Anthropogenic Perchlorate Increases since 1980 in the Canadian

Dec 22, 2017 - Jaeglé , L.; Yung , Y. L.; Toon , G. C.; Sen , B.; Blavier , J.-F. Balloon observations of organic and inorganic chlorine in the strat...
0 downloads 0 Views 1MB Size
Subscriber access provided by AUSTRALIAN NATIONAL UNIV

Article

Anthropogenic Perchlorate Increases since 1980 in the Canadian High Arctic Vasile I. Furdui, Jiancheng Zheng, and Andreea Furdui Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b03132 • Publication Date (Web): 22 Dec 2017 Downloaded from http://pubs.acs.org on December 27, 2017

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

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

Page 1 of 33

Environmental Science & Technology

1

Anthropogenic Perchlorate Increases since 1980 in

2

the Canadian High Arctic

3

Vasile I. Furdui1*, Jiancheng Zheng2,3, Andreea Furdui1

4

1

5

M9P 3V6, Canada

6

2

Geological Survey Canada, LMS, Natural Resources Canada, Ottawa, K1A 0E8, Canada

7

3

Department of Earth and Environmental Sciences, University of Ottawa, Ottawa, K1N 6N5,

8

Canada

9

*Corresponding author e-mail: [email protected]

10

Ontario Ministry of the Environment and Climate Change, 125 Resources Road, Toronto, ON,

KEYWORDS: Perchlorate, Arctic, Ice Core, Agassiz, Anthropogenic Source

11

12

ABSTRACT

13

An ice core of 15.5 meters retrieved from Agassiz Ice Cap (Nunavut, Canada) in April 2009 was

14

analyzed for perchlorate to obtain a temporal trend in the recent decades and to better understand

15

the factors affecting High Arctic deposition. The continuous record dated from 1936 to 2007,

16

covers the periods prior to and during the major atmospheric releases of organic chlorine species

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 33

17

that affected the stratospheric ozone levels. Concentrations and yearly fluxes of perchlorate and

18

chloride showed a significant correlation for the 1940-1959 period, suggesting a predominant

19

tropospheric formation by lightning. While concentration of chloride remained unchanged from

20

1940s until 2009, elevated levels of perchlorate were observed after 1979. A lack of significant

21

increases in either sulfate or chloride between 1980 and 2001 suggests that the effect of volcanic

22

activities on the perchlorate at the study site during this period could be insignificant. Therefore,

23

the elevated perchlorate in the ice could most likely be attributed to anthropogenic activities that

24

influenced perchlorate sources and formation mechanisms after 1979. Our results show that

25

anthropogenic contribution could be responsible for 66% of perchlorate found in the ice.

26

Although with some differences in trends and amounts, deposition rate found in this study is

27

similar to those observed at Devon Island (Nunavut, Canada), Eclipse Icefield (Yukon, Canada)

28

and Summit Station (Greenland). Methyl chloroform, a chlorinated solvent largely used after

29

1976, peaked in the atmosphere in 1990 and has a much shorter atmospheric life than

30

chlorofluorocarbons (CFCs). This study proposes methyl chloroform (CH3CCl3) as the

31

significant anthropogenic source of perchlorate in the Canadian High Arctic between 1980 and

32

2000, with HCFC-141b (Cl2FC-CH3), a relatively short-lived CFC probably responsible for a

33

slower decrease in perchlorate deposition after the late 1990s. The presence of aerosols in the

34

stratosphere appears to suppress perchlorate production after 1974. As both methyl chloroform

35

and HCFC-141b had no new significant emissions after 2003, deposition of perchlorate in High

36

Arctic is expected to remain at pre-1980 levels.

37 38

ACS Paragon Plus Environment

2

Page 3 of 33

Environmental Science & Technology

39

INTRODUCTION

40

Perchlorate, a stable anion in aqueous solution, is a widespread contaminant in surface and

41

ground water that can inhibit the iodide uptake by the thyroid 1, 2. Anthropogenic sources of

42

perchlorate contamination in surface and ground water include the production and usage of solid

43

rocket propellant, blasting agents, explosives, fireworks, road flares and its presence as an

44

impurity in Chilean fertilizer, sodium chlorate and bleach solutions 3, 4. The occurrence of

45

perchlorate in High Arctic precipitation 5 and Antarctic soil 6 was reported almost simultaneously

46

in early 2010. The first High Arctic study on perchlorate was carried out by analyzing samples

47

collected from a snow pit from Devon Island (Nunavut, Canada) covering 1996-2005 period 5.

48

Furdui and Tomassini 5 confirmed the existence of naturally formed perchlorate in a 2000-year

49

sample collected from the Agassiz Ice Cap (Nunavut, Canada) in the High Arctic. Occurrence of

50

natural perchlorate and its atmospheric deposition as a source was reported in salt deposits from

51

Atacama desert 7-9, desert soils 10, 11, ground water 10, 12, Mars 13, Moon and asteroids 14.

52

Atmospheric Formation of Perchlorate. In 1975, Simonaitis and Heicklen suggested that

53

HClO4 may be a more efficient sink than HCl for stratospheric chlorine 15. Jaeglé et al. 16 found

54

in 1996 a significant discrepancy between the total inorganic chlorine and the sum of measured

55

HCl, ClONO2 and HOCl in the stratosphere. They predicted HClO4 concentrations of 8–15 ppt at

56

16–19 km altitudes for a post-Pinatubo volcanic activity enhanced stratospheric aerosol levels.

57

Murphy and Thompson 17 confirmed the presence of 0.5–5 ppt ClO4- at 19 km altitude, while no

58

detectable perchlorate was found even though larger chlorine peaks were detected in the

59

tropospheric aerosols. Dasgupta et al.

60

precursors, chlorine radicals and chloride ions.

9

proposed two possible atmospheric perchlorate

ACS Paragon Plus Environment

3

Environmental Science & Technology

61

Page 4 of 33

Previous Ice Core Studies. The deposition pattern for various species relevant to this study, 18, 19

62

including perchlorate, can be reconstructed using samples collected from ice caps

63

and Tomassini observed seasonality for Devon Ice Cap samples and based on a correlation

64

between perchlorate and total ozone they considered that perchlorate was formed from chlorine

65

radicals in the stratosphere 5. Volcanic eruptions were proposed as a possible natural source of

66

increased atmospheric perchlorate for periods with higher deposition levels of sulfate 5. Discrete

67

ice core samples from the Eclipse Icefield, Yukon Territory, Canada

68

perchlorate and major anions, confirming their seasonal variation and concentration levels

69

comparable to those reported for Devon Island samples. An increase in perchlorate deposition

70

after 1980 observed at Eclipse Icefield was also confirmed by Rao et al.

71

non-polar ice-core from the Upper Fremont Glacier, (Wyoming, USA), even though Fremont ice

72

core contained lower perchlorate concentrations than the Eclipse ice core. For the period 2005-

73

2007, the yearly deposition rates for both ice cores were lower than the yearly deposition rates in

74

wet precipitation across North America21. Peterson et al.

75

drilled near Summit Station (Greenland) and presented perchlorate and sulfate data for the pre-

76

industrial period of 1705-1752 and the recent period of 1950-2006. Perchlorate and sulfate

77

depositions peaked simultaneously at Summit Station (Greenland) during El Chicón (1982) and

78

Pinatubo (1991) eruptions 22. Analyzing a West Antarctic Ice Sheet Divide snow pit, Crawford

79

et al.

80

ozone hole did not result in increased perchlorate concentrations, confirming the stratospheric

81

perchlorate origin during this period and the directly proportional relationship between ozone

82

and perchlorate observed in Devon snow pit 5. Jiang et al. recently investigated the perchlorate in

83

a South Pole firm core covering the 1920-2004 period 24 and observed an increase in perchlorate

23

22

20

20

. Furdui

were analyzed for

studying a discrete

analyzed samples from an ice core

found that penetration of UV radiation into the troposphere during the stratospheric

ACS Paragon Plus Environment

4

Page 5 of 33

Environmental Science & Technology

20, 22

84

after 1980, confirming the trend observed in the Arctic

. Perchlorate at the South Pole was

85

strongly correlated with the equivalent effective stratospheric chlorine (EESC), which peaked in

86

1996-1997, although perchlorate had three separate maximums between 1990 and 2004,

87

suggesting that other factors may have been involved. Spatial variability was also confirmed in

88

the Antarctic snow, with lower perchlorate concentrations observed at sites with higher

89

accumulation rates.

90

Stratospheric Organic Chlorine Species. Most chlorine enters the stratosphere as organic

91

chlorine species, which are converted to inorganic chlorine species like HCl, ClONO2 and ClO

92

through photochemical oxidation 25. While the majority of organic chlorine compounds found in

93

the atmosphere have anthropogenic sources, methyl chloride

94

chlorinated hydrocarbon released by marine algae, plants, fungi 27 and biomass burning 28. Other

95

natural chlorine-containing trace atmospheric gases are dichloromethane (CH2Cl2), chloroform

96

(CHCl3), trichloroethylene (ClCH=CCl2) and perchloroethylene (Cl2C=CCl2), all found together

97

in 1999 at 120 parts per trillion by volume (pptv), while methyl chloride was found alone at 540

98

pptv 29. All natural chlorinated hydrocarbons have atmospheric lifetimes shorter than two years

99

29

26

is the dominant natural

. Molina and Rowland identified in 1974 the potential of chlorofluorocarbons (CFCs) to

100

photodissociate and release chlorine atoms in the stratosphere 30, with destructive effects on the

101

stratospheric ozone. Main CFCs found in the stratosphere include CFC-11 (CCl3F), CFC-12

102

(CCl2F2), CFC-113 (Cl2FC-CClF2), all extremely stable with atmospheric lifetimes longer than

103

50 years. Chlorinated solvents are also important anthropogenic contributors of atmospheric

104

organic chlorine, with methyl chloroform (CH3CCl3) as the dominant compound largely used

105

after 1976 as a tropospheric inert replacement to trichloroethylene

106

organically bound chlorine in the atmosphere was of anthropogenic origin

31

. By 1980, 78% of 32

. Total organic

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 33

107

chlorine peaked in the troposphere in 1993-1994, declining initially at a rate of 1% per year and

108

reducing to a rate of 0.5% in recent years

109

chlorine occurred during the late 1990s as a result of the implementation of the 1987 Montreal

110

Protocol, which targeted production of ozone-depleting substances. Peak stratospheric total

111

chlorine occurred between 1996 and 1997 33, 34.

112

The objective of this study is to identify the perchlorate precursors and factors responsible for the

113

increases of perchlorate observed from 1980 in Arctic and Antarctic ice core studies. It is

114

important to identify whether natural (e.g. volcanic activity) or anthropogenic (e.g. increased

115

organic chlorine species in the atmosphere) sources have the main influence on the observed

116

deposition rates. Meanwhile, a proper identification of the precursor chlorine species will allow

117

development of models to better predict future atmospheric perchlorate trends. This study reports

118

a continuous long-term temporal trend of perchlorate in Canadian High Arctic snow with

119

simultaneous measurements of chlorate and chloride, covering an atmospheric deposition period

120

from 1936 to 2007 at the Agassiz Ice Cap. The influence of the stratospheric aerosol

121

concentration on perchlorate deposition was analyzed using historical records of the stratospheric

122

optical depth 35, 36.

33, 34

. An observed decrease in stratospheric total

123 124

EXPERIMENTAL SECTION

125

Field Sampling and Laboratory Sample Processing

126

In late April 2009, two parallel, approx. 15.5-meter short cores, about 1.5 m apart, were retrieved

127

from the Agassiz Ice Cap (80.7ºN and 73.1 ºW at 1670 m above sea level) located in Nunavut,

ACS Paragon Plus Environment

6

Page 7 of 33

Environmental Science & Technology

128

Canada as shown in Figure 1. For orientation purpose, sites of previous studies are also included

129

in the figure. One core was used for archive reconstruction of total mercury, plutonium, ion

130

chemistry and perchlorate, while the other was used to examine density and melt

131

percentage/stratigraphy for more precisely setting up depth–age relationship. A larger 14 cm

132

barrel corer was used in order to achieve a thorough decontamination by removing outer layers.

133

The core for archive reconstruction was split on site into two halves, with one half for total

134

mercury quantification. The second half was carefully packed, section by section, in double

135

layered lay-flat tubes and stored on site in three coolers, buried in a snow trench at -19 ºC or

136

lower. Once samples were returned to Resolute Bay, they were stored frozen in a walk-in freezer

137

and shipped back to the Geological Survey Canada laboratory in Ottawa via First Air frozen

138

cargo shipping. Samples were stored in Ottawa in deep freezers (-18 ºC or lower) until further

139

processing and analysis. All samples for this study were handled only with titanium and

140

polyethylene tools during processing and decontamination 37. Further sub-sampling for

141

perchlorate and ion studies was carried out in cold and room temperature clean rooms, all Class-

142

1000 with Class-100 working station/benches at the Natural Resources Canada facility in

143

Ottawa 38. A total of 287 samples were transferred to the Ministry of the Environment and

144

Climate Change laboratory in Toronto for analysis.

145 146

Instrumental Analysis

147

Certified standards for ClO4-, ClO3-, BrO3-, Br- were purchased as a custom mixture (100 mg/L)

148

from Inorganic Ventures (Christiansburg, VA, USA). Mass-labeled (18O) perchlorate and

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 33

149

individual 1000 mg/L certified reference standards of the major ions (Cl-, NO3- ,SO42- and F-)

150

were purchased from Sigma-Aldrich (Oakville, ON, Canada).

151

Major anions (Cl-, NO3- ,SO42- and F-) were analyzed using a capillary ion chromatograph (IC)

152

instrument (ICS5000, Thermo Scientific Dionex, Mississauga, ON, Canada), with following

153

modules: binary capillary pump, eluent generator module (KOH), anion capillary electrolytic

154

suppressor (ACES300) and conductivity detector. Isocratic separation (26 mM OH-) was

155

obtained in 12.5 min for 0.4 µL injection volume using a 10 µL min-1 flow rate on AG18

156

capillary guard column and AS18 capillary separation column. Analysis of trace level anions

157

(ClO4-, ClO3- and Br-) was performed using an ICS5000+ ion chromatograph (Thermo Scientific

158

Dionex, Mississauga, ON, Canada) coupled to API 3200TM triple quadrupole mass spectrometer

159

(SCIEX, Concord, ON, Canada). The IC consisted of an autosampler, dual pumps, an auxiliary

160

pump (0.5 mL min-1 for suppressor regeneration), an eluent generator (EluGen KOH cartridge)

161

and an ion suppressor (ASRS-300 2 mm). Separation was achieved in 18 minutes using 500 µL

162

injection volume, AG20 guard column (2mm i.d. x 50 mm) and AS20 separation column (2 mm

163

i.d. x 250 mm) from Dionex (Thermo Scientific). A gradient separation was used with 10 mM

164

OH- initial concentration for the IC eluent, followed after 60 s by a 3 minute ramp to 80 mM OH-

165

, an 11 minute hold and reverting to initial conditions at 16 min. The flow rate through the IC

166

was 300 µL min-1 and a post-column flow of 200 µL min-1 HPLC grade methanol (Fisher

167

Scientific) was added using a tee, prior to entering the mass spectrometer. The electrospray

168

source was operated in negative mode with corresponding parameters as follows: spray voltage -

169

3500V, gas temperature 450 °C, curtain gas 50, nebulizer gas 50 and turbo gas 60. The mass

170

spectrometer was operated in multiple reaction monitoring (MRM) mode, with the related

171

parameters optimized for each analyte as previously published 5. Analyst software version 1.6.2

ACS Paragon Plus Environment

8

Page 9 of 33

Environmental Science & Technology

172

(SCIEX) and DCMS Link version 2.12 (Thermo Scientific) were used for data acquisition and

173

quantitation. Statistical analysis was done using OriginPro® (OriginLab Corporation,

174

Northampton, MA, USA).

175

Perchlorate was the only ion measured using an internal standard approach (Cl18O4-), while the

176

other ions were measured using an external calibration curve. Analytical blank levels were below

177

detection limits (see Table S2 in Supporting Information (SI)) for both chlorate and perchlorate.

178

Two QC solutions (5 and 50 ng/L) were run after every 30 analyzed samples to monitor

179

instrument performance.

180 181

RESULTS AND DISCUSSIONS

182 183

Ice Dating and Major Ion Trends

184

Ion chemistry, ice/firn stratigraphy, density profiles, snow accumulation rates and melting layers

185

were used for dating samples. Age assignment was based on ion seasonal cycle counting and

186

confirmed by the summer melting layers. Further confirmation of the age assignment was also

187

provided by the two plutonium peaks during the years 1958/1959 and 1963. The bottom ice core

188

sample corresponds to 1936. Based on the two accurate and outstanding plutonium markers,

189

dating error was estimated to be between 6 and 12 months 37.

190

Correlation coefficients were calculated at 0.05 level using 2-tailed test of significance with

191

OriginPro® for 1936-2007, 1940-1959, 1960-1979 and 1980-2007, using both yearly fluxes and

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 33

192

concentrations. The results are presented in Tables S1-S8 (SI). Nitrate spikes correspond

193

directly with the sulfate spikes and nitrate concentrations were lower than sulfate concentrations

194

for all samples (Figure S1, SI). Although sulfate and nitrate showed a correlation for the entire

195

period of 1936-2007, the strongest correlation occurred between 1940 and 1959 (r=0.93,

196

p=2.96×10-9). Increasing anthropogenic influence on sulfate and nitrate from the 1940s until the

197

late 1980s was previously reported in studies of ice cores from Greenland, the Canadian Arctic

198

and Svalbard 39, 40 . A similar trend was observed in the current study, but with a slight decrease

199

in nitrate after the year 2000. Based on the work done by Goto-Azuma and Koerner 39, this study

200

attributes the anthropogenic sulfate and nitrate on Agassiz Ice Cap to Eurasia.

201 202

Perchlorate Trends

203

Yearly deposition rates were relatively unchanged for perchlorate until 1979, with a median of

204

0.42 µg m-2 year-1, followed by increased depositions from 1980 to 2007 with a median of 1.14

205

µg m-2 year-1 (see Figures 3 and S4, SI). Seasonal variation can be seen through the full dataset

206

plot with a resolution of 3-5 samples per year (Figure 2). Although resolution is not high, results

207

from this study are consistent with and supportive to the previous work by others for post 1980s

208

for wet deposition 21 and ice cores with higher resolution (up to 26 samples per year) 5, 20, 22. Due

209

to year-by-year variation of precipitation patterns and possible uneven distribution of

210

atmospheric deposition (e.g. scouring after deposition), seasonal variation could be more

211

pronounced in some years, such as certain pre-1980s years: 1938, 1940, 1958-1959, 1969 and

212

1971-1972 (Figures 2, S1 and S2, SI). Variability in perchlorate concentrations for the pre-1980

213

layers is higher at Agassiz than at Eclipse Icefield 20, South Pole and Antarctic Dome A 24.

ACS Paragon Plus Environment

10

Page 11 of 33

Environmental Science & Technology

214

Annual perchlorate fluxes achieved in this study along with those from previous publications are

215

presented in Figures 1 and 4. These figures include data from Devon Ice Cap (actual data set) 5

216

for 1996-2005, Eclipse Icefield (estimated from published figures) for 1970-1972, 1983-1984

217

and 2000 20, and Summit Station (estimated from published figures) for 1966-2006 22. For the

218

reported years, the highest values of perchlorate occurred for the Eclipse Icefield (Figure 4),

219

while the lowest were observed at Summit Station. During the period of 1966-2006, total

220

perchlorate deposition at Agassiz Ice Cap was twice as much as the deposition at Summit

221

Station. For the period between 1996 and 2005, total perchlorate deposition at Agassiz Ice Cap

222

accounted for only about two thirds (66%) of the total deposition at Devon Ice Cap. Location

223

specific effects on perchlorate concentration in wet precipitation were previously observed at

224

lower latitudes as well, with concentrations decreasing with relative distance from coast while no

225

relationship was observed for the total deposition 21. Yearly deposition of perchlorate at the

226

Eclipse Icefield was found to be the highest among the four locations. For the limited non-

227

continuous Eclipse record, the total deposition of perchlorate for 1970-1972, 1983-1984 and

228

2000 was 10 µg m-2, while for the same years the total deposition at Agassiz Ice Cap was 3.97

229

µg m-2 (2.5 times lower than that at Eclipse) and the total deposition at Summit Station was 2.4

230

µg m-2 (4 times lower than that at Eclipse). Currently available data indicates that perchlorate

231

deposition in the Arctic is location specific, confirming the spatial variability also observed in

232

the Antarctic snow record 24.

233

As shown in Figure 4, both Agassiz Ice Cap and Summit Station show a similar increasing trend

234

of perchlorate deposition beginning in 1979 until 1990-1992, after which it gradually decreased

235

towards the early 1970s levels. A gradual increase in perchlorate after 1970 was also observed by

236

Jiang et al. 24 at the South Pole, with a sharp increase around 1990, followed by a decrease after

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 33

237

2000. Based on the observed temporal trend of both perchlorate and nitrate (Figure S1), post

238

depositional loss of perchlorate is not suspected for Agassiz samples, in contrast to the samples

239

from South Pole 24. The median of yearly perchlorate fluxes for Agassiz Ice Cap from 1966-1978

240

was 0.48 µg m-2 (n=13), which is four times higher than the corresponding Summit Station

241

median. The median of yearly perchlorate fluxes for Agassiz Ice Cap from 1979-2001 was 1.25

242

µg m-2 (n=23), which is 2.5 times higher than the corresponding Summit Station median. Based

243

on the median values, perchlorate deposition at Agassiz Ice Cap was 2.6 times higher for 1979-

244

2001 than for 1966-1978, while Summit Station had a 3.8 times higher deposition for

245

comparison of the same periods. Total perchlorate deposition on Agassiz Ice Cap between 1979

246

and 2001 was 27.9 µg m-2, almost twice as much as the total amount deposited on Summit

247

Station (14.5 µg m-2). Based on early studies done by Koerner et al. 41, accumulation rate at the

248

Agassiz Ice Cap sampling site was 10 cm yr-1 (water equivalent), a much lower rate than at the

249

other sites used for comparison. While the accumulation rate at Summit is 21 cm yr-1 (water

250

equivalent) 42, which is slightly more than twice the rate at Agassiz, the rate at Eclipse Icefield is

251

138 cm yr-1 (water equivalent) 43, which is over 13 time higher than that at Agassiz. Furthermore,

252

there are differences in air masses received at each site. While Agassiz Ice Cap receives its air

253

masses mainly from Eurasia, Summit Station receives its sources from North America and

254

Eurasia 39 and Eclipse Ice Field receives its sources from the Pacific, especially in the winter

255

season 43. All those differences of site conditions may contribute to the differences observed in

256

perchlorate concentrations.

257 258

ACS Paragon Plus Environment

12

Page 13 of 33

Environmental Science & Technology

259

Volcanic Emissions and Perchlorate Deposition

260

Volcanic emissions were previously proposed as a factor favoring perchlorate formation 5, 20, 22.

261

The lack of significant increases in sulfate and chloride between 1980 and 2001 for Agassiz Ice

262

Core (see Figures 3, S4, S5 and S7, SI) suggest that volcanic activity had minimal influence

263

during the peak perchlorate period. As shown in Figure 3, perchlorate deposition spikes can be

264

clearly identified at 1980-1981, 1987, 1990, 1993, 1996, 1999 and 2005. Meanwhile, there was

265

no observable increase in perchlorate prior, during or after the outstanding chloride and sulfate

266

spikes (between 1972 and 1979). While chloride had the highest yearly deposition during 1975-

267

1976 (seven times higher than the average yearly chloride deposition of the entire 1936-2007

268

period), the highest deposition of sulfate was observed in 1977 (Figure 3). More careful

269

examination of the plots from Figure 3 indicates that sulfate concentration simultaneously

270

increased with perchlorate during the years 1987, 1990, 1996, 1999 and 2005, while sulfate-only

271

spikes were observed in 1982, 1985 and 1994. The sources of short-term spikes of sulfate and

272

chloride for Agassiz deposition could be volcanic eruptions from the Aleutian Islands (Augustine

273

1963, 1971, 1977, 2005; Pavlof 1983, Mount Redoubt 1989-1990), Kuril Islands (Alaid 1981)

274

and Kamchatka Peninsula (Kliuchevskoi 1987). The same volcanic influence from Alaskan,

275

Aleutian and Kamchatkan eruptions was observed in an ice core from the Eclipse Icefield,

276

Yukon Territory, Canada 44, even though sources of air masses received at Eclipse Icefield and

277

Agassiz Ice Cap may not be exactly the same. Bluth et al. 45 concluded that smaller eruptions at

278

higher latitudes can produce similar stratospheric impacts to the dense rock equivalent eruptions

279

from lower latitude volcanoes. In contrast to Agassiz, sulfate deposition at Summit Station

280

showed influence from major lower latitude volcanic activity 22, with the highest yearly

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 33

281

depositions of sulfate and perchlorate corresponding to El Chichón (1982) and Pinatubo (1992)

282

eruptions 22.

283

The Augustine volcanic eruption of 1976 released a large quantity of chlorine, with a significant

284

fraction being injected into the stratosphere as HCl 46. However, HCl injected into the

285

stratosphere is usually washed out immediately by water from the volcanic plumes 25. The 1976-

286

1977 chloride peak at Agassiz was not observed closer to the Augustine volcano at Eclipse

287

Icefield 44, suggesting that a volcanic source of chloride during this period is improbable. The

288

high level of Na+ observed during the period of 1976-1977 (unpublished data from 37) makes sea

289

salt the most likely source for the peak chloride levels. Our data suggest that volcanic activity

290

may not significantly increase perchlorate deposition on Agassiz Ice Cap even though mid-

291

latitude volcanic eruptions can be responsible for increased perchlorate levels in local

292

precipitation.

293 294

Chlorate in High Arctic Depositions

295

Chlorate levels were similar to the perchlorate levels for the entire period of 1936-2007, except

296

for two chlorate peaks observed in 1943 and 1990 (Figures 2 and S1, SI). Chlorate is less stable

297

than perchlorate, but even the oldest samples in this study, had chlorate and perchlorate at similar

298

levels. There are, however, two time periods, one in the late 1970s and one in the early 1990s,

299

when chlorate exceeded perchlorate in both concentrations and yearly fluxes. Murphy and

300

Thompson 17 confirmed existence of perchlorate in 1998 stratospheric aerosols, and they

301

considered the observed ClO-, ClO2- and ClO3- ions in the mass spectra as fragments from ClO4-,

302

but chlorate was probably the major source of the observed ClO3- ions. However, all these ClOx-

ACS Paragon Plus Environment

14

Page 15 of 33

Environmental Science & Technology

303

fragments were only observed in the lower stratosphere, but not in the troposphere. A

304

stratospheric source for chlorate and perchlorate is also supported by the correlation observed

305

between chlorate and perchlorate between 1980 and 2007 (r=0.64, p=2.69×10-4 for yearly fluxes

306

and r=0.49, p=9.3×10-9 for concentrations), although this was not confirmed for the 1996–2005

307

Devon ice record 5. In aqueous phase experiments, chlorate and perchlorate were formed

308

simultaneously from chlorine and oxy-chlorine species 47. Jaeglé et al. 16 suggested that HClO4

309

was responsible for 20-60% of the difference between total inorganic chlorine and measured

310

species (HCl, ClONO2 and HOCl). However, considering that both chlorate and perchlorate

311

seem to have stratospheric sources and are present at similar concentrations in Arctic ice core

312

samples (Agassiz and Devon 5), chlorate might be equally important to perchlorate for accurate

313

stratospheric chlorine mass balance.

314 315

Sources of Perchlorate in High Arctic

316

There are no spikes in the temporal distribution of perchlorate deposition at Agassiz site that can

317

be associated to the PEPCON perchlorate factory explosion in 1988 or the Space Shuttle

318

program (see page SI-22, SI). For example, during the period between 1981 and 1985, the

319

number of space shuttle launches constantly increased, but perchlorate spiked in 1980-1981,

320

followed by lower depositions rates until 1985. As there is no known use of perchlorate in the

321

High Arctic, atmospheric formation should be entirely responsible for the perchlorate measured

322

in the ice cores. For 1940–1959 and 1980–2007 time periods, significant correlations were

323

observed respectively for yearly fluxes of perchlorate-nitrate (see Tables S1-S8, SI), suggesting

324

that chlorine nitrate (ClONO2), known to strongly couple stratospheric chlorine and nitrogen

325

cycles 48, may be involved in perchlorate formation. However, due to the much higher

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 33

326

concentration of nitrate than perchlorate, other anthropogenic sources of nitrate have to be

327

considered. As presented by Catling et al. 49, nitric acid is similar to perchloric and sulfuric acids,

328

being the terminal condensed phase that result from the photochemical oxidation of the nitrogen

329

bearing gases.

330

Yearly fluxes of chloride and perchlorate showed significant correlation at the 0.05 level of

331

significance (n=20, Pearson correlation, 2-tailed test, r =0.88 and p=2.55 ×10-7) only for values

332

between 1940 and 1959 (Figure S3, SI). This was supported by a significant correlation between

333

chloride and perchlorate concentrations for the same period (n=71, Pearson correlation, 2-tailed

334

test, r =0.56 and p=4.06 ×10-7). This period is characterized by constant, low stratospheric

335

aerosol levels (refer to Northern Hemisphere optical depth values in Figure 7) and relatively low

336

level emissions of anthropogenic organic chlorine compounds 34, including methyl chloroform

337

(Figure 8) and other compounds 32. As there are no known common atmospheric sources of

338

chloride and perchlorate, this suggests that perchlorate was formed from a chloride ion precursor,

339

although a very small fraction of chloride seems to be converted to perchlorate. Laboratory

340

simulated lightning produced perchlorate at 3 orders of magnitude lower than the input Cl-

341

concentration, with moisture inhibiting formation of ClO4- 50. Lower energy storms are less

342

efficient in formation of ClO4-, possibly explaining the 4 orders of magnitude difference

343

observed in this study between perchlorate and Cl-.

344

The significant chloride-perchlorate correlation observed for the 1940-1959 indicates that

345

chloride was the predominant tropospheric source of perchlorate during this period, while in the

346

following years other precursors/pathways became more relevant. A similar chloride-perchlorate

347

correlation previously observed in a 1996-2005 subset of samples corresponding to summer peak

ACS Paragon Plus Environment

16

Page 17 of 33

Environmental Science & Technology

348

ClO4- concentration samples collected from Devon Island 5 confirmed the presence of this

349

tropospheric perchlorate pathway in Arctic depositions from recent years.

350

A standard deviation (SD) of 0.068 was calculated for the differences between measured and

351

estimated fluxes for 1940-1959 with ±3×SD limits, which are presented in Figure 5. Based on

352

measured yearly chloride fluxes (Figures 3 and S3, SI), a much higher level of perchlorate would

353

be expected in 1975-1976 and a much lower level in the following years (Figure 5). For the

354

unusually high level of chloride in 1975-1976, with elevated Na+ levels confirming the sea salt

355

origin 37, chloride was not significantly associated with an increase in perchlorate. A local source

356

of chloride alone is not sufficient for perchlorate formation in the High Arctic where

357

thunderstorms and lightning flashes, which are the key factors for this formation, are unusual 51.

358

We believe this is the reason why no perchlorate increase corresponds to the outstanding chloride

359

flux peak found in 1975-1976 5, 9. After 1980, the measured perchlorate flux was much higher

360

than the predicted perchlorate flux based on the relationship between perchlorate and chloride

361

from the data between 1940 and 1959. For the period of 1980-2001, the total measured

362

deposition of perchlorate was 26.9 µg m-2 while the predicted deposition was 9 µg m-2, which

363

indicates that 66% of perchlorate was formed from a different precursor or mechanism compared

364

to the one involved during the period of 1940-1959. While chloride levels remained mostly

365

unchanged compared to those from the 1940s, there was an increase of more than 100% in

366

perchlorate fluxes during the 1980s and 1990s (Figure 6). An even higher increase in perchlorate

367

(250%) and chlorate (400%) occurred in 1990s, while chloride levels remained nearly unchanged

368

since the 1940s. These patterns suggest that environmental conditions changed after 1979, either

369

favoring the atmospheric production of perchlorate from chloride or by the emergence of a new

370

precursor.

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 33

371 372

Influence of the Aerosols on Perchlorate Deposition

373

Halogens, as well as 30 other elements, were found on aerosols collected from the upper

374

troposphere and stratosphere 52, with some elements possibly having catalytic activity towards

375

perchlorate formation. After volcanic eruptions, the concentration of aerosols in the atmosphere

376

usually increases and their variation can be monitored by measuring optical depth. The

377

stratospheric optical depth values at 550 nm available for the Northern Hemisphere 35, 36, are

378

shown in Figure 7 together with perchlorate and chlorate fluxes. While the highest stratospheric

379

optical depth peaks were observed in 1983 and 1992, perchlorate deposition peaked two years

380

earlier in each case (1981 and 1990), followed by lower perchlorate depositions in 1982-1984

381

and 1991-1995. The yearly perchlorate values have 6 – 12 months dating error and the monthly

382

optical depth values have a much smaller dating error. Therefore, this suggests that the increase

383

in perchlorate levels observed for the 1980-2000 period occurred prior to the years with

384

enhanced aerosol levels. In the case of the Summit Station record, perchlorate depositions during

385

the two peaks of stratospheric optical depth (1983 to 1985 and 1992 to 1993) were lower than in

386

the corresponding previous years (Figure 4). The lack of any linear relationship between the

387

Northern Hemisphere stratospheric optical depth and perchlorate indicate that other factors are

388

involved in the perchlorate deposition at Aggasiz. The effect of elevated aerosol on chlorate

389

deposition is unclear.

390 391

ACS Paragon Plus Environment

18

Page 19 of 33

Environmental Science & Technology

392

Organic Chlorine Species as Perchlorate Precursors

393

During the preindustrial era, the short-term input of atmospheric chloride were sea salt and

394

volcanic eruptions 19. As a natural organic source of chlorine, methyl chloride was released at

395

relatively constant levels during and after preindustrial years 32. The naturally emitted methyl

396

chloride may be a stratospheric perchlorate precursor since its atmospherically released atomic

397

chlorine enters the ClOx cycle 48. After 1960, the content of organic chlorine in the atmosphere

398

began to increase, reaching a maximum level in the troposphere between 1993 and 1994 and a

399

maximum level in stratosphere between 1996 and 1997 33, 34 (Figures S15-S16, SI). This was

400

followed by a decrease in atmospheric organic chlorine, although at a much slower rate (0.5-1%

401

per year). This relatively small decrease in organic chlorine alone cannot explain the decrease in

402

perchlorate that occurred by the year 2002, when perchlorate deposition temporarily reached

403

1977-1979 levels. Emissions of CFCs began to decrease in the 1990s, however, those already

404

emitted could remain in the stratosphere for the next 40-150 years due to their relatively long

405

lifetime.

406

Methyl chloroform (CH3CCl3) is one of the few organic chlorine compounds that experienced a

407

sharp decrease in emissions beginning in the early 1990s 53. This decrease in emissions was

408

immediately followed by a decrease in atmospheric concentrations of CH3CCl3 due to its much

409

shorter lifetime (5 year) 34, as observed at Alert (Nunavut, Canada). Methyl chloroform was

410

used as a solvent after 1955 54 and its production increased in 1976 when it replaced

411

trichloroethylene as a tropospheric inert solvent 31. By 2011, the atmospheric burden of methyl

412

chloroform declined to 5% of its peak value reached in 1990 55. Based on a 15% transfer into the

413

stratosphere, where through photolysis methyl chloroform releases Cl, McConnell and Schiff 31

414

estimated a 20% ozone depletion potential from this solvent as compared with the ozone

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 33

415

depletion of CFCs released at 1973 rates. Another short-lived CFC with similar structure to

416

methyl chloroform (Cl3C-CH3) is HCFC-141b (Cl2FC-CH3). The emissions of HCFC-141b

417

peaked around 2000, decreasing to no emissions after 2006. It was measured at Alert (Nunavut,

418

Canada)

419

HCFC-141b had no new significant emissions after 2003. Kutsuna et al. 56 identified the

420

potential degradation of methyl chloroform to CH2=CCl2 by losing HCl on aluminosilica clay

421

minerals when illuminated with wavelengths longer than 300 nm. Kutsuna et al. therefore,

422

suggested its potential as a tropospheric sink, since CH2=CCl2 is very short-lived (one day) in the

423

atmosphere 34. Based on this new mechanism, when aerosol particles are present, methyl

424

chloroform is removed from the atmosphere, potentially reducing Cl release and ozone depletion

425

in the stratosphere. While other factors can be also involved, the described methyl chloroform

426

decomposition may have played a role in the lower perchlorate deposition levels observed during

427

the years with peak aerosol levels (stratospheric optical depth peaks for 1975, 1979, 1982-1984

428

and 1991-1994, Figure 7). The same effect was also observed at Summit Station 22 after 1983

429

and 1991 (Figure 4). As the stratospheric optical depth did not reach the same high levels prior to

430

1974, the effect of aerosols on perchlorate formation is harder to observe under lower methyl

431

chloroform emissions. In 1969, when optical depth values were comparable with 1975 level and

432

methyl chloroform emissions were lower (50% of 1975 level and 25% of 1991 level) 57, there is

433

no indication of suppression on perchlorate deposition, suggesting that after 1974 the aerosols

434

were able to efficiently remove chlorine species involved in perchlorate formation. Perchlorate

435

deposition pattern at Agassiz Ice Cap is similar to the methyl chloroform emission pattern

436

(Figure 8) but with a 4-6 years delay because methyl chloroform has an average retention time of

53

showing an increasing trend from 1993 (Figure 8). Both methyl chloroform and

ACS Paragon Plus Environment

20

Page 21 of 33

Environmental Science & Technology

437

5 years in the atmosphere 34. The same difference (Figure 8) is also observed between methyl

438

chloroform emissions (1990-1996) and measured concentration at Alert (1995-2001).

439

Jiang et al. 24 found that the perchlorate in the Antarctic snow follows the variation of the

440

equivalent effective stratospheric chlorine (EESC) , however, this cannot be confirmed in the

441

High Arctic. The Southern Hemisphere experienced different patterns of aerosol exposure, with a

442

much lower aerosol peak between 1982 and 1985, though the peak between 1991 and 1996 is

443

similar between the two Hemispheres. The much lower methyl chloroform concentrations

444

measured in the 1970s 54 in the Southern Hemisphere can be also responsible for slower

445

increases of perchlorate deposition in Antarctic snow compared to the situation in High Arctic

446

snow between 1970 and late 1980s.

447

Since 1996, methyl chloroform emissions have dropped to pre-1970 levels, with no significant

448

emissions of methyl chloroform and HCFC-141b by 2003. Assuming no major increases in

449

emissions of CFCs with atmospheric lifetimes between 0.5 and 10 years, the perchlorate

450

deposition in the Arctic is not expected to increase back to 1990 levels observed at Agassiz and

451

Summit Station.

452 453

ASSOCIATED CONTENT

454

Yearly depositions, correlation coefficients and scatter matrices for yearly fluxes and

455

concentrations of perchlorate, chlorate, chloride, sulfate, nitrate, bromide and fluoride are

456

included in the Supporting Information file.

457

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 33

458

ACKNOWLEDGMENT

459

Joe McConnell from Desert Research Institute (Reno, NV, USA) is thanked for providing

460

plutonium information. We thank our three anonymous reviewers for their helpful reviews of the

461

manuscript. Fengrong Sun and Robert Tooley from the Ontario Ministry of the Environment and

462

Climate Change, are thanked for their help in capillary ion chromatography analysis and

463

manuscript review, respectively.

464 465 466 467 468 469 470

Figure 1. Map showing the current sampling location (Agassiz) and locations used in previous

471

studies (Devon 5, Eclipse Icefield 20 and Summit Station 22). Total depositions of perchlorate are

472

calculated for the period of 1966-2006 (black) and 1996-2005 (red). Map data: Google,

473

DigitalGlobe.

474

Figure 2. Concentrations of perchlorate and chlorate found in ice core samples from Agassiz Ice

475

Cap, Nunavut, Canada.

ACS Paragon Plus Environment

22

Page 23 of 33

Environmental Science & Technology

476

Figure 3. Yearly fluxes of perchlorate, chloride and sulfate from Agassiz. The highest chloride

477

deposition was found during 1975-1976, seven times higher than average of the entire studied

478

time period from 1936 to2007 while the highest deposition of sulfate was observed in 1977.

479

Figure 4. Yearly flux of perchlorate from current study (Agassiz), and from previous

480

publications, (calculated yearly fluxes from Devon 5, estimated yearly fluxes from Eclipse

481

Icefield 20 and Summit, Greenland 22). Stratospheric optical depth from Northern Hemisphere

482

from Sato et al. 35 and NASA Goddard Institute for Space Studies 36

483

Figure 5. Difference in perchlorate yearly fluxes between measured and estimated values using

484

an equation derived from the 1940-1959 values (red dots). Chloride from local sea salt source

485

does not form perchlorate under High Arctic conditions (excluded values for 1975 and 1976),

486

while non-chloride sources are responsible for large differences observed after 1980. The two red

487

lines represent the ±3 standard deviations of the 1940-1959 difference values.

488

Figure 6. Decadal changes in total depositions, calculated from 1940-1949 total deposition for

489

perchlorate, chlorate, chloride, sulfate and nitrate. Total depositions for 2000s are based on 8

490

years yearly fluxes as no data is available for 2008 and 2009.

491

Figure 7. Trend of Stratospheric Optical Depth at 550 nm from Northern Hemisphere, yearly

492

perchlorate and chlorate deposition. Optical depth data were taken from Sato et al. 35 and NASA

493

Goddard Institute for Space Studies 36.

494

Figure 8. Methyl chloroform global emissions57 as kt per year, methyl chloroform (CH3CCl3),

495

HCFC-141b (CH3CCl2F) as dry air mole fractions measured at Alert (Nunavut, Canada)53 and

496

yearly perchlorate deposition from Agassiz.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 33

497 498

REFERENCES

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

1. Wolff, J., Perchlorate and the thyroid gland. Pharmacological Reviews 1998, 50, (1), 89105. 2. National Research, C.; Division on, E.; Life, S.; Board on Environmental, S.; Toxicology; Committee to Assess the Health Implications of Perchlorate, I., Health Implications of Perchlorate Ingestion. 2005; p 1-260. 3. Urbansky, E. T.; Brown, S. K.; Magnuson, M. L.; Kelty, C. A., Perchlorate levels in samples of sodium nitrate fertilizer derived from Chilean caliche. Environmental Pollution 2001, 112, (3), 299-302. 4. Gu, B.; Coates, J. D., Perchlorate: Environmental occurrence, interactions and treatment. 2006; p 1-411. 5. Furdui, V. I.; Tomassini, F., Trends and sources of perchlorate in Arctic snow. Environmental Science and Technology 2010, 44, (2), 588-592. 6. Kounaves, S. P.; Stroble, S. T.; Anderson, R. M.; Moore, Q.; Catling, D. C.; Douglas, S.; McKay, C. P.; Ming, D. W.; Smith, P. H.; Tamppari, L. K.; Zent, A. P., Discovery of Natural Perchlorate in the Antarctic Dry Valleys and Its Global Implications. Environmental Science & Technology 2010, 44, (7), 2360-2364. 7. Ericksen, G. E., The Chilean nitrate deposits. American Scientist 1983, 71, (4), 366-374. 8. Susarla, S.; Collette, T. W.; Garrison, A. W.; Wolfe, N. L.; McCutcheon, S. C., Perchlorate identification in fertilizers. Environmental Science and Technology 1999, 33, (19), 3469-3472. 9. Dasgupta, P. K.; Martinelango, P. K.; Jackson, W. A.; Anderson, T. A.; Tian, K.; Tock, R. W.; Rajagopalan, S., The Origin of Naturally Occurring Perchlorate:  The Role of Atmospheric Processes. Environmental Science & Technology 2005, 39, (6), 1569-1575. 10. Jackson, W. A.; Böhlke, J. K.; Andraski, B. J.; Fahlquist, L.; Bexfield, L.; Eckardt, F. D.; Gates, J. B.; Davila, A. F.; McKay, C. P.; Rao, B.; Sevanthi, R.; Rajagopalan, S.; Estrada, N.; Sturchio, N.; Hatzinger, P. B.; Anderson, T. A.; Orris, G.; Betancourt, J.; Stonestrom, D.; Latorre, C.; Li, Y.; Harvey, G. J., Global patterns and environmental controls of perchlorate and nitrate co-occurrence in arid and semi-arid environments. Geochimica et Cosmochimica Acta 2015, 164, 502-522. 11. Lybrand, R. A.; Bockheim, J. G.; Ge, W.; Graham, R. C.; Hlohowskyj, S. R.; Michalski, G.; Prellwitz, J. S.; Rech, J. A.; Wang, F.; Parker, D. R., Nitrate, perchlorate, and iodate co-occur in coastal and inland deserts on Earth. Chemical Geology 2016, 442, 174-186. 12. Rajagopalan, S.; Anderson, T. A.; Fahlquist, L.; Rainwater, K. A.; Ridley, M.; Jackson, W. A., Widespread presence of naturally occurring perchlorate in high plains of Texas and New Mexico. Environmental Science and Technology 2006, 40, (10), 3156-3162. 13. Hecht, M. H.; Kounaves, S. P.; Quinn, R. C.; West, S. J.; Young, S. M. M.; Ming, D. W.; Catling, D. C.; Clark, B. C.; Boynton, W. V.; Hoffman, J.; DeFlores, L. P.; Gospodinova, K.; Kapit, J.; Smith, P. H., Detection of perchlorate and the soluble chemistry of martian soil at the phoenix lander site. Science 2009, 325, (5936), 64-67.

ACS Paragon Plus Environment

24

Page 25 of 33

539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583

Environmental Science & Technology

14. Jackson, W. A.; Davila, A. F.; Sears, D. W. G.; Coates, J. D.; McKay, C. P.; Brundrett, M.; Estrada, N.; Böhlke, J. K., Widespread occurrence of (per)chlorate in the Solar System. Earth and Planetary Science Letters 2015, 430, 470-476. 15. Simonaitis, R.; Heicklen, J., Perchloric acid: A possible sink for stratospheric chlorine. Planetary and Space Science 1975, 23, (11), 1567-1569. 16. Jaeglé, L.; Yung, Y. L.; Toon, G. C.; Sen, B.; Blavier, J.-F., Balloon observations of organic and inorganic chlorine in the stratosphere: The role of HClO4 production on sulfate aerosols. Geophysical Research Letters 1996, 23, (14), 1749-1752. 17. Murphy, D. M.; Thomson, D. S., Halogen ions and NO+ in the mass spectra of aerosols in the upper troposphere and lower stratosphere. Geophysical Research Letters 2000, 27, (19), 3217-3220. 18. Delmas, R. J., Environmental information from ice cores. Reviews of Geophysics 1992, 30, (1), 1-21. 19. Legrand, M.; Mayewski, P., Glaciochemistry of polar ice cores: A review. Reviews of Geophysics 1997, 35, (3), 219-243. 20. Rao, B. A.; Wake, C. P.; Anderson, T.; Jackson, W. A., Perchlorate depositional history as recorded in North American ice cores from the eclipse icefield, Canada, and the Upper Fremont Glacier, USA. Water, Air, and Soil Pollution 2012, 223, (1), 181-188. 21. Rajagopalan, S.; Anderson, T.; Cox, S.; Harvey, G.; Cheng, Q.; Jackson, W. A., Perchlorate in Wet Deposition Across North America. Environmental Science & Technology 2009, 43, (3), 616-622. 22. Peterson, K. M.; Cole-Dai, J.; Brandis, D. L.; Manandhar, E., Assessment of anthropogenic contribution to perchlorate in the environment using an ice core record. In ACS Symposium Series, 2015; Vol. 1198, pp 175-185. 23. Crawford, T. Z.; Kub, A. D.; Peterson, K. M.; Cox, T. S.; Cole-Dai, J., Reduced perchlorate in West Antarctica snow during stratospheric ozone hole. Antarctic Science 2017, 29, (3), 292-296. 24. Jiang, S.; Cox, T. S.; Cole-Dai, J.; Peterson, K. M.; Shi, G., Trends of perchlorate in Antarctic snow: Implications for atmospheric production and preservation in snow. Geophysical Research Letters 2016, 43, (18), 9913-9919. 25. von Clarmann, T., Chlorine in the stratosphere. Atmosfera 2013, 26, (3), 415-458. 26. Fabian, P.; Borchers, R.; Penkett, S. A.; Prosser, N. J. D., Halocarbons in the stratosphere. Nature 1981, 294, (5843), 733-735. 27. Paul, C.; Pohnert, G., Production and role of volatile halogenated compounds from marine algae. Natural Product Reports 2011, 28, (2), 186-195. 28. Khalil, M. A. K.; Rasmussen, R. A., Atmospheric methyl chloride. Atmospheric Environment 1999, 33, (8), 1305-1321. 29. Khalil, M. A. K.; Moore, R. M.; Harper, D. B.; Lobert, J. M.; Erickson, D. J.; Koropalov, V.; Sturges, W. T.; Keene, W. C., Natural emissions of chlorine-containing gases: Reactive Chlorine Emissions Inventory. Journal of Geophysical Research Atmospheres 1999, 104, (D7), 8333-8346. 30. Molina, M. J.; Rowland, F. S., Stratospheric sink for chlorofluoromethanes: Chlorine atomc-atalysed destruction of ozone. Nature 1974, 249, (5460), 810-812. 31. McConnell, J. C.; Schiff, H. I., Methyl chloroform: Impact on stratospheric ozone. Science 1978, 199, (4325), 174-177.

ACS Paragon Plus Environment

25

Environmental Science & Technology

584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629

Page 26 of 33

32. Fabian, P.; Gömer, D., The vertical distribution of halocarbons in the stratosphere. Fresenius' Zeitschrift für Analytische Chemie 1984, 319, (8), 890-897. 33. Rinsland, C. P.; Mahieu, E.; Zander, R.; Jones, N. B.; Chipperfield, M. P.; Goldman, A.; Anderson, J.; Russell, J. M.; Demoulin, P.; Notholt, J.; Toon, G. C.; Blavier, J. F.; Sen, B.; Sussmann, R.; Wood, S. W.; Meier, A.; Griffith, D. W. T.; Chiou, L. S.; Murcray, F. J.; Stephen, T. M.; Hase, F.; Mikuteit, S.; Schulz, A.; Blumenstock, T., Long-term trends of inorganic chlorine from ground-based infrared solar spectra: Past increases and evidence for stabilization. Journal of Geophysical Research: Atmospheres 2003, 108, (D8), n/a-n/a. 34. WMO Scientific Assessment of Ozone Depletion; World Meteorological Organization Global Ozone Research and Monitoring Project - Report No. 55, Geneva; https://www.esrl.noaa.gov/csd/assessments/ozone/2014/, 2014. 35. Sato, M.; Hansen, J. E.; McCormick, M. P.; Pollack, J. B., Stratospheric aerosol optical depths, 1850–1990. Journal of Geophysical Research: Atmospheres 1993, 98, (D12), 2298722994. 36. NASA Goddard Institute for Space Studies, Stratospheric Aerosol Optical Thickness, Optical Depth at 550 nm, 2017, https://data.giss.nasa.gov/modelforce/strataer/. 37. Zheng, J., Archives of total mercury reconstructed with ice and snow from Greenland and the Canadian High Arctic. Science of the Total Environment 2015, 509–510, 133-144. 38. Zheng, J.; Pelchat, P.; Vaive, J.; Bass, D.; Ke, F., Total mercury in snow and ice samples from Canadian High Arctic ice caps and glaciers: A practical procedure and method for total Hg quantification at low pgg-1 level. Science of the Total Environment 2014, 468-469, 487-494. 39. Goto-Azuma, K.; Koerner, R. M., Ice core studies of anthropogenic sulfate and nitrate trends in the Arctic. Journal of Geophysical Research Atmospheres 2001, 106, (D5), 4959-4969. 40. Isaksson, E.; Pohjola, V.; Jauhiainen, T.; Moore, J.; Pinglot, J. F.; Vaikmaäe, R.; van de Wal, R. S. W.; Hagen, J. O.; Ivask, J.; Karlöf, L.; Martma, T.; Meijer, H. A. J.; Mulvaney, R.; Thomassen, M.; van den Broeke, M., A new ice-core record from Lomonosovfonna, Svalbard: viewing the 1920-97 data in relation to present climate and environmental conditions. Journal of Glaciology 2001, 47, (157), 335-345. 41. Koerner, R. M.; Fisher, D. A.; Goto-Azuma, K., A 100 year record of ion chemistry from Agassiz Ice Cap Northern Ellesmere Island NWT, Canada. Atmospheric Environment 1999, 33, (3), 347-357. 42. Johnsen, S. J.; Clausen, H. B.; Dansgaard, W.; Fuhrer, K.; Gundestrup, N.; Hammer, C. U.; Iversen, P.; Jouzel, J.; Stauffer, B.; Steffensen, J. P., Irregular glacial interstadials recorded in a new Greenland ice core. Nature 1992, 359, (6393), 311-313. 43. Fisher, D. A.; Wake, C.; Kreutz, K.; Yalcin, K.; Steig, E.; Mayewski, P.; Anderson, L.; Zheng, J.; Rupper, S.; Zdanowicz, C.; Demuth, M.; Waszkiewicz, M.; Dahl-Jensen, D.; GotoAzuma, K.; Bourgeois, J. B.; Koerner, R. M.; Sekerka, J.; Osterberg, E.; Abbott, M. B.; Finney, B. P.; Burns, S. J., Stable isotope records from Mount Logan, eclipse ice cores and nearby Jellybean Lake. Water cycle of the North Pacific over 2000 years and over five vertical kilometres: Sudden shifts and tropical connections. Geographie Physique et Quaternaire 2004, 58, (2-3), 337-352. 44. Yalcin, K.; Wake, C. P.; Germani, M. S., A 100-year record of North Pacific volcanism in an ice core from Eclipse Icefield, Yukon Territory, Canada. Journal of Geophysical Research: Atmospheres 2003, 108, (D1), AAC 8-1-AAC 8-12. 45. Bluth, G. J. S.; Rose, W. I.; Sprod, I. E.; Krueger, A. J., Stratospheric Loading of Sulfur From Explosive Volcanic Eruptions. The Journal of Geology 1997, 105, (6), 671-684.

ACS Paragon Plus Environment

26

Page 27 of 33

630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662

Environmental Science & Technology

46. Johnston, D. A., Volcanic Contribution of Chlorine to the Stratosphere: More Significant to Ozone Than Previously Estimated? Science 1980, 209, (4455), 491-493. 47. Rao, B.; Anderson, T. A.; Redder, A.; Jackson, W. A., Perchlorate formation by ozone oxidation of aqueous chlorine/oxy-chlorine species: Role of ClxOy radicals. Environmental Science and Technology 2010, 44, (8), 2961-2967. 48. Holloway, A. M.; Wayne, R. P., Atmospheric Chemistry. The Royal Society of Chemistry: 2010. 49. Catling, D. C.; Claire, M. W.; Zahnle, K. J.; Quinn, R. C.; Clark, B. C.; Hecht, M. H.; Kounaves, S., Atmospheric origins of perchlorate on mars and in the atacama. Journal of Geophysical Research E: Planets 2010, 115, (1). 50. Rao, B.; Mohan, S.; Neuber, A.; Jackson, W. A., Production of perchlorate by laboratory simulated lightning process. Water, Air, and Soil Pollution 2012, 223, (1), 275-287. 51. Schumann, U.; Huntrieser, H., The global lightning-induced nitrogen oxides source. Atmos. Chem. Phys. 2007, 7, (14), 3823-3907. 52. Murphy, D. M.; Thomson, D. S.; Mahoney, M. J., In Situ Measurements of Organics, Meteoritic Material, Mercury, and Other Elements in Aerosols at 5 to 19 Kilometers. Science 1998, 282, (5394), 1664-1669. 53. National Oceanic and Atmospheric Administration Earth System Research Laboratory, Halocarbons & other Atmospheric Trace Species Group; https://www.esrl.noaa.gov/gmd/hats/, 2017. 54. Lovelock, J. E., Methyl chloroform in the troposphere as an indicator of OH radical abundance [3]. Nature 1977, 267, (5606), 32. 55. Rigby, M.; Prinn, R. G.; O'Doherty, S.; Montzka, S. A.; McCulloch, A.; Harth, C. M.; Mühle, J.; Salameh, P. K.; Weiss, R. F.; Young, D.; Simmonds, P. G.; Hall, B. D.; Dutton, G. S.; Nance, D.; Mondeel, D. J.; Elkins, J. W.; Krummel, P. B.; Steele, L. P.; Fraser, P. J., Reevaluation of the lifetimes of the major CFCs and CH3CCl3 using atmospheric trends. Atmos. Chem. Phys. 2013, 13, (5), 2691-2702. 56. Kutsuna, S.; Takeuchi, K.; Ibusuki, T., Laboratory study on heterogeneous degradation of methyl chloroform (CH3CCl3) on aluminosilica clay minerals as its potential tropospheric sink. Journal of Geophysical Research Atmospheres 2000, 105, (D5), 6611-6620. 57. Krol, M.; van Leeuwen, P. J.; Lelieveld, J., Global OH trend inferred from methylchloroform measurements. Journal of Geophysical Research: Atmospheres 1998, 103, (D9), 10697-10711.

663 664 665 666 667

ACS Paragon Plus Environment

27

Environmental Science & Technology

668

Page 28 of 33

Figure 1.

669 670 671 672 673 674 675 676

ACS Paragon Plus Environment

28

Page 29 of 33

677

Environmental Science & Technology

Figure 2.

678 679

Figure 3.

680 681 682 683

ACS Paragon Plus Environment

29

Environmental Science & Technology

684

Page 30 of 33

Figure 4.

685 686 687

Figure 5.

688

689 690

ACS Paragon Plus Environment

30

Page 31 of 33

691

Environmental Science & Technology

Figure 6.

692 693 694

Figure 7.

695 696

ACS Paragon Plus Environment

31

Environmental Science & Technology

Page 32 of 33

697 698

Figure 8.

180

CH3CCl3+HCFC141B

160

CH3CCl3

140

HCFC141B

120 100

5.0

4.0 CH3CCl3 Emissions

3.0

Perchlorate

80

2.0

60 40

1.0

20

700

0 1960

Yearly Flux Perchlorate (µg m-2)

Dry Air Mole Fraction (ppt) 15% of CHCCl3 Emissions (kt)

699

0.0 1970

1980

1990

2000

2010

701 702 703 704 705 706 707

ACS Paragon Plus Environment

32

Page 33 of 33

708

Environmental Science & Technology

TOC:

709

710

ACS Paragon Plus Environment

33