Study of Iron-Bearing Dolomite Dissolution at ... - ACS Publications

Aug 2, 2017 - Fe hydroxides.5 In the mining industry, oxidation of sulfur .... Chemical Compositions of Dolomite Determined by Triacid Attack (HCl, HN...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Rochester | River Campus & Miner Libraries

Article

Study of iron-bearing dolomite dissolution at various temperatures: Evidence for the formation of secondary nanocrystalline iron-rich phases on the dolomite surface Mathieu Debure, Pascal Andreazza, Aurélien Canizarès, Sylvain Grangeon, Catherine Lerouge, Paul Mack, Benoit Madé, Patrick Simon, Emmanuel Véron, Fabienne Warmont, and Marylene Vayer ACS Earth Space Chem., Just Accepted Manuscript • DOI: 10.1021/ acsearthspacechem.7b00073 • Publication Date (Web): 02 Aug 2017 Downloaded from http://pubs.acs.org on August 12, 2017

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

ACS Earth and Space Chemistry 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 28

ACS Earth and Space Chemistry 1

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

Study of iron-bearing dolomite dissolution at various

2

temperatures: Evidence for the formation of secondary

3

nanocrystalline iron-rich phases on the dolomite surface

4

Mathieu Debure(1,*), Pascal Andreazza(2), Aurélien Canizarès(3), Sylvain Grangeon(1),

5

Catherine Lerouge(1), Paul Mack(4), Benoît Madé(5), Patrick Simon(3), Emmanuel

6

Veron(3), Fabienne Warmont(2), Marylène Vayer(2)

7

(1) BRGM – French Geological Survey, Water, Environment and Ecotechnologies Division, Storage

8

and Deep Geological Settings Unit; 3, avenue Claude Guillemin - BP 36009, 45060 Orléans Cedex 2 -

9

France.

10

(2) ICMN - UMR 7374 CNRS - Université d’Orléans 1b rue de la Férollerie CS 40059, 45071 Orléans

11

cedex 2, France.

12

(3) CEMHTI, UPR 3079 CNRS, - Université d’Orléans, CS90055, 45071 Orléans, France.

13

(4) Thermo Fisher Scientific, Imberhorne Lane, East Grinstead, West Sussex, RH19 1UB UK

14

(5) Andra, 1 – 7 rue Jean Monnet, 92298 Châtenay-Malabry, France

15

Abstract

16

We investigated the dissolution of a natural Fe-containing dolomite [Ca1.003Mg0.972Fe0.024Mn0.002(CO3)2]

17

under acidic conditions (pH 3-5.5) with atomic force microscopy (AFM) at 20 °C, and with batch

18

dissolution experiments at 80 °C. Dolomite dissolution proceeded by two identified mechanisms:

19

removal of dolomite layers through spreading and coalescence of etch pits nucleated at defect points,

20

and stepped retreat from surface edges. The dolomite dissolution rate increased when pH decreased

21

(from 0.410 nm s-1 at pH 3 to 0.035 nm s-1 at pH 5). Rates calculated from edge retreat (vedges) and

22

from etch-pit spreading rates (vsum) were consistent; the etch-pit digging rate was almost 10 times

23

slower than its spreading rate. Nanocrystalline secondary phases precipitated in the course of

24

dolomite dissolution at pH 3 and 80 °C were identified as (nano)hematite, ferrihydrite and an ankerite

25

like mineral using X-ray diffraction, transmission electron microscopy, MicroRaman and X-ray

26

photoelectron spectrometry. In addition, Mg enrichment of the surface layer was observed at 80 °C.

27

The characterizations performed at a nanocrystalline scale highlighted the role played by impurities in

28

the dolomite dissolution/precipitation scheme and evidenced that the preponderant mechanism

29

explaining the incongruent dolomite dissolution is secondary phase precipitation from major and minor

30

elements initially present in the pristine dolomite.

*

Corresponding author: BRGM – French Geological Survey, Water, Environment and Ecotechnologies Division, Storage and Deep Geological Settings Unit, 3, avenue Claude Guillemin - BP 36009, 45060 Orléans Cedex 2 France. Tel: +33 238 643 177; Fax: +33 238 644 797; E-mail: [email protected]

ACS Paragon Plus Environment

ACS Earth and Space Chemistry

Page 2 of 28 2

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

31

Keywords: Dolomite, nanocrystalline secondary phases, Fe oxyhydroxide, Ca-Mg-Fe carbonates, in

32

situ AFM experiments.

33

Graphical abstract for TOC

34

Fe

Mg Ca

Fe oxyhydroxides

Fe2++ O2

Fe Ca

Ca

Mg Mg Ca

Fe Fe3+ Edge retreat Mg Mg Mg 2+ Ca Fe2+ Ca Fe Mg Mg Edge pit Ca Fe2+ Ca Fe2+ spreading

35

Ca-Fe-Mg carbonate

Fe3+

Fe2+ Mg Mg Mg Ca Fe2+ Ca Fe2+ Mg Mg Ca Fe2+ Ca Fe2+

Edge pit digging

36

ACS Paragon Plus Environment

Page 3 of 28

ACS Earth and Space Chemistry 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 49 50 51 52 53 54 55 56 57 58 59 60

37

1. Introduction

38

Carbonate minerals play an important pH-buffering role1 in both natural and engineered systems.

39

Understanding and quantifying calcite (CaCO3) and dolomite (CaMg(CO3)2) dissolution and

40

precipitation mechanisms as a function of environmental conditions is necessary if we are to make

41

predictions about the evolution of natural systems containing carbonates such as aquifers in karst ,

42

marine water chemistry, and the CO2 global cycle , and engineered systems such as in CO2

43

sequestration or radioactive waste geological disposal. In the oil and geothermal industries,

44

hydrochloric acid injections in wells promote carbonate dissolution in reservoirs and improve

45

performance related to porosity and permeability6-8. In CO2 sequestration projects, the high amount of

46

injected CO2 reduces the pH and dissolves carbonate minerals in carbonate reservoirs9. Conversely, a

47

CO2 injection in basaltic rock leads to the precipitation of Ca-Mg-Fe carbonates and Fe hydroxides . In

48

the mining industry, oxidation of sulfur minerals due to natural conditions and/or anthropogenic

49

disruption (excavation, pumping) cause acid mine drainage and metal release (Pb, As, Hg, etc.) into

50

the environment, acidify the soils and consequently induce carbonate dissolution if those are present

51

in the soil10. The behavior of carbonates in acidic fields is also studied for desalination and water

52

delivery purposes. Inherently, desalination treatment leads to demineralization. So carbonate

53

dissolution is investigated as a way of remineralizing water to protect the distribution system from

54

corrosion damage11. In nuclear waste management, transitional acidic events are expected to occur

55

because of the oxidation of pyrite in the clay host-rocks , and carbonate mineral dissolution is

56

expected to buffer the pH in the host-rock pore water. In these systems, dolomite is often present in

57

significant amounts, and participates in controlling pore water composition and pH through dissolution

58

and precipitation processes.

59

The dolomite dissolution rate depends on several parameters: saturation state, pH, temperature, ionic

60

strength, ion concentrations, and hydrodynamic conditions

61

extensively investigated over the years in environmental fields3. The stoichiometry of dolomite

62

dissolution as a function of conditions is still under debate as well as the mechanisms involved in fluid

63

rock interaction

. Dolomite incongruent dissolution has been explained by the lower hydration

64

energy of Ca

2+

compared to Mg2+, which leads to a lower stability of Ca2+ at the dolomite/water

65

interface and to a Ca-depleted layer at the dolomite surface

66

reported the formation of a MgCO3 phase at the dolomite surface involving a dissolution/precipitation

67

scheme as evidenced for wollastonite21 and for other minerals20,

68

investigated dolomite surface reactivity exposed to a continuous flow of supersaturated solutions. The

69

precipitation of dolomite layers without any other secondary phase precipitation occurred only from a

70

fluid highly oversaturated with respect to magnesite, calcite and dolomite. Indeed, one might expect

71

that other Mg or Ca-bearing phases may have formed but they were not observed. Such contradictory

72

results underline the complexity of dolomite dissolution and the complex reactions occurring at the

73

dolomite surface. Moreover, there is a lack in the understanding of trace-element behavior during

74

natural dolomite dissolution, and in particular of Fe behavior, although this is a common trace-element

75

incorporated in the dolomite lattice.

2

3-5

5

12

13-15

. Most of these parameters have been

14-23

13, 14, 19, 22

. More recently, Urosevic, et al.

ACS Paragon Plus Environment

24

. Berninger, et al.

25

17

lately

ACS Earth and Space Chemistry

Page 4 of 28 4

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

76

Overall, there has been to our knowledge no systematic characterization of the secondary phases

77

formed in batch reaction experiments15, 16, 26. Studies focusing on nanoscale resolution have provided

78

new insights into the kinetics and mechanisms of dolomite dissolution and precipitation on almost pure

79

dolomite without discussing the role of trace elements. Those works found evidence to explain

80

dolomite dissolution by either preferential release of Ca over Mg or secondary phase precipitation but

81

did not establish the role of impurities inside the structure, or a combination of several mechanisms. It

82

is usually admitted in literature that a concentration gradient appears at the mineral surface during its

83

dissolution that explains the secondary phase precipitation. However, other even minor mechanisms

84

such as preferential release of an ion can occur. Thus, nanoscale characterizations are required to

85

clearly identify the products of dolomite dissolution to infer the reaction mechanism.

86

This study examined the dissolution of a Fe-bearing dolomite under various pH (3 to 5.5) and

87

temperature (20 and 80 °C) conditions, and with reaction time ranging from 30 minutes to 21 days. We

88

studied the secondary phase precipitation that occurred during the experiments. The objective was to

89

better understand the congruence or incongruence observed during dolomite dissolution, and to

90

determine quantitatively the crystal-dissolution rates, the reaction path, how Fe influences the phases

91

formed in situ and how these impact the dolomite dissolution and precipitation scheme.

92

2. Materials and methods

93

2.1. Natural dolomite

94

Our dolomite sample came from the French Geological Survey (BRGM) collection (sampled in Ariège,

95

southern France). Chemical compositions, determined with a tri-acid attack (HCl, HNO3, HF) on bulk

96

powder and by twenty spot analyses of the crystal surface using an electron microprobe, are in good

97

agreement (Table 1). The trace element composition is shown in Table S1 of the supporting

98

information. The dolomite was cleaved along the (104) surface, with this atomic plane being indexed

99

relative to the model proposed by Anthony, et al.

27

. A smooth, bright plane was obtained.

100

Table 1. Chemical compositions of dolomite determined by tri-acid attack (HCl, HNO3, HF) on bulk powder

101

and by spot analyses using an electron microprobe. The data is given in atoms per formula unit Ca (Fe,

102

Mg, Mn)(CO3)2).

Analysis technique

Ca

Fe

Mg

Mn

Tri-acid

0.972 ± 0.049 0.032 ± 0.002 0.984 ± 0.049 0.002 ± 0.002

Microprobe

1.002 ± 0.023 0.024 ± 0.003 0.972 ± 0.024 0.002 ± 0.002

103

2.2. Experiments

104

Dolomite dissolution was investigated using two types of experiment (Table 2). Atomic force

105

microscopy (AFM) experiments were used to follow the dissolution mechanism in situ. These

106

experiments were performed at ambient temperature, at three pHs, and were short. Batch experiments

107

were performed at 80 °C, for longer contact times (30 min, 1 day and 21 days), and at pH80°C ~3, for

108

which the highest dolomite surface reactivity was observed. The temperature of 80 °C was chosen

ACS Paragon Plus Environment

Page 5 of 28

ACS Earth and Space Chemistry 5

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

109 110

based on previous experimental studies on dolomite dissolution (e.g. Gautelier, et al. al.

28

and Berninger, et al.

15

, Pokrovsky, et

25

) and temperatures attained in an oil reservoir or CO2 injection well29, 30.

111

Table 2. Experimental conditions of the dolomite dissolution tests under various environmental

112

conditions (the name of the experiment includes type of experiment, pH and time expressed in hours).

Time

Experiment

pHsolution

(days:hours:min)

Surface 2

(mm )

Volume of solution

Temperature (°C)

(mL)

Environment

AFM-pH3-2.4

00:02:25

3.0 ± 0.1

5.0 ± 0.04

0.7 ± 0.01

20 ± 1

HCl

AFM-pH5-3.1

00:03:07

4.9 ± 0.1

3.9 ± 0.04

0.7 ± 0.01

20 ± 1

HCl

AFM-pH5-20

00:20:00

5.5 ± 0.1

5.6 ± 0.04

0.7 ± 0.01

20 ± 1

Pure water

Batch-pH3-0.5

00:00:30

2.8 ± 0.1

120.5 ± 0.04 44.8 ± 0.06

80 ± 1

HCl

Batch-pH3-24.3

01:00:20

2.8 ± 0.1

36.2 ± 0.04

45.0 ± 0.06

80 ± 1

HCl

Batch-pH3-504

21:00:00

2.9 ± 0.1

23.2 ± 0.04

45.0 ± 0.06

80 ± 1

HCl

113

2.2.1.

114

Atomic force microscopy experiments were carried out in Teflon cells with an inner volume of 0.7 mL

115

at room temperature (20 °C). Each solution consisted of ultrapure water (resistivity = 18.2 mΩ cm)

116

prepared immediately before the experiment to avoid equilibration with the ambient atmosphere. The

117

amount of carbonate and bicarbonate ions in solution can thus be considered as negligible initially and

118

as only coming from dolomite dissolution together with Ca, Mg and Fe. The solution pH was adjusted

119

with HCl. The (104) dolomite surface was put in contact with the solution for different time periods

120

depending on the pH. The surface exposed to the solution was around 5 mm (Table 2).

121

Laboratory observations and measurements used an AFM equipped with a Molecular Imaging Pico+

122

fluid cell, working in contact mode under ambient conditions (T = 20 °C). The AFM images were

123

collected using Si3N4 contact tips and analyzed with the Gwyddion software (Version 2.44).

124

Measurements of step-retreat velocity (or etch-pit spreading rate, vsum) were made from sequential

125

images scanned in the same direction. One type of step has an acute angle of 78° (-) because the

126

step-edge atoms of the upper terrace overhang those below the step edge31. The other type of step is

127

exposed opposite and has an obtuse angle of 102° (+). In all etch pits, the equivalent steps are

128

adjacent, whereas nonequivalent steps are parallel. The retreat velocity was given by vsum = (v+ + v-),

129

where v+ and v- represent the retreat velocities of + and - steps, respectively17, 32, 33. These two types

130

of steps retreat to opposite directions with different velocities as the mineral surface dissolves, and

131

hence, the dissolution of dolomite is anisotropic. The vsum values were calculated by measuring the

132

length increase per unit time between opposite parallel steps in sequential images. Overall dissolution

133

rates, RAFM (in mol cm s ), were calculated according to equation 1 given by Shiraki, et al. :

134

 =

Atomic Force Microscopy

2

-2

-1

31





∆

  

ACS Paragon Plus Environment

(1)

ACS Earth and Space Chemistry

Page 6 of 28 6

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

3

-1

135

where Vdol is the molar volume of dolomite (64.34 cm mol ), Δh is the difference in height along the

136

step orientation in sequential images after their time duration, and L is the length between the center

137

of the etch pit and its edge.

138

The macroscopic dolomite dissolution rates (RMac) were calculated when possible with equation 2:  =

139

 .

(2)

140

where [Ca] is the final Ca concentration in mol L-1, V the volume of the solution in L, A the surface area

141

exposed to the solution in m , and t the total duration time of the experiment.

142

2.2.2.

143

Batch tests were performed in 50 mL PTFE reactors at 80 °C. The solutions were prepared following

144

the same protocol as AFM experiments. The amounts of solution and surfaces of dolomite in contact

145

with the solutions are given in table 2. At the end of the experiments, the reactors were opened, the

146

solutions were filtered through 0.1 µm PVDF filters, and then pH and alkalinity were measured if the

147

pH during the experiment got higher than 4.5. The monoliths were removed from the reactor, carefully

148

rinsed with deionized water and dried for one day at room temperature before characterization.

149

2.3. Solution analyses

150

The pH measurement of batch solutions was carried out with a Mettler Toledo Seven multi pH meter

151

using NIST 1.7, 4, 7 and 9 buffers. Cations (Ca2+, Mg2+ and total Fe) were analyzed in solution by

152

inductively coupled plasma atomic emission spectroscopy (ICP-AES, Jobin Yvon), or by mass

153

spectroscopy (ICP-MS, Thermo Fisher Scientific). Chloride anions were analyzed by ionic

154

chromatography (HPLC, Dionex). Element concentrations in solution were determined with an

155

uncertainty of 3%. Alkalinity was measured using a Titrando 905 and a Dosino 800 equipped with a

156

5 mL syringe (Metrohm) to inject the HCl solution (10-3 mol L-1) into the sample. The alkalinity was then

157

calculated with the Gran method described by Neal 34.

158

2.4. Solid characterizations

159

Scanning electron microscopy (SEM) observations were done with a Low Vacuum – Field Emission

160

Scanning Electron Microscope (LV-FE-SEM) TESCAN Mira3XMU (TESCAN, Brno, Czech Republic)

161

coupled with an EDAX TEAM system equipped with a SDD detector APOLLO XPP. Detectors used for

162

SEM observations were Dual BSE/CL Tescan detectors for backscattered electrons (BSE).

163

Observations were performed at HV = 25 kV.

164

Transmission electron microscopy (TEM) observations were done with a Philips CM20 operated at

165

200 kV. Prior to observation, samples were scratched with a PTFE spatula and then dry-deposited on

166

a carbon-coated copper grid.

2

Batch experiments

ACS Paragon Plus Environment

Page 7 of 28

ACS Earth and Space Chemistry 7

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

167

Spot analyses of initial dolomite were performed on monoliths covered with a carbon coating, using a

168

CAMEBAX SX Five electron microprobe with an accelerating voltage of 15 kV, a beam current of

169

12 nA, and a 1 µm beam diameter. Peak and background counting times were 10 s for Ca, Mg, Mn

170

and Fe. Detection limits were 20 ppm for Ca, 8 ppm for Mg, 44 ppm for Mn and 40 ppm for Fe.

171

Standards used included both well-defined natural minerals and synthetic oxides. Matrix corrections

172

were done with a ZAF software program .

173

The Raman spectra were recorded at room temperature using a Renishaw InVia Reflex Raman

174

microscope to characterize dolomite and the secondary phases formed after reaction. A 633 nm He–

175

Ne laser was initially employed and Raman spectra were acquired with an acquisition time of 1 to 30 s

176

(depending on the phase fluorescence) for a relative wavenumber ranging from 100 cm-1 to 1400 cm-1.

177

Measurements were done with a 20 mW laser power using the Renishaw StreamLine

178

In this imaging mode, the laser focuses as a line on the sample via a cylindrical lens, decreasing the

179

power density of laser to around 200 µW on each analyzed point on the sample and avoiding phase

180

transformation by laser heating as observed by El Mendili et al. on maghemite

35

TM

36, 37

38

configuration.

and potentially

181

observed by de Faria, et al.

182

were used.

183

X-ray diffraction was performed at fixed incidence of 5° to enhance the signal arising from the surface

184

of the sample. Using a D8 ADVANCE diffractometer with Cu Kα radiation (8041 eV) at 40 mA and

185

40 kV, we selected step sizes of 0.04° with time-by-step fixed at 40 s over a 2θ range of 13° to 60°.

186

The small angle X-ray scattering (SAXS) experiments were performed on the “Xeuss” Xenocs X-ray

187

scattering set-up with a monochromatic X-ray beam at 8041 eV . X-ray scattering measurements

188

were performed versus the components of the wave-vector transfer (transferred momentum) q = ki - kf,

189

defined by the incident ki and the scattered kf wave vectors. To enhance the scattering signal coming

190

from the dolomite surface, an X-ray grazing incidence angle of 0.22° was selected. The two-

191

dimensional SAXS pattern was recorded using a high sensitivity 2D detection (Pilatus 300K hybrid

192

pixel detector) placed at a distance of 2.52 m from the sample, perpendicular to the edge. The

193

scattered beam intensities were collected in a range between [0; 3.5 nm ], by averaging ten frames to

194

improve the counting statistics. To extract morphological information, we analyzed two sections of 2D

195

patterns, or fully integrated the scattering signal. From these 1D data, the contribution of the form

196

factor could be extracted with a size evaluation

197

X-ray photoelectron spectra (XPS) were recorded on a Thermo Fisher Scientific ESCALAB250

198

Surface Science Instrument spectrometer, at a vacuum of 2×10

199

radiation (1486.6 eV) was employed, which was obtained by bombarding the Al anode with an

200

electron gun operating at a beam current of 10 mA and an accelerating voltage of 15 kV. The

201

spectrometer energy-scale calibration and the charge correction considered that the C 1 s signal of the

202

contaminating carbon (C−C or C−H bonds) was centered at 284.8 eV. The detection limit for each

203

element was evaluated at about 0.1 At% and the accuracy for high concentrations about at 1 At%.

on wüstite. A 100x high numerical aperture and a 600 lines/mm grating

39

-1

40, 41

.

−9

ACS Paragon Plus Environment

mbar. Monochromatic Al Kα

ACS Earth and Space Chemistry

Page 8 of 28 8

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

204

2.5. Saturation index calculations

205

Saturation indices were calculated with PHREEQC v342,

206

Thermodynamic constants for the primary phases and potential secondary phases are given in Table

207

S2 (supporting information). Only potential phases are shown, but no phases in the databases were

208

discarded.

209

3. Results

210

3.1. Solution analyses and saturation index calculation

211

Batch experiments performed at 80 °C with three reaction times and for an initial solution pH80°C ~ 3

212

showed that the pH remained acidic after 24 hours of contact, but was near neutral after 504 hours of

213

contact (Table 3). Ca and Mg concentrations increased in solutions congruently (within uncertainties)

214

according to dolomite dissolution. In the meantime, Fe concentration increased between 30 min and 1

215

day, and then was below the detection limit once a steady state was reached (Batch-pH3-504). In the

216

AFM experiments, the congruence between Ca and Mg was also observed at pH 3 but not at pH 5;

217

AFM-pH5-3.1 displayed a higher Ca concentration in solution than Mg.

43

using the Thermoddem44 database.

-1

218

Table 3. Solution composition at the end of the experiments in mol L (QL: quantification limit,

219

QLFe = 4.8 10 mol L for AFM-pH3-2.4 and AFM-pH5-3.1, 1.8 10 mol L for Batch-pH3-24.3 and Batch-

-6

-1

-6

-7

220

-1

-1

pH3-504 and 3.6 10 mol L for Batch-pH3-0.5).

Experiment

Final pH

Alkalinity

Ca

Mg -5

Fe -5

(1.70±0.07) 10-6

Batch-pH3-0.5

2.82

n.m.

(8.48±0.34) 10

Batch-pH3-24.3

3.00

n.m.

(3.60±0.14) 10-4 (3.44±0.14) 10-4 (9.13±0.37) 10-6

Batch-pH3-504

7.85

(3.2±0.13) 10

AFM-pH3-2.4

3.25

n.m.

-4

-4

(7.24±0.29) 10

-4

(1.58±0.06) 10

-5

(7.82±0.31) 10

-4