Microsecond Molecular Dynamics Simulation of ... - ACS Publications

Aug 16, 2016 - Microsecond Molecular Dynamics Simulation of Methane Hydrate. Formation in Humic-Acid-Amended Sodium Montmorillonite. Haoqing Ji,...
0 downloads 0 Views 2MB Size
Subscriber access provided by Northern Illinois University

Article

Microsecond molecular dynamics simulation of methane hydrate formation in the humic acid amended Na-montmorillonite Haoqing Ji, Guozhong Wu, Mucong Zi, and Daoyi Chen Energy Fuels, Just Accepted Manuscript • DOI: 10.1021/acs.energyfuels.6b01544 • Publication Date (Web): 16 Aug 2016 Downloaded from http://pubs.acs.org on August 21, 2016

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

Energy & Fuels 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 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

1

Microsecond molecular dynamics simulation of methane hydrate

2

formation in the humic acid amended Na-montmorillonite

3 4





Haoqing Ji , Guozhong Wu , Mucong Zi, Daoyi Chen *

5 6

Division of Ocean Science and Technology, Graduate School at Shenzhen, Tsinghua

7

University, Shenzhen 518055, China

1

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

8

ABSTRACT:

9

Natural gas hydrate in marine sediments is a promising energy resource while the

10

atomic level understanding of its formation on the organo-mineral complex remains

11

limited. Microsecond molecular dynamics simulations were performed to investigate

12

the methane hydrate growth in the sodium montmorillonite interlayer in presence of

13

natural sediment organic matter (Leonardite humic acid, LHA) at mass concentration

14

of 2% and 11%, respectively. The hydrate growth were characterized by the global

15

and local four-body order parameter, surface distribution function, snapshots of

16

molecular configurations and face-saturated incomplete cage analysis. It clearly

17

demonstrated the kinetic inhibition effects of LHA on hydrate formation on clay

18

minerals especially when the self-aggregation of LHA took place at high

19

concentration. Overall results highlighted the role of methane adsorption on LHA

20

aggregates on the observed inhibition phenomenon, which changed the pathway of

21

gas molecules by complex dynamic processes such as aggregates deformation, cage

22

break and cage re-formation.

23 24

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

25

1. INTRODUCTION

26

Natural gas hydrate (NGH) is a promising candidate for energy resource with a carbon

27

quantity twice more than the combination of all fossil fuels.1 Methane hydrate is the most

28

widespread type of NGH in which methane molecules are encaged in the water cages

29

formed by hydrogen bonds. It was estimated that the hydrate resource in the marine

30

sediments outnumbered that in the permafrost environment by more than two-order of

31

magnitude.2 Geochemical data further indicated that the smectite, a group of 2:1 clay

32

minerals such as sodium montmorillonite (Na-MMT), was the most abundant mineral in the

33

oceanic hydrate-bearing sediments.3,

34

explore the interactions between gas hydrate and clay minerals. It is known that the mineral

35

surface can facilitate hydrate crystallization and stabilize the hydrate cages by providing

36

nucleation sites.5,

37

intercalation of NGH in the clay mineral interlayers, because the hydrate formation varied

38

against the surface hydrophobicity,7 surface charge,8 surface area,9 pore size distribution,10

39

particle size and chemical compositions of minerals.11

6

4

Therefore, there are increasing investigations to

Nevertheless, scientific knowledge remains incomplete about the

40

Moreover, it becomes more complicated when natural soil organic matters (e.g. humic

41

substances, lignins and compounds with amide and amine groups) are present in the clays.

42

For example, it was demonstrated that the organic matters inhibited the hydrate formation

43

kinetics by forming hydrogen bonds with water and disrupting the hydrogen bond network

44

of water.12, 13 They might also inhibit the hydrate re-formation from the partial hydrate cages

45

and dissolved gas by eliminating the memory effect which is a phenomenon that gas and

46

water with previous hydrate history is easier to form hydrates.14, 3

ACS Paragon Plus Environment

15

By contrast, the

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

47

hydrophobic groups like alkyl chains in the organic matters decreased the inhibition effects

48

and even accelerated the hydrate growth by strengthening the local water structure.12 More

49

interestingly, experimental results demonstrated that the soil organic matter played a key

50

role as kinetic inhibitor during hydrate formation, but it turn to enhance the hydrate

51

nucleation kinetics by up to one-order of magnitude after adsorption on the Na-MMT

52

compared with that on bare Na-MMT or in bulk water.16-18 These results implied the

53

existence of complex interactions between gas hydrate and organo-minerals, which

54

demands future works for better understanding the hydrate evolution mechanism in the

55

hydrate-bearing clay sediments.

56

Additionally, majority of current studies attributed the inhibition or promotion effects of

57

organic matters to their interactions with water molecules. Relatively less attention was paid

58

to their influences on the gas transport, because the organic matters previously used were

59

relatively small molecules without taking into account the formation of large clusters by

60

self-aggregation. Humic substances such as humic acids in the seawater or oceanic

61

sediments have a strong tendency to aggregate from monomer to supramolecule by direct

62

assembling or molecular bridging mediated by water or multivalent cation.19 This process

63

was critical for the adsorption and diffusion of hydrocarbons, which was supposed to affect

64

the hydrate formation and disassociation.

65

On this basis, we performed molecular dynamics (MD) simulations in this study to

66

investigate the influence of humic acids on the methane hydrate formation in the Na-MMT

67

interlayer at atomic level. Specific objectives were to identify the distribution of gas and

68

water around the monomer or nanoaggregates of humic acids, and evaluate its effects on the 4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

69

Energy & Fuels

global and local rate of hydrate formation in the organo-clay pore water.

70 71

2. METHODOLOGY

72

2.1 Modeling and simulation. MD simulations were performed using the open source

73

software Gromacs (version 5.0.5).20 The montmorillonite unit cell with a stoichiometry of

74

[Al3Mg1][Si8O20][OH]4 was used in this study (a = 0.516 nm, b = 0.897 nm, c = 0.935 nm, α

75

= 91.2°, β = 100.5°, γ = 89.6°), which was obtained from the American Mineralogist crystal

76

structure database.21 It was expanded to a supercell structure (9 × 4 × 1). A liquid slab

77

(initial thickness: 3.3 - 3.5 nm) consisted of 36 sodium ions, 1628 water and 160 methane

78

was intercalated in the expanded montmorillonite (interlayer space: 4.3 nm). The sodium

79

ions were used to compensate the negative charges in the montmorillonite, while the

80

molecular ratio of methane to water was chosen to ensure enough water for hydrating all the

81

methane and interlayer cations.5 Hydrate cages were not pre-built in the initial configuration

82

in order to minimize artificial effects during model construction. Organo-clay models were

83

also constructed by loading one and six Leonardite humic acid (LHA, C31H26O12) molecules

84

into the Na-MMT interlayer, respectively.22 The corresponding mass concentration of LHA

85

in the pore water in these two scenarios was 2% and 11%, respectively.

86

The ClayFF force field was used to model Na-MMT.23 The corresponding bond and

87

nonbond parameters are listed in Tables S1 and S2, respectively, in the supporting

88

information. The LHA was modelled using the CHARMM36 force field which had proven

89

to be compatible with ClayFF.24, 25 Parameters for LHA were manually adjusted according

90

to analogue structures in the CHARMM36 force field database. Partial charges in LHA 5

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

91

molecule are illustrated in Figure S1. Methane was modelled using the united-atom

92

Lennard-Jones model while the TIP4P-ice model was used for water.26 Lorentz-Berthelot

93

mixing rules were implemented to calculate the cross interactions between different

94

molecules.27 Short-range interactions were truncated at 1.2 nm. Long-range electrostatic

95

interactions were calculated using the particle mesh Ewald method with a Fourier spacing

96

of 0.12 nm.28 The leap-frog algorithm with a time step of 1 fs was used to integrate the

97

motion equations.29

98

Energy minimization was performed using the steepest descent algorithm before

99

simulation, which was followed by 200 ps NPT runs for equilibration under constant

100

temperature (250 K) and pressure (500 bar). This step was followed by 3 µs production run

101

at NPT ensemble using the Nose-Hoover thermostat and Parrinello-Rahman barostat.30, 31

102

Semi-isotropic pressure coupling was used to barostat the system along the z direction

103

separately from the x and y dimensions. This allowed the the pressure component normal to

104

the clay layer to fluctuate independently from the tangential pressure components.

105

Three-dimentional periodic boundary condition was applied throughout simulations.

106 107

2.2 Data Analysis. The degree of methane hydrate formation was quantified by the four-body order parameter F4φ,32 which was calculated as follows: F



1 =  cos 3φ  

108

where n is the total number of H2O-H2O pairs with the distance between oxygen atoms less

109

than 3.5 Å, φi is the H-O···O-H torsion angle between oxygen atoms and two outermost

110

hydrogen atoms in the ith H2O-H2O pair. The average F4φ values for ice, liquid water and

111

hydrate are -0.4, -0.04, 0.7, respectively.33 6

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

112

The face-saturated incomplete cage analysis (FSICA) method was used to determine the

113

changes in the cage water and gas methane during simulation.34 Particularly, a gas methane

114

molecule was identified if the number of the surrounding water molecules was less than 16

115

within a sphere of 0.54 nm radius.

116

The intensity of LHA aggregation was quantified by the intermolecular contacts, which

117

was counted if the minimum distances between each two LHA molecules were ≤ 0.5 nm as

118

suggested by Zhu et al.35

119 120

3. RESULTS AND DISCUSSION

121

3.1 Influence of LHA monomer on the hydrate formation. The equilibrium

122

conformations of different molecules in the Na-MMT interlayer are shown in Fig. 1. The

123

corresponding density profiles along the normal direction of the Na-MMT surface are

124

shown in Fig. 2. Two sharp peaks were found at 1.0 and 4.1 nm, respectively, suggesting the

125

formation of a dense water film on the Na-MMT surface. This portion of water was

126

supposed to contribute little to the formation of methane hydrate because (i) the thermal

127

motion of sodium ions on the Na-MMT surface would decrease the water activity and

128

disrupt the formation of hydrogen-bonded water networks,16, 36 and (ii) the hydrophobic and

129

nonpolar methane molecules on the hydrophilic Na-MMT surface were inadequate for

130

hydrate formation (Fig. 2).

131

It appeared that the LHA monomer was immobilized by the surrounding methane hydrate

132

cages. It was unlikely to move towards the Na-MMT surface although it contained a couple

133

of hydroxyl and carboxyl groups with affinity to the oxygen and siloxy on the Na-MMT 7

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

134

surface. Nevertheless, the overall results demonstrated the kinetic inihibition effects of the

135

LHA on the methane hydrate formation in the Na-MMT pore. For example, the global F4φ

136

parameter during the initial 600 ns decreased by up to 16% when one LHA molecule was

137

added in the pore water (Fig. 3A). This was also supported by the less number of cage water

138

(Fig. 3B) and higher fractions of gas methane (Fig. 3C) after the addition of LHA. The

139

effects of space constraints after LHA addition on the decreased rate of hydrate formation

140

could be excluded, because the simulation box was flexible to stretch throughout the NPT

141

simulation. This inihibition tendency was inconsistent with Kyung et al.16 which reported

142

the accelerated hydrate induction on the organo-mineral complexes in presence of glycine.

143

The different findings were due to the fact that the thermal flucturation of Na+ on MMT

144

surface was suppressed after coordinating with the –COO- in the zwitter-ionic glycine, but

145

the LHA in this study was neutral without significantly altering the distribution of Na+ (Fig.

146

2).

147

In order to identify whether such inihibition effects were resulted from the changes in the

148

thickness of the water film aforementioned, we divided the pore water into various slices

149

according to the distance to Na-MMT surface. Each two neighbour slices shared an overlap

150

region of 0.27 nm width and the local F4φ parameter was then calculated in each slice. As

151

shown in Fig. 4, slice “n” represents the water slice ranging from 0.03 (n-1) to 0.3 + 0.03

152

(n-1) nm from the Na-MMT surface. The thickness of the water slice with little contribution

153

to the hydrate formation was determined by the position where the local F4φ values

154

increased suddenly. For example, the F4φ values almost kept constant in the first five slices

155

with less than 0.42 nm far from the Na-MMT surface, while explicit increase in the F4φ was 8

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

156

noticed in the sixth slice. It suggested that the water film on the Na-MMT surface was about

157

0.42 nm thick, which varied little after LHA addition (Fig. 4). Nevertheless, the hydroxyl

158

and carboxyl groups in the LHA molecule were expected to interact with water and

159

therefore delay the hydrate formation. As shown in Fig. 5A, the first peak was found at 0.28

160

nm in the RDF between oxygen in LHA (OL) and oxygen in water (OW), corresponding to

161

the distance between two hydrogen-bonded oxygen atoms. Similar phenomenon was

162

reported by Xu et al.,13 which attributed the inhibition effects of natural product pectin on

163

hydrate formation to the hydrogen bonding. It was inferred that the contribution of this

164

mechanism to the inhibition effects of LHA was less pronounced than pectin at the same

165

concentration, because the mass fraction of oxygen in LHA (32%) was lower than that in

166

pectin (55%).

167

For better elaborating the influence of LHA-water interactions on the dynamics of local

168

hydrate formation, we divided the overall water into several sections according to the

169

distance to the geometry center (COG) of LHA (Fig. 6). Particularly, water within a sphere

170

of 1.2 nm radius from the COG of LHA was defined as adjacent water while the rest was

171

defined as nonadjacent water. The cut-off value of 1.2 nm was selected because it was

172

enough to include all the water molecules being affected by LHA molecules, since the

173

distance between the two outermost atoms in a LHA molecule was about 1.8 nm and a

174

hydrogen bond length was about 0.3 nm. Overall, the closer to the LHA, the slower rate of

175

hydrate formation for the adjacent water. It should be noted that the nonadjacent water also

176

included the aforementioned water film near the Na-MMT surface where hydrate rarely

177

formed, therefore the average F4φ value for nonadjacent water was lower than that of the 9

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

178

adjacent water especially at the end of the simulation (Fig. 6A). An interesting finding was

179

that the hydrate formation rate for the adjacent water decreased remarkably at 200 ns and

180

greatly fluctuated until 1600 ns. It suggested that the inhibition by hydrogen-bonds

181

formation aforementioned was not the predominant mechanism at the late-stage of hydrate

182

growth. Otherwise, the hydrate growth pattern for the adjacent water should have been

183

similar to that of the nonadjacent water without substantial fluctuation, because the number

184

of hydrogen bonds between LHA and water was almost constant over the course of

185

simulation (Fig. 5C).

186

A possible reason for this finding was that the presence of LHA affected the fate and

187

transport of methane in the pore water. In order to identify these changes, the surface

188

distribution function (SDF) for methane near the LHA surface was plotted in Fig. 7A. The

189

first three peaks were characterized at around 0.35, 0.6 and 1.0 nm, respectively.

190

Accordingly, three compartments were defined including the internal, middle and external

191

regions for the partitioning of methane molecules. Results clearly demonstrated the

192

movement of methane away from the LHA surface throughout the simulation. For example,

193

the intensity of the first peak decreased with time and even disappeared during the last 1000

194

ns, which resulted in an increase in the intensity of the remaining two peaks (Fig. 7A). More

195

details could be gained from the changes in the number of methane molecules with time in

196

each region (Fig. 8A). The sharp peak located at about 140 ns in the number of methane in

197

the internal region suggested a fast initial accumulation of methane near the LHA surface,

198

which was mainly attributed to the affinity of methane to the hydrophobic functional groups

199

such as phenyl and alkyl groups in the LHA molecule. This portion of methane molecules 10

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

200

were unlikely to form hydrate directly, because they were too close to the LHA surface to

201

form complete cages and the number of the surrounding water molecules was also

202

inadequate to form hydrate structures. To confirm this, a snapshot of the water molecules

203

within 0.6 nm from each methane molecule in the internal region is shown in Fig. 9A. The

204

distance of 0.6 nm was chosen because the average cavity for a standard 512 cage (the most

205

common cage type for methane hydrate) was approximately 0.4 nm, while 0.6 nm was large

206

enough to involve the possible hydrate structure around a methane molecule. Results

207

indicated that none of the methane molecules accumulated near the LHA surface was

208

surrounded by an explicit water cage (Fig. 9A). Additionally, the methane in the internal

209

region become less stable when the hydrate started to form in the external region that would

210

facilitate the separation of methane from the LHA surface by the cage adsorption

211

mechanisms.37 Therefore, there was a quick decrease in the number of methane in the

212

internal region (Fig. 8A). However, the decreasing rate became much slower after 200 ns.

213

This was contributed by the formation of explicit hydrate cages in the internal region, which

214

made it harder for methane molecules to escape from LHA surface to the neighbor region

215

(Fig. 9B) due to the increased diffusion resistance of methane.38 A continuous diffusion of

216

methane molecules would result in the destruction and reconstruction of the hydrate cages

217

along the diffusion pathway, which accounted for the significant fluctuations found in the

218

local F4φ curves after 200 ns (Fig. 6A). As is known, the breakup and re-formation or

219

re-arrangements of water cages is a normal phemomenon during hydrate growth.39, 40 The

220

release and diffusion of methane would facilitate this process, because it would influence

221

the stable environment between water cages and bulk water pre-built by the adsorbed 11

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

222

methane.41

223

3.2 Influence of LHA nanoaggregates on the hydrate formation. Results demonstrated

224

that the kinetic inhibition effects of LHA on hydrate formation was more pronounced at

225

higher concentration. The global F4φ value during the initial 400 ns was up to 10% smaller

226

when six LHA molecules than a LHA monomer was added in the Na-MMT pore water (Fig.

227

3A). Similar trend was observed in the percentage of cage water (Fig. 3B). Additionally, the

228

distribution of the six LHA molecules close to the Na-MMT surface led to much smaller

229

values of local F4φ for the water adjacent to LHA aggregates than to a LHA monomer (Fig.

230

6B), because the water film close to the Na-MMT surface was hard for hydrate formation

231

and its thickness was not influenced by LHA concentration (Fig. 4).

232

The enhanced inhibition at high concentration was partially attributed to the increased

233

association of water with LHA, because the number of hydrogen bonds between LHA and

234

water increased by about 5-fold when the mass fraction of LHA increased from 2% to 11%

235

(Fig. 5C). However, the enhanced inhibition was obviously not proportional to the increased

236

number of hydrogen-bonds. It was inferred that the inhibition resulted from hydrogen

237

bonding mechanism was counterbalanced to some extent by the formation of LHA

238

nanoaggregates. This was evidenced by an increase in both the degree of LHA aggregation

239

(Fig. 10) and the local F4φ for the water adjacent to LHA (Fig. 6B) at around 1350 ns.

240

Results also suggested that the increased local F4φ was due to the desorption of methane

241

from the LHA driven by LHA aggregation. As shown in Fig. 8B, a sharp decrease in the

242

number of methane molecules in the internal region of LHA was found at 1200 ns, which

243

was not observed when only a LHA monomer was present in the pore water. Moreover, the 12

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

244

structure deformation of LHA aggregates also resulted in the release of methane. For

245

instance, the methane molecules were trapped in a small hydrophobic space formed by four

246

LHA molecules at 540 ns (Fig. 9C), which made it hard for the methane to transport out

247

from the internal region resulting in the appearance of a plateaus between 540 and 1000 ns

248

(Fig. 8B). The number of methane in the internal region restarted to decrease after 1200 ns

249

when the structure of LHA aggregates become less tight (Fig. 9D).

250

In order to track the pathway of the released methane, we randomly selected one of such

251

methane and compared the snapshots at different time (Fig. 11). It clearly showed the

252

moving of the released methane towards the hydrate cages near the LHA surface (399 - 404

253

ns), local breakdown of the hydrate cages (413 ns), penetration of methane through the

254

resulted open channel (443 ns) and re-formation of the hydrate cages (452 ns). A closer

255

examination of the above processes indicated the transition from two linked hexagonal faces

256

to an octagonal face when the methane approached the cages (Figs. 11A and B). Such

257

structure transformation was favorable for methane penetration, because there was a large

258

potential mean force barrier at the center of the hexagonal faces preventing methane from

259

crossing which completely disappeared in the octagonal face.42 It should be noted that the

260

penetration of methane was along a zigzag pathway accompanied by the fast break down

261

and re-formation of cage faces in different directions. Nevertheless, it was an irreversible

262

process that the passed methane was unlikely to transport back to the LHA surface when the

263

bottom cage faces were self-enclosed (Fig. 11E). The above findings suggested that the

264

hydrate formation at the late-stage was mainly inhibited by the mass transfer resistance for

265

methane through the hydrate cages already formed in the external regions. 13

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

266 267

4. CONCLUSIONS

268

Overall results demonstrated the inhibited methane hydrate growth in the Na-MMT after

269

amendment of humic acid. A summary of the processess influencing methane hydrate

270

growth observed in this study is shown in Fig. 12. The presence of LHA had little influence

271

on the distribution of sodium ions or the thickness of the water films close to the MMT

272

surface that were adverse for hydrate formation. It also suggested that the disruption of

273

water networks by forming hydrogen-bonds with pore water was not the predominant

274

mechanism for the observed inhibition by LHA addition. Instead, it was mainly contributed

275

by the adsorption of methane on LHA which made it hard for methane to form hydrate

276

directly near LHA surface or adsorbed by the water cages already formed in the external

277

region. This portion of methane tended to release from LHA again, which was driven by the

278

LHA aggregation, deformation of LHA aggregates and the competition between LHA

279

adsorption and cage adsorption. The circuitous diffusion of the released methane through

280

the solid cages by cage break and cage re-formation became the main factor limiting the

281

continue growth of hydrate.

282 283

SUPPORTING INFORMATION

284

Nonbond parameters for the CLAYFF force field (Tables S1), bond parameters for the

285

CLAYFF force field (Tables S2), partial charges for atoms in typical functional groups in

286

LHA molecule (Figure S1)

287 14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

288

AUTHOR INFORMATION

289



290

Corresponding Author

291

*E-mail: [email protected]. Telephone/Fax: +86-0755-26036290.

292

Notes

293

The authors declare no competing financial interest.

These authors contributed equally to this work

294 295

ACKNOWLEDGEMENTS

296

This study was finacially supported by National Natural Science Foundation of China (No.

297

21307069) and the Economy, Trade and Information Commission of Shenzhen Municipality

298

(Project Nos. HYCYPT20140507010002 and 201411201645511650).

299 300

REFERENCES

301

1.

302

resource: Prospects and challenges. Applied Energy 2016, 162, 1633-1652.

303

2.

304

2005, 19, (2), 459-470.

305

3.

306

Sylva, S. P., Biological formation of ethane and propane in the deep marine subsurface. Proceedings of the

307

National Academy of Sciences 2006, 103, (40), 14684-14689.

308

4.

309

Blake Ridge with downhole geochemical log measurements, Proceedings of the Ocean Drilling Program,

Chong, Z. R.; Yang, S. H. B.; Babu, P.; Linga, P.; Li, X.-S., Review of natural gas hydrates as an energy

Klauda, J. B.; Sandler, S. I., Global distribution of methane hydrate in ocean sediment. Energy & Fuels

Hinrichs, K.-U.; Hayes, J. M.; Bach, W.; Spivack, A. J.; Hmelo, L. R.; Holm, N. G.; Johnson, C. G.;

Collett, T. S.; Wendlandt, R. F. In Formation evaluation of gas hydrate-bearing marine sediments on the

15

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

310

Scientific Results, 2000; 2000; pp 199-215.

311

5.

312

and molecular dynamics simulations. The Journal of Physical Chemistry B 2003, 107, (10), 2281-2290.

313

6.

314

Langmuir 2012, 28, (20), 7730-7736.

315

7.

316

Clathrate Hydrates Formed within Hydrophilic and Hydrophobic Media: Kinetics of Dissociation and

317

Distortion of Host Structure. The Journal of Physical Chemistry C 2013, 117, (14), 7081-7085.

318

8.

319

suspensions. Environmental science & technology 2009, 43, (15), 5908-5914.

320

9.

321

in the presence of Bentonite Clay Suspension. Chemical Engineering & Technology 2013, 36, (5), 810-818.

322

10. Smith, D. H.; Wilder, J. W.; Seshadri, K., Methane hydrate equilibria in silica gels with broad pore‐size

323

distributions. AIChE Journal 2002, 48, (2), 393-400.

324

11. Heeschen, K. U.; Schicks, J. M.; Oeltzschner, G., The promoting effect of natural sand on methane

325

hydrate formation: Grain sizes and mineral composition. Fuel 2016, 181, 139-147.

326

12. Sa, J.-H.; Kwak, G.-H.; Lee, B. R.; Park, D.-H.; Han, K.; Lee, K.-H., Hydrophobic amino acids as a new

327

class of kinetic inhibitors for gas hydrate formation. Scientific reports 2013, 3.

328

13. Xu, P.; Lang, X.; Fan, S.; Wang, Y.; Chen, J., Molecular Dynamics Simulation of Methane Hydrate

329

Growth in the Presence of the Natural Product Pectin. The Journal of Physical Chemistry C 2016, 120, (10),

330

5392-5397.

331

14. Zeng, H.; Wilson, L. D.; Walker, V. K.; Ripmeester, J. A., Effect of antifreeze proteins on the nucleation,

Park, S.-H.; Sposito, G., Do montmorillonite surfaces promote methane hydrate formation? Monte Carlo

Bai, D.; Chen, G.; Zhang, X.; Wang, W., Nucleation of the CO2Hydrate from Three-Phase Contact Lines.

Takeya, S.; Fujihisa, H.; Gotoh, Y.; Istomin, V.; Chuvilin, E.; Sakagami, H.; Hachikubo, A., Methane

Lamorena, R. B.; Lee, W., Effect of pH on carbon dioxide hydrate formation in mixed soil mineral

Kumar Saw, V.; Udayabhanu, G. N.; Mandal, A.; Laik, S., Methane Hydrate Formation and Dissociation

16

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

332

growth, and the memory effect during tetrahydrofuran clathrate hydrate formation. Journal of the American

333

Chemical Society 2006, 128, (9), 2844-2850.

334

15. Zeng, H.; Walker, V. K.; Ripmeester, J. A., Approaches to the Design of Better Low‐Dosage Gas

335

Hydrate Inhibitors. Angewandte Chemie 2007, 119, (28), 5498-5500.

336

16. Kyung, D.; Lim, H.-K.; Kim, H.; Lee, W., CO2 Hydrate Nucleation Kinetics Enhanced by an

337

Organo-Mineral Complex Formed at the Montmorillonite–Water Interface. Environmental Science &

338

Technology 2015, 49, (2), 1197-1205.

339

17. Lamorena, R. B.; Kyung, D.; Lee, W., Effect of organic matters on CO2 hydrate formation in Ulleung

340

Basin sediment suspensions. Environmental science & technology 2011, 45, (14), 6196-6203.

341

18. Lee, K.; Lee, S.; Lee, W., Stochastic nature of carbon dioxide hydrate induction times in

342

Na-montmorillonite and marine sediment suspensions. Int. J. Greenhouse Gas Control 2013, 14, 15-24.

343

19. Schaumann, G. E.; Thiele-Bruhn, S., Molecular modeling of soil organic matter: squaring the circle?

344

Geoderma 2011, 166, (1), 1-14.

345

20. Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A. E.; Berendsen, H. J., GROMACS: fast,

346

flexible, and free. Journal Of Computational Chemistry 2005, 26, (16), 1701-1718.

347

21. Downs, R. T.; Hall-Wallace, M., The American Mineralogist crystal structure database. American

348

Mineralogist 2003, 88, (1), 247-250.

349

22. Niederer, C.; Goss, K. U., Quantum chemical modeling of humic acid/air equilibrium partitioning of

350

organic vapors. Environmental Science & Technology 2007, 41, (10), 3646-3652.

351

23. Cygan, R. T.; Liang, J.-J.; Kalinichev, A. G., Molecular Models of Hydroxide, Oxyhydroxide, and Clay

352

Phases and the Development of a General Force Field. Journal of Physical Chemistry B 2004, 108, (4),

353

1255-1266. 17

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

354

24. Vanommeslaeghe, K.; Hatcher, E.; Acharya, C.; Kundu, S.; Zhong, S.; Shim, J.; Darian, E.; Guvench, O.;

355

Lopes, P.; Vorobyov, I.; Mackerell, A. D., CHARMM general force field: A force field for drug-like molecules

356

compatible with the CHARMM all-atom additive biological force fields. Journal Of Computational Chemistry

357

2010, 31, (4), 671-690.

358

25. Wright, L. B.; Walsh, T. R., First-principles molecular dynamics simulations of NH 4+ and CH3COO−

359

adsorption at the aqueous quartz interface. Journal of Chemical Physics 2012, 137, (22), 224702.

360

26. Abascal, J.; Sanz, E.; Fernández, R. G.; Vega, C., A potential model for the study of ices and amorphous

361

water: TIP4P/Ice. The Journal of chemical physics 2005, 122, (23), 234511.

362

27. Allen, M. P.; Tildesley, D. J., Computer simulation of liquids. Clarendon Press: Oxford, 1987.

363

28. Darden, T.; York, D.; Pedersen, L., Particle mesh Ewald: An N⋅log (N) method for Ewald sums in large

364

systems. Journal Of Chemical Physics 1993, 98, (12), 10089-10092.

365

29. Van Gunsteren, W.; Berendsen, H., A leap-frog algorithm for stochastic dynamics. Molecular Simulation

366

1988, 1, (3), 173-185.

367

30. Hoover, W. G., Canonical dynamics: equilibrium phase-space distributions. Physical Review A 1985, 31,

368

(3), 1695-1697.

369

31. Parrinello, M.; Rahman, A., Polymorphic transitions in single crystals: A new molecular dynamics

370

method. Journal of Applied physics 1981, 52, (12), 7182-7190.

371

32. Rodger, P.; Forester, T.; Smith, W., Simulations of the methane hydrate/methane gas interface near

372

hydrate forming conditions conditions. Fluid phase equilibria 1996, 116, (1), 326-332.

373

33. Moon, C.; Hawtin, R.; Rodger, P. M., Nucleation and control of clathrate hydrates: insights from

374

simulation. Faraday discussions 2007, 136, 367-382.

375

34. Guo, G.-J.; Zhang, Y.-G.; Liu, C.-J.; Li, K.-H., Using the face-saturated incomplete cage analysis to 18

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Energy & Fuels

376

quantify the cage compositions and cage linking structures of amorphous phase hydrates. Physical Chemistry

377

Chemical Physics 2011, 13, (25), 12048-12057.

378

35. Zhu, X.; Chen, D.; Wu, G., Molecular dynamic simulation of asphaltene co-aggregation with humic acid

379

during oil spill. Chemosphere 2015, 138, 412-421.

380

36. Lamorena, R. B.; Lee, W., Formation of carbon dioxide hydrate in soil and soil mineral suspensions with

381

electrolytes. Environmental science & technology 2008, 42, (8), 2753-2759.

382

37. Guo, G.-J.; Li, M.; Zhang, Y.-G.; Wu, C.-H., Why can water cages adsorb aqueous methane? A potential

383

of mean force calculation on hydrate nucleation mechanisms. Physical Chemistry Chemical Physics 2009, 11,

384

(44), 10427-10437.

385

38. Zhong, D.; Yang, C.; Liu, D.; Wu, Z., Experimental investigation of methane hydrate formation on

386

suspended water droplets. Journal of Crystal Growth 2011, 327, (1), 237-244.

387

39. Walsh, M. R.; Koh, C. A.; Sloan, E. D.; Sum, A. K.; Wu, D. T., Microsecond simulations of spontaneous

388

methane hydrate nucleation and growth. Science 2009, 326, (5956), 1095-1098.

389

40. English, N. J.; Lauricella, M.; Meloni, S., Massively parallel molecular dynamics simulation of formation

390

of clathrate-hydrate precursors at planar water-methane interfaces: Insights into heterogeneous nucleation. The

391

Journal of chemical physics 2014, 140, (20), 204714.

392

41. Guo, G.-J.; Zhang, Y.-G.; Liu, H., Effect of methane adsorption on the lifetime of a dodecahedral water

393

cluster immersed in liquid water: a molecular dynamics study on the hydrate nucleation mechanisms. The

394

Journal of Physical Chemistry C 2007, 111, (6), 2595-2606.

395

42. Liu, C.-J.; Zhang, Z.-C.; Zhang, Z.-G.; Zhang, Y.-G.; Guo, G.-J., Effects of cage type and adsorption face

396

on the cage–methane adsorption interaction: Implications for hydrate nucleation studies. Chemical Physics

397

Letters 2013, 575, 54-58. 19

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

398 399

20

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Energy & Fuels

400

A

C

B

401 402 403 404

Fig. 1 Equilibrium conformations of simulation systems (a) without LHA, (b) with 2% LHA and (c) with 11% LHA. Na+: grey; Si: yellow; O: red ; H: white; Al: blue; Mg: pink; CH4: cygan; LHA: red sticks; H2O: blue lines; hydrogen bonds: blue dashed lines.

405

21

ACS Paragon Plus Environment

Energy & Fuels

406

120 Number density (nm-3)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Page 22 of 32

C

+

Na LHA CH4

80

O in H2O

60 40 20 0

407 408

B

A 100

0

1

2

3

4

5 0

1

2 3 4 Z-coordinate (nm)

5 0

1

2

3

4

Fig. 2 Number density profiles of species in the systems (a) without LHA, (b) with 2% LHA, and (c) with 11% LHA.

22

ACS Paragon Plus Environment

5

Page 23 of 32

409 80

50

B

A

0.3 0.2 0.1

C

40

60

Gas methane (%)

Cage water (%)

0.4

F4ϕ

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Energy & Fuels

40

20

0% LHA 2% LHA 11% LHA

30 20 10

0.0

0

0 0

600

1200

1800

2400

Time (ns)

3000

0

600

1200

1800

2400

Time (ns)

3000

0

600

1200

1800

2400

Time (ns)

410

Fig. 3 Changes in the (a) global F4φ, (b) percentage of cage water, and (c) percentage of gas methane. Curves in (a) and (b) were smoothed using

411

the 30-point adjacent-averaging method.

412 413 414

23

ACS Paragon Plus Environment

3000

Energy & Fuels

415

0.06

0.04

F 4ϕ

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 32

water on top Na-MMT 0.02

0.00

water on bottom Na-MMT

-0.02 1

2

3

4

5

6

7

8

9

10

Slice 416

Fig. 4 F4φ order parameters of different water slices in the systems without LHA

417

(square), with 2% LHA (circle) and with 11% LHA (triangle), respectively.

418 419

24

ACS Paragon Plus Environment

Page 25 of 32

420

A

6

OL-OW OW-OW

B

OL-OW OW-OW

5

5

4

g(r)

4 3

3

2

2

1

1

0 0.0

90

Number of H-bonds

6

g(r)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Energy & Fuels

0.3

0.6

0.9

r (nm)

1.2

1.5

1.8

C

80

2% LHA 11% LHA

70 60 50 40 10

0 0.0

0.3

0.6

0.9

1.2

1.5

r (nm)

1.8

0 0

500

1000

1500

2000

2500

3000

Time (ns)

421

Fig. 5 Radial distribution functions between oxygen in water (OW) and oxygen in LHA (OL) and between OW and OW in the system with (a) 2%

422

and (b) 11% LHA. Number of hydrogen bonds between LHA and water is shown in (c).

423 424

25

ACS Paragon Plus Environment

Energy & Fuels

425

0.5

0.4

0.4

0.3

F4ϕ

0.3

F4ϕ

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Page 26 of 32

0.2

0.2

A

0.0 0

0.1

r=0-0.9 nm r=0-1.0 nm r=0-1.1nm r=0-1.2 nm r>1.2 nm

0.1

500

1000

1500

2000

2500

B

0.0

3000

0

500

1000

1500

2000

2500

Time (ns)

Time (ns) 426

r=0-0.9 nm r=0-1.0 nm r=0-1.1 nm r=0-1.2 nm r>1.2nm

Fig. 6 Local F4φ for systems with (a) 2% LHA and (b) 11% LHA. Curves were smoothed by 30-point adjacent-averaging method.

427 428

26

ACS Paragon Plus Environment

3000

Page 27 of 32

429 430

160

internal

middle

external

120

120

internal

80 40

0-200 ns 200-1000ns 1000-2000 ns 2000-3000 ns

A

0 0.0

external

0.3

0.6

0.9

1.2

1.5

60 0-200 ns 200-1000ns 1000-2000 ns 2000-3000 ns

30 0 0.0

1.8

B 0.3

0.6

0.9

1.2

1.5

1.8

r (nm)

r (nm) 431 432 433 434

middle

90

g(r)

g(r)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Energy & Fuels

Fig. 7 Surface distribution functions for CH4 on (a) LHA monomer and (b) LHA aggregates. Curves were smoothed by 30-point adjacent-averaging method.

27

ACS Paragon Plus Environment

Energy & Fuels

435

A

50

B

70

internal region middle region external region

Number of methane

60

Number of methane

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Page 28 of 32

40 30 20 10

internal region middle region external region

60 50 40 30 20 10

0 0

500

1000

1500

2000

2500

3000

0

1000

1500

2000

2500

Time (ns)

Time (ns) 436 437

500

Fig. 8 Number of methane molecules in different regions from the COG of (a) LHA monomer and (b) LHA aggregates

28

ACS Paragon Plus Environment

3000

Page 29 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Energy & Fuels

B

A

200 ns

140 ns

D

C

540 ns 438 439

1200 ns

Fig. 9 Selected snapshots of the simulation systems with (a-b) 2% and (c-d) 11% LHA. White ball: methane; White line: water; Red line: LHA 29

ACS Paragon Plus Environment

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Number of intermolecular contacts

Energy & Fuels

Page 30 of 32

14 12 10 8 6 4 2 0 0

440 441

500

1000

1500 2000 Time (ns)

2500

3000

Fig. 10 Number of intermolecular contacts between LHA molecules

30

ACS Paragon Plus Environment

Page 31 of 32

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Energy & Fuels

A

B

399 ns

404 ns

E

D

C

443 ns

413 ns 442 443

452 ns

Fig. 11 Snapshots of methane (yellow ball) penetration through hydrate cages after release from the LHA surface. (Red : LHA; White: hydrate cages; Green: specific cage faces. Remainng molecules are not shown to highlight the structure of interest) 31

ACS Paragon Plus Environment

Energy & Fuels

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 32

mineral surface water film (~ 0.4 nm thick) methane

(a) nucleation & growth (b) CH4 adsorption (c) HA aggregation (d) H-bonding (e) HA deformation (f) CH4 desorption (g) cage adsorption (h) cage break (i) cage re-formation (j) CH4 diffusion

a

b

g

pore water

h i

f c e HA

j

f

d water

water film (~ 0.4 nm thick) mineral surface 444 445

Fig. 12 Schema of the processes influencing methane hydrate formation in the clay

446

pores in presence of humic acid (HA)

32

ACS Paragon Plus Environment