Influence of Dissolved Metals on N-Nitrosamine Formation under

Sep 3, 2015 - As the prime contender for postcombustion CO2 capture technology, amine-based scrubbing has to address the concerns over the formation o...
0 downloads 9 Views 738KB Size
Subscriber access provided by UNIV OF CAMBRIDGE

Article

Influence of Dissolved Metals on N-Nitrosamine Formation Under Amine-based CO2 Capture Conditions Zimeng Wang, and William A. Mitch Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b03085 • Publication Date (Web): 03 Sep 2015 Downloaded from http://pubs.acs.org on September 12, 2015

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 36

Environmental Science & Technology

Influence of Dissolved Metals on N-Nitrosamine Formation Under Amine-based CO2 Capture Conditions

Zimeng Wang and William A. Mitch*

Department of Civil and Environmental Engineering, Stanford University, Stanford, CA, United States

*Corresponding author Jerry Yang and Akiko Yamazaki Environment & Energy Building 473 Via Ortega Stanford, California 94305, United States Tel: 650-725-9298 Fax: 650-723-7058 Email: [email protected]

Revised Manuscript Submitted to Environmental Science & Technology

August 2015

ACS Paragon Plus Environment

Environmental Science & Technology

1

Abstract

2

As the prime contender for post-combustion CO2 capture technology, amine-based scrubbing

3

has to address the concerns over the formation of potentially carcinogenic N-nitrosamine

4

byproducts from reactions between flue gas NOx and amine solvents. This bench-scale study

5

evaluated the influence of dissolved metals on the potential to form total N-nitrosamines in the

6

solvent within the absorber unit and upon a pressure-cooker treatment that mimics desorber

7

conditions. Among six transition metals tested for the benchmark solvent monoethanolamine

8

(MEA), dissolved Cu promoted total N-nitrosamine formation in the absorber unit at

9

concentrations permitted in drinking water, but not the desorber unit. The Cu effect increased

10

with oxygen concentration. Variation of the amine structural characteristics (amine order, steric

11

hindrance, -OH group substitution and alkyl chain length) indicated that Cu promotes N-

12

nitrosamine formation from primary amines with hydroxyl or carboxyl groups (amino acids), but

13

not from secondary amines, tertiary amines, sterically hindered primary amines, or amines

14

without oxygenated groups. Ethylenediaminetetraacetate (EDTA) suppressed the Cu effect. The

15

results suggested that the catalytic effect of Cu may be associated with the oxidative degradation

16

of primary amines in the absorber unit, a process known to produce a wide spectrum of

17

secondary amine products that are more readily nitrosatable than the pristine primary amines,

1

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36

Environmental Science & Technology

18

and that can form stable N-nitrosamines. This study highlighted an intriguing linkage between

19

amine degradation (operational cost) and N-nitrosamine formation (health hazards), all of which

20

are challenges for commercial-scale CO2 capture technology.

21

22

Introduction

23

Carbon capture and sequestration is one of the necessary strategies to mitigate climate

24

change associated with anthropogenic greenhouse gas emissions to the atmosphere. Fossil-fuel

25

combustion power plants are the largest point sources for CO2 emissions. In the near term, the

26

amine-based scrubbing process remains one of the most mature and economically efficient

27

carbon capture technologies for post-combustion applications.1 Over the last few years, several

28

pilot plants and test facilities have been evaluated in North America, Europe and Australia.2-4 In

29

an amine-based post-combustion CO2 capture process, flue gases (typically 4-12% CO2, 4-10%

30

O2, and several hundred ppmv NOx) pass through an absorber column where a countercurrent

31

amine solvent absorbs CO2 by reacting to form a covalent carbamate (for primary and secondary

32

amines) or ionic bicarbonate complex (for tertiary amines and some sterically-hindered amines).

33

The CO2-loaded amine solution then proceeds to a desorber column, where high temperature

34

(usually above 100 °C) reverses the reaction, regenerating lean amine that returns to the absorber 2

ACS Paragon Plus Environment

Environmental Science & Technology

35

and releasing CO2 for compression and geological storage. While a number of alternative amine

36

solvents have been proposed,5 monoethanolamine (MEA) is the industry benchmark due to its

37

low cost and well-documented physical and chemical properties.

38

To improve the economic and environmental sustainability of amine-based carbon capture

39

technologies, extensive research6,

40

degradation,8-12 (2) metallic leaching from equipment corrosion13-15, fly ash16 and metal-based

41

corrosion inhibitors17,

42

amines and reactive nitrogen species (e.g., NOx and nitrite).19, 20 Emerging evidence suggests that

43

the three processes are interdependent in scenarios relevant to carbon capture conditions.

18

7

is being performed in three areas: (1) oxidative solvent

(3) formation of carcinogenic byproducts from the reaction between

44

Amine solvent degradation is a concern due to the high cost of the solvent. Previous research

45

suggests that solvent degradation and metal leaching are synergistic processes. Amine

46

carbamates and oxidative degradation products of amines (e.g., heat stable salts, particularly

47

oxalic acid) can exacerbate metal leaching from equipment21 and from fly ash16, 22, 23 by acting as

48

conductive electrolytes and chelating agents for metals. Carbon capture facilities may

49

accumulate up to mmol/L concentrations of dissolved Fe, Ni, Cu and Cr over 1000 hours of

50

operation.22, 24, 25 On the other hand, it is proposed that oxidative degradation of amines proceeds

51

by free radicals, deriving from the decomposition of trace amounts of organic hydroperoxides.26,

3

ACS Paragon Plus Environment

Page 4 of 36

Page 5 of 36

Environmental Science & Technology

52

27

53

oxidative degradation of amines. Several transition metals that can come from equipment

54

corrosion,14,

55

chromium26, 29 and manganese26, were reported to be catalysts for oxidative degradation of MEA.

56

The earliest report of MEA oxidative degradation catalyzed by transition metals was during the

57

1960s in the context of a submarine CO2 scrubber, in which the catalytic activity of Cu was

58

higher than for iron, nickel and chromium.30, 31 Field tests also found a correlation between MEA

59

degradation (represented by NH3 emission rates) and dissolved metal concentrations in the

60

solvent.32

Hydroperoxide decomposition can be catalyzed by certain transition metals, accelerating

17

fly ash,16 and corrosion inhibitors18, including iron21,

28, 29

, copper21,

28, 29

,

61

An emerging concern for amine-based post-combustion CO2 capture is the formation of

62

potentially carcinogenic N-nitrosamines and N-nitramines from reactions between the amine

63

solvent and flue gas NOx.6,

64

nitrosamines was orders of magnitude higher than N-nitramines.19,

65

indicated that N-nitrosamines were more mutagenic than their N-nitramine analogues.34 Previous

66

biomedical literature on N-nitrosamine formation from the reaction between nitrite and amines

67

under acidic conditions has suggested that N2O3 is the responsible nitrosating agent35, and our

68

recent results also indicated that N2O3 (formed by the reaction of NO with NO2) is the dominant

7

We reported that within the capture unit, formation of N-

4

ACS Paragon Plus Environment

33

In vitro toxicity tests

Environmental Science & Technology

69

nitrosating agent within the absorber.33 Within the desorber, where gaseous species are depleted,

70

the high temperature conditions promote nitrosation of the amines by nitrite despite the high

71

pH.36-38 The amine order dictates the rate and pathways of N-nitrosamine formation.39 Only

72

secondary amines can form N-nitrosamines. Nitrosation of tertiary amines forms an unstable N-

73

nitrosamine intermediate that de-alkylates to form nitrosatable secondary amines. Nitrosation of

74

primary amines also forms a highly unstable N-nitrosamine intermediate, which decays nearly

75

instantaneously to nitrogen gas and a carbocation. Under both absorber and desorber conditions,

76

the rates of total N-nitrosamine formation were in the order: secondary ≈ tertiary >> primary

77

amines.

78

Recent research suggests a connection between oxidative amine degradation and formation

79

of N-nitrosamine byproducts. The small but appreciable formation of N-nitrosamines from

80

primary amine solvents is believed to result from the formation of secondary amines as

81

degradation products of the primary amines.38-40 Relevant degradation pathways of primary

82

amines that lead to secondary amines include oxidation (probably the primary pathway in the

83

absorber41) and carbamate polymerization (occurring at desorber temperatures6). Thus, for

84

primary amines there is a potential coupling between N-nitrosamine formation and amine

85

degradation.

5

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36

Environmental Science & Technology

86

The effect of transition metals on N-nitrosamine formation under amine-based post-

87

combustion carbon capture conditions has not been evaluated. Because transition metals promote

88

amine solvent degradation, and formation of secondary amines from degradation of primary

89

amine solvents enables N-nitrosamine formation, transition metals may enhance N-nitrosamine

90

formation. Additionally, transition metals may oxidize NO to form nitrosyl cation (+NO), a

91

potent but short-lived nitrosating agent42. Although readily hydrolyzed to nitrite at carbon

92

capture relevant pH, +NO may react with nitrite to form the nitrosating agent, N2O3.43 Moreover,

93

transition metals can form complexes with +NO that also are nitrosating agents.44 Previous

94

biomedical literature has characterized metal-catalyzed formation of N-nitrosamines in vivo,45 in

95

which the coordination46,

96

nitrosation reactions. A catalytic effect of cupric ion was previously documented for

97

nitrosodimethylamine (NDMA) formation from 1,1-dimethylhydrazine oxidation by oxygen.49, 50

98

The goal of this research was to evaluate the effect of transition metals on N-nitrosamine

99

formation during amine-based carbon capture. Bench-scale experiments were performed with the

100

following objectives: (1) to screen the transition metals that may promote N-nitrosamine

101

formation from the benchmark MEA solvent, (2) to understand the effects of amine structural

102

characteristics, O2 and metal concentrations on N-nitrosamine formation, 3) to compare the

47

and redox capabilities48 of transition metals contributed to the

6

ACS Paragon Plus Environment

Environmental Science & Technology

103

importance of transition metal promotion of N-nitrosamine formation within the absorber and

104

desorber, and 4) to evaluate the effectiveness of a transition metal chelating agent for mitigating

105

N-nitrosamine formation.

106 107

Materials and Methods

108

Materials. Deionized water was used in all experimental procedures including the

109

preparation of amine solvents. MEA and nine other amines varying in amine order, steric

110

hindrance, functional group substitution and alkyl chain length were evaluated (Figure 1).

111

Praxair nitrogen (99.995%), carbon dioxide (99.99%), oxygen (99.6%), NO (2580 ppmv in

112

nitrogen) and NO2 (2543 ppmv in nitrogen) were used as received. The flow rates of NO and

113

NO2 were set by mass flow controllers and those of the other major gases were controlled by

114

volumetric flow meters. The dissolved metal stock solutions (CuCl2, FeCl3, MnCl2, ZnSO4,

115

NiSO4, CrCl3) were prepared by dissolution into deionized water. More details about the amines

116

and metals are provided in the Supporting Information.

117

Bench-scale CO2 absorber column setup. The flue gas-amine reactions were simulated in

118

an absorber column (5 cm diameter, 40 cm length), containing 6 mm borosilicate glass Raschig

119

rings. A 2.5 M amine stock solution (200 mL) mixed by a magnetic stir bar was recirculated

7

ACS Paragon Plus Environment

Page 8 of 36

Page 9 of 36

Environmental Science & Technology

120

through the absorber column at 15 mL/min counter-current to the synthetic flue gas. The

121

absorber column and solvent reservoir were maintained at 48 °C, a representative absorber

122

temperature during postcombustion carbon capture.24 The amine solutions were self-buffered,

123

but amino acid solutions were adjusted to 1 pH unit above the pKa of the amine group using

124

sodium hydroxide. The synthetic flue gas (6 L/min) was constituted by N2, O2, CO2, NO, NO2

125

and water vapor. The 12% humidity was set by purging the N2, O2 and CO2 mixture through a

126

pre-wetting column maintained at 48 °C. NO and NO2 were added downstream of the pre-

127

wetting column. The baseline condition for flue gas composition was 4 % CO2, 14 % O2, 15

128

ppmv NO, 3 ppmv NO2 and 6.5 % H2O, with N2 making up the balance of the gas. Due to the

129

lack of desorber, the system was pre-equilibrated with the synthetic flue gas without NOx for 1

130

hour, which was verified to be sufficient to reach steady CO2 loading (i.e., rich conditions).33

131

Then NOx was introduced to the synthetic flue gas. The liquid amine samples were collected at

132

various time points after NOx addition and analyzed for nitrite and total N-nitrosamine

133

concentrations. Because monitoring of the volume in the amine reservoir indicated that it

134

declined during the experiment due to evaporation, the results are reported as the rate of

135

formation of reaction products by mass rather than by concentration. Error bars represent the

136

standard deviation of experimental replicates (n = 2-4). The effect of dissolved metals on N-

8

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 36

137

nitrosamine formation was evaluated by adding aliquots of concentrated metal stock solutions

138

into the amine reservoir in the midst of the experiments. The pH shift of the amine solution upon

139

addition of dissolved metals was negligible.

140

Pressure cooker treatment to mimic desorber conditions. To simulate desorber conditions,

141

selected fresh amines were maintained at 48 °C in a water bath and mixed with aliquots of

142

sodium bicarbonate. A 1-hour water bath treatment enabled formation of carbamates to simulate

143

the loaded amines entering desorber units. Mixing amines with bicarbonate in this fashion was

144

previously verified to form carbamates.36 Then the amines were titrated with aliquots of sodium

145

nitrite and dissolved metals. To exclude the variation of pH, all the solutions were adjusted to pH

146

10.4, representative of desorber conditions.24 Those reactors were heated to 120 °C in a pressure

147

cooker at 15 psi for 6 hours. Nitrite and total N-nitrosamines were measured before and after

148

pressure-cooker treatment.

149

Analyses. Nitrite was measured by the N-(1-naphthyl)ethylenediamine dihydrochloride

150

colorimetric method.51 CO2 loadings of solvent amines were measured following a barium

151

titration method previously described.52 Among the 10 different amines tested, only morpholine

152

(MOR) and diethanolamine (DEA), the secondary amines, react directly to produce N-

153

nitrosamines (i.e., N-nitrosomorpholine and N-nitrosodiethanolamine). To form stable N-

9

ACS Paragon Plus Environment

Page 11 of 36

Environmental Science & Technology

154

nitrosamines, the other amines, particularly the 7 primary amines, require transformation to

155

secondary amines. Because a wide array of potentially nitrosatable amine products may form

156

during degradation of these parent amines, a total N-nitrosamine (TONO) analysis was used as

157

the primary analytical method to quantify N-nitrosamine formation. The TONO assay was

158

described previously.19 Briefly, after pretreatment with sulfamic acid to remove nitrite, which

159

interferes with the TONO assay, a sample is injected into a heated reaction chamber containing

160

an acidic tri-iodide solution, which causes cleavage of the N-NO bond in N-nitrosamines. NO is

161

purged from the reaction chamber by a stream of nitrogen gas into a chemiluminescence detector

162

(EcoPhysics CLD 88sp). The total N-nitrosamine concentration is quantified based upon the

163

moles of NO liberated, using NDMA to construct a standard curve. Interference of dissolved

164

metals with the TONO assay was ruled out by the absence of signal from control samples

165

containing dissolved metals only, dissolved metals mixed with amines and nitrite, and samples

166

collected during a control experiment involving sparging the synthetic flue gas into a metal

167

solution in the absence of amines. Selected samples from MOR experiments were also analyzed

168

for nitrosomorpholine concentration by high performance liquid chromatography (HPLC) with

169

UV detection using a method described in our previous publication.33

170

10

ACS Paragon Plus Environment

Environmental Science & Technology

171

Page 12 of 36

Results

172

Screening tests for metals using MEA. We screened the effects of several metals that are

173

relevant to equipment corrosion and fly ash particulates on N-nitrosamine formation from MEA

174

within the absorber column. Under the baseline flue gas conditions (4 % CO2, 14 % O2, 15 ppmv

175

NO, 3 ppmv NO2 and 6.5 % H2O), the total N-nitrosamine formation rate was 0.062 (± 0.026)

176

nmol/s. Injection of 5 µM and 100 µM Cu during the experiment promoted total N-nitrosamine

177

formation rates instantaneously by factors of 5 and 20, respectively (Figure 2). Note that 5 µM

178

Cu is below the 20.5 µM (1.3 mg/L) Primary Maximum Contaminant Level (MCL) for Cu in

179

U.S. drinking waters.53 Although Cu sorption to borosilicate class is possible, the high amine

180

concentration in the solvent would favor complexation with the amines.54 Similar experiments

181

performed with all other metals (Fe, Mn, Ni, Zn and Cr) did not show any appreciable effect of

182

the metals on N-nitrosamine formation rates.

183

Importance of Oxygen. Absorber experiments performed with various combinations of Cu

184

and O2 concentrations suggested that the promotional effect of Cu on total N-nitrosamine

185

formation rates from MEA is dependent on the oxygen content in the flue gas (Figure 3). In the

186

absence of Cu and O2, the total N-nitrosamine formation rate was 0.020 (± 0.001) nmol/s.

11

ACS Paragon Plus Environment

Page 13 of 36

Environmental Science & Technology

187

Increasing O2 concentration to 6 % and 14 % increased the N-nitrosamine formation rate by

188

factors of 2 and 3, respectively.

189

In the absence of O2 in the flue gas, addition of Cu had a minimal effect. Even though a

190

slight increase of N-nitrosamine formation rate (less than a factor of 4) was suggested at 100 µM

191

Cu, it is probably attributable to the fact that our MEA solvent reservoir was not completely free

192

of O2 uptake from the ambient air as the experiments were conducted in a well-vented fume hood.

193

At 6 % O2, 5 µM and 100 µM Cu increased the N-nitrosamine formation rate by factors of 2.7

194

and 15, respectively. These values are lower than the factors of 5 and 20 observed at 14 % O2,

195

yet still substantial.

196

The results suggest that the effects of Cu and O2 are synergistic with respect to N-nitrosamine

197

formation form MEA. Previous studies reported that MEA degradation kinetics appeared to be

198

first order with respect to the concentration of O2.55 The similar dependence of total N-

199

nitrosamine formation rates from MEA on O2 concentration (Figure 3) suggests that the N-

200

nitrosamine formation is associated with MEA degradation. In contrast, previous research

201

demonstrated that the flue gas O2 concentration in the absorber did not impact N-nitrosamine

202

formation from morpholine (MOR), a secondary amine that can form N-nitrosamines directly.33

12

ACS Paragon Plus Environment

Environmental Science & Technology

203

These results suggest that N-nitrosamine formation from MEA derives from secondary amines

204

generated as degradation products of MEA38, 39 via reactions promoted by O2 and Cu.

Page 14 of 36

205

Importance of amine structure. To evaluate the extent to which the Cu promotion of N-

206

nitrosamine formation is specific to MEA, the solvent amine structure was varied to capture

207

differences in (1) amine order (i.e., primary, secondary, and tertiary), (2) steric hindrance of the

208

amino group in primary alkanolamines, (3) substituent functional groups in primary amines, and

209

(4) the alkyl chain length in primary amines (Figure 1). In the absence of Cu, the total N-

210

nitrosamine formation rates for primary amines were one or more orders of magnitude lower

211

than those for two secondary amines (MOR and DEA) and one tertiary amine (TEA) (Figure 4A),

212

concurring with the results of previous research.36 For primary amines, substituting the hydroxyl

213

group of MEA with an amine group (EDA), increasing the alkyl chain length (3AP), or

214

introducing steric hindrance (AMP and TRIS) had little effect on total N-nitrosamine formation

215

compared with the benchmark MEA in the absence of Cu. The two primary amino acids (GLY

216

and ALA) featured the lowest total N-nitrosamine formation rates. The trends of those rates for

217

N-nitrosamine formation, as well as for nitrite formation (Figure 4B), were consistent with

218

previous results.39, 56

13

ACS Paragon Plus Environment

Page 15 of 36

Environmental Science & Technology

219

The promotional effect of Cu on total N-nitrosamine formation was observed only for the

220

primary amines, MEA, GLY, ALA and 3AP (Figure 4A). Addition of Cu had no impact on

221

nitrite formation rates for any of the tested amines (Figure 4B), indicating that Cu does not

222

impact the absorption of NOx from the gas phase. As a secondary amine, MOR can form stable

223

N-nitrosamines directly. Indeed, for samples collected from MOR experiments, the

224

concentrations of N-nitrosomorpholine (NMOR) by HPLC matched well with the concentration

225

of total N-nitrosamines.36, 37 For all other amines tested, the original amines must be converted to

226

secondary amines to produce stable N-nitrosamines. Transformation of tertiary amines to

227

nitrosatable secondary amines occurs through dealkylation of an unstable nitrosated tertiary

228

amine intermediate releasing an aldehyde and a secondary amine.57 The rate of total N-

229

nitrosamine formation from the tertiary amine, TEA, was also not impacted by Cu, suggesting

230

that the dealkylation step was not catalyzed by Cu.

231

Sterically hindered primary amines are attractive solvents compared to MEA because they

232

require lower temperatures to regenerate and they are more resistant to oxidative degradation.8,

233

58-60

234

the promotional effect of Cu for total N-nitrosamine formation. The two amino acids (GLY,

235

ALA), which lack steric hindrance and feature substitution of the hydroxyl group of MEA by

Our results showed that introducing steric hindrance to MEA (i.e., AMP, TRIS) eliminated

14

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 36

236

carboxyl groups, demonstrated a significant promotional effect of Cu for N-nitrosamine

237

formation. However, substituting the hydroxyl group of MEA by an amine group (EDA)

238

eliminated the Cu effect. The amine 3AP, which retains the hydroxyl group of MEA but bears an

239

additional alkyl carbon, also exhibited a higher total N-nitrosamine formation rate after addition

240

of Cu. Together, the results suggest that promotion of N-nitrosamine formation by Cu is

241

applicable to primary amines with oxygenated functional groups that lack steric hindrance.

242

Interestingly, GLY and ALA are the intuitive oxidation products of MEA and 3AP, respectively

243

(i.e., oxidation of the hydroxyl group to a carboxyl group). While the N-nitrosamine formation

244

rate for GLY was promoted by Cu, it was an order of magnitude less than for MEA exposed to

245

Cu.

246

Test of “NO + Cu” as a potential nitrosating agent. Previous studies reported that NO and

247

Cu react to form +NO, a potent nitrosating agent for secondary amines at alkaline pH,48,

248

following a reaction such as

249

Cu2+ + NO + R2NH → Cu+ + H+ + R2N-N=O

61

(1)

250

This “reductive nitrosation” reaction is particularly appreciated for its biochemical function for

251

NO transport and signaling as part of the mammalian immune response.62 In amine-based carbon

252

capture processes, it is of interest to investigate if the lack of any effect of Cu in the MOR

15

ACS Paragon Plus Environment

Page 17 of 36

Environmental Science & Technology

253

experiments was simply due to the presence of NO2 in the synthetic flue gas, which might have

254

masked the impacts, if any, of oxidation of NO by Cu. In authentic flue gas, NO2 is usually

255

easier to remove than NO during pre-treatments due to its higher solubility. Downstream of the

256

absorber column, where most of the NO2 is likely scavenged, NO is probably the dominant

257

reactive nitrogen species in the washwater unit.33 We performed a control experiment using 0.05

258

M MOR (to simulate a diluted amine concentration in the washwater33, 39) in which the flue gas

259

was free of NO2 but contained NO, O2, CO2 and N2. The addition of 100 µM Cu did not induce

260

any change in the N-nitrosamine formation rate (Figure S1). Although the reductive nitrosation

261

mechanism enabled by Cu and NO is physically possible, it does not appear to be as important in

262

post-combustion carbon capture systems as in biomedical systems. Because flue gas always

263

contains unconsumed oxygen which can partially oxidize NO to NO2, N2O3 can form as the

264

predominant nitrosating agent.33, 63 Thus, addition of Cu to initiate reaction 1, and allowing Cu to

265

cycle between the +I and +II oxidation states by oxidation of Cu(I) by oxygen does not

266

contribute appreciably to N-nitrosamine formation compared with N2O3.

267

Cu effects quenched by a chelating agent. Addition of a strong chelating agent could

268

suppress the promotional effect of Cu on N-nitrosamine formation. While spiking 100 µM Cu

269

accelerated the formation of N-nitrosamines in MEA by a factor of 20, adding 1 mM EDTA

16

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 36

270

dramatically reduced the N-nitrosamine formation rate (Figure 5). The usefulness of EDTA as an

271

inhibitor of Cu catalysis for MEA oxidative degradation has been recognized since the 1960s

272

with reference to a CO2 scrubber in submarines.30,31 Recent studies performed under post-

273

combustion CO2 capture conditions confirmed the role of EDTA in suppressing Cu- and Fe-

274

catalyzed MEA oxidative degradation, although the inhibition capacity declined over time.28, 29

275

Our results indicate that the inhibitory effect of EDTA also is applicable to Cu-catalyzed N-

276

nitrosamine formation from MEA.

277

Metal effects under desorber conditions. Under desorber conditions, where nitrite is the

278

nitrosating agent,38 the effect of 100 µM metals on N-nitrosamine formation rates was compared

279

between MEA and MOR. Previously, we had demonstrated that nitrosamine formation was more

280

important in the absorber for primary amines, but more important in the desorber for secondary

281

and tertiary amines.39 Six hours of pressure-cooker treatment completely consumed an initial 5

282

mM nitrite when the MOR concentration was 2.5 M as in the absorber experiments. Those

283

experiments produced 5 mM N-nitrosomorpholine, as confirmed by both TONO and HPLC

284

analyses, consistent with previous literature that nitrosation of secondary amines by nitrite is

285

stoichiometric.36, 38 Therefore, a lower MOR concentration (25 mM) was used to examine the

17

ACS Paragon Plus Environment

Page 19 of 36

Environmental Science & Technology

286

effect of dissolved metals. No effect of any dissolved metal (100 µM) on the formation of N-

287

nitrosamines was observed for MOR (Figure 6).

288

Compared with the absorber conditions, similar results were obtained from the metal

289

screening tests under the desorber conditions using 2.5 M MEA (Figure 6). None of the metals

290

affected total N-nitrosamine formation rates, except for 100 µM Cu, for which a 2.5-fold increase

291

in total N-nitrosamine formation was observed. Nevertheless, the promotional effect of Cu under

292

desorber conditions, where nitrite was the nitrosating agent, was much less pronounced

293

compared with the 20-fold increase under absorber conditions, where flue gas containing oxygen

294

and NOx was continuously delivered to the MEA solvent. It was also noted that the amount of

295

total N-nitrosamine formation (on the order of µM) was orders of magnitude less than the

296

consumption of nitrite (on the order mM). These results are consistent with nitrosation of MEA

297

leading predominantly to formation of nitrogen gas and a carbocation, with only a small fraction

298

of the nitrite reacting with the traces of secondary amine degradation products of MEA to form

299

stable N-nitrosamines.

300

Interestingly, the pressure cooker treatments generated the same order-of-magnitude total N-

301

nitrosamine concentration from MOR and MEA, even though the MOR concentration was three

302

orders of magnitude lower than MEA. This confirmed again the significantly lower potential for

18

ACS Paragon Plus Environment

Environmental Science & Technology

303

N-nitrosamine formation from primary amines than secondary amines under desorber

304

conditions.39

Page 20 of 36

305

Discussion

306

An indirect role for Cu in catalyzing N-nitrosamine formation. The formation of N-

307

nitrosamines from MEA and other primary amines relies on an initial transformation via

308

oxidative degradation to secondary amines that can be nitrosated.40 Therefore the catalytic effect

309

of Cu may relate to either formation of the secondary amines or nitrosation of the secondary

310

amines. MOR and DEA are among the possible oxidative degradation products of MEA.11, 26

311

Moreover, MOR is resistant to oxidative degradation.64 The lack of Cu effect on total N-

312

nitrosamine formation from MOR and DEA (Figure 4 and S1) indicates that Cu does not

313

catalyze the nitrosation of secondary amines under carbon capture conditions.

314

The correlation of the promotion of N-nitrosamine formation by Cu from MEA with oxygen

315

concentration (Figure 3) suggests that Cu promotes the oxidative degradation of MEA to form an

316

array of products, a portion of which are secondary amines that can react with N2O3 in the

317

absorber column to produce N-nitrosamines. Previous research has indicated that oxygen

318

promotes the degradation of MEA to form an array of secondary amine products.8, 11, 65 The wide

319

array of products that are produced renders the characterization of N-nitrosamine formation

19

ACS Paragon Plus Environment

Page 21 of 36

Environmental Science & Technology

320

pathways difficult, and necessitated the use of the total N-nitrosamine assay to quantify N-

321

nitrosamine production. Indeed, the Cu effect was not observed for sterically-hindered primary

322

amines, which resist oxidative degradation because they lack α-hydrogens; α-hydrogens play a

323

role in the formation of an imine, an initial step in an oxidative degradation pathway following

324

electron abstraction from the lone electron pair on the amine nitrogen (see Scheme S1).10, 60

325

Additionally, the Cu effect was only observed for primary amines with oxygenated

326

functional groups. Previous research on MEA oxidative degradation pathways indicated that the

327

hydroxyl group of MEA reacts with carboxylic acids (which are known MEA degradation

328

products) to form esters. Ring cyclization followed by an SN2 reaction with a second MEA

329

produces secondary amine products, such as hydroxyethylethylenediamine (HEEDA).8, 11, 41, 59

330

The carboxyl groups of GLY and ALA may play a similar role in producing secondary amines

331

via oxidative degradation. In contrast, EDA which does not contain an oxygenated functional

332

group, and did not exhibit a promotional effect on N-nitrosamine formation by Cu. Nevertheless,

333

definition of the specific reactions involved in Cu promotion of N-nitrosamine formation from

334

primary amines requires further characterization of the oxidative degradation pathways of

335

primary amines, which is beyond the scope of this study.

20

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 36

336

While our results indicated the Cu catalysis of total N-nitrosamine formation relates to the

337

promotion of oxidative degradation, Cu likely promotes a specific subset of degradation

338

pathways leading to secondary amine formation. Fe and Mn were also reported to catalyze MEA

339

oxidative degradation, as indicated by MEA loss, and instantaneously enhanced formation rates

340

of organic acids and NH3.21, 26, 28, 29 However, Fe and Mn do not promote total N-nitrosamine

341

formation, suggesting that these metals promote the formation of products other than secondary

342

amines. Formation of N-nitrosamines from tertiary amines proceeds through a nitrosative

343

dealkylation step to produce secondary amines. However, Cu did not promote N-nitrosamine

344

formation from TEA (Figure 4). Previous research indicating similar total N-nitrosamine

345

formation rates between tertiary amines and their corresponding secondary amines suggests that

346

the dealkylation reaction is relatively efficient under absorber conditions,39 such that catalysis by

347

Cu, if any, would not significantly enhance the observed N-nitrosamine formation rates.

348

Implications for carbon capture practice. Cu was the only one of six transition metals that

349

promoted N-nitrosamine formation from primary amines. The low potential for N-nitrosamine

350

formation from primary amines has been considered an attractive feature compared to secondary

351

and tertiary amines. However, the N-nitrosamine formation rates from MEA in the presence of

352

Cu were similar to those of secondary and tertiary amines. Although for secondary and tertiary

21

ACS Paragon Plus Environment

Page 23 of 36

Environmental Science & Technology

353

amines, nitrite-induced nitrosamine formation in the desorber is more important than that in the

354

absorber, primary amines were found to form more nitrosamines in the absorber than in the

355

desorber.39,

356

absorber.

66

The catalytic effect of Cu can further enhance nitrosamine formation in the

357

Carbon capture operations employing primary amine solvents will need to beware of

358

potential sources of Cu entering solvents and washwaters, where reactions of residual NOx with

359

amines accumulating in the washwater may also form nitrosamines.33 While Cu is usually

360

considered as a relatively minor component in stainless steel, it is a common element in flue gas

361

particulates from coal combustion processes.22, 67 Although Cu concentrations in carbon capture

362

processes may vary from case to case depending on the specific fuel and flue gas composition,

363

Cu was observed in lean amine and reclaimer waste sludge at appreciable concentrations (ppm

364

level) that may be comparable with Fe.11, 68 Additionally, as Cu-based corrosion inhibitors (e.g.,

365

CuCO3) are widely applied for protecting stainless steel equipment,18 their promotion of

366

oxidative degradation of amine solvents has been recognized,21,

367

catalyze nitrosamine formation as illustrated in the present study. EDTA effectively suppressed

368

the Cu effect. Development of alternative oxidation inhibitors for amine solvents should evaluate

369

the extent to which they can mitigate Cu catalysis for N-nitrosamine formation.70 Lastly, the use

22

ACS Paragon Plus Environment

28, 29, 69

and they may also

Environmental Science & Technology

370

of tap water or surface waters to constitute solvents or washwaters may be a concern given that

371

Cu promoted nitrosamine formation from primary amines at concentrations lower than the

372

Maximum Contaminant Level permitted in drinking water.

Page 24 of 36

373 374

Acknowledgements

375

This work was funded by the Stanford Woods Institute for the Environment. Valuable

376

discussions with personnel from the CO2 Capture Mongstad Project and Dr. Ning Dai (currently

377

at the State University of New York at Buffalo) are appreciated. Comments and suggestions of

378

three anonymous reviewers and Associate Editor Rich Valentine improved the quality of an

379

earlier version of this paper.

380 381 382 383

Supporting Information This supporting information (one figure, one table, and additional details about the materials) is available free of charge via the Internet at http://pubs.acs.org.

384 385

23

ACS Paragon Plus Environment

Page 25 of 36

Environmental Science & Technology

386

References

387

(1)

Rochelle, G. T., Amine scrubbing for CO2 capture. Science 2009, 325 (5948), 1652-1654.

388

(2)

Maree, Y.; Nepstad, S.; de Koeijer, G., Establishment of knowledge base for emission regulation

389

for the CO2 technology centre Mongstad. Energy Procedia 2013, 37, 6265-6272.

390

(3)

391

of the rich‐split process modification at an Australian‐based post combustion CO2 capture pilot plant.

392

Greenhouse Gas. Sci. Technol. 2012, 2 (5), 329-345.

393

(4)

394

capture and storage in new fossil fuel-fired power plants. Environ. Sci. Technol. 2014, 48 (14), 7723-

395

7729.

396

(5)

397

Carbon dioxide postcombustion capture: a novel screening study of the carbon dioxide absorption

398

performance of 76 amines. Environ. Sci. Technol. 2009, 43 (16), 6427-6433.

399

(6)

400

scale postcombustion capture of CO2 with monoethanolamine solvent: key considerations for solvent

401

management and environmental impacts. Environ. Sci. Technol. 2012, 46 (7), 3643-3654.

402

(7)

403

beginning of 2013—A Review. ACS Appl. Mater. Interfaces 2015, 7 (4), 2137-2148.

404

(8)

405

review. Int. J. Greenh. Gas. Con. 2012, 10, 244-270.

406

(9)

407

monoethanolamine. Ind. Eng. Chem. Res. 2010, 50 (2), 667-673.

408

(10)

409

41 (17), 4178-4186.

410

(11)

411

a CO2 capture facility. Energy Fuels 2003, 17 (4), 1034-1039.

412

(12)

413

other gas treating processes. Int. J. Greenh. Gas. Con. 2011, 5 (1), 1-6.

414

(13)

415

absorption process using aqueous amine solutions. Ind. Eng. Chem. Res. 1999, 38 (10), 3917-3924.

416

(14)

417

positions of an amine-based CO2 Capture Pilot Plant. Ind. Eng. Chem. Res. 2012, 51 (19), 6714-6721.

Cousins, A.; Cottrell, A.; Lawson, A.; Huang, S.; Feron, P. H., Model verification and evaluation

Clark, V. R.; Herzog, H. J., Assessment of the US EPA’s determination of the role for CO2

Puxty, G.; Rowland, R.; Allport, A.; Yang, Q.; Bown, M.; Burns, R.; Maeder, M.; Attalla, M.,

Reynolds, A. J.; Verheyen, T. V.; Adeloju, S. B.; Meuleman, E.; Feron, P., Towards commercial

Dutcher, B.; Fan, M.; Russell, A. G., Amine-based CO2 capture technology development from the Gouedard, C.; Picq, D.; Launay, F.; Carrette, P.-L., Amine degradation in CO2 capture. I. A Sexton, A. J.; Rochelle, G. T., Reaction products from the oxidative degradation of Chi, S.; Rochelle, G. T., Oxidative degradation of monoethanolamine. Ind. Eng. Chem. Res. 2002, Strazisar, B. R.; Anderson, R. R.; White, C. M., Degradation pathways for monoethanolamine in Bedell, S. A., Amine autoxidation in flue gas CO2 capture—Mechanistic lessons learned from Veawab, A.; Tontiwachwuthikul, P.; Chakma, A., Corrosion behavior of carbon steel in the CO2 Gao, J.; Wang, S.; Sun, C.; Zhao, B.; Chen, C., Corrosion behavior of carbon steel at typical

24

ACS Paragon Plus Environment

Environmental Science & Technology

418

(15)

419

aluminized nickel coating in a carbon dioxide capture process using aqueous monoethanolamine. Corros.

420

Sci. 2011, 53 (11), 3666-3671.

421

(16)

422

monoethanolamine degradation during CO2 capture. Int. J. Greenh. Gas. Con. 2014, 25 (0), 102-108.

423

(17)

424

piperazine. Ind. Eng. Chem. Res. 2009, 48 (20), 9299-9306.

425

(18)

426

in gas scrubbing plant. U.S. Patent 4,690,740, 1987.

427

(19)

428

nitrosamine and nitramine formation from NOx reactions with amines during amine-based carbon dioxide

429

capture for postcombustion carbon sequestration. Environ. Sci. Technol. 2012, 46 (17), 9793-9801.

430

(20)

431

use of amines in carbon capture and storage (CCS). Chem. Soc. Rev. 2012, 41 (19), 6684-6704.

432

(21)

433

capture conditions. Ind. Eng. Chem. Res. 2004, 43 (20), 6400-6408.

434

(22)

435

coal? Energy Procedia 2014, 63 (0), 1944-1956.

436

(23)

437

Impact of flue gas contaminants on monoethanolamine thermal degradation. Ind. Eng. Chem. Res. 2013,

438

53 (2), 553-563.

439

(24)

440

300 kg/h post-combustion capture test plant using monoethanolamine. In 2nd Post Combustion Capture

441

Conference (PCCC2), Bergen, Norway, 2013.

442

(25)

443

2013, 37, 1912-1923.

444

(26)

445

Texus, Austin, 2013.

446

(27)

447

degradation inhibitors on the oxidative and thermal degradation of monoethanolamine in postcombustion

448

CO2 capture. Ind. Eng. Chem. Res. 2014, 53 (47), 18121-18129.

449

(28)

450

monoethanolamine in CO2 capture processes. Ind. Eng. Chem. Res. 2005, 45 (8), 2513-2521.

451

(29)

452

monoethanolamine. Int. J. Greenh. Gas. Con. 2009, 3 (6), 704-711.

Page 26 of 36

Sun, Y.; Remias, J. E.; Peng, X.; Dong, Z.; Neathery, J. K.; Liu, K., Corrosion behaviour of an

Chandan, P.; Richburg, L.; Bhatnagar, S.; Remias, J. E.; Liu, K., Impact of fly ash on Nainar, M.; Veawab, A., Corrosion in CO2 capture process using blended monoethanolamine and Cringle, D. C.; Pearce, R. L.; Dupart, M. S. Method for maintaining effective corrosion inhibition Dai, N.; Shah, A. D.; Hu, L.; Plewa, M. J.; McKague, B.; Mitch, W. A., Measurement of

Nielsen, C. J.; Herrmann, H.; Weller, C., Atmospheric chemistry and environmental impact of the Goff, G. S.; Rochelle, G. T., Monoethanolamine degradation: O2 mass transfer effects under CO2 Schallert, B.; Neuhaus, S.; Satterley, C. J., Is fly ash boosting amine losses in carbon capture from Huang, Q.; Thompson, J.; Bhatnagar, S.; Chandan, P.; Remias, J. E.; Selegue, J. P.; Liu, K.,

Unterberger, S.; Riederer, A.; Gunnesch, G.; Kurtz, M.; Hafer, C., Operational experience from a

Nielsen, P. T.; Li, L.; Rochelle, G. T., Piperazine degradation in pilot plants. Energy Procedia Voice, A. K. Amine oxidation in carbon dioxide capture by aqueous scrubbing. The University of Léonard, G.; Voice, A.; Toye, D.; Heyen, G., Influence of dissolved metals and oxidative

Goff, G. S.; Rochelle, G. T., Oxidation inhibitors for copper and iron catalyzed degradation of Sexton, A. J.; Rochelle, G. T., Catalysts and inhibitors for oxidative degradation of 25

ACS Paragon Plus Environment

Page 27 of 36

Environmental Science & Technology

453

(30)

Blachly, C. H.; Ravner, H., Stabilization of monoethanolamine solutions in carbon dioxide

454

scrubbers. J. Chem. Eng. Data 1966, 11 (3), 401-403.

455

(31)

456

additive package for the stabilization of monoethanolamine solutions; NRL-MR-1598; Naval Research

457

Laboratory: Washington DC, USA, 1965.

458

(32)

459

and ammonia emissions from amine based post combustion carbon capture: Lessons learned from field

460

tests. Int. J. Greenh. Gas. Con. 2013, 13 (0), 72-77.

461

(33)

462

in postcombustion CO2 capture systems. Environ. Sci. Technol. 2014, 48 (13), 7519-7526.

463

(34)

464

nitrosamines and nitramines associated with amine-based carbon capture and storage. Environ. Sci.

465

Technol. 2014, 48 (14), 8203-8211.

466

(35)

467

Toxicol. Appl. Pharmacol. 1975, 31 (3), 325-351.

468

(36)

469

nitrite and piperazine in CO2 capture. Environ. Sci. Technol. 2013, 47 (7), 3528-3534.

470

(37)

471

understand potential solvent based CO2 capture process reactions. Environ. Sci. Technol. 2013, 47 (10),

472

5481-5487.

473

(38)

474

desorber temperatures. Environ. Sci. Technol. 2014, 48 (15), 8777-8783.

475

(39)

476

potential relevant to postcombustion CO2 capture systems. Environ. Sci. Technol. 2013, 47 (22), 13175-

477

13183.

478

(40)

479

in the flue gas degradation of MEA. Energy Procedia 2011, 4 (0), 1566-1573.

480

(41)

481

H. F., Comparison of MEA degradation in pilot-scale with lab-scale experiments. Energy Procedia 2011,

482

4, 1652-1659.

483

(42)

484

nitrosodimethylamine (NDMA) as a drinking water contaminant: a review. Environ. Eng. Sci. 2003, 20

485

(5), 389-404.

486

(43)

487

441.

Blachly, C.; Ravner, H. Studies of submarine carbon dioxide scrubber operation: effect of an

Mertens, J.; Lepaumier, H.; Desagher, D.; Thielens, M.-L., Understanding ethanolamine (MEA)

Dai, N.; Mitch, W. A., Effects of flue gas compositions on nitrosamine and nitramine formation Wagner, E. D.; Osiol, J.; Mitch, W. A.; Plewa, M. J., Comparative in vitro toxicity of

Mirvish, S. S., Formation of N-nitroso compounds: Chemistry, kinetics, and in vivo occurrence. Goldman, M. J.; Fine, N. A.; Rochelle, G. T., Kinetics of N-nitrosopiperazine formation from Chandan, P. A.; Remias, J. E.; Neathery, J. K.; Liu, K., Morpholine nitrosation to better

Fine, N. A.; Goldman, M. J.; Rochelle, G. T., Nitrosamine formation in amine scrubbing at Dai, N.; Mitch, W. A., Influence of amine structural characteristics on N-nitrosamine formation

Fostås, B.; Gangstad, A.; Nenseter, B.; Pedersen, S.; Sjøvoll, M.; Sørensen, A. L., Effects of NOx Lepaumier, H.; da Silva, E. F.; Einbu, A.; Grimstvedt, A.; Knudsen, J. N.; Zahlsen, K.; Svendsen,

Mitch, W. A.; Sharp, J. O.; Trussell, R. R.; Valentine, R. L.; Alvarez-Cohen, L.; Sedlak, D. L., N-

Ridd, J. H., Nitrosation, diazotisation, and deamination. Q. Rev. Chem. Soc. 1961, 15 (4), 41826

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 36

488

(44)

Ford, P. C.; Lorkovic, I. M., Mechanistic aspects of the reactions of nitric oxide with transition-

489

metal complexes. Chem. Rev. 2002, 102 (4), 993-1018.

490

(45)

491

Nitroso Compounds Analysis and Formation, Walker, E. A.; Bogovski, P.; Criciute, L., Eds. International

492

Agency for Research on Cancer: Lyon, 1976.

493

(46)

Wilkinson, G., The long search for stable transition metal alkyls. Science 1974, 185, 109-112.

494

(47)

Maltz, H.; Grant, M. A.; Navaroli, M. C., Reaction of nitroprusside with amines. J. Org. Chem.

495

1971, 36 (2), 363-364.

496

(48)

497

copper-catalysed reaction of nitric oxide and diethylamine (I). Recl. Trav. Chim. Pays-Bas 1965, 84 (3),

498

357-371.

499

(49)

500

water. Factors controlling the formation of 1,1-dimethylnitrosamine. Chemosphere 1984, 13 (4), 549-559.

501

(50)

502

air and hydrogen peroxide. Chemosphere 1994, 29 (7), 1577-1590.

503

(51)

504

wastewater. American Public Health Association, American Water Works Association, and Water

505

Environment Association: Washington DC, USA, 1998.

506

(52)

507

Anal. Chem. 1969, 41 (12), 1709-1710.

508

(53)

National Primary Drinking Water Regulations. http://water.epa.gov/drink/contaminants/#List

509

(54)

Kittel, J.; Fleury, E.; Vuillemin, B.; Gonzalez, S.; Ropital, F.; Oltra, R., Corrosion in

510

alkanolamine used for acid gas removal: From natural gas processing to CO2 capture. Mater. Corros.

511

2012, 63 (3), 223-230.

512

(55)

513

of 2-ethanolamine: The effect of oxygen concentration and temperature on product formation. Int. J.

514

Greenh. Gas. Con. 2013, 18, 88-100.

515

(56)

516

2014, 63, 830-847.

517

(57)

518

(5), 1147-1157.

519

(58)

520

Degradation in the Presence of CO2. Ind. Eng. Chem. Res. 2009, 48 (20), 9061-9067.

521

(59)

522

Mechanisms. Ind. Eng. Chem. Res. 2009, 48 (20), 9068-9075.

Keefer, L. K., Promotion of N-nitrosation reactions by metal complexes. In Environmental N-

Brackman, W.; Smit, P. J., Studies in homogeneous catalysis: Kinetics and mechanism of the

Banerjee, S.; Pack Jr, E. J.; Sikka, H.; Kelly, C. M., Kinetics of oxidation of methylhydrazines in Lunn, G.; Sansone, E. B., Oxidation of 1,1-dimethylhydrazine (UDMH) in aqueous solution with Cleceri, L.; Greenberg, A.; Eaton, A., Standard methods for the examination of water and

Weiland, R. H.; Trass, O., Titrimetric determination of acid gases in alkali hydroxides and amines.

Vevelstad, S. J.; Grimstvedt, A.; Elnan, J.; da Silva, E. F.; Svendsen, H. F., Oxidative degradation

Fine, N. A.; Rochelle, G. T., Absorption of nitrogen oxides in aqueous amines. Energy Procedia Smith, P. A.; Loeppky, R. N., Nitrosative cleavage of tertiary amines. J. Am. Chem. Soc. 1967, 89 Lepaumier, H.; Picq, D.; Carrette, P.-L., New Amines for CO2 Capture. I. Mechanisms of Amine Lepaumier, H.; Picq, D.; Carrette, P.-L., New Amines for CO2 Capture. II. Oxidative Degradation 27

ACS Paragon Plus Environment

Page 29 of 36

Environmental Science & Technology

523

(60)

Wang, T.; Jens, K.-J., Oxidative degradation of aqueous 2-amino-2-methyl-1-propanol solvent for

524

postcombustion CO2 capture. Ind. Eng. Chem. Res. 2012, 51 (18), 6529-6536.

525

(61)

526

oxide reduction of copper(II) complexes and ligand nitrosation. Inorg. Chem. 2011, 50 (23), 11868-

527

11876.

528

(62)

529

2010.

530

(63)

531

solutions containing piperazine: a low‐energy collisional behaviour study. Rapid Commun. Mass

532

Spectrom. 2010, 24 (24), 3567-3577.

533

(64)

534

for CO2 capture. IV. Degradation, corrosion, and quantitative structure property relationship model. Ind.

535

Eng. Chem. Res. 2012, 51 (18), 6283-6289.

536

(65)

537

experimental setup on amine degradation. Int. J. Greenh. Gas. Con. 2014, 28 (0), 156-167.

538

(66)

539

blended amines for CO2 capture. Int. J. Greenh. Gas. Con. 2015, 39, 329-334.

540

(67)

541

species in coal combustion flue gas. Environ. Sci. Technol. 2002, 36 (7), 1561-1573.

542

(68)

Environmental impact of solvent scrubbing of CO2; 2006/14; IEA GHG Programme: 2006.

543

(69)

Sexton, A. J.; Rochelle, G. T., Catalysts and inhibitors for MEA oxidation. Energy Procedia 2009,

544

1 (1), 1179-1185.

545

(70)

546

and corrosion using multifunctional additive. Energy Procedia 2014, 63 (0), 814-821.

Kalita, A.; Kumar, P.; Deka, R. C.; Mondal, B., Role of ligand to control the mechanism of nitric

Ignarro, L. J., Nitric Oxide: Biology and Pathobiology. 2nd ed.; Academic Press Burlington, MA, Jackson, P.; Attalla, M. I., N‐Nitrosopiperazines form at high pH in post‐combustion capture

Martin, S.; Lepaumier, H.; Picq, D.; Kittel, J.; de Bruin, T.; Faraj, A.; Carrette, P.-L., New amines

Vevelstad, S. J.; Grimstvedt, A.; Knuutila, H.; da Silva, E. F.; Svendsen, H. F., Influence of Voice, A. K.; Hill, A.; Fine, N. A.; Rochelle, G. T., Nitrosamine formation and mitigation in Pavageau, M.-P.; Pécheyran, C.; Krupp, E. M.; Morin, A.; Donard, O. F. X., Volatile metal

Chandan, P. A.; Rogers, F.; Bhatnagar, S.; Landon, J.; Liu, K., Minimizing solvent degradation

547 548

28

ACS Paragon Plus Environment

Environmental Science & Technology

Monoethanolamine (MEA)

Industry Benchmark Amine Order

Page 30 of 36

Substitution

Steric Hindrance

Chain Length

Ethylenediamine (EDA)

Morpholine (MOR) 2-Amino-2-methyl1-propanol (AMP)

Diethanolamine (DEA)

Triethanolamine (TEA)

Glycine (GLY)

2-Amino-2(hydroxymethyl)-1,3propanediol (TRIS)

3-Aminopropanol (3AP)

β-Alanine (ALA)

549 550 551 552

Figure 1. Structures of the nine model solvent amines exhibiting four different types of structural variations from the benchmark solvent monoethanolamine (MEA).

553

29

ACS Paragon Plus Environment

Page 31 of 36

Environmental Science & Technology

Total N-Nitrosamine (µmol)

20

5 µM

No Cu

100 µM

15

10

5

0

0

2

4 Time (h)

6

8

554 555 556 557

Figure 2. Accumulation of total N-nitrosamines in an absorber experiment at 48 °C during which dissolved Cu was spiked into the 2.5 M MEA solvent. The flue gas composition was at its baseline condition.

558 559

30

ACS Paragon Plus Environment

Environmental Science & Technology

2 1.5

561 562 563 564

0.041

0.5

0.062

1

0 560

0 µM Cu 5 µM Cu 100 µM Cu

0.020 0.022 0.076

Total N-Nitrosamine Formation Rate (nmol/s)

2.5

Page 32 of 36

0% 6% 14 % Oxygen in Flue Gas (%)

Figure 3. Formation rates of total N-nitrosamines in the absorber experiments at 48 °C as a function of Cu and O2 concentrations. All experiments were performed with 2.5 M MEA and, with the exception of O2, the baseline flue gas composition. Error bars represent standard deviations from replicate experiments (n = 2 – 4).

565 566

31

ACS Paragon Plus Environment

Page 33 of 36

Environmental Science & Technology

567

570 571

0.1 0.01

12

(b)

No Cu

100 M Cu

10 8 6 4 2 0

MEA MOR DEA TEA AMP TRIS EDA GLY ALA 3AP

100 M Cu

Nitrite Formation Rate (nmol/s)

No Cu

14

1

0.001

568 569

(a)

MEA MOR DEA TEA AMP TRIS EDA GLY ALA 3AP

Total N-Nitrosamine Formation Rate (nmol/s)

10

Figure 4. Formation rates of total N-nitrosamines (panel a) and nitrite (panel b) in the absorber experiments at 48 °C. All experiments were performed with 2.5 M amine and the baseline flue gas composition. Error bars represent standard deviations from replicate experiments (n = 2 – 4).

572 573 574 575

32

ACS Paragon Plus Environment

Environmental Science & Technology

20

578

100 µM Cu

100 µM Cu + 1 mM EDTA

15 10 5 0

576 577

Baseline

Total N-Nitrosamine (µmol)

25

Page 34 of 36

0

2

4 Time (h)

6

8

Figure 5. Formation of total N-nitrosamines over time from 2.5 M MEA with the baseline flue gas composition in an absorber experiment at 48 °C.

579

33

ACS Paragon Plus Environment

Page 35 of 36

Environmental Science & Technology

580

Final Concentration of Total N-Nitrosamine (µM)

12

25 mM MOR

10 8 6 4 2 0

581

2.5 M MEA

Blank Cu Fe Mn Zn

Ni

Cr

582

Figure 6. Final concentrations of total N-nitrosamines from the pressure-cooker treatment (2 bar,

583

120 °C, 6 hours) starting with 2.5 M MEA or 25 mM MOR with 100 µM of various metals at pH 10.4. Horizontal lines provide visual guides for the N-nitrosamine formation in the metal-free control experiments. Initial nitrite = 5 mM. Final nitrite = 3.75 mM for MEA and 5 mM for MOR. The CO2 loadings (as NaHCO3 added) were 0.5 and 0.2 for MEA and MOR respectively. Error bars represent the range from duplicate experiments.

584 585 586 587 588

34

ACS Paragon Plus Environment

Environmental Science & Technology

589

Table of Contents Art

590

35

ACS Paragon Plus Environment

Page 36 of 36