Reductive Dechlorination of Trichloroethene by Zero-valent Iron

Nov 8, 2016 - Department of Civil, Environmental, and Construction Engineering, Texas Tech University, Lubbock, Texas 79409, United States ... (3, 9, ...
0 downloads 7 Views 838KB Size
Subscriber access provided by University of Otago Library

Article

Reductive Dechlorination of Trichloroethene by Zero-valent Iron Nanoparticles: Reactivity Enhancement through Sulfidation Treatment Yanlai Han, and Weile Yan Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b03997 • Publication Date (Web): 08 Nov 2016 Downloaded from http://pubs.acs.org on November 8, 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.

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

150x55mm (96 x 96 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

1

Reductive Dechlorination of Trichloroethene by Zero-valent Iron

2

Nanoparticles: Reactivity Enhancement through Sulfidation Treatment

3

Yanlai Han1, Weile Yan1,*

Page 2 of 33

4 5

1

Department of Civil, Environmental, and Construction Engineering, Texas Tech University,

6

Lubbock, TX, 79409, USA

7 8 9 10 11 12 13 14 15 16 17 18 19 20 21

Corresponding author. Tel: +1 806 834 3478; Fax: +1 806 742 3449

22

Email address: [email protected]

23

1 ACS Paragon Plus Environment

Page 3 of 33

Environmental Science & Technology

24

Abstract

25

Zero-valent iron nanoparticles (nZVI) synthesized in the presence of reduced sulfur compounds

26

have been shown to degrade trichloroethene (TCE) at significantly higher rates. However, the

27

applicability of sulfidation as a general means to enhance nZVI reactivity under different particle

28

preparation conditions and the underlying cause for this enhancement effect are not well

29

understood. In this study, the effects of sulfidation reagent, time point of sulfidation, and sulfur

30

loading on the resultant particles were assessed through TCE degradation experiments. Up to 60-

31

fold increase in TCE reaction rates was observed upon sulfidation treatment, with products being

32

fully dechlorinated hydrocarbons. While the reactivity of these sulfur-treated nZVI (S-nZVI) was

33

relatively unaffected by the sulfidation reagent (viz., sodium sulfide, dithionite, or thiosulfate) or

34

the sequence of sulfidation relative to iron reduction, TCE reaction rates were found to depend

35

strongly on sulfur to iron ratio. At a low sulfur loading, TCE degradation was accelerated with

36

increasing sulfur dose. The rate constant reached a limiting value, however, as the sulfur to iron

37

mole ratio was greater than 0.025. Different from previous propositions that iron sulfidation

38

leads to more efficient TCE or tetrachloroethene (PCE) degradation by enabling depassivation of

39

iron surface, affording catalytic pathways, or facilitating electron transfer, we show that the role

40

of sulfur in nZVI lies essentially in its ability to poison hydrogen recombination, which drives

41

surface reactions to favor reduction by atomic hydrogen. This implies that the reactivity of S-

42

nZVI is contaminant-specific and is selective against the background reaction of water reduction.

43

As the effect of sulfur manifests through surface processes, sulfidation represents a broadly

44

applicable surface modification approach to modulate or increase the reactivity of nZVI for

45

treating TCE and other related contaminants.

46 2 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 33

47

1. Introduction

48

Due to widespread historical applications in a broad range of industrial and commercial

49

processes and their persistence again natural attenuation [1, 2], trichloroethene (TCE) and

50

tetrachloroethene (PCE) in the form of dissolved chemicals or non-aqueous phase liquid (NAPL)

51

are the most frequently encountered contaminants at the U.S. superfund sites [3,4, 5]. Compared

52

to microbial-mediated reduction of PCE and TCE, which tends to produce toxic intermediates

53

such as dichloroethenes (DCEs) and vinyl chloride (VC), abiotic dechlorination undergoes

54

predominantly a reductive elimination pathway to yield completely dechlorinated products of

55

benign nature (e.g., acetylene, ethene, or ethane) [6,7,8]. With recently reported TCE and PCE

56

transformation by reduced iron minerals under field conditions [9,10], there is a surge of interest

57

in using indigenous or engineered abiotic materials to improve the remediation performance of

58

sites impacted by chlorinated ethenes. To this end, two broad categories of iron materials have

59

been evaluated. Ferrous-containing minerals such as iron sulfides (e.g., mackinawite or pyrite),

60

magnetite, and green rust are able to reduce chlorinated ethenes to acetylene [7,11,12,13,14].

61

These Fe(II)-containing minerals are naturally present in subsurface soils, or their formation can

62

be stimulated under conditions favorable for biologically mediated sulfate or Fe(III) reduction in

63

processes known as in situ biogeochemical treatment [3,9,10]. However, reductive

64

dechlorination on ferrous minerals is relatively slow and the minerals appear to possess limited

65

reduction capacities [12,15], thus it requires high mass loadings of the solids to outcompete the

66

less desirable biological reduction pathways [16]. Another form of iron materials extensively

67

studied for the degradation of chlorinated ethenes is zero-valent iron (ZVI) [17-21]. Various

68

forms of ZVI, including iron granules and powder, colloidal iron nanoparticles (nZVI), and

69

bimetallic iron carrying a small amount of catalyst metal (e.g., Pd-Fe and Ni-Fe) have been 3 ACS Paragon Plus Environment

Page 5 of 33

Environmental Science & Technology

70

studied in the past. In spite of the intrinsic reactivity of ZVI materials, corrosion of iron in

71

aqueous solutions causes spontaneous surface passivation and the catalyst additives on ZVI are

72

prone to deactivation by common groundwater solutes [22, 23,24,25].

73

Recent studies have reported enhanced reactivity of ZVI towards chlorinated contaminants in the

74

presence of sulfur compounds [26,27,28,29,30,31]. Hassan [26] observed that iron filings

75

containing sulfur impurities were more efficient at TCE degradation than high purity iron. Butler

76

and Hayes noted enhanced reduction of chlorinated ethenes when the reaction mixture was

77

amended with sulfide ion [11]. By adding sodium dithionite into the synthesis broth of nZVI,

78

Kim et al. created a Fe/FeS nanocomposite material with up to 20-fold increases in TCE removal

79

rates [27,28]. Similarly, iron nanoparticles that had been conditioned in dilute sulfide or

80

dithionite solutions were found to degrade TCE [29,31] and 1,2-dichloroethane [30] more rapidly.

81

While multiple explanations have been postulated on the origin of the enhanced reactivity caused

82

by iron sulfidation, including a catalytic effect ascribed to the iron sulfides formed on the particle

83

surface [26], more efficient charge transfer mediated by the sulfides [27, 29,32], and increased

84

depassivation of iron surface [33,34], these views remain largely hypothetical awaiting

85

experimental verification. Moreover, variations in experiment conditions, type of iron substrates

86

used, and sulfidation procedures in these studies preclude the identification of critical factors

87

controlling the reactivity of sulfur-modified iron. As a result, the broader implications of

88

sulfidation as a means to increase the performance of ZVI materials for the treatment of

89

chlorinated contaminants are unclear.

90

The objective of this study was to examine the effects of sulfidation on the physicochemical

91

characteristics of nZVI and their reactivity in TCE dechlorination experiments. Our choice of

92

nanosized ZVI stems from their consistent quality and an ability to manipulate particle synthesis 4 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 33

93

conditions to accommodate different sulfidation procedures. While the focus of the present study

94

was on nZVI, our ongoing investigations suggest that sulfidation is applicable to other forms of

95

iron materials such as commercial ZVI products, thus the findings presented here will lend

96

relevant insights into the general role of sulfur in modulating the reactivity of ZVI materials.

97

2. Materials and methods

98

2.1 Chemicals and Materials

99

All chemicals used are listed in the Supporting Information. Deoxygenated deionized-distilled

100

water (DDI), prepared by purging DDI with N2 for 30 min, was used in all procedures including

101

material synthesis and TCE dechlorination experiments. nZVI was synthesized using the

102

borohydride reduction method [23]. Sulfur-treated nZVI particles (denoted as S-nZVI) were

103

prepared using two approaches. The first approach follows that of Kim et al 27, which involves

104

amending an appropriate amount of sulfidation reagent to an Fe(III) solution prior to the addition

105

of borohydride and is referred to as pre-synthesis S-nZVI. Three common sulfur compounds

106

were evaluated as sulfidation reagents in this study, namely, sodium sulfide (Na2S), sodium

107

dithionite (Na2S2O4), and sodium thiosulfate (Na2S2O3). The dose of the sulfur compound was

108

varied such that the mole ratio of the sulfur reagent to the initial concentration of ferric salt in the

109

synthesis mixture (denoted as S/Fe mole ratio) was the range of 1.25 x 10-3 to 0.75. In the

110

second approach, a sulfidation reagent was dosed into the synthesis mixture at 20 min after the

111

onset of Fe(III) reduction via the addition of borohydride. The resultant particles are denoted as

112

post-synthesis S-nZVI. Amorphous iron sulfide (FeS) was synthesized in the lab following the

113

method by Butler and Hayes [35]. All iron sulfide or sulfided iron particles were used

114

immediately in subsequent experiments upon preparation.

5 ACS Paragon Plus Environment

Page 7 of 33

Environmental Science & Technology

115

To investigate the mechanism of sulfur-induced reactivity improvement, a small amount of

116

arsenic-modified nZVI was prepared following the same post-synthesis method as that of S-

117

nZVI, except that the sulfur compound (sodium thiosulfate) was replaced with equivalent moles

118

of sodium arsenite (NaAsO2).

119

2.2. TCE dechlorination experiments

120

Batch TCE dechlorination experiments were performed to compare the reactivity of S-nZVI

121

prepared under different conditions. All experiments were conducted in 45-mL EPA vials

122

containing 30 mL of aqueous solution and the balance as headspace. The initial pH of all

123

solutions was adjusted to between 7.8-8.2 using dilute NaOH or HCl to simulate the typical pH

124

in groundwater. The solutions were amended with 5 g/L of particles (dry weight). The vials were

125

capped with PTFE-lined mininert valves. Experiments were started by injecting a small volume

126

of TCE stock solution in methanol to reach an initial TCE concentration of 25 mg/L. The

127

reactors were placed on a wrist-action shaker at 250 rpm at 22 +/- 1 oC. Control experiments

128

without iron materials or with unmodified nZVI were performed in parallel. Periodically, an

129

aliquot (25 – 50 μL) of headspace gas was withdrawn using a gastight syringe. The samples were

130

directly injected into a GC-FID system (Agilent 6890) equipped with an Agilent PoraPlot Q

131

column (25 m x 0.32 mm) to analyze for the concentrations of TCE, chlorinated intermediates

132

(not detected in this study), acetylene, ethene, ethane, and longer chain hydrocarbons (up to C6).

133

The analysis conditions are described in SI. This method provides adequate separation between

134

TCE and the daughter products. TCE calibration line was constructed by headspace analysis of

135

TCE aqueous standard solutions prepared in the same type of vials as the experimental reactors.

136

Calibrations for C2-C6 hydrocarbons were performed using commercial gas standards as

137

mentioned in the SI. The results were used to compute their total concentrations in the reaction 6 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 33

138

vials after accounting for partition between headspace and aqueous phases using the respective

139

Henry’s Law constants 36, 37 (Table S1). Details on H2 evolution measurements and isotope

140

fractionation analysis of TCE during reactions with S-nZVI are available in the SI.

141

The solids were subject to microscopic, crystallographic, and surface chemistry characterizations

142

and the details are described in the SI.

143

3. Results and discussion

144

3.1. Effect of sulfidation conditions

145

The effect of sulfidation reagent on the reactivity of nZVI for TCE dechlorination was evaluated.

146

The particles shown in Figure 1 were prepared using different sulfur compounds at a constant

147

S/Fe mole ratio of 0.05 following the post-synthesis sulfidation procedure. As a comparison, the

148

inset of Figure 1 shows TCE degradation by fresh nZVI prepared under equivalent conditions but

149

without exposure to any sulfur reagent. It is evident that all sulfur-amended nZVI displayed

150

remarkable improvements in TCE dechlorination rates. The observed mass-normalized pseudo-

151

first-order reaction rate constants (km) of various s-nZVI were approximately 60 folds higher

152

than that of the untreated nZVI (Table 1). Distribution of products was qualitatively similar

153

among different S-nZVI, with ethene being the dominant product, accompanied by lesser

154

amounts of ethane, acetylene, and heavier hydrocarbons (mixture of C3-C6 alkanes and alkenes)

155

(Table 1). Dichloroethene isomers (DCEs) and vinyl chloride (VC), common intermediates

156

generated by hydrogenolysis reactions, were not detected in nZVI or S-nZVI systems during the

157

course of experiments, which agrees with prior studies that reduction of TCE on abiotic surfaces

158

occurs predominantly via a dichloro-elemination pathway bypassing the formation of chlorinated

159

intermediates [21,38]. 7 ACS Paragon Plus Environment

Page 9 of 33

Environmental Science & Technology

160

In prior studies, dithionite and sulfide ions have been employed to restore the reactivity of

161

passivated ZVI [39] or to synthesize Fe(0)/FeS nanocomposite materials [27,29,31]. Aqueous

162

sulfide (as H2S or HS- at near neutral pH) is a corrosive chemical and its attaking on iron results

163

in deposition of a layer of iron sulfide (FeS) on the surface. Hydrolysis of dithionite in acidic

164

solutions gives rise to thiosulfate and sulfite [27, 40] (R1). At an alkaline pH, dithionite may

165

hydrolyze via another pathway producing sulfite and sulfide [39, 41] (R2). Disproportionation of

166

thiosulfate leads to the formation of elemental sulfur and sulfite (R3) [42]. Elemental sulfur may

167

react with iron directly (R4) or convert to sulfide that subsequently binds with iron to form FeS

168

[43,44]. As FeS is cathodic to Fe(0), its formation propels further corrosion of Fe(0).

169

2S2O42- + H2O  S2O32- + 2HSO3-

(R1)

170

S2O42- + 6OH- 5SO32- + S2- + 3H2O

(R2)

171

S2O32-  S0 + SO32-

(R3)

172

Fe(0) + S0  FeS

(R4)

173

In principle, sulfidation of the ZVI material can be achieved with the use of either thiosulfate,

174

dithionite, or free sulfide. Dithionite is a fairly strong reductant, especially under alkaline pH

175

[39,45], for which it has been proposed as a reductant to prepare nZVI [46]. Thiosulfate does not

176

have an as strong reducing capability, but it readily decomposes to release elemental sulfur or

177

sulfide (R3), and the former is reduced in the presence of Fe(0) to sulfide [42]. Thus thiosulfate

178

effectively serves as a source of sulfide in aqueous nZVI suspension. The sulfide salt used, Na2S,

179

is highly hygroscopic and tends to absorb moisture and CO2 in the air, posing material storage

180

and handling difficulties. Furthermore, the rapid release of toxic fume upon addition of a sulfide

181

chemical raises process safety concerns. In field applications, the above considerations are 8 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 33

182

significant enough to justify the choice of dithionite or thiosulfate. Considering the availability of

183

thiosulfate (both as a synthetic chemical and a naturally occuring sulfur compound) and the

184

concern that excess dose of dithionite may consume Fe(0) [31], thiosulfate was chosen as the

185

sulfidation reagent in all subsequent experiments.

186

In the method proposed by Kim et al. [27], dithionite was introduced into the synthesis solutions

187

prior to Fe(III) reduction by borohydride. More recently, post-synthesis sulfidation involving

188

reacting pre-formed nZVI in sulfide solutions has been employed [29,47]. To assess whether the

189

time point of suflidation exerts an effect on the nature of the particles formed, we prepared pre-

190

and post-synthesis S-nZVI using thiosulfate at the same S/Fe mole ratio (0.05). The morphology

191

of the solids emerging from the two preparations was considerably different. Under TEM, the

192

pre-synthesis S-nZVI consists of a heterogeneous mixture of spherical particles that are typical

193

of solution-derived iron nanoparticles (indicated by a white arrow) together with some cubic (red

194

arrow) and platy (blue arrow) structures that resemble iron sulfides or oxides (Figure 2b). In

195

contrast, the post-synthesis S-nZVI sample in Figure 2c shows more uniform structure

196

characterized by spherical particles aggregating in string-like clusters. The appearance was akin

197

to that of the unmodified nZVI prepared in our earlier studies [48,49]. The surface chemistry of

198

the pre- and post-synthesis S-nZVI was analyzed with X-ray photoelectron spectroscopy (XPS).

199

The S 2p spectra of pre- and post-synthesis S-nZVI are shown in Figures 2d and 2e. The spectra

200

were fitted with S 2p3/2 and S 2p1/2 spin-orbit doulets that are sperated by 1.2 eV with an intensity

201

ratio of 2:1 [50]. Peak assignement was based on literature reported binding energies of sulfide

202

minerals [50, 51,52] and the spectra of reference materials acquired under the same conditions as

203

the samples. Pre-synthesis S-nZVI carried prodominantly monosulfide (S2-) and disulfide (S22-),

204

accounting for 63 and 37 atomic percents (at.%) of total sulfur species, respectively. The surface 9 ACS Paragon Plus Environment

Page 11 of 33

Environmental Science & Technology

205

of post-synthesis S-nZVI consists mainly of S2- (34 at.%) and S22- (46 at.%), with S22-

206

contributing a higher portion than that in the pre-synthesis S-nZVI. The post-synthesis sample

207

also features a group of low-rising peaks in the binding energy range of 163.3 – 164.3 eV,

208

corresponding to polysulfides (Sn2-) and possibly elemental sulfur, and a sulfate (SO42-)

209

component at the highest binding energy (167.6 eV). Comparison of the two S 2p spectra

210

suggests that oxidation of the sulfur precursor has occcurred to a greater extent during the post-

211

synthesis sulfidation process.

212

X-ray diffraction analysis (Figure S2) detected the presence of Fe(0) and magnetite in both types

213

of S-nZVI, with the post-synthesis S-nZVI exhibiting a lower degree of crystallinity due to

214

broadening of diffraction peaks. No mono-, di-, or polysulfides of iron can be discerned in the

215

diffraction spectra, which in conjunction with the XPS analysis confirms that sulfide formation

216

on the surface is amorphous. This observation is consistent with the notion that rapid corrosion

217

of iron in sulfidic water tends to produce poorly ordered iron sulfides [53].

218

We noticed that post-synthesis S-nZVI can be efficiently separated from the aqueous phase

219

through a filtration step, while the pre-synthesis S-nZVI had a significant portion of the solids

220

passing through a 0.2 μM filter. This observation was likely caused by a tendency of post-

221

synthesis nZVI to form aggregates and the presence of fine, loose iron sulfide or oxide particles

222

in the pre-synthesis S-nZVI as suggested by TEM images.

223

In spite of significant structrual differences, the two forms of S-nZVI exhibit similar reactive

224

behavior in TCE dechlorination experiments. km for pre-synthesis and post-synthesis S-nZVI

225

was 0.9 ± 0.1 x 10-3 and 0.8 ± 0.05 x 10-3 L/g-min, respectively, and the product composition

226

matches closely with each other (Table 1). In view of the uniform texture of particles prepared

10 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 33

227

via the post-synthesis method and their amenability to fast solid/liquid separation, this method

228

was adopted to produce S-nZVI in all subsequent experiments.

229

3.2. Effect of sulfur to iron ratio

230

A series of S-nZVI were prepared using different doses of thiosulfate such that the S/Fe mole

231

ratio in the synthesis solutions varied in the range of 1.25 x 10-3 to 0.75. The concentration of

232

thiosulfate after S-nZVI synthesis was measured, and the amount of sulfur deposited on nZVI,

233

estimated based on thiosulfate consumption, correlates well with the initial S/Fe mole ratio

234

(Table S2 in SI). It was found that the loading of sulfur on iron has a strong impact on TCE

235

degradation rates. With particles prepared using the highest sulfur dose (S/Fe = 0.75, Figure 3a),

236

ethene and ethane were the dominant products, accounting for 70% and 17%, respectively, of

237

total products identified (product yields were determined at approximately 90% TCE conversion).

238

Close inspection of Figure 3a indicates there was an accumulation of acetylene during the initial

239

phase of the reaction, nonetheless, its concentration declined over time accompanied by

240

concurrent increases in ethene and ethane concentrations. Partially dechlorinated intermediates

241

(i.e., DCEs or VC) were not detected in the headspace mixture. C3-C6 hydrocarbons contributed

242

to a minor fraction (11%) of the products formed. In comparison, particles treated with the

243

lowest sulfur dose, corresponding to a S/Fe mole ratio of 1.25 x 10-3, produced ethene, ethane,

244

and C3-C6 products (Figure 3b), and acetyelne was below detection limit at any sampling point.

245

Overall, a carbon recovery (as C2 equivalent) of 50% to 90% was achieved for all S-nZVI used

246

in this study. Incomplete carbon recovery has been noticed in prior studies of TCE

247

dechlroination using nZVI [38] or iron sulfide materials [7,11]. As noted in later discussion, the

248

missing carbon is likely products of acetylene polymerization reactions, which are affinitive to

11 ACS Paragon Plus Environment

Page 13 of 33

Environmental Science & Technology

249

metal surfaces, thus their quantities cannot be reliably measured using the headspace sampling

250

method.

251

The results of control experiments, namely, the reactions of TCE with unmodified nZVI and pure

252

FeS prepared from aquoues precipitation, were shown in Figure 3c and 3d. Greater than 90%

253

TCE degradtion was achieved within a time frame of 0.3 to 2 days when S-nZVI was the

254

reductant, whereas it required 21 days to attain a similar extent of TCE removal by nZVI.

255

Degradation of TCE in FeS suspsension was even slower, with only 25% TCE being degraded in

256

19 days. The apparent mass-normalized reaction rates of various solids vary by approximately

257

three orders of magnitude (Table 1). The composition of product mixture is similar for the

258

original nZVI and those receiving a low dose of sulfur (Figures 3b and 3c). At a high sulfur dose,

259

considerable accumulation of acetylene during the intermeidate stage of the reaction was

260

observed and ethene was the dominant final product (Figure 3a). Contrary to nZVI or S-nZVI,

261

TCE degradation by FeS yielded exclusively acetylene. This slow transformation of TCE to

262

acetylene by FeS without further hydrogenation of aceytlene agrees with earlier findings by other

263

investigators [7,11].

264

As all TCE degradation data conform to a first-order rate model, the effect of sulfur loading on

265

TCE reduction kinetics was assessed by plotting the mass-normalized rate constant, km, against

266

the S/Fe mole ratio. The results, shown in Figure 4a, reveal a biphasic trend. When thiosulfate

267

was applied at a small dose (S/Fe < 0.025), more rapid TCE dechlorination occurred with

268

increasing S/Fe ratio. However, when the S/Fe mole ratio exceeds 0.025, the rate constant levels

269

out approaching a limiting value with increasing sulfur loading. The highest rate constant was

270

1.3 x 10-3 L/g-min, in comparison to 2.2 x 10-5 and 1.5 x 10-6 L/g-min achieved by unmodified

271

nZVI and FeS, respectively. The effect of S/Fe ratio on product distribution is depicted in Figure 12 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 33

272

4b. The final yields of ethene or ethane, defined as the amount of product formed over TCE

273

consumed determined at the point of 90% TCE conversion, do not bear a strong correlation with

274

the S/Fe ratio. However, the maximum accumulation of acetylene during the course of TCE

275

transformation is strongly affected by the sulfur dose. For particles prepared under a low sulfur

276

loading condition (S/Fe < 0.025), no acetylene was detected in the product mixture, whereas

277

particles containing a higher sulfur dose caused a substantial buildup of acetylene before its

278

gradual conversion to downstream products.

279

Acetylene is a reactive chemical and it readily undergoes polymerization reactions on metal

280

surfaces to form longer chain hydrocarbons [54], or in the presence of a hydrogen source,

281

hydrogenates to more saturated products. Our separate experiments reacting acetylene with nZVI

282

and S-nZVI confirm that both solids are able to rapidly transform acetylene into ethene, ethane,

283

and higher order hydrocarbons (Figure S3). Notably, significant gaps in carbon recovery were

284

also observed during these reactions, which was attributed to deposition of non-volatile

285

polymerized products (>C6) on the particle surface. FeS, as expected from previous studies, did

286

not show any appreciable reaction with acetyelene. The composition of TCE daughter products

287

and the reactivity of acetylene towards different iron materials suggest that TCE reduction on

288

nZVI, S-nZVI, or FeS shares an identical pathway of β-elimination leading to acetylene

289

formation [6, 7]. In the presence of nZVI or S-nZVI, acetylene is further converted to ethene,

290

ethane, and higher order hydrocarbons, while it remains intact on FeS. The reaction pathway and

291

its pertinent kinetic parameters are shown schematically in Figure 5. Although multiple steps are

292

invovled in TCE transformation to acetylene, including surface adsorpton of TCE and

293

conversion of chloroacetylene (i.e., the immediate product of TCE β-elimination) to acetylene,

294

the rate of these sequential steps can be captured by a single kinetic parameter (k1) that 13 ACS Paragon Plus Environment

Page 15 of 33

Environmental Science & Technology

295

represents the rate-limiting step. Note that in Figure 5 we consider ethene and ethane as being

296

formed via two parallel pathways instead of going through sequential hydrogenation because

297

ethene hydrogenation by nZVI or S-nZVI was exceedingly slow (Figure S4) and the ratio of

298

ethene to ethane remained constant during TCE degradation. The values of rate constants in

299

Figure 5 were estimated from TCE reduction and acetylene hydrogenation data and are

300

summarized in Table 2. It is intersting to note that sulfiation of nZVI effectively increases the

301

value of k1, but the treatment has no enhancement effect on k2 or k3 value. Thus, the effect of

302

sulfidation is specific for TCE conversion to acetylene and it does not accelerate the subsequent

303

hydrogenation steps.

304

3.3. Role of sulfidation in iron reactivity improvement

305

A pertinent question then arises on why the incorporation of sulfur into nZVI substrate would

306

cast such a prominent effect on TCE dechlorination. Earlier studies suggest that reduced sulfur

307

compounds such as free sulfide and dithionite are able to depassivate iron surface by reducing

308

Fe(III) to Fe(II) leading to disintegration of the native oxide layer and/or the formation of Fe(II)-

309

containing oxides (e.g., magnetite) that have greater charge transfer abilities [39,53,45].

310

Depassivation effect alone is, however, unable to account for our findings here, since treating

311

nZVI with dilute acid or amending the nZVI suspension with ascobate (a reductant of Fe(III))

312

[55] did not bring about substantial improvements in TCE reduction rates compared to the

313

freshly synthesized particles (Figure S5). This suggets that the presence of sulfided iron is

314

necessary to enable the large increases in TCE degradation rates. It has been proposed that iron

315

sulfide may catalyze PCE or TCE reduction by ZVI [26]. Nevertheless, our analysis of carbon

316

isotope fractionation during TCE experiments did not record consistent shifts in TCE bulk

317

enrichment factors of the unmodified nZVI and its sulfur-treated counterparts (Table S3), nor did 14 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 33

318

saturating the reaction mixture with H2 bring about more rapid TCE removal, refuting a possible

319

catalytic role played by the sulfide formation.

320

An alternative explanation points to enhanced electron transfer from the Fe(0) core to the

321

solution phase by FeS surface deposits owing to its good electron conducting ability [27,32]. The

322

proposition is supported by the obsevatin of increased anodic currents in recent electrochemical

323

investigations [53,56], however, more rapid iron oxidation may not beget higher rates of TCE

324

transformation since the latter reaction is not limited by electron transfer but the availability of

325

atomic hydrogen [57,58]. Furthermore, the rate enhancement effect caused by accelerated iron

326

corrison is expected to apply to other reducible contaminants, such as carbon tetrachloride (CT,

327

CCl4), whose reduction is governed by a direct electron transfer process [57,59]. To this end, we

328

evaluated reactions of CT with nZVI and sulfided nZVI. The results reveal nearly identical

329

performance by nZVI and those receiving varying levels of sulfur dose (Figure S6). The effect of

330

sulfidation is therefore specific for TCE dechlorination and cannot be ascribed to a general cause

331

related to increased iron corrosion.

332

In the catalysis literature, sulfur is a potent poison of hydrogen recombination reactions on metal

333

surfaces [60,61]. In the case of iron, corrosion in anaerobic water consists of two fundamental

334

processes, namely, the transfer of electrons to protons resulting in surface-adsorbed hydrogen

335

atoms, and the recombination of hydrogen adatoms to form molecular hydrogen that bubbles off

336

the surface. The addition of sulfur on metal surface inhibits hydrogen recombindation and as a

337

result, slows down H2 evolution. This forces more atomic hydrogen to remain on the surface or

338

penetrate into the bulk substrate. Such effect has been investigated extensively and is known to

339

cause hydrogen embrittlement and stress-induced cracking of steel [62,63,64]. Nonetheless,

340

when iron is used as a chemical reductant, inhibition of hydrogen recombination would favor 15 ACS Paragon Plus Environment

Page 17 of 33

Environmental Science & Technology

341

reactions involving atomic hydrogen, and we believe this effect is the primary mechanism

342

responsible for the remarkable improvements in TCE degradtion rates, since the reduction of

343

chlorinated ethenes on iron are considered to be predominantly mediated by surface adsorbed

344

atomic hydrogen [58,59]. This action of sulfur is consistent with the highly specific effect of

345

sulfidation on TCE reduction relative to that of chlorinated compounds undertaking different

346

reduction mechanisms (e.g., carbon tetrachloride). It also corroborates with recent findings that

347

H2 production was surpressed in the presence of S-nZVI, in contrary to what would be expected

348

if the increase in reactivity is contributed by enhanced iron corrosion [29,31]. In this study,

349

H2 production by S-nZVI was also evaluated under conditions relevant to TCE dechlorination

350

experiments. The results (Figure 6a) clearly demonstrate that sulfur amendment exerts a strong

351

impact on the rate of H2 evolution. The trend of H2 production shown in Figure 6a, that there is a

352

notable decrease in H2 generation rate when S/Fe ratio increases from 0.01 to 0.05 while further

353

increase in sulfur loading does not give rise to significant reduction in H 2 generate rate, agrees

354

with the effect of S dosage on TCE dechlorination kinetics (Figure 4a). An additional argument

355

in support of the poisoning effect of sulfur comes from TCE reduction by nZVI loaded with

356

arsenic (As-nZVI), another potent deactivator of H recombination reactions [63]. The As-nZVI

357

was prepared using the same protocol as that of S-nZVI except that sodium arsenite was

358

employed in place of sodium thiosulfate. The results indicate that arsenic-modified nZVI

359

exhibited similar reactive behavior as the sulfur-modified particles towards TCE (Figure 6b),

360

releasing ethene as the dominant product. Thus, sulfur does not act as a direct facilitator of TCE

361

dechlorination reaction, but rather modifies iron surface chemistry to favor the production of a

362

key reactive species invovled in TCE reduction. When sulfur is deposited on the iron surface, it

363

induces dissolution of the native oxide and causes the surface to be more favorable for atomic

16 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 33

364

hydrogen adsorption. Meantime, inhibtion of hydrogen recombination promotes the build-up of

365

atomic hydrogen on the iron surface and its reaction with TCE. The process is illustrated

366

schematically in Figure 7.

367

Finally, the dependence of the rate constant on S/Fe ratio in Figure 4a suggests the existence of

368

an optimal surface sulfur coverage. Exceeding this optimal loading, there is no further increase in

369

reactivity, possibly due to the formation of polysulfides, discrete FeS precipitates, or oxidized

370

sulfur species that do not contribute as effectively towards reactivity enhancement. The relatively

371

low concentration of sulfur required to achieve this optimal loading implies effective sulfidation

372

can be attained in dilute solutions of sulfur reagents.

373

3.4. Environmental implications

374

ZVI is one of the most frequently applied and environmentally benign reductants for treating a

375

broad vareity of water contaminants. The ZVI chemistry has been studied extensively for

376

chlorinated ethenes. Degradation of TCE by iron alone is relativley slow, and iron reactivity

377

tends to be short-lived due to rapid passivation in air or water. Attemps to improve the

378

performance of ZVI in the past has largely concentrated on a group of bimetallic ZVI particles.

379

The catalyst metals amended on iron surface, such as Pd or Ni, are able to improve the rates of

380

contaminant reduction by catalyzing the activation of H2 [65,66]. The sulfidation method

381

examined here represents a different approach to modify ZVI reactivity. Instead of serving as a

382

catalyst, sulfur poisons a parallel reaction that competes with TCE for the electron source (i.e.,

383

water reduction and H2 evolution), thereby increasing the accessibility of atomic hydrogen for

384

TCE reduction. Although under optimal conditions, the reactivity of these sulfur-modified ZVI

385

may fall short of that of the highly reactive bimetallic materials (for example, km of fresh Pd-

17 ACS Paragon Plus Environment

Page 19 of 33

Environmental Science & Technology

386

nZVI prepared in our lab is approximately 1.4 x 10 -2 L/g-min [25]), iron sulfidation promises

387

some crucial advantages from an enviornmental chemistry perspective. It circumvents the use of

388

catalyst metals that are expensive or toxic to aquatic environment. Inhibition of hydrogen

389

evolution effectively favors the reduction of the target contaminant (e.g., TCE) against the

390

prevalent background reaction between iron and water. The last point will be of interest in large-

391

scale field implementations as material efficiency and longevity in environmental matrices are

392

important considerations in those circumstances. As the effects of sulfur manifest primarily

393

through surface processes, sulfdation can be applied as a surface treatment procedure to pre-

394

synthesized nZVI, and the method should in principle be extendable to other ZVI materials,

395

including bulk iron granules or filing that are frequently used in remediation applications.

396

Further investigations on sulfidation of other forms of ZVI material are underway.

397

Reduced sulfur compounds are ubiquitous in anoxic environment. Partially reduced sulfur anions

398

such as thiosulfate, polysulfides, and sulfite may arise as intermediates during sulfide oxidation

399

and their interconversion is strongly coupled with biogeochemical processes [67,68]. Interstingly,

400

these species are known to be strong inhibitor of catalytic systems including the bimetallic Pd-Fe

401

[25]. From this viewpoint, iron sulfidation not only offers a material that can sustain its reactivity

402

in underground matrix where reduced sulfur ligands are abundant, but also suggests possibilities

403

to optimize the outcome of iron-based remediation technologies via biogeochemical

404

manipulations.

405

18 ACS Paragon Plus Environment

Environmental Science & Technology

406

Acknowlegement

407

The authors acknowledge the start-up fund by TTU and funding support from the National

408

Science Foundation (CHE-1611465). The authors appreciate assistance from Drs. Juske Horita

409

and Kaz Suroweic at TTU for carbon isotope and H2 analyses.

410

Supporting Information

411

Data of XRD characterization, carbon isotope analysis, TCE degradation products by S-nZVI

412

prepared under different synthesis conditions, acetylene and ethene hydrogenation, and carbon

413

tetrachloride reduction by S-nZVI are available in the supporting information. The Supporting

414

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

415

Reference

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.

2.

3.

4.

5.

6.

7. 8.

Page 20 of 33

Doherty, R. E., A History of the Production and Use of Carbon Tetrachloride, Tetrachloroethylene, Trichloroethylene and 1,1,1-Trichloroethane in the United States: Part 1--Historical Background; Carbon Tetrachloride and Tetrachloroethylene. Environmental Forensics 2000, 1, (2), 69-81. Doherty, R. E., A History of the Production and Use of Carbon Tetrachloride, Tetrachloroethylene, Trichloroethylene and 1,1,1-Trichloroethane in the United States: Part 2--Trichloroethylene and 1,1,1-Trichloroethane. Environmental Forensics 2000, 1, (2), 83-93. He, Y. T.; Wilson, J. T.; Su, C.; Wilkin, R. T., Review of Abiotic Degradation of Chlorinated Solvents by Reactive Iron Minerals in Aquifers. Ground Water Monitoring and Remediation 2015, 35, (3), 57-75. Stroo, H. F.; Leeson, A.; Marqusee, J. A.; Johnson, P. C.; Ward, C. H.; Kavanaugh, M. C.; Sale, T. C.; Newell, C. J.; Pennell, K. D.; Lebron, C. A.; Unger, M., Chlorinated Ethene Source Remediation: Lessons Learned. Environmental Science & Technology 2012, 46, (12), 6438-6447. Stroo, H. F.; Unger, M.; Ward, C. H.; Kavanaugh, M. C.; Vogel, C.; Leeson, A.; Marqusee, J. A.; Smith, B. P., Remediating chlorinated solvent source zones. Environmental Science & Technology 2003, 37, (11), 224A-230A. Roberts, A. L.; Totten, L. A.; Arnold, W. A.; Burris, D. R.; Campbell, T. J., Reductive elimination of chlorinated ethylenes by zero valent metals. Environmental Science & Technology 1996, 30, (8), 2654-2659. Butler, E. C.; Hayes, K. F., Kinetics of the transformation of trichloroethylene and tetrachloroethylene by iron sulfide. Environmental Science & Technology 1999, 33, (12), 2021-2027. Lojkasek-Lima, P.; Aravena, R.; Shouakar-Stash, O.; Frape, S. K.; Marchesi, M.; Fiorenza, S.; Vogan, J., Evaluating TCE Abiotic and Biotic Degradation Pathways in a Permeable Reactive Barrier Using Compound Specific Isotope Analysis. Ground Water Monitoring and Remediation 2012, 32, (4), 53-62.

19 ACS Paragon Plus Environment

Page 21 of 33

440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490

Environmental Science & Technology

9.

10. 11.

12.

13. 14.

15. 16.

17. 18.

19. 20.

21. 22.

23. 24. 25.

26. 27.

28.

Kennedy, L. G.; Everett, J. W.; Becvar, E.; DeFeo, D., Field-scale demonstration of induced biogeochemical reductive dechlorination at Dover Air Force Base, Dover, Delaware. Journal of Contaminant Hydrology 2006, 88, (1-2), 119-136. Navy Facilities Engineering Command (NAVFAC) In Situ Biogeochemical Transformation Processes for Treating Contaminated Groundwater; 2014. Butler, E. C.; Hayes, K. F., Factors influencing rates and products in the transformation of trichloroethylene by iron sulfide and iron metal. Environmental Science & Technology 2001, 35, (19), 3884-3891. Lee, W.; Batchelor, B., Abiotic reductive dechlorination of chlorinated ethylenes by iron-bearing soil minerals. 1. Pyrite and magnetite. Environmental Science & Technology 2002, 36, (23), 51475154. Lee, W.; Batchelor, B., Abiotic, reductive dechlorination of chlorinated ethylenes by iron-bearing soil minerals. 2. Green rust. Environmental Science & Technology 2002, 36, (24), 5348-5354. Jeong, H. Y.; Anantharaman, K.; Han, Y.-S.; Hayes, K. F., Abiotic Reductive Dechlorination of cisDichloroethylene by Fe Species Formed during Iron- or Sulfate-Reduction. Environmental Science & Technology 2011, 45, (12), 5186-5194. Lee, W. J.; Batchelor, B., Reductive capacity of natural reductants. Environmental Science & Technology 2003, 37, (3), 535-541. Butler, E. C.; Chen, L. X.; Darlington, R., Transformation of Trichloroethylene to Predominantly Non-Regulated Products under Stimulated Sulfate Reducing Conditions. Ground Water Monitoring and Remediation 2013, 33, (3), 52-60. Gillham, R. W.; Ohannesin, S. F., Enhanced Degradation of Halogenated Alphatics by Zero-Valent Iron. Ground Water 1994, 32, (6), 958-967. Li, X.-q.; Elliott, D. W.; Zhang, W.-x., Zero-valent iron nanoparticles for abatement of environmental pollutants: Materials and engineering aspects. Critical Reviews in Solid State and Materials Sciences 2006, 31, (4), 111-122. Crane, R. A.; Scott, T. B., Nanoscale zero-valent iron: Future prospects for an emerging water treatment technology. Journal of Hazardous Materials 2012, 211, 112-125. Liu, Y. Q.; Majetich, S. A.; Tilton, R. D.; Sholl, D. S.; Lowry, G. V., TCE dechlorination rates, pathways, and efficiency of nanoscale iron particles with different properties. Environmental Science & Technology 2005, 39, (5), 1338-1345. Song, H.; Carraway, E. R., Catalytic hydrodechlorination of chlorinated ethenes bly nanoscale zerovalent iron. Applied Catalysis B-Environmental 2008, 78, (1-2), 53-60. Sarathy, V.; Tratnyek, P. G.; Nurmi, J. T.; Baer, D. R.; Amonette, J. E.; Chun, C. L.; Penn, R. L.; Reardon, E. J., Aging of iron nanoparticles in aqueous solution: Effects on structure and reactivity. Journal of Physical Chemistry C 2008, 112, (7), 2286-2293. Han, Y.; Yan, W., Bimetallic Nickel-Iron Nanoparticles for Groundwater Decontamination: Effect of Groundwater Constituents on Surface Deactivation. Water Research 2014, 66, 149-159. Lim, T. T.; Zhu, B. W., Effects of anions on the kinetics and reactivity of nanoscale Pd/Fe in trichlorobenzene dechlorination. Chemosphere 2008, 73, (9), 1471-1477. Han, Y.; Liu, C.; Horita, J.; Yan, W., Trichloroethene hydrodechlorination by Pd-Fe bimetallic nanoparticles: Solute-induced catalyst deactivation analyzed by carbon isotope fractionation. Applied Catalysis B-Environmental 2016, 188, 77-86. Hassan, S. M., Reduction of halogenated hydrocarbons in aqueous media: I. Involvement of sulfur in iron catalysis. Chemosphere 2000, 40, (12), 1357-1363. Kim, E.-J.; Kim, J.-H.; Azad, A.-M.; Chang, Y.-S., Facile Synthesis and Characterization of Fe/FeS Nanoparticles for Environmental Applications. Acs Applied Materials & Interfaces 2011, 3, (5), 1457-1462. Kim, E.-J.; Murugesan, K.; Kim, J.-H.; Tratnyek, P. G.; Chang, Y.-S., Remediation of Trichloroethylene by FeS-Coated Iron Nanoparticles in Simulated and Real Groundwater: Effects of Water Chemistry. Industrial & Engineering Chemistry Research 2013, 52, (27), 9343-9350. 20 ACS Paragon Plus Environment

Environmental Science & Technology

491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540

Page 22 of 33

29. Rajajayavel, S. R. C.; Ghoshal, S., Enhanced reductive dechlorination of trichloroethylene by sulfidated nanoscale zerovalent iron. Water Research 2015, 78, 144-153. 30. Garcia, A. N.; Boparai, H. K.; O'Carroll, D. M., Enhanced Dechlorination of 1,2-Dichloroethane by Coupled Nano Iron-Dithionite Treatment. Environmental Science & Technology 2016, 50, (10), 5243-5251. 31. Fan, D.; Johnson, G. O.; Tratnyek, P. G.; Johnson, R. L., Sulfidation of Nano Zerovalent Iron (nZVI) for Improved Selectivity during In- Situ Chemical Reduction (ISCR). Environmental Science & Technology 2016, (50), 9558-9565. 32. Turcio-Ortega, D.; Fan, D.; Tratnyek, P. G.; Kim, E.-J.; Chang, Y.-S., Reactivity of Fe/FeS Nanoparticles: Electrolyte Composition Effects on Corrosion Electrochemistry. Environmental Science & Technology 2012, 46, (22), 12484-12492. 33. Lipczynskakochany, E.; Harms, S.; Milburn, R.; Sprah, G.; Nadarajah, N., Degradation of carbon tetrachloride in the presence of iron and sulfur-containing compounds Chemosphere 1994, 29, (7), 1477-1489. 34. Hansson, E. B.; Odziemkowski, M. S.; Gillham, R. W., Influence of Na2S on the degradation kinetics of CCl4 in the presence of very pure iron. Journal of Contaminant Hydrology 2008, 98, (34), 128-134. 35. Butler, E. C.; Hayes, K. F., Effects of Solution Composition and pH on the Reductive Dechlorination of Hexachloroethane by Iron Sulfide. Environmental Science & Technology 1998, 32, (9), 12761284. 36. Yaws, C. L., Yaws' Handbook of Thermodynamic and Physical Properties of Chemical Compounds. In Knovel. 37. Williams, M. L., CRC Handbook of Chemistry and Physics, 76th edition. Occupational and Environmental Medicine 1996, 53, (7), 504-504. 38. Arnold, W. A.; Roberts, A. L., Pathways and kinetics of chlorinated ethylene and chlorinated acetylene reaction with Fe(O) particles. Environmental Science & Technology 2000, 34, (9), 17941805. 39. Xie, Y.; Cwiertny, D. M., Use of Dithionite to Extend the Reactive Lifetime of Nanoscale ZeroValent Iron Treatment Systems. Environmental Science & Technology 2010, 44, (22), 8649-8655. 40. Lister, M. W.; Garvie, R. C., Sodium dithionite, decomposition in aqueous solution and in the solid state. Canadian Journal of Chemistry-Revue Canadienne De Chimie 1959, 37, (9), 1567-1574. 41. Greenwood, N. N.; Earnshaw, A., Chemistry of the Elements. 2nd ed.; Elsevier Butterworth Heinemann 1997. 42. Kappes, M.; Frankel, G. S.; Sridhar, N.; Carranza, R. M., Reaction Paths of Thiosulfate during Corrosion of Carbon Steel in Acidified Brines. Journal of the Electrochemical Society 2012, 159, (4), C195-C204. 43. Macdonald, D. D.; Roberts, B.; Hyne, J. B., Corrosion of carbon-steel by wet elemental sulfur Corrosion Science 1978, 18, (5), 411-425. 44. Schmitt, G., Effect of elemental sulfur on corrosion in sour gas systems Corrosion 1991, 47, (4), 285-308. 45. Szecsody, J. E.; Fruchter, J. S.; Williams, M. D.; Vermeul, V. R.; Sklarew, D., In situ chemical reduction of aquifer sediments: Enhancement of reactive iron phases and TCE dechlorination. Environmental Science & Technology 2004, 38, (17), 4656-4663. 46. Sun, Q.; Feitz, A. J.; Guan, J.; Waite, T. D., Comparison of the reactivity of nanosized zero-valent iron (nZVI) particles produced by borohydride and dithionite reduction of iron salts Nano 2008, 3, (5), 341-349. 47. Fan, D.; Anitori, R. P.; Tebo, B. M.; Tratnyek, P. G.; Pacheco, J. S. L.; Kukkadapu, R. K.; Engelhard, M. H.; Bowden, M. E.; Kovarik, L.; Arey, B. W., Reductive Sequestration of Pertechnetate ((TcO4)-Tc-99) by Nano Zerovalent Iron (nZVI) Transformed by Abiotic Sulfide. Environmental Science & Technology 2013, 47, (10), 5302-5310.

21 ACS Paragon Plus Environment

Page 23 of 33

541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589

Environmental Science & Technology

48. Martin, J. E.; Herzing, A. A.; Yan, W. L.; Li, X. Q.; Koel, B. E.; Kiely, C. J.; Zhang, W. X., Determination of the oxide layer thickness in core-shell zerovalent iron nanoparticles. Langmuir 2008, 24, (8), 4329-4334. 49. Yan, W.; Herzing, A. A.; Kiely, C. J.; Zhang, W.-x., Nanoscale zero-valent iron (nZVI): Aspects of the core-shell structure and reactions with inorganic species in water. Journal of Contaminant Hydrology 2010, 118, (3-4), 96-104. 50. Smart, R. S.; Skinner, W. M.; Gerson, A. R., XPS of sulphide mineral surfaces: Metal-deficient, polysulphides, defects and elemental sulphur. Surface and Interface Analysis 1999, 28, (1), 101-105. 51. Herbert, R. B.; Benner, S. G.; Pratt, A. R.; Blowes, D. W., Surface chemistry and morphology of poorly crystalline iron sulfides precipitated in media containing sulfate-reducing bacteria. Chemical Geology 1998, 144, (1-2), 87-97. 52. Mullet, M.; Boursiquot, S.; Abdelmoula, M.; Genin, J. M.; Ehrhardt, J. J., Surface chemistry and structural properties of mackinawite prepared by reaction of sulfide ions with metallic iron. Geochimica Et Cosmochimica Acta 2002, 66, (5), 829-836. 53. Hansson, E. B.; Odziemkowski, M. S.; Gillham, R. W., Formation of poorly crystalline iron monosulfides: Surface redox reactions on high purity iron, spectroelectrochemical studies. Corrosion Science 2006, 48, (11), 3767-3783. 54. Egloff, G.; Lowry, C. D.; Schaad, R. E., Polymerization and decomposition of acetylene hydrocarbons. J. Phys. Chem. 1932, 36, (5), 1457-1520. 55. Suter, D.; Banwart, S.; Stumm, W., Dissolution of hydrous iron(III) oxides by reductive mechanisms Langmuir 1991, 7, (4), 809-813. 56. Kim, E.-J.; Kim, J.-H.; Chang, Y.-S.; Turcio-Ortega, D.; Tratnyek, P. G., Effects of Metal Ions on the Reactivity and Corrosion Electrochemistry of Fe/FeS Nanoparticles. Environmental Science & Technology 2014, 48, (7), 4002-4011. 57. Li, T.; Farrell, J., Reductive dechlorination of trichloroethene and carbon tetrachloride using iron and palladized-iron cathodes. Environmental Science & Technology 2000, 34, (1), 173-179. 58. Elsner, M.; Hofstetter, T. B., Current Perspectives on the Mechanisms of Chlorohydrocarbon Degradation in Subsurface Environments: Insight from Kinetics, Product Formation, Probe Molecules, and Isotope Fractionation. In Aquatic Redox Chemistry, Tratnyek, P. G.; Grundl, T. J.; Haderlein, S. B., Eds. 2011; Vol. 1071, pp 407-439. 59. Li, T.; Farrell, J., Electrochemical investigation of the rate-limiting mechanisms for trichloroethylene and carbon tetrachloride reduction at iron surfaces. Environmental Science & Technology 2001, 35, (17), 3560-3565. 60. Oudar, J., Sulfur adsorption and poisoning of metallic catalysts Catalysis Reviews-Science and Engineering 1980, 22, (2), 171-195. 61. Burke, M. L.; Madix, R. J., Hydrogen on Pd(100)-S The effect of sulfur on precursor mediated adsorption and desorption. Surface Science 1990, 237, (1-3), 1-19. 62. Radhakrishnan, T.; Shreir, L., Permeation of hydrogen through steel by electrochemical transfer-I. Influence of catalytic poisons. Electrochimica Acta 1966, 11, 1007-1021. 63. Berkowitz, B. J.; Horowitz, H. H., The role of H2S in the corrosion and hydrogen embrittlement of steel Journal of the Electrochemical Society 1982, 129, (3), 468-474. 64. Berkowitz, B. J.; Heubaum, F. H., The role of hydrogen in sulfide stress cracking of low-alloy steels Corrosion 1984, 40, (5), 240-245. 65. Alonso, F.; Beletskaya, I. P.; Yus, M., Metal-mediated reductive hydrodehalogenation of organic halides. Chemical Reviews 2002, 102, (11), 4009-4091. 66. Urbano, F. J.; Marinas, J. M., Hydrogenolysis of organohalogen compounds over palladium supported catalysts. Journal of Molecular Catalysis a-Chemical 2001, 173, (1-2), 329-345. 67. Luther, G. W.; Church, T. M.; Scudlark, J. R.; Cosman, M., Inorganic and organic sulfur cycling in salt-marsh pore waters. Science 1986, 232, (4751), 746-749.

22 ACS Paragon Plus Environment

Environmental Science & Technology

590 591 592

Page 24 of 33

68. Cantrell, K. J.; Yabusaki, S. B.; Engelhard, M. H.; Mitroshkov, A. V.; Thornton, E. C., Oxidation of H2S by iron oxides in unsaturated conditions. Environmental Science & Technology 2003, 37, (10), 2192-2199.

593

23 ACS Paragon Plus Environment

Page 25 of 33

Environmental Science & Technology

594 595 596 597 598

Figure 1. TCE dechlorination by S-nZVI prepared using different sulfidation reagents. The initial mole ratio of sulfidation reagent to iron was fixed at 0.05. Initial TCE concentration was 25 mg/L. Inset shows TCE degradation by fresh nZVI without sulfidation treatment. The particle dose was 5 g/L in all experiments.

599 600 601 602 603 604 605 606 607 608 609

24 ACS Paragon Plus Environment

Environmental Science & Technology

(a)

0.8

(b) Pre-synthesis S-nZVI

Pre-synthesis S-nZVI Post-synthesis S-nZVI

0.6

C/C0

(d) CPS (a.u.)

1.0

Page 26 of 33

S22- S2-

Pre-synthesis S-nZVI

100 nm

(e)

0.4

(c) Post-synthesis S-nZVI

0.2

S22-

Post-synthesis S-nZVI SO42-

Sn2-

S2-

0.0

0

200 Time (min)

400 50 nm

169 167 165 163 161 159 Binding Energy (eV)

610 611 612 613 614

Figure 2. (a) TCE dechlorination by S-nZVI that receives sulfidation treatment at different stages of particle synthesis. (b) and (c) TEM micrographs of the particles used in (a). (d) and (e) XPS S 2p3/2 spectra of the corresponding particles in (a). All materials were prepared with a S/Fe mole ratio of 0.05.

615 616 617 618 619 620

25 ACS Paragon Plus Environment

Page 27 of 33

Environmental Science & Technology

621 622 623 624

Figure 3. TCE dechlorination and product formation in reactor of (a) S-nZVI with low sulfur dose, (b) S-nZVI with high sulfur dose, and (c) unmodified nZVI, and (d) FeS. The particle dose was 5 g/L.

625 626

26 ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 33

km (10-3 L/g-min)

(a) 1

0.1 nZVI 0.01 0.0001

0.01

1

S/Fe mole ratio 0.7

Product Yield

0.6 0.5

Acetylene (max yield) Ethene (final yield) ethane (final yield)

(b)

0.4 0.3 0.2 0.1

0.0 0.0001 627 628 629

0.01

1

S/Fe mole ratio Figure 4. Effect of S/Fe mole ratio on (a) TCE degradation rate and (b) product yields. The particle dose was 5 g/L.

630

27 ACS Paragon Plus Environment

Page 29 of 33

Environmental Science & Technology

Cl

H C

C

Cl

Cl

k1 H

k2 H H 631 632 633 634

C

Polymerization products

H

k3 H

C

C

C H

H H H C C H H H

Figure 5. Proposed reaction pathways of TCE decomposition on S-nZVI. Only experimentally observed intermediates or products are shown. Dashed line indicates possible involvement of multiple reaction steps. Values of reaction rate constants (k1 – k3) are tabulated in Table 2.

635

28 ACS Paragon Plus Environment

Environmental Science & Technology

H2 production, mmole

0.6

nZVI

0.5

Page 30 of 33

(a)

S-nZVI (S/Fe=0.01) S-nZVI (S/Fe=0.05)

0.4

S-nZVI (S/Fe=0.25)

0.3 0.2 0.1 0.0

0

24

48

72

Time, h

636

1.0

(b)

S-nZVI (S/Fe = 0.05)

As-nZVI (As/Fe = 0.05)

C/Co

0.8 0.6 0.4 0.2 0.0

0

200 Time (min)

400

637 638 639 640

Figure 6. (a) H2 production by nZVI and S-nZVI of varying S/Fe mole ratio. (b) TCE degradation by S-nZVI and arsenic-modified nZVI (As-nZVI) at an As or S to Fe mole ratio of 0.05. Particle dose was 5 g/L in all experiments.

641

29 ACS Paragon Plus Environment

Page 31 of 33

Environmental Science & Technology

Unmodified nZVI H+

H• H•

H•

e-

Fe(0)

H2

H+

C2HCl3 C2H2

H+

H• H•

e-

Sulfur-treated nZVI H+

C2HCl3 C2H2

H• H• H• H• H• Fe Sulfide e-

Fe(0)

H+

H2

H+

slow

H• H•

e-

642 643 644

Figure 7. Schematics of reactions on nZVI and S-nZVI.

645

30 ACS Paragon Plus Environment

Environmental Science & Technology

646 647

Table 1. Pseudo-first-order rate constants and product distribution of TCE dechlorination by nZVI and S-nZVI prepared under different conditions Particle type FeS

nZVI

S-nZVI (thiosulfate)

S-nZVI (dithionite)

S-nZVI (sulfide)

S-nZVI (thiosulfate)

648 649 650 651 652

Page 32 of 33

Sulfidation condition

Products

Yielda

Carbon recovery

km (10-3 L g-1 min-1 )c

(%)b Ethene Ethane Acetylene

N.D. N.D. 126

C3-C6

N.D.

Ethene Ethane

17 6.5 N.D. 5.7 44 10 4.5 7.2 33 8.2 10 9.6 39 9.4 7.2 7.5 45 10 1.3 8.1

Acetylene C3-C6 Ethene Post-synthesis sulfidation, S/Fe Ethane Acetylene = 0.05 C3-C6 Post-synthesis Ethene sulfidation, S/Fe Ethane = 0.05 Acetylene C3-C6 Post-synthesis Ethene sulfidation, S/Fe Ethane = 0.05 Acetylene C3-C6 Ethene Pre-synthesis sulfidation, S/Fe Ethane Acetylene = 0.05 C3-C6

107

0.0024 ± 0.0007

36

0.023± 0.008

68

0.80 ± 0.05

68

0.78 ± 0.21

67

1.04 ± 0.13

70

0.90 ± 0.08

a

calculated from product formation over TCE consumption. b sum of products (as C2 equivalents) and TCE remain. Both a and b were determined at the point of ca. 90% TCE conversion or, for slow reactions, the last sampling point. c mass-normalized pseudo-first-order rate constants, uncertainties represent 95% confidence intervals.

653

31 ACS Paragon Plus Environment

Page 33 of 33

654 655

Environmental Science & Technology

Table 2. Mass-normalized rate constants of steps involved in TCE degradation by S-nZVI of varying S/Fe mole ratio S/Fe mole ratio

656

k1 (10-3 L g-1 min-1)a

k2 (10-3 L g-1 min-1)b

k3(10-3 L g-1 min-1)b

0.00125 0.18 1.03 0.22 0.05 1.16 0.68 0.15 0.25 1.11 0.60 0.11 a obtained from TCE degradation experiments. b estimated from acetylene hydrogenation experiments.

657

32 ACS Paragon Plus Environment