Influence of Atmospheric Processes on the Solubility and Composition

Jun 10, 2016 - Athena uses a nonlinear, least-squares minimization approach to fit spectra of unknown samples with standard material spectra, computin...
0 downloads 9 Views 483KB Size
Subscriber access provided by UNIV OF CAMBRIDGE

Article

Influence of atmospheric processes on the solubility and composition of iron in Saharan dust Amelia F. Longo, Yan Feng, Barry Lai, William M Landing, Rachel U. Shelley, Athanasios Nenes, Nikolaos Mihalopoulos, Kalliopi Violaki, and Ellery D. Ingall Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b02605 • Publication Date (Web): 10 Jun 2016 Downloaded from http://pubs.acs.org on June 12, 2016

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

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

Environmental Science & Technology

8

10%

7

8%

6

6%

5

4%

4

2%

3

0%

Travel Time (days)

ACS Paragon Plus Environment

Fe Solubility (%)

pH

Page 1 of 31

Environmental Science & Technology

Page 2 of 31

1 2

Influence of atmospheric processes on the solubility and composition of iron in Saharan dust

3 4 5

Amelia F. Longo1, Yan Feng2, Barry Lai3, William M. Landing4, Rachel U. Shelley4§, Athanasios Nenes1,5-7, Nikolaos Mihalopoulos7,8, Kalliopi Violaki8, and Ellery D. Ingall1*

6

1

Drive, Atlanta, GA 30332-0340, USA

7 8

2

3

4

Department of Earth, Ocean and Atmospheric Science, Florida State University, Tallahassee, FL 32306, USA

13 14

Advanced Photon Source, Argonne National Laboratory, 9700 S. Cass Avenue, Argonne, IL 60439, USA

11 12

Enviromental Science Division, Argonne National Laboratory, 9700 S. Cass Avenue, Argonne, IL 60439, USA

9 10

School of Earth and Atmospheric Sciences, Georgia Institute of Technology, 311 Ferst

5

School of Chemical and Biomolecular Engineering, Georgia Institute of Technology, 311 Ferst Drive, Atlanta, GA 30332-0340, USA.

15 16

6

Foundation for Research and Technology, Hellas, Patras 70013, Greece.

17

7

National Observatory of Athens, Penteli GR-15236, Greece

18

8

Department of Chemistry, University of Crete, Iraklion 71003, Greece

19

§

Present Address: Laboratoire des Sciences de l'Environnement Marin, Institut

20 21

Universitaire Européen de la Mer, Plouzané, 29280, France *Corresponding Author Contact Information:404-894-3883, [email protected]

22 23

21 ACS Paragon Plus Environment

Page 3 of 31

Environmental Science & Technology

24

Keywords Aerosol composition, iron, Saharan dust, GEOTRACES, particle aging

25

Abstract Aerosol iron was examined in Saharan dust plumes using a combination of iron near-

26

edge X-ray absorption spectroscopy and wet chemical techniques. Aerosol samples were

27

collected at three sites located in the Mediterranean, the Atlantic, and Bermuda to characterize

28

iron at different atmospheric transport lengths and time scales. Iron(III) oxides were a component

29

of aerosols at all sampling sites and dominated the aerosol iron in Mediterranean samples. In

30

Atlantic samples, iron(II & III) sulfate, iron(III) phosphate, and iron(II) silicates were also

31

contributors to aerosol composition. With increased atmospheric transport time, iron(II) sulfates

32

are found to become more abundant, aerosol iron oxidation state became more reduced, and

33

aerosol acidity increased. Atmospheric processing including acidic reactions and photo-reduction

34

likely influence the form of iron minerals and oxidation state in Saharan dust aerosols and

35

contribute to increases in aerosol iron solubility.

36 37 38

22 ACS Paragon Plus Environment

Environmental Science & Technology

39

Page 4 of 31

1.0 Introduction

40

Iron is a key micronutrient that is vital for all organisms and controls primary productivity in

41

approximately 30% of the world’s oceans.1 As a result, the relationship between iron and marine

42

microorganisms has been extensively studied most notably through iron fertilization experiments

43

in which the sequestration of carbon dioxide was examined.2-5 These studies show that the

44

addition of soluble iron to extensive ocean regions increases primary productivity, which is

45

reflected in the formation of carbon-sequestering algae blooms.3-7 The solubility of aerosol iron

46

is also a key factor in the formation of reactive oxygen species in aqueous phase.8-10 Reactive

47

oxygen species are involved in numerous gas and aqueous phase reactions and have been

48

associated with negative health effects.11 The potential impacts of soluble aerosol iron on ocean

49

nutrient budgets12, 13 and reactive oxygen species formation8-10 underscore the need to understand

50

controls and transformations in aerosol iron solubility during atmospheric transport.

51

The Sahara Desert contains the largest and most active sources of dust globally.14 The

52

Saharan Air Layer, a hot, dry layer of the atmosphere that can overlay dense marine air,

53

transports Saharan dust from North Africa, where it is uplifted and entrained as aerosol, to as far

54

as North America. The vast transport distances make Saharan dust a key source of nutrients to

55

the Mediterranean Sea, much of the North Atlantic Ocean, and even as far as the Gulf of

56

Mexico.15, 16 As a result, this influential dust source has been extensively studied as a nutrient

57

reservoir.17-19 One trend has emerged for aerosol iron solubility: when total iron is low, the

58

relative solubility of iron is high, and when total iron is high, the relative solubility of iron is

59

low.20 The factors that control the solubility of iron in mineral dust, however, remain unclear.

60

Particle size,21 physical processes,22 compositional, and oxidation state variations as well as

61

particle aging processes, including photo-reduction,23-25 organic ligand association,25,

26

and

23 ACS Paragon Plus Environment

Page 5 of 31

Environmental Science & Technology

62

proton reactions,25, 27-31 have all been proposed as controls of aerosol iron dissolution, but little

63

direct evidence for these has been collected for Saharan dust.

64

In laboratory studies, proton reactions have been shown to enhance solubility of recalcitrant

65

mineral phases25,

31, 32

66

evidence of acidic reactions is only circumstantial for environmental samples.32,

67

atmospheric proton reactions, acidic species may overcome the carbonate buffer capacity of an

68

aerosol particle, allowing the pH to reach values of 1 to 2 in aqueous solutions surrounding

69

aerosol particles.37,

70

particle is rich in calcium carbonate allowing acidic conditions to form more readily in some

71

particles.39 Under these acidic conditions, iron is more soluble.36 Models have estimated photo-

72

chemical and organic ligand-promoted dissolution to increase solubility of iron by an order of

73

magnitude and are predicted to increase total dissolved iron deposition by as much as 75%

74

globally.40 Because of the uncertainty surrounding particle aging processes, proton reactions are

75

the most common reaction type included in atmospheric models. This variability in modeled

76

global soluble iron flux impedes understanding of aerosol iron in marine ecosystems and

77

interactions between the atmosphere and ocean biogeochemical cycles.41, 42

38

as well as precipitate iron-rich reactive nanoparticles;33,

34

however,

35, 36

During

However because of the heterogeneous nature of aerosols, not every

78

Observational studies of aerosol composition and aging processes often employ bulk

79

measurements of iron composition and oxidation state.24-26, 43 Composition of aerosol iron has

80

often been confined to oxidation state44, 45 or sequential extraction procedures, which categorize

81

iron as water-exchangeable, easily reducible, or oxidizable.46 Spectroscopic techniques allow for

82

more detailed aerosol iron characterization, such as identification of specific compound or

83

mineral classes.47-50 Here, bulk chemical characterizations are combined with synchrotron-based

84

elemental mapping and spectroscopy to analyze Saharan dust aerosols. The results of these

24 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 31

85

experiments are then combined with a thermodynamic analysis of the major soluble ions present

86

in the sample for the determination of its acidity.

87

2.0 Methods

88

2.1 Sample Collection

89

Samples were collected at three locations, Greece, Atlantic Ocean, and Bermuda (Figure S1).

90

Sampling stations were selected to capture Saharan dust at different transport length and time

91

scales but do not represent points along a single atmospheric flow path. Ideally, sampling would

92

continuously follow a single plume as it travelled through the atmosphere; however, these

93

experiments are cost prohibitive requiring aircraft or multiple ship based sampling sites.

94

Therefore, sampling sites were chosen to capture characteristic atmospheric climatology, which

95

represents the typical pathways that Saharan dust takes through the atmosphere. These three

96

sampling sites are located in the common transport pathway of Saharan dust and receive dust

97

laden air masses on seasonal cycles. At Finokalia Research Station, located on the island of

98

Crete, Greece, North African dust can be found any time between October and May; however,

99

December to April constitutes the primary Saharan dust season.16 Saharan dust is also

100

transported episodically across the Atlantic Ocean. These dust events, occurring between May

101

and September, originate in North Africa and are transported on the trade winds to Bermuda.51-55

102

Although samples were not collected following a single dust plume, this data set examines

103

multiple dust events that impact either the Mediterranean Sea or the Atlantic Ocean. These

104

sampling sites will be compared, generally, for differences between sampling sites, and general

105

characteristics of aerosol iron minerals at individual geographical locations in these Saharan dust

106

transport pathways.

25 ACS Paragon Plus Environment

Page 7 of 31

Environmental Science & Technology

107

The Greek samples, closest to the Saharan dust source, were collected at the Finokalia

108

Research Station (35.32˚N, 25.67˚E), a site isolated from both local and regional influences on

109

the island of Crete, Greece, located 70 km from the nearest major city.56 Samples of particulate

110

matter with an aerodynamic diameter < 10 µm, PM10, were collected on Teflon filters without

111

any pretreatment using a virtual impactor with an operational flow rate of 16.7 L/min. More

112

details are reported in Koulouri et al. (2008).57 Seven samples were collected at the Finokalia site

113

over a one to three day period from 2009-2011 during which North African air masses

114

dominated.

115

Twenty-four hour particle collections were completed as part of the 2011 GEOTRACES 58

116

Atlantic transect

. Because the GEOTRACES samples were collected over a short period of

117

time and relatively near the source, they likely provide the purest sampling of Saharan air

118

masses. The GEOTRACES samples are collected from three sites; however, the sampling

119

locations are very close together, and the dust was from a single dust event, sampled over three

120

days. Therefore, the GEOTRACES sample sites are referred to collectively. The three samples

121

were collected at the following locations: Sample 7899, 22.37˚N 35.62˚W; Sample 7946,

122

20.88˚N 32.62˚W; Sample 8004, 19.43˚N, 29.38˚W. Total suspended particle (TSP) samples

123

were collected using Whatman-41 filters and a Tisch 5170-VBL high volume sampler operated

124

on average with a face velocity of ~100 cm/sec.58 The aerosol sampler was wind speed (> 0.5

125

m/s) and sector-controlled (+/- 60˚ from bow) to avoid contamination from ship’s exhaust. The

126

Whatman 41 (W41) filters were cleaned by (1) soaking in 0.5 M HCl (Optima) for 24 hours, (2)

127

rinsing thoroughly with ultra-high purity water (UHP water, 18.2 MΩ*cm), (3) soaking in a

128

second 0.5 M HCl (Optima) bath for 24 h, (4) rinsing with UHP water until the pH of the water

26 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 31

129

returned to the pH of UHP water (~ pH 5.6), (5) placing on a plastic mesh rack inside a HEPA

130

laminar flow bench until dry and (6) storage in two zipper-seal bags until use.

131

Weekly samples were collected at the Bermuda Institute of Ocean Sciences (32.24˚N,

132

64.87˚W), which represent the longest transport of Saharan dust in our collection. Sampling was

133

conducted from June 2011 to June 2012. These samples were collected using a Tisch 5170-

134

VBL high-volume Total Suspended Particle (TSP) aerosol sampler using Whatman-41 filters

135

and with a face velocity of ~100 cm/sec. The Whatman-41 filters were cleaned as detailed above.

136

Of the fifty-two weekly-integrated samples that were collected, eight samples were largely

137

influenced by Saharan dust events. All aerosol samples were stored in the dark at -20oC until

138

analysis.

139

Aerosols sampled from Finokalia Research Station represent particles with an aerodynamic

140

diameter of less than 10 µm (PM10). GEOTRACES and BIOS samples represent Total

141

Suspended Particles (TSP), meaning they collect particles with an aerodynamic diameter of less

142

than 10 µm but also any particles with an aerodynamic diameter of greater than 10 µm. However,

143

particles larger than this will often settle out of during long range atmospheric transport.59 So,

144

while Finokalia Research Station samples represent a different size fraction from the

145

GEOTRACES and BIOS samples, the actual particle size ranges collected are likely similar.

146

2.2 HYSPLIT

147

Hybrid Single Particle Lagrangian Integrated Trajectory Model (HYSPLIT)60,

61

back

148

trajectories were completed for each sample using the GDAS half-degree meteorological archive

149

in order to confirm the geographic origin of the air masses sampled (Figure S1). For the

150

Bermuda Institute of Ocean Sciences (BIOS) and GEOTRACES samples, HYSPLIT back

151

trajectories were calculated at 0 m, 500 m and 1000 m above ground level to confirm the North 27 ACS Paragon Plus Environment

Page 9 of 31

Environmental Science & Technology

152

African provenance. At the Finokalia Research Station, HYSPLIT back trajectories were

153

calculated at 1000 m and 3000 m above ground level for five days (approximately atmospheric

154

life time of dust aerosols) preceding sample collection. HYSPLIT back trajectories computed at

155

1000 m show the origin of air masses within the boundary layer; this height is chosen, rather than

156

the more conventional heights of 0 m or 500 m to avoid orographic influences at this site.

157

HYSPLIT back trajectories were computed at 3000 m to identify dust events. At Finokalia

158

Research Station, dust is either homogeneously distributed from 0 to about 3000 m, during

159

spring and autumn dust events, or is found in a layer between 2500 – 4000 m during summer and

160

autumn dust events.62 While the HYSPLIT back trajectories do not guarantee that pure end

161

members were sampled, they help to demonstrate that most air masses contained North African

162

origins during the sampling time periods. Additionally, all samples had the distinctive red-

163

orange color of Saharan aerosols.

164

The HYSPLIT back trajectories were also used to estimate the residence time of dust

165

aerosols in the atmosphere from generation in the Sahara Desert to collection. For this

166

calculation, a five day HYSPLIT back trajectory was completed in ensemble mode at 3000 m.

167

Since each sample was a multiday collection, HYSPLIT back trajectories were completed for

168

each day during the collection period. The mean distance travelled was then determined for each

169

sample and taken as the average distance travelled in five days by the air mass. The average

170

distance travelled in five days was used to estimate a velocity for each sample (average distance

171

travelled in five days divided by five days time). Finally, the distance travelled from source to

172

collection was estimated. The source region in the Sahara Desert was taken as a single location

173

(30˚N, 8.5˚E). The distances travelled were estimated as follows: Finokalia Research Station,

174

1800 km; GEOTRACES, 2100 km; and Bermuda Institute of Ocean Science, 6900 km. Using

28 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

175

these distances and the estimated velocities, an approximation of travel time was generated for

176

each sample. These approximate travel times are reasonable for Saharan aerosols which have

177

been known to travel for as many as 30 days before deposition.63

178

2.3 Total and soluble iron

179

Water soluble iron, hereafter referred to as soluble iron, was determined for all samples using

180

the ferrozine technique.64 Handling of the filters was conducted under a HEPA-filter laminar

181

flow hood. A quarter filter was extracted into 10 mL of deionized water (>18 MΩ) in a trace

182

metal grade centrifuge tube by 30 min of ultrasonication. The extraction was followed by the

183

removal of insoluble particles from the extract by a 0.2 µm pore size syringe filter. Any

184

dissolved Fe(III) in the sample was reduced to Fe(II) through the addition of 40 µM solution of

185

hydroxylamine hydrochloride added in a 1:100 ratio of reagent to filter extract. Samples were

186

allowed to react with the hydroxylamine hydrochloride overnight. A 5 mM ferrozine solution

187

which forms a colored complex with dissolved Fe(II) was then added in a 1:100 ratio of reagent

188

to filter extract. After a ten minute reaction, samples were spectrophotometrically measured at

189

562 nm.36 The detection limit for this method (5.6 ng/mL) was two orders of magnitude lower

190

than the sample signals. Total iron was measured for GEOTRACES and Bermuda Institute of

191

Ocean Science samples using protocols described in Morton et al. (2013).58,

192

limits for this methods are between 0.46 – 0.62 ng/m3.66 For BIOS and GEOTRACES samples,

193

total iron was also measured via elemental densities collected from X-ray fluorescence maps to

194

ensure total iron could be reliably determined through synchrotron-based techniques,47 and for

195

Finokalia Research Station samples, total iron was only determined using elemental densities

196

collected from X-ray fluorescence maps. The detection limit of this method is approximately 1

197

ng/m3.

65, 66

Detection

29 ACS Paragon Plus Environment

Page 11 of 31

198

Environmental Science & Technology

2.4 Soluble Ions

199

Ion chromatography was used to measure major soluble ions for each aerosol sample

200

according to Bardouki et al. (2003).67 Briefly, approximately 1/8th of each filter was extracted

201

into 10 mL of nanopure water using an ultrasonic bath for 45 min. The filter extracts were then

202

syringe filtered (PALL IC Acrodisc (PES), 0.45 µm, 13 mm) to remove any insoluble species.

203

The acquired filtered solutions were analyzed for anions (Cl−, Br−, NO3−, SO42-, HPO42-, C2O42-)

204

and cations (NH4+, K+, Na+, Mg2+, Ca2+). Anions were determined using a Dionex-500 ion

205

chromatograph equipped with an Ion Pac AS4A-SC column and an AG4A-SC precolumn, with

206

an ASRS-300 suppressor. Anion separation was conducted with isocratic elution of NaHCO3

207

(3.4 mM)/Na2CO3 (3.6 mM) as an eluent and a flow of 1.5 mL/min. For cations, an Ion Pac

208

CS12A column and a CG12A guard column were used, with a CSRS-300 suppressor, under

209

isocratic elution of 20 mM MSA (methanesulfonic acid) at a flow rate of 1.0 mL/min. The

210

detection limit of the analysis was 20, 12, 40, 12 and 40 ppb for NH4+, K+, Na+, Mg2+, and Ca2+,

211

respectively, while the corresponding detection limit for all anions (Cl−, Br−, NO3− , SO42-,

212

HPO42-, and C2O42-) was 20 ppb.

213

2.5 pH modelling

214

ISORROPIA was used to estimate the aerosol acidity of each aerosol sample.68,

69

This

215

thermodynamic model will output the concentration of species in the gas and aerosol

216

(solid/liquid) phases, pH, and aerosol water. This approach to aerosol pH modelling will only

217

determine whether the samples are acidic or circum neutral. To compute the pH, knowledge of

218

the relative humidity and temperature is needed; therefore, values characteristic of a humid

219

marine boundary layer were assumed (80-90% Relative humidity; 293K). Input values for major

220

aerosol species can be found in Table S1.

30 ACS Paragon Plus Environment

Environmental Science & Technology

221

Page 12 of 31

2.6 Synchrotron Analysis

222

Aerosol composition was determined using the micro X-ray fluorescence capabilities at

223

Argonne National Laboratory, Advanced Photon Source, Station 2-ID-D. In this study X-ray

224

fluorescence microscopy and X-ray Absorption Near Edge Structure (XANES) spectroscopy

225

were used to examine the iron in aerosol. These techniques work by providing incident X-rays of

226

sufficient energy to eject the electrons from the innermost electron shell of atoms. Subsequently,

227

an outer shell electron may relax into the vacated position, emitting a characteristic fluorescence

228

signal. The type, oxidation state, and structural arrangement of atoms in a particle are revealed as

229

the ejected electron interacts with neighboring atoms and are reflected in XANES spectra70.

230

Therefore, Fe-XANES spectra provide information on both oxidation state and the composition

231

associated with the element of interest.

232

An energy dispersive Si-drift detector (Vortex EM, with a 50 mm2 sensitive area, and a 12.5

233

µm Be window; SII NanoTechnology, Northridge, CA) was used to measure X-ray fluorescence

234

of the samples. All measurements were conducted under a helium atmosphere in order to

235

minimize absorption and fluorescence caused by beam interactions with air. A randomly selected

236

area of each filter sample (0.5 cm2) was placed over a slot of an aluminum sample mount for

237

analysis. The sample was initially analyzed in microscopic X-ray fluorescence mode to identify

238

iron-containing particles on the filter. In this mode, the X-ray beam with a diameter of ca. 200

239

nm was scanned over a randomly selected filter area (typically 40 to 50 µm2) at a step size of 0.4

240

µm and 0.5 s dwell time to produce an elemental distribution map of the filter. The resulting

241

elemental data were compiled to identify iron-containing particles (>1 µm diameter) for Fe

242

XANES spectroscopy.

31 ACS Paragon Plus Environment

Page 13 of 31

Environmental Science & Technology

243

Fe XANES spectroscopy was conducted in bulk and individual particle modes. In bulk Fe

244

XANES mode, X-ray focusing optics (zone plate and order sorting aperture) were removed

245

allowing characterization of iron present in ca. 0.4 mm2 filter areas. In individual particle mode,

246

iron-containing particles identified in elemental distribution maps were characterized with a

247

focused beam with a 0.4 µm spot size. Of course, every particle could not be examined because

248

of time constraints. Based on elemental maps, particles were selected to attempt to characterize

249

as many different elemental iron associations as possible. Individual particle XANES

250

spectroscopy represent the composition of an entire particle, which can be a mixture of several

251

mineral phases. In both bulk and individual particle modes, XANES scans were collected from

252

7100 to 7180 eV in 0.5 eV steps with 0.5−3 s dwell time per step.

253

2.7 Data Reduction and Analysis

254

Oxidation state was determined for each sample by determining the pre-edge centroid of

255

each Fe XANES spectra. The pre-edge feature has been shown to be the most reliable

256

indication of oxidation state, and by interpolating between pre-edge centroid position %

257

Fe(II) of standards, the relative abundance of Fe(II) can be determined for environmental

258

samples.47, 71, 72

259

Spectral data were normalized to account for variations in the incoming X-ray flux and

260

processed using the Athena software package.73 In total, 105 spectra were collected. In

261

order to extract the most useful information from this number of spectra, data reduction

262

techniques were employed. A combination of principal component analysis, PCA, and k-

263

means clustering were used to group like spectra.74 The PCA algorithm used, built into

264

Athena,73 generates a scree plot where the number of principal components can be

265

determined. Like spectra were merged73 to create an average spectra representing a cluster,

32 ACS Paragon Plus Environment

Environmental Science & Technology

266

and linear combination fitting was then completed for each cluster’s Fe XANES spectra

267

using Athena.73 These data reduction techniques were used on each sampling site, with the

268

spectra being divided into bulk and individual particle Fe XANES.

269

The linear combination fitting algorithm in Athena uses a database of known iron

270

mineral standards (e.g., iron oxides, organics, silicates, sulfides, and sulfates)75 run on the

271

same synchrotron beamline, in this case, to determine the phases contributing to an

272

unknown aerosol Fe XANES spectra. An iterative process was used to refine the standard

273

database used to model the sample spectra. First, exotic iron minerals unlikely to be major

274

components of aerosol iron were excluded from the database.47 Second, the database was

275

narrowed through elimination of standards with low contribution (i.e. less than 10%) or

276

poor fit (i.e. high R-factor) during initial linear combination fitting. As the acidic reactions

277

and proto-reduction reactions are taking place on the surface of aerosols, very small changes

278

in composition may be below the detection limit of these methods. This could lead to

279

insufficient quantities of a specific mineral or compound in a sample and an

280

underestimation of the specific compound during linear combination fitting;76 thus, iron

281

composition was generalized to express broad chemical classes: iron(III) oxides, iron(III)

282

sulfates, iron(III) phosphates, iron(II) sulfates, and iron(II) silicates. Different minerals

283

within these classes can have widely varying chemical properties. For iron(III) phosphates,

284

it is mainly the hydration that differs; however, iron(III) oxide minerals such as hematite

285

and goethite are in the same category, but have iron solubilities that differ by an order of

286

magnitude.21 Therefore, it is worth noting that in this study that goethite was the iron oxide

287

mineral likely present in the aerosol. Athena uses a non-linear, least-squares minimization

288

approach to fit spectra of unknown samples with standard material spectra, computing an

Page 14 of 31

33 ACS Paragon Plus Environment

Page 15 of 31

Environmental Science & Technology

289

error term, the R-factor, to quantify the goodness of a given linear combination fit. The

290

linear combination of standards that yielded the lowest R-factor reflected the best fit.73

291

3.0 Results

292

To put the results in context, it is important to visualize the typical transport paths of air

293

masses originating in the Sahara. Air masses carrying Saharan dust to Finokalia Research

294

Station travel a relatively short distance across the Mediterranean (~1800 km). GEOTRACES

295

samples collected off the west coast of Africa, near the Cape Verde Islands, also represent a

296

relatively short travel distance (~2100 km). The air masses that deliver Saharan dust to the Cape

297

Verde islands and surrounding waters may continue their westward journey across the Atlantic

298

Ocean, reaching Bermuda after atmospheric transport of approximately 6900 km from source.

299

Sampling stations do not represent different points along a single atmospheric flow path; here,

300

we use the different sampling sites to explore the effects of a range of atmospheric transport

301

times on iron properties in aerosols.

302

3.1 Iron Solubility

303

Saharan dust solubility varies at the three different sampling sites, with the highest solubility

304

at the Bermuda Institute of Ocean Sciences (BIOS) site and the lowest at the GEOTRACES

305

collection site (Figure 1).

306

concentration of total iron, averaging 196 ng/m3 (Table S2). Both the Finokalia Research Station

307

and GEOTRACES sampling sites had significantly higher total aerosol iron concentrations

308

(averaging 735 ng/m3 and 1580 ng/m3, respectively). At each collection site, the soluble iron was

309

on the same order of magnitude with mean values ranging from 5.0-8.75 ng/m3. Moreover, when

310

the soluble iron is normalized to the total iron, the differences between each sampling site

311

become apparent. The BIOS contained on average 2.55% soluble iron, with values ranging from

While BIOS exhibited the highest solubility, it had the lowest

34 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 31

312

1.7 – 8.9% (Figure 1). In contrast, the GEOTRACES samples had consistently low solubility,

313

averaging 0.41%. Samples from the Finokalia Research Station fell between these two locations

314

containing on average 1.19% soluble iron; this station also exhibited a wide range of iron

315

solubilities from 0.4 – 1.5% (Figure 1). Solubility generally increased with the estimated travel

316

time from the source. BIOS samples, which had the longest estimated travel time, had higher

317

iron solubility than the GEOTRACES and Finokalia Research Station samples, which both short

318

estimated travel times. However, there is not a clear statistically significant relationship between

319

estimated travel time and iron solubility or total iron concentrations (Figure 1). This increased

320

solubility could be attributed to the decreases seen in pH as the travel time increases (Table S2).

321

The pH drops from circum neutral to less than four pH units after approximately 15 days in the

322

atmosphere, which corresponds to a marked increase in solubility.

323

3.2 Iron Composition

324

The results of the k-means cluster analysis and linear combination fitting reveal the aerosol

325

iron composition at the Finokalia Research Station to be uniform at the bulk level (Figure 2;

326

Table S3). At the GEOTRACES collection sites, each bulk sample was determined to be

327

compositionally distinct despite the fact that the collections sampled three sequential days. At the

328

bulk level, iron from the BIOS collection site was represented by two compositions. Individual

329

particles sampled from Finokalia Research Station all have the same composition; similarly,

330

individual particles sampled from the BIOS collection also exhibited a uniform composition. At

331

the GEOTRACES sample site, the composition of the individual particles sampled was much

332

more varied having one of three different iron compositions (Table S3).

333

In contrast to the GEOTRACES and BIOS samples, the Finokalia Research Station samples

334

were on average dominated by iron(III) oxides (>90%) (Figure 2). The GEOTRACES and BIOS

35 ACS Paragon Plus Environment

Page 17 of 31

Environmental Science & Technology

335

samples contained significantly lower fractions of iron(III) oxides (57% and 41%, respectively)

336

on average. In the GEOTRACES and BIOS aerosol samples, iron(III) oxides were found along

337

with lesser portions (70%) with the remaining fraction being iron(II) silicates, and GEOTRACES individual

343

particles contained iron(III) phosphates and iron(II & III) sulfates. (Figure 2)

344

3.3 Oxidation State

345

Iron in Saharan dust exhibits different oxidation states across the different sampling sites.

346

The Finokalia Research Station samples contained the most oxidized iron with more than 85% of

347

the iron as iron(III), while the BIOS and GEOTRACES aerosol samples contained

348

approximately 60% iron(III). The iron oxidation state, shown using the pre-edge centroid

349

position, is negatively correlated (R2=0.54) with estimated travel time, suggesting that the longer

350

an aerosol remains in the atmosphere, the more reduced the iron becomes (Figure 3). Although

351

iron(II) is generally considered more soluble than iron(III), the fractional solubility of iron did

352

not correlate (R2 = 0.14) with oxidation state (Table S2).

353

4.0 Discussion

354

The variation of iron chemistry in Saharan dust during atmospheric transport can be seen

355

most clearly in the oxidation state, with aerosol iron being more reduced with longer transport

356

times. The photo-reduction of iron during atmospheric transport could explain this trend. Photo-

357

reduction pathways for aerosol iron have been demonstrated for both ligand-assisted and

36 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 31

358

unassisted mechanisms in laboratory settings.77, 78 Once in aqueous form, the reducible fraction

359

of iron(III) can be quickly converted to iron(II) by ligand assisted photo-reduction mechanisms,

360

with this reducible fraction having important implications for bioavailability.78, 79 A correlation

361

between the oxidation state and the travel time is evident, but not strong (Figure 3; R2=0.54).

362

This could be attributed to differences within the different mineral classes we have defined. Even

363

within the class of iron(III) oxides, different minerals have variable susceptibilities to

364

photochemical processing.24 Alternatively, reduced iron could be an original component in the

365

BIOS and GEOTRACES samples regardless of travel time, but this would require a significant

366

source of reduced iron in Saharan soil, which is unlikely in this arid environment. Despite

367

iron(II) being more soluble than iron(III), there was no correlation between the oxidation state of

368

iron and the solubility. Although Zhu et al. (1997) did not find iron(II) sulfates, the general lack

369

or correlation between aerosol iron solubility and iron oxidation state is consistent.80 Reductive

370

processes, implied by the change in oxidation state, would require relatively short time frames in

371

these Saharan samples, on the order of one to three days. The short time scale of the reductive

372

processes is consistent with the findings of Majestic et al. (2007),44 although they did not find

373

iron(II) sulfates. If iron phases susceptible to photo-reduction are completely reduced in this

374

short time frame, then other transformation mechanisms, such as proton-induced iron dissolution,

375

may become increasingly important on longer time scales. Nitric acid suppresses the formation

376

of iron(II) at low pH, therefore pH can also act as a control of oxidation state of aerosol iron.81 In

377

these samples, nitrate is negatively correlated with the relative abundance of iron(II) (R2 = 0.75;

378

Table S4), where high concentrations of nitrate correspond to more iron(III). Nitrate has also

379

been suggested to significantly increase the reactivity of iron oxides.82 These complex

380

interactions suggest that the various particle aging mechanisms, such as acidic reactions and

37 ACS Paragon Plus Environment

Page 19 of 31

Environmental Science & Technology

381

photo-reduction, may not be working simultaneously. A shift in mechanisms offers a possible

382

explanation for the lack of a clear correlation between solubility and oxidation state of aerosol

383

iron in the long-range transport of dust aerosols.

384

The changes in oxidation state were accompanied by variations in the composition of iron in

385

Saharan air masses. Bulk aerosol iron from the GEOTRACES and BIOS locations contained on

386

average only ~50% iron(III) oxides. For aerosol samples, the individual particle composition was

387

often different from the bulk composition seen at a sampling site, suggesting that the

388

composition of the large particles (>1µm) is different from the fine particles (1 µm). X-ray fluorescence imaging shows that aerosol phosphorus is sometimes

400

collocated with iron and almost always with calcium (Figure S2). This suggests the iron

401

phosphates are likely formed as a result of the aggregation and subsequent acidification of

402

calcium phosphate containing particles and iron containing particles. At the BIOS site, the iron

403

sulfates account for nearly 50% of the total iron, but appear only in the bulk phases, suggesting

38 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 31

404

that iron sulfates are likely to be found in small particles (15 days), which

408

corresponds to the aerosols with the more soluble iron. This suggests that proton reactions are

409

solubilizing the iron during atmospheric transport. Our observations of iron sulfates36 and

410

phosphates provide further support that acidic reactions are occurring. In these Saharan dust

411

samples, a lower pH is correlated with more iron(II) (R2 = 0.50; Table S4). Because nitrate

412

seems to inhibit the formation of iron(II), other acidic species, such as sulfate, are likely

413

responsible for this correlation. This relationship helps support the hypothesis that iron(II)

414

sulfates are formed as a result of acidic reactions in the atmosphere. Sulfates have also been

415

shown to preferentially accumulate on aluminosilicate-rich dust if the dust air mass contacts a

416

sulfate-rich air mass during transport.35 Correlations between non-seasalt sulfate and soluble Fe

417

have been suggested as evidence for sulfuric acid reactions that solubilize iron.41,

418

correlations were not significant in our samples (R2 = 0.22); however, the absence of correlation

419

does not exclude sulfuric acid as a means of iron solubilization in our samples because non-

420

seasalt sulfate concentrations may not be an adequate indicator of aerosol acidity.37,

421

solubilization of iron by sulfuric acid might explain the presence of iron sulfates, which could

422

precipitate out of solution as these acidic reactions occur. However, these reactions can be

423

buffered by minerals such as carbonates found in Saharan dust,30, 85 so the buffer capacity of the

424

particles must be overcome for these reactions to proceed. The buffering effect of minerals

425

present in Saharan dust helps explain the finding of secondary phosphates from acidic reactions

83

These

84

The

39 ACS Paragon Plus Environment

Page 21 of 31

Environmental Science & Technology

426

as well32 as the iron(III) oxides in particles where acids were not able to overcome the buffering

427

capacity.

428

As the largest source of dust to the Atlantic Ocean, it is important to consider iron inputs

429

from the Sahara Desert. The variations in oxidation state suggest that a reductive process is

430

taking place during atmospheric transport while formation of iron phosphate and sulfate phases

431

offer possible support for proton reactions. However, these two together are not fully sufficient

432

to explain differences in aerosol iron solubility seen at the short and long range sampling

433

locations. For example, iron sulfate phases are generally considered a soluble form of iron,

434

however samples which contained significant iron sulfates never contain more than 10% soluble

435

iron, suggesting that the iron sulfates are not fully dissolved. This could be a result of the

436

heterogeneous nature of aerosols, where the most labile phases may not remain on the surface of

437

the particle where they can be readily dissolved. Furthermore, physical processes,22 particle

438

size,21 and ligand promoted iron dissolution,86 not explored in this study, are also factors that

439

may contribute to the variations in aerosol iron solubility that were observed at the different

440

sampling locations. Overall, these findings suggest that a combination of factors affects aerosol

441

iron solubility during long-distance atmospheric transport and emphasize the need to consider

442

reductive mechanisms as well as proton-induced solubilization of aerosol iron in modeling

443

studies.

40 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 31

444

Acknowledgments This material is based upon work supported by the National Science

445

Foundation under Grants OCE-1357375 (EDI), OCE-0929919 (WML), and OCE-1034764

446

(WML). The data used to produce these results is available upon request to the corresponding

447

author. Any opinions, findings, and conclusions or recommendations expressed in this material

448

are those of the authors and do not necessarily reflect the views of the National Science

449

Foundation. Use of the Advanced Photon Source and support to YF and BL are provided by

450

Argonne National Laboratory under the U.S. Department of Energy contract No. DE-AC02-

451

06CH11357. NM and KV acknowledge support from European Union (European Social Fund)

452

and Greek national funds through the Operational Program "Education and Lifelong Learning"

453

of the National Strategic Reference Framework Research Funding Program, ARISTEIA.

454

Supporting Information This article is accompanied by additional tables and figures available

455

in the Supporting Information. This information is available free of charge via the Internet at

456

http://pubs.acs.org/.

41 ACS Paragon Plus Environment

Page 23 of 31

Environmental Science & Technology

457

References

458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502

1. Moore, C. M.; Mills, M. M.; Arrigo, K. R.; Berman-Frank, I.; Bopp, L.; Boyd, P. W.; Galbraith, R.; Geider, R. J.; Guieu, C.; Jaccard, S. L.; Jickells, T. D.; La Roche, J.; Lenton, T. M.; Mahowald, N. M.; Maranon, E.; Marinov, I.; Moore, J. K.; Nakatsuka, T.; Oschlies, A.; Saito, M. A.; Thingstad, T. F.; Tsuda, A.; Ulloa, O., Processes and patterns of oceanic nutrient limitation. Nat. Geosci. 2013, 6, 701 - 710. 2. Martin, J. H., Glacial-Interglacial CO2 Change: The Iron Hypothesis. Paleoceanography 1990, 5, (1), 1-13. 3. de Jong, J.; Schoemann, V.; Lannuzel, D.; Croot, P.; de Baar, H.; Tison, J.-L., Natural iron fertilization of the Atlantic sector of the Southern Ocean by continental shelf sources of the Antarctic Peninsula. Journal of Geophysical Research-Biogeosciences 2012, 117. 4. Coale, K. H.; Johnson, K. S.; Chavez, F. P.; Buesseler, K. O.; Barber, R. T.; Brzezinski, M. A.; Cochlan, W. P.; Millero, F. J.; Falkowski, P. G.; Bauer, J. E.; Wanninkhof, R. H.; Kudela, R. M.; Altabet, M. A.; Hales, B. E.; Takahashi, T.; Landry, M. R.; Bidigare, R. R.; Wang, X. J.; Chase, Z.; Strutton, P. G.; Friederich, G. E.; Gorbunov, M. Y.; Lance, V. P.; Hilting, A. K.; Hiscock, M. R.; Demarest, M.; Hiscock, W. T.; Sullivan, K. F.; Tanner, S. J.; Gordon, R. M.; Hunter, C. N.; Elrod, V. A.; Fitzwater, S. E.; Jones, J. L.; Tozzi, S.; Koblizek, M.; Roberts, A. E.; Herndon, J.; Brewster, J.; Ladizinsky, N.; Smith, G.; Cooper, D.; Timothy, D.; Brown, S. L.; Selph, K. E.; Sheridan, C. C.; Twining, B. S.; Johnson, Z. I., Southern ocean iron enrichment experiment: Carbon cycling in high- and low-Si waters. Science 2004, 304, (5669), 408-414. 5. Buesseler, K. O.; Andrews, J. E.; Pike, S. M.; Charette, M. A., The effects of iron fertilization on carbon sequestration in the Southern Ocean. Science 2004, 304, (5669), 414-417. 6. Boyd, P. W.; Watson, A. J.; Law, C. S.; Abraham, E. R.; Trull, T.; Murdoch, R.; Bakker, D. C. E.; Bowie, A. R.; Buesseler, K. O.; Chang, H.; Charette, M.; Croot, P.; Downing, K.; Frew, R.; Gall, M.; Hadfield, M.; Hall, J.; Harvey, M.; Jameson, G.; LaRoche, J.; Liddicoat, M.; Ling, R.; Maldonado, M. T.; McKay, R. M.; Nodder, S.; Pickmere, S.; Pridmore, R.; Rintoul, S.; Safi, K.; Sutton, P.; Strzepek, R.; Tanneberger, K.; Turner, S.; Waite, A.; Zeldis, J., A mesoscale phytoplankton bloom in the polar Southern Ocean stimulated by iron fertilization. Nature 2000, 407, (6805), 695-702. 7. Coale, K. H.; Johnson, K. S.; Fitzwater, S. E., A massive phytoplankton bloom induced by an ecosystem-scale iron fertilization experiment in the equitorial Pacific Ocean. Nature 1996, 383, 495-501. 8. Park, S.; Nam, H.; Chung, N.; Park, J.; Lim, Y., The role of iron in reactive oxygen species generation from diesel exhaust particles. Toxicology in Vitro 2006, 20, (6), 851-857. 9. Shafer, M. M.; Perkins, D. A.; Antkiewicz, D. S.; Stone, E. A.; Quraishi, T. A.; Schauer, J. J., Reactive oxygen species activity and chemical speciation of size-fractionated atmospheric particulate matter from Lahore, Pakistan: an important role for transition metals. Journal of Environmental Monitoring 2010, 12, (3), 704-715. 10. Zhang, Y.; Schauer, J. J.; Shafer, M. M.; Hannigan, M. P.; Dutton, S. J., Source apportionment of in vitro reactive oxygen species bioassay activity from atmospheric particulate matter. Environmental Science & Technology 2008, 42, (19), 7502-7509. 11. Tao, F.; Gonzalez-Flecha, B.; Kobzik, L., Reactive oxygen species in pulmonary inflammation by ambient particulates. Free Radical Biology & Medicine 2003, 35, (4), 327-340. 12. Mahowald, N.; Baker, A. R.; Bergametti, G.; Brooks, N.; Duce, R. A.; Jickells, T. D.; Kubilay, N.; Prospero, J. M.; Tegen, I., Atmospheric global dust cycle and iron inputs to the ocean. Global Biogeochemical Cycles 2005, 19, (GB4025). 13. Jickells, T. D.; An, Z. S.; Andersen, K. K.; Baker, A. R.; Bergametti, G.; Brooks, N.; Cao, J. J.; Boyd, P. W.; Duce, R. A.; Hunter, K. A.; Kawahata, H.; Kubilay, N.; laRoche, J.; Liss, P. S.; Mahowald, N.; Prospero, J. M.; Ridgwell, A. J.; Tegen, I.; Torres, R., Global iron connections between desert dust, ocean biogeochemistry, and climate. Science 2005, 308, (5718), 67-71.

42 ACS Paragon Plus Environment

Environmental Science & Technology

503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550

Page 24 of 31

14. Engelstaedter, S.; Tegen, I.; Washington, R., North African dust emissions and transport. EarthScience Reviews 2006, 79, (1-2), 73 - 100. 15. Lenes, J. M.; Prospero, J. M.; Landing, W. M.; Virmani, J. I.; Walsh, J. J., A model of Saharan dust deposition to the eastern Gulf of Mexico. Mar. Chem. 2012, 134, 1-9. 16. Ganor, E.; Mamane, Y., Transport of Saharan dust across the Eastern Mediterranean. Atmospheric Environment 1982, 16, (3), 581-587. 17. Baker, A. R.; Weston, K.; Kelly, S. D.; Voss, M.; Streu, P.; Cape, J. N., Dry and wet deposition of nutrients from the tropical Atlantic atmosphere: Links to primary productivity and nitrogen fixation. Deep-Sea Res. Part I-Oceanogr. Res. Pap. 2007, 54, 1704-1720. 18. Herut, B.; Zohary, T.; Krom, M. D.; Mantoura, R. F. C.; Pitta, P.; Psarra, S.; Rassoulzadegan, F.; Tanaka, T.; Thingstad, T. F., Response of East Mediterranean surface water to Saharan dust: On-board microcosm experiment and field observations. Deep-Sea Res. Part II-Top. Stud. Oceanogr. 2005, 52, (2223), 3024-3040. 19. Herut, B.; Collier, R.; Krom, M. D., The role of dust in supplying nitrogen and phosphorus to the Southeast Mediterranean. Limnology and Oceanography 2002, 47, (3), 870-878. 20. Sholkovitz, E. R.; Sedwick, P. N.; Church, T. M.; Baker, A. R.; Powell, C. F., Fractional solubility of aerosol iron: Synthesis of a global-scale data set. Geochimica Et Cosmochimica Acta 2012, 89, 173-189. 21. Cartledge, B. T.; Marcotte, A. R.; Herckes, P.; Anbar, A. D.; Majestic, B. J., The impact of particle size, relative humidity, and sulfur dioxide on iron solubility in simulated atmospheric marine aerosols. Environmental Science & Technology 2015, 49, (12), 7179-7187. 22. Baker, A. R.; Jickells, T. D., Mineral particle size as a control on aerosol iron solubility. Geophysical Research Letters 2006, 33, (17). 23. Zhuang, G.; Yi, Z.; Duce, R. A.; Brown, P. R., Link between iron and sulphur cycles suggested by detection of Fe(II) in remote marine aerosols. Nature 1992, 335, (6360), 537-539. 24. Cornell, R. M.; Schindler, P. W., Photochemical dissolution of goethite in acid/oxalate solution. Clays and Clay Minerals 1987, 35, (5), 347-352. 25. Wiederhold, J. G.; Kraemer, S. M.; Teutsch, N.; Borer, P. M.; Halliday, A. N.; Kretzschmar, R., Iron isotope fractionation during proton-promoted, ligand-controlled, and reductive dissolution of goethite. Environmental Science & Technology 2006, 40, (12), 3787-3793. 26. Xu, N.; Gao, Y., Characterization of hematite dissolution affected by oxalate coating, kinetics and pH. Applied Geochemistry 2008, 23, (4), 783-793. 27. Buck, C. S.; Landing, W. M.; Resing, J. A.; Measures, C., The solubility and deposition of aerosol Fe and other trace elements in the North Atlantic Ocean: Observations from the A16N CLIVAR/CO2 repeat hydrography section. Mar. Chem. 2010, 120, (1-4), 57-70. 28. Buck, C. S.; Landing, W. M.; Resing, J. A.; Lebon, G. T., Aerosol iron and aluminum solubility in the northwest Pacific Ocean: Results from the 2002 IOC cruise. Geochemistry, Geophysics, Geosystems 2006, 7, (4). 29. Meskhidze, N.; Chameides, W. L.; Nenes, A., Dust and pollution: A recipe for enhanced ocean fertilization? Journal of Geophysical Research-Atmospheres 2005, 110, (D3). 30. Meskhidze, N.; Chameides, W. L.; Nenes, A.; Chen, G., Iron mobilization in mineral dust: Can anthropogenic SO2 emissions affect ocean productivity? Geophysical Research Letters 2003, 30, (21). 31. Journet, E.; Desboeufs, K. V.; Caquineau, S.; Colin, J. L., Mineralogy as a critical factor of dust iron solubility. Geophysical Research Letters 2008, 35, (7). 32. Nenes, A.; Krom, M. D.; Mihalopoulos, N.; Van Cappellen, P.; Shi, Z.; Bougiatioti, A.; Zarmpas, P.; Herut, B., Atmospheric acidification of mineral aerosols: a source of bioavailable phosphorus for the oceans. Atmospheric Chemistry and Physics 2011, 11, (13), 6265-6272. 33. Shi, Z.; Krom, M. D.; Bonneville, S.; Benning, L. G., Atmospheric processing outside clouds increases soluble iron in mineral dust. Environmental Science & Technology 2015, 49, (3), 1472-1477. 43 ACS Paragon Plus Environment

Page 25 of 31

551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596

Environmental Science & Technology

34. Shi, Z.; Krom, M. D.; Bonneville, S.; Baker, A. R.; Jickells, T. D.; Benning, L. G., Formation of iron nanoparticles and increase in iron reactivity in mineral dust during simulated cloud processing. Environmental Science & Technology 2009, 43, (17), 6592-6596. 35. Sullivan, R. C.; Guazzotti, S. A.; Sodeman, D. A.; Prather, K. A., Direct observations of the atmospheric processing of Asian mineral dust. Atmospheric Chemistry and Physics 2007, 7, 1213-1236. 36. Oakes, M.; Ingall, E. D.; Lai, B.; Shafer, M. M.; Hays, M. D.; Liu, Z. G.; Russell, A. G.; Weber, R. J., Iron Solubility Related to Particle Sulfur Content in Source Emission and Ambient Fine Particles. Environmental Science & Technology 2012, 46, (12), 6637-6644. 37. Seinfeld, J. H.; Pandis, S. N., Atmospheric chemistry and physics: from air pollution to climate change. John Wiley & Sons Inc.: New York, 2006. 38. Ito, A.; Feng, Y., Role of dust alkalinity in acid mobilizaiton of iron Atmospheric Chemistry and Physics 2010, 10, 9237-9250. 39. Shi, Z.; Krom, M. D.; Jickells, T. D.; Bonneville, S.; Carslaw, K. S.; Mihalopoulos, N.; Baker, A. R.; Benning, L. G., Impacts on iron solubility in the mineral dust by processes in the source region and the atmosphere: A review. Aeolian Research 2012, 5, 21-42. 40. Johnson, M. S.; Meskhidze, N., Atmospheric dissolved iron deposition to the global oceans: effects of oxalate-promoted dissolution, photochemical redox cycling, and dust mineralogy. Geoscientific Model Development 2013, 6, 1137 - 1155. 41. Baker, A. R.; Croot, P. L., Atmospheric and marine controls on aerosol iron solubility in seawater. Mar. Chem. 2010, 120, 4-13. 42. Blain, S.; Bonnet, S.; Guieu, C., Dissolved iron distribution in the tropical and subtropical South Eastern Pacific. Biogeosciences 2008, 5, 269-280. 43. Cornell, R. M.; Schwertmann, U., The iron oxides, structure, properties, reactions, occurence and uses. John Wiley: Hokoken, NJ, 1996. 44. Majestic, B. J.; Schauer, J. J.; Shafer, M. M., Application of synchrotron radiation for measurement of iron red-ox speciation in atmospherically processed aerosols. Atmospheric Chemistry and Physics 2007, 7, (10), 2475-2487. 45. Fittschen, U. E. A.; Meirer, F.; Streli, C.; Wobrauschek, P.; Thiele, J.; Falkenberg, G.; Pepponi, G., Characterization of atmospheric aerosols using Synchroton radiation total reflection X-ray fluorescence and Fe K-edge total reflection X-ray fluorescence-X-ray absorption near-edge structure. Spectrochimica Acta Part B - Atomic Spectroscopy 2008, 63, (12), 1489-1495. 46. Dabek-Zlotorzynska, E.; Kelly, M.; Chen, H.; Chakrabarti, C. L., Evaluation of capillary electrophoresis combined with a BCR sequential extraction for determining distribution of Fe, Zn, Cu, Mn, and Cd in airborne particulate matter. Analytica Chimica Acta 2003, 498, (1-2), 175-187. 47. Oakes, M.; Weber, R. J.; Lai, B.; Russell, A.; Ingall, E. D., Characterization of iron speciation in urban and rural single particles using XANES spectroscopy and micro X-ray fluorescence measurements: investigating the relationship between speciation and fractional iron solubility. Atmospheric Chemistry and Physics 2012, 12, (2), 745-756. 48. Hoffmann, P.; Dedik, A. N.; Ensling, J.; Weinbruch, S.; Weber, S.; Sinner, T.; Gutlich, P.; M., O. H., Speciation of iron in atmospheric aerosol samples. Journal of Aerosol Science 1996, 27, (2), 325-337. 49. Moffet, R. C.; Furutani, H.; Rodel, T. C.; Henn, T. R.; Sprau, P. O.; Laskin, A.; Uematsu, M.; Gilles, M. K., Iron speciation and mixing in single aerosol particles from Asian continental outflows. Journal of Geophysical Research 2012, 117, (D07204). 50. Weber, S.; Hoffmann, P.; Ensling, J.; Dedik, A. N.; Weinbruch, S.; Miehe, G.; Gutlich, P.; Ortner, H. M., Characterization of iron compounds from urban and rural aerosol sources. Journal of Aerosol Science 2000, 31, (8), 987-997.

44 ACS Paragon Plus Environment

Environmental Science & Technology

597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644

Page 26 of 31

51. Prospero, J. M.; Ginoux, P.; Torres, O.; Nicholson, S.; Gill, T. E., Environmental characterization of global sources of atmospheric soil dust identified with the NIMBUS 7 Total Ozone Mapping Spectrometer (TOMS) absorbing aerosol product. Reviews of Geophysics 2002, 40, (1), 1002. 52. Moore, J. K.; Doney, S. C.; Glover, D. M.; Fung, I. Y., Iron cycling and nutrient-limitation patterns in surface waters of the World Ocean. Deep-Sea Research Part II - Tropical Stidies in Oceanography 2002, 49, (1-3), 463-507. 53. Moore, J. K.; Doney, S. C.; Lindsay, K., Upper ocean ecosystem dynamics and iron cycling in a global three-dimensional model. Global Biogeochemical Cycles 2004, 18, (4), 21. 54. Sedwick, P. N.; Church, T. M.; Bowie, A. R.; Marsay, C. M.; Ussher, S. J.; Achilles, K. M.; Lethaby, P. J.; Johnson, R. J.; Sarin, M. M.; McGillicuddy, D. J., Iron in the Sargasso Sea (Bermuda Atlantic Timeseries Study region) during summer: Eolian imprint, spatiotemporal variability, and ecological implications. Global Biogeochemical Cycles 2005, 19, (4). 55. Zhenglong, T.; Ollivier, P.; Veron, A.; Church, T. M., Atmospheric Fe deposition modes at Bermuda and the adjacent Sargasso Sea. Geochemistry Geophysics Geosystems 2008, 9, (8), Q08007. 56. Markaki, Z.; Oikonomou, K.; Kocak, M.; Kouvarakis, G.; Chaniotaki, A.; Kubilay, N.; Mihalopoulos, N., Atmospheric deposition of inorganic phosphorus in the Levantine Basin, eastern Mediterranean: Spatial and temporal variability and its role in seawater productivity. Limnology and Oceanography 2003, 48, (4), 1557-1568. 57. Koulouri, E.; Saarikoski, S.; Theodosi, C.; Markaki, Z.; Gerasopoulos, E.; Kouvarakis, G.; Mäkelä, T.; Hillamo, R.; Mihalopoulos, N., Chemical composition and sources of fine and coarse aerosol particles in the Eastern Mediterranean. Atmospheric Environment 2008, 42, (26), 6542-6550. 58. Shelley, R. U.; Morton, P. L.; Landing, W. M., Elemental ratios and enrichment factors in aerosols from the US-GEOTRACES north Atlantic transects. Deep-Sea Research Part I-Tropical Studies in Oceanography 2015, 116, 262-272. 59. Baker, A. R.; Jickells, T. D.; Witt, M.; Linge, K. L., Trends in the solubility of iron, aluminium, manganese and phosphorus in aerosol collected over the Atlantic Ocean. Mar. Chem. 2006, 98, (1), 4358. 60. Izquierdo, R.; Benitez-Nelson, C. R.; Masque, P.; Castillo, S.; Alastuey, A.; Avila, A., Atmospheric phosphorus deposition in a near-coastal rural site in the NE Iberian Peninsula and its role in marine productivity. Atmospheric Environment 2012, 49, 361-370. 61. Draxier, R. R.; Hess, G. D., An overview of the HYSPLIT_4 modelling system for trajectories, dispersion and deposition. Australian Meteorological Magazine 1998, 47, (4), 295-308. 62. Kalivitis, N.; Gerasopoulos, E.; Vrekoussis, M.; Kouvarakis, G.; Kubilay, N.; Hatzianastassiou, N.; Vardavas, I.; Mihalopoulos, N., Dust transport over the eastern Mediterranean derived from Total Ozone Mapping Spectrometer, Aerosol Robotic Network, and surface measurements. Journal of Geophysical Research-Atmospheres 2007, 112, (D3). 63. Mahowald, N.; Albani, S.; Kok, J. F.; Engelstaedter, S.; Scanza, R.; Ward, D. S.; Flanner, M. G., The size distribution of desert dust aerosols and its impact on the Earth system. Aeolian Research 2014, 15, 53-71. 64. Stookey, L. L., Ferrozine - A new spectrophotometric reagent for iron. ANALYTICAL CHEMISTRY 1970, 42, (7). 65. Kadko, D.; Landing, W. M.; Shelley, R. U., A novel tracer technique to quantify the atmospheric flux of trace elements to remote ocean regions. J. Geophys. Res.-Oceans 2015, 119. 66. Morton, P. L.; Landing, W. M.; Hsu, S. C.; Milne, A.; Aguilar-Islas, A. M.; Baker, A. R.; Bowie, A. R.; Buck, C. S.; Gao, Y.; Gichuki, S.; Hastings, M. G.; Hatta, M.; Johansen, A. M.; Losno, R.; Mead, C.; Patey, M. D.; Swarr, G.; Vandermark, A.; Zamora, L. M., Methods for the sampling and analysis of marine aerosols: results from the 2008 GEOTRACES aerosol intercalibration experiment. Limnology and Oceanography 2013, 11, 62-78. 45 ACS Paragon Plus Environment

Page 27 of 31

645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691

Environmental Science & Technology

67. Bardouki, H.; Liakakou, H.; Economou, C.; Sciare, J.; Smolik, J.; Zdimal, V.; Eleftheriadis, K.; Ladaridis, M.; Dye, C.; Mihalopoulos, N., Chemical composition of size-resolved atmospheric aerosols in the eastern Mediterranean during summer and winter. Atmospheric Environment 2003, 37, (2), 195-208. 68. Fountoukis, C.; Nenes, A., ISORROPIA II: a computationally efficient thermodynamic equilibrium model for K+-Ca2+-Mg2+-Nh(4)(+)-Na+-SO42--NO3--Cl--H2O aerosols. Atmospheric Chemistry and Physics 2007, 7, (17), 4639-4659. 69. Nenes, A.; Pandis, S. N.; Pilinis, C., ISORROPIA: A new thermodynamic equilibrium model for multiphase multicomponent inorganic aerosols. Aquat. Geochem. 1998, 4, (1), 123-152. 70. Ingall, E.; Brandes, J.; Diaz, J.; de Jonge, M.; Paterson, D.; McNulty, I.; Elliott, W.; Northrup, P., Phosphorus K-edge XANES spectroscopy of mineral standards. Journal of Synchrotron Radiation 2011, 18, 189-197. 71. Wilke, M.; Farges, F.; Petit, P. E.; Brown, G. E.; Martin, F., Oxidation state and coordination of Fe in minerals: An FeK-XANES spectroscopic study. American Mineralogist 2001, 86, (5-6), 714-730. 72. Bajt, S.; Sutton, S. R.; Delaney, J. S., X-ray microprobe analysis of iron oxidation-states in silicates and oxides using X-ray -absorption near-edge structure (XANES). Geochimica Et Cosmochimica Acta 1994, 58, (23), 5209-5214. 73. Ravel, B.; Newville, M., ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT. Journal of Synchrotron Radiation 2005, 12, 537-541. 74. Marcus, M. A.; Lam, P. J., Visualising Fe speciation diversity in ocean particulate samples by micro X-ray absorption near-edge spectroscopy. Environmental Chemistry 2014, 11, (1), 10-17. 75. Ingall, E.; Diaz, J.; Longo, A.; Oakes, M.; Finney, L.; Vogt, S.; Lai, B.; Yager, P. L.; Twining, B. S.; Brandes, J., Role of biogenic silica in the removal of iron from Antarctic Seas. Nature Commun. 2013, 4:1981 76. Hesterberg, D., Chapter 11 - Macroscale Chemical Properties and X-Ray Absorption Spectroscopy of Soil Phosphorus. In Developments in Soil Science, Balwant, S.; Markus, G., Eds. Elsevier: 2010; Vol. Volume 34, pp 313-356. 77. Siefert, R. L.; Webb, S. M.; Hoffmann, M. R., Determination of photochemically available iron in ambient aerosols. Journal of Geophysical Research 1996, 10, (D9), 14441-14449. 78. Siefert, R. L.; Pehkonen, S. O.; Erel, Y.; Hoffmann, M. R., Iron photochemistry of aqueous suspensions of ambient aerosol with added organic acids. Geochimica Et Cosmochimica Acta 1994, 58, (15), 3217-3279. 79. Chen, Y.; Siefert, R. L., Determination of various types of labile atmospheric iron over remote oceans. Journal of Geophysical Research-Atmospheres 2003, 108, (D24). 80. Zhu, X. R.; Prospero, J. M.; Millero, F. J., Diel variability of soluble Fe (II) and soluble total Fe in North African dust in the trade winds at Barbados. Journal of Geophysical Research 1997, 102, (D17), 21297 - 21305. 81. Cwiertny, D. M.; Baltrusaitis, J.; Hunter, G. J.; Laskin, A.; Scherer, M. M.; Grassian, V. H., Characterization and acid-mobilization study of iron-containing mineral dust source materials. Journal of Geophysical Research 2008, 113, (D05202). 82. Kong, L. D.; Zhao, X.; Sun, Z. Y.; Yang, Y. W.; Fu, H., B.; Zhang, S. C.; Cheng, T. T.; Yang, X.; Want, L.; Chen, J. M., The effects of nitrate on the heterogeneous uptake of sulfur dioxide on hematite. Atmospheric Chemistry and Physics 2014, 14, 9451-9467. 83. Srinivas, B.; Sarin, M. M.; Kumar, A. K., Impact of anthropogenic sources on aerosol iron solubility over the Bay of Bengal and the Arabian Sea. Biogeochemistry 2012, 110, 257-268. 84. Hennigan, C. J.; Izumi, J.; Sullivan, A. P.; Weber, R. J.; Nenes, A., A critical evaluation of proxy methods use to estimate the acidity of atmospheric particles. Atmospheric Chemistry and Physics 2015, 15, 2775-2790.

46 ACS Paragon Plus Environment

Environmental Science & Technology

692 693 694 695 696 697

Page 28 of 31

85. Ito, A.; Feng, Y., Iron mobilization in North African Dust. Procedia Environmental Sciences 2011, 6, 27-34. 86. Wozniak, A. S.; Shelley, R. U.; Sleighter, R. L.; Abdulla, H. A. N.; Morton, P. L.; Landing, W. M.; Hatcher, P. G., Relationships among aerosol water soluble organic matter, iron and aluminum in European, North African, and Marine air masses from the 2010 US GEOTRACES cruise. Mar. Chem. 2013, 154, 24-33.

698 699 700

47 ACS Paragon Plus Environment

Page 29 of 31

Environmental Science & Technology

701 10%

2000

8%

1500

6%

1000

4%

500

2%

0

0%

Soluble Fe (%)

Total Fe (ng/m3)

2500

Travel Time (days) Total Fe (ng/m3)

Fe Solubility (%)

702 703

Figure 1: Aerosol samples from the Finokalia Research Station ( ) had the shortest estimated

704

travel times, followed by the GEOTRACES samples ( ), and the BIOS samples ( ). While the

705

BIOS sample consistently had the highest aerosol iron solubility, there was no significant

706

correlation between the estimated travel time and either total iron (R2 = 0.35) or fractional iron

707

solubility (R2 = 0.18).

708 709

48 ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 31

710 1

Factional Relative Abundance

0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

711

Finokalia Bulk

Finokalia Particle

GEOTRACES GEOTRACES Bulk Particle

BIOS Bulk

BIOS Particle

712

Figure 2: The mean bulk and individual particle iron composition of samples from each

713

sampling site. Iron(III) oxides ( ) dominate in the Finokalia Research Station samples, with

714

traces of iron(II) silicates ( ) and iron (II) sulfates ( ) in the bulk and individual particles,

715

respectively (7 bulk sample spectra; 11 individual particle spectra). In the GEOTRACES

716

samples, iron(III) sulfates ( ) and iron(III) phosphates ( ) appeared (3 bulk sample spectra; 17

717

individual particle sample spectra). The bulk samples from the BIOS site contained significant

718

contributions from iron sulfates; while individual particles were a mix of iron(II) silicates and

719

iron(III) oxides (8 bulk sample spectra; 18 individual particle spectra).

720 721

49 ACS Paragon Plus Environment

Page 31 of 31

Environmental Science & Technology

722 30.0

R2 = 0.54

Travel Time (days)

25.0 20.0 15.0 10.0 5.0 0.0 7.1134 7.1136 7.1138

7.114

7.1142 7.1144 7.1146

PrePre-Edge Centroid (keV)

723 724

Figure 3: Travel time, estimated using HYSPLIT back trajectories, is negatively correlated with

725

the pre-edge centroid, a reliable indicator of oxidation state that increases as iron transitions from

726

iron(II) to iron(III).

727 728

50 ACS Paragon Plus Environment