Role of Sulfur Trioxide (SO3) in Gas-Phase Elemental Mercury

Feb 25, 2019 - As a drug, insulin is hard to swallow. Well, not literally. Oral insulin medications exist, and people... BUSINESS CONCENTRATES ...
0 downloads 0 Views 397KB Size
Subscriber access provided by ECU Libraries

Remediation and Control Technologies

Role of Sulfur Trioxide (SO3) in Gas-Phase Elemental Mercury Immobilization by Mineral Sulfide Zequn Yang, Hailong Li, Wenqi Qu, Mingguang Zhang, Yong Feng, Jiexia Zhao, Jianping Yang, and Kaimin Shih Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b07317 • Publication Date (Web): 25 Feb 2019 Downloaded from http://pubs.acs.org on March 4, 2019

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

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

Page 1 of 24

Environmental Science & Technology

2

Role of Sulfur Trioxide (SO3) in Gas-Phase Elemental Mercury Immobilization by Mineral Sulfide

3 4

Zequn Yanga, Hailong Lib*, Wenqi Qub, Mingguang Zhangb, Yong Fenga, Jiexia Zhaob, Jianping Yangb,

5

Kaimin Shiha**

1

6 7

a. Department of Civil Engineering, The University of Hong Kong, Hong Kong, Hong Kong SAR,

8

China

9

b. School of Energy Science and Engineering, Central South University, Changsha, 410083, China

10 11

Revision Submitted to

12

Environmental Science & Technology

13 14

*To whom correspondence should be addressed:

15

TEL: +86-18670016725

16

E-mail: [email protected]

17 18

**To whom correspondence should be addressed:

19

TEL: +852-2859-2973

20

Email: [email protected]

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 24

22

ABSTRACT: Mineral sulfide based sorbents were superior alternatives to traditional activated carbons

23

for elemental mercury (Hg0) immobilization in industrial flue gas. A systematical study concerning the

24

influence of sulfur trioxide (SO3) on Hg0 adsorption over a nano-sized copper sulfide (Nano-CuS) was

25

for the first time conducted. SO3 was found to significantly inhibit the Hg0 removal over Nano-CuS

26

partially because SO3 oxidized the reduced sulfur species (sulfide) with high affinity to mercury to its

27

oxidized sulfur species (sulfate). Moreover, a brand new “oxidation-reduction” mechanism that led to a

28

simultaneous oxidation of sulfide and reduction of mercury on the immobilized mercury sulfide (HgS)

29

was responsible for the inhibitory effect. Even though the released Hg0 from the reduction of mercury in

30

HgS could be oxidized by SO3 into its sulfate form (HgSO4) and re-captured by the sorbent, the

31

“oxidation-reduction” mechanism still compromised the Hg0 capture performance of the Nano-CuS

32

because HgSO4 deposited on the sorbent surface could be easily leached out when environmentally

33

exposed. These new insight into the role of SO3 in Hg0 capture over Nano-CuS can help to determine

34

possible solutions and facilitate the application of mineral sulfide sorbents as outstanding alternatives to

35

activated carbons for Hg0 immobilization in industrial flue gas.

36

Keywords: Mercury; Mineral sulfide; Sulfur trioxide; Adsorption; Flue gas

ACS Paragon Plus Environment

Page 3 of 24

38

Environmental Science & Technology

INTRODUCTION

39

Mercury is a notorious environmental hazard that attracts worldwide attention for decades owing to its

40

toxicity, persistence, transportability and bioaccumulation.1-5 On August 2017, the Minamata

41

Convention came into force to urge the implementation of ever rigorous standards among its 128

42

signatories for limiting mercury emission from anthropogenic sources.2, 6 Generally, mercury enriched

43

in fuels and raw materials emits in three forms from boiler/furnace after combustion/smelting, i.e.,

44

elemental mercury (Hg0), oxidized mercury (Hg2+) and particulate-bound mercury (Hgp).7-10 The Hg2+

45

and Hgp are easily to be removed from flue gas co-benefitted by the equipment of air pollution control

46

devices (APCDs) including desulfurization and particulate-control systems.11-13 On the contrary, Hg0 is

47

extremely difficult to be degraded due to its high volatility and insolubility in water. Thus, Hg0 persists

48

in flue gas and acts as the main mercury species discharged from stack to immediate environment.14-16

49

Reduced sulfur species (S0/S2-) is an ideal active candidate to treat Hg0 polluted flue gas attributed to

50

the relatively high affinity between S0/S2- and Hg0.17 Activated carbon (AC) impregnated by element

51

sulfur exhibited significantly enhanced Hg0 capture capacity (2.3 mg/g) compared to that of bare AC.18,

52

19

53

renowned for its extreme stability when exposed to complex environment.20,

54

sulfur still showed relatively limited adsorption capacity to Hg0, which caused exceedingly high

55

operation cost of the activated carbon injection (ACI) strategy.22-24 The adoption of carbonaceous

56

material also introduced overwhelming carbon content in fly ash that impeded its reusability as raw

57

concrete materials.17 To overcome these drawbacks, nano-sized mineral sulfides were recently

58

developed and demonstrated to establish outstanding performance for Hg0 immobilization in flue gas,25-

59

30

60

with Hg0 adsorption capacity reaching as high as 122.4 mg/g. Moreover, the mineral sulfides exhibited

61

neutral even beneficial effect on the reusability of fly ash as concrete feed.29 With these advantages,

62

mineral sulfides were recognized as promising alternatives to traditional ACs for Hg0 pollution control

63

in various industrial processes.

Moreover, the existence of sulfur species warrants the conversion of Hg0 into mercury sulfide (HgS), 21

However, elemental

among which the nano-sized copper sulfide (Nano-CuS) exhibited the best adsorption performance,

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 24

64

Industrial flue gas contains complicated gas compositions like hydrogen chloride (HCl), water vapor

65

(H2O), nitrogen oxides (NOx) and sulfur oxides (SOx). They may exert different impacts on Hg0

66

adsorption over mineral sulfides with mechanisms complex and varied. For instance, NOx inhibited the

67

Hg0 adsorption over nano-sized zinc sulfide mainly because NOx oxidized the sorbent surface.25 Gas-

68

phase H2O suppressed the adsorption performance of Nano-CuS due to both of the active sites

69

prevention and competitive adsorption effects.22 SO2 was found to have neutral effect on Hg0 adsorption

70

over mineral sulfides,29 which further justified the candidature of the mineral sulfides as efficient Hg0

71

trap, since SO2 widely exists in combustion/smelting flue gas,8, 31 and its concentration is relatively high

72

for many cases. In non-ferrous smelting flue gas, the content of SO2 generally exceeds 2.0 vt%

73

attributed to the burning of sulfide ores.32 In oxy-fuel combustion technique, which was regarded as a

74

promising alternative to traditional air combustion strategy due to its significance in reducing carbon

75

dioxide (CO2) emission, high-concentration of SO2 (2-2.5 vt%) accumulates in flue gas due to the flue

76

gas recycle.33

77

Besides SO2, depending on the burning factors and feed materials, sulfur trioxide (SO3) is another

78

form of sulfur species partitioned from fuels and raw materials after combustion/smelting and

79

concentrated in flue gas.8,

80

smelting results in both high concentration of SO2 and SO3 in flue gas.8 On the other hand, for coal

81

combustion, the selective catalytic reduction (SCR) system partially oxidized SO2 into SO3.34 This

82

makes the SO3 concentration at the inlet of the electrostatic precipitator (ESP), where is an ideal site for

83

sorbent injection, doubled compared to that at the inlet of SCR system.36 Oxy-fuel combustion flue gas

84

even carries pronouncedly high content of catalyzed SO3 attributed to the abundance of SO2.33, 37, 38 SO3

85

was found to suppress the Hg0 removal over bare AC due to the competitive adsorption and depletion of

86

surface active sites even at very low concentration (20 ppm).39, 40 Moreover, the SO3, with extremely

87

strong oxidizability,41 is highly probable to oxidize the reduced sulfur species to its oxidized species that

88

exhibits less affinity to Hg0. As suggested by these perspectives, it is of considerable importance to

89

conduct a systematical research to investigate the effect of SO3 on Hg0 removal over sulfur based

34, 35

High content of sulfur species in sulfide ores used for non-ferrous

ACS Paragon Plus Environment

Page 5 of 24

Environmental Science & Technology

90

materials, especially for mineral sulfides, where the sulfur species in sorbent and adsorbate are both in

91

their reduced forms.

92

Herein, a systematical study concerning with the influence of SO3 on Hg0 adsorption over mineral

93

sulfide based materials is for the first time reported in this work with the mechanisms discussed in detail.

94

Nano-CuS with the best performance among various mineral sulfides was adopted to represent the

95

mineral sulfide sorbents. This work provides an in-depth comprehension of the interaction between the

96

SO3, mineral sulfide sorbent and the product after Hg0 adsorption, which is critical for designing novel

97

gas phase Hg0 sorbents and developing optimal sorbent applications in a real-world condition.

98

EXPERIMENTAL AND COMPUTATIONAL METHODS

99

Sorbent Preparation. Please refer to the Supporting Information, the Sorbent preparation section.

100

Experimental setup and methods. The Hg0 adsorption over Nano-CuS was tested in a fixed-bed

101

reaction system (shown in Figure S1). Compressed gas cylinders containing nitrogen (N2), oxygen (O2)

102

and SO2 were adopted to generate desired gas atmospheres. SO3 was supplied by oxidizing SO2 over a

103

vanadium-titanium oxides based catalyst (VTBC) put in a quartz tube prior to the Hg0 removal reactor,

104

for which the concentration was adjusted by the inlet SO2 concentration with the conversion efficiency

105

of SO2 kept at 100%. Mass flow controllers was adopted to precisely control the total flow rate of 1 L

106

min-1. A constant feed of Hg0 was provided by a Dynacal Hg0 permeation device (VICI Metronics)

107

heated to an unchanged temperature. The Hg0 removal was conducted in a borosilicate glass reactor

108

with 10 mm inner diameter that was placed in a tubular furnace to control the reaction temperature. A

109

mercury analyzer (VM3000, Mercury Instruments, Inc.) was used to record the mercury concentration.

110

Before each test, the Hg0 carried by different flue gas atmospheres bypassed the reactor to record the

111

inlet Hg0 concentration (Cin). Then, the Hg0 feed passed through the reactor and the as-obtained Hg0

112

concentration was denoted as outlet Hg0 concentration (Cout). Hence, the Hg0 adsorption capacity (C, μg

113

Hg0/g sorbent) was calculated by equation (1):

114

C

1 t2 (Cin  Cout )  f  dt m t1

ACS Paragon Plus Environment

(1)

Environmental Science & Technology

Page 6 of 24

115

in which m is the mass of the sorbent (g), f is the total gas flow rate (m3/h), and t is the duration time of

116

the adsorption processes (h).

117

Four sets of experiments were conducted with parameters detailed in Table S1. Set I experiments

118

were designed to investigate the Hg0 removal over Nano-CuS under different gas atmospheres. In set II

119

experiments, the competitive adsorption between SO3 and Hg0 for active sites was studied. The fresh

120

Nano-CuS was pretreated under 500 μg/m3 Hg0 for 1 h to be the Hg-laden Nano-CuS. The pure HgS

121

was directly purchased from Aladdin Co. Ltd without any purification. In set III experiments, the Nano-

122

CuS pretreated under 4% O2 plus 1000 ppm SO2 in the presence or absence of 500 ppm SO3 carried by

123

pure N2 for 2 h was used for Hg0 removal to investigate the possible surface active sites damage effects

124

by SO3. Set IV experiments was designed to explore the possible reduction of HgS by the mixture of

125

SO2 and SO3. The fresh Nano-CuS was pretreated by 500 μg/m3 until equilibrium to be the Hg-saturated

126

Nano-CuS. In all five sets of the experiments, the adsorption temperature was fixed at 75 oC, the optimal

127

operation temperature of Nano-CuS,29 with the inlet Hg0 concentration kept as 100 μg/m3. The dosage

128

of Nano-CuS was 5 mg, much lower to that in a real-world condition, to magnify the influence of

129

different gas components.

130 131 132 133

Sorbent characterizations. Please refer to the Supporting Information, Sorbent characterization section. Temperature programmed desorption/decomposition (TPD). Please refer to the Supporting Information, the Temperature programmed desorption/decomposition section.

134

First principle calculation. The CuS model was in hexagonal structure with space group P63/mmc

135

(a=b=3.796 Å, c=16.360 Å, α=β =90°, γ=120°). An 8-layer slab with a (3×3) unit cell based on a

136

(3×3×1) primitive cell was used to model the structure of CuS(001) surface. A 15 Å vacuum region

137

between the slabs was constructed to avoid the spurious interactions. During the geometric optimization,

138

the bottom 3 layers were fixed and the rest were allowed for relaxation. All calculations were conducted

139

with the quantum mechanics based Dmol3 program package in Materials Studio 8.0. The exchange

140

correlation potential was determined by Perdew-Burke-Ernzerhoff (PBE) approximation in the place of ACS Paragon Plus Environment

Page 7 of 24

Environmental Science & Technology

141

the generalized gradient approximation (GGA) scheme.42 The interaction between valance electrons,

142

inner electrons and atomic nucleus was performed with the double numerical basis sets plus polarization

143

functional (DNP). The Monkhorst-Pack scheme k-points grid of 2×2×1 was used to simplify the

144

Brillouin zone in the cell of CuS (001) and the real space basis set functions are adjusted to be 4.0 Å.

145

The criteria for the tolerances of energy, force, displacement, and SCF convergence criteria are set as

146

2×10-5 Ha, 4×10-3 Ha Å-1, 5×10-3 Å, and 10-5, respectively. A Methfessel-Paxton smearing of 0.005 Ha

147

was used to improve calculation performance. Gas-phase species (Hg0 and SO3) molecules were also

148

optimized separately in a large crystal cell of 10×10×10 Å.

149 150

Mercury leaching test. Please refer to the Supporting Information, the Mercury leaching test section. RESULTS AND DISCUSSIONS

151

SO3 inhibited the Hg0 adsorption over Nano-CuS. As shown in Figure 1(a), the as-prepared Nano-

152

CuS established outstanding performance for Hg0 capture under pure N2, showing almost 100% Hg0

153

removal efficiency within a 4-h experiment. The adsorption capacity reaching 28.3 mg/g is the highest

154

value among various mineral sulfides developed yet for Hg0 decontamination from industrial flue gas.17,

155

26-28

156

over Nano-CuS, which exhibited 4-h uptake capacities of 28.4 and 28.1 mg/g, respectively. The

157

excellent resistance of Nano-CuS towards SO2 poisoning makes it a promising candidate for Hg0

158

adsorption from industrial flue gas containing high concentration of SO2. However, SO3 showed adverse

159

impact on Hg0 adsorption over Nano-CuS (shown in Figure 1(a)). The outlet Hg0 concentration climbed

160

to 17.6 μg/m3 at the end of a 4-h experiments when 100 ppm SO3 was added to the pure N2 gas flow.

161

The as-derived adsorption capacity was slightly suppressed from 28.3 mg/g under pure N2 to 26.1 mg/g

162

under 100 ppm SO3. When the addition of SO3 was increased from 100 to 500 ppm, a significant

163

inhibitory effect was observed with the outlet Hg0 concentration sharply increased within 4 h and

164

reached 86.7 μg/m3 at the end. This indicates that the Nano-CuS almost lost its activity for Hg0 capture

165

after in contact with 500 ppm SO3 for 4 h. Also, under this condition, the Hg0 adsorption capacity of

The additions of 4% O2 and 4% O2 plus 1000 ppm SO2 showed negligible effect on Hg0 removal

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 24

166

Nano-CuS within 4 h pronouncedly decreased to 16.8 mg/g compared to that of 28.3 mg/g under pure

167

N2.

168

It should be noticed that in a coal combustion case, Nano-CuS is a promising alternative to traditional

169

AC to be injected through the pipeline during the flue gas treatment process, which means that the

170

contacting time between the sorbent and flue gas is short. The 30 min timespan is a very important

171

parameter to evaluate the performance of the sorbent in a real-world condition because the residence

172

time of the injected sorbent that interacts with mercury is about 3-5 s in the case of an electrostatic

173

precipitator and approximately 25 min in the case of a fabric filter.25 As shown in Figure 1(a), the

174

adsorption performance of Nano-CuS was just negligibly influenced within 30 min when 100 ppm SO3

175

was present. The concentration of 100 ppm was even higher than the highest SO3 concentration in a real

176

coal combustion flue gas.36 Thus, even though Hg0 capture over Nano-CuS was limited by the presence

177

of SO3, the results still suggest that Nano-CuS is a promising candidate to be used in an injection

178

strategy for Hg0 immobilization in coal combustion flue gas. However, for a fix-bed strategy that was

179

generally applied to realize a better economical friendliness, the long-exposure to SO3 may cause

180

accumulating poisoning effect on the sorbent. To avoid the sorbent deactivation and discuss the possible

181

solutions in such cases, investigating the involved mechanisms for the inhibitive effect of SO3 is

182

indispensable.

183

SO3 repelled negligible Hg0 from the sorbent surface. Competitive adsorption is a renowned reason

184

accounted for the prohibitive influence of flue gas components on the performance of Hg0 traps. For

185

example, SO2 was found to deactivate metal oxide for Hg0 adsorption by repelling Hg0 from the metal

186

oxide surface.43 Gas-phase H2O suppressed the Hg0 uptake capacity of sulfur-impregnated AC by as

187

much as 25% due to the occupation of surface active sites.44 Moreover, SO3 was also proven to compete

188

for the adsorption sites with Hg0 on traditional AC.40 For mineral sulfides, although SO2 was

189

demonstrated to show no competitive effect with Hg0,29 the role of SO3 for this perspective remains

190

unknown. Previous study reported that a very low concentration of SO3 showed significant adverse

191

influence on Hg0 removal over AC due to competitive adsorption,40 which indicates that SO3 may ACS Paragon Plus Environment

Page 9 of 24

Environmental Science & Technology

192

exhibit stronger capability for competition than SO2. If this is the same case for Nano-CuS, the weakly

193

adsorbed Hg0 on Hg-laden Nano-CuS will be repelled out by the addition of SO3. However, similar to

194

that for SO2, 500 ppm SO3 repelled no Hg0 from the Hg-laden Nano-CuS surface (shown in Figure S2).

195

This evidences that the competitive adsorption negligibly accounted for the deactivation of Nano-CuS

196

for Hg0 capture in the presence of SO3, and suggests that the mercury over mineral sulfides surface was

197

primarily anchored by sulfide species to form HgS chemisorbed on the sorbent surface with negligible

198

weakly adsorbed mercury exists. On the contrary, for traditional AC, considerable amount of mercury

199

was physiosorbed on the sorbent surface and easily to be repelled out when SO3 was added.45

200

SO3 oxidized sulfide on Nano-CuS into sulfate. As shown in Figure 2, the Nano-CuS pretreated by

201

4% O2 plus 1000 ppm SO2 showed unchanged Hg0 removal performance in a 4-h test compared with the

202

fresh sample. However, when pretreated in the extra presence of 500 ppm SO3, the Hg0 removal

203

performance of the pretreated Nano-CuS was significantly suppressed. The outlet Hg0 concentration

204

rapidly increased during the experiments and reached 50.6 μg/m3 at the end of 4 h. This gives us strong

205

hints that the sorbent surface was permanently deactivated by SO3. Considering the oxidizability of SO3

206

and the reduced nature of sulfur species on Nano-CuS, it is highly probable that the sulfide was oxidized

207

to high-valence sulfur species that exhibits mediocre affinity towards Hg0.

208

To further identify the as-formed sulfur species over Nano-CuS after pretreated by SO3, X-ray

209

photoelectron spectroscopy (XPS) analysis was conducted and the results are shown in Figure 3. The S

210

2p peaks for fresh and spent Nano-CuS centering at 162.2, 163.5 and 164.4 eV were assigned to the

211

characteristic peaks of monosulfide, S-S dimer and polysulfide, respectively (Figure 3(a)).29 The spike

212

locating at 168-172 eV was due to the presence of sulfite (SO32-) and/or sulfate (SO42-) species.46 For the

213

fresh sample, sulfur primarily exists as sulfide with no impurity like elemental sulfur detected, which

214

indicates that the as-prepared Nano-CuS was almost in its pure phase. The existence of minority of

215

oxidized sulfur species was probably due to the residual of CuSO4 precursor. For the fresh Nano-CuS,

216

the surface ratio between the reduced sulfurs species and oxidized sulfur species was derived from the

217

peak integration to be around 9:1. However, after pretreated by SO3, the amount of SO42- and/or SO32ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 24

218

significantly increased compared with that on fresh Nano-CuS and accompanied by the corresponding

219

peak intensity decline of the sulfide species. For the pretreated sample, the surface ratio between the

220

sulfide species and SO42-/SO32- decreased to 5:4. These results suggest that SO3 efficiently converted the

221

reduced sulfur forms on Nano-CuS to oxidized sulfur forms (SO42-/SO32-) that is inert for Hg0 removal.

222

The oxidation of the Nano-CuS sorbent was also evidenced by the O 1s spectrum shown in Figure 3(b).

223

The O 1s peak for the pretreated Nano-CuS was notably intensified compared to that for the fresh Nano-

224

CuS, which is in agreements with SO42- and/or SO32- formation on the sorbent surface. The binding

225

energy difference occurred between different copper species may be resulted from the coordination and

226

charge changes when copper was bonded to different functional groups.47, 48 In the Cu 2p pattern (shown

227

in Figure 3(c)), the peak centering at 932.4 eV was attributed to the Cu2+ on CuS while the peak locating

228

from 934-936 eV was due to the presence of CuSO4.49 For the fresh Nano-CuS, the ratio between the

229

CuS and CuSO4 was 9:1. For the pretreated CuS, the ratio between CuS and CuSO4 decreased to 5:4.

230

These results are in line with the S 2p spectra and indicate that sulfide species in Nano-CuS has been

231

oxidized to its highest valence, i.e., SO42-, instead of SO32-, by SO3.

232

SO2-TPD is an effective method to identify the as-formed sulfur compounds by their characteristic

233

decomposition temperature. As shown in Figure 4, the Nano-CuS pretreated by 500 ppm SO3 only

234

exhibited one decomposition peak to release SO2 from 50 to 750 oC, which centers around 680 oC. The

235

characteristic peak indexes to the decomposition of CuSO4,7 which further proves that CuSO4 is the

236

final destination of the sulfide species on Nano-CuS after pretreated by SO3. The inserted Fourier

237

transform-infrared (FTIR) spectra show two spikes in the wavenumber range from 800-2400 cm-1, of

238

which the peak locating at 1150 cm-1 was resulted from the existence of SO42- while the peak from

239

1600-1800 cm-1 was assigned to the surface water molecule.25, 50 No separate peak centering at 950 cm-1

240

and belonging to SO32- was detected. These results demonstrate that SO3 with strong oxidizability

241

directly oxidized the Nano-CuS sorbent into CuSO4 with no intermediate observed. The as-formed

242

sulfate irreversibly damaged the Hg0 removal performance of the Nano-CuS due to the inhibition of

243

Hg0 adsorption by sulfate.51 ACS Paragon Plus Environment

Page 11 of 24

Environmental Science & Technology

244

SO3 compromised the stability of adsorbate (HgS) over Nano-CuS. As shown by Equation (1),

245

SO3 has the potential to oxidize sulfide in HgS and simultaneously reduce oxidized mercury into its

246

elemental form. HgS + 2SO3 (g) → Hg0 (g) + 3SO2 (g) ΔG (75 oC) = -90.2 kJ/mol

247

(1)

248

The negative Gibbs free energy indicates that the reaction is thermodynamically spontaneous under 75

249

oC.

250

when SO3 passed through. This seems “contradictory” with the thermodynamic results that suggests

251

reaction (1) definitely processes to form Hg0. To reveal the microcosmic mechanisms behind the

252

“contradiction”, the first principle calculation based on density functional theory (DFT) was conducted

253

to illustrate the microcosmic role of SO3 in reducing HgS into Hg0 and compromising the adsorption

254

capacity.

However, as shown in Figure S3, no Hg0 was detected downstream of the Hg-saturated Nano-CuS

255

Figure 5(a) shows a stable structure of Hg0 immobilized by a monosulfide over CuS surface in the

256

absence of SO3 molecule. The bond length of the as-formed HgS was 2.47 Å, close to the experimental

257

value of the bond length of HgS (2.58 Å).52 The Mulliken charge of the adsorbed mercury was derived

258

to be 0.15 e, indicating the electron loss of Hg0 after captured by CuS. The partial density of states

259

(PDOS) analysis of mercury and the monosulfide (shown in Figure S4(a)) further evidenced the

260

interaction between them. The overlap of the peaks between -5 to -6 eV indicates the strong interaction

261

between the d orbital of mercury and the s and p orbitals of the monosulfide. Moreover, the interaction

262

also introduced a weak orbital hybridization between -10 to -12 eV. These results demonstrate that

263

mercury was stably immobilized by the monosulfide to form HgS. Then, a SO3 molecule was used to

264

attack the stable structure, and the final products after geometric optimization was shown in Figure 5(b).

265

After the geometric optimization, one O atom obviously dissociated from SO3 and recombined with the

266

monosulfide dissociated from the sorbent surface to form an O-S-Hg intermediate. The left two O atoms

267

and one S atom was repelled away to form an O-S-O structure with the bond angle of 117.7o and bond

268

length of 1.48 and 1.47 Å, respectively. These parameters are in line with the experimental results

269

showed that the bond length of SO2 is 1.43 Å and the bond angle between O-S-O is 118o.53 The bond ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 24

270

length of Hg-S in the O-S-Hg intermediate was slightly prolonged to 2.49 Å and Mulliken charge of the

271

adsorbed mercury was decreased by 0.01 e to be 0.14 e, which suggests a non-obvious reduction of the

272

mercury after attacked by one SO3 molecule. The reaction process after one HgS was attacked by one

273

SO3 can be derived as followed: HgS + SO3 (g) → Hg-S-O + SO2 (g)

274

(2)

275

However, after the HgS was attacked by two SO3 molecules (shown in Figure 5(c)), the distance

276

between the mercury and the initial monosulfide was increased obviously to 3.27 Å. The PDOS analysis

277

of mercury and the initial monosulfide dissociated from the CuS surface shows that there was no

278

interaction between them (Figure S4(b)). These results indicate a full dissociation of the HgS and the

279

formation of Hg0. The monosulfide is fully oxidized by SO3 into a SO2 molecule and the adsorbed

280

mercury was reduced to its elemental form. The overall reaction after the HgS was attacked by another

281

SO3 can be expressed as:

282

Hg-S-O + SO3 (g) → Hg0 + 2SO2 (g)

283

Reaction (1) is the sum of reaction (2) and (3). The initial monosulfide expected to react with Hg0 was

284

fully oxidized into SO2 and co-existed with another two SO2 molecules due to the dissociation of SO3

285

(shown in Figure 5(d)), which molecularly prove that the adsorbed mercury can be successfully

286

transformed by SO3 to Hg0 over CuS surface. However, these results still disagree with our

287

experimental results that shows no Hg0 was detected when the Hg-saturated Nano-CuS was attacked by

288

SO3.

(3)

289

As SO3 can oxidize Hg0 into HgSO4,41 it is highly probable that, even though the released Hg0

290

successfully escaped from the sorbent, the abundant SO3 will oxidize it into its sulfate salt and deposit

291

on the sorbent surface. The Hg-TPD results of Nano-Cus pretreated under different conditions

292

evidenced this point (shown in Figure 6(a)). An extra characteristic peak locating at 550 oC appeared

293

after pretreated in the presence of SO3, which indexed to the decomposition peak of HgSO4.54 The

294

extremely high decomposition temperature of mercury sulfate (HgSO4) makes it stable in an air

295

condition. However, the solubility of HgSO4 would result in its extensively environmental leaching as ACS Paragon Plus Environment

Page 13 of 24

Environmental Science & Technology

296

compared to HgS. As shown in Figure 6(b), the Nano-CuS pretreated in the co-presence of Hg0 and SO3

297

showed more than 200 folds higher leachability in simulated environmental condition as compared to

298

that pretreated in the absence of SO3. The presence of SO3 in gas flow inhibits the re-emission of the

299

reduced Hg0 into air, but its re-emission into aqueous environment is inevitable. The non-observation of

300

Hg0 when the Hg-saturated Nano-CuS was attacked by SO3 was mainly attributed to the existence of

301

SO3 transformed the escaped Hg0 into a water-soluble HgSO4.

302

In sum, SO3 in high concentration was found to significantly inhibited the Hg0 adsorption over Nano-

303

CuS because the sulfation of Nano-CuS surface and the transformation of immobilized adsorbate. This

304

would not compromise the candidature of mineral sulfides as alternative to activated carbon for Hg0

305

removal from coal combustion process because the SO3 concentration is relatively low and the contact

306

time is extremely short. However, for fixed-bed Hg0 removal applications in other scenarios like

307

nonferrous industry flue gas containing relatively high-concentration of both Hg0 and SO3, a feasible

308

solution is worthy to be discussed further to abate the detrimental effect of SO3, like adding sacrifice

309

agent to react with SO3 and protect the sorbents. The results obtained in this study will provide valuable

310

reference for optimizing the design of mineral sulfide based sorbents in the future that can be applied for

311

Hg0 decontamination from various industrial processes.

312

ASSOCIATED CONTENT

313

Supporting

Information.

Experimental

details

regarding

sorbent

preparation,

sorbent

314

characterization, mercury temperature programmed desorption/decomposition, mercury leaching test,

315

Table S1, Figure S1-S4. This material is available free of charge via the Internet at http://pubs.acs.org.

316

AUTHOR INFORMATION

317

Corresponding Author

318

*TEL: +86-18670016725; E-mail: [email protected]

319

**TEL: +852-2859-2973; Email: [email protected]

320

ORCID

321

Hailong Li: 0000-0003-0652-6655 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 24

322

Kaimin Shih: 0000-0002-6461-3207

323

ACKNOWLEDGEMENTS

324

This project was supported by the National Natural Science Foundation of China (51776227), Natural

325

Science Foundation of Hunan Province, China (2018JJ1039, 2018JJ3675), and the Research Council of

326

Hong Kong (17257616, C7044-14G, and T21-771/16R).

327

References

328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372

1. Pacyna, E. G.; Pacyna, J. M.; Steenhuisen, F.; Wilson, S., Global anthropogenic mercury emission inventory for 2000. Atmos. Environ. 2006, 40, (22), 4048-4063. 2. Kong, L.; Zou, S.; Mei, J.; Geng, Y.; Zhao, H.; Yang, S., Outstanding Resistance of H2S-Modified Cu/TiO2 to SO2 for Capturing Gaseous Hg0 from Nonferrous Metal Smelting Flue Gas: Performance and Reaction Mechanism. Environ. Sci. Technol. 2018, 52, (17), 10003-10010. 3. Xu, H.; Qu, Z.; Zong, C.; Huang, W.; Quan, F.; Yan, N., MnOx/Graphene for the Catalytic Oxidation and Adsorption of Elemental Mercury. Environ. Sci. Technol. 2015, 49, (11), 6823-6830. 4. Chen, W.; Pei, Y.; Huang, W.; Qu, Z.; Hu, X.; Yan, N., Novel Effective Catalyst for Elemental Mercury Removal from Coal-Fired Flue Gas and the Mechanism Investigation. Environ. Sci. Technol. 2016, 50, (5), 2564-2572. 5. Chen, J.; Li, C.; Li, S.; Lu, P.; Gao, L.; Du, X.; Yi, Y., Simultaneous removal of HCHO and elemental mercury from flue gas over Co-Ce oxides supported on activated coke impregnated by sulfuric acid. Chem. Eng. J 2018, 338, 358-368. 6. Gao, L.; Li, C.; Zhang, J.; Du, X.; Li, S.; Zeng, J.; Yi, Y.; Zeng, G., Simultaneous removal of NO and Hg0 from simulated flue gas over CoOx-CeO2 loaded biomass activated carbon derived from maize straw at low temperatures. Chem. Eng. J 2018, 342, 339-349. 7. Yang, Z.; Li, H.; Liu, X.; Li, P.; Yang, J.; Lee, P.H.; Shih, K., Promotional effect of CuO loading on the catalytic activity and SO2 resistance of MnOx/TiO2 catalyst for simultaneous NO reduction and Hg0 oxidation. Fuel 2018, 227, 79-88. 8. Liu, W.; Xu, H.; Liao, Y.; Quan, Z.; Li, S.; Zhao, S.; Qu, Z.; Yan, N., Recyclable CuS sorbent with large mercury adsorption capacity in the presence of SO2 from non-ferrous metal smelting flue gas. Fuel 2019, 235, 847-854. 9. Li, H.; Zhang, W.; Wang, J.; Yang, Z.; Li, L.; Shih, K., Coexistence of enhanced Hg0 oxidation and induced Hg2+ reduction on CuO/TiO2 catalyst in the presence of NO and NH3. Chem. Eng. J. 2017, 330, 1248-1254. 10. Li, G.; Wang, S.; Wu, Q.; Wang, F.; Ding, D.; Shen, B., Mechanism identification of temperature influence on mercury adsorption capacity of different halides modified bio-chars. Chem. Eng. J 2017, 315, 251-261. 11. Yang, J.; Zhao, Y.; Zhang, J.; Zheng, C., Regenerable Cobalt Oxide Loaded Magnetosphere Catalyst from Fly Ash for Mercury Removal in Coal Combustion Flue Gas. Environ. Sci. Technol. 2014, 48, (24), 14837-14843. 12. Li, H.; Zhang, W.; Wang, J.; Yang, Z.; Li, L.; Shih, K., Copper slag as a catalyst for mercury oxidation in coal combustion flue gas. Waste Manage. 2018, 74, 253-259. 13. Liu, X.; Jiang, S.; Li, H.; Yang, J.; Yang, Z.; Zhao, J.; Peng, H.; Shih, K., Elemental mercury oxidation over manganese oxide octahedral molecular sieve catalyst at low flue gas temperature. Chem. Eng. J 2019, 356, 142-150. 14. Zhang, L.; Zhuo, Y.; Chen, L.; Xu, X.; Chen, C., Mercury emissions from six coal-fired power plants in China. Fuel Process. Technol. 2008, 89, (11), 1033-1040. 15. Li, H.; Zhao, J.; Zhang, W.; Yang, J.; Wang, J.; Zhang, M.; Yang, Z.; Li, L.; Shih, K., NH3 inhibits mercury oxidation over low-temperature MnOx/TiO2 SCR catalyst. Fuel Process. Technol. 2018, 176, 124-130. 16. Shen, B.; Zhu, S.; Zhang, X.; Chi, G.; Patel, D.; Si, M.; Wu, C., Simultaneous removal of NO and Hg0 using Fe and Co co-doped Mn-Ce/TiO2 catalysts. Fuel 2018, 224, 241-249. 17. Li, H.; Zhu, L.; Wang, J.; Li, L.; Shih, K., Development of Nano-Sulfide Sorbent for Efficient Removal of Elemental Mercury from Coal Combustion Fuel Gas. Environ. Sci. Technol. 2016, 50, (17), 9551-9557. 18. Korpiel, J. A.; Vidic, R. D., Effect of Sulfur Impregnation Method on Activated Carbon Uptake of Gas-Phase Mercury. Environ. Sci. Technol. 1997, 31, (8), 2319-2325. 19. Liu, W.; Vidić, R. D.; Brown, T. D., Optimization of Sulfur Impregnation Protocol for Fixed-Bed Application of Activated Carbon-Based Sorbents for Gas-Phase Mercury Removal. Environ. Sci. Technol. 1998, 32, (4), 531-538. 20. Zhao, J.; Li, H.; Yang, Z.; Zhu, L.; Zhang, M.; Feng, Y.; Qu, W.; Yang, J.; Shih, K., Dual Roles of Nano-Sulfide in Efficient Removal of Elemental Mercury from Coal Combustion Flue Gas within a Wide Temperature Range. Environ. Sci. Technol. 2018, 52, (21), 12926-12933. 21. Li, H.; Zhang, M.; Zhu, L.; Yang, J., Stability of mercury on a novel mineral sulfide sorbent used for efficient mercury removal from coal combustion flue gas. Environ. Sci. Pollut. Res. 2018, 25, (28), 28583-28593.

ACS Paragon Plus Environment

Page 15 of 24

373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432

Environmental Science & Technology

22. Yang, Z.; Li, H.; Liao, C.; Zhao, J.; Feng, S.; Li, P.; Liu, X.; Yang, J.; Shih, K., Magnetic Rattle-Type Fe3O4@CuS Nanoparticles as Recyclable Sorbents for Mercury Capture from Coal Combustion Flue Gas. ACS Appl. Nano Mater. 2018, 1, (9), 4726-4736. 23. Zhou, Q.; Duan, Y.; Chen, M.; Liu, M.; Lu, P.; Zhao, S., Effect of flue gas component and ash composition on elemental mercury oxidation/adsorption by NH4Br modified fly ash. Chem. Eng. J 2018, 345, 578-585. 24. Li, N.; Wei, H.; Duan, Y.; Tang, H.; Zhao, S.; Hu, P.; Ren, S., Experimental Study on Mercury Adsorption and Adsorbent Regeneration of Sulfur-Loaded Activated Carbon. Energy & Fuels 2018, 32, (10), 11023-11029. 25. Li, H.; Zhu, L.; Wang, J.; Li, L.; Lee, P.-H.; Feng, Y.; Shih, K., Effect of Nitrogen Oxides on Elemental Mercury Removal by Nanosized Mineral Sulfide. Environ. Sci. Technol. 2017, 51, (15), 8530-8536. 26. Zhao, H.; Yang, G.; Gao, X.; Pang, C. H.; Kingman, S. W.; Wu, T., Hg0 Capture over CoMoS/γ-Al2O3 with MoS2 Nanosheets at Low Temperatures. Environ. Sci. Technol. 2016, 50, (2), 1056-1064. 27. Li, H.; Zhu, W.; Yang, J.; Zhang, M.; Zhao, J.; Qu, W., Sulfur abundant S/FeS2 for efficient removal of mercury from coal-fired power plants. Fuel 2018, 232, 476-484. 28. Liao, Y.; Chen, D.; Zou, S.; Xiong, S.; Xiao, X.; Dang, H.; Chen, T.; Yang, S., Recyclable Naturally Derived Magnetic Pyrrhotite for Elemental Mercury Recovery from Flue Gas. Environ. Sci. Technol. 2016, 50, (19), 10562-10569. 29. Yang, Z.; Li, H.; Feng, S.; Li, P.; Liao, C.; Liu, X.; Zhao, J.; Yang, J.; Lee, P.-H.; Shih, K., Multiform Sulfur Adsorption Centers and Copper-Terminated Active Sites of Nano-CuS for Efficient Elemental Mercury Capture from Coal Combustion Flue Gas. Langmuir 2018, 34, (30), 8739-8749. 30. Zhao, H.; Mu, X.; Yang, G.; George, M.; Cao, P.; Fanady, B.; Rong, S.; Gao, X.; Wu, T., Graphene-like MoS2 containing adsorbents for Hg0 capture at coal-fired power plants. Appl. Energy 2017, 207, 254-264. 31. Yang, S.; Guo, Y.; Yan, N.; Wu, D.; He, H.; Xie, J.; Qu, Z.; Jia, J., Remarkable effect of the incorporation of titanium on the catalytic activity and SO2 poisoning resistance of magnetic Mn-Fe spinel for elemental mercury capture. Appl. Catal. B: Environ. 2011, 101, (3), 698-708. 32. Wu, Q.; Wang, S.; Hui, M.; Wang, F.; Zhang, L.; Duan, L.; Luo, Y., New Insight into Atmospheric Mercury Emissions from Zinc Smelters Using Mass Flow Analysis. Environ. Sci. Technol. 2015, 49, (6), 3532-3539. 33. Spörl, R.; Maier, J.; Belo, L.; Shah, K.; Stanger, R.; Wall, T.; Scheffknecht, G., Mercury and SO3 Emissions in Oxy-fuel Combustion. Energy Procedia 2014, 63, 386-402. 34. Srivastava, R. K.; Miller, C. A.; Erickson, C.; Jambhekar, R., Emissions of Sulfur Trioxide from Coal-Fired Power Plants. Journal of the Air Waste Manage. Associa. 2004, 54, (6), 750-762. 35. Mitsui, Y.; Imada, N.; Kikkawa, H.; Katagawa, A., Study of Hg and SO3 behavior in flue gas of oxy-fuel combustion system. Inter. J. Greenhouse Gas Control 2011, 5, S143-S150. 36. Krishnakumar, B.; Niksa, S., Predicting the impact of SO3 on mercury removal by carbon sorbents. Proceed. Combust. Ins. 2011, 33, (2), 2779-2785. 37. Yang, J.; Zhao, Y.; Chang, L.; Zhang, J.; Zheng, C., Mercury Adsorption and Oxidation over Cobalt Oxide Loaded Magnetospheres Catalyst from Fly Ash in Oxyfuel Combustion Flue Gas. Environ. Sci. Technol. 2015, 49, (13), 8210-8218. 38. Fleig, D.; Andersson, K.; Johnsson, F., Influence of Operating Conditions on SO3 Formation during Air and OxyFuel Combustion. Ind. Eng. Chem. Res. 2012, 51, (28), 9483-9491. 39. Zhuang, Y.; Martin, C.; Pavlish, J.; Botha, F., Cobenefit of SO3 reduction on mercury capture with activated carbon in coal flue gas. Fuel 2011, 90, (10), 2998-3006. 40. Presto, A. A.; Granite, E. J., Impact of Sulfur Oxides on Mercury Capture by Activated Carbon. Environ. Sci. Technol. 2007, 41, (18), 6579-6584. 41. Li, H.; Li, Y.; Wu, C.-Y.; Zhang, J., Oxidation and capture of elemental mercury over SiO2-TiO2-V2O5 catalysts in simulated low-rank coal combustion flue gas. Chem. Eng. J 2011, 169, (1), 186-193. 42. Delley, B., From molecules to solids with the DMol3 approach. J. Chem. Phys. 2000, 113, (18), 7756-7764. 43. Li, H.; Wu, C.-Y.; Li, Y.; Li, L.; Zhao, Y.; Zhang, J., Role of flue gas components in mercury oxidation over TiO2 supported MnOx-CeO2 mixed-oxide at low temperature. J. Hazard. Mater. 2012, 243, 117-123. 44. Liu, W.; Vidic, R. D.; Brown, T. D., Impact of Flue Gas Conditions on Mercury Uptake by Sulfur-Impregnated Activated Carbon. Environ. Sci. Technol. 2000, 34, (1), 154-159. 45. Padak, B.; Wilcox, J., Understanding mercury binding on activated carbon. Carbon 2009, 47, (12), 2855-2864. 46. Romano, E. J.; Schulz, K. H., A XPS investigation of SO2 adsorption on ceria-zirconia mixed-metal oxides. Appl. Surf. Sci. 2005, 246, (1), 262-270. 47. Espinós, J. P.; Morales, J.; Barranco, A.; Caballero, A.; Holgado, J. P.; González-Elipe, A. R., Interface Effects for Cu, CuO, and Cu2O Deposited on SiO2 and ZrO2. XPS Determination of the Valence State of Copper in Cu/SiO2 and Cu/ZrO2 Catalysts. J. Phys. Chem. B 2002, 106, (27), 6921-6929. 48. Shenasa, M.; Sainkar, S.; Lichtman, D., XPS study of some selected selenium compounds. J. Electron Spect. Related Phen. 1986, 40, (4), 329-337. 49. Hayez, V.; Franquet, A.; Hubin, A.; Terryn, H., XPS study of the atmospheric corrosion of copper alloys of archaeological interest. Surf. Inter. Anal. 2004, 36, (8), 876-879. 50. Guo, X.; Xiao, H.-S.; Wang, F.; Zhang, Y.-H., Micro-Raman and FTIR Spectroscopic Observation on the Phase Transitions of MnSO4 Droplets and Ionic Interactions between Mn2+ and SO42−. J. Phys. Chem. A 2010, 114, (23), 6480-6486. ACS Paragon Plus Environment

Environmental Science & Technology

433 434 435 436 437 438 439 440

Page 16 of 24

51. Olson, E. S.; Crocker, C. R.; Benson, S. A.; Pavlish, J. H.; Holmes, M. J., Surface Compositions of Carbon Sorbents Exposed to Simulated Low-Rank Coal Flue Gases. J. Air Waste Manage. Associa. 2005, 55, (6), 747-754. 52. Filatov, M.; Cremer, D., Revision of the Dissociation Energies of Mercury Chalcogenides-Unusual Types of Mercury Bonding. ChemPhysChem 2004, 5, (10), 1547-1557. 53. Post, B.; Schwartz, R. S.; Fankuchen, I., The crystal structure of sulfur dioxide. Acta Crystallo 1952, 5, (3), 372-374. 54. Rumayor, M.; Fernandez-Miranda, N.; Lopez-Anton, M. A.; Diaz-Somoano, M.; Martinez-Tarazona, M. R., Application of mercury temperature programmed desorption (HgTPD) to ascertain mercury/char interactions. Fuel Process. Technol. 2015, 132, 9-14.

ACS Paragon Plus Environment

Page 17 of 24

Environmental Science & Technology

442

List of Figures:

443

Figure 1. (a) breakthrough curves and (b) derived adsorption capacity of Nano-CuS for Hg0 removal

444

under different conditions.

445

Figure 2. 4-h Hg0 breakthrough curves of fresh and spent Nano-CuS pretreated under different

446

conditions.

447

Figure 3. XPS spectra of (a) S 2p, (b) O 1s and (c) Cu 2p of fresh and spent Nano-CuS.

448

Figure 4. SO2-TPD profile of fresh and pretreated Nano-CuS (inserted with the FTIR spectra of fresh

449

and pretreated Nano-CuS).

450

Figure 5. Reduction mechanism of HgS by SO3 on a molecular level: (a) Hg0 immobilized by a

451

monosulfide; (b) surface HgS attacked by a SO3; (c) surface HgS attacked by another SO3; and (d) as-

452

formed three SO2 in panel (c).

453

Figure 6. (a) Hg-TPD of spent Nano-CuS pretreated under different conditions; and (b) leaching of

454

mercury from different samples.

ACS Paragon Plus Environment

Environmental Science & Technology

456

Figure 1. (a) 4-h breakthrough curves and (b) derived adsorption capacity of Nano-CuS for Hg0

457

removal under different conditions.

Outlet Hg0 concentration (g/m3)

100

80

60

40

20

0

0

60

120 Time (min)

458 30

180

240

(b)

Pure N2 N2 + 4% O2 N2 + 4% O2 + 1000 ppm SO2 N2 + 4% O2 + 1000 ppm SO2 + 100 ppm SO3 N2 + 4% O2 + 1000 ppm SO2 + 500 ppm SO3

25 Hg0 uptaken by CuS (mg/g)

(a)

Pure N2 N2 + 4% O2 N2 + 4% O2 + 1000 ppm SO2 N2 + 4% O2 + 1000 ppm SO2 + 100 ppm SO3 N2 + 4% O2 + 1000 ppm SO2 + 500 ppm SO3

20 15 10 5 0

459

0

40

80

120 160 Time (min)

200

240

ACS Paragon Plus Environment

Page 18 of 24

Page 19 of 24

Environmental Science & Technology

461

Figure 2. 4-h Hg0 breakthrough curves of fresh and spent Nano-CuS pretreated under different

462

conditions.

Outlet Hg0 concentration (g/m3)

100

80

60

40

20

0

463

Fresh Pretreated by 4% O2 + 1000 ppm SO2 Pretreated by 4% O2 + 1000 ppm SO2 + 500 ppm SO3

0

60

120 Time (min)

180

240

ACS Paragon Plus Environment

Environmental Science & Technology

465

Figure 3. XPS spectra of (a) S 2p, (b) O 1s and (c) Cu 2p of fresh and spent Nano-CuS (a) S 2p S4+/S6+

S2-

Sx2- S22-

Intensity (a.u.)

Spent Nano-CuS

Fresh Nano-CuS

176

172

168 164 Binding energy (eV)

466

160

156

(b) O 1s

Intensity (a.u.)

Spent Nano-CuS

Fresh Nano-CuS

540

535 530 Binding energy (eV)

467 (c) Cu 2p

525

Cu2+ in CuS

Cu2+ in CuSO4

Intensity (a.u.)

Spent Nano-CuS

Fresh Nano-CuS

940

468

938

936 934 932 Binding energy (eV)

930

928

ACS Paragon Plus Environment

Page 20 of 24

Page 21 of 24

Environmental Science & Technology

470

Figure 4. SO2-TPD profile of fresh and pretreated Nano-CuS (inserted with the FTIR spectra of

471

fresh and pretreated Nano-CuS).

Absorbance (a.u.)

Intensity (a.u.)

Fresh Pretreated with N2 + 4% O2 + 1000 ppm SO2 + 500 ppm SO3

Pretreated Nano-CuS Fresh Nano-CuS

800

100

472

CuSO4

200

1200 1600 2000 Wavenumber (cm-1)

300

2400

400 500 600 Temperature (oC)

700

800

900

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 24

474

Figure 5. Reduction mechanism of HgS by SO3 on a molecular level: (a) Hg0 immobilized by a

475

monosulfide; (b) surface HgS attacked by a SO3; (c) surface HgS attacked by another SO3; and (d)

476

as-formed three SO2 in panel (c).

477

ACS Paragon Plus Environment

Page 23 of 24

Environmental Science & Technology

479

Figure 6. (a) Hg-TPD of spent Nano-CuS pretreated under different conditions; and (b) leaching

480

of mercury from different samples. (a)

HgS

CuS pretreated by Hg0 with SO3 Mercury signal (a.u.)

CuS pretreated by Hg0 without SO3

HgSO4

481

50

150

250

350 Temperature (oC)

450

550

650

Leaching mercury concentration (g/L)

160

482

140

(b)

120 100 80 60 40 4 3 2 1 0

CuS pretreated by Hg0 without SO3

CuS pretreated by Hg0 with SO3

ACS Paragon Plus Environment

Environmental Science & Technology

484

Table of contents

485

ACS Paragon Plus Environment

Page 24 of 24