Role of Electrochemical Surface Potential and ... - ACS Publications

Page 2 irradiation induces alterations in the atomic structure of almandine by ... and aluminate octahedra, wherein no O atoms are shared between two ...
0 downloads 0 Views 6MB Size
Subscriber access provided by Technical University of Munich University Library

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Role of Electrochemical Surface Potential and Irradiation on Garnet-type Almandine’s Dissolution Kinetics Yi-Hsuan Hsiao, Erika Callagon La Plante, N. M. Anoop Krishnan, Howard A Dobbs, Yann Le Pape, Narayanan Neithalath, Mathieu Bauchy, Jacob N. Israelachvili, and Gaurav N. Sant J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b04459 • Publication Date (Web): 04 Jul 2018 Downloaded from http://pubs.acs.org on July 7, 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 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018)

Role of Electrochemical Surface Potential and Irradiation on Garnet-type Almandine’s Dissolution Kinetics 1 2 3 4 5

Yi-Hsuan Hsiao (*,a), Erika Callagon La Plante (*,a), N. M. Anoop Krishnan (*,†,‡), Howard A. Dobbs (§), Yann Le Pape (**), Narayanan Neithalath (††), Mathieu Bauchy (†), Jacob Israelachvili (§,‡‡), Gaurav Sant* (*,§§,***)

6

Abstract

7

Nanoscale resolved quantifications of almandine’s (Fe3Al2(SiO4)3) dissolution rates across a

8

range of pH’s (1 ≤ pH ≤ 13) – established using vertical scanning interferometry (VSI) – reveal

9

that its dissolution rate achieves a minimum around pH 5. This minimum coincides with

10

almandine’s point of zero charge (PZC). These trends in almandine’s dissolution can be

11

estimated using the Butler-Volmer (BV) equation that reveals linkages between surface

12

potentials and dissolution rates, demonstrating proton- and hydroxyl-promoted breakage of Si–

13

O bonds. In contrast with well-polymerized silicates, the dissolution of almandine can also occur

14

through the rupture of its cationic bonds. This behavior is reflected in the observed influences

15

of irradiation on its dissolution kinetics. Molecular dynamics simulations highlight that



Laboratory for the Chemistry of Construction Materials (LC2), Department of Civil and Environmental Engineering, University of California, Los Angeles, CA, USA † Laboratory for the Physics of Amorphous and Inorganic Solids (PARISlab), Department of Civil and Environmental Engineering, University of California, Los Angeles, CA, USA ‡ Department of Civil Engineering, Indian Institute of Technology Delhi, Hauz Khas, New Delhi, India § Department of Chemical Engineering, University of California, Santa Barbara, CA, USA ** Oak Ridge National Laboratory, Oak Ridge, TN, USA †† School of Sustainable Engineering and the Built Environment, Arizona State University, Tempe, AZ, USA ‡‡ California Nanosystems Institute, University of California, Santa Barbara, CA, USA §§ Department of Materials Science and Engineering, University of California, Los Angeles, CA, USA *** California Nanosystems Institute, University of California, Los Angeles, CA, USA * Corresponding author: G. Sant, Phone: (310) 206-3084, Email: [email protected]

*

a

Both authors contributed equally to this work.

ACS Paragon Plus Environment

Page 1

The Journal of Physical 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 2 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 16

irradiation induces alterations in the atomic structure of almandine by reducing the

17

coordination state of the cations (Fe2+ and Al3+); thereby enhancing its reactivity by a factor of

18

two. This is consistent with the minor change induced in the structure of almandine’s silicate

19

backbone, whose surface charge densities produce the observed pH dependence (and rate

20

control) of dissolution rates. These findings reveal the influential roles of surface potential

21

arising from solution pH, and atomic scale alterations on affecting the reactivity of garnet-type

22

silicates.

23



24



25



26



27



28



29



30



31



32



33



34



35



36



37



38



ACS Paragon Plus Environment

Page 2

Page 3 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 39

1.

Introduction and background

40

Almandine (Fe3Al2(SiO4)3) – a nesosilicate mineral and an end-member of the garnet family

41

(X3Y2(SiO4)3), consists of isolated silicate tetrahedra connected by cations of mixed valence (X =

42

Fe2+, Y = Al3+). It is found in abundance in metamorphic1,2 and igneous rocks.3 For this reason,

43

garnet-containing sands find extensive use in engineering applications,4–8 e.g., in the form of

44

siliceous mineral aggregates (sand and stone) in concrete – a mixture composed of cement,

45

aggregates and water,4 in water filtration5 and as abrasives in waterjet cutting.6,8 Since the

46

durability of composites such as concrete depends on the durability of its constituents, there is

47

a need to understand how the chemical environment (e.g., solution pH), and external stimuli

48

(e.g., radiation), which may induce changes in its structure, may affect its properties. This is of

49

special relevance in facilities such as nuclear power plants (NPPs) wherein environmental

50

variables – e.g., exposure to irradiation and the alkaline environment – could compromise

51

concrete’s chemical durability.9–12

52 53

Almandine is composed of mixed-valence cations (Fe2+ and Al3+) and isolated silicate tetrahedra

54

and aluminate octahedra, wherein no O atoms are shared between two Si atoms or two Al

55

atoms (see Figure 1). Due to the isolated (non-percolating) nature of the silicate tetrahedra,

56

almandine is substantially less polymerized than framework silicates such as quartz and albite.14

57

As a result, during dissolution, isolated silicate tetrahedra and aluminate octahedra in

58

nesosilicates are released intact from reacting surfaces following protonation (or hydrolysis) of

59

bonds between Fe2+ and structural oxygens.15 This is unlike the case of tectosilicates (i.e., which

60

feature percolated silicate tetrahedra in which all O atoms are shared between Si atoms) such

ACS Paragon Plus Environment

Page 3

The Journal of Physical 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 4 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 61

as quartz and feldspar wherein bond hydrolysis needs to occur in the silicate tetrahedra to

62

release aqueous silicon species. In addition, almandine’s dissolution is also induced by ion-

63

exchange or hydrolysis in acidic or caustic media, respectively, processes which control its

64

dissolution behavior as a function of the pH, as typical of complex silicates.16

65

Figure 1: The crystal structure of pristine almandine (Fe3Al2(SiO4)3) showing isolated Si tetrahedra (dark blue), Al octahedra (light blue), Fe atoms (yellow) and O atoms (red) as visualized using VESTA.13 66 67

The rate of release of aqueous Si is typically regarded as reflecting a mineral’s dissolution rate.17

68

Several studies have shown similarities between the mechanisms of nesosilicate’s dissolution,

69

wherein hydrated cations leave a dissolving mineral surface, and ligand exchange reactions, in

70

which a hydrated cation complex exchanges with protons or hydroxyls in water.15 In almandine,

71

Fe2+ cations detach earliest producing an appropriate charge on almandine’s surface due to the

72

remaining Si and Al sites. As such, the observed rates of dissolution reflect the combined effects

73

of cation release and pH-dependent hydrolysis. This is significant as surface charge variations as

74

a function of pH may offer guidance regarding almandine’s pH-dependent dissolution rates.

75

ACS Paragon Plus Environment

Page 4

Page 5 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 76

Usually, the disordering of a silicate’s atomic structure, when induced at constant composition

77

(e.g., due to pressure, temperature or ballistic shocks), is expected to result in elevations of

78

chemical reactivity, as indicated by aqueous dissolution rates—although there are exceptions

79

depending on the solution’s composition or the silicate’s atomic structure.18 The vast majority

80

of such disordering-induced reactivity amplifications have been studied in framework (i.e.,

81

percolated) silicates wherein silicate tetrahedra are intimately connected to each other. It is

82

unknown whether silicates consisting of isolated (non-percolating) silicate tetrahedra may show

83

similar sensitivity to atomic disordering. In order to better understand these aspects, the

84

dissolution behavior of disordered (irradiated) almandine was examined using vertical scanning

85

interferometry (VSI) from observations of surface retreat. Molecular dynamics (MD) simulations

86

and topological constraint theory (TCT) were used to explain the dissolution trends in the

87

context of changes in atomic structure, e.g., in terms of the coordination number, bond length,

88

and bond angle distributions. It is shown that nesosilicates feature relatively small changes in

89

chemical reactivity following irradiation – since their isolated silicate tetrahedra show little if

90

any change in structural connectivity. This behavior is consistent with the dissolution

91

mechanism revealed from pH-dependent measurements of dissolution rates. Taken together,

92

these new insights offer guidance regarding the role of the electrochemical and ballistic

93

environment on affecting the durability of garnet-type silicates in aggressive environments.

94 95

2.

Materials and methods

96

2.1. Sample preparation and ion-irradiation

ACS Paragon Plus Environment

Page 5

The Journal of Physical 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 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 97

Naturally occurring almandine was sourced from Ward’s Natural Science Company.19 The

98

samples were sliced using a low speed diamond saw (IsoMetTM 1000, Buehler Inc.) into a size of

99

1 cm × 1 cm × 1 mm. The pristine (non-irradiated) samples were embedded in a cold-cured inert

100

acrylic resin (EpoxiCure Resin, PN 203430128: Buehler Inc.). After 24 hours, the resin-mounted

101

samples were successively mechanically polished using 400, 600, 800 and 1200 grit SiC

102

abrasives, and then using diamond paste (starting from 6, 3, 1 down to 0.25 µm). The polished

103

almandine samples had a surface roughness (Sa) of 100 nm over a field of view (FoV) of 1.3 µm

104

× 1.3 µm as assessed using vertical scanning interferometry (VSI).

105 106

In addition, polished samples prepared for Ar+ ion implantation were sent to the Michigan Ion

107

Beam Laboratory (MIBL). Ion implantation was conducted using 400 keV Ar+ ions at room

108

temperature to a total fluence (dose) of 1.0 × 1014 ions/cm2. Ar+ ion irradiation was selected to

109

simulate neutron irradiation in nuclear power plants because it offers precise control of the

110

temperature and dose, and is known to result in similar alterations to the crystal structure at

111

terminal disordering.20,21 The Ar+ implantation depth was calculated using “Stopping and Range

112

of Ions in Matter” (SRIM) to be on the order of 420 nm, assuming the samples are defect-free

113

single crystals. The calculation assumes that the density of almandine is 4.30 g/cm3, and that it

114

features a composition of Fe3Al2(SiO4)3.22 It should be noted however that the natural

115

almandine used herein contains minor impurities such as biotite and amphibole as detected

116

using polarized light microscopy (Figure 2), which is expected to result in a penetration depth

117

greater than 420 nm. The irradiated almandine samples were then adhered to a cylindrical resin

118

mount to facilitate handling.

ACS Paragon Plus Environment

Page 6

Page 7 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 119

(a) (b) Figure 2: Representative polarized light microscopy images of a thin section of the almandine sample shown in: (a) plane-polarized light, and, (b) cross-polarized light showing impurities of biotite and amphibole. The sample is dominantly composed of almandine, as evidenced from petrographic analysis. However, the images shown here represent a magnified region selected to demonstrate the presence of impurities. 120 121

2.2. Material characterization

122

2.2.1. X-ray Diffraction (XRD)

123

Finely powdered almandine was mounted on a rotating sample holder, and analyzed using

124

powder x-ray diffraction (XRD). The diffraction patterns were collected using a Bruker D8-

125

Advance powder x-ray diffractometer in θ-θ Bragg-Brentano geometry, using Cu-Kα radiation (λ

126

= 1.5406 Å) at an accelerating voltage of 40 kV, and a beam intensity of 40 mA. Each pattern

127

was recorded for a total duration of 16 minutes at a step size of 0.02° in the range of scattering

128

angles, θ, between 10° to 70°. The XRD pattern of the almandine sample is similar to that of

129

pure almandine reported in the literature,23 albeit showing the presence of biotite and

130

amphibole impurities (see Figures 2-3).

131 132

2.2.2. Thin section analysis using polarized light microscopy

ACS Paragon Plus Environment

Page 7

The Journal of Physical 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 8 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 133

The almandine sample was sectioned and trimmed to a thickness of 30 µm (Wagner

134

Petrographic) and examined using a polarized light microscope (Leica DM750P).24 This analysis

135

revealed impurities in the almandine matrix in the form of biotite and amphibole (Figure 2), in

136

agreement with the x-ray diffraction patterns shown in Figure 3.

137

Figure 3: The x-ray diffraction patterns of the natural (non-irradiated) almandine powder showing the presence of biotite and amphibole impurities. 138 139

2.3. Zeta potential measurement

140

The zeta potential of finely powdered almandine particulates (d50 ≈ 20 µm) was measured by

141

suspending 1 g of particulates in 100 mL Milli-Q deionized water (>18 MΩ·cm) that was

142

conditioned to an appropriate pH (1, 2, 4, 6, 10, 12 and 13) using NaOH (basic regime) or HCl

143

(acidic regime). The suspensions were sonicated for 1 minute after which 1.5 mL of the

144

suspension containing the particulates was placed into a cuvette for zeta potential

145

measurements. Zeta potentials were measured using a ZetaPALS analyzer from Brookhaven

146

Instruments Corporation25 that used an acoustic electrophoresis (i.e., streaming potential)

ACS Paragon Plus Environment

Page 8

Page 9 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 147

method. At each pH, ten discrete measurements of the zeta potential were obtained which

148

were then averaged to obtain statistically representative data.

149 150

2.4. Measuring almandine’s dissolution rates using vertical scanning interferometry (VSI)

151

Almandine was dissolved under flowing conditions using a flow-through cell (in-line diffusion

152

cell, PermeGear) at 25 ± 3 °C at each solution pH. The solutions were prepared by adding NaOH

153

or HCl to Milli-Q water (>18 MΩ·cm), and the pH of each solution was measured using a

154

ThermoFisher Scientific Orion Versa Star Pro pH Benchtop Multiparameter Meter calibrated

155

over the range 1 ≤ pH ≤ 13 at 25 ± 3 °C. The solutions were pumped into the flow-through cell

156

using a peristaltic pump (Bio-Rad)26 at flow rates ranging between 0-to-18 mL/minute.

157 158

To correlate atomic (structural) alterations following irradiation to dissolution rates, the

159

dissolution rates of pristine and irradiated almandine were measured using vertical scanning

160

interferometry (VSI) by quantifying the surface retreat due to dissolution following contact with

161

flowing solutions. It should be noted that a portion of the sample surface was covered with an

162

inert silicone mask (Silicone Solutions SS-38) to offer a “reference/inert area” from which the

163

extent of surface retreat of the dissolving (unmasked) area was measured. Following contact of

164

the almandine samples with the flowing solution for a selected contact time, the samples were

165

removed from the flow cell, and their surfaces were evacuated of any residual solution by

166

exposing them to a stream of compressed air. Thereafter, the silicone mask was peeled off to

167

expose the unreacted surface. In the last step, the sample’s surface topography was measured

168

using VSI (in air).

ACS Paragon Plus Environment

Page 9

The Journal of Physical 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 10 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 169 170

The nanoscale vertical resolution (1-2 nm) of VSI offers the ability to directly track the change in

171

height of the reacting surface with respect to a masked reference, as a function of the solution

172

exposure interval. This allows us to assess dissolution rates of reacting (dissolving, precipitating,

173

or corroding) materials at unparalleled resolution. VSI images were acquired using a 50× Mirau

174

objective (numerical aperture, N.A. = 0.55) which offers a lateral resolution of around 163 nm.

175

A single image field acquired using this objective features lateral dimensions of 170 µm × 170

176

µm. However, by stitching partially overlapping adjacent images, the field of view (FoV) can be

177

virtually expanded such that, herein, a stitched image field with dimensions of 1236 μm × 441

178

μm was acquired (N.B.: the image grid consisted of 10 × 3 images). All images were acquired

179

while accounting for the effects of tilt, i.e., by nulling the surfaces. The fringes on the sample

180

surface correspond to the light and dark bands that are produced by the – constructive

181

(amplitude enhances) or destructive (amplitude diminishes) – interference of the light source.

182

The operation of nulling the surfaces minimizes and distinguishes the number of fringes that

183

are visible inside the field of view (FoV) by adjusting the pitch (P) and roll (R) of the stage, thus

184

ensuring that the sample surface is level and is oriented perpendicular to the light source. The

185

analyses of acquired VSI images were carried out using Gwyddion (version 2.48)27 which

186

quantified the change in height of the un-masked surface area with respect to the “reference

187

area”. The change in height (Δh, nm) was divided by the solution contact time (Δt, h) to reveal

188

the dissolution rate (DR = Δh/Δt, nm/h). The uncertainty in dissolution rates acquired in this

189

manner is estimated to be on the order of 10% due to inherent uncertainties in VSI

ACS Paragon Plus Environment

Page 10

Page 11 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 190

measurements, e.g., low-frequency vibrations that may not be fully damped, or the effects of

191

temperature variations in the surrounding environment.

192



193

3.

194

MD simulations of irradiation-induced structural alterations in almandine were carried out

195

using LAMMPS28 following a previously developed approach.11,29–32 The pristine mineral was

196

simulated using the crystal structure of Armbruster et al.23 This unit cell was symmetrically

197

expanded in 3D to obtain a supercell consisting of 5760 atoms. Periodic boundary conditions

198

were imposed. First, this structure was equilibrated at 300 K using a Nosé-Hoover

199

thermostat33,34 for 20 ps. This ensured complete relaxation of the supercell. A constant

200

timestep of 1 fs was used during the initial relaxation of the structure.

Molecular dynamics (MD) simulations

201 202

To simulate ballistic impacts resulting from irradiation, an atom was chosen randomly from the

203

supercell. Note that weighted collision probabilities based on the neutron cross sections of the

204

elements were used while choosing the atom. Then, this chosen atom was accelerated with a

205

kinetic energy equivalent to that gained from an elastic collision with an incident neutron —

206

taken to be 600 eV in this case. The accelerated atom, also known as the primary knock-on

207

atom (PKA), then collided with other atoms in the supercell, called secondary knock-on atoms

208

(SKAs). The collisions of these atoms (PKA and SKAs) with other atoms in the supercell resulted

209

in a ballistic cascade. Ballistic cascades could result in high temperatures locally, which reduce

210

to ambient temperatures following relaxation. Therefore, to avoid any spurious effects of the

211

thermostat on the dynamics of the damage cascade, a spherical region with a radius of 10 Å

ACS Paragon Plus Environment

Page 11

The Journal of Physical 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 12 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 212

was created around the PKA. All atoms outside this spherical region were kept at a constant

213

temperature of 300 K by a Nosé-Hoover thermostat. Furthermore, the atoms inside the sphere

214

are treated within the NVE ensemble.35,36 This ensures that the dynamics of the damage

215

cascade, occurring within the spherical region, remain unaffected by the velocity scaling used

216

by the thermostat. It should be noted that the objective of defining a 10 Å cutoff is not to

217

distinguish the region of the system that can experience some local heating from the one that

218

cannot. In fact, the entire system is free to experience temporary heating (that is, if potential

219

energy is converted into kinetic energy). The Nosé-Hoover thermostat is used to simulate the

220

gradual energy dissipation that will occur due to heat exchange with the thermostat (i.e., due to

221

heat exchange of the excited zone with the bulk material in real situations). Rather, the

222

objective of defining a 10 Å-radius around the PKA wherein the NVE ensemble is enforced is to

223

ensure that the fast dynamics of the atoms colliding during the ballistic cascade is not affected

224

by any spurious effect of the thermostat. The size of this radius was determined by computing

225

the maximum spatial extent of the ballistic cascade occurring around the PKA for a given

226

incident energy. Further details regarding this methodology can be found in Wang et al. 2015.32

227

Here, a Buckingham potential parameterized by Teter and splined with the Ziegler-Biersack-

228

Littmark (ZBL) potential was used to capture realistic short-range repulsive interactions during

229

the ballistic cascades37. The choice of this potential was made since Teter’s force-field offers

230

excellent predictions of the structure of aluminosilicate glasses,38–40 while suitably describing

231

the crystal structure of pristine almandine.

232

ACS Paragon Plus Environment

Page 12

Page 13 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 233

Due to the high velocities and excessive collisions of atoms, a variable timestep was used during

234

the ballistic cascade to avoid numerical errors. Otherwise, a timestep of 1 fs was used. The

235

dynamics of the cascade was simulated for 15 ps. This was found to be long enough to ensure

236

the convergence of both the temperature and energy of the system. After each collision, the

237

system was further relaxed in the NPT ensemble at 300 K and zero pressure for another 5 ps.

238

This allowed complete relaxation of the structure. This process was repeated in an iterative

239

fashion, with different knock-on atoms, until the system exhibited saturation in terms of both

240

its enthalpy and density29. Any further irradiation was found to have no effect on the density or

241

enthalpy of the irradiated almandine structure.

242 243

4.

Results and discussion

244

4.1.

The dissolution behavior of pristine almandine in acidic and basic environments

245

The steady-state dissolution rates of pristine and irradiated almandine were measured using

246

VSI at a flow rate of 13.5 mL/min after a total solution contact time of 6 hours. This flow rate

247

was selected following measurements of dissolution rates over a wide range of flow rates since

248

it is known that both the flow rate and period of solution contact affect attainment of the

249

steady-state dissolution condition (see Figure 4a). The flow rate dependent dissolution analysis

250

indicated that: (a) for flow rates greater than 8 mL/min, the dissolution rate of almandine

251

became invariant with flow rate (e.g., see Figure 4a) indicating that steady-state dissolution has

252

been achieved; wherein the rate of release of constituent species from almandine is constant,

253

and, (b) flow rates greater than 8 mL/min were sufficient to prevent the formation and

254

accumulation of aluminum hydroxide (Al(OH)3) and ferrous hydroxide (Fe(OH)2) precipitates41

ACS Paragon Plus Environment

Page 13

The Journal of Physical 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 14 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 255

on almandine’s surface which could compromise measurements of surface topography, and

256

surface retreat using VSI. This is because, when causticity is abundant, following the equation

257

M2+ + nOH- → M(OH)n(s), where, M is a metal ion, almandine’s dissolution results in the

258

precipitation of Al(OH)3 and Fe(OH)2 on account of their low solubilities.

259 260

While dissolution is noted to be incongruent at short solvent contact times, at steady-state,

261

dissolution is congruent, i.e., stoichiometric, as evidenced by separate measurements of

262

elemental compositions of reacted solutions at acidic pH by inductively coupled plasma-optical

263

emission spectrometry (ICP-OES), consistent with Walther et al. (1996).42 The observed

264

incongruency during initial stages of dissolution is typical to aluminosilicates, and the

265

preferential release of either Al42–44 or Si45 initially has been previously observed at acidic pH.

266

The enhanced rate of Si release relative to Al in the almandine sample may be on account of the

267

size differences of silicate tetrahedra and aluminate octahedra, and the prevalence of isolated

268

silicate tetrahedra in almandine’s structure.

269 270

The dissolution of nesosilicates such as Ca2SiO4, Mg2SiO4, Be2SiO4, Ni2SiO4 and Co2SiO4 occurs

271

through mechanisms including the rupture of cationic bonds and release of silicate tetrahedra,

272

and hydrolysis of Si-O bonds within the silicate structural units.39,15 In general, the dissolution

273

rates in acidic media of silicates consisting of a single framework cation have been found to

274

scale in ascending order with the size of the framework cation (Figure 4b).15 Westrich et al.

275

explained this scaling in acidic solutions as being indicative of the relative ease of exchanging

276

the framework cation with protons during ion-exchange controlled dissolution.15 In proton-

ACS Paragon Plus Environment

Page 14

Page 15 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 277

promoted dissolution, bridging oxygens at the mineral surface are replaced by water molecules

278

or hydroxyl ions; the replaced hydrated ions are ultimately released from the mineral to the

279

aqueous solution. The rate of solvent exchange correlates with size of hydrated ions,

280

particularly since, larger cations have a larger surface area to accommodate more protons in

281

associative exchange, which make them easier to detach. A comparison of almandine’s

282

dissolution rate, a mixed-cation nesosilicate, with single-cation nesosilicates shows a trend that

283

is consistent with this concept (Figure 4b). This is indeed suggested by the work of Westrich et

284

al. which estimated dissolution rates for mixed-cation nesosilicates by a linear rule of mixtures

285

for solid-solution compositions (e.g., analogous to Vegard’s law)46 as shown in Equation (1):39

286

(a) (b) Figure 4: (a) The dissolution rate of almandine as a function of flow rate quantified using vertical scanning interferometry (VSI). Beyond a flow rate of 8 mL/min the dissolution rate plateaus demonstrating a flow rate independence. (b) The dissolution rates of nesosilicates as a function of their cationic radii. Taking into account the average cationic radius (Fe2+, Al3+) of almandine, the dissolution rate of almandine (i.e., a mixed cation nesosilicate) estimated using Equation (1) (red) closely matches its measured dissolution rates (green) in agreement with the broader trend that relates dissolution rates and ionic radii for nesosilicates (blue). 287

ACS Paragon Plus Environment

Page 15

The Journal of Physical 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 16 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 288

𝑟"𝐴$ 𝐵& (𝑆𝑖𝑂+ )- . = 𝐴01.% [𝑟(𝐴0 (𝑆𝑖𝑂+ )5 )] + 𝐵01.% [𝑟(𝐵8 (𝑆𝑖𝑂+ )9 )]

Eq. (1a)

289

𝐷; "𝐴$ 𝐵& (𝑆𝑖𝑂+ )- . = 𝐴01.% [𝐷; (𝐴0 (𝑆𝑖𝑂+ )5 )] + 𝐵01.% [𝐷; (𝐵8 (𝑆𝑖𝑂+ )9 )]

Eq. (1b)

290

where, r is the cation radius, Dr is the dissolution rate, A and B are metal cations, and at.%

291

represents the atomic fraction of a given cation in the mixed-cation composition. Of course, the

292

implicit condition that needs to be satisfied in the case of mixed-cation nesosilicates is that the

293

cations feature similar sizes. A similar relationship can be expected at high pH, such that cations

294

with a larger field strength (i.e., ratio of charge to radius) induce faster dissolution because of

295

their tendency to concentrate OH-, which causes bond polarization. The cation influence

296

detailed above describes the variation in rates for a given pH. Steady-state dissolution rates of

297

almandine also vary with pH, as shown in Figure 5(a). In general, it is seen that dissolution rate

298

achieves a minimum at around pH 5 – featuring increases on either side of this dissolution

299

minimum. Such dependencies of (alumino)silicate dissolution rates on solution pH have been

300

observed extensively wherein the dissolution rate scales as a function of {𝑎}? 47–49 where ‘{a}’

301

denotes the activity of an ion a, typically H+ or OH- species depending on the solution pH, and n

302

is an empirical fitting coefficient. When the logarithm of the dissolution rates is plotted as a

303

function of the solution pH – n is often noted to vary between 0.3 ≤ n ≤ 0.5 for silicates.42,50

304 305

Furthermore, it is noted that almandine’s zeta potential decreases systematically from acidic to

306

basic regimes with the point of zero charge (PZC) – sited around pH 5 – corresponding to the

307

point at which the surface charges of Al and Si surface sites are minimized.42 Due to the

308

presence of H2O, an oxide mineral’s surface is generally covered in varying degrees with

309

“surface hydroxyl groups”, e.g., Si–OH.51 For the case of pure silica, the minimum total surface

ACS Paragon Plus Environment

Page 16

Page 17 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 310

charge is affected only by H+ and OH- in the solution, without any specific adsorption on the

311

surface, and hence the PZC and isoelectric point (IEP) are equivalent. In acidic solutions, the

312

reaction between a silicate surface and the solution interface can be written as:

313

Si − OH + H + → Si − OH 2+

314

Therefore, positive charges are present on the mineral surface. When the surface contacts

315

alkaline solutions, the reaction can be written as:

316

Si − OH + OH − → H 2O + Si − O −

317

This explains the origin of the negative surface charge of silicates in basic solutions.51 Based on

318

studies on quartz and aluminum oxide, Walther et al.42,44 observed that the minimum in the

319

dissolution rate occurs at the pH where the sum of absolute charge from Si in quartz or Al in

320

aluminum oxide is minimized—i.e., pH 2.5 and 8.4 for quartz and corundum, respectively.42

Eq. (2a)

Eq. (2b)

321

(a) (b) (c) Figure 5: (a) The dissolution rates of pristine almandine measured at a flow rate of 13.5 mL/min showing the dissolution minimum in the mid-pH region (pH 5). (b) The zeta potential of pristine almandine measured as a function of solution pH. (c) The dissolution rate of pristine almandine as a function of its zeta potential showing that the dissolution minimum occurs at the pH where the sum of the absolute value of surface charge (Al and Si) is minimized; i.e., the point of zero charge (PZC, pH 5) is achieved. Also shown is a comparison of measured (blue circles) and estimated dissolution rate for pristine almandine (dashed red line) using Equation (3).

ACS Paragon Plus Environment

Page 17

The Journal of Physical 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 18 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 322 323

As such, for aluminosilicates, the dissolution minima are expected to be determined by

324

composition, i.e., the Al/Si ratio, which dictates the pH at which the surface features minimum

325

(zero) surface charge. Since alkalis and alkaline earth cations tend to leach rapidly from

326

aluminosilicates and be replaced by protons, it is the residual Al and Si left on the surface that

327

determine the value of this pH.42,52 For example, for Al/Si < 1 (e.g., almandine; Al/Si = 0.67) the

328

surface charge and dissolution rate minimum is observed at a pH between 4 and 6, as noted in

329

Figure 5(a-b). This pH-dependence of dissolution rate is controlled by the electrosorption of H+

330

and OH− ions on terminal surface Al or Si sites. Since pure SiO2 and Al2O3 exhibit a dissolution

331

minimum at pH 2.5 and 8.4, respectively, it is expected that, in acidic regimes (pH < 7), the

332

surface charge for SiO2 is close to 0, and is relatively low in magnitude compared to its value in

333

high pH conditions. The opposite is true for Al2O3, which exhibits high surface charge at low pH

334

compared to SiO2, and while it becomes increasingly negative in alkaline conditions, it remains

335

smaller than of SiO2.42,53,54 This analysis affords the use of surface charge, as indicated by zeta

336

potentials, to estimate the dissolution kinetics of silicates. It should be noted however that zeta

337

potential does not discriminate the identity of surface sites, and is only sensitive to the total

338

charge arising from both Si or Al sites on the mineral surface.

339 340

Recently, Kristiansen et al. showed that the dissolution kinetics of silica enhanced with applied

341

potential.55,56 Such a dissolution rate-applied potential scaling can be described using the

342

Butler-Volmer equation as shown in Equation (3) below:

343

𝐷; = −𝐶𝑒

CD

E ∆J FG H

Eq. (3a)

ACS Paragon Plus Environment

Page 18

Page 19 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) KL -M

R S FG O,U

Eq. (3b)

𝐶=

345

where, Dr is the mineral dissolution rate (mol/cm2/s), α is a transfer factor, and the product α

346

·∆U (mV) reveals the surface potential (note that in the fitting below, ∆U is the zeta potential,

347

mV),57 e/kBT is the “thermal voltage” which equals to 1/25.8 (mV-1) at room temperature (K),

348

where e is the charge on an electron, kB is Boltzmann’s constant (eV/K), T is the temperature

349

(K), and C (mol/cm2/s) is a material specific constant based on its molar mass (Mm, 497.75

350

g/mol) and density (ρ, 4.30 g/cm3), and where Ea is the activation energy of the dissolution

351

reaction (kJ/mol), z is the collision frequency (z, s-1), p is a steric factor (1/6), and A denotes the

352

surface area that is available for dissolution (m2). It should be noted that the negative sign

353

preceding the transfer factor is eliminated if the surface potential is positive.

NO

𝑒 PQ

C

344

354 355

By regression, Equation (3a) can be fitted to the dissolution rates of pristine almandine in acidic

356

and basic regimes, separately, to recover C (2.5) and α (3.8 and 1.6), respectively, as shown in

357

Figure 5(c). The fitting applies a transfer factor to describe the difference between the surface

358

potential and zeta potential (i.e., if the surface potential and zeta potential were equal, α = 1).57

359

The difference in numerical values of α reflects a change in dissolution mechanism (e.g., ion

360

exchange in acidic solution or hydrolysis in basic solution) with pH. The reasonableness of the

361

fitted values of the transfer factor can be assessed from knowledge of the electrical potential as

362

a function of distance from the particle surface as shown in Equation (4):58,59

363

ϕ = ϕ 0 e −κ x

364

⎛ 1 M 2 2 ∞⎞ κ =⎜ ∑ zi e ni ⎟ ⎝ εrε 0 kT 1 ⎠

Eq. (4a) 1/2

Eq. (4b)

ACS Paragon Plus Environment

Page 19

The Journal of Physical 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 20 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 365

where, 𝜑W (mV) is the surface potential at x = 0 (i.e., at the mineral surface, nm), 𝜅 is the inverse

366

of the Debye length60 (nm-1) which depends on the solution’s ionic strength, εrε0 is electrical

367

permittivity, zi is the valence of the ionic species, ni∞ is the number density in the bulk solution,

368

and M is the number of species (Na+, OH-, H+, Cl-). As such, 1/𝑒 C[$ = α, i.e., the transfer factor

369

in Equation (3a). At pH 13, for a 0.1 mol/L NaOH solution, setting x ≈ 0.3 nm to describe the

370

position of the slip results in a transfer factor: 1/𝑒 C[$ = 1.47 which is similar to the value

371

suggested by fitting Equation (3a) to the data in Figure 5(c) under alkaline conditions (α = 1.6).

372

This is a significant finding that supports the use of electrical surface potentials (or alternatively,

373

zeta potential) as a proxy of mineral dissolution rates. As such, it is evident that the dissolution

374

rate of a mineral correlates with the surface potential as described by a modified Butler–

375

Volmer equation. This suggests that mineral dissolution is fundamentally linked to the

376

electrochemistry of surfaces. Not only does this finding provide new insights into dissolution

377

dynamics in general, but it offers a means to potentially estimate dissolution rates from

378

measured zeta potentials, and vice versa. The universality of the observed relationship between

379

zeta potentials and dissolution rate is an important aspect that needs to be further clarified.

380

Nonetheless, taking the examples of the oxide surfaces of Al2O3 and SiO2,42 and of mica, silica

381

and zircon,55 it can be argued that the observed relationship is indeed – general, and broadly

382

applicable – i.e., to both simple, and complex oxide surfaces.

383 384 385

4.2.

Using MD simulations and topological constraint theory to assess dissolution rate alterations following structural changes induced by irradiation

ACS Paragon Plus Environment

Page 20

Page 21 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 386

To elucidate the effects of irradiation on chemical durability, the dissolution rate of irradiated

387

almandine was measured in acidic and basic media. In general, irradiation resulted in an

388

increase (up to 2x, where x is the dissolution rate of pristine almandine at a given pH) in

389

dissolution rate across all pH levels (Figure 6a). Importantly, the slope of the dissolution rate-pH

390

trendline is unaffected, whether almandine may be irradiated or not, indicating that

391

electrochemical controls on dissolution remained unaffected by radiation (see Section 4.1). This

392

also highlights that while irradiation can alter absolute dissolution rates, it does not alter the

393

rate controlling step(s) in dissolution (i.e., the dissolution mechanism/pathway). Previously,

394

Pignatelli et al. and Hsiao et al. have shown that silicate dissolution rates substantially elevate

395

following a mineral’s exposure (e.g., quartz, albite) to irradiation.11,12 This is on account of the

396

structural disordering/amorphization that is induced. However, both quartz and albite feature

397

highly polymerized structures wherein their silicate tetrahedra are intimately connected to

398

each other – forming a periodic framework. Consequently, irradiation induces substantial

399

depolymerization as a result of which quartz’s dissolution rate can elevate by up to 3 orders of

400

magnitude and albite’s dissolution rate can increase by up to 20 times at a fixed pH.12

401

Contrastingly, almandine shows a comparatively marginal increase (doubling) of its aqueous

402

dissolution rate following irradiation.

403 404

Since MD simulations provide direct access to the trajectory of atoms, they offer direct insight

405

into the nature of atomic alterations induced in almandine’s structure following irradiation.

406

Analysis of the pair distribution functions (PDF, g(r)), before and following irradiation, as shown

407

in Figure 6(b) indicates that almandine’s short-range order (SRO, 3 Å), a substantial smoothening/dissipation

410

of the PDF is observed, indicative of vitrification.62–64

411

(a) (b) Figure 6: (a) The dissolution rates of pristine and irradiated almandine as a function of the solution pH. The similar slopes in acidic and basic regimes clarify that irradiation does not induce a change in the dissolution mechanism, and the elevation in dissolution rates simply arises from a reduction in cation coordination. (b) The pair distribution functions (g(r)) of pristine and irradiated almandine. 412 413

The effects of vitrification on structural connectivity/rigidity can be quantified within the

414

framework of topological constraint theory (TCT).65,66 TCT reduces complex atomic networks to

415

a simple network consisting of nodes (atoms) and rigid-members (atomic bonds).67 This allows

416

computation of the rigidity of an atomic network as described by the average number of

417

topological constraints placed on a given atom in the network (nc, unitless). The constraints

418

herein include: (a) the radial bond-stretching (BS), and, (b) the angular bond-bending (BB)

419

variants which maintain bond lengths and bond angles fixed around their average values,

420

respectively. Following Maxwell’s stability criterion,68 atomic networks can be categorized as:

ACS Paragon Plus Environment

Page 22

Page 23 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 421

(1) flexible (𝑛] < 3), showing floppy internal modes of deformation,69 (2) stressed-rigid (𝑛] >

422

3), featuring eigenstress due to mutually incompatible constraints,70,71 or (3) isostatic (𝑛] = 3),

423

being rigid but free of eigenstress. This nature of analysis reveals that the number of constraint

424

per atom (nc) for almandine decreased from 2.85 to 2.18, from the pristine to the irradiated

425

state, respectively. This reduction in the number of constraints, or network rigidity, as also

426

reflected by the smoothening of the MRO is on account of: (1) a change in the coordination of

427

87.8% of Al atoms from octahedral to tetrahedral coordination, (b) a change in the coordination

428

of 74.4% of Fe-species from being 8-coordinated Fe to becoming 5-or-6 coordinated, and, (3)

429

the formation of non-bridging oxygen (NBO) atoms, by an amount equivalent to 33% of all O

430

atoms, which do not exist in pristine almandine.

431

(a) (b) Figure 7: Representative snapshots of the simulated atomic structure of: (a) pristine and (b) irradiated almandine. For clarity, only SiO44- units are shown (Si: yellow, O: red) to highlight the state of silicate connectivity through the network. 432 433

It is important to note, the decrease in the number of constraints (Δnc = 0.67) in almandine

434

results mostly from a decrease in cationic coordination number(s) rather than a change in the

ACS Paragon Plus Environment

Page 23

The Journal of Physical 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 24 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 435

atomic connectivity of silicate tetrahedra (e.g., a change in Si–O bond length or Si–O–Si bond

436

angle in the case of framework silicates). This is significant as, following Hsiao et al. (2017), Δnc

437

= 0.67 should result in an increase of dissolution rate by around 200 times for typical silicate-

438

based materials.12 However, a much more modest increase in dissolution rates (and hence

439

decrease in chemical durability) is noted. The breakdown of the previously suggested

440

relationship between dissolution rate and number of constraints per atom can be understood

441

from the fact that, in contrast to more polymerized silicates,72 irradiation does not induce

442

changes in the connectivity of almandine’s silicate backbone – but rather only eases the

443

breakage of cationic (Fe2+ and Al3+) bonds in the structure. In silicates, dissolution implies the

444

breakage of high-energy Si–O bonds, which often serves as the rate-limiting step of

445

dissolution.73,74 As shown in Figure 7(a), pristine almandine initially exhibits isolated SiO4 units

446

(i.e., no Si–O–Si bonds are present). Both before and following irradiation, dissolution can

447

proceed without requiring the breakage of Si–O bonds. As such, SiO44- units are directly

448

transferred into solution as rigid blocks.

449 450

Upon the disordering of almandine’s network following irradiation although some Si–O–Si

451

bonds form, they are limited. Hence, many SiO44- units remain isolated or form small clusters

452

(see Figure 7b) so that silicate chains (i.e., –Si–O–Si– bonds) do not percolate throughout the

453

atomic network. This is partially on account of the composition which tends to form Si–O–Al

454

bonds rather than Si–O–Si or Al–O–Al bonds – in partial compliance with Loewenstein’s “Al-

455

avoidance” rule. 75,76 Consequently, dissolution can also occur through another path such as

456

breakage of cationic bonds rather than Si–O bonds. It is the former process that is enhanced by

ACS Paragon Plus Environment

Page 24

Page 25 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 457

irradiation, particularly through changes in the cation’s coordination numbers. It should be

458

noted that the small silicate clusters (i.e., with 2 or 3 Si atoms) that form upon irradiation are

459

expected to be released into solution as rigid units (rather than individual SiO44- groups), which

460

explains the slight increase (around 2x) in dissolution rate. Hence, in nesosilicates, dissolution

461

rates are enhanced by facilitated extraction of their cations rather than by affecting the

462

connectivity of the silicate network.

463 464

5.

Summary and conclusions

465

This study presents new insights into almandine’s dissolution behavior in acidic and basic

466

regimes; and following exposure to radiation. First, dissolution kinetics are shown to be

467

correlated with a mineral’s surface charge. In particular, it is shown that the minimum

468

dissolution rate is achieved at pH 5, corresponding to the pH at which almandine, which has a

469

garnet-type structure, achieves a minimum (absolute) value of total Al and Si surface charge.

470

This allows the dissolution rates of almandine to be estimated using a modified Butler-Volmer

471

equation, highlighting the critical role of surface electrochemistry on dissolution processes.

472

Second, it is highlighted that almandine experiences relatively small elevations in dissolution

473

rates following irradiation, as compared to percolated silicates such as quartz and albite with

474

feature higher levels of polymerization. Molecular dynamics simulations show that unlike

475

framework silicates, silicate connectivity in the garnet-type almandine remains broadly

476

unaffected upon irradiation. Instead, the coordination numbers of the cations Fe2+ and Al3+

477

decreases, causing a modest, but constant, enhancement in rates across the entire solution pH

478

range. Thus, the mechanisms of dissolution, which include both breakage of cationic bonds and

ACS Paragon Plus Environment

Page 25

The Journal of Physical 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 26 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 479

subsequent release of intact silicate tetrahedra as well as pH-dependent hydrolysis of Si–O

480

bonds, are reflected in the observed irradiation effect on overall dissolution rate. As such, the

481

outcomes of this work offer new insights into the origin of, and rate controls on the reactivity of

482

non-percolated garnet-type silicates across a wide range of solution pH’s and specifically

483

highlight how surface electrochemistry and surface potential, and network polymerization

484

affect dissolution processes in natural and engineered material systems. This understanding

485

gained, i.e., of electrochemical controls on dissolution processes can be applied not only within

486

an earth science perspective, but also to a diversity of engineered and biological systems

487

including: nuclear waste disposal, glass corrosion, cementation processes, hard tissue systems,

488

and micro- and nanofluidic systems, in which an understanding of the controls on reactivity of

489

oxide surfaces is of considerable importance.

490 491

Acknowledgements

492

The authors acknowledge financial support for this research provisioned by the: Department of

493

Energy’s Nuclear Energy University Program (DOE-NEUP: DE-NE0008398), National Science

494

Foundation (CAREER Award: 1253269), the U.S. Department of Transportation (U.S. DOT)

495

through the Federal Highway Administration (DTFH61-13-H-00011) and University of California,

496

Los Angeles (UCLA). The contents of this paper reflect the views and opinions of the authors

497

who are responsible for the accuracy of data presented. This research was carried out in the

498

Laboratory for the Chemistry of Construction Materials (LC2), Molecular Instrumentation

499

Center, and Laboratory for the Physics of Amorphous and Inorganic Solids (PARISlab) at UCLA.

ACS Paragon Plus Environment

Page 26

Page 27 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 500

As such, the authors gratefully acknowledge the support that has made these laboratories and

501

their operations possible.

502 503

References

504

(1)

Garnet-Muscovite Assemblages. Contrib. Mineral. Petrol. 1981, 76, 92–97.

505 506

(2)

Cressey, G. Exsolution in Almandine-Pyrope-Grossular Garnet. Nature 1978, 271, 533– 534.

507 508

Ghent, E. D.; Stout, M. Z. Geobarometry and Geothermometry of Plagioclase-Biotite-

(3)

Keesmann, I.; Matthes, S.; Schreyer, W.; Seifert, F. Stability of Almandine in the System

509

FeO-(Fe2O3)-Al2O3-SiO2-(H2O) at Elevated Pressures. Contrib. Mineral. Petrol. 1971, 31,

510

132–144.

511

(4)

Mater. Sci. 1998, 21, 349–354.

512 513

(5)

Conley, W. R.; Hsiung, K. Design and Application of Multimedia Filters. J. Am. Water Works Assoc. 1969, 61, 97–101.

514 515

Banerjee, G. Beach and Minerals: A New Material Resource for Glass and Ceramics. Bull.

(6)

Barcova, K.; Mashlan, M.; Martinec, P. Mössbauer Study of the Thermal Behaviour of

516

Garnets Used in High-Energy Water Jet Technologies. Industrial Applications of the

517

Mössbauer Effect 2002, 463–469.

518

(7)

Olsen, D.W. Garnet, Industrial, U.S. Geological Survey Publications, 2002.

519

(8)

Kulekci, M. K. Processes and Apparatus Developments in Industrial Waterjet Applications.

520

Int. J. Mach. Tools Manuf. 2002, 42, 1297–1306.

ACS Paragon Plus Environment

Page 27

The Journal of Physical 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 28 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 521 522 523 524 525

(9)

Ichikawa, T.; Koizumi, H. Possibility of Radiation-Induced Degradation of Concrete by Alkali-Silica Reaction of Aggregates. J. Nucl. Sci. Technol. 2002, 39, 880–884.

(10) Ichikawa, T.; Kimura, T. Effect of Nuclear Radiation on Alkali-Silica Reaction of Concrete. J. Nucl. Sci. Technol. 2007, 44, 1281–1284. (11) Pignatelli, I.; Kumar, A.; Field, K. G.; Wang, B.; Yu, Y.; Le Pape, Y.; Bauchy, M.; Sant, G.

526

Direct Experimental Evidence for Differing Reactivity Alterations of Minerals Following

527

Irradiation: The Case of Calcite and Quartz. Sci. Rep. 2016, 6, 20155.

528

(12) Hsiao, Y.-H.; La Plante, E. C.; Krishnan, N. M. A.; Le Pape, Y.; Neithalath, N.; Bauchy, M.;

529

Sant, G. Effects of Irradiation on Albite’s Chemical Durability. J. Phys. Chem. A 2017, 121,

530

7835-7845.

531 532 533 534 535 536 537 538 539 540 541 542

(13) Momma, K.; Izumi, F. VESTA 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276. (14) Crundwell, F. K. The Mechanism of Dissolution of Forsterite, Olivine and Minerals of the Orthosilicate Group. Hydrometallurgy 2014, 150, 68–82. (15) Casey, W.H.; Westrich, H.R.; Control of Dissolution Rates of Orthosilicate Minerals by Divalent Metal-Oxygen Bonds, Nature 1992, 355, 6356. (16) Westrich, R. T.; Cygan, R. T.; Casey, W. H.; Zemitis, C.; Arnold, G.W. The Dissolution Kinetics of Mixed-Cation Orthosilicate Minerals. Am. J. Sci. 1993, 293, 869-893. (17) Dove, P. M. The Dissolution Kinetics of Quartz in Aqueous Mixed Cation Solutions. Geochim. Cosmochim. Acta 1999, 63, 3715–3727. (18) Hamilton, J. P.; Pantano, C. G.; Brantley, S. L. Dissolution of Albite Glass and Crystal. Geochim. Cosmochim. Acta 2000, 64, 2603–2615.

ACS Paragon Plus Environment

Page 28

Page 29 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 543

(19) Ward's Science. https://www.wardsci.com/store/ (accessed June 5 2018).

544

(20) Douillard, L.; Duraud, J. P. Swift Heavy Ion Amorphization of Quartz - A Comparative

545

Study of the Particle Amorphization Mechanism of Quartz. Nucl. Instrum. Methods Phys.

546

Res., Sect. B 1996, 107, 212−217.

547 548 549

(21) Snead, L. L.; Zinkle, S. J.; Hay, J. C.; Osborne, M. C. Amorphization of SiC under Ion and Neutron Irradiation. Nucl. Instrum. Methods Phys. Res., Sect. B 1998, 141, 123−132. (22) Stoller, R. E.; Toloczko, M. B.; Was, G. S.; Certain, A. G.; Dwaraknath, S.; Garner, F. A. On

550

the Use of SRIM for Computing Radiation Damage Exposure. Nucl. Instrum. Methods

551

Phys. Res. Sect. B Beam Interact. Mater. At. 2013, 310, 75–80.

552 553

(23) Armbruster, T.; Geiger, C.; A Lager, G. Single-Crystal X-Ray Structure Study of Synthetic Pyrope Almandine Garnets at 100 and 293 K. Am. Mineral. 1992, 77, 512-521.

554

(24) Leica Microsystems. http://www.leica-microsystems.com/home (accessed June 5 2018).

555

(25) Brookhaven Instruments Corporation. https://www.brookhaveninstruments.com/

556

(accessed June 5 2018).

557

(26) Bio-Rad. http://www.bio-Rad.com/ (accessed June 5 2018).

558

(27) Nečas, D.; Klapetek, P. Gwyddion: An Open-Source Software for SPM Data Analysis. Open

559 560 561

Phys. 2011, 10, 181–188. (28) Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys. 1995, 117, 1–19.

562

(29) Krishnan, N. M. A.; Wang, B.; Yu, Y.; Le Pape, Y.; Sant, G.; Bauchy, M. Enthalpy Landscape

563

Dictates the Irradiation-Induced Disordering of Quartz. Phys. Rev. X 2017, 7, 31019.

ACS Paragon Plus Environment

Page 29

The Journal of Physical 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 30 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 564

(30) Krishnan, N. A.; Wang, B.; Le Pape, Y.; Sant, G.; Bauchy, M. Irradiation-vs. Vitrification-

565

Induced Disordering: The Case of 𝜶-Quartz and Glassy Silica. J. Chem. Phys. 2017, 146,

566

204502.

567 568 569 570 571 572 573 574 575

(31) Devanathan, R.; Corrales, L. R.; Weber, W. J.; Chartier, A.; Meis, C. Molecular Dynamics Simulation of Disordered Zircon. Phys. Rev. B 2004, 69, 64115. (32) Wang, B.; Yu, Y.; Pignatelli, I.; Sant, G.; Bauchy, M. Nature of Radiation-Induced Defects in Quartz. J. Chem. Phys. 2015, 143, 24505. (33) Hoover, W. G. Canonical Dynamics: Equilibrium Phase-Space Distributions. Phys. Rev. A 1985, 31, 1695. (34) Nosé, S. A Unified Formulation of the Constant Temperature Molecular Dynamics Methods. J. Chem. Phys. 1984, 81, 511–519. (35) Douillard, L.; Duraud, J. P. Swift Heavy Ion Amorphization of Quartz — a Comparative

576

Study of the Particle Amorphization Mechanism of Quartz. Nucl. Instrum. Methods Phys.

577

Res. Sect. B Beam Interact. Mater. At. 1996, 107, 212–217.

578 579 580 581

(36) Douillard, L.; Duraud, J. P. Amorphization of α-Quartz under Irradiation. J. Phys. III 1996, 6, 1677–1687. (37) Biersack, J. P.; Ziegler, J. F. The Stopping and Range of Ions in Solids. In Ion Implantation Techniques Springer, Berlin, Heidelberg, 1982, 122–156.

582

(38) Bauchy, M. Structural, Vibrational, and Elastic Properties of a Calcium Aluminosilicate

583

Glass from Molecular Dynamics Simulations: The Role of the Potential. J. Chem. Phys.

584

2014, 141, 24507.

ACS Paragon Plus Environment

Page 30

Page 31 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 585 586 587 588 589 590 591 592 593 594 595

(39) Yu, Y.; Wang, M.; Zhang, D.; Wang, B.; Sant, G.; Bauchy, M. Stretched Exponential Relaxation of Glasses at Low Temperature. Phys. Rev. Lett. 2015, 115, 165901. (40) Cormack, A. N.; Du, J.; Zeitler, T. R. Sodium Ion Migration Mechanisms in Silicate Glasses Probed by Molecular Dynamics Simulations. J. Non-Cryst. Solids 2003, 323, 147–154. (41) Velbel, M. A. Natural Weathering Mechanisms of Almandine Garnet. Geology 1984, 12, 631–634. (42) Walther, J. V. Relation between Rates of Aluminosilicate Mineral Dissolution, pH, Temperature, and Surface Charge. Am. J. Sci. 1996, 296, 693–728. (43) Chou, L.; Wollast, R. Steady-State Kinetics and Dissolution Mechanisms of Albite. Am. J. Sci. 1985, 285, 963–993. (44) Carroll-Webb, S. A.; Walther, J. V. A Surface Complex Reaction Model for the pH-

596

Dependence of Corundum and Kaolinite Dissolution Rates. Geochim. Cosmochim. Acta

597

1988, 52, 2609–2623.

598 599 600 601 602 603 604 605

(45) Wieland, E.; Stumm, W. Dissolution Kinetics of Kaolinite in Acidic Aqueous Solutions at 25°C. Geochim. Cosmochim. Acta 1992, 56, 3339–3355. (46) Vegard, L. Die Konstitution der Mischkristalle und die Raumfüllung der Atome. Z. Für Phys. 1921, 5, 17–26. (47) Blum, A. E.; Lasaga, A. C. The Role of Surface Speciation in the Dissolution of Albite. Geochim. Cosmochim. Acta 1991, 55, 2193–2201. (48) Lasaga, A. C.; Kirkpatrick, R. J. In Reviews in Mineralogy; Vol. 8, Kinetics of Geochemical Processes. 3, 1981.

ACS Paragon Plus Environment

Page 31

The Journal of Physical 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 32 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 606 607 608 609 610 611 612 613 614 615 616 617 618

(49) Chou, L.; Wollast, R. Study of the Weathering of Albite at Room Temperature and Pressure with a Fluidized Bed Reactor. Geochim. Cosmochim. Acta 1984, 48, 2205–2217. (50) Lasaga, A. C. Chemical kinetics of water-rock interactions. J. Geophys. Res. 1984, 89, 4009-4025. (51) Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters; Stumm, W.; Morgan, J. J.; Wiley 2012. (52) Parks, G. A. Aqueous Surface Chemistry of Oxides and Complex Oxide Minerals. Am. Chem. Soc. 1967, 67, 121–160. (53) Brady, P. V.; Walther, J. V. Kinetics of Quartz Dissolution at Low Temperatures. Chem. Geol. 1990, 82, 253–264. (54) Dzombak, D. A.; Morel, F. M. M. Surface Complexation Modeling: Hydrous Ferric Oxide. New York, John Wiley, 1990, 393. (55) Kristiansen, K.; Valtiner, M.; Greene, G. W.; Boles, J. R.; Israelachvili, J. N. Pressure

619

Solution – The Importance of the Electrochemical Surface Potentials. Geochim.

620

Cosmochim. Acta 2011, 75, 6882–6892.

621

(56) Greene, G. W.; Kristiansen, K.; Meyer, E. E.; Boles, J. R.; Israelachvili, J. N. Role of

622

Electrochemical Reactions in Pressure Solution. Geochim. Cosmochim. Acta 2009, 73,

623

2862–2874.

624

(57) Valtiner, M.; Kristiansen, K.; Greene, G. W.; Israelachvili, J. N. Effect of Surface Roughness

625

and Electrostatic Surface Potentials on Forces between Dissimilar Surfaces in Aqueous

626

Solution. Adv. Mater. 2011, 23, 2294–2299.

627

(58) Hunter, R. J. In Zeta Potential in Colloid Science. Academic Press. 1981, 1–58.

ACS Paragon Plus Environment

Page 32

Page 33 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 628 629 630 631 632

(59) Ohshima, H. In Colloid and Interface Science in Pharmaceutical Research and Development. Elsevier, 2014, 1–28. (60) Debye, P.; Hfickel, E. The Interionic Attraction Theory of Deviations from Ideal Behavior in Solution, Phys. Z. 1923, 24, 185. (61) Stoller, R. E.; Toloczko, M. B.; Was, G. S.; Certain, A. G.; Dwaraknath, S.; Garner, F. A. On

633

the Use of SRIM for Computing Radiation Damage Exposure. Nucl. Instrum. Methods

634

Phys. Res. Sect. B Beam Interact. Mater. At. 2013, 310, 75–80.

635

(62) Wang, B.; Krishnan, N. M. A.; Yu, Y.; Wang, M.; Le Pape, Y.; Sant, G.; Bauchy, M.

636

Irradiation-Induced Topological Transition in SiO 2: Structural Signature of Networks’

637

Rigidity. J. Non-Cryst. Solids 2017, 463, 25−30.

638 639 640 641 642 643 644 645 646 647 648 649

(63) Bauchy, M.; Qomi, M. J. A.; Ulm, F. J.; Pellenq, R. J. M. Order and Disorder in Calcium−silicate−hydrate. J. Chem. Phys. 2014, 140, 214503. (64) Bauchy, M.; Guillot, B.; Micoulaut, M.; Sator, N. Viscosity and Viscosity Anomalies of Model Silicates and Magmas: A Numerical Investigation. Chem. Geol. 2013, 346, 47−56. (65) Phillips, J. C. Topology of Covalent Non-Crystalline Solids I: Short-Range Order in Chalcogenide Alloys. J. Non-Cryst. Solids 1979, 34, 153. (66) Bauchy, M. Topological Constraints and Rigidity of Network Glasses from Molecular Dynamics Simulations. Am. Ceram. Soc. Bull. 2012, 91, 34. (67) Phillips, J. C.; Thorpe, M. F. Constraint Theory, Vector Percolation and Glass Formation. Solid State Commun. 1985, 53, 699−702. (68) Maxwell, J.C. L. On the Calculation of the Equilibrium and Stiffness of Frames. London, Edinburgh, Dublin Philos. Mag. J. Sci. 1864, 27, 294−299.

ACS Paragon Plus Environment

Page 33

The Journal of Physical 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 34 of 35

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 650

(69) Thorpe, M. F. Continuous Deformations in Random Networks. J. Non-Cryst. Solids 1983, 57, 355.

651 652

(70) Wang, F.; Mamedov, S.; Boolchand, P.; Goodman, B.; Chandrasekhar, M. Pressure Raman

653

Effects and Internal Stress in Network Glasses. Phys. Rev. B: Condens. Matter Mater.

654

Phys. 2005, 71, 174201.

655

(71) Chubynsky, M. V.; Brier̀e, M. A.; Mousseau, N. Self- Organization with Equilibration: A Model for the Intermediate Phase in Rigidity Percolation. Phys. Rev. E 2006, 74, 016116.

656 657

(72) Bauchy, M.; Micoulaut, M. Percolative Heterogeneous Topological Constraints and Fragility in Glass-Forming Liquids. EPL Europhys. Lett. 2013, 104, 56002.

658 659

(73) Mascaraque, N.; Bauchy, M.; Smedskjaer, M. M. Correlating the Network Topology of

660

Oxide Glasses with Their Chemical Durability. J. Phys. Chem. B 2017, 121, 1139–1147.

661

(74) Mascaraque, N.; Bauchy, M.; Fierro, J. L. G.; Rzoska, S. J.; Bockowski, M.; Smedskjaer, M.

662

M. Dissolution Kinetics of Hot Compressed Oxide Glasses. J. Phys. Chem. B 2017, 121,

663

9063–9072.

664

(75) Loewenstein, W. The Distribution of Aluminum in the Tetrahedra of Silicates and Aluinates. Am. Mineral. 1954, 39, 92-96.

665 666

(76) Tossell, J.A. A Theoretical Study of the Molecular Basis of the Al Avoidance Rule and of

667

the Spectral Characteristics of Al-O-Al Linkages. Am. Mineral. 1993, 78, 911-920.

668 669 670 671 672 673 674



ACS Paragon Plus Environment

Page 34

Page 35 of 35 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

The Journal of Physical Chemistry

Prepared for Submission to the Journal of Physical Chemistry C (July 2018) 675

“TOC Graphic”

676



677



ACS Paragon Plus Environment

Page 35