A Raman-Based Imaging Method for Characterizing the Molecular

Jan 31, 2018 - A Raman-Based Imaging Method for Characterizing the Molecular Adsorption and Spatial Distribution of Silver Nanoparticles on Hydrated M...
0 downloads 8 Views 1MB Size
Subscriber access provided by Universitaetsbibliothek | Johann Christian Senckenberg

Article

A Raman-Based Imaging Method for Characterizing the Molecular Adsorption and Spatial Distribution of Silver Nanoparticles to Hydrated Mineral Surfaces Seth W Brittle, Daniel P. Foose, Kevin A O'Neil, Janice M Sikon, Jasmine K Johnson, Adam C Stahler, John David Ryan, Steven R Higgins, and Ioana Emilia Sizemore Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b04884 • Publication Date (Web): 31 Jan 2018 Downloaded from http://pubs.acs.org on February 1, 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 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.

Environmental Science & Technology 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 33

Environmental Science & Technology

2

A Raman-Based Imaging Method for Characterizing the Molecular Adsorption and Spatial Distribution of Silver Nanoparticles to Hydrated Mineral Surfaces

3 4

Seth W. Brittle, Daniel P. Foose, Kevin A. O’Neil, Janice M. Sikon, Jasmine K. Johnson, Adam C. Stahler, John Ryan, Steven R. Higgins, and Ioana E. Sizemore*

5 6

Department of Chemistry, Wright State University, 3640 Colonel Glenn Hwy., Dayton, OH, 45435

1

*Corresponding author: [email protected]

7 8

ABSTRACT

9

Although minerals are known to affect the environmental fate and transformation of heavy metal

10

ions, little is known about their interaction with the heavily-exploited silver nanoparticles (AgNPs).

11

Proposed here is a combination of hitherto under-utilized micro-Raman-based mapping and

12

chemometric methods for imaging the distribution of AgNPs to various mineral surfaces and

13

their molecular interaction mechanisms. The feasibility of the Raman-based imaging method was

14

tested on two macro- and micro-sized mineral models, muscovite (KAl2(AlSi3O10)(OH)2) and

15

corundum (α-Al2O3), under key environmental conditions (ionic strength and pH). Both AgNPs-

16

and AgNPs+ were found to covalently attach to corundum (pHpzc = 9.1) through the formation of

17

Ag-O-Al- bonds and thereby to potentially experience reduced environmental mobility. Because

18

label-free Raman imaging showed no molecular interactions between AgNPs- and muscovite

19

(pHpzc = 7.5), a label-enhanced Raman imaging approach was developed for mapping the scarce

20

spatial distribution of AgNPs- onto such mineral surfaces. Raman maps comprising of n = 625-

21

961 spectra for each sample/control were rapidly analyzed in Vespucci, a free open-source

22

software, and the results were confirmed via ICP-OES, AFM, and SEM-EDX. The proposed

23

Raman-based imaging requires minimum to no sample preparation, is sensitive, non-invasive,

24

cost-effective, and might be extended to other environmentally relevant systems.

25

ACS Paragon Plus Environment

Environmental Science & Technology

26

Keywords

27 28

Raman imaging, chemometrics, silver nanoparticles, corundum, muscovite, ICP-OES, AFM, SEM-EDX

29 30

INTRODUCTION

31

The expansion of nanotechnologies in consumer products has raised increasing concern

32

about their potential impact on human and environmental health. Special emphasis has been

33

placed on silver nanoparticles (AgNPs); 54% of the total number of consumer products

34

containing nanomaterials primarily harness the antimicrobial properties of AgNPs. About 1,230

35

tons of the total silver produced worldwide is allocated to the fabrication of AgNPs.1 The

36

increased use of AgNPs may result in a high Ag content for the biosolids produced by

37

wastewater treatment plants (WWTP). For example, a 2010 study reported that AgNPs released

38

from antibacterial athletic socks was significant2 but removable by WWTPs2. The resulting

39

biosolids may then be sold as fertilizer, thus providing a pathway for these engineered

40

nanomaterials to be introduced into soils through irrigation and rainfall. Numerous studies have

41

already examined the health effects of AgNPs on both aquatic and terrestrial organisms including

42

humans, and primarily attributed their toxicity to the release of Ag+ ions.3-5 However, the impact

43

AgNPs have on environmental health remains under investigation in view of their possible

44

transformations during the interaction with diverse soil components as well as the lack of cost-

45

and time-efficient methodologies for examining these aspects.5-12

46

Current research suggests that released AgNPs are most likely to be immobilized in soils

47

due to the adsorption to natural organic matter (NOM) and minerals.6–8 Most of these studies

48

focused on the adsorption of AgNPs to NOM because of their high chemical affinity to the

49

sulfhydryl-rich functional groups present in NOM.7–9 Because NOM constitutes only ~ 5% of

50

soil,10 it is important to also examine the interaction of AgNPs with the main soil component,

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

Environmental Science & Technology

51

minerals (~ 45%). Although there is extensive research available on the interactions between

52

metal ions and soil minerals, only a few studies were reported on nanometals.11–13 The research

53

that does exist on AgNPs is generally focused on simulations of ideal interaction scenarios or

54

employs advanced methodologies (e.g., X-ray absorption spectroscopy and electron microscopy)

55

that usually requires extensive sample preparation and expensive resources.8,14-25 Proposed here

56

is a combination of hitherto under-utilized micro-Raman-based mapping and chemometric

57

methods15–17 for the simple and cost-effective imaging of the spatial distribution of AgNPs to

58

various mineral surfaces and their molecular interaction mechanisms.

59

It was suggested that AgNPs released into the environment through wastewater,

60

fertilizers or landfill runoff might exist in multiple forms.8 AgNPs may persist in their nanoscale

61

form, oxidize to aqueous Ag+, aggregate or precipitate into salts (Ag2S, AgCl, Ag2SO4, AgOH,

62

Ag2CO3, and Ag2O).3,8,18 Thus, a methodology capable of observing AgNP distribution on

63

mineral surfaces and their molecular interaction mechanisms is essential to the determination of

64

their fate and transport. Raman and surface-enhanced Raman spectroscopy (SERS) offer a

65

unique approach because of their molecular fingerprinting and multiplex detection capabilities,

66

non-destructive and aqueous compatible nature, and little to no requirement for sample

67

preparation.19–30 Furthermore, modern Raman systems can collect multiple point spectra in x,y-

68

raster patterns (Raman maps) to micro- or nano-map molecular interactions across large surfaces.

69

The herein proposed Raman-based imaging approach is novel in that it offers the possibility of

70

performing either label-free or label-enhanced SERS measurements. In general, minerals (e.g.,

71

quartz, calcite, and corundum) have been decorated with noble metal nanoparticles (e.g., silver

72

and gold) for the development of new SERS platforms and in this context, they have been

73

examined through the acquisition of Raman control spectra.31–35 However, no approaches were

ACS Paragon Plus Environment

Environmental Science & Technology

74

yet proposed for the rigorous mapping of the molecular adsorption and spatial distribution

75

behavior of AgNPs onto environmentally relevant mineral surfaces. Here it is demonstrated that

76

the proposed Raman-based imaging methodology can achieve these goals for different types of

77

minerals. If no direct molecular interaction occurs between minerals and AgNPs, a Raman active

78

label of large scattering cross-section (e.g., rhodamine 6G (R6G) dye) and high affinity toward

79

AgNPs may be utilized to make AgNPs “Raman-visible” and to indirectly image their spatial

80

distribution onto mineral surfaces with increased sensitivity. Our group has already reported

81

single-molecule SERS detection events of R6G (1 fM) adsorbed onto Creighton AgNPs under

82

the excitation of visible laser lines and with less than 20 s acquisition times.23 When a target

83

species is located in the immediate vicinity of an individual AgNP or at the nano-sized interstitial

84

site of aggregated AgNPs, the SERS effect occurs and further boosts the sensitivity of the

85

Raman-based detection method to unprecedented levels.19–22 The SERS enhancement is largely

86

due to the increase in the magnitude of both the incident and the scattered electromagnetic fields

87

resulting from the excitation of localized surface plasmon resonances (LSPR) of AgNPs.21,23–26

88

Thus, the label-free and label-enhanced Raman-mapping of mineral surfaces exposed to AgNPs

89

results in the collection of a large number of SERS spectra characteristic to the direct or indirect

90

interaction of the two systems, respectively. An additional element of novelty in this study is the

91

coupling of Raman mapping with chemometric methods for the facile hyperspectral analysis of

92

large volumes of Raman data in a reproducible manner. This was achieved using a free, open-

93

source, stand-alone analysis software package, Vespucci.27

94

The feasibility of the Raman-mapping chemometric method was demonstrated on two

95

representative minerals in micro and macro form and unfunctionalized (Creighton) AgNPs- and

96

cetyltrimethylammonium bromide (CTAB)-functionalized (Sun method) AgNPs+. Corundum (α-

ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Environmental Science & Technology

97

Al2O3) and muscovite (KAl2(AlSi3O10)(OH)2) were selected because they contain several of the

98

most abundant elements in the earth’s crust (O, Al, and Si).10 In addition, both minerals are rock-

99

forming and encountered in streams and sands.10 While the Creighton synthesis is one of the

100

most widely-used bottom-up fabrication approaches of colloidal AgNPs of negative charge due

101

to its simplicity, time, and cost efficiency,14,29 the Sun method is generally employed for the

102

synthesis of AgNPs of positive charge.36, 37 A concentration of 1 mg L-1 of AgNPs was used in

103

order to ensure sub-monolayer coverage at the mineral surface and to surpass the maximum

104

contaminant level (MCL) set by the U.S. Environmental Agency (EPA) for Ag+ in drinking

105

water from both natural and anthropogenic sources (0.1 mg L-1).38 The interaction mechanism of

106

corundum and muscovite in macro form was investigated at the pH of the colloid (8.2 for

107

Creighton and 6.9 for Sun) because both minerals have a pH at point of zero charge (pHpzc of 9.1

108

for corundum39 and 7.5 for muscovite40) close to this pH value. For illustrative purposes, the

109

interaction mechanism between corundum in micro form and AgNPs was also investigated in the

110

pH range from 6 to 11, typical to the pH of soils (3.5–9)10,41. Aggregation and settling of AgNPs

111

in the corundum mixtures were noticed at smaller (< 6) or larger (> 11) pH values. In addition,

112

several other well-established techniques, inductively coupled plasma optical emission

113

spectroscopy (ICP-OES), atomic force microscopy (AFM), and scanning electron microscopy

114

coupled with energy dispersive X-Ray (SEM-EDX) were employed to confirm and complement

115

the Raman results.

116

MATERIALS AND METHODS

117

Single-crystal muscovite and corundum were purchased from Ward’s Earth Science and

118

Marketech International, respectively. Macro-muscovite (i.e., single-crystal) was prepared by

119

freshly cleaving 10 mm × 10 mm squares. Macro-corundum (i.e., single-crystal) consisted of

ACS Paragon Plus Environment

Environmental Science & Technology

120

cylindrical, single α-phase crystals cut and polished on the ( 1120 ) plane, 25 mm wide and 5 mm

121

thick. Micro-corundum was purchased as fused, 1-µm-sized particles (99% fused α-Al2O3)

122

from Alfa Aesar. High-quality (HQ) water (18.2 MΩ ·cm) was produced in a LabConco system

123

and utilized throughout the course of all experiments unless otherwise specified. All other

124

materials were purchased from Fisher Scientific and used without further modification unless

125

specified.

126

Synthesis and Characterization of AgNPs: Negatively charged, colloidal AgNPs were

127

fabricated using a slightly modified Creighton method in water.19,29,30 Specifically, a 2:1 mM

128

ratio of sodium borohydride (300 mL of 2 x 10-3 M of NaBH4) to silver nitrate (50 mL of 1 x 10-3

129

M of AgNO3) solutions was used to minimize the amount of excess reagents and byproducts.

130

Positively charged, colloidal AgNPs were synthesized using a slightly modified Sun method in

131

water through the reduction of AgNO3 (300 mL of 5 x 10-3 M) with CTAB (20 mL of 2 x 10-2

132

M) and NaBH4 (0.2 mL of 1% (w/v)).36,50 The physicochemical properties of AgNPs were then

133

characterized by Raman spectroscopy, ICP-OES, AFM, SEM-EDX, and UV-Vis absorption

134

spectroscopy.

135

Sample Preparation: The flat macro-corundum surfaces were sequentially washed with

136

acetone (HPLC grade), methanol (HPLC grade), and nitric acid (70% OPTIMA grade) in a sonic

137

bath for 10 min each. Next, they were annealed in air at 1250°C for 12 h to provide a “clean”

138

terraced surface. Both macro-corundum and macro-muscovite samples were then submerged in

139

the colloid (100 mL of 1 mg L-1 of AgNPs-) mixed with 100 µL of 5 M of sodium nitrate

140

(NaNO3), as an ionic strength adjuster in order to simulate the ion strength of freshwater

141

environmental samples. Macro-muscovite samples were removed after 30 min and dried with a

142

stream of nitrogen gas. Macro-muscovite received an additional exposure to rhodamine 6G

ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33

Environmental Science & Technology

143

(R6G), a standard SERS probe, in order to make adsorbed-AgNPs- Raman "active". Briefly, 10

144

µL of 10-4 M of R6G solution was dispensed onto the mineral covering a surface area of ~1 cm2.

145

Next, the surface was washed three times with water after 30 min incubation. Control groups

146

with ISA included mineral exposure to water, CTAB (for AgNPs+ only), AgNPs (positively or

147

negatively charged), or 10-3 M of R6G (for macro-muscovite only).

148

Approximately 1.2 g micro-corundum was incubated with 100 mL of 1 mg L-1 of AgNPs-

149

, 100 µL of 5 M of NaNO3, and pH adjusters (5-100 µL of 0.1 M of HNO3 or 0.1 M of sodium

150

hydroxide (NaOH)). This amount of micro-corundum yields submonolayer coverage when

151

assuming total AgNP- adsorption and a specific surface area (SSA) of 6.5 ± 0.2 m2 g-1. Similar

152

experiments were carried out with CTAB-functionalized AgNPs+ for illustrative purposes. The

153

pH was measured with a SevenGo Duo pro model pH meter that was calibrated daily with four

154

pH buffers (4.0, 7.0 10.0, and 12.0). After stirring for 30 min, the samples were centrifuged for 2

155

min at 5000 G in an AccuSpin Micro 17/17R model centrifuge. The centrifuge supernatants

156

containing free, unbound AgNPs- were saved for ICP-OES analysis, while the centrifuge pellets

157

consisting of corundum particles with bound AgNPs- were collected for Raman imaging. Control

158

groups with ISA included mineral exposed to AgNPs- and hydrated mineral only at each

159

experimental pH value (6, 7, 8, 9, 10, and 11).

160

Specific Surface Area (SSA) Analysis: BET42 surface area analysis was performed on

161

the corundum powder using a Micromeritics Tristar II 3020 surface area and porosity analyzer.

162

A 5-point N2 BET isotherm with relative pressures in the range P/Po = 0–0.25 was used for the

163

analysis, where Po is the saturation pressure. Prior to the analysis of the sample material, a

164

standard reference material specified as alumina powder was measured to validate the method

165

employed. The standard alumina was determined to have a specific surface area of 0.28 ± 0.02

ACS Paragon Plus Environment

Environmental Science & Technology

166

m2 g-1, which compared well with the specified surface area of 0.29 ± 0.02 m2 g-1. The Alfa-

167

Aesar corundum was found to have a specific surface area of 6.5 ± 0.2 m2 g-1, which is within the

168

range of 6-8 m2 g-1 specified by the supplier.

169

Raman Spectroscopy Analysis: Raman data was collected using a LabRam HR800

170

Raman system coupled to an Olympus BX41 confocal microscope (100x objective) and a

171

motorized stage. Samples were irradiated with HeNe (632.8 nm) and Nd:YAG (532.134 nm)

172

lasers with powers at the sample of 15 and 19 mW, respectively. The 532 nm excitation line is

173

close to the R6G resonance at 530 nm (resonant conditions). The backscattered photons were

174

measured using a thermoelectrically-cooled Andor CCD camera with a resolution of 1024×256

175

pixels. The following parameters were selected: 300 µm confocal hole, 600 grooves mm-1

176

holographic grating, 3 s (muscovite) and 1 s (macro-corundum) acquisition times, and 2 or 3

177

averaging cycles. Under these conditions, the spectral resolution, i.e., the distance between points

178

on the spectral abscissa, was ~1.18 cm-1. For each macro-muscovite sample, n = 900 spectra

179

were measured in a 150×150 µm grid, at 5 µm increments. For each macro-corundum sample, n

180

= 961 spectra were collected in a 31×31 µm grid, at 1 µm increments. The Raman system proved

181

successful in the collection of data with nano-spatial resolution but at higher time costs. All

182

spectra were collected in the 100–1700 cm-1 spectral range. Samples from the micro-corundum

183

experiment were smeared onto clean glass slides and 11×11 µm areas were mapped in 1 µm

184

increments for each sample to obtain a representative molecular picture of the interaction

185

between AgNPs- and corundum particles. Because increased dispersity was anticipated with

186

these non-flat samples, statistical confidence was strengthened by measuring n = 3 maps from n

187

= 3 individually prepared samples, totaling 9 maps for each pH (i.e., n = 1089 spectra for each

188

pH). For illustrative purposes, point Raman spectra were also collected from random areas on n

ACS Paragon Plus Environment

Page 8 of 33

Page 9 of 33

Environmental Science & Technology

189

= 3 independently prepared micro-corundum samples exposed to AgNPs+ in the presence of ISA,

190

at different pH values (6-11).

191

Raman Data Analysis: Raman maps were processed in Vespucci27. All spectra were

192

smoothed with a median filter (window size 7). Muscovite spectra were baseline corrected using

193

a rolling-ball approach43 in order to remove the fluorescent background and then normalized to

194

the 1-norm. Univariate analysis to determine peak area and intensity against a local linear

195

baseline was performed on the following regions of interest: the Ag-O stretch (235 cm-1) arising

196

from interaction between AgNPs and the corundum surface in corundum samples and several

197

R6G bands19,23 in the muscovite samples. Macro-corundum spectra were min/max normalized by

198

subtracting the minimum value of the spectrum from each value of the spectrum, then

199

subsequently dividing by the maximum value of the spectrum, so that the smallest value of each

200

spectrum was 0, and the largest value of each spectrum was 1. The micro-corundum data were

201

further analyzed using the Kruskal-Wallis test, with Dunn's test performed post-hoc to determine

202

the significance of differences in Raman signal between pH groups.

203

AFM analysis: Topographic images were created with an Agilent AFM operated in

204

intermittent contact mode (i.e., AC or Tapping Mode), in air. Cantilevers were obtained from

205

Nanoworld (NCHR, Pointprobe, non-contact mode) and were fabricated from single-crystal Si

206

and coated with Al, with a nominal resonance frequency of 320 kHz and a nominal force

207

constant of 42 N/m. Scanning speeds were typically set to 1–2 Hz and 2×2 µm images of

208

256×256 pixels were recorded to adequately display features between 1–100 nm (i.e., AgNPs-).

209

AgNP- adsorption was then interpreted by differentiating topographic profiles inbetween controls

210

and samples.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 33

211

ICP-OES Analysis: Quantitative characterization of the total Ag content reacted with the

212

micro-sized corundum was determined by difference using a Varian 710 ICP-OES system.

213

Briefly, original colloids and supernatant samples containing free, unbound AgNPs- were

214

chemically digested and diluted in trace metal grade nitric acid (HNO3) following the U.S. EPA

215

method 200.11.44,45 An eleven-point external calibration curve (0, 5, 10, 15, 20, 25, 50, 75, 100,

216

125, and 150 µg L-1) was then constructed, and the Ag concentrations were determined by

217

interpolation from the calibration curve.

218

SEM-EDX Analysis: Prior to the data collection with an FEI Quanta FEG 250 SEM

219

system, all samples were sputter coated with 2-nm of Ir to increase their stability during exposure

220

to relatively high beam voltages. Each sample was then loaded onto an aluminum stub using Cu

221

tape and exposed to a 6-kV electron beam at a working distance of 10-mm. An EDAX EDS

222

detector coupled to the microscope was utilized in the collection of spot size EDX spectra. At

223

least 20,000 counts were obtained from the dominant element present within each spectrum. A

224

holographic peak deconvolution was then employed to qualify spectra by ruling out noise and

225

validating surface signal.

226

RESULTS AND DISCUSSION Synthesis and Characterization of AgNPs: n = 3 batches of AgNPs were synthesized

227 228

for

each

experiment

and

were

characterized

in

229

recommendations.46,47 Both AgNPs- and AgNPs+ exhibited a surface plasmon resonance peak at

230

~400 nm in good agreement with the literature.50 ICP-OES revealed an average Ag colloidal

231

concentration of 15.4 ± 0.8 and 300.6 ±0.6 mg L-1 for AgNPs- and AgNPs+, respectively. Raman

232

spectroscopy confirmed a) the absence of organic impurities or silver oxide peaks in the colloids

233

and b) the presence of CTAB at the AgNP+ surface (Figure S1). Electron microscopy data

ACS Paragon Plus Environment

accordance

with

U.S.

EPA’s

Page 11 of 33

Environmental Science & Technology

234

showed that AgNPs- are spherical, have an average diameter of 14.1 ± 13.4 nm and a moderate

235

breadth size distribution in the 1-100 nm range.23 The specific AgNP- surface area was then

236

estimated to be 20.9 - 814.8 m2 g-1 based on this size range and assumed sphericalness.48 AgNPs+

237

were also found to be spherical but had a narrower size distribution (1-40 nm) than AgNPs-.50

238

Zeta-potential measurements demonstrated that both AgNPs are charged and stable at the

239

experimental pH of this study (ζ-potential of -41.47 mV at pH = 8.2 for AgNPs- and +34 mV at

240

pH = 7 for AgNPs+).14,48 These colloidal AgNPs contain less than 10% Ag+ ions.49,50

241

Macro-Muscovite: The label-free Raman spectra of macro-muscovite exposed to

242

AgNPs- indicated no direct covalent interaction between the two (Figure S2). Thus, label-

243

enhanced Raman spectra were acquired on these samples using R6G as a SERS label of large

244

scattering cross-section in order to indirectly image the AgNP- distribution on the macro-

245

muscovite surfaces. It is known that R6G cations have a high affinity towards the negatively-

246

charged AgNP surface, where they experience large Raman signal enhancements associated with

247

the SERS effect19. As a result, R6G fluorescence is quenched and detailed SERS spectra may be

248

collected from R6G molecules located in the immediate vicinity of individual AgNPs- or AgNP--

249

aggregates interacting with mineral surfaces.

250

Raman controls (n = 900 spectra) consisting of hydrated macro-muscovite surfaces (no

251

AgNPs- or R6G) exhibited solely the characteristic vibrational modes of muscovite (e.g., 170,

252

170, 195, 215, 267, 382, 413, 638, 707, 757, 915, 959, and 1120 cm-1), which were identified in

253

good agreement with the literature51-54 (Table S1). Macro-muscovite controls with just AgNPs-

254

yielded no detectable additional peaks and appeared vibrationally similar to muscovite control

255

(Figure S2). The macro-muscovite control sample exposed to 10-3 M of R6G (no AgNPs-)

256

displayed a fluorescent background with no characteristic Raman R6G peaks. In order to

ACS Paragon Plus Environment

Environmental Science & Technology

257

mitigate R6G fluorescence in the label-enhanced SERS spectra, R6G concentration was lowered

258

to 10-4 M (R6G footprint sufficient for AgNP- coverage). The R6G-AgNP--mineral samples

259

revealed significant differences with respect to controls: strong Raman spectral signatures of

260

R6G and muscovite were simultaneously detected (Figures 1, S2). The loadings of all PCs that

261

accounted for > 1% of variance exhibited both R6G and muscovite Raman signatures. The

262

presence of strong R6G peaks (e.g., 611, 776, 1310, 1361, 1508, and 1652 cm-1)19,23,55,56 indicate

263

enhancement had occurred in the areas where R6G-labeled AgNPs- have interacted with

264

muscovite. Because no other Raman bands (e.g., Ag-O stretching mode) were detected in the n =

265

3600 analyzed spectra, the interaction between AgNPs- and muscovite was deemed to be of a

266

non-covalent nature.

267 268

Figure 1. A) Raman images of four R6G-AgNP--macro-muscovite samples (25 x 25 µm each)

269

constructed from the first PC and B) the loading for the first PC corresponding to R6G and

ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

Environmental Science & Technology

270

muscovite signals. The sample was prepared by immersing muscovite into 1 mg L-1 of AgNPs-

271

for 30 min. Colors are mapped to the score for the first PC (accounting for 46% of variance),

272

with purple regions representing spectra with negative scores (R6G absorbed onto AgNPs-) and

273

brown regions representing spectra with positive scores (muscovite only). A larger score in the

274

first PC is associated with a stronger muscovite Raman signature relative to the R6G signature,

275

as seen in the plot of the first PC loading.

276 277

Macro-Corundum—The interaction mechanism and spatial distribution of AgNPs- on the

278

mineral surface were investigated via label-free Raman imaging (Figures 2 and S3). The

279

corundum controls exhibited the typical corundum peaks from literature (e.g., 379, 417, 429,

280

451, 576, 644, and 750 cm-1, Figure S3, Table S2)57-59, while corundum exposed to AgNPs-

281

revealed additional molecular vibrations in the 225-255 cm-1 spectral range, matching the

282

literature for Ag-O stretching modes.20,23,60 Therefore, a covalent attachment of AgNPs- to

283

corundum was inferred. An illustrative Raman image is given in Figure 2, where the gray-black

284

areas correspond to the presence of corundum only (quantified through the integrated Raman

285

area). The spectrum containing the most intense Ag-O peak from n = 961 analyzed spectra is

286

represented by the white pixel in the chemical image. The high sensitivity of the Raman-based

287

imaging methodology was demonstrated through the detection of scarce molecular interaction

288

events of AgNPs- on the macro-mineral surface. Further AFM and SEM-EDX analyses were

289

then carried out to complement the Raman results and to further interrogate the interaction

290

mechanism.

ACS Paragon Plus Environment

Environmental Science & Technology

291 292

Figure 2. A). Raman image of a AgNP--macro-corundum window (52 µm x 52 µm) prepared by

293

immersing corundum into 1 mg L-1 of AgNPs- for 30 min. The white pixel (4 µm2) corresponds

294

to the covalent attachment of AgNP(s)- to the mineral surface, while the gray-black pixels are

295

indicative of corundum alone. B) The corresponding Raman spectra illustrate the appearance of a

296

Ag-O stretching mode at 244 cm-1 and confirm the presence of corundum. The spectrum with the

297

minimum Euclidean distance to the mean of all spectra in the image is shown.

298

299

Topographic AFM profiles of corundum windows revealed a step and terrace structure on

300

the ( 1120 ) plane surface (Figure 3). Control hydrated mineral surfaces originally appeared

301

devoid of any nanoscale features but exhibited protrusions and holes characteristic to macro-

302

minerals (Figure 3A). The addition of ISA (NaNO3) to the control mineral surfaces resulted into

303

the formation of a few nanoscale features (n = ~2.5 per 1 µm2, 1-30 nm, Figure 3B). This is

304

probably due to the precipitation of the electrolyte from the thin water film that was not

ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

Environmental Science & Technology

305

completely removed with the stream of nitrogen gas during drying. The EDX elemental analysis

306

(Figure S3C) confirmed the chemical nature of the ISA-derived nanostructures (Na peak at 1.05

307

keV); they appeared as white features in the collected SEM images (Figure S3A). The corundum

308

windows were then submerged in 1 mg L-1 of colloidal AgNPs- for 30 min (with 0.005 M of

309

NaNO3 at pH 8) and dried again with nitrogen gas. The resultant AFM images of the exposed

310

windows showed AgNPs- of a larger size distribution (n > 10 per 1 µm2, 1-100 nm) on the

311

mineral surfaces without any real indication of avoidance of or preference to adsorption at the

312

step edges (white features, Figure 3C). SEM-EDX measurements (figure S4) further established

313

the presence of AgNPs- (Ag peak at 2.99 keV) to the mineral surface (O and Al peaks at 0.53

314

keV and 1.5 keV, respectively).

315

316

317

318

319

320

321

322

323

ACS Paragon Plus Environment

Environmental Science & Technology

324

Figure 3. 2 µm x 2 µm AFM topographic images of macro-corundum windows. A) Mineral

325

control: a water-washed and annealed mineral window displaying step-and-terrace structure with

326

minimal nano-sized features. B) Mineral control with ISA (no AgNPs-): the same window

327

submerged in 0.005 M of NaNO3 for 30 min leading to the formation of few nano-sized features

328

(white spots) due to the NaNO3 precipitation during the evaporation of the solution film on the

329

surface. C) Mineral sample with AgNPs-: the same window submerged in 1 mg L-1 of AgNPs-

330

generating additional, larger nano-sized features (white spots).

331

332

The macro-corundum experiment was conducted at pH = 8, but a wide range of pH

333

values are possible in natural aquatic systems. Because mineral surface charge and particle

ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33

Environmental Science & Technology

334

oxidation depend on pH, the next set of experiments were conducted on micro-corundum in the

335

absence and presence of AgNPs- and AgNPs+, in the 6-11 pH range. Furthermore, multiple

336

crystalline faces exposed on mineral grains are more common in natural systems. Thus, the

337

incorporation of a micro-mineral model can offer additional insight into the mineral-AgNP

338

interaction behavior where additional crystallographic planes other than the ( 1120 ) surface are

339

exposed.

340

Micro-corundum particles: As expected, no Ag-O peaks were observed in the micro-

341

corundum Raman controls similarly to macro-corundum (Figure 4A), when no AgNPs were

342

present. Upon exposure to AgNPs-, a Ag-O stretching band appeared at ~ 230 cm-1 at each pH

343

(Figure 4B), confirming the covalent interaction between AgNPs- and corundum (for both

344

micro- and macro-corundum). However, n > 0.73 more Ag-O interactions per µm2 were

345

detected in the micro- than macro-corundum samples probably due to the different number of

346

lattice planes exposed at the surface; AgNP- concentration and mineral surface area were

347

factored into these calculations. While macro-corundum has only a single lattice plane exposed

348

at the surface (an ordered lattice in the a-plane (112ത0)61), micro-corundum has many different

349

lattice planes that could potentially favor the Ag-O interactions.

350

To further analyze the interaction mechanism, the total number of Raman spectra

351

exhibiting the Ag-O stretching mode was plotted as a function of pH (Figure 4C). Although the

352

largest number of AgNP--corundum interactions was detected at pH ≥ 9 (n = 902-1015 spectra)

353

when compared to all other examined pH values (n = 796-862), the difference was not

354

statistically significant (p > 0.05). Thus, the adsorption process of AgNPs- was deemed not pH

355

dependent under the employed experimental conditions, i.e., 1.2 g of micro-corundum particles

356

(SSA ~ 6.5 ± 0.2 m2 g-1) exposed to 100 mL of 1 mg L-1 of negatively charged AgNPs- for 30

ACS Paragon Plus Environment

Environmental Science & Technology

357

min at each pH. The integrated area under each Ag-O peak was also estimated to further

358

confirm this pH independent behavior (Figure 4D). Again, the differences amongst all pH

359

values were not statistically significant (p > 0.05), despite pH 8 and 9 yielding the largest

360

average integrated area. Similar Ag-O interactions were detected between AgNPs+ and micro-

361

corundum at each pH (Figure S1), but no statistically significant trend could again be

362

established with the change in pH. It should be noted that possible trends in the integrated area

363

of the corresponding Ag-O stretching peak (235-250 cm-1) might have been skewed by the

364

presence of a small CTAB vibrational mode at ~238 cm-1 (Figure S1). Future work that

365

establishes the relative contributions of CTAB to the Ag-O band profile , such as a pH

366

dependence study on CTAB alone or the use of different AgNPs+ colloidal systems, would lead

367

to a better quantitative description of the Ag-O interactions.

ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33

Environmental Science & Technology

368 369

Figure 4. Raman data and their statistical analysis for AgNP--micro-corundum at all

370

experimental pH values (6 – 11). Average Raman spectra of A) corundum control with ISA and

371

B) corundum sample with ISA exposed to 1 mg L-1 of AgNPs- for 30 min at each pH. C) The

372

total number of Raman spectra of the AgNP--micro-corundum samples exhibiting Ag-O

373

stretching modes at each pH value. D) Mean of the integrated area of the Ag-O stretching mode

374

at each examined pH value with 95% confidence intervals. The values obtained at each pH

375

represent an average of n = 1089 spectra from n = 3 separate trials.

376 377

Because the Raman experiments conducted on both macro- and micro-corundum

378

revealed the appearance of Ag-O stretching modes, the surface complexation of AgNPs- to the

379

terminus corundum oxygen atoms (Ag-O-Al-) was considered to play a key role in their

ACS Paragon Plus Environment

Environmental Science & Technology

380

interaction. Furthermore, the highest chemisorption level was detected around a pH value of 9,

381

where corundum has approximately zero net surface charge and other interaction mechanisms

382

may be at play. This is illustrated in the proposed mechanistic scheme (Figure 5). In addition to

383

the favored Ag-O covalent interactions at pH ~ 9, AgNPs- may also experience electrostatic

384

repulsions and attractions to the negative and positive moieties, respectively, of the mineral

385

surface. The formation of hydrogen bonding may also occur at all pH values due to the

386

hydration of corundum. In fact, literature shows that the hydration of corundum is accompanied

387

by the formation of multiple hydroxyl terminations onto its surface,59 which can facilitate this

388

interaction. Some hydroxyl groups will undergo deprotonation above the pHpzc (pH = 9.1) of

389

corundum to create an overall negative charge, whereas positively-charged surface sites will be

390

in the majority due to surface site protonation below the pHpzc. Thus, electrostatic repulsions

391

and attractions are expected at pH values > 9.1 and < 9.1, respectively.

392 393

Figure 5. Schematic of the proposed interaction mechanisms between AgNPs- and micro-

394

corundum at the investigated pH values (6-11). Red atoms = Al, blue = O, gray = H, and green =

395

Ag. Drawing is for illustrative purposes and is not to scale.

ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

Environmental Science & Technology

396 397

To support the Raman analysis, the total Ag content on the mineral surface was measured

398

by ICP-OES. It was estimated by the difference between the Ag content of the initial AgNP-

399

colloid and that of the supernatant containing free AgNPs-, which were isolated from AgNPs-

400

bound to micro-corundum particles through centrifugation. This was interpreted as an average

401

percent of AgNPs- adsorbed onto micro-corundum. Quantitative measurements of the total

402

amount of Ag was also made for all controls at each pH; no Ag was detected in the absence of

403

AgNPs-. In good agreement with the Raman results, no significant dependency on pH was

404

observed for AgNP- adsorption to micro-corundum (Figure 6, p > 0.05). Again, this is likely due

405

to the multiple interaction mechanisms possibly occurring at each experimental pH. For example,

406

at pH > 9.1, corundum particles have an overall negative charge and are thereby expected to

407

experience electrostatic repulsions to the negatively charged AgNPs-. Nevertheless, over 80% of

408

the total amount of available AgNPs- was found to be adsorbed on the mineral surface at these

409

pH values (ICP-OES data). Thus, the chemisorption of AgNPs- to corundum through the

410

formation of Ag-O covalent bonds (Raman data) may offset to a certain degree the contributions

411

of the electrostatic repulsions to the overall interaction scheme. Future work that establishes the

412

relative amounts of AgNP- sorbed via electrostatic and chemisorption interactions, such as via a

413

study of the effects of ionic strength on sorption, would lead to a better quantitative description

414

of the speciation of sorbed AgNPs-. On the other hand, at pH < 9.1, corundum particles have an

415

overall positive charge and are electrostatically attracted to the negatively charged AgNPs-.

416

These electrostatic contributions should add up to the chemisorption ones leading to statistically

417

higher levels of AgNP- adsorption to corundum, which was not observed here (figure 6). Thus,

ACS Paragon Plus Environment

Environmental Science & Technology

418

additional interaction mechanisms may be at play here and should be the subject of future

419

investigations.

420 421

Figure 6. Average percent adsorption of AgNPs- to micro-corundum as a function of pH. The

422

Ag content was by determined by ICP-OES after 30 min incubation of 1.2 g of corundum

423

particles (SSA of 6.5 ± 0.2 m2 g-1) with 100 mL of 1 mg L-1 of AgNPs- followed by the

424

separation of free AgNPs- from corundum-bound AgNPs- through centrifugation. Error bars

425

represent the standard deviation of n = 9 independently prepared samples.

426 427

Overall, this study addresses a fundamental methodology gap by demonstrating the

428

capability of the Raman-based imaging and chemometrics approaches in characterizing the

429

molecular adsorption behavior and distribution of AgNPs to hydrated mineral surfaces with

430

nano- or micro-spatial resolution (available in modern Raman systems). This Raman method

431

requires minimum to no sample preparation, is sensitive and non-invasive. Furthermore, label-

432

enhanced Raman imaging could substitute for label-free Raman imaging to indirectly map the

433

distribution of AgNPs onto mineral surfaces, when no direct molecular interaction occurs

434

between the two. The proposed Raman method could be extended to a large variety of

435

environmentally relevant minerals (e.g., quartz, kaolinite, orthoclase, montmorillonite, hematite,

436

and calcite) and nanomaterials (e.g., Au, Cu, C, Zn, CdSe, TiO2, and so on). For example,

ACS Paragon Plus Environment

Page 22 of 33

Page 23 of 33

Environmental Science & Technology

437

Raman spectroscopy was successfully utilized in the identification of nano-TiO2 phases, crystal

438

size, quantum confinement, annealing effects, and so on. Furthermore, most of the mentioned

439

nanomaterials were found to enhance the Raman signal of analytes physisorbed or chemisorbed

440

on their surface through the SERS effect. The effects of key solution variables (e.g., pH and ionic

441

strength) and nanomaterial characteristics (e.g., size, concentration, surface charge, and surface

442

coating) could also be explored. In addition, both cationic and anionic Raman labels could be

443

employed to further investigate the interactions with mineral surfaces of permanent negative or

444

positive charge. The knowledge gained from such Raman-based imaging projects could raise

445

awareness with respect to the safe manufacture, processing and proper disposal of engineered

446

nanomaterials into the environment.

447 448

SUPPORTING INFORMATION

449

The Supporting Information is available free of charge on the ACS Publications website.

450

Raman spectra of micro-corundum exposed to CTAB-capped AgNPs+ with ISA and the

451

corresponding controls at different pH values, representative Raman spectra of muscovite

452

controls, a Raman image of a AgNP--macro-corundum control window, assignments of

453

the main Raman vibrational modes observed for the muscovite and corundum controls,

454

backscattered SEM images and EDX spectra of micro-corondum exposed to AgNPs- with

455

ISA and the corresponding controls.

456 457

ACKNOWLEDGEMENTS

458

The Chemistry Department at WSU and the U.S. National Science Foundation Award #1438340

459

are gratefully acknowledged for the financial support of this project. Garrett VanNess and Joseph

ACS Paragon Plus Environment

Environmental Science & Technology

460

Solch are thanked for their assistance in the maintenance and operation of the ICP-OES system

461

at WSU. Dr. Dhriti Nepal is acknowledged for providing the access to the SEM-EDX system at

462

Wright Patterson Air Force Base, Dayton, Ohio. The authors are grateful to Cody Fourman and

463

Garrett VanNess for conducting BET surface area analysis on the mineral samples. The authors

464

also acknowledge support from the Wright State University Research Challenge program for

465

financial support for the surface area analysis instrumentation.

466

467

REFERENCES

468

(1)

environment. Environ. Sci. Technol. 2008, 42 (12), 4447–4453; DOI 10.1021/ES7029637.

469

470

Mueller, N. C.; Nowack, B. Exposure modeling of engineered nanoparticles in the

(2)

Benn, T.; Cavanagh, B.; Hristovski, K.; Posner, J. D.; Westerhoff, P. The release of

471

nanosilver from consumer products used in the home. J. Environ. Qual. 2010, 39 (6),

472

1875–1882.

473

(3)

Luoma, S. N. Project on Emerging Nanotechnologies; PEN 15 September, 2008.

474

(4)

Fabrega, J.; Fawcett, S. R.; Renshaw, J. C.; Lead, J. R. Silver nanoparticle impact on

475

bacterial growth: Effect of pH, concentration, and organic matter. Environ. Sci. Technol.

476

2009, 43 (19), 7285–7290; DOI 10.1021/es803259g.

477

(5)

Lok, C.-N.; Ho, C.-M.; Chen, R.; He, Q.-Y.; Yu, W.-Y.; Sun, H.; Tam, P. K.-H.; Chiu, J.-

478

F.; Che, C.-M. Silver nanoparticles: partial oxidation and antibacterial activities. J. Biol.

479

Inorg. Chem. 2007, 12 (4), 527–534.; DOI 10.1007/s00775-007-0208-z.

480 481

(6)

Akaighe, N.; MacCuspie, R. I.; Navarro, D. A.; Aga, D. S.; Banerjee, S.; Sohn, M.; Sharma, V. K. Humic acid-induced silver nanoparticle formation under environmentally

ACS Paragon Plus Environment

Page 24 of 33

Page 25 of 33

Environmental Science & Technology

482

relevant conditions. Environ. Sci. Technol. 2011, 45 (9), 3895–3901; DOI

483

10.1021/es103946g.

484

(7)

Kanel, S. R.; Flory, J.; Meyerhoefer, A.; Fraley, D. J. L.; Roose, D.; Sizemore, I. E.;

485

Goltz, M. N. Influence of natural organic matter on fate and transport of silver

486

nanoparticles in saturated porous media: laboratory experiments and modeling. J.

487

Nanoparticle Res. 2015, 17 (3), 1-13; DOI 10.1007/s11051-015-2956-y.

488

(8)

Batley, G. E.; Kirby, J. K.; McLaughlin, M. J. Fate and risks of nanomaterials in aquatic

489

and terrestrial environments. Acc. Chem. Res. 2013, 46 (3), 854–862; DOI

490

10.1021/ar2003368.

491

(9)

Levard, C.; Reinsch, B. C.; Michel, F. M.; Oumahi, C.; Lowry, G. V; Brown Jr., G. E.

492

Sulfidation processes of PVP-coated silver nanoparticles in aqueous solution: Impact on

493

dissolution rate. Environ. Sci. Technol. 2011, 45 (12), 5260–5266; DOI

494

10.1021/es2007758.

495

(10)

http://www.physicalgeography.net/fundamentals/10t.html (accessed on August, 2017).

496

497

Pidwirny, M. Introduction to Soils

(11)

Prasad, M.; Saxena, S.; Amritphale, S. S.; Chandra, N. Kinetics and isotherms for aqueous

498

lead adsorption by natural minerals. Ind. Eng. Chem. Res. 2000, 39, 3034–3037; DOI

499

10.1021/ie9909082.

500

(12)

Mikutta, R.; Baumgärtner, A.; Schippers, A.; Haumaier, L.; Guggenberger, G.

501

Extracellular polymeric substances from Bacillus subtilis associated with minerals modify

502

the extent and rate of heavy metal sorption. Environ. Sci. Technol. 2012, 46 (7), 3866–

503

3873; DOI 10.1021/es204471x.

ACS Paragon Plus Environment

Environmental Science & Technology

504

(13)

Zaunbrecher, L. K.; Cygan, R. T.; Elliott, W. C. Molecular models of cesium and

505

rubidium adsorption on weathered micaceous minerals. J. Phys. Chem. A 2015, 119 (22),

506

5691–5700; DOI 10.1021/jp512824k.

507

(14)

El Badawy, A. M.; Luxton, T. P.; Silva, R. G.; Scheckel, K. G.; Suidan, M. T.; Tolaymat,

508

T. M. Impact of environmental conditions (pH, ionic strength, and electrolyte type) on the

509

surface charge and aggregation of silver nanoparticles suspensions. Environ. Sci. Technol.

510

2010, 44 (4), 1260–1266; DOI 10.1021/es902240k.

511

(15)

Cooper, J. B. Chemometric analysis of Raman spectroscopic data for process control

512

applications. Chemom. Intell. Lab. Syst. 1999, 46 (2), 231-247; DOI 10.1016/S0169-

513

7439(98)00174-9.

514

(16)

Liang, S.; Singh, M.; Dharmaraj, S.; Gam, L.-H. The PCA and LDA analysis on the

515

differential expression of proteins in breast cancer. Dis. Markers 2010, 29 (5), 231–242;

516

DOI 10.3233/DMA20100753.

517

(17)

Roggo, Y.; Edmond, A.; Chalus, P.; Ulmschneider, M. Infrared hyperspectral imaging for

518

qualitative analysis of pharmaceutical solid forms. Anal. Chim. Acta 2005, 535 (1–2), 79–

519

8; DOI 10.1016/j.aca.2004.12.037.

520

(18)

Lowry, G. V; Gregory, K. B.; Apte, S. C.; Lead, J. R. Transformations of nanomaterials in

521

the environment. Environ. Sci. Technol. 2012, 46 (13), 6893–6899; DOI

522

10.1021/es300839e.

523

(19)

Pavel, I. E.; Alnajjar, K. S.; Monahan, J. L.; Stahler, A.; Hunter, N. E.; Weaver, K. M.;

524

Baker, J. D.; Meyerhoefer, A. J.; Dolson, D. A. Estimating the analytical and surface

525

enhancement factors in surface-enhanced Raman scattering (SERS): A novel physical

ACS Paragon Plus Environment

Page 26 of 33

Page 27 of 33

Environmental Science & Technology

526

chemistry and nanotechnology laboratory experiment. J. Chem. Educ. 2012, 89 (2), 286–

527

290; DOI 10.1021/ed200156n.

528

(20)

Pavel, I.; McCarney, E.; Elkhaled, A.; Morrill, A.; Plaxco, K.; Moskovits, M. Label-free

529

SERS detection of small proteins modified to act as bifunctional linkers. J. Phys. Chem. C

530

2008, 112, 4880–4883; DOI 10.1021/jp710261y.

531

(21)

Vlčková, B.; Moskovits, M.; Pavel, I.; Šišková, K.; Sládková, M.; Šlouf, M. Single-

532

molecule surface-enhanced Raman spectroscopy from a molecularly-bridged silver

533

nanoparticle dimer. Chem. Phys. Lett. 2008, 455 (4), 131–134; DOI

534

10.1016/j.cplett.2008.02.078.

535

(22)

Waterhouse, G. I. N.; Bowmaker, G. A.; Metson, J. B. The thermal decomposition of

536

silver (I, III) oxide: A combined XRD, FT-IR and Raman spectroscopic study. Phys.

537

Chem. Chem. Phys. 2001, 3 (17), 3838–3845; DOI 10.1039/b103226g.

538

(23)

Dorney, K. A chemical free approach for increasing the biochemical surface-enhanced

539

Raman spectroscopy (SERS)-based sensing capabilities of colloidal silver nanoparticles.

540

Master of Science thesis, Wright State University, Dayton, OH, 2014.

541

(24)

Nie, S.; Emory, S. R. Probing single molecules and single nanoparticles by surface-

542

enhanced Raman scattering Science 1997, 275 (5303), 1102–1106; DOI

543

10.1126/science.275.5303.1102.

544

(25)

Xu, H.; Bjerneld, E. J.; Käll, M.; Börjesson, L. Spectroscopy of single molecule

545

hemoglobin molecules by surface enhanced Raman scattering. Phys. Rev. Lett. 1999, 83

546

(21), 4357–4360.

ACS Paragon Plus Environment

Environmental Science & Technology

547

(26)

Michaels, A. M.; Nirmal, M.; Brus, L. E. Surface enhanced Raman spectroscopy of

548

individual rhodamine 6G molecules on large Ag nanocrystals. J. Am. Chem. Soc. 1999,

549

121, 9932-9939; DOI 10.1021/ja992128q.

550

(27)

data analysis and imaging. J. Open Res. Softw. 2016, 4 (e4), 1-8 ; DOI 10.5334/jors.91.

551

552

Foose, D. P.; Sizemore, I. E. P. Vespucci: A free, cross-platform tool for spectroscopic

(28)

Stahler, A. C.; Monahan, J. L.; Dagher, J. M.; Baker, J. D.; Markopoulos, M. M.; Iragena,

553

D. B.; NeJame, B. M.; Slaughter, R.; Felker, D.; Burggraf, L. W.; Isaac, L. A. C.; Grossie,

554

D.; Gagnon, Z. E.; Sizemore I. E. P. Evaluating the abnormal ossification in tibiotarsi of

555

developing chick embryos exposed to 1.0 ppm doses of platinum group metals by

556

spectroscopic techniques. Bone 2013, 53, 421-429; DOI 10.1016/j.bone.2012.12.051.

557

(29)

Creighton, J. A.; Blatchford, C. G.; Albrecht, M. G. Plasma resonance enhancement of

558

Raman scattering by pyridine adsorbed on silver or gold sol particles of size comparable

559

to the excitation wavelength. J. Chem. Soc. Faraday Trans. 2 1979, 75, 790-798; DOI

560

10.1039/F29797500790.

561

(30)

Mulfinger, L.; Solomon, S. D.; Bahadory, M.; Jeyarajasingam, A. V; Rutkowsky, S. A.;

562

Boritz, C. Synthesis and study of silver nanoparticles. J. Chem. Educ. 2007, 84 (2), 322-

563

325; DOI 10.1021/ed084p322.

564

(31)

Pisarek, M.; Nowakowski, R.; Kudelski, A.; Holdynski, M.; Roguska, A.; Janik-Czachor,

565

M.; Kurowska-Tabor, E.; Sulka, G. D. Surface modification of nanoporous alumina layers

566

by deposition of Ag nanoparticles Effect of alumina pore diameter on the morphology of

567

silver deposit and its influence on SERS activity. Appl. Surf. Sci. 2015, 357, 1736-1742;

568

DOI 10.1016/j.apsusc.2015.10.011.

ACS Paragon Plus Environment

Page 28 of 33

Page 29 of 33

569

Environmental Science & Technology

(32)

Ossig, R.; Kwon, Y.-H.; Hubenthal, F.; Kronfeldt, H.-D. Naturally grown Ag

570

nanoparticles on quartz substrates as SERS substrate excited by a 488 nm diode laser

571

system for SERDS. Appl. Phys. B 2012, 106 (4), 835–839; DOI 10.1007/s00340-011-

572

4866-8.

573

(33)

Hwang, J.-S.; Chen, K.-Y.; Hong, S.-J.; Chen, S.-W.; Syu, W.-S.; Kuo, C.-W.; Syu, W.-

574

Y.; Lin, T. Y.; Chiang, H.-P.; Chattopadhyay, S.; Chen, K.-H.; Chen, L.-C. The

575

preparation of silver nanoparticle decorated silica nanowires on fused quartz as reusable

576

versatile nanostructured surface-enhanced Raman scattering substrates. Nanotech. 2010,

577

21 (025502), 1-6; DOI 10.1088/0957-4484/21/2/025502.

578

(34)

Yang, T.; Wang, E.; Wang, F.; Chou, K.; Hou, X. Fabrication of ordered mullite

579

nanowhisker array with surface enhanced Raman scattering effect. Sci. Rep. 2015, 5

580

(9690), 1-7; DOI 10.1038/srep09690.

581

(35)

Stetciura, I.; Markin, A. V.; Ponomarev, A. N.; Yakimansky, A. V.; Demina, T. S.;

582

Grandfils, C.; Volodkin, D. V.; Gorin, D. A. New surface-enhanced Raman scattering

583

platforms: Composite calcium carbonate microspheres coated with astralen and silver

584

nanoparticles. Langmuir 2013, 29, 4140-4147; DOI 10.1021/la305117t.

585

(36)

Sun, L.; Song, Y.; Wang, L.; Guo, C.; Sun, Y.; Liu, Z.; Li, Z. Ethanol-induced formation

586

of silver nanoparticle aggregates for highly active SERS substrates and application in

587

DNA detection. J. Phys. Chem. C. 2008, 112, 1415-1422; DOI 10.1021/jp075550z.

588

(37)

Tolaymat, T. M.; El Badawy, A. M.; Genaidy, A.; Scheckel, K. G.; Luxton, T. P.; Suidan, M.

589

2010. An evidence-based environmental perspective of manufactured silver nanoparticle in

590

syntheses and applications: A systematic review and critical appraisal of peer-reviewed scientific

ACS Paragon Plus Environment

Environmental Science & Technology

papers. Sci. Total Environ. 2010, 408, 999-1006.

591 592

(38)

U.S. E.P.A. Secondary Drinking Water Standards: Guidance for Nuisance Chemicals at

593

https://www.epa.gov/dwstandardsregulations/secondary-drinking-water-standards-

594

guidance-nuisance-chemicals (accessed September 2017).

595

(39)

theory. Geochim. Cosmochim. Acta 1994, 58 (14), 3123–3129.

596

597

Sverjensky, D. A. Zero-point-of-charge prediction from crystal chemistry and solvation

(40)

Yan, L.; Englert, A. H.; Masliyah, J. H.; Xu, Z. Determination of anisotropic surface

598

characteristics of different phyllosilicates by direct force measurements. Langmuir 2011,

599

27 (21), 12996-13007; DOI 10.1021/la2027829.

600

(41)

1996.

601

602

(42)

Brunauer, S.; Emmett, P. H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60 (2), 309-319.

603

604

McLAren, R. G.; Cameron, K. C. Soil Science, Second Edition, Oxford University Press,

(43)

Kneen, M. A.; Annegarn, H. J. Algorithm for fitting XRF, SEM and PIXE X-Ray spectra

605

backgrounds. Nucl. Instruments Methods Phys. Res. Sect. B Beam Interact. with Mater.

606

Atoms 1996, 109–110 (1), 209–213.

607

(44)

Environmental Monitoring Systems Laboratory; McDaniel, W. Methods for the Determination of Metals in Environmental Samples. 1996, pp 24–30.

608

609

(45)

Agency, U. S. E. P. A Fed. Regist. 1994, 4.

610

(46)

Montano, M. D.; Ranville, J.; Lowry, G. V; Blue, J.; Hiremath, N.; Koenig, S.; Tuccillo,

611

M. E.; Gardner, S. P. in Detection and characterization of engineered nanomaterials in

ACS Paragon Plus Environment

Page 30 of 33

Page 31 of 33

Environmental Science & Technology

612

the environment: Current state-of-the-art and future directions. published by U.S.

613

Environmental Protection Agency, 2014, No. August, 186.

614

(47)

Hanson, N.; Harris, J.; Joseph, L. A.; Ramakrishnan, K.; Thompson, T. in EPA needs to

615

manage nanomaterial risks more effectively, published by U.S. Environmental Protection

616

Agency office of Inspector General, December 29, 2011, Report No. 12-P-0162.

617

(48)

Dorney, K. M.; Baker, J. D.; Edwards, M. L.; Kanel, S. R.; O’Malley, M.; Sizemore, I. E.

618

P. Tangential flow filtration of colloidal silver nanoparticles: A "green" laboratory

619

experiment for chemistry and engineering students. J. Chem. Educ. 2014, 91 (7), 1044–

620

1049; DOI 10.1021/ed400686u.

621

(49)

Brittle, S. W.; Paluri, S. L. A.; Foose, D. P.; Ruis, M. T.; Amato, M. T.; Lam, N. H.;

622

Buttigieg, B.; Gagnon, Z. E.; Sizemore, I. E. Freshwater crayfish: A potential benthic-

623

zone indicator of nanosilver and ionic silver pollution. Environ. Sci. Technol. 2016, 50,

624

7056-7065; DOI 10.1021/acs.est.6b00511.

625

(50)

Paluri, S. L. A.; Ryan, J. D.; Lam, N. H.; Nepal, D.; Sizemore, I. E. Analytical-based

626

methodologies for examining the in vitro absorption, distribution, metabolism, and

627

elimination (ADME) of silver nanoparticles. Small 2017, 13 (1603093), 1-15; DOI

628

10.1002/smll.201603093.

629

(51)

muscovite. Am. Mineral. 1999, 84, 1041-1048.

630

631 632

McKeown, D. A.; Bell, M. I.; Etz, E. S. Vibrational analysis of the dioctahedral mica: 2M1

(52)

Tlili, A.; Smith, D. C.; Beny, J. -M.; Boyer, H. A Raman microprobe study of natural micas. Mineral. Mag. 1989, 53, 165-179.

ACS Paragon Plus Environment

Environmental Science & Technology

633

(53)

phyllosilicates. Indian J. Pure App. Phy. 2016, 54, 116-122.

634

635

(54)

Wang, A.; Freeman, J. J., Jolliff, B. L. Understanding the Raman spectral features of phyllosilicates. J. Raman Spectrosc. 2015, 46(10), 829-845; DOI 10.1002/jrs.4680.

636

637

Singha, M.; Singh, L. Vibrational spectroscopic study of muscovite and biotite layered

(55)

Trefry, J. C.; Monahan, J. L.; Weaver, K. M.; Meyerhoefer, A. J.; Markopolous, M. M.;

638

Arnold, Z. S.; Wooley, D. P.; Pavel, I. E. Size-selection and concentration of silver

639

nanoparticles by tangential flow ultrafiltration for SERS-based biosensors. J. Am. Chem.

640

Soc. 2010, 132, 10970-10972; DOI 10.1021/ja103809c.

641

(56)

Michaels, A. M.; Nirmal, M.; Brus, L. E. Surface enhanced Raman spectroscopy of

642

individual rhodamine 6G molecules on large Ag nanocrystals. J. Am. Chem. Soc. 1999,

643

121, 9932-9939; DOI 10.1021/ja992128q.

644

(57)

10091-1012; DOI 10.1063/1.1711980.

645

646

Porto, S. P. S.; Krishnan, R. S. Raman effect of corundum. J. Chem. Phys. 1967, 47 (11),

(58)

Pezzotti, G.; Zhu, W. Resolving stress tensor components in space from polarized Raman

647

spectra: polycrystalline alumina. Phys. Chem. Chem. Phys. 2015, 17 (4), 2608–2627; DOI

648

10.1039/c4cp04244a.

649

(59)

Hass, K. C.; Schneider, W. F.; Curioni, A.; Andreoni, W. The chemistry of water on

650

alumina surfaces: Reaction dynamics from first principles. Science 1998, 282 (5387),

651

265–268.

652 653

(60)

Martina, I.; Wiesinger, R.; Jemrih-Simbuerger, D.; Schreiner, M. Micro-Raman characterisation of silver corrosion products: Instrumental setup and reference database. e-

ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

Environmental Science & Technology

Preservation Sci. 2012, 9, 1–8.

654

655

(61)

Zabinski, J. S.; McDevitt, N. T. in Raman Spectra of Inorganic Compounds related to

656

solid state tribochemical studies, Report WL-TR-96-4034, Wright Laboratory, Wright-

657

Patterson Air Force Base, OH, 1996.

658

659

TOC ART

660

ACS Paragon Plus Environment