Heterogeneous Nucleation and Growth of Barium Sulfate at Organic

May 10, 2016 - “On demand” triggered crystallization of CaCO 3 from solute precursor species stabilized by the water-in-oil microemulsion. Tomasz ...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF PITTSBURGH

Article

Heterogeneous Nucleation and Growth of Barium Sulfate at Organic-Water Interfaces: Interplay between Surface Hydrophobicity and Ba2+ Adsorption Chong Dai, Andrew G. Stack, Ayumi Koishi, Alejandro Fernandez-Martinez, Sang Soo Lee, and Yandi Hu Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.6b01036 • Publication Date (Web): 10 May 2016 Downloaded from http://pubs.acs.org on May 14, 2016

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

Langmuir is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 32

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

Langmuir

1

Heterogeneous Nucleation and Growth of Barium Sulfate at Organic-Water

2

Interfaces: Interplay between Surface Hydrophobicity and Ba2+ Adsorption

3 4 5 6

Chong Dai1, Andrew G. Stack2, Ayumi Koishi3, Alejandro Fernandez-Martinez3, 4, Sang Soo Lee5, and Yandi Hu1* 1

7 8 9 10 11 12 13

2

Department of Civil & Environmental Engineering, University of Houston, Houston, TX 77004

Chemical Science Division, Oak Ridge National Laboratory, P.O. Box 2008, MS-6110, Oak Ridge, Tennessee 37831, United States 3

Univ. Grenoble Alpes, ISTerre, 38041 Grenoble, France 4

CNRS, ISTerre, 38041 Grenoble, France

14 15 16

5 Chemical Sciences and Engineering Division, Argonne National Laboratory, 9700 South Cass Avenue, Argonne, Illinois 60439, United States

17 18

*To Whom Correspondence Should Be Addressed

19

E-mail: yhu11@ uh.edu

20

Phone: (713)743-4285

21

Fax: (713)743-4260

22

http://www.cive.uh.edu/faculty/hu

23 24

Submitted: March 2016

25

Revised: May 2016

26

Langmuir

27

1

ACS Paragon Plus Environment

Langmuir

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

28

Abstract

29

Barium sulfate (BaSO4) is a common scale-forming mineral in natural and engineered

30

systems, yet, the rates and mechanisms of heterogeneous BaSO4 nucleation are not understood.

31

To address these, we created idealized interfaces on which to study heterogeneous nucleation

32

rates and mechanisms, that also are good models for organic-water interfaces: self-assembled

33

thin films terminated with different functional groups (i.e., –COOH, –SH, or mixed –SH &

34

COOH) coated on glass slides. BaSO4 precipitation on coatings from barite-supersaturated

35

solutions (saturation index, SI = 1.1) was investigated using grazing-incidence small angle X-ray

36

scattering. After reaction for 1 hr, little amount of BaSO4 formed on hydrophilic bare and –

37

COOH coated glasses; Meanwhile, BaSO4 nucleation were significantly promoted on

38

hydrophobic –SH and mixed –SH & COOH coatings. This is because substrate hydrophobicity

39

likely affected the interfacial energy and hence thermodynamic favorability of heterogeneous

40

nucleation. The heterogeneous BaSO4 nucleation and growth kinetics were found to be affected

41

by the amount of Ba2+ adsorption onto the substrate and incipient BaSO4 nuclei. The importance

42

of Ba2+ adsorption was further corroborated by the finding that precipitation rate increased under

43

[Ba2+]/[SO42-] concentration ratios > 1. These observations suggest that thermodynamic

44

favorability for nucleation is governed by substrate-water interfacial energy; while given

45

favorable thermodynamics, rate is governed by ion attachment to substrates and incipient nuclei.

46

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

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

47

Langmuir

Introduction

48

Barium sulfate (BaSO4) precipitation plays a crucial role in many natural environment

49

and industrial processes.1-5 In geological reservoirs, BaSO4 can precipitate onto rock surfaces

50

that can change the surface properties5, 6 and clog pores.7 During industrial processes, BaSO4 can

51

precipitate as unwanted scale on oil equipment and facilities as well as on membranes during

52

water treatment.1, 5, 8 In addition, during barite precipitation, it can incorporate toxic elements

53

(e.g., Ra, Sr) into its structure, so the solubility of a disordered phase (e.g., (Ra, Ba)SO4) often

54

governs transport of the contaminant.9

55

BaSO4 precipitation starts with nucleation (i.e., formation of initial BaSO4 nanoparticles),

56

and proceeds with the growth of BaSO4 through the attachment of Ba2+ and SO42- ions, or

57

perhaps other species, onto BaSO4 particles.10 Homogeneous BaSO4 precipitation from

58

supersaturated solution has been studied by monitoring changes in aqueous Ba2+ concentration or

59

turbidity, without distinction of the nucleation and growth stages. The growth rates of BaSO4

60

single crystals have also been studied by tracking step velocity based on atomic force

61

microscopy (AFM) measurements,10-16 or by morphological and chemical structural analysis.17

62

Although the homogeneous BaSO4 nucleation and growth has been extensively studied, few

63

studies have been conducted to understand heterogeneous BaSO4 nucleation and growth; e.g.,

64

Pina and et al. investigated the heterogeneous (Ba, Sr)SO4 nucleation and growth on a barite (001)

65

surface.18 However, to the best of our knowledge, the initial stages of heterogeneous BaSO4

66

nucleation and growth at organic-water interfaces have not been studied,18-20 leading to an

67

inability to predict factors that affect the rates of nucleation and growth. Furthermore, previous

68

studies have reported that aqueous [Ba2+]/[SO42-] concentration ratios can affect homogeneous

69

BaSO4 crystal growth in solution, where excess Ba2+ concentrations can promote the growth of

3

ACS Paragon Plus Environment

Langmuir

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

70

BaSO4 crystals more significantly than the excess SO42-,21, 22 and [Ba2+]/[SO42-] ratio strongly

71

affects step velocities of single crystals..22, 23 Computational studies have also supported the

72

concept of barium ion attachment as a rate controlling process for barite crystal growth.16,24, 25

73

Since these studies have examined growth of barite itself, it is therefore not known how the

74

interfacial interaction of barium and sulfate ions with a substrate will affect heterogeneous

75

nucleation and subsequent precipitation kinetics of BaSO4.

76

Heterogeneous nucleation and growth processes could be affected by the various

77

interfacial interactions among the aqueous species, the BaSO4 particles, and the substrate

78

surfaces. In addition to inorganic solid phases (e.g., minerals, metal), in many natural and

79

engineered aqueous environments, various organics are present forming coatings on minerals,

80

membranes, or engineered surfaces. These organic coatings can generate organic-water

81

interfaces and affect the interactions among substrates, BaSO4, and aqueous species, and

82

therefore can influence heterogeneous BaSO4 nucleation and growth at organic-water

83

interfaces.26-29 However, neither the rates nor the mechanisms of heterogeneous BaSO4

84

nucleation and growth at organic-water interfaces are well understood.

85

The objectives of this study are 2-fold: 1) to quantify heterogeneous BaSO4 nucleation

86

and growth at different organic-water interfaces, from solutions with various Ba2+/SO42- ratios;

87

and 2) To explore the interfacial interactions among Ba2+ and SO42- ions, substrates, and BaSO4

88

precipitates and understand the fundamental interfacial processes controlling heterogeneous

89

BaSO4 nucleation and growth. To achieve these objectives, grazing-incidence small angle X-ray

90

scattering (GISAXS) experiments were utilized to quantify BaSO4 precipitation on bare and

91

organic coated SiO2, and multiple interfacial techniques were used to investigate the

92

physicochemical interactions at organic-water interfaces.

4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

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

Langmuir

93

Experimental Section

94

Preparation of Self-Assembled Thin Films on Glass

95

Glass slides (75 × 25 × 1 mm, Superfrost Plus Micro Slide, VWR international) were cut

96

into small pieces (10 × 10 × 1 mm). The composition and cleaning procedure of the glass slides

97

can be found in our previous publications.30, 31 (3-Triethoxysilyl)propyl succinic acid (TSA) was

98

purchased from Gelest Inc and (3-Mercaptopropyl) trimethoxysilane (MT) was purchased from

99

Sigma-Aldrich. To prepare –SH or –COOH terminated self-assembled thin films, the glass

100

samples were submerged in a mixed solution with 20 wt% TSA or MT and 80 wt% toluene. To

101

prepare the mixed –COOH & –SH terminated self-assembled thin film, glass samples were

102

submerged in a mixed solution, with 10 wt% TSA, 10 wt% MT, and 80 wt% toluene. After 24 h,

103

the coated substrates were rinsed with acetone, to remove the excess amount of loosely-bonded

104

molecules. The bare glass and substrates coated with self-assembled thin films were labeled as

105

“SiO2”, “-COOH”, “-SH” and “-SH & COOH”.

106

The surface morphology of self-assembled thin films was characterized using atomic

107

force microscopy (AFM, Figure S6),32-34 while the roughness, thickness and coverage were

108

estimated using X-ray reflectivity (XR, at beamline 33-BM-C, Advanced Photon Source, APS).

109

Also, surface densities of carboxyl group on –COOH and mixed –SH & COOH coated substrates

110

were measured using Toluidine Blue O method (TBO). Detailed information of these

111

measurements can be found in the Supporting Information.

112

BaSO4 Precipitation Experiments

113

Two sets of barium sulfate (BaSO4) precipitation experiments were conducted. The first

114

set of experiments was conducted in solutions containing 0.04 mM BaCl2 and 0.04 mM Na2SO4

115

on bare and organic coated glass substrates, with fixed saturation index (SI = log(aBaaSO4/Ksp))

5

ACS Paragon Plus Environment

Langmuir

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

116

and [Ba2+]/[SO42-] = 1. The second set of experiments was conducted under varying

117

[Ba2+]/[SO42-] ratios (115:1, 10:1, 1:10, and 1:115, Table 1) on –SH coated glass substrates using

118

a fixed saturation index. As calculated by Geochemist’s Workbench (GWB, Release 9.0,

119

Aqueous Solutions LLC), saturation indices (SI, Table 1, Ksp,barite = 10-9.98) of all solutions with

120

respect to BaSO4 were 1.1, and the solutions’ pH values were ~ 5.6. The calculated pH values

121

agreed well with measured pH values for all solutions, indicating that the thermodynamic

122

database used was appropriate.

123

Grazing Incidence Small-Angle X-ray Scattering (GISAXS) Measurements

124

For each GISAXS experiment, a freshly prepared substrate was placed in a custom built

125

GISAXS cell.35-38 Before each measurement, the substrate surface was aligned to the center of

126

the incident X-ray beam. Then, 0.7 mL of each freshly mixed solution (Table 1) was injected

127

into the cell and the GISAXS measurements started (due to closing the experimental hutch,

128

approximately 2 minutes passed from solution preparation to the initiation of data collection).

129

During measurements, a small incident angle (0.10°) was chosen to yield high reflectivity of the

130

incident X-rays with the energy of 14 keV,39 and scattering from particles formed on the

131

substrates was collected by the 2D detector. Scattering patterns were collected every 2 min for 1

132

h after the data collection had begun. Experiments were performed at the Advanced Photon

133

Source (APS), Argonne National Laboratory (ANL, Beamline 12 ID-B), Argonne, IL.

134

All GISAXS data reduction was conducted with GISAXS-SHOP macro, Igor Pro 6.34

135

(WaveMetrics, Inc., Oregon). The first image of each GISAXS experiment was chosen as the

136

background, which was subtracted from all later scattering images. Based on the two-

137

dimensional (2D) GISAXS images, the barium sulfate particles were polydisperse and there were

138

no obvious scattering patterns along vertical directions. For simplicity, the shape of particles was

6

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

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

Langmuir

139

approximated to a low-resolution, highly symmetric shape, such as a sphere. Therefore, the 2D

140

GISAXS images were converted to one-dimensional (1D) scattering curves, i.e., intensity I

141

plotted vs. wave factor q (Å-1), by cropping along the Yoneda wing (dqz = 0.007 Å-1), where the

142

X-ray scattering signal is the strongest.40, 41 Finally, particle sizes (in radius) were obtained by

143

fitting the scattering curves using a log-normal model of non-interacting spherical particles, with

144

the IRENA package.42 Lorentz-corrected intensity curves (i.e., I × q2 vs. q, Figure S5) were

145

plotted, and the invariant ( Q = ∫ I ( q ) q 2 dq ) evolution (which is proportional to the total particle



0

146

volume of the precipitates) was calculated.43 More detailed information about GISAXS data

147

analysis can be found in our previous publications.30, 31

148

Water Contact Angle Measurements

149

Water contact angle (WCA) measurements were conducted using Dataphysics OCA

150

15EC for bare and organic coated glass substrates. The substrates were placed horizontally on a

151

stage. A drop of distilled water (3 µl) was deposited using a Gastight 500 uL precision syringe

152

(Hamilton Co., USA); meanwhile, a video acquisition was obtained using a CCD camera.

153

Zeta Potential Measurements

154

Two sets of electrophoretic mobility (EM) measurements (Zetasizer Nanoseries, Malvern

155

Instruments) were conducted for understanding the interactions among ions, BaSO4 particles,

156

and substrates. The first set of EM measurements aimed to measure the zeta potential of BaSO4

157

particles in solutions. To generate enough scattering intensity, solutions (Table 1) were prepared

158

and concentrated ~100 times by centrifuging using centrifugal filters (Millipore). The

159

concentrated solutions were then injected into a capillary cell (DTS1070 folded capillary cell,

160

Malvern Instruments) where zeta potentials (ζ) of BaSO4 particles were measured. 7

ACS Paragon Plus Environment

Langmuir

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

161

In the second set, we aimed to measure zeta potential of bare and organic coated glass

162

samples. Cleaned glasses were cut into rectangular pieces (4 mm × 7 mm), adhered to a holder,

163

and then coated with organics as described earlier. After coating, the sample holder was placed

164

in the surface zeta potential cell (ZEN1020, Malvern Instruments) between two electrodes, and a

165

coarse height adjustment was done by an alignment tool. Then, the zeta potential cell was

166

inserted into the standard cuvette, and 1.2 mL tracer solution of graphene oxide (GO)

167

nanoparticles (pH = 5.7 ± 0.1) was injected into the cuvette. During surface zeta potential

168

measurements, the apparent electrophoretic mobility of the GO tracer particles was measured at

169

125, 250, 500, 750, and 3000 µm from the substrate surface, and the zeta potentials of the

170

substrates were calculated. Triplicate measurements were conducted with good reproducibility.

171

More information on surface zeta potential measurement and calculations can be found in

172

Supporting Information.

173

Ba2+ and SO42- Ion Adsorption

174

Ba2+ and SO42- ion adsorption was investigated using dynamic light scattering (DLS),

175

quartz crystal microbalance with dissipation (QCM-D, Q-sense E1, Biolin Scientific), and

176

attenuated total reflectance Fourier transform infrared spectroscopy (ATR-FTIR). Surface zeta

177

potentials of bare or organic coated glass samples were measured in the presence of 0.04 mM

178

BaCl2 or Na2SO4 at pH =5.7 ± 0.1. The differences in surface zeta potentials of substrates in the

179

presence and absence of 0.04 mM BaCl2 or Na2SO4 were evaluated to indicate Ba2+ or SO42- ion

180

adsorption.

181

Sensors coated with SiO2 (QSX 303, Q-Sense) were purchased from Biolin Scientific,

182

and were coated with self-assembled thin films following the same procedures of coating self-

183

assembled thin films on glass substrates. Then the sensors were placed into the QCM-D flow 8

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

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

Langmuir

184

module, and mounted on the Q-sense chamber platform. For initial stabilization, ultrapure water

185

was pumped through the flow module at 0.2 mL/min. Then, the testing solution with 0.1 mM

186

BaCl2 or Na2SO4 at pH =5.7 ± 0.1 was pumped into the flow module at a constant flow rate of

187

0.2 mL/min. The arrows in the Figures of QCM-D measurements indicate switch points from

188

ultrapure water to the testing solutions (Figure 4, Figure S3 and S4 in Supporting Information).

189

Changes in resonance frequency (∆f) and energy dissipation (∆D) of the sensors were measured,

190

which indicated mass changes on the sensors caused by ion adsorption. More details of QCM-D

191

measurements can be found in Supporting Information and our previous publication.30

192

Attenuated total reflectance Fourier transform infrared (ATR-FTIR) spectroscopy was

193

also used to detect Ba2+ or SO42- adsorption on self-assembled thin films by looking at changes in

194

their vibrational frequencies. Self-assembled thin film -coated glasses were exposed to 0.1 M

195

BaCl2 or Na2SO4 solutions for 2 hrs with solutions’ pH values adjusted to 5.7 ± 0.1. Then, ATR-

196

FTIR measurements were performed using a Nicolet iS 10 (Thermo Scientific) instrument

197

equipped with a single-reflection diamond crystal.

198

Results and Discussion

199

Heterogeneous BaSO4 Precipitation on Bare and Coated Glasses Measured by GISAXS

200

GISAXS scattering intensities from BaSO4 particles on different substrates are plotted

201

against scattering vector (q, Å-1) (Figure 1) with time-dependent changes resulting from the

202

formation of BaSO4 at the surfaces. Almost no change was observed on bare and –COOH coated

203

SiO2 surfaces, indicating a low density of BaSO4 particles. Contrarily, the scattering intensity on

204

–SH and the mixed –SH & COOH coatings continuously increased throughout the 60 min

205

experiments, indicating continuous BaSO4 precipitation at these two coated surfaces. The

9

ACS Paragon Plus Environment

Langmuir

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 32

206

GISAXS observations suggested the promoted heterogeneous BaSO4 precipitation on –SH and

207

the mixed –SH & COOH coatings.

208

Also, there is a notable difference in the time-dependent changes between these two

209

systems. For –SH coated surfaces, significant enhancements to the scattering intensity were

210

observed mostly in the lower q region, indicating increases in particle size. In contrast, for the

211

mixed –SH & COOH coated substrates, the enhancements were almost proportional over the

212

entire q range, indicating increase in number of particles without significant change in size.

213

Particle size evolutions throughout the 60 min experiments were obtained from model fitting.

214

The average radii of BaSO4 particles on –SH coated substrate grew from 1.3 ± 0.2 to 8.1 ± 0.6

215

nm over the course of 60 min; On the other hand, the BaSO4 particles on mixed –SH & COOH

216

coated substrate grew slowly from 1.4 ± 0.2 to 3.3 ± 0.4 nm in the same amount of time (Figure

217

2A). The evolution of total particle volumes throughout the 60 min reactions are shown in Figure

218

2B. Interestingly, the smaller BaSO4 precipitates on the mixed –SH & COOH coating had larger

219

total particle volumes compared with those on –SH coating (Figure 2B), indicating a larger

220

number of particles on the mixed –SH & COOH coating. As nucleation will increase the number

221

of particles and growth will increase particle size, we concluded that heterogeneous BaSO4

222

nucleation was faster on the mixed –SH & COOH coatings, while growth was faster on –SH

223

coating.

224

To understand the favored heterogeneous BaSO4 precipitation on –SH and the mixed –

225

SH & COOH coatings than the bare and –COOH coated glasses, also to understand the different

226

nucleation and growth rates of BaSO4 on –SH and the mixed –SH & COOH coatings, varied

227

interfacial techniques were employed to understand the physicochemical interactions at the

228

organic-water interfaces. The mechanisms controlling heterogeneous BaSO4 nucleation and

10

ACS Paragon Plus Environment

Page 11 of 32

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

Langmuir

229

growth are discussed from both thermodynamic and kinetic points of view in the following

230

subsections.

231

Substrate Hydrophobicity Controls Thermodynamic Favorability of Heterogeneous BaSO4

232

Nucleation

233

Nucleation is the first step during mineral precipitation, and according to classical

234

nucleation theory (CNT), at constant supersaturation, differences in free energy barriers for

235

heterogeneous nucleation result from changes in interfacial energy (∆Ginterface), as ∆  =   +   −   (1)

236

where Alp and Asp are the surface areas of the precipitate–liquid and precipitate–substrate

237

interfaces; σlp, σsp, and σls are the interfacial energies of precipitates–liquid, precipitate–substrate,

238

and liquid–substrate. For the BaSO4-water interface, free energies of σlp = 135-137 mJ/m2 have

239

been observed,44, 45 although much lower values of 20-25 mJ/m2 have also been reported for

240

individual surfaces.46,

241

substrates. The term σsp is related with the structural mismatch between the substrate and the

242

BaSO4 particles, but this is not known for BaSO4-organic interfaces. The term σls has been

243

reported to correlate well with the hydrophobicity of substrates.48 Based on water contact angle

244

measurements, the hydrophobicity of substrates followed the order: -SH & COOH (89.5 ± 0.6°) >

245

-SH (66.4 ± 0.4°) > -COOH (26.4 ± 3.6°) > SiO2 surface (8.3 ± 0.5°) (Figure S1), indicating that

246

the values of σls followed the order: -SH & COOH > -SH > -COOH > SiO2 surface. Based on

247

eqn. (1), a higher σls correlates with a lower interfacial energy for heterogeneous nucleation. The

248

high σls for –SH and mixed –SH&COOH coatings resulted in the lower nucleation free energy

249

barriers, which therefore promoted heterogeneous BaSO4 nucleation on the coatings (Figure 1C

250

and 1D). On the other hand, the low σls for bare and –COOH coated substrates inhibited

47

Nevertheless, the σlp value will be independent of the presence of

11

ACS Paragon Plus Environment

Langmuir

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

251

heterogeneous BaSO4 precipitation on them. The water contact angle and GISAXS

252

measurements (Figure 1) suggested that substrate hydrophobicity, σls, was a key parameter

253

controlling the thermodynamic favorability of heterogeneous BaSO4 nucleation. This conclusion

254

is also in agreement with our previous study on ferrihydrite nucleation, which was also found to

255

be controlled by substrate hydrophobicity.35

Page 12 of 32

256

It is speculated that the hydrophobicity of the self-assembled thin films can be attributed

257

to both molecular and structural properties of the systems. We first considered the difference in

258

charged state between the two functional groups (–SH and –COOH). The pKa values of –SH and

259

–COOH groups were reported to be in the ranges of 8.4–10.3 and 4.2–4.8, respectively.49-53

260

Accordingly, under our experimental condition (pH = 5.7 ± 0.1), less than 1% -SH groups were

261

deprotonated to –S-, resulting in the high hydrophobicity of the –SH coating. Meanwhile, more

262

than 99 % -COOH were deprotonated to –COO-, resulting in the low hydrophobicity of the –

263

COOH coating. Hydrophobicity can also be controlled by surface morphology. However, our

264

TBO, AFM, and XR measurements (Figure S6, S8, and Table S1) showed that all the self-

265

assembled thin films had similar surface coverage, thickness, and roughness, suggesting that they

266

were not the main factors that led to the large hydrophobicity differences among the substrates.

267

Therefore, the differences in substrate hydrophobicity were mainly due to the differences in

268

terminal functional groups of the organic films. Details of TBO, AFM, and XR measurements

269

and results can be found in Supporting Information.

270

In summary, from a thermodynamic point of view, substrate hydrophobicity determines if

271

heterogeneous BaSO4 nucleation on substrates is favorable or not. Therefore, heterogeneous

272

nucleation occurred only on highly hydrophobic surfaces (i.e., -SH and –SH & COOH coatings),

273

which likely lowered the interfacial energy. Contrarily, the low hydrophobicity of –COOH

12

ACS Paragon Plus Environment

Page 13 of 32

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

Langmuir

274

terminated substrate resulted in high interfacial energy, thus making heterogeneous BaSO4

275

nucleation unfavorable.

276

Amount of Ba2+ Adsorption onto Substrates Affected Heterogeneous BaSO4 Nucleation and

277

Growth Rates

278

The favored heterogeneous BaSO4 precipitation on –SH and the mixed –SH & COOH

279

coatings was explained from thermodynamic point of view in the above subsection. In this

280

subsection, mechanisms controlling the different nucleation and growth rates of BaSO4 on –SH

281

and the mixed –SH & COOH coatings was explored from kinetic point of view. Just as growth of

282

existing barite crystals is enacted by Ba2+ and SO42- adsorption onto reactive sites,10, 24, 25 so

283

heterogeneous barite nucleation on the organic coatings might be enhanced by Ba2+ and SO42-

284

ions adsorption at the interfaces. In stagnant aqueous environments, like batch reactors with

285

fixed amounts of Ba2+ and SO42- ions, the ion adsorption onto substrates may compete with ion

286

adsorption onto pre-existing barite crystals, and thus can affect the rates of heterogeneous BaSO4

287

nucleation and growth. The relationship between nucleation density and/or rate and ion sorption

288

onto the substrates was investigated using three interfacial techniques: QCM-D, ATR-FTIR, and

289

surface zeta potential measurements. QCM-D provides quantitative measurements of the mass of

290

ions interacting with a surface; ATR-FTIR provides confirmation of inner-sphere adsorption; and

291

surface zeta potential measurements show how surface charge densities are affected by ion

292

sorption.

293

QCM-D measurements performed under a flowing Na2SO4 solution show that no

294

observable SO42- adsorption onto any substrates occurred (Figure S4 in Supporting Information).

295

In contrast, while no Ba2+ adsorption was observed onto bare SiO2, significant amounts of Ba2+

13

ACS Paragon Plus Environment

Langmuir

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 32

296

ion adsorbed onto –COOH, -SH, and mixed –SH & COOH coated sensors, with values of 23.5 ±

297

2.9, 6.2 ± 1.5, and 30.6 ± 8.8 ng/cm2, respectively (Figure 4).

298

ATR-FTIR experiments (Figure 5 and S7) also showed no changes in the spectra from

299

the substrates exposed to Na2SO4 solution (Figure S7), whereas new vibrational modes appeared

300

in the range of 2980-3020 cm-1 for the –COOH and mixed –SH&COOH coatings exposed to

301

BaCl2 solutions (Figure 5), indicating that Ba2+ adsorption occurred in an inner-sphere state

302

bound to these two self-assembled thin films (Figure 5). On the contrary, no interaction was

303

observed for Ba2+ with bare SiO2 substrate, and only slight interaction was observed for Ba2+

304

with –SH coating.

305

Zeta potentials of substrates before (ζ) and after (ζ-Ba2+) exposed to BaCl2 solution are

306

summarized in Table 2. Zeta potentials of bare and –SH coated SiO2 in the presence of 0.04 mM

307

Ba2+ (-67.8 ± 1.7 mV and -87.9 ± 8.6 mV, respectively) were similar to those measured without

308

Ba2+ (-64.6 ± 2.6 mV and -82.3 ± 12.5 mV, respectively), indicating the number of charged sites

309

was not significantly affected by Ba2+ adsorption. For –COOH and –SH & COOH coated SiO2,

310

the surface zeta potentials in the presence of 0.04 mM Ba2+ were less negative (-60.3 ± 0.35 and -

311

67.2 ± 6.9 mV, respectively) than in the absence of Ba2+ (-72.6 ± 4.7 and -84.3 ± 6.0 mV,

312

respectively), indicating a significant amount of surface sites were occupied by adsorbed Ba2+.

313

In summary, great consistency was achieved among multiple interfacial techniques: no

314

SO42- adsorption was observed onto any substrate, and the amount of Ba2+ adsorption onto

315

substrates followed the order: on mixed –SH & COOH > –COOH > –SH. The mechanisms

316

controlling Ba2+ adsorption on self-assembled thin films were considered. Previous studies have

317

shown that divalent ions (Ca2+ and Mg2+) can adsorb onto deprotonated –COO- forming –

318

COOCa+ and –COOMg+ species.54, 55 Considering the pKa values of –SH and –COOH groups,4914

ACS Paragon Plus Environment

Page 15 of 32

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

Langmuir

319

53

320

-COOH were deprotonated to –COO-. Therefore, significant amounts of Ba2+ adsorption onto –

321

COOH and mixed –SH &COOH coatings occurred. Furthermore, based on TBO method, surface

322

densities of carboxyl groups on –COOH (i.e., 1.6 ± 0.6 µg/cm2) and mixed –COOH & SH (i.e.,

323

1.3 ± 0.8 µg/cm2) coated substrates were similar, indicating similar amounts of –COOH coatings

324

on both substrates. Therefore, the amount of Ba2+ adsorption onto mixed –SH & COOH coatings

325

was higher than on –COOH coating. The mechanisms controlling the surface coverages of –SH

326

or –COOH on substrates from a mixed –SH & COOH precursor solution were not clear, which is

327

an important direction under our further investigation for future publications.

328

under our experimental condition, less than 0.01% -SH were deprotonated to –S-, while > 99 %

Previous studies reported that Ba2+ dehydration and adsorption, instead of SO42-, onto

329

BaSO4 particles, controls the growth of BaSO4 in solution.11,

330

nucleation and growth studied here, the mixed –SH & COOH coatings can adsorb significant

331

amount of Ba2+. The adsorbed Ba2+ may have acted as nucleation centers and promoted

332

heterogeneous BaSO4 nucleation on the mixed –SH & COOH. As our experiments were

333

conducted in a stagnant system with limited amount of Ba2+, Ba2+ adsorption onto the mixed –

334

SH & COOH coatings resulted in less available Ba2+ to attach to incipient BaSO4 particles, thus

335

slowed down BaSO4 growth. For the –SH coating, there was only a small amount of Ba2+

336

adsorbed, thus perhaps creating fewer nuclei on the coating, which resulted in faster

337

heterogeneous BaSO4 growth. Note that the -COOH coating had a significant amount of Ba2+

338

adsorbed, but did not display substantial nucleation measured by GISAXS. This finding is

339

interesting because it suggests that extensive ion adsorption onto the substrate alone cannot drive

340

nucleation, it must coincide with a thermodynamically favorable substrate-nucleus interaction.

341

Ba2+ Adsorption onto BaSO4: Effects of Aqueous Ba2+/SO42- Ratio on BaSO4 Growth

15

ACS Paragon Plus Environment

56-58

For heterogeneous BaSO4

Langmuir

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 32

342

To confirm the effects of differences in barium and sulfate ion adsorption on

343

heterogeneous BaSO4 precipitation, BaSO4 precipitation (SI = 1.1) on –SH was quantified under

344

various aqueous [Ba2+]/[SO42-] ratios (115:1, 10:1, 1:1, 1:10, and 1:115).

345

[Ba2+]/[SO42-] = 1 (where particles grew from 1.3 to 8.1 nm within 60 min), with [Ba2+]/[SO42-] >

346

1, faster growth of BaSO4 particles was observed (Figure 2 and 3). As the reaction went on, the

347

increases in both particle size and particle volume occurred more rapidly (Figure 2). Adversely,

348

with [Ba2+]/[SO42-] < 1, the precipitation of BaSO4 particles on –SH coated substrates were

349

slower than [Ba2+]/[SO42-] = 1, with slower increases in both particle sizes and particle volumes

350

(Figure 2 and 3). Dehydration and attachment of Ba2+ has been shown to be a rate limiting step

351

for single crystal BaSO4 growth.24 The observation of heterogeneous precipitation of BaSO4

352

particles on the –SH coating was enhanced with excess amount of Ba2+, is consistent with barium

353

ion attachment also being rate limiting during heterogeneous nucleation and/or initial stages of

354

growth.

355

Summary and Conclusions

Compared with

356

Existing barite growth models do not reflect the initial stages of nanoparticle nucleation

357

and growth processes, especially at organic-water interfaces, but instead tend to focus on the

358

homogeneous growth of existing crystals in solution. In this study, using GISAXS, the kinetics

359

of heterogeneous BaSO4 nucleation and growth at organic-water interfaces under varied aqueous

360

[Ba2+]/[SO42-] ratios were quantified for the first time.

361

Multiple interfacial characterization techniques were employed to probe the interfacial

362

interactions among aqueous ions, the barite particles and the substrates, and the controlling

363

mechanisms of heterogeneous BaSO4 precipitation were systematically explored. Heterogeneous

364

BaSO4 nucleation was found to correlate first with substrate hydrophobicity, which likely affects

16

ACS Paragon Plus Environment

Page 17 of 32

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

Langmuir

365

the favorability of heterogeneous nucleation. Given a sufficiently hydrophobic substrate, inner-

366

sphere Ba2+ adsorption, confirmed by three interfacial techniques, was found to strongly affect

367

both heterogeneous BaSO4 nucleation and subsequent particle growth. Ba2+ adsorption onto the

368

substrates seemingly creates a higher density of nucleation sites; while Ba2+ dehydration and

369

adsorption onto BaSO4 particles has been previously shown to control BaSO4 growth rates.

370

Based on our findings, one can envision a process-based augmentation of classical nucleation

371

theory where the thermodynamic favorability for nuclei formation (based on interfacial energies)

372

determines if nucleation can proceed, but given a favorable interaction, nucleation rates are

373

governed by adsorption of mineral’s constituent ions onto a substrate (i.e., determines nucleation

374

density), and subsequent growth is driven by the attachment and detachment of ions from

375

solution onto the incipient nuclei.

376

Acknowledgements

377

This material is primarily based upon work supported by the U.S. Department of Energy,

378

Office of Science, Office of Basic Energy Sciences, Chemical Sciences, Geosciences, and

379

Biosciences Division, and support C. Dai, A.G. Stack, S.S. Lee, and Y. Hu. A. Koishi and A.

380

Fernandez-Martinez acknowledge funding from the EC2CO (CNRS-INSU) NUCLEATION

381

program, Labex OSUG@2020 (Investissements d’avenir – ANR10 LABX56), and the PICS

382

(CNRS) No. 06736. This project has been funded in part with funds from the State of Texas as

383

part of the program of the Texas Hazardous Waste Research Center. The contents do not

384

necessarily reflect the views and policies of the sponsor nor does the mention of trade names or

385

commercial products constitute endorsement or recommendation for use. We thank Dr. Xiaobing

386

Zuo and Dr. Byeongdu Lee for valuable discussion on GISAXS experiments and data analysis at

387

beamline 12ID-B. Use of the facilities at beamlines Sector 12-ID-B and 33-BM-C at APS was

17

ACS Paragon Plus Environment

Langmuir

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

388

supported by the US Department of Energy, Office of Science, Office of Basic Energy Science,

389

under Contract No. DE-AC02-06CH11357.

390

Supporting Information:

391

Experimental operation and data analysis details, water contact angle measurements of

392

substrates (Figure S1), GISAXS scattering intensities and Lorentz corrected intensities (Figure

393

S2 and S5), QCM-D measurements with NaCl and Na2SO4 solutions (Figure S3 and S4), AFM

394

(Figure S6), ATR-FTIR (Figure S7), and X-ray reflectivity measurements (Figure S8 and Table

395

S1) of substrates. This material is available free of charge via the Internet at http://pubs.acs.org.

18

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

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

Langmuir

396

References

397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439

1. Fan, C.; Kan, A.; Zhang, P.; Lu, H.; Work, S.; Yu, J.; Tomson, M. Scale prediction and inhibition for oil and gas production at high temperature/high pressure. SPE. J. 2012, 17, 379392. 2. Fan, C.; Kan, A. T.; Zhang, P.; Tomson, M. B. Barite nucleation and inhibition at 0 to 200 degree C with and without thermodynamic hydrate inhibitors. SPE. J. 2011, 16, 440-450. 3. Jones, F.; Radomirovic, T.; Ogden, M. I. Effect of solution silicate on the precipitation of barium sulfate. Cryst. Growth Des. 2012, 12, 3057-3065. 4. Kan, A. T.; Tomson, M. B. Scale prediction for oil and gas production. SPE. J. 2012, 17, 362-378. 5. Merdhah, A. B. B.; Yassin, A. A. M. Laboratory study on precipitation of barium sulfate in Malaysia sandstone cores. The open petroleum engineering 2009, 2, 1-11. 6. Naehr, T.; Stakes, D.; Moore, W. Mass wasting, ephemeral fluid flow, and barite deposition on the California continental margin. Geology 2000, 28, 315-318. 7. Frenier, W. W.; Ziauddin, M. Formation, Removal, and Inhibition of Inorganic Scale in the Oilfield Environment; Society of Petroleum Engineers2008. 8. Stamatakis, E.; Chatzichristos, C.; Sagen, J.; Stubos, A. K.; Palyvos, I.; Muller, J.; Stokhan, J. A. An integrated radiotracer approach for the laboratory evaluation of scale inhibitors performance in geological environments. Chem. Eng. Sci. 2006, 61, 7057-7067. 9. Martin, A.; Crusius, J.; McNee, J. J.; Yanful, E. The mobility of radium-226 and trace metals in pre-oxidized subaqueous uranium mill tailings. Appl. Geochem. 2003, 18, 1095-1110. 10. Pina, C. M.; Becker, U.; Risthaus, P.; Bosbach, D.; Putnis, A. Molecular-scale mechanisms of crystal growth in barite. Nature 1998, 395, 483-486. 11. Benton, W. J.; Collins, I. R.; Grimsey, I. M.; Parkinson, G. M.; Rodger, S. A. Nucleation, growth and inhibition of barium sulfate-controlled modification with organic and inorganic additives. Faraday Discuss. 1993, 95, 281-297. 12. Godinho, J. R.; Stack, A. G. Growth Kinetics and Morphology of Barite Crystals Derived from Face-Specific Growth Rates. Cryst. l Growth Des. 2015, 15, 2064-2071. 13. Judat, B.; Kind, M. Morphology and internal structure of barium sulfate—derivation of a new growth mechanism. J Colloid Interface Sci. 2004, 269, 341-353. 14. Marchisio, D. L.; Barresi, A. A.; Garbero, M. Nucleation, growth, and agglomeration in barium sulfate turbulent precipitation. AIChE journal 2002, 48, 2039-2050. 15. Higgins, S. R.; Bosbach, D.; Eggleston, C. M.; Knauss, K. G. Kink dynamics and step growth on barium sulfate (001): a hydrothermal scanning probe microscopy study. J. Phys. Chem. 2000, 104, 6978-6982. 16. Becker, U.; Risthaus, P.; Bosbach, D.; Putnis, A. Selective attachment of monovalent background electrolyte ions and growth inhibitors to polar steps on sulfates as studied by molecular simulations and AFM observations. Mol. Simul. 2002, 28, 607-632. 17. Weber, J.; Barthel, J.; Brandt, F.; Klinkenberg, M.; Breuer, U.; Kruth, M.; Bosbach, D. Nano-structural features of barite crystals observed by electron microscopy and atom probe tomography. Chem. Geol. 2016. 18. Pina, C. M.; Enders, M.; Putnis, A. The composition of solid solutions crystallising from aqueous solutions: the influence of supersaturation and growth mechanisms. Chem. Geol. 2000, 168, 195-210.

19

ACS Paragon Plus Environment

Langmuir

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

440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484

Page 20 of 32

19. Pina, C.; Putnis, A.; Astilleros, J. The growth mechanisms of solid solutions crystallising from aqueous solutions. Chem. Geol. 2004, 204, 145-161. 20. Pina, C.; Putnis, A. The kinetics of nucleation of solid solutions from aqueous solutions: a new model for calculating non-equilibrium distribution coefficients. Geochim. Cosmochim. Acta 2002, 66, 185-192. 21. Kucher, M.; Babic, D.; Kind, M. Precipitation of barium sulfate: Experimental investigation about the influence of supersaturation and free lattice ion ratio on particle formation. Chem. Eng. Process. 2006, 45, 900-907. 22. Kowacz, M.; Putnis, C. V.; Putnis, A. The effect of cation:anion ratio in solution on the mechanism of barite growth at constant supersaturation: Role of the desolvation process on the growth kinetics. Geochim. Cosmochim. Acta 2007, 71, 5168-5179. 23. Bracco, J. N. Growth of Sparingly-Soluble AB-type Minerals as a Function of their A: B Ratio. Wright State University2015. 24. Stack, A. G.; Raiteri, P.; Gale, J. D. Accurate rates of the complex mechanisms for growth and dissolution of minerals using a combination of rare-event theories. J. Am. Chem. Soc. 2012, 134, 11-14. 25. Piana, S.; Jones, F.; Gale, J. D. Assisted desolvation as a key kinetic step for crystal growth. J. Am. Chem. Soc. 2006, 128, 13568-13574. 26. Mann, S.; Archibald, D. D.; Didymus, J. M.; Douglas, T.; Heywood, B. R.; Meldrum, F. C.; Reeves, N. J. Crystallization at Inorganic-organic Interfaces: Biominerals and Biomimetic Synthesis. Science 1993, 261, 1286-1292. 27. Rieke, P. C.; Marsh, B. D.; Wood, L. L.; Tarasevich, B. J.; Liu, J.; Song, L.; Fryxell, G. E. Aqueous solution deposition kinetics of iron oxyhydroxide on sulfonic acid terminated selfassembled monolayers. Langmuir 1995, 11, 318-326. 28. Heywood, B. R.; Mann, S. Template-directed inorganic crystallization: oriented nucleation of barium sulfate under Langmuir monolayers of an aliphatic long chain phosphonate. Langmuir 1992, 8, 1492-1498. 29. Zhao, X.; Yang, J.; McCormick, L. D.; Fendler, J. Epitaxial formation of lead sulfide crystals under arachidic acid monolayers. J. Phys. Chem. 1992, 96, 9933-9939. 30. Dai, C.; Hu, Y. Fe(III) Hydroxide Nucleation and Growth on Quartz in the Presence of Cu(II), Pb(II), and Cr(III): Metal Hydrolysis and Adsorption. Environ. Sci. Technol. 2015, 49, 292-300. 31. Dai, C.; Zuo, X.; Cao, B.; Hu, Y. Homogeneous and Heterogeneous (Fex, Cr1-x)(OH)3 Precipitation: Implications for Cr Sequestration. Environ. Sci. Technol. 2016, 50, 1741-1749. 32. Uchida, E.; Uyama, Y.; Ikada, Y. Sorption of low-molecular-weight anions into thin polycation layers grafted onto a film. Langmuir 1993, 9, 1121-1124. 33. Chua, K.-N.; Lim, W.-S.; Zhang, P.; Lu, H.; Wen, J.; Ramakrishna, S.; Leong, K. W.; Mao, H.-Q. Stable immobilization of rat hepatocyte spheroids on galactosylated nanofiber scaffold. Biomaterials 2005, 26, 2537-2547. 34. Ray, J. R.; Lee, B.; Baltrusaitis, J.; Jun, Y.-S. Formation of iron (III)(hydr) oxides on polyaspartate-and alginate-coated substrates: effects of coating hydrophilicity and functional group. Environ. Sci. Technol. 2012, 46, 13167-13175. 35. Hu, Y.; Neil, C.; Lee, B.; Jun, Y.-S. Control of heterogeneous Fe (III)(hydr) oxide nucleation and growth by interfacial energies and local saturations. Environ. Sci. Technol. 2013, 47, 9198-9206.

20

ACS Paragon Plus Environment

Page 21 of 32

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

485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529

Langmuir

36. Hu, Y.; Lee, B.; Bell, C.; Jun, Y.-S. Environmentally abundant anions influence the nucleation, growth, ostwald ripening, and aggregation of hydrous Fe (III) oxides. Langmuir 2012, 28, 7737-7746. 37. Jun, Y.-S.; Lee, B.; Waychunas, G. A. In situ observations of nanoparticle early development kinetics at mineral− water interfaces. Environ. Sci. Technol. 2010, 44, 8182-8189. 38. Hu, Y.; Li, Q.; Lee, B.; Jun, Y.-S. Aluminum affects heterogeneous Fe (III)(Hydr) oxide nucleation, growth, and Ostwald ripening. Environ. Sci. Technol. 2013, 48, 299-306. 39. Henke, B. L.; Gullikson, E. M.; Davis, J. C. X-ray interactions: photoabsorption, scattering, transmission, and reflection at E= 50-30,000 eV, Z= 1-92. Atomic data and nuclear data tables 1993, 54, 181-342. 40. Lee, B.; Seifert, S.; Riley, S. J.; Tikhonov, G.; Tomczyk, N. A.; Vajda, S.; Winans, R. E. Anomalous grazing incidence small-angle x-ray scattering studies of platinum nanoparticles formed by cluster deposition. J. Chem. Phys 2005, 123, 074701. 41. Renaud, G.; Lazzari, R.; Revenant, C.; Barbier, A.; Noblet, M.; Ulrich, O.; Leroy, F.; Jupille, J.; Borensztein, Y.; Henry, C. R. Real-time monitoring of growing nanoparticles. Science 2003, 300, 1416-1419. 42. Ilavsky, J.; Jemian, P. R. Irena: tool suite for modeling and analysis of small-angle scattering. J. Appl. Crystallogr. 2009, 42, 347-353. 43. Li, Y.; Gao, T.; Chu, B. Synchrotron SAXS studies of the phase-separation kinetics in a segmented polyurethane. Macromolecules 1992, 25, 1737-1742. 44. Prieto, M.; Putnis, A.; Fernandez-Diaz, L. Crystallization of solid solutions from aqueous solutions in a porous medium: zoning in (Ba, Sr) SO 4. Geol.Mag. 1993, 130, 289-299. 45. Sangwal, K. On the estimation of surface entropy factor, interfacial tension, dissolution enthalpy and metastable zone-width for substances crystallizing from solution. J. Cryst. Growth 1989, 97, 393-405. 46. Chibowski, E.; Hołysz, L. Influence of tetradecylamine chloride on the surface freeenergy components and flotability of barite. J. Mater. Sci. 1992, 27, 5221-5228. 47. Stack, A. G. Molecular Dynamics Simulations of Solvation and Kink Site Formation at the {001} Barite− Water Interface†. J. Phys. Chem. C 2009, 113, 2104-2110. 48. Dong, J.; Mao, G.; Hill, R. M. Nanoscale aggregate structures of trisiloxane surfactants at solid-liquid interface. Langmuir 2004, 20, 2695-2700. 49. Denison, P.; Jones, F.; Watts, J. X-ray photoelectron spectroscopic analysis of bariumlabelled carbon fibre surfaces. J. Mater. Sci. 1985, 20, 4647-4656. 50. Muto, S.; Ogata, H.; Kamiya, Y. Oxidative addition of organic acids to dioxygen complexes and formation of hydrogen peroxide. Chem. Lett. 1975, 4, 809-812. 51. O'neill, P. Pulse radiolytic study of the interaction of thiols and ascorbate with OH adducts of dGMP and dG: implications for DNA repair processes. Radiat. Res. 1983, 96, 198210. 52. Sober, H. E. Handbook of Biochemistry: The chemial Rubber Co., 1968. 53. Myung, S.; Lee, M.; Kim, G. T.; Ha, J. S.; Hong, S. Large ‐ Scale “ Surface ‐ Programmed Assembly” of Pristine Vanadium Oxide Nanowire‐Based Devices. Adv. Mater. 2005, 17, 2361-2364. 54. Tang, C. Y.; Huang, Z.; Allen, H. C. Binding of Mg2+ and Ca2+ to palmitic acid and deprotonation of the COOH headgroup studied by vibrational sum frequency generation spectroscopy. J. Phys. Chem. B 2010, 114, 17068-17076.

21

ACS Paragon Plus Environment

Langmuir

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

530 531 532 533 534 535 536 537 538 539 540 541 542

Page 22 of 32

55. Ito, T. Ion-channel-mimetic sensor for trivalent cations based on self-assembled monolayers of thiol-derivatized 4-acyl-5-pyrazolones on gold. J. Electroanal. Chem. 2001, 495, 87-97. 56. Kucher, M.; Babic, D.; Kind, M. Precipitation of barium sulfate: Experimental investigation about the influence of supersaturation and free lattice ion ratio on particle formation. CHem. Eng. Process 2006, 45, 900-907. 57. Fischer, R. B.; Rhinehammer, R. B. Rapid precipitation of barium sulfate. Anal. Chem. 1953, 25, 1544-1548. 58. Wong, D.; Jaworski, Z.; Nienow, A. Effect of ion excess on particle size and morphology during barium sulphate precipitation: an experimental study. Chem. Eng. Sci. 2001, 56, 727-734. 59. Pulido, T.; Adzerikho, I.; Channick, R. N.; Delcroix, M.; Galiè, N.; Ghofrani, H.-A.; Jansa, P.; Jing, Z.-C.; Le Brun, F.-O.; Mehta, S. Macitentan and morbidity and mortality in pulmonary arterial hypertension. N. Engl. J. Med. 2013, 369, 809-818.

543

22

ACS Paragon Plus Environment

Page 23 of 32

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

Langmuir

Table 1. Solution Conditions

544

Aqueous Ba2+/SO42ratio (R) R = 1:1 R = 10:1 R = 115:1 R = 1:10 R = 1:115

BaCl2, mM

Na2SO4, mM

ISa, mM

pHb

SIc

0.040 0.120 0.400 0.012 0.004

0.040 0.012 0.003 0.120 0.460

0.24 0.39 1.20 0.39 1.37

5.61 5.61 5.61 5.62 5.62

1.1 1.1 1.1 1.1 1.1

545

Note: ISa: Ionic Strength. pHb: initial pH values, calculated by GWB, which were consistent with

546

measured pH values. SIc: Saturation indices (SI= Log (Q/Ksp) with respect to barite, which were

547

calculated using GWB.

548

23

ACS Paragon Plus Environment

Langmuir

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

Table 2. Substrate Properties

549

Water contact ζa, ζ-Ba2+ b, Ba2+ adsorption, angle (°) mV mV ng/cm2 SiO2 8.3 ± 0.5 -64.6 ± 2.6 -67.8 ± 1.7 N/A -COOH coating 26.4 ± 3.6 -72.6 ± 4.7 -60.3 ± 0.35 23.5 ± 2.9 -SH coating 66.4 ± 0.4 -82.3 ± 12.5 -87.9 ± 8.6 6.2 ± 1.5 -SH & COOH coating 89.5 ± 0.6 -84.3 ± 6.0 -67.2 ± 6.9 30.6 ± 8.8 a 2+ 2+ b Note: ζ : The zeta potentials of substrates in the absence of Ba ; ζ-Ba : The zeta potentials of Substrates

550 551

Page 24 of 32

substrates in the presence of 0.04 mM Ba2+.

552 553

554

555

24

ACS Paragon Plus Environment

Page 25 of 32

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

Langmuir

556

List of Figures

557

Figure 1. GISAXS scattering intensities generated by BaSO4 particles formed in the mixed

558

solutions of 0.04 mM BaCl2 and 0.04 mM Na2SO4 (pH = 5.6 ± 0.2) on bare glass SiO2

559

(Figure 1A, SiO2), and self-assembled thin films terminated with -COOH (Figure 1B),

560

-SH (Figure 1C), and mixed -SH & COOH (Figure 1D). The experimental data are

561

shown as colored dots, while the fitted curves are shown as black lines. Particle

562

growth are indicated by the red arrows (Figure 1C and 1D), where the peak positions

563

shift to lower q ranges.

564

Figure 2. Under solution conditions listed in Table 1, radii (Figure 2A) and relative total

565

volumes (Figure 2B) of BaSO4 particles form on -SH and mixed -SH & COOH coated

566

substrates. For data obtained with high aqueous Ba2+/SO42- ratio (R = 10:1 and 115:1),

567

the sizes of BaSO4 particles at later stage are too big and are out of GISAXS

568

measurement range. Therefore, their size and volume cannot be calculated.

569

Figure 3. GISAXS scattering intensities generated by BaSO4 particles formed on –SH coatings

570

from solutions with varying aqueous Ba2+/SO4 ratios (R = 115:1 in panel A; R = 10:1

571

in panel B; R= 1:10 in panel C; R = 1:115 in panel D). The experimental data are

572

shown as colored dots, while the fitted curves are shown as black lines

573

Figure 4. QCM-D measurements with 0.1 mM BaCl2 solution and sensors coated with SiO2

574

(Figure A), and self-assembled thin films terminated with -COOH (Figure B), -SH

575

(Figure C), and mixed -SH & COOH (Figure D). The decreases in frequency (∆f)

576

indicate the adsorption of Ba2+ ions onto sensors.

25

ACS Paragon Plus Environment

Langmuir

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 32

577

Figure 5. ATR-FTIR measurements of bare glass and self-assembled thin films on -coated

578

substrates in water and 0.1 mM BaCl2 solution. In the frequency range from 2830-2970

579

cm-1, characteristic vibrational frequencies of symmetric and antisymmetric stretching

580

modes of υ(CH2) are presented.59 The vibrational spectrum of pure SiO2 shows the

581

presence of a residual organic compound, probably due to traces of ethanol adsorbed after

582

cleaning. The spectra of organic-coated substrates are slightly different from that of pure

583

SiO2, with maxima for the stretching modes shifted to higher frequencies than that of the

584

residual organic from SiO2. Compared with spectra measured in water, the spectra for –

585

SH, -COOH, and mixed –SH&COOH coatings measured in 0.1 mM BaCl2 solution show

586

additional peaks in the range of 2980-3020 cm-1, indicating Ba2+ adsorption onto the

587

substrates.

588 589

26

ACS Paragon Plus Environment

Page 27 of 32

Figure 1.

10 10 10

4 min 8 min 14 min 20 min 30 min 40 min 60 min

4

3

2

1

10

4

10

3

10

2

10

1

5 6

2

3 4 5 6 -1

0.01 q_xy (Å

)

10

0.1

5 6

2

0.01

3

q_xy (Å

4 5 6 -1

)

D. -SH&COOH

C. -SH

B. -COOH

5

10

10 10 10 10

0.1

5

10 Intensity (counts)

10

A. SiO2

5

Intensity (counts)

10

Intensity (counts)

590

Intensity (counts)

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

Langmuir

4

3

2

1 5 6

2

0.01

3 4 5 6 -1

q_xy (Å

591 592

27

ACS Paragon Plus Environment

)

10 10 10 10

0.1

5

4

3

2

1 5 6

2

3

4 5 6

0.01 q_xy (Å-1)

0.1

Langmuir

593

Figure 2.

B. Total particle volume

-SH R=1:1 -SH & COOH R=1:1 -SH R=115:1 -SH R=10:1 -SH R=1:10 -SH R=1:115

12 10 8 6 4 2

Total particle volume (relative unit)

A. Particle size

Radius, nm

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

Page 28 of 32

0 0 594

10

20

30

40

50

0.6 0.5 0.4 0.3 0.2 0.1 0.0

60

0

10

20

30

40

Time, min

Time, min

595

28

ACS Paragon Plus Environment

50

60

Page 29 of 32

Figure 3.

A. -SH (115:1) 10 10 10 10 10

6 5 4 3 2

B. -SH (10:1) 7

10

10

6

10

Intensity (counts)

10

4 min 8 min 14 min 20 min 30 min 40 min 60 min

Intensity (counts)

10

7

5

10

4

10

3

10

2

10

1

1

10 5 6

2

0.01

3

4 5 6 -1

q_xy (Å

)

5 6

0.1

2

0.01

3

q_xy (Å

10 10 10 10 10

4 5 6

0.1

-1

10

)

C. -SH (1:10)

7

D. -SH (1:115) 7

10

6

6

Intensity (counts)

596

Intensity (counts)

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

Langmuir

5 4 3 2 1

10

5

10

4

10

3

10

2

10

1

5 6

2

0.01

3

q_xy (Å

597 598

29

ACS Paragon Plus Environment

4 5 6 -1

)

10

0.1

5 6

2

0.01

3

4 5 6 -1

q_xy (Å

)

0.1

Langmuir

599

Figure 4.

A. Glass

B. -COOH

4 ∆ f (HZ)

∆ f (HZ)

4

0.1 mM BaCl2

2 0

-4

-4

10

20

30

Time, min

0

10 20 Time, min

C. -SH

D. –SH & COOH

∆ f (HZ)

4 2

0.1 mM BaCl2

0

-2 -4 0

0.1 mM BaCl2

0

-2

0

600

2

-2

∆ f (HZ)

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 32

10 20 Time, min

30

4 0.1 mM BaCl2 2 0 -2 -4 0 10 20 Time, min

30

ACS Paragon Plus Environment

30

30

Page 31 of 32

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

Langmuir

Figure 5.

31

ACS Paragon Plus Environment

Langmuir

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

Page 32 of 32

TOC

Heterogeneous nucleation

Hydrophobicity

Substrate coatings

BaSO4

Ba2+ and SO42- ions

32

ACS Paragon Plus Environment