4 Years after the Deepwater Horizon Spill: Molecular Transformation

Jul 28, 2016 - ... C.; Mauseth , G.; Challenger , G.; Taylor , E. Extent and degree of shoreline oiling: Deepwater Horizon oil spill, Gulf of Mexico, ...
0 downloads 0 Views 5MB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Four Years after the Deepwater Horizon Spill: Molecular Transformation of Macondo Well Oil in Louisiana Salt Marsh Sediments Revealed by FT-ICR Mass Spectrometry Huan Chen, Aixin Hou, Yuri E. Corilo, Qianxin Lin, Jie Lu, Irving A Mendelssohn, rui zhang, Ryan Patrick Rodgers, and Amy M McKenna Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b01156 • Publication Date (Web): 28 Jul 2016 Downloaded from http://pubs.acs.org on July 29, 2016

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

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

Page 1 of 28

1 2 3 4

Environmental Science & Technology

Four Years after the Deepwater Horizon Spill: Molecular Transformation of Macondo Well Oil in Louisiana Salt Marsh Sediments Revealed by FTICR Mass Spectrometry

5 6 7 8 9 10 11 12 13 14

Huan Chen†, Aixin Hou‡, Yuri E. Corilo†, Qianxin Lin§, Jie Lu┴, Irving A. Mendelssohn§, Rui Zhang‡, Ryan P. Rodgers†,┴, and Amy M. McKenna†,* †

National High Magnetic Field Laboratory, Florida State University ,1800 East Paul Dirac Dr., Tallahassee, FL 32310-4005 ‡

Department of Environmental Sciences, School of the Coast and Environment, Louisiana State University, Baton Rouge, LA 70803 USA. §

Department of Oceanography and Coastal Sciences, School of the Coast and Environment, Louisiana State University, Baton Rouge, LA 70803 USA.

15 16

┴Future Fuels Institute, Florida State University, 1800 East Paul Dirac Drive,

17 18

*To whom correspondence should be addressed. Tel.: +1 850 644 4809; Fax +1 850 644 1366. E-mail address: [email protected]

19

Re-Submitted to Environ. Sci. Technol. (es-2015-015659m)

20

■ ABSTRACT

21

Gulf of Mexico saltmarsh sediments were heavily impacted by Macondo Well Oil (MWO)

22

released from the 2010 Deepwater Horizon (DWH) oil spill. Detailed molecular-level

23

characterization of sediment extracts collected over 48 months post-spill highlights the

24

chemical complexity of highly polar, oxygen-containing compounds that remain

25

environmentally persistent. Electrospray ionization (ESI) Fourier transform ion cyclotron

26

resonance mass spectrometry (FT-ICR MS) combined with chromatographic pre-

27

fractionation correlates bulk chemical properties to elemental compositions of oil

28

transformation products as a function of time. Carboxylic acid incorporation into parent

29

MWO hydrocarbons detected in sediment extracts (corrected for mass loss relative to

30

C30 hopane), proceeds with an ~3-fold increase in O2 species after 9 months to a

31

maximum of ~5.5-fold after 36 months, compared to the parent MWO. More importantly,

Tallahassee, Florida 32310-4005, United States

ACS Paragon Plus Environment

July 5, 2016

Environmental Science & Technology

32

higher-order oxygenated compounds (O4-O6) not detected in the parent MWO, increase

33

in relative abundance with time, as lower order oxygenated species are transformed into

34

highly polar, oxygen-containing compounds (Ox, where x>3). Here, we present the first

35

molecular-level characterization of temporal compositional changes that occur in

36

Deepwater Horizon derived oil contamination deposited in a saltmarsh ecosystem from

37

9 to 48 mos post-spill, and identify highly oxidized Macondo well oil compounds that are

38

not detectable by routine GC-based techniques.

39 40 41 42 43 44 45

Keywords: Deepwater Horizon, petroleum, petroleomics, crude oil, Fourier transform, ion cyclotron resonance, gas chromatography, time series

2

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

46

Environmental Science & Technology

■ INTRODUCTION

47

The explosion and sinking of the Deepwater Horizon (DWH) oil platform in the

48

summer of 2010 released an judicially determined 3.1 million barrels of crude oil into the

49

Gulf of Mexico ecosystem,1 the largest marine oil spill in United States’ history.2 Point-

50

source releases of anthropogenic pollution, such as oil spills, threaten coastal habitats3

51

and have been linked to immediate animal die-offs, altered animal behaviors,4,

52

persistence of oil-derived compounds in coastal food webs.6, 7 The oil released from the

53

Macondo Well impacted over 1000 linear miles of shoreline, and contaminated 463 total

54

miles of saltmarsh (Louisiana: 436 miles), Mississippi: 21 miles) and Alabama: 6 miles).8

55

Coastal salt marshes provide vital services to coastal ecosystems, including shoreline

56

protection, carbon sequestration, and water quality enhancement, and generate

57

revenue from fisheries, agriculture, and recreational tourism.6,

58

particularly sensitive to oil contamination due to low wave action and high capacity of

59

sediments to absorb organic contaminants,11 and accelerated shoreline erosion has

60

been reported since the DWH event.6

61

DWH-derived

compounds

that

reached

the

9, 10

shore

5

and

Salt marshes are

underwent

rapid

62

biodegradation in salt marsh sediments,12-14 but the long-term impact of environmentally

63

persistent petroleum compounds and their weathering products remains unknown.15-17

64

Conventional characterization of Deepwater Horizon contamination relies almost

65

exclusively on gas chromatography-based techniques18-26 that only account for a

66

fraction of the mass of pre-spill Macondo Well Oil (MWO) on a molecular basis.27 Abiotic

67

and biotic weathering processes create oxygenated transformation products that

68

account for 60~80% of extractable material19, which further challenge GC-based

69

characterization techniques due to increased polarity and decreased volatility.19, 20, 28, 29

70

Recent studies address the toxicological impact of weathered oil contaminants not

71

detected by GC-based techniques4,

72

techniques to identify molecular transformations that occur to petroleum hydrocarbons

73

after environmental release.

28

and highlight the need for advanced analytical

74

Recent GC-based reports suggest that the level of hydrocarbons extracted from

75

salt marsh samples from Barataria Bay had reached background levels within 18 mos

76

post-spill mostly due to biodegradation by indigenous microbial communities.6,

3

ACS Paragon Plus Environment

12

Environmental Science & Technology

77

However, a more recent report found that oil residues remained detectable up to 36

78

months in the top 2 cm of sediment of heavily oiled marshes and have been heavily

79

degraded.26 GC-based techniques are limited to compounds with volatility below ~400

80

°C to prevent column degradation. The formation of oxygenated transformation products

81

through biodegradation/photo-oxidation decreases the applicability of GC-based

82

techniques due to the low volatility and high polarity of MWO oxidized products.

83

The compositional complexity of crude oil challenges all analytical techniques,

84

and the term “unresolved complex mixture” (UCM) describes the raised baseline hump

85

observed in gas chromatograms of petroleum.30 Recent reports demonstrate the

86

potential of using Fourier Transform Ion Cyclotron Resonance Mass Spectrometry (FT-

87

ICR MS) to target heavily weathered petroleum compounds without boiling point

88

limitations27, 31-33 FT-ICR MS routinely achieves ultrahigh resolving power (m/∆m50% =

89

1,000,000 at m/z 500, in which ∆m50% is the mass spectral peak width at half-maximum

90

peak height), sufficient to separate and identify more than 80,000 unique elemental

91

compositions in a single mass spectrum for weathered oil29 and identify molecular-level

92

transformations that occur during degradation.27, 29 Early environmental applications of

93

FT-ICR MS report an increased abundance of high molecular weight (>200-300 Da),

94

highly oxidized transformation products.34-37 Lemkau et. al applied ultrahigh resolution

95

FT-ICR MS to catalogue compositional changes that occur to oil residues derived from

96

the 2007 M/V Cosco Busan heavy fuel oil (HFO) spill in San Francisco Bay over 617

97

days and observed an increased abundance of condensed aromatic, oxygen-containing

98

compounds relative to the parent HFO.31,

99

applied to elemental compositions derived from FT-ICR MS characterization of field

100

samples identified a suite of recalcitrant polar petroleum markers that facilitate source

101

identification of oil spilled from two possible storage tanks ruptured when the M/V Cosco

102

Busan struck the San Francisco--Oakland Bay Bridge.33 More recently, FT-ICR MS

103

characterization was applied to parent Deepwater Horizon oil,27 and identified unique

104

ketone transformation products of weathered oil that reached Pensacola Beach in

105

2010.32

38

Principal Component Analysis (PCA)

106

More than four years after the Deepwater Horizon disaster, salt marsh sediments

107

along the Gulf of Mexico coast remain contaminated with Macondo Well oil compounds.

4

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Environmental Science & Technology

108

We combine comprehensive two-dimensional gas chromatography with mass

109

spectrometry (GC×GC MS) and flame ionization detection (GC×GC-FID), and

110

electrospray ionization (ESI) FT-ICR mass spectrometry to Deepwater Horizon oil

111

contamination extracted from the same location in northern Barataria Bay from 9 to 48

112

mos post spill and catalogue oxidation patterns that occur to oil compounds in salt

113

marsh sediments. Highly polar (acidic), oxygen-containing hydrocarbons are compared

114

to pre-spill oil, and we identify recalcitrant oxygen-containing compounds derived from

115

Deepwater Horizon oil degradation at the molecular level.

116 117

■ EXPERIMENTAL METHODS

118

Sample Collection. Figure 1 shows a map of northern Barataria Basin of Louisiana,

119

USA, approximately 8 km x 5 km (coordinates N 29.44060° - 29.47459°, W 89.88492° -

120

89.94647°), with 21 sampling locations (seven each that received heavy oiling (Sites 1-

121

7), moderate oiling (sites 8-14), and no observed oiling or reference (sites 15-21), based

122

on the Shoreline Cleanup Assessment Technique (SCAT) Program (28 April 2010), field

123

observations, and total petroleum hydrocarbon measurements.39, 40 No cleanup events

124

at the sampling sites were noted during this period. Surface sediments (0-2 cm) were

125

collected in pre-autoclaved glass jars and frozen (-80 °C) at Louisiana State University,

126

LA. For each time point, all seven samples from reference, moderately, or heavily oiled

127

sites were collected and homogenized respectively to obtain a composite sample.

128

Heavily oiled sites were collected at 9, 18, 24, 31, 36, 41, 43 and 48 mos post-spill.

129

Moderately oiled sites sampled at 9, 41 and 48 mos from seven moderately oiled

130

locations, and reference sites collected and homogenized from seven reference

131

locations (Table S-1). The composite samples were shipped in EPA certified glass jars

132

to the National High Magnetic Field Laboratory (Tallahassee, FL, USA) for analysis.

133 134

Sediment Extraction. Approximately 20 g of sediment were homogenized with an

135

equal weight of sodium sulfate drying agent, loaded into a 30 x 100 mm cellulose

136

thimble, and Soxhlet extracted with dichloromethane for 4 h, desolvated under dry

137

nitrogen (N2) and weighed.41

5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 28

138

Bulk Elemental Analysis. Complete experimental methods can be found in the

139

supporting material. Briefly, elemental analysis was performed on MWO and sediment

140

extracts collected from heavily oiled sites at 9, 18, 24, 31, 36, 41, and 48 mos post spill

141

on a Thermo Finnigan Elemental Analyzer (FLASH EA 112, San Jose, CA, USA). Due

142

to sample limitations, moderately oiled and reference sites were not analyzed. Detailed

143

description can be found in Supporting Information.

144

GC×GC-MS and GC×GC-FID. Comprehensive two-dimensional gas chromatagraphy

145

(GC×GC) and

146

collected from heavily oiled sites at 9, 18, 24, 31, 36, 41, and 48 mos post spill. Each

147

sample was analyzed with GC×GC coupled to a time-of-flight mass spectrometry (ToF-

148

MS) for the identification of petroleum hydrocarbons. GC×GC with a flame ionization

149

detector (GC×GC-FID) was used for the quantitation of the petroleum biomarkers. The

150

fraction losses of total oil and individual compounds were calculated by nomalizing the

151

peak area of individual compound to that of 17α(H),21β(H)-hopane (herein refered to as

152

C30-hopane) and comparing to the corresponding values to MWO.42,

153

experimental details and calculations are provided in Supporting Information.

154

Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Macondo well

155

crude oil was provided by BP oil company through the Gulf Coast Restoration

156

Organization. All solvents were HPLC grade (Sigma-Aldrich Chemical Co., St. Louis,

157

MO). Prior to mass spectral analysis, MWO and extracts were diluted in HPLC grade

158

toluene to make a stock solution (1 mg/ mL) that was further diluted with equal parts

159

(vol: vol) methanol spiked with 2 % (by volume) formic acid for positive ion electrospray

160

ionization (ESI) or 0.25 % (by volume) tetramethylammonium hydroxide (CAS no. 75-

161

59-2, TMAH, 25 % by weight in methanol) for negative ion ESI FT-ICR MS. Sample

162

preparation method, ionization source conditions, and detailed FT-ICR MS conditions

163

can be found in Supporting Information.

elemental analysis were performed on MWO and sediment extracts

164 165

6

ACS Paragon Plus Environment

43

Complete

Page 7 of 28

Environmental Science & Technology

166

■ RESULTS AND DISCUSSION

167 168

Extraction of MWO from Saltmarsh Sediments. Table S-1 shows the total extracted

169

mass of oil isolated from 20 g each of heavily oiled, moderately oiled, and reference

170

sites.

171

sediments occurs between 9 mos (26 wt%) and 48 mos (2.4 wt%), indicating oil removal

172

from marshes with time. This suggests the majority of MWO deposition occurred in the

173

early months after the spill, and oil residues were removed or transformed with time.

174

However, after 48 mos the total extracted oil is ~ 438 times greater than reference sites,

175

indicating the long-term persistence of oil-like compounds in the marshes. Sediments

176

that received moderate oiling contained approximately the same mass of oil residue

177

after 9 mos (0.470 g), compared to heavily oiled sites after 48 mos (0.482 g). This

178

suggests that residual MWO contamination is still present in salt marsh sediments four

179

years after the spill in concentrations that exceed non-oiled reference sites, even though

180

the majority of the oil has been removed or degraded.

A 10-fold decrease in the total amount of oil extracted from heavily oiled

181 182

(Insert Figure 1 here)

183 184

Gas Chromatography of MWO and Sediment Extracts. Figure S-1 shows gas

185

chromatograms of MWO and petroleum residues collected from 9 to 48 mos at heavily

186

oiled saltmarsh sites. In the first 9 mos, the complete loss of low molecular weight

187

aromatic hydrocarbons and normal saturated hydrocarbons (below ~nC20) through

188

evaporation, water-washing and other weathering processes occurred relative to the

189

MWO. After 9 mos, the appearance of an unresolved complex mixture occurs based on

190

GC analysis, and by 48 mos, few or almost no heavy alkanes are identifiable by GC

191

analysis above the UCM (Figure S-1), in agreement with previous GC-based studies on

192

the DWH spill, where the removal of n-alkanes, isoprenoids, and two-ring PAHs were

193

reported in beach samples collected within 18 mos of the spill.19,

194

sediment extracts indicate the presence of weathered oil components, but detection is

195

limited to compounds that are volatile below ~400 °C.27, 30

7

ACS Paragon Plus Environment

24

GC analysis of

Environmental Science & Technology

196

Oil Depletion Estimated by GC×GC-FID Analysis. Each GC×GC-FID image was

197

base-plane subtracted and normalized to the conserved biomarker C30-hopane43 to

198

estimate degradation in sediment extracts.42, 44, 45 Figure 2 shows the calculated loss of

199

total petroleum hydrocarbons (TPH) and chrysenes (C1-C3) normalized to C30 hopane

200

for 9, 24, 36 and 48 mos. After 9 mos, ~45% of the parent MWO is removed, and after

201

48 mos, ~63% of the TPHs have been removed. The percentage of total oil and

202

individual compound losses assume that the C30-hopane is conserved across the

203

sampling time, although C30-hopane can be degraded under certain conditions.3-5

204

Between 36 and 48 mos, no significant weathering loss is observed, which indicates

205

either oil is continually depleted up to 36 mos and then slows down, or C30-hopane

206

begins to degrade with time. The long term fate of heavily degraded oil relies on reliable

207

fingerprinting methods when the biomarker fidelity is questionable, which is the subject

208

of this study.

209

For chrysenes, the extent of loss increases with degree of alkylation. Chrysene is

210

depleted by 58% after 9 mos, but only 64% after 48 mos. However, ~88% of C3-

211

chrysene is lost at 48 mos (Figure 2). The loss of alkyl-substituted chrysenes prior to

212

less alkylated analogs corresponds to photo-degradation rather than biodegradation.46

213

No obvious difference of Chrysene losses are observed from 9 to 48 mos.

214

(Insert Figure 2 here)

215

Elemental Analysis of Heavily Contaminated Sediment Extracts. Bulk elemental

216

analysis of sediments that received heavy oiling identify global changes in weight

217

percent carbon, hydrogen, nitrogen, oxygen and sulfur (Table S-2) compared to parent

218

MWO. The total mass fraction of sulfur increases after 48 mos to 0.9% by weight

219

compared to 0.5% by weight in parent MWO. Nitrogen content remains constant from

220

parent MWO (0.3% by weight) to 48 mos (0.3% by weight). The most dramatic change

221

in bulk composition occurs in the oxygen content of field samples. The total mass

222

fraction for samples collected in January 2011 indicates a more than 7-fold increase in

223

oxygen (3.9% by weight) compared to the parent MWO (0.5% by weight) after 9 mos,

224

which corresponds to oxidative transformation products.19, 20, 31, 32, 38 The oxygen content

225

increases to 4.7% by weight after 24 mos, but remains relatively constant across the

226

time series. The significant increase of oxygen content indicates oxidative processes

8

ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

Environmental Science & Technology

227

dominated oil weathering during the first few months after the spill, and produced

228

oxidized transformation products that persist in the environment.19 Therefore, we focus

229

on characterization of oxygen-containing compounds and molecular transformations

230

that occur as a function of time.

231 232

Acidity Increases in Field Samples Revealed by Aminopropyl Silica (APS)

233

Isolation. Bulk compositional changes indicate that oxygen content is increasing, but do

234

not identify the chemical functionality of oxygen transformation products. Since toxicity

235

changes with chemical functionality, identification of acidic functional groups is critical to

236

the long-term ecological impact of MWO in coastal areas. Previous studies identified

237

ketone transformation products in MWO-affected beach sands, and naphthenic acids

238

inherent in crude oil have known toxicity to marine bacteria,47 phytoplankton,48

239

zooplankton,49 plants,50 fish,51, 52 and mammals53 at low concentrations (2.5 – 5 mg L-

240

1 54

241

water-solubility.55 Isolation and fractionation of crude oil compounds on aminopropyl

242

silica (APS) has been reported to identify acidic functional groups in parent MWO and

243

selected field samples.56 Briefly, carboxylic acid-containing compounds interact via

244

strong hydrogen-bonding with APS. Non-acidic compounds elute with dichloromethane

245

(DCM) in fraction 1, and moderately acidic species elute with DCM-methanol in fraction

246

2. Carboxylic acids elute in fraction 3 with the addition of formic acid.56 Table 1 shows

247

fractional percent yields and total percent recovery for APS fractions for parent MWO

248

and field samples collected 9 mos, 36 mos and 48 mos postspill. Due to sample

249

limitation, we did not have enough extracted materials to run the separation in triplicates.

250

The previous published data show that the standard error of APS extraction is no

251

greater than 1.1% based on triplicate analysis.56 Parent MWO is a light, Louisiana oil

252

with high API° gravity (35.2)57, and low total acid number, and the majority of the

253

compounds (62%) recovered as nonpolar compounds (i.e., non-acids) in fraction 1 and

254

less than 1% elute in fraction 2 (0.7%) and fraction 3 (0.42%).15, 57 (Parent MWO is a

255

light crude with a high abundance of highly volatile, low molecular weight petroleum

256

hydrocarbons at atmospheric conditions, and approximately ~17% of MWO compounds

257

evaporated during the desolvation process determined independently. Therefore, the

),

and low molecular weight naphthenic acids are highly toxic47 due to increased

9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 28

258

total recovered yield for parent MWO includes the total mass lost through evaporation

259

under dry nitrogen plus the dry weight of each fraction.) After 9 mos, 3% of compounds

260

elute as moderately acidic species, and ~6% contain one or more carboxylic acid

261

moieties. Between 36 and 48 mos, moderately acidic compounds remain constant

262

between ~3.8-4%, and carboxylic acids increase slightly from ~8% to ~9% yield. This

263

could indicate secondary oxidation of moderately acidic compounds to create

264

compounds

265

characterization and future toxicity of heavily oxidized, acidic transformation products

266

will probe the long-term impact of these compounds to saltmarsh ecosystems.

that

contain

one

or more

carboxylic

acids.

Detailed

molecular

267 268 269

270 271 272 273 274 275 276

Table 1. Percent Recovery of Each Fraction and Total Percent Recovery for APS Acid Extractions on MWO and Selected Field Samples MWO 9 mos 36 mos 48 mos % of Fraction1 a

62.72

70.10

73.80

75.18

% of Fraction2 b

0.70

3.05

3.82

4.00

% of Fraction 3 c

0.42

6.88

8.30

9.44

Total % Recovery

63.84+17.12d

80.03

85.92

88.62

a. b. c.

d.

Fraction 1 denotes non-acid compounds eluted with 100% (v) DCM. Fraction 2 denotes non-acid compounds eluted with 50/50 (v/v) DCM: MeOH Fraction 3 denotes acid-containing compounds fraction eluted with 50/50/5 (v/v/v)DCM: MeOH: FA The % of volatile compounds in the unweathered MWO which was evaporated under dry nitrogen. The total percentage mass recovery of MWO reached 80.96% after taking this portion into consideration.

277 278

Acidic Speciation by Negative-ion ESI FT-ICR MS. Negative-ion electrospray

279

ionization selectively ionizes acidic species (e.g., carboxylic acids, pyrrolic nitrogen)

280

through deprotonation, and combined with FT-ICR mass spectrometry provides

281

elemental composition assignment to polar compounds in crude oil that are inaccessible

282

by conventional GC based techniques due to volatility limitations27,

283

separation shows that the acidity of the field samples increases, we focus on examining

284

the evolution of acidic species upon weathering by negative-ion ESI, which selectively

10

ACS Paragon Plus Environment

56, 58

Since APS

Page 11 of 28

Environmental Science & Technology

285

targets

acidic

species.

Speciation

by

positive-ion

ESI,

which

targets

286

functionalities, shows very little change, and is provided in Supporting Information.

basic

287 288

Molecular Weight Distribution. Figure S-2 shows broadband negative-ion ESI FT-ICR

289

mass spectra for parent MWO and select field samples. Broadband MWO and field

290

samples span a similar molecular weight range (~m/z 250-950), with an increased

291

abundance of low molecular weight compounds between m/z 400-450 after 9 mos

292

relative to MWO. After 36 and 48 mos, low molecular weight compounds

293

(300 1,000,000) for each peak

317

identified at m/z 501 for each time point.

318 319

Heteroatom Class Distribution for MWO. When combined with Kendrick mass

320

sorting, FT-ICR MS provides sub-ppm mass measurement accuracy sufficient for

321

elemental composition assignment34. Molecular formulae can be grouped into

322

heteroatom classes (e.g., nitrogen, sulfur, oxygen) based on relative abundance to

323

elucidate global compositional trends that occurred field samples compared to parent

324

MWO. Figure 3 shows the heteroatom class distribution for species >0.5% relative

325

abundance for the MWO and field samples (9 to 48 mos) based on triplicate analysis

326

and normalized to the most abundant peak in each mass spectrum. Compounds

327

containing a single nitrogen atom (N1) correspond to pyrrolic nitrogen in negative-ion

328

ESI, and comprise the most abundant class (~35%) in parent MWO, followed by

329

hydrocarbons (HC,~10%), O1 (~5%) and N1O1 (~5%) as previously reported.27 The N1

330

class decreases to ~4% after 9 mos, and ~1% after 36 mos. Similarly, the hydrocarbon

331

class (HC), which corresponds to five-member ring hydrocarbons (e.g., fluorene)

332

decrease in abundance over time (~5% at 9 mos to ~2% at 41 mos) and are not

333

detected after 43 mos.

334 335

(Insert Figure 3 here)

336 337

Increased Abundance of Oxygen Classes in Field Samples. The rapid decline in

338

relative abundance of nonpolar hydrocarbon (HC) and pyrrolic nitrogen (N1) classes

339

(Figure 3) is matched by a dramatic increase in species that contain oxygen.

340

Heteroatom classes that contain one or more oxygen atoms are highlighted in Figure S-

341

3. Field samples contain higher weight percent oxygen than MWO, which could suggest

342

numerous oxygen functionalities (i.e., ketones, aldehydes, carboxylic acids) that create

343

polyfunctional transformation products, as previously reported32. The most abundant

344

class in the 9 mos sample correspond to carboxylic acids (O2) followed by O1, O3 and

345

O4 classes. Compounds with one oxygen (O1) correspond to phenolic compounds, and

346

decrease in relative abundance after 9 mos, with no species detected above 0.5%

12

ACS Paragon Plus Environment

Page 13 of 28

Environmental Science & Technology

347

relative abundance after 43 mos. The increase in O2 species (corrected for mass loss

348

relative to C30 hopane), proceeds over 36 months, with an increase of ~3-fold after 9

349

months to a maximum of ~5.5-fold after 36 months, but decreases to ~3-fold after 48

350

months as compared to the parent MWO. Higher order oxygen classes (O2 – O4)

351

increase in abundance concurrent with a decreased abundance of low heteroatom-

352

containing compounds (e.g., hydrocarbons, N1 and O1) between 9 and 36 mos,59 and

353

corresponds to oxidation of O1 compounds (e.g., alcohols, aldehyde intermediates) to

354

form carboxylic acids through biodegradation.31-33 Between 41 and 48 mos, O2 - O3

355

classes decrease and O4 – O6 increase in abundance. This is in accordance with the

356

results revealed by APS separation that filed samples are becoming more acidic over

357

time (Table 1), indicating secondary oxidation of moderately acidic compounds to

358

create compounds that contain one or more carboxylic acids.

359 360

Hydrogen Deficiency (DBE) versus Carbon Number for Acidic Species in Field

361

Samples. Graphical images of hydrogen deficiency displayed as double bond

362

equivalents (DBE, the number of rings plus double bonds to carbon, calculated from

363

elemental composition CcHhNnOoSs) versus carbon number allow rapid visualization of

364

compositional changes between samples for all members of a heteroatom class

365

simultaneously.60 Figure 4 shows isoabundance-contoured plots of DBE versus carbon

366

number for O1 – O5 classes derived from negative-ESI FT-ICR MS of parent MWO and

367

field samples collected at 9 and 48 mos. Compounds that contain one oxygen atom (O1)

368

correspond to phenolic compounds and decrease in abundance after 9 mos (Figure S-

369

3). Within the O2 and O3 classes, a decrease of ~7 DBE occurs across a similar carbon

370

number range (C20 – C70), and indicates oxidation of aliphatic chains to form carboxylic

371

acids that results in the formation of low DBE compounds not observed in parent MWO.

372

Between 9 and 31 mos, the DBE distribution decreases by 1-2 DBE, and compounds

373

with fewer than 3 oxygen atoms (0.5% relative abundance) for moderately and heavily oiled sites collected at 9, 41 and

405

48 mos. The relative abundance of O2 species in moderately after 9 mos nearly doubles

406

compared to heavily oiled sites from the same time point, with the O2 class at 12%

407

(heavily oiled) to 26% (moderately oiled), concurrent with an increased abundance of

408

higher order oxygen species (O3-O6). After 41 mos, the relative abundance of O2 14

ACS Paragon Plus Environment

Page 15 of 28

Environmental Science & Technology

409

compounds in moderately oiled sediments was nearly 30% compared to 20% in heavy

410

contamination. Increased oil contamination decreases microbial activity; therefore,

411

biodegradation occurs more rapidly at moderately oiled sites,62 in agreement with FT-

412

ICR MS results. The materials extracted from the homogenized composites collected

413

from non-oiled reference sites shown in Figure 1 (7 sites) resemble typical natural

414

organic matter (Figure S-5) and provide a background to assess oil contamination.63, 64

415 416

This study represents the first temporal characterization of MWO transformation

417

products at the molecular level. Here, we identify persistent MWO transformation

418

products and catalogue compositional changes that occur to parent MWO in Barataria

419

Bay saltmarsh sediments up to 4 years after the Deepwater Horizon oil spill. Highly

420

polar, acidic oxygenated transformation products derived from biodegradation of

421

Macondo Well oil in coastal salt marshes span a wide range of chemical functionalities,

422

and remain environmentally persistent. The increased abundance of highly polar

423

compounds of low volatility is ideally suited for FT-ICR MS can address the complexity

424

of these highly polar, multifunctional oxidized transformation products at the molecular

425

level. Compositional images of DBE versus carbon number indicate that degradation

426

processes convert nonpolar, saturated compounds into highly polar, carboxylic acid

427

compounds that appear over time, and subsequent oxidation converts lower order

428

oxygen compounds (O1-O3) to higher order oxygen compounds (O4-O6) after 36 mos.

429

Bulk fractionation on aminopropyl silica attributes the increased oxygen content

430

obtained from elemental analysis of field samples to an increase of moderately acidic

431

and carboxylic acid moieties over time. Van Krevelen diagrams indicate that initial

432

oxidation during the first 9 mos was non-selective across a wide range of O:C ratios,

433

and shifts aromatic hydrocarbon degradation after 41 mos. Oil weathering pattern

434

revealed by GC×GC-FID analysis and FT-ICR MS indicate both photo-oxidation and

435

biodegradation contribute to the molecular modification of weathered oil in surface

436

layers of saltmarsh sediments. Moderately oiled sites undergo rapid microbial

437

degradation compared to the heavily oiled sites, and indicate that heavy oil

438

contamination hinders microbial activity.

15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 28

439

■Future Work

440

Sediment samples in this study were split in half and preserved for microbial analysis at

441

Louisiana State University. The chemical data presented here will be correlated with the

442

soil microbial community structure and function to provide molecular level insight on

443

how the oil contamination perturbed the indigenous microbial community, and how the

444

microbes responded to the perturbation. In addition, future studies will apply

445

methodologies to Deepwater Horizon oil residues deposited in coastal marshes, buried

446

in tidal beach sediments, and washed ashore as aggregated “tar balls” to further probe

447

the chemical diversity of oxygen compounds formed through MWO degradation.

448

Microcosm studies will also be compared to “real world” samples to disentangle biotic

449

and abiotic modification and the long-term fate of MWO.

450

■ASSOCIATED CONTENT

451

Supporting Information.

452

Additional methods, figures and tables, as noted in the text, are provided. This material

453

is available free of charge via the Internet at http://pubs.acs.org.

454

■ACKNOWLEDGEMENTS

455 456 457 458 459 460 461 462

Work supported by NSF Division of Materials Research through DMR-11-57490, The Gulf of Mexico Research Initiative to the Deep-C Consortium and to Louisiana State University through Ocean Leadership subaward SA 13-30/GoMRI-013, and the State of Florida. The authors thank Nathan K. Kaiser, John P. Quinn, and Greg T. Blakney for their continued assistance with instrumental maintenance and data analysis, Jacqueline M. Jarvis and Rebecca L. Beasley for technical assistance, as well as Alan G. Marshall, Steven M. Rowland and David C. Podgorski for helpful discussions.

16

ACS Paragon Plus Environment

Page 17 of 28

Environmental Science & Technology

463

■References

464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507

1. McNutt, M. K.; Camilli, R.; Crone, T. J.; Guthrie, G. D.; Hsieh, P. A.; Ryerson, T. B.; Savas, O.; Shaffer, F., Review of flow rate estimates of the Deepwater Horizon oil spill. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (50), 20260-20267. 2.Camilli, R.; Di Iorio, D.; Bowen, A.; Reddy, C. M.; Techet, A. H.; Yoerger, D. R.; Whitcomb, L. L.; Seewald, J. S.; Sylva, S. P.; Fenwick, J., Acoustic measurement of the Deepwater Horizon Macondo well flow rate. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (50), 20235-20239. 3. Jackson, J. B.; Cubit, J. D.; Keller, B. D.; Batista, V.; Burns, K.; Caffey, H. M.; Caldwell, R. L.; Garrity, S. D.; Getter, C. D.; Gonzalez, C., Ecological effects of a major oil spill on Panamanian coastal marine communities. Science 1989, 243, (4887), 37-44. 4. Incardona, J. P.; Swarts, T. L.; Edmunds, R. C.; Linbo, T. L.; Aquilina-Beck, A.; Sloan, C. A.; Gardner, L. D.; Block, B. A.; Scholz, N. L., Exxon Valdez to Deepwater Horizon: comparable toxicity of both crude oils to fish early life stages. Aquat. Toxicol. 2013, 142-143, 303-316. 5. Garshelis, D. L.; Johnson, C. B., Prolonged recovery of sea otters from the Exxon Valdez oil spill? A re-examination of the evidence. Mar. Pollut. Bull. 2013, 71, (1), 7-19. 6. Silliman, B. R.; van de Koppel, J.; McCoy, M. W.; Diller, J.; Kasozi, G. N.; Earl, K.; Adams, P. N.; Zimmerman, A. R., Degradation and resilience in Louisiana salt marshes after the BP– Deepwater Horizon oil spill. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (28), 11234-11239. 7. Jewett, S.; Dean, T.; Smith, R.; Blanchard, A., Exxon Valdez oil spill: impacts and recovery in the soft-bottom benthic community in and adjacent to eelgrass beds. Marine ecology. Progress series 1999, 185, 59-83. 8. Tao, Z.; Bullard, S.; Arias, C., High numbers of Vibrio vulnificus in tar balls collected from oiled areas of the north-central Gulf of Mexico following the 2010 BP Deepwater Horizon oil spill. Ecohealth 2011, 8, (4), 507-511. 9. Mitsch, W. J.; Gosselink, J. G., The value of wetlands: importance of scale and landscape setting. Ecol. Econ. 2000, 35, (1), 25-33. 10. Engle, V. D., Estimating the provision of ecosystem services by Gulf of Mexico coastal wetlands. Wetlands 2011, 31, (1), 179-193. 11. Venosa, A. D.; Lee, K.; Suidan, M. T.; Garcia-Blanco, S.; Cobanli, S.; Moteleb, M.; Haines, J.R.; Tremblay, G.; Hazelwood, M., Bioremediation and Biorestoration of a Crube OilContaminated Freshwater Wetland on the St.Lawrence River. Bioremediat. J. 2002, 6, (3), 261281. 12. Mahmoudi, N.; Porter, T. M.; Zimmerman, A. R.; Fulthorpe, R. R.; Kasozi, G. N.; Silliman, B. R.; Slater, G. F., Rapid degradation of Deepwater Horizon spilled oil by indigenous microbial communities in Louisiana saltmarsh sediments. Environ. Sci. Technol. 2013, 47, (23), 1330313312. 13. Beazley, M. J.; Martinez, R. J.; Rajan, S.; Powell, J.; Piceno, Y. M.; Tom, L. M.; Andersen, G. L.; Hazen, T. C.; Van Nostrand, J. D.; Zhou, J., Microbial community analysis of a coastal salt marsh affected by the Deepwater Horizon oil spill. PLoS One 2012, 7, (7), e41305. 14. King, G.; Kostka, J.; Hazen, T.; Sobecky, P., Microbial Responses to the deepwater horizon oil spill: from coastal wetlands to the deep sea. Ann. Rev. Mar. Sci. 2015, 7, 377-401. 15. Reddy, C. M.; Arey, J. S.; Seewald, J. S.; Sylva, S. P.; Lemkau, K. L.; Nelson, R. K.; Carmichael, C. A.; McIntyre, C. P.; Fenwick, J.; Ventura, G. T.; Van Mooy, B. A.; Camilli, R., Composition and fate of gas and oil released to the water column during the Deepwater Horizon oil spill. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (50), 20229-20234. 17

ACS Paragon Plus Environment

Environmental Science & Technology

508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553

Page 18 of 28

16. Valentine, D. L.; Fisher, G. B.; Bagby, S. C.; Nelson, R. K.; Reddy, C. M.; Sylva, S. P.; Woo, M. A., Fallout plume of submerged oil from Deepwater Horizon. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, (45), 15906-15911. 17. Ryerson, T. B.; Camilli, R.; Kessler, J. D.; Kujawinski, E. B.; Reddy, C. M.; Valentine, D. L.; Atlas, E.; Blake, D. R.; de Gouw, J.; Meinardi, S.; Parrish, D. D.; Peischle, J.; Seewald, J. S.; Warneke, C., Chemical Data Quantify Deepwater Horizon hydrocarbon flow rate and environmental distribution. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (50), 20246-20253. 18. White, H. K.; Hsing, P. Y.; Cho, W.; Shank, T. M.; Cordes, E. E.; Quattrini, A. M.; Nelson, R. K.; Camilli, R.; Demopoulos, A. W.; German, C. R.; Brooks, J. M.; Roberts, H. H.; Shedd, W.; Reddy, C. M.; Fisher, C. R., Impact of the Deepwater Horizon oil spill on a deep-water coral community in the Gulf of Mexico. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (50), 20303-20308. 19. Aeppli, C.; Carmichael, C. A.; Nelson, R. K.; Lemkau, K. L.; Graham, W. M.; Redmond, M. C.; Valentine, D. L.; Reddy, C. M., Oil weathering after the Deepwater Horizon disaster led to the formation of oxygenated residues. Environ. Sci. Technol. 2012, 46, (16), 8799-8807. 20. Hall, G. J.; Frysinger, G. S.; Aeppli, C.; Carmichael, C. A.; Gros, J.; Lemkau, K. L.; Nelson, R. K.; Reddy, C. M., Oxygenated weathering products of Deepwater Horizon oil come from surprising precursors. Mar. Pollut. Bull. 2013, 75, (1-2), 140-149. 21. Turner, R. E.; Overton, E. B.; Meyer, B. M.; Miles, M. S.; McClenachan, G.; Hooper-Bui, L.; Engel, A. S.; Swenson, E. M.; Lee, J. M.; Milan, C. S.; Gao, H., Distribution and recovery trajectory of Macondo (Mississippi Canyon 252) oil in Louisiana coastal wetlands. Mar. Pollut. Bull. 2014, 87, (1–2), 57-67. 22. Carmichael, C. A.; Arey, J. S.; Graham, W. M.; Linn, L. J.; Lemkau, K. L.; Nelson, R. K.; Reddy, C. M., Floating oil-covered debris from Deepwater Horizon: identification and application. Environ. Res. Lett 2012, 7, 015301. 23. Lewan, M.; Warden, A.; Dias, R.; Lowry, Z.; Hannah, T.; Lillis, P.; Kokaly, R.; Hoefen, T.; Swayze, G.; Mills, C., Asphaltene content and composition as a measure of Deepwater Horizon oil spill losses within the first 80days. Org. Geochem. 2014, 75, 54-60. 24. Gros, J.; Reddy, C. M.; Aeppli, C.; Nelson, R. K.; Carmichael, C. A.; Arey, J. S., Resolving biodegradation patterns of persistent saturated hydrocarbons in weathered oil samples from the Deepwater Horizon disaster. Environ. Sci. Technol. 2014, 48, (3), 1628-1637. 25. Looper, J. K.; Cotto, A.; Kim, B. Y.; Lee, M. K.; Liles, M. R.; Ni Chadhain, S. M.; Son, A., Microbial community analysis of Deepwater Horizon oil-spill impacted sites along the Gulf coast using functional and phylogenetic markers. Env. Sci. Process. Impact. 2013, 15, (11), 2068-2079. 26. Atlas, R. M.; Stoeckel, D. M.; Faith, S. A.; Minard-Smith, A.; Thorn, J. R.; Benotti, M. J., Oil Biodegradation and Oil-degrading Microbial Populations in Marsh Sediments Impacted by Oil from the Deepwater Horizon Well Blowout. Environ. Sci. Technol. 2015, 49, (14), 83568366. 27. McKenna, A. M.; Nelson, R. K.; Reddy, C. M.; Savory, J. J.; Kaiser, N. K.; Fitzsimmons, J. E.; Marshall, A. G.; Rodgers, R. P., Expansion of the analytical window for oil spill characterization by ultrahigh resolution mass spectrometry: Beyond gas chromatography. Environ. Sci. Technol. 2013, 47, (13), 7530-7539. 28. Incardona, J. P.; Vines, C. A.; Anulacion, B. F.; Baldwin, D. H.; Day, H. L.; French, B. L.; Labenia, J. S.; Linbo, T. L.; Myers, M. S.; Olson, O. P., Unexpectedly high mortality in Pacific herring embryos exposed to the 2007 Cosco Busan oil spill in San Francisco Bay. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, (2), E51-E58.

18

ACS Paragon Plus Environment

Page 19 of 28

554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598

Environmental Science & Technology

29. McKenna, A. M.; Williams, J. T.; Putman, J. C.; Aeppli, C.; Reddy, C. M.; Valentine, D. L.; Lemkau, K. L.; Kellerman, M. Y.; Savory, J. J.; Kaiser, N. K.; Marshall, A. G.; Rodgers, R. P., Unprecedented ultrahigh resolution FT-ICR mass spectrometry and parts-per-billion mass accuracy enable direct characterization of nickel and vanadyl porphyrins in petroleum from natural seeps. Energy Fuels 2014, 28, 2454-2464. 30. Frysinger, G. S.; Gaines, R. B.; Xu, L.; Reddy, C. M., Resolving the unresolved complex mixture in petroleum-contaminated sediments. Environ. Sci. Techn. 2003, 37, (8), 1653-1662. 31. Lemkau, K. L.; McKenna, A. M.; Podgorski, D. C.; Rodgers, R. P.; Reddy, C. M., Molecular evidence of heavy-oil weathering following the M/V Cosco Busan spill: Insights from Fourier transform ion cyclotron resonance mass spectrometry. Environ. Sci. Technol. 2014, 48, (7), 3760-3767. 32. Ruddy, B. M.; Huettel, M.; Kostka, J. E.; Lobodin, V. V.; Bythell, B. J.; McKenna, A. M.; Aeppli, C.; Reddy, C. M.; Nelson, R. K.; Marshall, A. G., Targeted petroleomics: analytical investigation of Macondo well oil oxidation products from Pensacola Beach. Energy Fuels 2014, 28, (6), 4043-4050. 33. Corilo, Y. E.; Podgorski, D. C.; McKenna, A. M.; Lemkau, K. L.; Reddy, C. M.; Marshall, A. G.; Rodgers, R. P., Oil spill source identification by principal component analysis of electrospray ionization Fourier transform ion cyclotron resonance mass spectra. Anal. Chem. 2013, 85, (19), 9064-9069. 34. Rodgers, R. P.; Schaub, T. M.; Marshall, A. G., Petroleomics: MS Returns to Its Roots. Anal. Chem. 2005, 77, (1), 20 A-27 A. 35. Rodgers, R. P.; Blumer, E. N.; Freitas, M. A.; Marshall, A. G., Jet Fuel Chemical Composition, Weathering, and Identification as a Contaminant at a Remediation Site, Determined by Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Anal. Chem. 1999, 71, 5171-5176. 36. Hughey, C. A.; Minardi, C. S.; Galasso-Roth, S. A.; B., P. G.; Mapolelo, M. M.; Rodgers, R. P.; Marshall, A. G.; Ruderman, D. L., Napthenic acids as indicators of crude oil biodegradation in soil based on semi-quantitative electrospray ionization Fourier transform ion cyclotron resonance mass spectrometry. Rapid Commun. Mass Spectrom. 2008, 22, 3968-3976. 37. Rodgers, R. P.; Blumer, E. N.; Freitas, M. A.; Marshall, A. G., Complete compositional monitoring of the weathering of transportation fuels based on elemental compositions from Fourier transform ion cyclotron resonance mass spectrometry. Environ. Sci. Technol. 2000, 34, 1671-1678. 38. Lemkau, K. L.; Peacock, E. E.; Nelson, R. K.; Ventura, G. T.; Kovecses, J. L.; Reddy, C. M., The M/V Cosco Busan spill: Source identification. Mar. Pollut. Bull. 2010, 60, (11), 2123-2129. 39. Lin, Q.; Mendelssohn, I. A., Impacts and recovery of the Deepwater Horizon oil spill on vegetation structure and function of coastal salt marshes in the northern Gulf of Mexico. Environ. Sci. Technol. 2012, 46, (7), 3737-43. 40. Michel, J.; Owens, E. H.; Zengel, S.; Graham, A.; Nixon, Z.; Allard, T.; Holton, W.; Reimer, P. D.; Lamarche, A.; White, M.; Rutherford, N.; Childs, C.; Mauseth, G.; Challenger, G.; Taylor, E., Extent and degree of shoreline oiling: Deepwater Horizon oil spill, Gulf of Mexico, USA. PLoS One 2013, 8, (6), e65087. 41. ASTM D5369-93(2008)e1, Standard Practice for Extraction of Solid Waste Samples for Chemical Analysis Using Soxhlet Extraction. In ASTM International: West Conshohocken, PA, 2008.

19

ACS Paragon Plus Environment

Environmental Science & Technology

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

Page 20 of 28

42. Aeppli, C.; Nelson, R. K.; Radović, J. R.; Carmichael, C. A.; Valentine, D. L.; Reddy, C. M., Recalcitrance and degradation of petroleum biomarkers upon abiotic and biotic natural weathering of Deepwater Horizon oil. Environ. Sci. Technol. 2014, 48, (12), 6726-6734. 43. Prince, R. C.; Elmendorf, D. L.; Lute, J. R.; Hsu, C. S.; Haith, C. E.; Senius, J. D.; Dechert, G. J.; Douglas, G. S.; Butler, E. L., 17α(H)-21β(H)-hopane as a conserved internal marker for estimating the biodegradation of crude oil. Environ. Sci. Technol. 1994, 28, (1), 142-145. 44. Venosa, A.; Suidan, M.; King, D.; Wrenn, B., Use of hopane as a conservative biomarker for monitoring the bioremediation effectiveness of crude oil contaminating a sandy beach. J. Ind. Microbiol. Biotechnol. 1997, 18, (2-3), 131-139. 45. Fingas, M., Handbook of Oil Spill Science and Technology. John Wiley & Sons: 2014. 46. Prince, R.C.; Garrett, R.M.; Bare, R.E.; Grossman, M.J.; Townsend, T.; Suflita, J.M.; Lee, K.; Owens, E.H.; Sergy, G.A.; Braddock, J.F., The roles of photooxidation and biodegradation in long-term weathering of crude and heavy fuel oils. Spill Science & Technology Bulletin 2003, 8, (2), 145-156. 47. Frank, R.A.; Kavanagh, R.; Kent Burnison, B.; Arsenault, G.; Headley, J.V.; Peru, K.M.; Van Der Kraak, G.; Solomon, K.R., Toxicity assessment of collected fractions from an extracted naphthenic acid mixture. Chemosphere 2008, 72, (9), 1309-1314. 48. Leung, S.S.C.; MacKinnon, M.D.; Smith, R. E., Aquatic reclamation in the Athabasca, Canada, oil sands: naphthenate and salt effects on phytoplankton communities. Environ. Toxicol. Chem. 2001, 20, (7), 1532-1543. 49. Frank, R. A.; Fischer, K.; Kavanagh, R.; Burnison, B. K.; Arsenault, G.; Headley, J. V.; Peru, K. M.; Kraak, G. V. D.; Solomon, K. R., Effect of Carboxylic Acid Content on the Acute Toxicity of Oil Sands Naphthenic Acids. Environ. Sci. Technol. 2008, 43, (2), 266-271. 50. Kamaluddin, M.; Zwiazek, J. J., Naphthenic acids inhibit root water transport, gas exchange and leaf growth in aspen (Populus tremuloides) seedlings. Tree Physiol. 2002, 22, (17), 12651270. 51. Young, R.; Orr, E.; Goss, G.; Fedorak, P., Detection of naphthenic acids in fish exposed to commercial naphthenic acids and oil sands process-affected water. Chemosphere 2007, 68, (3), 518-527. 52. Peters, L.E.; MacKinnon, M.; Van Meer, T.; van den Heuvel, M. R.; Dixon, D., Effects of oil sands process-affected waters and naphthenic acids on yellow perch Perca flavescens and Japanese medaka embryonic development. Chemosphere 2007, 67, (11), 2177-2183. 53. Rogers, V.V.; Wickstrom, M.; Liber, K.; MacKinnon, M.D., Acute and subchronic mammalian toxicity of naphthenic acids from oil sands tailings. Toxicol. Sci. 2002, 66, (2), 347355. 54. Dokholyan, B.; Magomedov, A., The effect of sodium naphthenate on the viability and physiological and biochemical indices of fish. Voprosy Ikhtiologii 1984, 23, (6), 1013-1019. 55. Stanford, L.A.; Kim, S.; Klein, G.C.; Smith, D.F.; Rodgers, R.P.; Marshall, A.G., Identification of Water-Soluble Heavy Crude Oil Organics. Acidic and Basic NSO Compounds in Fresh Water and Sea Water by Electrospray Ionization Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Environ. Sci. Technol. 2007, 41, 2696-2702. 56. Rowland, S. M.; Robbins, W. K.; Corilo, Y. E.; Marshall, A. G.; Rodgers, R. P., Solid-Phase Extraction Fractionation To Extend the Characterization of Naphthenic Acids in Crude Oil by Electrospray Ionization Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Energy Fuels 2014, 28, (8), 5043-5048.

20

ACS Paragon Plus Environment

Page 21 of 28

644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689

Environmental Science & Technology

57. Atlas, R.M., Hazen, T.C., Oil biodegradation and bioremediation: a tale of the two worst spills in U.S. history. Environ. Sci. Technol. 2011, 45, (16), 6709-6715. 58. Hughey, C. A.; Rodgers, R.P.; Marshall, A.G.; Qian, K.; Robbins, W.K., Identification of acidic NSO compounds in crude oils of different geochemical origins by negative ion electrospray Fourier transform ion cyclotron resonance mass spectrometry. Org. Geochem. 2002, 33, (7), 743-759. 59. Mapolelo, M.M.; Rodgers, R.P.; Blakney, G. T.; Yen, A. T.; Asomaning, S.; Marshall, A. G., Characterization of naphthenic acids in crude oils and naphthenates by electrospray ionization FT-ICR mass spectrometry. Int. J. Mass Spectrom. 2011, 300, (2-3), 149-157. 60. McLafferty, F.W.; Turecek, F., Interpretation of Mass Spectra. 4th ed. ed.; University Science Books: Mill Valley, CA, 1993; p 371. 61. Kim, S.; Kramer, R.W.; Hatcher, P.G., Graphical method for analysis of ultrahigh-resolution broadband mass spectra of natural organic matter, the van Krevelen diagram. Anal. Chem. 2003, 75, (20), 5336-5344. 62. Leahy, J.G.; Colwell, R.R., Microbial degradation of hydrocarbons in the environment. Microbiological reviews 1990, 54, (3), 305-315. 63. Kujawinski, E.B.; Freitas, M.A.; Zang, X.; Hatcher, P.G.; Green-Church, K.B.; Jones, R.B., The application of electrospray ionization mass spectrometry (ESI MS) to the structural characterization of natural organic matter. Org. Geochem. 2002, 33, (3), 171-180. 64. Schmidt, F.; Elvert, M.; Koch, B. P.; Witt, M.; Hinrichs, K.-U., Molecular characterization of dissolved organic matter in pore water of continental shelf sediments. Geochim. Cosmochim. Acta 2009, 73, (11), 3337-3358. FIGURE LEGENDS Figure 1. Map of northern Barataria Basin of Louisiana, USA, approximately 8 km x 5 km (coordinates N 29.44060° - 29.47459°, W 89.88492° - 89.94647°), with 21 sampling locations (seven each that received heavy oiling (Sites 1-7), moderate oiling (sites 814), and no observed oiling or reference (sites 15-21), based on the Shoreline Cleanup Assessment Technique (SCAT) Program (28 April 2010), field observations, and total petroleum hydrocarbon measurements. No cleanup events at the sampling sites were noted during this period. Figure 2. Calculated losses of total petroleum hydrocarbon and chrysenes relative to MWO based on the conservation of hopane for field samples extracted from 9, 24, 36 and 48 months after the DWH spill (total, eq1; Chrysene eq2;). The C1, C2, C3 indicates the number of alky carbons on the parent molecule. Figure 3. Heteroatom class distribution of major species (>1% relative abundance) derived from negative ESI FT-ICR mass spectra of parent MWO Macondo well oil and oil contaminants over 48 mos post-spill with a TMAH-modified solvent system. Bars indicate standard errors (N = 3) Figure 4. Negative ESI-derived isoabundance-contoured plots of double bond equivalents (DBE), versus carbon number for the O1 – O5 classes for parent MWO and

21

ACS Paragon Plus Environment

Environmental Science & Technology

690 691 692 693 694 695 696 697 698

Page 22 of 28

field samples collected from 9 to 48 mos post-spill. Each compositional image is normalized to the most abundant species within that heteroatom class for each mass spectrum. Relative abundance weighted average carbon numbers and DBE values for each sample are listed in Table 1. Figure 5. Van Krevelen diagrams for the O1 –O8 classes of the acidic polar species in the parent MWO and field samples collected at heavily contaminate sites analyzed by negative ion ESI FT-ICR MS.

22

ACS Paragon Plus Environment

Page 23 of 28

Environmental Science & Technology

Bay Chene Fleur

Figure 1

Wilkinson Bay 21

17

18

20 19 16 15

10

11

14 12 13

7

5

8 1

Heavily Oiled

9 2

Moderately Oiled Reference

4

6

Bay Jimmy

3

Barataria ACS ParagonBay Plus Environment 1Km

Environmental Science & Technology

% Loss Relative to MWO

9 mos 90

24 mos

80

36 mos

Page 24 of 2 28 Figure

48 mos

70 60 50 40 30 20 10 0

ACS Paragon Plus Environment

TPH

Chrysene

C1-Chrysenes

C2-Chrysenes C3-Chrysenes

Page 25 of 28

Figure 3

Environmental Science & Technology

a)

40

MWO

18 mos

30

24 mos 31 mos 36 mos

20

41 mos 43 mos

O4S1

O2S1

O6

O5

O4

O3

O2

O1

N1O4

N1O3

N1O2

N1O1

N1

0

O3S1

48 mos

10

HC

% Relative Abundance

9 mos

MWO

b)

9 mos

25 % Relative Abundance

18 mos 24 mos

20

31 mos 36 mos

15

41 mos 43 mos

10

48 mos

5 0

O1

O2

O3

O4

ACS Paragon Plus Environment

Heteroatom Class

O5

O6

Figure 4 30

MWO

9 mos

Environmental Science & Technology

18 mos

24 mos

31 mos

36 mos

41 mos

43 mos

No Species detected ≥ 0.5%

No Species detected ≥ 0.5%

20 10

Page 26 of 28

48 mos

O1

30 20

O2

10

DBE

30 20

O3

10 30 20 10

No Species detected ≥0.5%

O4

No Species detected ≥ 0.5%

O5

30 20 10 0

20 40 60 Relative Abundance (% total)

20 80 20

40

60

80 20

40

60

80

20

40

60 80 20 40 60 80 20 ACS Paragon Plus Environment

Carbon Number

40

60

80 20

40

60

80 20

40

60

80 20

40

60

80

Figure 5 28 Page 27 of

Environmental Science & Technology

2.0

(a). MWO

(e). 36 mos

(b). 9 mos

(f). 41 mos

(c). 24 mos

(g). 43 mos

(d). 31 mos

(h). 48 mos

1.5 1.0 0.5 0.0 2.0 1.5 1.0 0.5

H/C

0.0 2.0 1.5 1.0 0.5 0.0 2.0 1.5 1.0 0.5 0.0 0.0

0.1

0.2

0.3

0.4

0.0

O/C Relative Abundance (% total) ACS Paragon Plus Environment

0.1

0.2

0.3

0.4

25

MWO

Environmental Science & Technology Page 289 mos of 28

% Relative Abundance

18 mos

20

24 mos 31 mos 36 mos

15

41 mos 43 mos 48 mos

10 5

ACS Paragon Plus Environment 0

O1

O2

O3

O4

Heteroatom Class

O5

O6