentonite has been studied through a 440

series of numerical simulations. Various chemical reactions that may occur due to the 441

presence of excess hydrogen gas have been considered. 442

The results show the high buffering capacity of bentonite buffer to accommodate the amount 443

of gas generated considering a realistic corrosion rate of a steel canister over a period of 444

10,000 years. The presence of accessory minerals was found to have an important role in 445

mitigating the excess hydrogen ions, thus increasing the dissolution capacity of the system to 446

gas. Development of various forms of aqueous complexations between the inorganic 447

components and hydrogen ions has also been found to be important in buffering the excess 448

hydrogen evolved. 449

(16)

16 Based on the results obtained, it is postulated that the presence of various chemical 450

components in the clay buffer may influence the transport and fate of the hydrogen gas. It is 451

suggested that this investigation of gas flow/geochemical interactions has provided some new 452 insights. 453 454 Acknowledgement 455

Financial support from Arup, in the form of a PhD Fellowship to the third author, is 456 gratefully acknowledged. 457 458 References 459

1. Åkesson, M., Birgersson, M., & Hökmark, H. 2005. EBS task force benchmark 460

calculations 1.1 and 1.2. EBS Task Force, Barcelona. 461

2. Appelo, C. A .J. & Postma, D. 2005. Geochemistry, Groundwater and Pollution. A.A. 462

Balkema, Rotterdam. 463

3. Börgesson, L. & Hernelind, J. 1999. Coupled thermo-hydro-mechanical calculations of 464

the water saturation phase of a KBS-3 deposition hole. Influence of hydraulic rock

465

properties on the water saturation phase. SKB TR-99-41. Swedish Nuclear Fuel and

466

Waste Management Co, Stockholm, Sweden. 467

4. Bradbury, M. & Baeyens, B. 2002. Porewater chemistry in compacted re-saturated MX-468

80 bentonite: physicochemical characterisation and geochemical modelling. Nagra

469

Technical Report NTB 01– 08, Wettingen, Switzerland.

470

5. De Windt, L., Marsal, F., Corvisier, J., & Pellegrini, D. 2014. Modeling of oxygen gas 471

diffusion and consumption during the oxic transient in a disposal cell of radioactive 472

waste, Applied Geochemistry, 41, 115-127. 473

6. Graham, J., Halayko, K. G., Hume, H., Kirkham, T., Gray, M. & Oscarson, D. 2002. A 474

capillarity-advective model for gas break-through in clays. Engineering Geology 64 (2-475

3), 273-286. 476

7. Ho, C. & Webb, S. 2006. Gas Transport in Porous Media. Springer, AA Dordrecht, The 477

Netherlands. 478

(17)

17 8. Horseman, S. T., Harrington, J. F. & Sellin, P. 1999. Gas migration in clay barriers. 479

Engineering Geology 54 (1-2), 139-149.

480

9. Huertas, F., Fuentes-Cantillana, J. L., Jullien, F., Rivas, P., Linares, J., Farina, P., 481

Ghoreychi, M., Jockwer, N., Kickmaier, W., Martines, M. A., Samper, J., Alonso, E., & 482

Elorza, F. J. 2000. Full-scale engineered barriers for a deep geological repository for 483

high-level radioactive waste in crystalline host rock (FEBEX project). European

484

Commission, Nuclear science and technology series, Report EUR 19147 EN. 485

10. Lassin, A., Azaroual, M. & Mercury L., 2005. Geochemistry of unsaturated soil systems: 486

aqueous speciation and solubility of minerals and gases in capillary solutions. 487

Geochimica et Cosmochimica Acta 69, 5187-5201.

488

11. Lassin, A., Dymitrowska, M., & Azaroual, M., 2011. Hydrogen solubility in pore water 489

of partially saturated argillites: Application to Callovo-Oxfordian clayrock in the context 490

of a nuclear waste geological disposal, Physics and Chemistry of the Earth 36, 1721– 491

1728. 492

12. Masum, S. A. 2012. Modelling of reactive gas transport in unsaturated soil. A coupled 493

Thermo-Hydro-Chemical-Mechanical Approach, PhD Thesis, Cardiff University, UK.

494

13. Masum, S. A., Vardon, P. J., Thomas, H. R., Chen, Q. & Nicholson, D. 2012. 495

Multicomponent gas flow through compacted clay buffer in a higher activity radioactive 496

waste disposal facility. Mineralogical Magazine 75 (8), 3337-3344. 497

14. Millington, R. J. & Quirk, J. M. 1961. Permeability of porous solids. Transactions of 498

Faraday Society 57, 1200–1207.

499

15. Papillon, F., Jullien, M., & Bataillon, C. 2003. Carbon steel behaviour in compacted 500

clay: two long term tests for corrosion prediction. In Prediction of Long Term Corrosion

501

Behaviour in Nuclear Waste Systems. European Federation of Corrosion Publication , 502

36, 439-454. 503

16. Pusch, R., Ranhagen, L. & Nilsson, K. 1985. Gas migration through MX-80 bentonite. 504

Nagra Technical Report NTB 85-36, Wettingen, Switzerland. 505

17. Parkhurst, D. L. & Appelo, C. A. J. 1999. User’s guide to PHREEQC (version 2), U.S. 506

Geological Survey, Water Resource Investigation Report, 99-4259. 507

(18)

18 18. Sedighi, M. 2011. An investigation of Hydro-geochemical processes in coupled thermal, 508

hydraulic, chemical and mechanical behaviour of unsaturated soils. PhD thesis, Cardiff

509

University, UK. 510

19. Sedighi, M., Thomas, H. R. & Vardon, P. J. 2012. On the Reactive Transport of 511

Chemicals in Unsaturated Expansive Soils, 4th International Conference on Problematic

512

Soils, 21-23 September 2012, Wuhan, China, 211-218. 513

20. Seetharam, S.C., Thomas, H.R., & Cleall, P.J. 2007. Coupled thermo-hydro-chemical-514

mechanical model for unsaturated soils-Numerical algorithm. International Journal of 515

Numerical Methods in Engineering, 70, 1480-1511.

516

21. Smart, N.R., Rance, A.P., & Fennell, A.H. 2006. Expansion due to the anaerobic 517

corrosion of iron. SKB Technical Report TR-06-41, Stockholm, Sweden.

518

22. Spycher, N.F. & Rreed, M.H. 1988. Fugacity coefficients of H2, CO2, CH4, H2O and of 519

H2O-CO2-CH4 mixtures: A virial equation treatment for moderate pressures and 520

temperatures applicable to calculations of hydrothermal boiling. Geochimica et 521

Cosmochimica Acta, 52, 739-749

522

23. Taniguchi, N., Kawasaki, M., Kawakami, S., & Kubota, M., 2004. Corrosion behaviour 523

of carbon steel in contact with bentonite under anaerobic condition. In Prediction of long

524

Term Corrosion in Nuclear Waste Systems. Proc. 2nd Int. Workshop, Nice, September 525

2004 (European Federation of Corrosion and Andra), 24-34. 526

24. Taylor, R. & Krishna, R. 1993. Multicomponent Mass Transfer Wiley, New York. 527

25. Thomas, H. R. & He, Y. 1998. Analysis of coupled heat, moisture and air transfer in a 528

deformable unsaturated soil, Géotechnique 45 (4), 677-689. 529

26. Thomas, H. R., Sedighi, M. & Vardon, P. J. 2012. Diffusive reactive transport of 530

multicomponent chemicals under coupled thermal, hydraulic, chemical and mechanical 531

conditions. Journal of Geotechnical and Geological Engineering, 30 (4), 841-857. 532

27. Thomas, H. R., Cleall, P. J., Chandler, N., Dixon, D., & Mitchell, H. P. 2003. Water 533

infiltration into a large-scale in-situ experiment in an underground research laboratory. 534

Géotechnique, 53(2), 207-224.

535

28. Truche, L., Berger, G., Destrigneville, C., Guillaume, D., Giffaut, E., & Jacquot, E. 536

2009. Experimental reduction of aqueous sulphate by hydrogen under hydrothermal 537

(19)

19 conditions: implication for the nuclear waste storage, Geochimica et Cosmochimica Acta 538

Geochim. 73, 4824-4835.

539

29. van Genuchten, M.Th. 1980. A closed-form equation for predicting the hydraulic 540

conductivity of unsaturated soils. Soil Science Society of America Journal, 44, 892-898. 541

30. Vardon, P. J., Thomas H. R., Masum S. A., Chen, Q. & Nicholson, D. 2013. Simulation 542

of repository gas migration in a bentonite buffer, Engineering and Computational 543

Mechanics, Special Issue Nuclear, 167, 13-22.

544

31. Xu, T., Senger, R. & Finsterle, S. 2008. Corrosion-induced gas generation in a nuclear 545

waste repository: Reactive geochemistry and multiphase flow effects. Applied 546

Geochemistry, 23, 3423-3433.

547

32. Yong, R. N., Pusch, R. & Nakano, M. 2010. Containment of High-Level Radioactive and 548

Hazardous Solid Wastes with Clay Barriers. Spon Press, New York.

549 550

List of Tables 551

552

Table 1: Mineralogical/geochemical composition of MX-80 (adopted form Bradbury and 553

Baeyens, 2002). 554

Table 2: Pore water composition of compacted MX-80. 555

Table 3: Quantities of minerals and exchangeable ions in compacted MX-80. 556

Table 4: Geochemical reactions considered in the simulations and the equilibrium constants 557

(log ) of the reactions. 558

559

List of Figures 560

561

Fig. 1: The schematic diagram of the initial and boundary conditions for the simulation 562

scenario with 10% initial water content. 563

Fig. 2: Hydrogen evolution at the injection face for 10,000 years. 564

Fig. 3: Evolution of pH in the buffer at various water contents. 565

Fig. 4: Redox behaviour during the simulation period of 10,000 years. 566

Fig. 5: Evolution of gypsum in the buffer due to hydrogen influx. 567

Fig. 6: Evolution of calcite in the buffer due to hydrogen influx over a period of 10,000 year. 568

(20)

569 570 571

Fig.1.

(21)

572 573 574 575 576 Fig.2. Fig.3. 21

(22)

577 578 579 580 581 Fig.4. Fig.5. 22

(23)

582 583 584 585 Fig.6. 23

Cytaty

Powiązane dokumenty

Some examples of polynomial interpolation methods are the Newton and cubic spline methods, which will be used to estimate and predict the gas content of hydrogen and iodine in

Nitrate concentrations in transitional season I ranged from 2,538-3,966 mg / L and had a difference during second transitional season which ranged from 0.94 to 1.39 mg / L (Fig.2) The

Łamanie Alicja Załęcka Streszczenia w języku angielskim Autorzy Streszczenia w języku rosyjskim Stanisław Karpionok. Streszczenia w języku włoskim

During the period of the sea level stabilization, the redoxcline in the Salic-Tidalic Gleysols (Thionic, Siltic) dropped below 10-15 cm enhancing mobility of the trace elements

In the second stanza (a quatrain) in the third verse, in the place of the clausula acting as rhyme, a monosyllabic word occurs instead of an expected disyllabic one,

The field-scale experiment at an artificial embankment with controlled rainfall shows clear indication that water infiltration progresses to a deeper level from the surface

Literackich reminiscencji można znaleźć na stronicach Stawu sporo12 1 3 , tym jednak, co przesądza o interpretacji powieści Riemizowa jako całości, jest

Samodzielna Brygada Strzelców Karpackich, prasa Polskich Sił Zbrojnych na Zachodzie, życie kulturalne Polskich Sił Zbrojnych na Zachodzie, Marian Hemar, poezja żołnierska,