Effects of Phosphonate Structures on Brine–Biotite Interactions under

View: ACS ActiveView PDF | PDF | PDF w/ Links | Full Text HTML ... Citation data is made available by participants in Crossref's Cited-by Linking serv...
0 downloads 0 Views 1MB Size
Subscriber access provided by READING UNIV

Article

Effects of Phosphonate Structures on Brine-Biotite Interactions under Subsurface Relevant Conditions Lijie Zhang, Doyoon Kim, and Young-Shin Jun ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.8b00075 • Publication Date (Web): 12 Jul 2018 Downloaded from http://pubs.acs.org on July 20, 2018

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 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 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.

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 29 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

ACS Earth and Space Chemistry

Effects of Phosphonate Structures on Brine−Biotite Interactions under Subsurface Relevant Conditions

Lijie Zhang, Doyoon Kim, and Young-Shin Jun* Department of Energy, Environmental and Chemical Engineering, Washington University, St. Louis, MO 63130

Address: One Brookings Drive, Campus Box 1180 E-mail: [email protected] Phone: (314)935-4539 Fax: (314)935-7211 http://encl.engineering.wustl.edu/ Submitted: July 2018

ACS Earth and Space Chemistry

*Corresponding Author

ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

1

Page 2 of 29

ABSTRACT

2

Phosphonates have been widely used as scale inhibitors in energy-related

3

subsurface operations, where their performance is greatly affected by interactions with

4

rocks and minerals. However, information about commonly used phosphonate scale

5

inhibitor–shale interactions is limited. In this study, using Fe-bearing mica (biotite) as a

6

model phyllosilicate mineral, the effects of three common phosphonates, namely IDMP

7

(iminodi(methylene)phosphonate),

8

DTPMP (diethylenetriaminepenta(methylene)phosphonate), were studied at 95 °C and 102

9

atm of CO2. During the experiments (0−70 h), IDMP remained stable, while NTMP and

10

DTPMP were degraded and released phosphate, formate, and new phosphonates with

11

smaller molecular weights. Due to the differences in chelating capability, IDMP, with the

12

fewest phosphonate functional groups, promoted biotite dissolution mainly through surface

13

complexation, and DTPMP, with the most functional groups, promoted biotite dissolution

14

mainly through aqueous complexation. Furthermore, the presence of phosphonates

15

enhanced secondary precipitation of P, Fe, and Al-bearing minerals, and their phosphonate

16

structures affected the morphologies, phases, and distributions of secondary precipitates.

17

As a result of phosphonate−biotite interactions (mainly due to surface adsorption), the

18

biotite surfaces became much more hydrophilic. This study provides new insight into

19

structure-dependent phosphonate−mineral interactions, and the results have important

20

implications for the safety and efficiency of energy-related subsurface operations.

21

Keywords: Engineered subsurface operations; phosphonate degradation; biotite

22

dissolution; secondary precipitation; wettability alteration

NTMP

(nitrilotris(methylene)phosphonate),

1 ACS Paragon Plus Environment

and

Page 3 of 29 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

ACS Earth and Space Chemistry

23

INTRODUCTION

24

Phosphonates are widely used as scale inhibitors in a variety of technical and

25

industrial applications, such as cooling water systems, textile production, and subsurface

26

oil/gas recovery.1-3 In particular, during unconventional oil/gas recovery from shale

27

reservoirs led by horizontal drilling and hydraulic fracturing, phosphonates have been used

28

to inhibit mineral scale formation that can reduce the porosity and permeability of the

29

fractured rocks.2-3 The average concentrations of scale inhibitors in fracking fluids are

30

around 0.023% by weight percent (~0.5 mM).4 To ensure effective scale inhibition,

31

phosphonates need to be used in concentrations higher than threshold concentrations.5

32

However, it has been reported that phosphonates could have strong chemical reactions with

33

mineral surfaces, resulting in significant mass loss in solutions and affecting the

34

performance of scale inhibition.1, 6 Moreover, phyllosilicates are abundant minerals in shale

35

reservoirs, which are the main subsurface reservoirs for unconventional oil and gas

36

recovery.7-8

37

phosphonate concentrations in the solution, or they can also impact the physico-chemical

38

properties of phyllosilicates. Therefore, it is important to understand the interactions

39

between phosphonates and phyllosilicate minerals.

Thus,

phosphonate−phyllosilicate

interactions

can

affect

available

40

Phosphonates contain C-P and/or C-N bonds and usually have more than one

41

phosphonate functional group (-PO3H2). A wide range of phosphonates with various

42

numbers of phosphonate functional groups has been used in oil/gas production fields.3, 9-10

43

So far, the adsorption of phosphonates onto mineral surfaces and their secondary mineral

44

precipitation have been investigated,6, 11-15 and their adsorption depends on phosphonate

45

structure, concentration, pH, temperature, and mineralogy. For example, the adsorption and

2 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

46

complexation of phosphonates onto goethite surfaces is related to their structure, and

47

phosphonate adsorption decreases as the number of phosphonate functional groups

48

increases.11-12,

49

dissolution are still unclear. In addition, phosphonates are strong chelating agents that can

50

form aqueous complexes with metal ions, with the chelating ability depending on the

51

number of functional groups.16-18 Aqueous complexation can in turn promote mineral

52

dissolution by reducing the apparent saturation in solution and increasing the mineral’s

53

solubility.19 Considering both surface adsorption and aqueous complexation, the structures

54

of phosphonates may affect mineral dissolution and subsequent secondary mineral

55

precipitation. However, information on the relationship between phosphonate structure and

56

mineral dissolution and precipitation is limited.

15

However, the effects of phosphonate surface adsorption on mineral

57

The chemical stability of phosphonates also affects their scale inhibition

58

performance. Under ambient conditions, structure-dependent degradation of phosphonates

59

has been reported.18,

60

temperature and pressure, limited oxygen, and darkness), the degradation of phosphonates

61

with different structures has not been fully investigated. Further, phosphonate degradation

62

can change an original structure by releasing phosphate and formate to form a new

63

phosphonate structure. Then, the newly formed phosphonate and phosphate ions may affect

64

scale inhibition, mineral dissolution, and secondary precipitation in a different way. Hence,

65

the chemical stability of phosphonates may also affect phosphonate−mineral interactions.

66

Chemical reactions of minerals are known to alter their surface wettability, which

67

is an important parameter controlling the distribution and transport of reactive fluids at

68

subsurface sites.21-23 The adsorption of phosphonates with different structures, and the

20

However, under subsurface relevant conditions (e.g., high

3 ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29 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

ACS Earth and Space Chemistry

69

mineral dissolution and subsequent surface precipitation may further affect mineral

70

wettability. To better understand the impacts of phosphonate applications at subsurface

71

sites, we also need to investigate the consequent wettability alterations of minerals by

72

reactions with phosphonates.

73

Therefore, the goal of this study was to examine the interactions between

74

phosphonates with different numbers of phosphonate functional groups and biotite, a

75

model phyllosilicate mineral, under subsurface relevant conditions (95 °C and 102 atm of

76

CO2). The chemical stability of phosphonates, biotite dissolution, secondary mineral

77

precipitation, and mineral wettability alteration were characterized. Then, the mechanisms

78

of phosphonate−biotite interactions were further elucidated by considering phosphonate

79

structure-dependent degradation, surface complexation, aqueous complexation, and

80

extents of secondary precipitation. The new findings from this study will benefit our

81

understanding of the impacts of different phosphonates on the geochemical reactions of

82

minerals and the subsequent wettability alteration by their interactions. The information

83

can help evaluate the applications of phosphonates with different structures and choose

84

more appropriate phosphonates for scale inhibition.

85

EXPERIMENTAL SECTION

86

Mineral and chemicals

87

Biotite, purchased from Ward’s Natural Science, had an elemental composition of

88

K0.91Na0.08Ca0.05Mg1.72Mn0.06Fe1.12Ti0.12Al1.00Si2.98O10(F0.51(OH)0.49)2, as characterized by

89

X-ray fluorescence.24 Biotite flakes were prepared by cleaving along the {001} basal

90

planes and cutting them into 1.0 cm × 0.8 cm rectangles with a thickness of 80 ± 10 m.

91

The flakes were sonicated in acetone, ethanol, and isopropanol successively for 5 min each 4 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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 6 of 29

92

to remove organic matter, and then washed with deionized (DI) water (resistivity > 18.2

93

MΩ·cm, Barnstead Ultrapure Water System, Dubuque, IA) and dried with high purity

94

nitrogen gas. The clean biotite flakes were stored in dust-free tubes for further batch

95

experiments. Three phosphonates with different numbers of phosphonate functional groups,

96

namely iminodi(methylene)phosphonate

97

(NTMP), and diethylenetriaminepenta(methylene)phosphonate (DTPMP), were obtained

98

from Sigma-Aldrich (Technical grade, St. Louis) and used as received (Table 1 and Table

99

S1).

100

(IDMP),

nitrilotris(methylene)phosphonate

Table 1 Abbreviations, chemical formula, and structures of phosphonates used in this study Acronym Chemical Formula

Number of Functional Groups

Structure OH

O P

IDMP

C2H9NO6P2

HO

HO HN

2

O P OH

OH

O P

HO

HO

NTMP

N

C3H12NO9P3

O

3

P OH

HO

P

OH

O OH

O

P

HO

DTPMP

HO

HO N

C9H28N3O15P5

N N

HO

P

OH

OH

O P OH

5

P

O

101

OH

O

P

HO

O

Batch experiments

102

All batch experiments were conducted in a benchtop reactor (Parr Instrument Co.,

103

IL, Figure S1), simulating conditions relevant to engineered subsurface operations where

104

CO2 is injected for enhanced oil/gas production or geologic CO2 sequestration (95 °C and

5 ACS Paragon Plus Environment

Page 7 of 29 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

ACS Earth and Space Chemistry

105

102 atm CO2).24-25 Phosphonate solutions with a concentration of 0.5 mM were prepared

106

in 0.5 M NaCl, mimicking the high salinity of subsurface brine and controlling the ionic

107

strength.26 The acid impurities (mainly HCl, Table S1) in the phosphonates were

108

neutralized and the initial pH of the reaction solutions was adjusted to be 5.6 under ambient

109

conditions, which resulted in an initial pH of around 3.2 at 95 °C under 102 atm CO2 (for

110

a more detailed description of the experimental conditions, refer to S2 in the Supporting

111

Information). As given in Table S2, this in situ pH was close to that of control experiments

112

without adding any phosphonate. The preparation process was open to air, allowing oxygen

113

to be dissolved in the solution. Biotite flakes were placed in PTFE tubes containing 4 mL

114

of the prepared solutions, and the tubes were sealed in the reactor without access to light.

115

Triplicate samples for each batch were prepared, and duplicate batch experiments were

116

conducted. After reaction for a desired elapsed time (3, 8, 22, 44, and 70 h), the experiments

117

were stopped, and the reactor was slowly degassed and cooled within 30 min. The

118

degassing process was the same for all experiments. Thus, the differences among

119

experiments resulted solely from the reactions.

120

Analysis of aqueous solutions

121

Aqueous complexation was analyzed with ultraviolet-visible spectroscopy (UV-

122

Vis, Thermo Scientific Evolution 60S) by scanning the wavelength range of 200–500 nm

123

for the reaction solutions after filtering them through 0.2 µm polypropylene membranes.

124

Then the solutions were acidified in 1% trace metal nitric acid (HNO3). The acidified

125

solutions were measured for aqueous cation concentrations and total phosphorus by

126

inductively coupled plasma-optical emission spectrometry (ICP-OES, PerkinElmer

127

Optima 7300 DV). To analyze the concentrations of phosphate and formate, two

6 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

128

degradation products of phosphonates, ion chromatography (IC, Thermo Scientific Dionex

129

ICS-1600) was used.

130

Characterization of reacted biotite samples

131

The reacted biotite flakes were rinsed with DI water right after they were taken

132

from the reactor, and then dried with high purity nitrogen gas. Their surface morphologies

133

were measured using atomic force microscopy (AFM, Nanoscope V Multimode SPS,

134

Veeco). Areas of 50 µm × 50 µm on the sample surfaces were scanned in contact mode

135

with a nonconductive silicon nitride probe (tip radius of 10 nm, DNP-S10, Bruker) at a rate

136

of 0.999 Hz and a deflection point of 1.975 V. The AFM images were then analyzed with

137

Nanoscope software (v7.20) and representative height mode images were reported.

138

Scanning electron microscopy (SEM, Nova NanoSEM 230) was also used to characterize

139

the reacted biotite basal surfaces. An electron accelerating voltage of 10.00 kV was used

140

for imaging. The elemental composition of reacted biotite was measured by energy

141

dispersive X-ray spectroscopy (EDX).

142

Identification of secondary mineral phases

143

To identify the phases of secondary minerals formed on biotite surfaces and in

144

solutions, high-resolution transmission electron microscopy (HR-TEM, JEOL-2100F) was

145

used. The reacted biotite flake was sonicated in ethanol for 5 min to detach particles from

146

the surface into the ethanol. Then a droplet of the suspension was placed on a

147

Formvar/carbon coated-Cu grid (Electron Microscopy Science, PA) for imaging surface

148

precipitates. To image particles formed in solution, a droplet of the reacted solution before

149

filtration was placed on a TEM grid after centrifugation in DI water for five times, each

150

time for 5 min at 5,000 rpm. To prevent unexpected precipitation during sample drying, 7 ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29 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

ACS Earth and Space Chemistry

151

the centrifugation was performed to remove dissolved ionic species. EDX provided

152

information about elemental compositions. Electron diffraction patterns and lattice fringes

153

were obtained to identify the secondary mineral phases. In addition, the polypropylene

154

membranes used for filtering the solutions were washed with DI water, dried, and examined

155

by SEM for detached biotite layers and particles formed in the solutions.

156

Measurement of water contact angles on biotite samples

157

Right after the reacted biotite flakes were rinsed with DI water and dried with high

158

purity nitrogen gas, and before AFM measurements or secondary mineral phase

159

identification, contact angles were measured using a contact angle analyzer (Surface

160

Electro Optics, Phoenix 300) under ambient conditions. A biotite sample was placed in the

161

sample stage, and then a droplet of DI water was deposited on the basal surface by a syringe

162

needle. The droplet on the mineral surface was imaged and its contact angle was measured

163

(Figure S2). The measurement was repeated at least six times for each experimental

164

condition to obtain the average contact angle.

165

RESULTS AND DISCUSSION

166

Degradation of phosphonates

167

The chemical stabilities of the three phosphonates in the presence of biotite under

168

the high temperature and high pressure conditions were different. Significant degradation

169

of NTMP and DTPMP was observed, as shown by the evolution of phosphate and formate

170

concentrations in Figure 1. It has been reported that, under ambient conditions, reactive

171

oxygen species (ROS) can be generated by interaction between Fe(II) associated with

172

phyllosilicates (e.g., biotite and nontronite) and dissolved molecular oxygen, and the ROS

173

can oxidize organic compounds.27-30 The degradation of NTMP and DTPMP here could 8 ACS Paragon Plus Environment

ACS Earth and Space Chemistry

174

also result from reaction with reactive oxygen species formed through oxygenation of

175

biotite flakes. Changes in phosphate and formate concentrations were not detected in the

176

IDMP reaction system, indicating that IDMP was stable and its degradation was not

177

significant over the course of the reaction. This observation was consistent with a report

178

that, under ambient conditions, NTMP could be degraded to form IDMP with one

179

phosphate released, while the further abiotic degradation of IDMP was considerably slower,

180

due to the difference in structure.31 Specifically, the degradation of phosphonates could

181

start with one-electron abstraction from N, followed by the release of phosphate and

182

formation of formate.20, 32-34 The nitrogen (N) in NTMP and DTPMP connects to three -

183

CH2- groups, while the N in IDMP connects two -CH2- groups and one -H, thus it would

184

be more difficult to abstract one electron from the N in IDMP to initiate its degradation.

185

The concentrations of phosphate and formate normalized by the number of phosphonate

186

functional groups in NTMP (3 groups) and DTPMP (5 groups) were close, suggesting their

187

similar initial susceptibilities to be degraded.

Normalized Conc. (mM)

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 10 of 29

188

0.3 0.2

Formate_NTMP Phosphate_NTMP Formate_DTPMP Phosphate_DTPMP

0.1 0.0 0

20

40

60

Reaction Time (h)

9 ACS Paragon Plus Environment

80

Page 11 of 29 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

ACS Earth and Space Chemistry

189

Figure 1 Concentrations of phosphate and formate from NTMP and DTPMP degradation

190

with biotite, normalized by their numbers of functional groups, at 95 °C and 102 atm CO2.

191

Error bars are the standard deviations of triplicate samples in duplicate experiments. IDMP

192

degradation was not detected.

193

In addition, in both the NTMP and DTPMP systems, the concentrations of

194

phosphate and formate increased quickly within 22 h reaction, and then slowed. Two

195

mechanisms can provide potential reasons for the evolution of phosphonate degradation:

196

First, the release of phosphate and formate by phosphonate degradation forms new

197

phosphonates with fewer functional groups. The newly formed phosphonates (e.g., IDMP

198

formed from NTMP) can be more stable than the original NTMP and DTPMP,18, 31 leading

199

to slower subsequent degradation. Second, the slower phosphonate degradation after

200

reaction times longer than 22 h could result from the lower oxygen level in the reactor,

201

because the reaction consumed oxygen and the reactor was sealed and isolated from air. It

202

is also possible that the produced phosphate could be involved in secondary precipitation

203

with cations released from biotite dissolution,22, 35 thus inhibiting any further increase of

204

its aqueous concentration. However, the increase of formate concentration was also limited

205

after reaction times longer than 22 h, while formate was not supposed to form significant

206

precipitation. For this reason, we conclude that secondary mineral precipitation is not a

207

main mechanism. Note that during the degradation of phosphonates, phosphate is usually

208

first released, and formate is formed by further oxidation or hydrolysis of degradation

209

intermediates.20, 32-34 Considering the limited oxygen level as the reaction went on, it is

210

reasonable that the early degradation product phosphate increased faster than formate.

10 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

211

Phosphonates promoted biotite dissolution

212

The presence of phosphonates affected the evolution of aqueous cation

213

concentrations, but without significant differences among phosphonates, as shown by the

214

ICP-OES results in Figure 2. Regarding interlayer K, phosphonates did not affect its

215

aqueous concentrations compared to those observed for the control system. This result

216

occurred because of similar extents of ion-exchange reactions in the systems, where the

217

initial pH and ionic strength of the reaction solutions were adjusted to be the same.22, 24, 36

218

For Si and Mg, higher concentrations were observed for all the three phosphonate systems

219

than for the control, suggesting that phosphonates promoted biotite dissolution. However,

220

the aqueous concentrations of Al and Fe in phosphonate systems were generally lower than

221

those in the control system. The solubility constants (Ksp) of Fe- and Al-bearing

222

phosphonate/phosphate minerals are on the order of 10-19 or lower, suggesting that they are

223

quite insoluble.37-40 We have also reported that the presence of phosphate and DTPMP

224

promoted precipitation of secondary Fe- and Al-bearing minerals.22, 35 Note that the ICP-

225

OES measured the net concentrations of cations from biotite dissolution and secondary

226

mineral precipitation. Thus, we hypothesized that IDMP and NTMP or released phosphate

227

from NTMP degradation promoted precipitation of secondary Fe- and Al-bearing minerals,

228

as in the DTPMP system, thereby leading to the lower aqueous Al and Fe concentrations.

229

To investigate the mechanisms underlying the indistinguishable net Si and Mg

230

concentrations and the slightly different Al and Fe concentrations, we then considered

231

phosphonate structure-dependent aqueous complexation and surface complexation

232

affecting mineral dissolution and secondary mineral precipitation.

11 ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29 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

ACS Earth and Space Chemistry

233

234

Figure 2 Aqueous concentrations of K, Si, Al, Mg, and Fe from biotite dissolution

235

experiments with 0.5 mM of different phosphonates, at 95 °C and 102 atm CO2. Error bars

236

are the standard deviations of triplicate samples in duplicate experiments.

237

Aqueous complexation was identified by the UV-Vis spectra of the reacted

238

solutions (Figure S3). In both the NTMP and DTPMP reaction solutions, a peak between

239

250–260 nm was observed, which indicated Fe(III)-phosphonate aqueous complexes.35, 41

240

Aqueous complexation between cations and organic ligands could reduce the activity of

241

cations and increase the apparent solubility of minerals.19 Our previous study reported that

242

the aqueous complexation of Fe(III)-DTPMP contributed to promoting biotite

243

dissolution.35 By the same mechanism, aqueous complexation of Fe(III)-NTMP could also

244

promote biotite dissolution. However, in the IDMP system, the peak did not show up

245

clearly, indicating that aqueous complexation was not significant. Figure 3 shows the

246

absorbance of Fe(III)-phosphonate aqueous complexes at 260 nm in the three phosphonate

247

systems over time. Consistent with the literature, DTPMP containing five phosphonate 12 ACS Paragon Plus Environment

ACS Earth and Space Chemistry

248

functional groups had the largest extent of aqueous complexation with Fe(III), then NTMP,

249

and IDMP.42-44 Thus, DTPMP was expected to promote the most biotite dissolution,

250

although this was not the case on the basis of the evolution of measured aqueous cation

251

concentrations (Figure 2). Hence, in addition to aqueous complexation, other mechanisms,

252

such as surface complexation and secondary mineral precipitation, should also be

253

considered and are discussed in later sections.

Absorbance at 260 nm

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 14 of 29

254

2.0 1.5

IDMP NTMP DTPMP

1.0 0.5 0.0 0

20 40 60 Reaction Time (h)

80

255

Figure 3 Evolution of aqueous complexation over time in the phosphonate systems after

256

reactions with biotite at 95 °C and 102 atm of CO2.

257

3.3 Structure-dependent effects of phosphonates on biotite surface morphology

258

alteration and secondary mineral precipitation

259

In addition to the aqueous phase, phosphonates significantly altered the

260

morphology of the basal surfaces of biotite. The alterations were dependent on phosphonate

261

structures, as shown by the AFM images in Figure 4 and SEM images in Figure S4.

262

Compared with the control sample, the presence of IDMP significantly promoted

263

secondary mineral precipitation on biotite basal surfaces, even at the early reaction times

13 ACS Paragon Plus Environment

Page 15 of 29 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

ACS Earth and Space Chemistry

264

of 3 h and 8 h (Figure 4). However, within 8 h reaction, the formation of surface

265

precipitation was less extensive on the NTMP and DTPMP samples than that on the IDMP

266

samples. In the presence of IDMP, the SEM image of a 3 h biotite sample (Figure S4)

267

shows spherical particles in addition to fibrous illite precipitates as we identified in our

268

previous work.45 EDX analyses of the reacted biotite surfaces indicated that, after 3 h

269

reaction, the main elemental composition of biotite remained, while there was a higher

270

phosphorus (P) amount on the IDMP sample surface (Figure S5). Considering that IDMP

271

contains the fewest phosphonate functional groups, the EDX observations suggest that,

272

compared to the other phosphonates, there was higher extent of surface IDMP

273

complexation and adsorption, and consequently more significant surface precipitation. The

274

greater surface adsorption by IDMP than NTMP and DTPMP was consistent with previous

275

studies that phosphonate adsorption decreases with increasing phosphonate functional

276

groups.11-12,

277

subsequent secondary mineral precipitation on mineral surfaces. After 22 h reaction, many

278

precipitates formed on the NTMP and DTPMP samples. The difference in induction times

279

of the surface precipitation in different phosphonate systems might result from different

280

extents of available aqueous complexation and surface complexation. The strongest

281

aqueous complexation in the DTPMP system, indicated by UV-Vis absorption data (Figure

282

3), increased mineral apparent solubility and slowed the surface precipitation. SEM images

283

of 22 h and 44 h phosphonate samples (Figure S4) show cracked biotite layers, bending

284

outwards and detaching from biotite basal surfaces. The bent layers were probably due to

285

adsorption of large phosphonate molecules.35

15

IDMP surface adsorption might promote both biotite dissolution and

14 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

286 287

Figure 4 Height mode AFM images of biotite basal surfaces after reaction at 95 °C and

288

102 atm CO2. The AFM images are 50 µm × 50 µm. To highlight differences, the height

289

scale is 200 nm for the images with blue diamonds and is 60 nm for the other images. The

290

DTPMP images are from reference 35. Copyright 2018 American Chemical Society.

15 ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29 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

ACS Earth and Space Chemistry

291 292

Figure 5 TEM measurements of particles formed (A) on biotite surfaces and (B) in

293

solutions after 70 h reaction at 95 °C and 102 atm CO2. The DTPMP images are adapted

294

from reference 35. Copyright 2018 American Chemical Society.

295

To identify the phases of surface precipitates, particles formed on biotite were

296

detached and measured with TEM (Figure 5A). Lattice fringe information indicated that

297

the particles formed on biotite surface after reaction with IDMP were P, Fe, and Al-rich

298

amorphous phases, while the surface particles in NTMP and DTPMP samples were

299

crystallized. The d-spacing of 2.73 Å in surface particles in the NTMP sample highly

300

suggested the presence of an P, Fe, and Al-bearing phosphate mineral, vauxite 16 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

301

(FeAl2(PO4)2(OH)2).46 In the DTPMP system, the surface particles were reported to be a

302

mixture of gibbsite (Al(OH)3), strengite (FePO4), and berlinite (AlPO4).35 Thermodynamic

303

calculations of mineral saturation states were not possible because detailed thermodynamic

304

data for IDMP, NTMP, and DTPMP are not available, while these can strongly affect the

305

apparent solubility by strong complexation with cations and protons. Still, from the

306

experimental observations, the differences in the secondary mineral phases were probably

307

due to the absence or presence of phosphate: NTMP and DTPMP degraded to release

308

phosphate, while IDMP did not.

309

Detached biotite layers and homogeneously precipitated amorphous Fe- and Al-

310

bearing minerals in solution were reported in the 0.5 mM DTPMP system.35 In the present

311

work, particles in the reacted solutions with IDMP and NTMP were also collected and

312

measured by SEM (Figure 6). After 44 h reaction, in the IDMP system (Figure 6A2),

313

abundant microscale platy particles and aggregates of needle-like particles were observed,

314

and both mainly contained P, Fe, and Al (Figure S7). In the NTMP and DTPMP systems

315

(Figure 6A3 and A4), we did not observe the aggregates of needle-like particles. After 70 h

316

reaction, in the NTMP system (Figure 6B3), in addition to the platy-like microscale

317

particles and tiny spherical particles, aggregates of needle-like particles were formed. This

318

was probably resulted from the accumulation of IDMP as a consequence of NTMP

319

degradation. Therefore, the morphology of secondary precipitates also depends on the

320

structure of phosphonates. Compared with the IDMP and NTMP systems, the amount of

321

homogeneous secondary precipitates was much higher in the DTPMP system (Figure 6).

322

There were also tiny spherical particles, which were not resolved clearly by SEM (Figure

323

6). Further, to identify the tiny spherical particles, TEM was utilized. Similar to those in

17 ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29 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

ACS Earth and Space Chemistry

324

the DTPMP system, the results indicated that the particles in the IDMP and NTMP systems

325

were amorphous phases (Figure 5B), with abundant P, Fe, and Al (Figure S6). Consistently,

326

the precipitation of secondary Fe-bearing phases might lead to lower aqueous complexation

327

between Fe(III) and phosphonates (i.e., NTMP and DTPMP), as shown in Figure 3.

328 329

Figure 6 SEM measurements of samples collected on membranes used for filtering

330

solutions after (A) 44 h and (B) 70 h reactions. The images in the third row were taken

331

from the corresponding samples, indicated by the same color borders. The blue arrows

332

indicate detached biotite layers, orange arrows indicate needle-like aggregates, yellow

333

arrows indicate platy particles, green arrows indicate tiny spherical particles (aggregates).

334

The EDX analyses of these particles are shown in Figure S7. The DTPMP image is adapted

335

from reference 35. Copyright 2018 American Chemical Society.

336

Taken together, phosphonates affected both biotite dissolution and precipitation,

337

with similar net effects, as indicated by the evolution of aqueous cation concentrations. 18 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

338

However, the mechanisms underlying the effects of phosphonates were dependent on

339

phosphonate structures. In the presence of phosphonates, precipitation of P, Fe, and Al-

340

bearing minerals was enhanced, leading to lower aqueous Fe and Al concentrations than in

341

the control experiments. Phosphonates containing more phosphonate functional groups had

342

larger extents of aqueous complexation, promoting biotite dissolution and later

343

heterogeneous surface precipitation. On the other hand, IDMP, containing the fewest

344

functional groups, had greater surface complexation, promoting both biotite dissolution

345

and faster surface precipitation. There was more significant homogeneous precipitation in

346

the DTPMP system than in the IDMP and NTMP systems. The IDMP structure is key to

347

controlling the formation of needle-like aggregated particles and phosphate released from

348

phosphonate degradation is key to controlling the crystallinity of surface precipitates. As a

349

result of degradation, NTMP decomposed to form IDMP and behaved like IDMP. Due to

350

limitations of both the current experimental results and the literature on phosphonate

351

surface complexation, it remains unclear how configurations of phosphonate surface

352

complexation and molecular pathways specifically control the morphologies and phases of

353

secondary precipitates.

354

Enhanced surface wettability by phosphonate adsorption

355

Phosphonate−biotite interactions significantly impacted biotite wettability, which

356

is an important mineral property affecting the flow of reactive fluids during subsurface

357

operations. This change is evident in the alterations of contact angles shown in Figure 7.

358

The contact angles of control samples without any phosphonates were around 28°.

359

However, for biotite samples after reaction with phosphonates, much lower contact angles

360

(around 11°) were observed, showing that phosphonate−biotite reactions enhanced biotite

19 ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29 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

ACS Earth and Space Chemistry

361

surface wettability. Surface roughness enhances wettability by supporting and stabilizing

362

thicker water films on hydrophilic surfaces.21, 47-48 In AFM images (Figure 4), the surface

363

roughness of the reacted biotite samples differed for different phosphonates, particularly

364

after 3 h and 8 h of reaction. In other words, the surface roughness increased with time, in

365

particular for the NTMP and DTPMP samples. However, the contact angles for

366

phosphonate samples did not change over time, which was consistent with our report for

367

DTPMP samples 35. Therefore, surface roughness was not the main contributor to the lower

368

contact angles observed for phosphonate samples. It has been reported that adsorption of

369

phosphate and DTPMP significantly enhances biotite surface wettability by exposing

370

hydrophilic groups, with a stronger effect by DTPMP adsorption.22, 35, 49 However, IDMP

371

was stable and no phosphate from IDMP degradation was detected during the course of

372

reaction. Hence, the enhanced wettability of biotite samples in the IDMP system was

373

attributed mainly to the adsorption of IDMP on biotite. The number of phosphonate

374

functional groups in NTMP is in between those of IDMP and DTPMP, thus we can assume

375

that NTMP adsorption also played a dominant role in enhancing the wettability of biotite.

376

The time-independent contact angles observed for phosphonate samples indicated that a

377

small amount adsorption of phosphonates exposed enough amount of hydrophilic

378

functional groups to enhance the surface wettability, and more adsorption after a longer

379

reaction time did not further change the contact angles.

20 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

Contact angle (Deg.)

ACS Earth and Space Chemistry

380

Page 22 of 29

40 30

Control NTMP

IDMP DTPMP

20 10 0 0

20 40 60 Reaction Time (h)

80

381

Figure 7 Wettability analyses of biotite basal surfaces after reaction with different

382

phosphonates over time at 95 °C and 102 atm CO2. The error bars are standard deviations

383

of six measurements from triplicate samples in duplicate experiments.

384

Interestingly, the effects of phosphonate adsorption on wettability alteration were

385

not dependent on the structure of the phosphonates. The structure-independent

386

enhancement effects by the three phosphonates were consistent with the time-independent

387

effects of phosphonates that adsorption of even a small amount of phosphonate sufficed to

388

make biotite surfaces more hydrophilic. Moreover, although we observed a slightly higher

389

amount of P on biotite surfaces after reaction with IDMP for 3 h (Figure S5), similar

390

amounts of phosphonate functional groups might be exposed on biotite surfaces among the

391

three phosphonates after longer reaction times. The changes in total P concentrations after

392

70 h (Figure S8) were about 80%, 50%, and 30% respectively for IDMP, NTMP, and

393

DTPMP, all corresponding to an approximate 0.8 mM loss of phosphonate functional

394

groups, due to both adsorption and precipitation. Based on the evolution of aqueous cation 21 ACS Paragon Plus Environment

Page 23 of 29 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

ACS Earth and Space Chemistry

395

concentrations, we roughly assumed that the total amounts of secondary precipitates in the

396

three phosphonate systems were very close. Therefore, the numbers of phosphonate

397

functional groups exposed on biotite surfaces after phosphonate adsorption could also be

398

close, resulting in similar extents on wettability alteration.

399

CONCLUSIONS

400

This study presented experimental evidence on the chemical stability of

401

phosphonates and structure-dependent effects of phosphonates in brine−biotite interactions

402

and discussed their dominant mechanisms. IDMP was stable under subsurface relevant

403

conditions, while NTMP and DTPMP degraded, forming new phosphonates and releasing

404

phosphate and formate. The degradation of phosphonates reduces their aqueous

405

concentrations and alters their structures, which may affect their performance in scale

406

inhibition. Furthermore, phosphonates promoted biotite dissolution and enhanced

407

secondary precipitation of P, Fe, and Al-bearing minerals. With increasing numbers of

408

phosphonate functional groups in phosphonate structures, their effects change from surface

409

complexation-dominant to aqueous complexation-dominant. Phosphonates with different

410

structures also affected the morphologies, phases, and distributions of secondary

411

precipitates, which can further influence the porosity and permeability of reservoir rocks

412

to different extents.

413

The wettability of biotite was found to be enhanced by phosphonates, however,

414

their effects were structure-independent. Adsorption of a small amount of phosphonates on

415

biotite surfaces exposed enough hydrophilic hydroxyl groups to enhance wettability. The

416

altered wettability will impact the transport of related reactive fluids in subsurface sites.

22 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

417

The information provided here can be useful to evaluate phosphonate applications and

418

benefit future energy-related engineered subsurface operations.

419

SUPPORTING INFORMATION

420

Details on high temperature and high pressure experimental setup (S1),

421

thermodynamic calculations of pH and in situ high temperature and high pressure pH

422

measurements (S2), contact angle measurement (S3), UV-Vis analyses (S4), SEM and

423

TEM measurements (S5), evolution of total aqueous P concentrations (S6), and accuracy

424

of the experimental apparatus (S7).

425

ACKNOWLEDGMENTS

426

This work was supported by the Center for Nanoscale Controls on Geologic CO2,

427

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

428

Science, Office of Basic Energy Sciences, via Grant DE-AC02-05CH11231. We also thank

429

the National Science Foundation’s CAREER Award (EAR-1057117). The authors

430

acknowledge Washington University’s Nano Research Facility for use of SEM

431

and ICP-OES. We also thank Professor James Ballard for carefully reviewing our

432

manuscript.

23 ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29 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

ACS Earth and Space Chemistry

REFERENCES 1. 2.

3. 4. 5. 6.

7.

8.

9.

10.

11. 12. 13.

14.

15.

Nowack, B., Environmental chemistry of phosphonates. Water Res. 2003, 37 (11), 2533-2546. Ketrane, R.; Saidani, B.; Gil, O.; Leleyter, L.; Baraud, F., Efficiency of five scale inhibitors on calcium carbonate precipitation from hard water: effect of temperature and concentration. Desalination 2009, 249 (3), 1397-1404. Dyer, S.; Anderson, C.; Graham, G., Thermal stability of amine methyl phosphonate scale inhibitors. J. Pet. Sci. Eng. 2004, 43 (3), 259-270. Tollefson, J., Secrets of fracking fluids pave way for cleaner recipe. Nature 2013, 501 (7466), 146. Lake, L. W., Enhanced oil recovery. Prentice Hall: Englewood Cliffs, NJ, 1989. Yan, F.; Zhang, F.; Bhandari, N.; Wang, L.; Dai, Z.; Zhang, Z.; Liu, Y.; Ruan, G.; Kan, A.; Tomson, M., Adsorption and precipitation of scale inhibitors on shale formations. J. Pet. Sci. Eng. 2015, 136, 32-40. Bourg, I. C., Sealing shales versus brittle shales: a sharp threshold in the material properties and energy technology uses of fine-grained sedimentary rocks. Environ. Sci. Technol. Lett. 2015, 2 (10), 255-259. Kargbo, D. M.; Wilhelm, R. G.; Campbell, D. J., Natural gas plays in the Marcellus Shale: Challenges and potential opportunities. Environ. Sci. Technol. 2010, 44 (15), 5679-5684. Zhang, B.; Zhang, L.; Li, F.; Hu, W.; Hannam, P. M., Testing the formation of Ca– phosphonate precipitates and evaluating the anionic polymers as Ca–phosphonate precipitates and CaCO3 scale inhibitor in simulated cooling water. Corros. Sci. 2010, 52 (12), 3883-3890. Zhang, P.; Shen, D.; Ruan, G.; Kan, A. T.; Tomson, M. B., Mechanistic understanding of calcium–phosphonate solid dissolution and scale inhibitor return behavior in oilfield reservoir: formation of middle phase. Phys. Chem. Chem. Phys. 2016, 18 (31), 21458-21468. Nowack, B.; Stone, A. T., Adsorption of phosphonates onto the goethite–water interface. J. Colloid Interface Sci. 1999, 214 (1), 20-30. Nowack, B.; Stone, A. T., Competitive adsorption of phosphate and phosphonates onto goethite. Water Res. 2006, 40 (11), 2201-2209. Jordan, M.; Sorbie, K.; Ping, J.; Yuan, M. D.; Todd, A.; Hourston, K. In Phosphonate Scale Inhibitor Adsorption/Desorption and the Potential for Formation Damage in Reconditioned Field Core, SPE Formation Damage Control Symposium, Society of Petroleum Engineers: 1994. Jordan, M.; Sorbie, K.; Chen, P.; Armitage, P.; Hammond, P.; Taylor, K., The design of polymer and phosphonate scale inhibitor precipitation treatments and the Importance of precipitate solubility in extending squeeze lifetime. In International Symposium on Oilfield Chemistry, Society of Petroleum Engineers: 1997. Rott, E.; Minke, R.; Steinmetz, H., Removal of phosphorus from phosphonate-loaded industrial wastewaters via precipitation/flocculation. Journal of Water Process Engineering 2017, 17, 188-196. 24 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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

16.

17.

18.

19.

20. 21.

22.

23. 24.

25.

26.

27.

28.

29. 30.

Tomson, M.; Kan, A.; Oddo, J., Acid/base and metal complex solution chemistry of the polyphosphonate DTPMP versus temperature and ionic strength. Langmuir 1994, 10 (5), 1442-1449. Sawada, K.; Miyagawa, T.; Sakaguchi, T.; Doi, K., Structure and thermodynamic properties of aminopoly-phosphonate complexes of the alkaline-earth metal lons. J. Chem. Soc., Dalton Trans. 1993, (24), 3777-3784. Nowack, B.; Stone, A. T., Degradation of nitrilotris (methylenephosphonic acid) and related (amino) phosphonate chelating agents in the presence of manganese and molecular oxygen. Environ. Sci. Technol. 2000, 34 (22), 4759-4765. Yang, Y.; Ronzio, C.; Jun, Y.-S., The effects of initial acetate concentration on CO2– brine–anorthite interactions under geologic CO2 sequestration conditions. Energy Environ. Sci. 2011, 4 (11), 4596-4606. Schowanek, D.; Verstraete, W., Hydrolysis and free radial mediated degradation of phosphonates. J. Environ. Qual. 1991, 20 (4), 769-776. Zhang, L.; Kim, Y.; Jung, H.; Wan, J.; Jun, Y.-S., Effects of Salinity-Induced Chemical Reactions on Biotite Wettability Changes under Geologic CO2 Sequestration Conditions. Environ. Sci. Technol. Lett. 2016, 3 (3), 92-97. Zhang, L.; Kim, D.; Kim, Y.; Wan, J.; Jun, Y.-S., Effects of Phosphate on Biotite Dissolution and Secondary Precipitation under Conditions Relevant to Engineered Subsurface Processes. Phys. Chem. Chem. Phys. 2017, 19, 29895-29904. Rezaee, R.; Saeedi, A.; Iglauer, S.; Evans, B., Shale alteration after exposure to supercritical CO2. Int. J. Greenhouse Gas Control 2017, 62, 91-99. Zhang, L.; Jun, Y.-S., Distinctive Reactivities at Biotite Edge and Basal Planes in the Presence of Organic Ligands: Implications for Organic-Rich Geologic CO2 Sequestration. Environ. Sci. Technol. 2015, 49 (16), 10217-10225. Jun, Y.-S.; Zhang, L.; Min, Y.; Li, Q., Nanoscale Chemical Processes Affecting Storage Capacities and Seals during Geologic CO2 Sequestration. Acc. Chem. Res. 2017, 50 (7), 1521-1529. Kharaka, Y. K.; Thordsen, J. J.; Hovorka, S. D.; Seay Nance, H.; Cole, D. R.; Phelps, T. J.; Knauss, K. G., Potential environmental issues of CO2 storage in deep saline aquifers: Geochemical results from the Frio-I Brine Pilot test, Texas, USA. Appl. Geochem. 2009, 24 (6), 1106-1112. Yuan, S.; Liu, X.; Liao, W.; Zhang, P.; Wang, X.; Tong, M., Mechanisms of electron transfer from structrual Fe (II) in reduced nontronite to oxygen for production of hydroxyl radicals. Geochim. Cosmochim. Acta 2018, 223, 422-436. Tong, M.; Yuan, S.; Ma, S.; Jin, M.; Liu, D.; Cheng, D.; Liu, X.; Gan, Y.; Wang, Y., Production of abundant hydroxyl radicals from oxygenation of subsurface sediments. Environ. Sci. Technol. 2015, 50 (1), 214-221. Zeng, Q.; Dong, H.; Wang, X.; Yu, T.; Cui, W., Degradation of 1, 4-dioxane by hydroxyl radicals produced from clay minerals. J. Hazard. Mater. 2017, 331, 88-98. Liu, X.; Yuan, S.; Tong, M.; Liu, D., Oxidation of trichloroethylene by the hydroxyl radicals produced from oxygenation of reduced nontronite. Water Res. 2017, 113, 7279.

25 ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29 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

ACS Earth and Space Chemistry

31.

32.

33.

34.

35.

36. 37.

38. 39. 40.

41.

42.

43.

44.

45.

46.

Steber, J.; Wierich, P., Properties of aminotris (methylenephosphonate) affecting its environmental fate: degradability, sludge adsorption, mobility in soils, and bioconcentration. Chemosphere 1987, 16 (6), 1323-1337. Jaisi, D. P.; Li, H.; Wallace, A. F.; Paudel, P.; Sun, M.; Balakrishna, A.; Lerch, R. N., Mechanisms of Bond Cleavage during Manganese Oxide and UV Degradation of Glyphosate: Results from Phosphate Oxygen Isotopes and Molecular Simulations. J. Agric. Food Chem. 2016, 64 (45), 8474-8482. Aguila, A.; O'Shea, K. E.; Tobien, T.; Asmus, K.-D., Reactions of hydroxyl radical with dimethyl methylphosphonate and diethyl methylphosphonate. A fundamental mechanistic study. J. Phys. Chem. A 2001, 105 (33), 7834-7839. O'Shea, K. E.; Beightol, S.; Garcia, I.; Aguilar, M.; Kalen, D. V.; Cooper, W. J., Photocatalytic decomposition of organophosphonates in irradiated TiO2 suspensions. J. Photochem. Photobiol., A 1997, 107 (1-3), 221-226. Zhang, L.; Kim, D.; Jun, Y.-S., The Effects of Phosphonate-Based Scale Inhibitor on Brine−Biotite Interactions under Subsurface Conditions. Environ. Sci. Technol. 2018, 52 (10), 6042-6049. Sparks, D. L., Environmental soil chemistry. Elsevier Science: London, UK, 2003. Friedfeld, S. J.; He, S.; Tomson, M. B., The temperature and ionic strength dependence of the solubility product constant of ferrous phosphonate. Langmuir 1998, 14 (13), 3698-3703. Iuliano, M.; Ciavatta, L.; De Tommaso, G., On the solubility constant of strengite. Soil Sci. Soc. Am. J. 2007, 71 (4), 1137-1140. Ferguson, J. F.; King, T., A model for aluminum phosphate precipitation. J. Water Pollut. Control Fed. 1977, 646-658. Wang, X.; Hu, Y.; Tang, Y.; Yang, P.; Feng, X.; Xu, W.; Zhu, M., Phosphate and phytate adsorption and precipitation on ferrihydrite surfaces. Environmental Science: Nano 2017, 4 (11), 2193-2204. Matthus, E.; De Oude, N.; Bolte, M.; Lemaire, J., Photodegradation of ferric ethylenediaminetetra (methylenephosphonic acid)(EDTMP) in aqueous solution. Water Res. 1989, 23 (7), 845-851. Martell, A. E.; Smith, R. M.; Motekaitis, R., NIST critically selected stability constants of metal complexes database. National Institute of Science and Technology: 1998. Deluchat, V.; Bollinger, J.-C.; Serpaud, B.; Caullet, C., Divalent cations speciation with three phosphonate ligands in the pH-range of natural waters. Talanta 1997, 44 (5), 897-907. Popov, K.; Rönkkömäki, H.; Lajunen, L. H., Critical evaluation of stability constants of phosphonic acids (IUPAC technical report). Pure Appl. Chem. 2001, 73 (10), 1641-1677. Hu, Y.; Ray, J. R.; Jun, Y. S., Biotite-brine interactions under acidic hydrothermal conditions: fibrous illite, goethite, and kaolinite formation and biotite surface cracking. Environ. Sci. Technol. 2011, 45 (14), 6175-6180. Chichagov, A. V., Information-calculating system on crystal structure data of minerals (MINCRYST). In Kristallografiya, 1990; Vol. 35, pp 610-616. 26 ACS Paragon Plus Environment

ACS Earth and Space Chemistry 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.

48. 49.

Kim, T. W.; Tokunaga, T. K.; Bargar, J. R.; Latimer, M. J.; Webb, S. M., Brine film thicknesses on mica surfaces under geologic CO2 sequestration conditions and controlled capillary pressures. Water Resour. Res. 2013, 49 (8), 5071-5076. Nakae, H.; Inui, R.; Hirata, Y.; Saito, H., Effects of surface roughness on wettability. Acta Mater. 1998, 46 (7), 2313-2318. Plumridge, T. H.; Steel, G.; Waigh, R. D., Geometry-based simulation of the hydration of small molecules. PhysChemComm 2000, 3 (8), 36-41.

27 ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29 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

ACS Earth and Space Chemistry

Table of Contents

28 ACS Paragon Plus Environment