Geochemistry of Dissolved Organic Matter in a ... - ACS Publications

May 6, 2016 - Tables S3 and S4 for 1H NMR section integrals according ..... Degradation of terrestrially derived macromolecules in the Amazon. River. ...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF PITTSBURGH

Article

Geochemistry of Dissolved Organic Matter in a Spatially Highly Resolved Groundwater Petroleum Hydrocarbon Plume Cross-Section Sabine Eva-Maria Dvorski, Michael Gonsior, Norbert Hertkorn, Jenny Uhl, Hubert Müller, Christian Griebler, and Philippe Schmitt-Kopplin Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00849 • Publication Date (Web): 06 May 2016 Downloaded from http://pubs.acs.org on May 10, 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

Environmental Science & Technology

1

Geochemistry of Dissolved Organic Matter in a Spatially Highly

2

Resolved Groundwater Petroleum Hydrocarbon Plume Cross-

3

Section

4 5

Sabine E.-M. Dvorski a, Michael Gonsior b, Norbert Hertkorn a, Jenny Uhl a, Hubert Müller c, Christian Griebler c, Philippe Schmitt-Kopplin* a, d

6

a

7

Analytical BioGeoChemistry, D-85764 Neuherberg, Germany

8 9

b

Helmholtz Zentrum München - German Research Center for Environmental Health, Research Unit

University of Maryland Center for Environmental Science, Chesapeake Biological Laboratory, Solomons, Maryland 20688, USA

10

c

11

Groundwater Ecology, D-85764 Neuherberg, Germany

12 13

d

Helmholtz Zentrum München - German Research Center for Environmental Health, Institute of

Technische Universität München, Chair of Analytical Food Chemistry, D-85354 FreisingWeihenstephan, Germany

14 15

* Corresponding author: Philippe Schmitt-Kopplin, Helmholtz Zentrum München, - German Research

16

Center for Environmental Health, Research Unit Analytical BioGeoChemistry, Ingolstädter

17

Landstraße 1, 85764 Neuherberg, Germany, phone: (+49) 089 3187 3246, fax: (+49) 089 3187 3358,

18

Email: [email protected]

19

20

Page 1 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 28

21

Abstract

22

At numerous groundwater sites worldwide, natural dissolved organic matter (DOM) is quantitatively

23

complemented with petroleum hydrocarbons. To date, research has been focused almost exclusively

24

on the contaminants, but detailed insights of the interaction of contaminant biodegradation, dominant

25

redox processes, and interactions with natural DOM are missing. This study linked on-site high

26

resolution spatial sampling of groundwater with high resolution molecular characterization of DOM

27

and its relation to groundwater geochemistry across a petroleum hydrocarbon plume cross-section.

28

Electrospray- and atmospheric pressure photoionization (ESI, APPI) ultra-high resolution mass

29

spectrometry (FT-ICR-MS) revealed a strong interaction between DOM and reactive sulfur species

30

linked to microbial sulfate reduction, i.e. the key redox process involved in contaminant

31

biodegradation. Excitation emission matrix (EEM) fluorescence spectroscopy in combination with

32

Parallel Factor Analysis (PARAFAC) modeling attributed DOM samples to specific contamination

33

traits. Nuclear magnetic resonance (NMR) spectroscopy evaluated the aromatic compounds and their

34

degradation products in samples influenced by the petroleum contamination and its biodegradation.

35

Our orthogonal high resolution analytical approach enabled a comprehensive molecular level

36

understanding of the DOM with respect to in situ petroleum hydrocarbon biodegradation and

37

microbial sulfate reduction. The role of natural DOM as potential co-substrate and detoxification

38

reactant may improve future bioremediation strategies.

39

Page 2 of 28

ACS Paragon Plus Environment

Page 3 of 28

Environmental Science & Technology

40

Introduction

41

Natural dissolved organic matter (DOM) is an essential energy source for heterotrophic organisms in

42

aquatic ecosystems.1 Its concentration and composition are pivotal parameters related to water

43

quality.2 Groundwater ecosystems are typically poor in organic carbon and energy and, consequently,

44

systems of low productivity.3 DOM concentrations in pristine oxic groundwater are as little as 0.1

45

mg/L to a few mg/L.4 However, numerous groundwater sites worldwide are contaminated with

46

organic chemicals and exhibit significantly higher DOM concentrations.5,

47

groundwater contaminants are petroleum hydrocarbons with benzene, toluene, ethylbenzene and

48

xylenes (BTEX) and polycyclic aromatic hydrocarbons (PAHs) being the most common.7,

49

contaminated systems, natural DOM may act as a reactant, chelator, and sorbent for hydrophobic

50

organic contaminants influencing their bioavailabilty, mobility, and fate.1, 9 Thus far, detailed insight

51

into the role of natural DOM for petroleum hydrocarbon contaminated groundwater ecosystems is

52

missing on a molecular level.

53

From the point source, pollutants partly dissolve in groundwater and are transported along with the

54

water flow, thereby forming a contaminant plume.5 The fate and transformation of organic

55

contaminants in a mature plume are related to several abiotic processes such as dilution, diffusion, and

56

microbial degradation.10 Current research underlined that contaminant plumes in porous aquifers are

57

characterized by steep and small-scale gradients in the range of centimeters to decimeters with high

58

biodegradation activities occurring at the fringes of the plume.11, 12 This “plume fringe concept” is

59

founded on the depletion of dissolved electron acceptors in the plume core.5, 13 Biodegradation with

60

oxygen, nitrate, or sulfate reduction can then only take place at the fringes of the plume where

61

electron acceptors are replenished from surrounding groundwater by dispersion and diffusion. In the

62

plume core, biodegradation may alternatively be driven by reduction of solid electron acceptors such

63

as Fe(III)- or Mn(IV), and also by methanogenesis.14

6

Prominent organic

8

In

Page 3 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 28

64

Natural DOM interacts with not only organic pollutants but also inorganic reactive species that may

65

directly be related to the dominant terminal redox processes coupled to contaminant biodegradation.

66

Reactive sulfur species such as sulfide are a prominent example. In fact, sulfate reduction is the

67

quantitatively most important redox process in petroleum hydrocarbon contaminated aquifers.10 An

68

improved fundamental understanding of contaminant biodegradation bottlenecks and development of

69

successful remediation strategies need detailed investigations of the mutual influence, the dominating

70

redox reactions and products, and characteristics and distribution of contamination as well as its

71

interaction with natural DOM. As natural DOM is a complex organic mixture, often consisting of

72

thousands of organic compounds, high resolution non-targeted analytical characterization by means of

73

Fourier transform ion cyclotron resonance mass spectrometry (FT-ICR-MS), excitation emission

74

matrix (EEM) fluorescence spectroscopy, and nuclear magnetic resonance (NMR) spectroscopy are

75

techniques of choice for an adequate DOM assessment.1, 15-19

76

The diverse molecular characteristics of complex DOM can be unraveled using FT-ICR-MS as shown

77

in previous studies.1, 20 Here, the application of negative mode electrospray ionization ((-)ESI) FT-

78

ICR-MS led to a significantly improved mechanistic understanding of natural biotic and abiotic as

79

well as anthropogenic processes that characterize DOM in aquatic ecosystems.21-26 Recently, the FT-

80

ICR-MS analysis of DOM alteration in wastewater treatment and drinking water plants improved the

81

assessment of water quality.27-29 While (-)ESI FT-ICR-MS is selective towards DOM molecules

82

containing carboxylic acids and other polar functional groups, positive mode atmospheric pressure

83

ionization (+)APPI FT-ICR-MS favors the detection of aromatic and less polar compounds.30-33

84

(+)APPI FT-ICR-MS has also been used to characterize nonpolar complex mixtures such as shale oils,

85

asphaltenes, and PAHs and is targeted towards molecules with π-electron systems.34-37

86

Complementary to FT-ICR-MS, EEM fluorescence spectroscopy in conjunction with Parallel Factor

87

Analysis (PARAFAC) is a powerful tool used to study fluorescent dissolved organic matter

88

(FDOM).38, 39 PARAFAC modeling has been utilized to study water treatment and recycling schemes Page 4 of 28

ACS Paragon Plus Environment

Page 5 of 28

Environmental Science & Technology

89

as well as processes in natural aquatic systems.15,

40-44

90

molecules in mineral oils and their degradation products, EEM fluorescence spectroscopy is useful to

91

assess distribution, influence, and transformation of petroleum products in the environment.45-48

92

Additionally, 1D and 2D NMR spectroscopy offers quantitative in-depth information of proton and

93

carbon chemical environments from which extended substructures can be assembled.1,

94

combined use of NMR and FT-ICR-MS produced an informative assessment of DOM structural

95

details according to ecosystem characteristics.1,

96

employed for the analysis of oils from various sources and to study the interaction of natural organic

97

matter and organic pollutants.50-53

98

The aim of this study was to combine on-site high resolution spatial sampling across a well-defined

99

petroleum hydrocarbon plume with high resolution molecular characterization of DOM in an effort to

100

study the interaction of the contaminants in addition to redox processes and their reaction products

101

with natural DOM. We performed comprehensive analyses of solid phase extracted DOM (SPE-

102

DOM) from groundwater of the petroleum hydrocarbon contaminated sandy aquifer in Düsseldorf

103

Flingern by means of high-field (-)ESI and (+)APPI FT-ICR-MS, NMR, and EEM fluorescence

104

spectroscopy.

105

Methods

106

Site description and sampling

107

Groundwater was sampled in September 2013 at the former gasworks site in Düsseldorf Flingern,

108

Germany (51°13’19.83” N; 6°49’05.44” E). The aquifer is predominantly composed of quaternary

109

sand and gravel. The disposal of tar oil in the shallow subsurface for many decades (i.e. during

110

operation from 1893 – 1967) caused a contaminant plume mainly consisting of BTEX and PAHs. A

111

detailed description of the test site is provided elsewhere.11, 12

18, 19, 49

Owing to the high abundance of aromatic

17, 18

The

Furthermore, NMR spectroscopy has been

Page 5 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 28

112

Groundwater samples were collected from a high resolution multilevel well (HR-MLW), which

113

allowed a spatial vertical resolution of 3 to 10 cm. Thirty-two small volume samples were taken

114

simultaneously to prevent mixing of water from different depths. A list of sampling depths is provided

115

in the Supporting Information (SI). The sampling scheme using this HR-MLW is described in detail

116

elsewhere.11, 12

117

Groundwater geochemistry and occurrence of (parent) contaminants

118

Methods and detailed protocols for the analysis of aromatic hydrocarbons, i.e., BTEX and selected

119

PAHs, are given by Anneser et al.11 Basic physical-chemical parameters such as pH and redox

120

potential (EH) were determined in freshly collected groundwater samples with field sensors (WTW,

121

Weilheim, Germany). Major ions, e.g., sulfate, were analyzed within 2 days by ion chromatography

122

(Dionex DC-100, Idstein, Germany) from samples stored in the dark without headspace and cooled.

123

Fe(II) concentrations were measured by the ferrozine assay.54 For sulfide determination, 200 µL of

124

sample were fixed in duplicates in 1 mL of a 2% (w/v) zinc acetate solution and analyzed on-site

125

following a protocol, which was adapted to smaller volumes, from Cline.55

126

Solid phase extraction of DOM

127

DOM from the 0.2 µm filtered groundwater samples was isolated by an established solid phase

128

extraction (SPE) method described by Dittmar et al.56 A detailed description is given in the SI

129

Methods.

130

Ultra-high resolution mass spectrometry

131

Ultra-high resolution spectra were acquired with a Bruker (Bremen, Germany) SolariX Qe FT-ICR-

132

MS equipped with a 12 Tesla superconducting magnet. Reproducibility of the mass spectra was

133

assured by measuring SPE-DOM from 250 mL and 1 L samples collected at the same depths, also

134

indicating no extraction volume effect. Additionally, FT-ICR mass spectra of samples taken in close

135

vicinity to each other (enabled by the high spatial resolution sampling in the centimeter range)

136

resemble in principle composition, indicating indirectly reproducibility of FT-ICR mass spectra. Page 6 of 28

ACS Paragon Plus Environment

Page 7 of 28

Environmental Science & Technology

137

Further details on the mass spectra acquisition and formula assignment from mass spectra are

138

described in the SI. Hierarchical cluster analyses of assigned FT-ICR-MS peaks were performed with

139

Hierarchical Clustering Explorer Version 3.0 and R.

140

FT-ICR-MS is a semi-quantitative method owing to the lack of calibration standards in non-targeted

141

analysis and differential ionization efficiency in complex and chemically diverse mixtures. The signal

142

intensities of mass peaks are not necessarily directly proportional to their respective concentration;

143

however increase and decrease of mass peak intensities follow the changes in relative concentrations.

144

To address the limited detection of some compounds due to low ionization efficiency and suppression,

145

we performed two orthogonal ionization modes. Additionally, ionization and mass independent

146

quantitative 1H NMR and optical EEM fluorescence spectroscopy was utilized.

147

EEM Fluorescence Spectroscopy and PARAFAC Modeling

148

Excitation emission matrix fluorescence spectroscopy of SPE-DOM was measured using a Jobin

149

Horiba Instruments (Kyoto, Japan) Aqualog spectrofluorometer. Details on the measurements and

150

PARAFAC modeling, which were performed by the drEEM MATLAB toolbox, are given in the SI

151

(Figures S1, S2).38 Principal component analysis (PCA) of Fmax values was carried out with SIMCA-

152

P 9.0 to evaluate the influence of the components on the individual samples.

153

NMR

154

1D and 2D 1H and 13C NMR spectra were acquired from re-dissolved SPE-DOM in CD3OD at 283 K

155

with a Bruker (Karlsruhe, Germany) AV III 800 spectrometer operating at B0 = 18.7 Tesla with

156

Bruker standard pulse sequences and cryogenic detection. A detailed description of the NMR

157

acquisition conditions is described in the SI.

158

Results and Discussion

159

Biogeochemical gradients and distribution patterns of major contaminants and signature metabolites

160

provide critical information on the redox processes involved in transformation of contaminants and Page 7 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 28

161

active biodegradation pathways.5 Typically, biodegradation in organic contaminant plumes in porous

162

aquifers is most pronounced at the plume fringes as repeatedly confirmed by (1) steep and opposing

163

gradients of contaminants and electron acceptors such as sulfate, (2) peaking of the metabolic end

164

products sulfide and Fe(II) from sulfate and iron reduction, respectively, and (3) bacterial cell counts

165

and contaminant specific stable carbon isotope values at the plume fringes.5, 11, 13

166

Samples collected from the petroleum hydrocarbon contaminated aquifer in Düsseldorf Flingern

167

showed steep small-scale gradients for BTEX and naphthalene as the major contaminants (Figure

168

1A). The contaminant plume, at the time of sampling, was located at the groundwater table extending

169

to 0.8 m beneath (6.51–7.31 m below surface (bls)). According to the scheme of Anneser et al., the

170

contaminant plume was subdivided into three compartments (Figure 1).57 The first, the plume core,

171

was defined as the zone exhibiting BTEX concentrations >50% of the observed maximum of 616

172

µmol/L (6.61–7.01 m bls). The second compartment, the upper plume fringe (6.51–6.61 m bls), was

173

bound at the upper side by the groundwater table including the capillary fringe, which was located

174

right above the groundwater table and only partially filled with water. The lower plume fringe, which

175

was the third compartment and located below the plume core, extended from 7.01–7.31 m bls.

176

PAHs, in particular naphthalene, exhibited a high abundance (maximum of 89.5 µmol/L at 6.78 m

177

bls) in the plume core (Figures 1A, B). The three-ring PAHs exhibited the highest abundances ~0.3 m

178

deeper (Figures 1A, B). Acenaphthene was present at approximate concentrations of 7 µmol/L all

179

along the sampled vertical profile (Figure 1B). Previous studies at this site had shown a comparable

180

distribution of the major contaminants, however, with slightly different concentrations of the

181

individual compounds.11, 57

182

Redox potentials between +89 mV at the groundwater table and −149 mV in the plume core indicated

183

strictly anoxic conditions except for the capillary fringe (Figure 1D). The pH profile showed elevated

184

values of up to pH 8.05 in the plume core (Figure 1D). Sulfate as the predominant electron acceptor

185

was almost absent in the plume core and showed a steep counter gradient to the major contaminants at Page 8 of 28

ACS Paragon Plus Environment

Page 9 of 28

Environmental Science & Technology

186

the lower plume fringe. Outside of the contaminant plume, sulfate exhibited a rather constant high

187

concentration of approximately 1.5 mmol/L (Figure 1C). Thus, hydrocarbon degradation was sulfate-

188

limited in the plume core. Sulfide peaked at concentrations of up to 0.17 mmol/L in the plume core

189

and lower plume fringe, indicating a hot spot of sulfate reduction between 6.75 and 7.21 m bls. The

190

discrepancy between the amount of sulfate reduced within the contaminated area and the free sulfide

191

detected in the water is explained by precipitation of reduced sulfur with iron as well as incorporation

192

of reduced sulfur into DOM (see below). As already shown in previous studies11,

193

reduction was the key redox process involved in biodegradation of organic contaminants in the

194

Düsseldorf Flingern aquifer. However, in contrast to earlier surveys, sulfate, sulfide, and BTEX

195

profiles indicated biodegradation by sulfate reduction taking place not only at the lower plume fringe

196

but also in parts of the plume core. These changing patterns may be related to temporal dynamics of

197

contaminant concentrations and biodegradation activities as evaluated by contaminant-specific stable

198

isotope analysis (CSIA) and microbial community analysis (unpublished data). Extraordinary high

199

Fe(II) concentrations of up to 0.38 mmol/L were found close to the groundwater table, thereby

200

hinting, in accordance with previous studies11, at iron reduction to be a relevant redox process at the

201

upper plume fringe.

202

(-)ESI FT-ICR-MS of DOM

203

DOM in pristine groundwater is generally composed of organic molecules from natural carbon

204

sources, such as plants, microbes and fungi, whereas DOM at contaminated sites considerably derives

205

from the anthropogenic carbon input.2 (-)ESI FT-ICR-MS spectra from SPE-DOM showed distinct

206

compositional variation along the aquifer (Figure S3B). In general, sample similarity followed spatial

207

proximity, and hierarchical cluster analysis revealed three distinct subclusters, 6.51–6.64 m bls, 6.67–

208

7.06 m bls, and 7.08–10.20 m bls (Figure 2A). The extensive chemical dissimilarity between the

209

subclusters is depicted in Figure 2B. The van Krevelen diagrams and ring charts display the most

210

representative molecular formulas (correlation coefficient r > 0.9) for each subcluster. More details on

211

the selection of characteristic molecular formulas are given in the SI.

12, 57, 58

, sulfate

Page 9 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 28

212

The differences between (-)ESI FT-ICR mass spectra of the SPE-DOM isolates corresponded to the

213

respective prevailing and clearly distinct zone-dependent redox chemistry (Figures 1, S3B). The

214

specific DOM chemistry in the upper part of the aquifer, including the upper fringe, plume core, and 5

215

cm of the lower plume fringe (cf. blue and red subcluster, Figure 2), was dominated by sulfur organic

216

molecules. This is in agreement with sulfate reduction being the major redox process in the plume and

217

its fringes.12 The samples from 6.67–7.06 m bls (red subcluster, Figure 2) were dominated by CHOS

218

compounds with a contribution to the total intensity of assigned molecular formulas of up to 52%

219

(Figure S4A). This was most likely a result of high sulfide concentrations in these depths generated by

220

sulfate reduction (Figure 1C). The relative proportion of CHO compounds from 6.67–7.06 m bls

221

decreased (Figure S4A), thus, indicating that CHOS compounds were formed from CHO compounds.

222

The formation of sulfur organic compounds at elevated sulfide concentrations is a common

223

biogeochemical process

224

DOM and model organic compounds.63-66 Sulfide reacted with organic matter functionalities such as

225

single and activated double bonds (Michael systems, Quinones) at the boundary of the contamination

226

plume.63-66 Therefore, we anticipate direct reaction of highly nucleophilic sulfide with DOM

227

functional entities to result in CHOS molecules. CHOS compounds might have been also formed

228

directly during sulfate reduction. Sulfide concentrations in the upper plume fringe were low because

229

of the supposed dynamic re-oxidation of sulfide to sulfate close to the capillary fringe.67 Furthermore,

230

the Fe(II) concentration was elevated in the upper fringe (Figure 1), which likely led to precipitation

231

of sulfide with Fe(II) and implied a dominance of iron over sulfate reduction in this zone.12 Thus,

232

DOM chemistry in the upper fringe and the topmost centimeters of the plume core was affected

233

differently compared to the area ~20 cm below. Einsiedl et al. found elevated sulfite concentrations

234

close to the capillary fringe; thus, characteristic CHOS compounds from 6.51–6.64 m bls (blue

235

subcluster, Figure 2) potentially resulted from sulfonation reactions.67 With respect to contaminants,

236

the influence of ionizable aromatic molecules (H/C < 1) was higher in the top centimeters close to the

237

groundwater table, thereby indicating fewer degraded contaminants present inside the upper plume

59-62

that has also been confirmed in several laboratory experiments with

Page 10 of 28

ACS Paragon Plus Environment

Page 11 of 28

Environmental Science & Technology

238

fringe, which was likely due to the limited access to electron acceptors (Figures 1D, 2).11 A

239

substantial portion of characteristic CHNO and especially CHNOS compounds suggested significant

240

nitrogen DOM chemistry in the upper plume fringe where dissolved sulfide was almost absent.

241

The characteristic mass peaks of SPE-DOM from the deep zone below 7.08 m bls showed mainly

242

CHO and CHNO compounds in the polyphenolic region of the van Krevelen diagram (1.5 > H/C >

243

0.7 and 0.7 > O/C > 0.1) (Figure 2B). Hence, DOM characteristics of less contaminated deep zone

244

groundwater showed similarities to previously published (-)ESI FT-ICR-MS results of groundwater

245

DOM.68, 69 The comparably low number of characteristic mass peaks for deep zone SPE-DOM arose

246

from the fact that these samples represented general DOM patterns of the sampling site, which was

247

only marginally impacted by petroleum contaminants, degradation products, and intense sulfur redox

248

chemistry (Figure S3A).

249

(+)APPI FT-ICR-MS of DOM

250

To address limited sensitivity of some compounds due to low ionization efficiency and suppression in

251

(-)ESI FT-ICR-MS we performed (+)APPI FT-ICR-MS as a nearly orthogonal ionization mode

252

(Figure S3C).30-32 (+)APPI is the method of choice for the detection of molecules with extended,

253

conjugated π-electron systems such as petroleum contaminants and their degradation products.30 As

254

an effect of the sample preparation, which was not optimized for the extraction of nonpolar

255

contaminants but for DOM extraction, a mere 6% of the annotated molecular formulas of the whole

256

data set were oxygen-free molecules. Therefore, only the results of the oxygen containing CHNOS

257

molecular formulas are presented and discussed.

258

Hierarchical cluster analysis of (+)APPI FT-ICR mass peaks and the associated van Krevelen

259

diagrams of characteristic molecular formulas of the subclusters showed that all samples above 7.06

260

m bls were dominated by aromatic and even condensed aromatic molecules (low H/C and O/C ratios)

261

(Figures 3B, 3C). These molecules are typically derived from petroleum, thus the segmentation of the

262

cluster was clearly driven by the elevated contamination input in the top ~50 cm beneath the Page 11 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 28

263

groundwater table (6.54 m bls). Analogous to the (-)ESI FT-ICR-MS results, the proportion of

264

characteristic CHNO and CHNOS compounds in the contamination plume and fringes was greater in

265

the samples taken close to the groundwater table further corroborating the hypothesis that organic

266

nitrogen chemistry played a role in this depth region (Figure 3C). Furthermore, (+)APPI FT-ICR-MS

267

showed that a large proportion of characteristic formulas were sulfur organic compounds, which was

268

in accordance with the (-)ESI FT-ICR-MS results and the intensive bacterial sulfate reduction

269

(Figures 3B, 3C, S5).12,

270

formulas did not show any clear trend in the highly sulfidic region (Figure S4B). Thus, CHOS

271

compounds formed as a result of bacterial sulfate reduction were proposed to derive from highly

272

functionalized DOM molecules (i.e. carboxylic acids, carbonyl derivatives, alcohols,…), which have

273

a greater (-)ESI ionization efficiency.30

274

On the contrary, SPE-DOM isolates from the deeper zone of the aquifer were characterized by a large

275

combined abundance (greater 80% of the overserved characteristic mass peaks describing the overall

276

diversity) of CHO and CHNO compounds with 1.5 > H/C > 1 and 0.6 > O/C > 0.2 (Figure 3C, note

277

that due to limitations in the visualization, CHO compounds are partially hidden by the CHNO

278

compounds); therefore, corresponding to APPI FT-ICR mass spectra of DOM from previous

279

studies.32, 31 This implied that DOM from the lower fringe and below was presumably not heavily

280

affected by the contamination and its degradation products at the time of sampling.

281

Field and culture degradation studies of BTEX, PAHs and heteroatom PAHs at the same or similar

282

conditions were able to identify typical degradation products.7,

283

molecular formula annotation combined with the profound literature knowledge, the respective mass

284

peaks most probably derive from these known degradation products. Figure 3D shows the relative

285

mass peak intensities for the four highest abundant putative degradation products originating from

286

BTEX, PAHs and heteroatom PAHs as well as their methylated relatives methylnaphthyl-2-

287

methylsuccinic acid (MNMS, C16H16O4), naphthyl-2-methyl-succinic acid (NMS, C15H14O4),

67

However, the intensity-weighted contribution of assigned molecular

71-75

Due to the high accuracy in

Page 12 of 28

ACS Paragon Plus Environment

Page 13 of 28

Environmental Science & Technology

288

methylnaphthoic acid (MNA, C12H10O2), and acenaphthene-5-carboxylic acid (AC, C13H10O2).7, 70-74

289

The depth profile was in accordance with the distribution of plume marker compounds,

290

hydrochemical parameters, and redox chemistry markers (Figures 1, 3D). Moreover, MNMS, NMS,

291

AC, and an unknown compound (C15H12O2) represented the highly abundant characteristic masses of

292

the 6.75 to 7.03 m bls subcluster (Figure 3C). Hence, the initial steps of biodegradation, which have

293

been identified in pervious degradation studies

294

plume core where bacterial sulfate reduction was highly active.

295

EEM fluorescence spectroscopy of DOM

296

EEM-PARAFAC modeling was validated for a three component model and the relative contribution

297

of each component was identified for the SPE-DOM samples taken along the aquifer (Figures 4, S1,

298

S2). The first component (C1) showed an excitation/emission (Ex/Em) maximum at 237(288)/360

299

nm, which was similar to the PARAFAC components modeled in Deepwater Horizon oil spill studies

300

resembling oil and its degradation products as well as BTEX- and PAH-like fluorescence (Table

301

S2).46, 47, 48 The second component (C2) identified at an Ex/Em maximum of 255(315)/420 nm, which

302

was in accordance with humic-like fluorescence, has been found in many PARAFAC studies on

303

organic matter.75, 44, 76 Also, PARAFAC studies on the Deepwater Horizon oil spill showed similar

304

components, which are described as humic-like therein (Table S2).45-48 The third component (C3)

305

exhibited its Ex/Em maximum at 255/333 nm matching previously published components that were

306

reported as oil, BTEX, and PAH related (Table S2).46, 48 High bacterial activity was expected to show

307

protein-like fluorescence, however this signal was absent from zones of highly active sulfate reducing

308

bacteria. In fact, the typical protein-like fluorescence has never been observed under anoxic

309

conditions from sulfate reducing bacteria but has often been associated with aerobic heterotrophic

310

bacteria.77 At this point, it is not clear why anaerobic conditions would prevent the release of water

311

soluble proteins that contain tyrosine and/or tryptophan, however this was not the aim of this study.

312

Furthermore, Wang et al. 2015 showed significant quenching of tyrosine and tryptophan fluorescence

313

as well as shift of the peak position in the presence of humic substances.78

7, 71-75

, were mainly located in the lower part of the

Page 13 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 28

314

To distinguish fluorescence spectra of SPE-DOM dominated by contamination-related versus typical

315

DOM fluorescence, PCA of Fmax values of the components was performed (Figure 4). PCA revealed

316

that SPE-DOM from 6.59 to 6.96 m bls were mainly driven by C1, which was annotated to oil, BTEX

317

and, PAH, in addition to its degradation products. The samples from 7.01 to 7.21 m bls showed a

318

strong correlation with C3, which was also described as oil, BTEX, PAH, and its degradation

319

products. Therefore, C3 fluorescence most likely derived from further degraded products and three-

320

ring PAHs that were more abundant in this region (Figure 1B). Samples in close vicinity of the

321

groundwater table and the ones from the deep zone (highlighted in green, Figure 4B) were

322

characterized by C2, showing that their fluorescence was humic-like and not strongly affected by the

323

contamination. The ratios of C1/C2 and C3/C2, which indicated oil/humic contributions to SPE-DOM

324

composition, were elevated for the samples ranging from 6.61–7.08 m bls. Therefore, EEM

325

fluorescence spectroscopy also successfully differentiated SPE-DOM influenced by the contamination

326

plume.

327

NMR spectroscopy of DOM

328

In addition to FT-ICR-MS spectrometry, with its already discussed susceptibility to selective

329

ionization and suppression effects, and EEM fluorescence spectroscopy, which is only able to detect

330

fluorophores, NMR spectroscopy was utilized. NMR does not suffer from any of these potential

331

distortions and provides quantitative depiction of relevant structural features in complex organic

332

mixtures.18 But 1H and

333

EEM fluorescence spectroscopy and is rather limited in its capability to provide in depth information

334

about nitrogen and sulfur containing compounds compared to FT-ICR mass spectrometry.18 1H NMR

335

spectroscopy of selected SPE-DOM sampling positions that were indicative of changing geochemical

336

parameters and contamination levels provided detailed structural information and substantial variance

337

according to zonation (Figures 1, 5). These variations at first indicated variable contributions of

338

common bulk DOM with broad NMR resonances and of petroleum degradation products showing

339

sharp NMR resonances with well resolved splittings from J-couplings (Figure 5) (cf. Tables S3, S4 for

13

C NMR spectroscopy is less sensitive compared to both FT-ICR-MS and

Page 14 of 28

ACS Paragon Plus Environment

Page 15 of 28

Environmental Science & Technology

340

1

341

units of observed DOM and petroleum degradation products is listed in Table S5.

342

In general, the relative contribution of aromatic structures (δ(1H) = 9.5–7.0 ppm) ranged up to 30% in

343

the plume core and fringes, while it was below 15% for SPE-DOM close to the groundwater table

344

(6.64 m bls) and the deeper zone (10.20 m bls) (Tables S3, S4). Figure S6 depicts overlaid 1H NMR

345

spectra adjusted to common amplitude of the aliphatic bulk DOM signature (δ(1H) = 0.8–2.0 ppm;

346

Figure S6A) and alternatively area-normalized 1H NMR spectra across the entire range of chemical

347

shifts (δ(1H) = 0.5–9.5 ppm; Figure S6B). 1H NMR spectra clearly showed a nearly constant but

348

somewhat variable background bulk DOM signature present in all SPE-DOM (Figure S6A) and

349

highly variable proportions of PAH and its degradation products. In particular, not only the ratios of

350

aromatic protons (δ(1H) > 7 ppm) and those of HOOC-CαH-protons (δ(1H) ~ 2.1–2.7 ppm) differed

351

quite substantially, but the line shapes within these sections of chemical shifts also greatly differed.

352

This demonstrated a huge variance of PAH-degradation products present at each individual depth of

353

sampling and again testified to the unique capability of 1H NMR spectroscopy in the elucidation of

354

complex environmental unknowns.16-19

355

As the petroleum contamination arose mainly from BTEX and naphthalene, only a very small

356

percentage of the NMR total integral derived from multi (>3) ring condensed aromatics with δ(1H) >

357

8.3 ppm. Most of the aromatic structures were found in the chemical shift region δ(1H) = 8.3–7.3

358

ppm, which was indicative of single and two-ring aromatics, six membered N-heterocycles, and

359

especially oxidized aromatics (i.e., mainly polycarboxylated). These types of structural features were

360

in general matching the contaminants and their degradation products.7, 70 The two SPE-DOM from the

361

top and bottom of the aquifer, while assumed representative of common groundwater DOM,

362

nevertheless showed elevated proportions of petroleum derivatives compared to pristine DOM

363

levels.69 This background concentration was drastically increased in the plume core, and the co-

364

occurrence of contaminants and their degradation products was confirmed by co-occurring sharp

H NMR section integrals according to DOM substructures). A detailed description of key structural

Page 15 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

365

NMR resonances at δ(1H) > 7 ppm (aromatics) and δ(1H) ~ 2.1–2.7 ppm (aliphatic carboxylic acids).

366

Additionally, J-resolved (JRES) and total correlation (TOCSY) spectroscopy indicated the presence of

367

multiring polycyclic aromatic compounds. Most of them showed two rings (Figure S7), in accordance

368

with common composition of groundwater contamination by petroleum products.11, 12 Furthermore,

369

aliphatic cross peaks in TOCSY and methylene-edited 1H, 13C distortion less enhanced by polarization

370

transfer heteronuclear single quantum coherence (DEPT HSQC) NMR spectra showed typical DOM

371

signatures as well as succinic acid side chains attached to aromatics (Figure S8) that are typical for

372

PAH degradation products such as NMS, which were also detected in the (+)APPI FT-ICR mass

373

spectra (Figure 3D).7, 70, 72, 73 The proportion of non-functionalized monoaromatic structures (δ(1H) =

374

7.3–7.0 ppm) was lower but might be underestimated because of losses during the sample preparation

375

as a consequence of their high volatility.

376

The region δ(1H) = 7.0–6.5 ppm of Csp2H likely originated from OH and OR substituted aromatics of

377

common SPE-DOM and partially indicated aromatics with fused alicyclic units, which are

378

constituents of partly biodegraded mineral oil.52 Otherwise, conjugated (δ(1H) = 6.5–6.0 ppm) and

379

isolated double bonds (δ(1H) = 6.0–5.1 ppm) likely represented natural product derivatives present in

380

groundwater SPE-DOM.

381

The slight depletion of carbohydrates (δ(1H) = 4.9–3.1 ppm) in SPE-DOM from depths with high

382

contamination levels might be an artifact of the sample preparation as PPL cartridges discriminate

383

against carbohydrates.19 Also, the region of carboxyl-rich alicyclic molecules (CRAM) from δ(1H) =

384

3.1–1.9 ppm showed no clear trend because of the interference of CRAM and contamination

385

degradation products.

386

Contaminant abundance and that of “original” SPE-DOM signatures were inversely related across the

387

aquifer (Figure S6B, radar diagram). However, the region of non-functionalized aliphatics (CCCH;

388

δ(1H) = 1.9–0.5 ppm), a major contributor of common SPE-DOM, maintained its core integrity with

389

limited but clearly recognizable variance (Figure 5A) across the entire range of sampling.19 Page 16 of 28

ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Environmental Science & Technology

390

DOM characteristics and contamination biodegradation

391

Non-targeted high resolution analytics in combination with spatially highly resolved sampling

392

revealed DOM signatures from petroleum contaminants and specific degradation products (Figures 3–

393

5, Tables S2–S5, Figures S6–S8). Additionally, our approach enabled a comprehensive molecular

394

level understanding of natural DOM and its role in petroleum hydrocarbon bioremediation. Previous

395

studies at the investigated site identified bacterial sulfate reduction as the major redox process

396

involved in the contaminant biodegradation by isotope ratio analysis.11,

397

demonstrates that the DOM characteristics were highly affected by bacterial sulfate reduction. Both

398

(+)APPI FT-ICR-MS, which favors the detection of petroleum derived compounds, and (-)ESI FT-

399

ICR-MS, which is preferably applied for the characterization of natural DOM, revealed characteristic

400

CHOS compounds at the depths where sulfate reduction is active. Organosulfur compounds were

401

found by far more abundant in the (-)ESI FT-ICR mass spectra (Fig. S4). Thus, reduced sulfur mainly

402

reacts with natural DOM. CHOS compounds represented >50% of the total mass peak intensity of the

403

corresponding assigned molecular formulas derived from (-)ESI FT-ICR mass spectra in the zone

404

with maximum sulfide concentrations (Figures 1, 2, S4A). Especially in zones where Fe(II) and other

405

complexation partners were limited, sulfide remained in solution. Contaminant biodegradation is

406

limited by the availability of sulfate as electron acceptor at the investigated site. Elevated sulfide

407

concentrations may even further decelerate petroleum biodegradation at low sulfate concentrations;

408

indeed, high concentrations of sulfide are toxic to many microbes.79 In this context, natural DOM may

409

act as a beneficial key player in the mediation of sulfide resulting from bacterial sulfate reduction. In

410

consequence, natural DOM may be applied in bioremediation. Future work will thus address the

411

linkage of contaminant biodegradation, redox processes, DOM, and the microbial communities.

412

Acknowledgements

413

We thank C. Kellermann and K. Hörmann for their contribution in the field. We are grateful to J.

414

Luek for support with the EEM fluorescence spectroscopy measurements. The authors S.E.-M. D. and

12

The current study

Page 17 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 28

415

H.M. thank the Helmholtz WasserZentrum München for funding. This is contribution (xxxx, to be

416

filled in after acceptance of manuscript) of the University of Maryland Center for Environmental

417

Science.

418

Supporting Information Available

419

The SI contains extended information on methods (Table S1 and Figures S1, S2) and results and

420

discussion (Tables S2–S5 and Figures S3–S8). This material is available free of charge via the

421

Internet at http://pubs.acs.org.

422

TOC/Abstract Art

423

Page 18 of 28

ACS Paragon Plus Environment

Page 19 of 28

Environmental Science & Technology

424

425 426 427 428 429 430 431 432

Figure 1. Vertical concentration profiles of main contaminants (A) BTEX and (B) PAHs, (C) sulfide, sulfate and Fe(II) as well as (D) the hydrochemical parameters pH and redox potential (depth in m below surface (m bls)). The black dashed line depicts the groundwater table. Values represent single or means of duplicate and triplicate measurements (error bars depict standard deviations), respectively. The assumed contamination plume is highlighted in gray and the plume core defined by >50% of the maximal BTEX concentration is depicted by the solid gray line. For further information, see text.

433

Page 19 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 28

434 435 436 437 438 439 440 441

Figure 2. (A) Hierarchical cluster analysis of the assigned (-)ESI FT-ICR-MS derived molecular formulas observed in SPE-DOM along the aquifer with samples named according to their sampling depth (bls = below surface). (B) Van Krevelen diagrams depict the most representative characteristic molecular formulas (correlation coefficient > 0.9; cf. SI) for the three main subclusters, representing the specific DOM chemistry. Elemental composition of the molecular formulas are represented as CHO (blue), CHNO (orange), CHOS (green), and CHNOS (red), and the bubble area depicts the mean mass peak intensity within the respective subcluster.

442 443

Page 20 of 28

ACS Paragon Plus Environment

Page 21 of 28

Environmental Science & Technology

444 445 446 447 448 449 450 451 452 453 454 455

Figure 3. (A) Hierarchical cluster analysis of the assigned (+)APPI FT-ICR-MS derived molecular formulas observed in SPE-DOM along the aquifer with samples named according to their sampling depth. (B) Van Krevelen diagram of shared characteristic molecular formulas of the two upper subclusters, which were highly affected by the contamination plume and (C) van Krevelen diagrams of the characteristic molecular formulas for the three main subclusters representing the specific SPEDOM chemistry. Elemental composition of the molecular formulas are represented as CHO (blue), CHNO (orange), CHOS (green), and CHNOS (red), and the bubble area depicts the mean mass peak intensity within the respective subcluster. Note that the size of the three and four most abundant CHO (blue) bubbles of the purple (B) and red subcluster (C) had to be down-sized because of their very high mass peak intensity. (D) (+)APPI FT-ICR-MS derived abundance profile of main petroleum contamination degradation products.

456

Page 21 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 28

457 458 459 460 461

Figure 4. (A) EEM fluorescence spectra of three PARAFAC-derived components C1–C3. (B) PCA of SPE-DOM along the aquifer based on the Fmax values of the PARAFAC components. Samples are labeled according to their sampling depth in m bls. H corresponds to 250 mL samples, and L corresponds to 1 L samples.

462

Page 22 of 28

ACS Paragon Plus Environment

Page 23 of 28

Environmental Science & Technology

463 464 465 466 467

Figure 5. (A) 1H NMR spectra (800 MHz, CD3OD) of selected SPE-DOM with depth provided and color coded according to similarity of contamination patterns and geochemical parameters. (B) Bar charts indicate relative 1H NMR section integrals according to substructures shown in Tables S3 and S4 and the main structural sections are depicted with dotted lines.

468

Page 23 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 28

469

References

470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518

1. Minor, E. C.; Swenson, M.; Mattson, B. M.; Oyler, A. R., Structural characterization of dissolved organic matter: a review of current techniques for isolation and analysis. Environ. Sci.: Processes Impacts 2014, 16, 2064-2079; DOI 10.1039/C4EM00062E. 2. Shen, Y.; Chapelle, F. H.; Strom, E. W.; Benner, R., Origins and bioavailability of dissolved organic matter in groundwater. Biogeochemistry 2014, 122 (1), 61-78; DOI 10.1007/s10533-0140029-4. 3. Griebler, C.; Lueders, T., Microbial biodiversity in groundwater ecosystems. Freshwater Biol. 2009, 54 (4), 649-677; DOI 10.1111/j.1365-2427.2008.02013.x. 4. Fitts, C. R. Groundwater Science (Second Edition); Elsevier: Amsterdam, 2012. 5. Meckenstock, R. U.; Elsner, M.; Griebler, C.; Lueders, T.; Stumpp, C.; Aamand, J.; Agathos, S. N.; Albrechtsen, H. J.; Bastiaens, L.; Bjerg, P. L.; Boon, N.; Dejonghe, W.; Huang, W. E.; Schmidt, S. I.; Smolders, E.; Sorensen, S. R.; Springael, D.; van Breukelen, B. M., Biodegradation: Updating the Concepts of Control for Microbial Cleanup in Contaminated Aquifers. Environ. Sci. Techol. 2015, 49 (12), 7073-7081; DOI 10.1021/acs.est.5b00715. 6. Vogt, C.; Richnow, H. H., Bioremediation via in situ microbial degradation of organic pollutants. Advances in biochemical engineering/biotechnology 2014, 142, 123-46; DOI 10.1007/10_2013_266. 7. Jobelius, C.; Ruth, B.; Griebler, C.; Meckenstock, R. U.; Hollender, J.; Reineke, A.; Frimmel, F. H.; Zwiener, C., Metabolites indicate hot spots of biodegradation and biogeochemical gradients in a high-resolution monitoring well. Environ. Sci. Techol. 2011, 45 (2), 474-481; DOI 10.1021/es1030867. 8. Eberhardt, C.; Grathwohl, P., Time scales of organic contaminant dissolution from complex source zones: coal tar pools vs. blobs. J. Contam. Hydrol. 2002, 59 (1-2), 45–66; DOI 10.1016/S01697722(02)00075-X. 9. Moeckel, C.; Monteith, D. T.; Llewellyn, N. R.; Henrys, P. A.; Pereira, M. G., Relationship between the Concentrations of Dissolved Organic Matter and Polycyclic Aromatic Hydrocarbons in a Typical U.K. Upland Stream. Environ. Sci. Techol. 2014, 48 (1), 130–138; DOI 10.1021/es403707q. 10. Wiedemeier, T. H.; Rifai, H. S.; Newell, C. J.; Wilson, J. T. Natural Attenuation of Fuels and Chlorinated Solvents in the Subsurface; John Wiley & Sons, Inc.: New York, 1999. 11. Anneser, B.; Einsiedl, F.; Meckenstock, R. U.; Richters, L.; Wisotzky, F.; Griebler, C., Highresolution monitoring of biogeochemical gradients in a tar oil-contaminated aquifer. Appl. Geochem. 2008, 23 (6), 1715-1730; DOI 10.1016/j.apgeochem.2008.02.003. 12. Anneser, B.; Pilloni, G.; Bayer, A.; Lueders, T.; Griebler, C.; Einsiedl, F.; Richters, L., High Resolution Analysis of Contaminated Aquifer Sediments and Groundwater—What Can be Learned in Terms of Natural Attenuation? Geomicrobiol. J. 2010, 27 (2), 130-142; DOI 10.1080/01490450903456723. 13. Bauer, R. D.; Maloszewski, P.; Zhang, Y.; Meckenstock, R. U.; Griebler, C., Mixingcontrolled biodegradation in a toluene plume--results from two-dimensional laboratory experiments. J. Contam. Hydrol. 2008, 96 (1-4), 150-168; DOI 10.1016/j.jconhyd.2007.10.008. 14. Foght, J., Anaerobic biodegradation of aromatic hydrocarbons: pathways and prospects. J. Mol. Microbiol. Biotechnol. 2008, 15 (2-3), 93-120; DOI 10.1159/000121324. 15. Stubbins, A.; Lapierre, J. F.; Berggren, M.; Prairie, Y. T.; Dittmar, T.; Del Giorgio, P. A., What's in an EEM? Molecular Signatures Associated with Dissolved Organic Fluorescence in Boreal Canada. Environ. Sci. Techol. 2014, 48 (18), 10598-10606; DOI 10.1021/es502086e. 16. Simpson, A. J.; Simpson, M. J.; Soong, R., Nuclear Magnetic Resonance Spectroscopy and Its Key Role in Environmental Research. Environ. Sci. Techol. 2012, 46 (21), 11488-11496; DOI 10.1021/es302154w. 17. Simpson, M. J.; Simpson, A. J. NMR Spectroscopy: A Versatile Tool for Environmental Research; John Wiley & Sons Inc: New York, 2014. Page 24 of 28

ACS Paragon Plus Environment

Page 25 of 28

519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570

Environmental Science & Technology

18. Hertkorn, N.; Ruecker, C.; Meringer, M.; Gugisch, R.; Frommberger, M.; Perdue, E. M.; Witt, M.; Schmitt-Kopplin, P., High-precision frequency measurements: indispensable tools at the core of the molecular-level analysis of complex systems. Anal. Bioanal. Chem. 2007, 389 (5), 1311-27; DOI 10.1007/s00216-007-1577-4. 19. Hertkorn, N.; Harir, M.; Koch, B. P.; Michalke, B.; Schmitt-Kopplin, P., High-field NMR spectroscopy and FTICR mass spectrometry: powerful discovery tools for the molecular level characterization of marine dissolved organic matter. Biogeosciences 2013, 10 (3), 1583-1624; DOI 10.5194/bg-10-1583-2013. 20. Koch, B. P.; Dittmar, T.; Witt, M.; Kattner, G., Fundamentals of Molecular Formula Assignment to Ultrahigh Resolution Mass Data of Natural Organic Matter. Anal. Chem. 2007, 79 (4), 1758-1763; DOI 10.1021/ac061949s. 21. Kellerman, A. M.; Dittmar, T.; Kothawala, D. N.; Tranvik, L. J., Chemodiversity of dissolved organic matter in lakes driven by climate and hydrology. Nat. Commun. 2014, 5 (3804), 1-8; DOI 10.1038/ncomms4804. 22. Ward, N. D.; Keil, R. G.; Medeiros, P. M.; Brito, D. C.; Cunha, A. C.; Dittmar, T.; Yager, P. L.; Krusche, A. V.; Richey, J. E., Degradation of terrestrially derived macromolecules in the Amazon River. Nat. Geosci. 2013, 6, 530–533; DOI 10.1038/ngeo1817. 23. Lechtenfeld, O. J.; Kattner, G.; Flerus, R.; McCallister, S. L.; Schmitt-Kopplin, P.; Koch, B. P., Molecular transformation and degradation of refractory dissolved organic matter in the Atlantic and Southern Ocean. Geochim. Cosmochim. Acta 2014, 126, 321-337; DOI 10.1016/j.gca.2013.11.009. 24. Dubinenkov, I.; Flerus, R.; Schmitt-Kopplin, P.; Kattner, G.; Koch, B. P., Origin-specific molecular signatures of dissolved organic matter in the Lena Delta. Biogeochemistry 2014, 123 (1-2), 1-14; DOI 10.1007/s10533-014-0049-0. 25. Gonsior, M.; Schmitt-Kopplin, P.; Bastviken, D., Depth-dependent molecular composition and photo-reactivity of dissolved organic matter in a boreal lake under winter and summer conditions. Biogeosciences 2013, 10, 6945-6956; DOI 10.5194/bg-10-6945-2013. 26. Minor, E. C.; Steinbring, C. J.; Longnecker, K.; Kujawinski, E. B., Characterization of dissolved organic matter in Lake Superior and its watershed using ultrahigh resolution mass spectrometry. Org. Geochem. 2012, 43, 1-11; DOI 10.1016/j.orggeochem.2011.11.007. 27. Gonsior, M.; Schmitt-Kopplin, P.; Stavklint, H.; Richardson, S. D.; Hertkorn, N.; Bastviken, D., Changes in Dissolved Organic Matter during the Treatment Processes of a Drinking Water Plant in Sweden and Formation of Previously Unknown Disinfection Byproducts. Environ. Sci. Techol. 2014, 48 (21), 12714–12722; DOI 10.1021/es504349p. 28. Gonsior, M.; Zwartjes, M.; Cooper, W. J.; Song, W.; Ishida, K. P.; Tseng, L. Y.; Jeung, M. K.; Rosso, D.; Hertkorn, N.; Schmitt-Kopplin, P., Molecular characterization of effluent organic matter identified by ultrahigh resolution mass spectrometry. Water Res. 2011, 45 (9), 2943-2953; DOI 10.1016/j.watres.2011.03.016. 29. Cortés-Francisco, N.; Caixach, J., Molecular Characterization of Dissolved Organic Matter through a Desalination Process by High Resolution Mass Spectrometry. Environ. Sci. Techol. 2013, 47 (17), 9619-9627; DOI 10.1021/es4000388. 30. Sleighter, R. L.; Hatcher, P. Fourier Transform Mass Spectrometry for the Molecular Level Characterization of Natural Organic Matter: Instrument Capabilities, Applications, and Limitations. In Fourier Transforms - Approach to Scientific Principles; Nikolic, G., Ed.; InTech: Rijeka 2011; pp 295-320. 31. D'Andrilli, J.; Dittmar, T.; Koch, B. P.; Purcell, J. M.; Marshall, A. G.; Cooper, W. T., Comprehensive characterization of marine dissolved organic matter by Fourier transform ion cyclotron resonance mass spectrometry with electrospray and atmospheric pressure photoionization. Rapid Commun. Mass Spectrom. 2010, 24 (5), 643-650; DOI 10.1002/rcm.4421. 32. Hertkorn, N.; Frommberger, M.; Witt, M.; Koch, B. P.; Schmitt-Kopplin, P.; Perdue, E. M., Natural Organic Matter and the Event Horizon of Mass Spectrometry. Anal. Chem. 2008, 80 (23), 8908-8919; DOI 10.1021/ac800464g. Page 25 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621

Page 26 of 28

33. Podgorski, D. C.; McKenna, A. M.; Rodgers, R. P.; Marshall, A. G.; Cooper, W. T., Selective ionization of dissolved organic nitrogen by positive ion atmospheric pressure photoionization coupled with Fourier transform ion cyclotron resonance mass spectrometry. Anal. Chem. 2012, 84 (11), 50855090; DOI 10.1021/ac300800w. 34. Bae, E.; Na, J.-G.; Chung, S. H.; Kim, H. S.; Kim, S., Identification of about 30 000 Chemical Components in Shale Oils by Electrospray Ionization (ESI) and Atmospheric Pressure Photoionization (APPI) Coupled with 15 T Fourier Transform Ion Cyclotron Resonance Mass Spectrometry (FT-ICR MS) and a Comparison to Conventional Oil. Energy & Fuels 2010, 24 (4), 2563-2569; DOI 10.1021/ef100060b. 35. Purcell, J. M.; Hendrickson, C. L.; Rodgers, R. P.; Marshall, A. G., Atmospheric Pressure Photoionization Fourier Transform Ion Cyclotron Resonance Mass Spectrometry for Complex Mixture Analysis. Anal. Chem. 2006, 78 (16), 5906–5912; DOI 10.1021/ac060754h. 36. Purcell, J. M.; Merdrignac, I.; Rodgers, R. P.; Marshall, A. G.; Gauthier, T.; Guibard, I., Stepwise Structural Characterization of Asphaltenes during Deep Hydroconversion Processes Determined by Atmospheric Pressure Photoionization (APPI) Fourier Transform Ion Cyclotron Resonance (FT-ICR) Mass Spectrometry. Energy & Fuels 2010, 24 (4), 2257-2265; DOI 10.1021/ef900897a. 37. Jiang, B.; Liang, Y.; Xu, C.; Zhang, J.; Hu, M.; Shi, Q., Polycyclic aromatic hydrocarbons (PAHs) in ambient aerosols from Beijing: characterization of low volatile PAHs by positive-ion atmospheric pressure photoionization (APPI) coupled with Fourier transform ion cyclotron resonance. Environ. Sci. Techol. 2014, 48 (9), 4716-4723; DOI 10.1021/es405295p. 38. Murphy, K. R.; Stedmon, C. A.; Graeber, D.; Bro, R., Fluorescence spectroscopy and multiway techniques. PARAFAC. Anal. Methods 2013, 5 (23), 6541–6882; DOI 10.1039/c3ay41160e. 39. Stedmon, C. A.; Markager, S.; Bro, R., Tracing dissolved organic matter in aquatic environments using a new approach to fluorescence spectroscopy. Mar. Chem. 2003, 82 (3–4), 239254; DOI 10.1016/S0304-4203(03)00072-0. 40. Sanchez, N. P.; Skeriotis, A. T.; Miller, C. M., A PARAFAC-based long-term assessment of DOM in a multi-coagulant drinking water treatment scheme. Environ. Sci. Techol. 2014, 48 (3), 15821591; DOI 10.1021/es4049384. 41. Li, W. T.; Chen, S. Y.; Xu, Z. X.; Li, Y.; Shuang, C. D.; Li, A. M., Characterization of dissolved organic matter in municipal wastewater using fluorescence PARAFAC analysis and chromatography multi-excitation/emission scan: a comparative study. Environ. Sci. Techol. 2014, 48 (5), 2603-2609; DOI 10.1021/es404624q. 42. Murphy, K. R.; Hambly, A.; Singh, S.; Henderson, R. K.; Baker, A.; Stuetz, R.; Khan, S. J., Organic matter fluorescence in municipal water recycling schemes: toward a unified PARAFAC model. Environ. Sci. Techol. 2011, 45 (7), 2909-2916; DOI 10.1021/es103015e. 43. Fasching, C.; Behounek, B.; Singer, G. A.; Battin, T. J., Microbial degradation of terrigenous dissolved organic matter and potential consequences for carbon cycling in brown-water streams. Sci. Rep. 2014, 4 (4981), 1-7; DOI 10.1038/srep04981. 44. Huang, S.; Wang, Y.; Lei Tong; Ma, T.; Wang, Y.; Liu, C.; Zhao, L., Linking groundwater dissolved organic matter to sedimentary organic matter from a fluvio-lacustrine aquifer at Jianghan Plain, China by EEM-PARAFAC and hydrochemical analyses. Sci. Total Environ. 2015, 529 (1), 131-139; DOI 10.1016/j.scitotenv.2015.05.051. 45. Bianchi, T. S.; Osburn, C.; Shields, M. R.; Yvon-Lewis, S.; Young, J.; Guo, L.; Zhou, Z., Deepwater horizon oil in gulf of Mexico waters after 2 years: transformation into the dissolved organic matter pool. Environ. Sci. Techol. 2014, 48 (16), 9288-9297; DOI 10.1021/es501547b. 46. Zhou, Z.; Guo, L., Evolution of the optical properties of seawater influenced by the Deepwater Horizon oil spill in the Gulf of Mexico. Environ. Res. Lett. 2012, 7 (2), 1-12; DOI 10.1088/1748-9326/7/2/025301. 47. Zhou, Z.; Guo, L.; Shiller, A. M.; Lohrenz, S. E.; Asper, V. L.; Osburn, C. L., Characterization of oil components from the Deepwater Horizon oil spill in the Gulf of Mexico using Page 26 of 28

ACS Paragon Plus Environment

Page 27 of 28

622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671

Environmental Science & Technology

fluorescence EEM and PARAFAC techniques. Mar. Chem. 2013, 148, 10-21; DOI 10.1016/j.marchem.2012.10.003. 48. Mendoza, W. G.; Riemer, D. D.; Zika, R. G., Application of fluorescence and PARAFAC to assess vertical distribution of subsurface hydrocarbons and dispersant during the Deepwater Horizon oil spill Environ. Sci.: Processes Impacts 2013, 15 (5), 1017-1030; DOI 10.1039/c3em30816b. 49. Lechtenfeld, O. J.; Hertkorn, N.; Shen, Y.; Witt, M.; Benner, R., Marine sequestration of carbon in bacterial metabolites. Nat. Commun. 2015, 6 (6711), 1-8; DOI 10.1038/ncomms7711. 50. Mazzei, P.; Piccolo, A., Interactions between natural organic matter and organic pollutants as revealed by NMR spectroscopy. Magnetic Resonance in Chemistry 2015, 53 (9), 667-678; DOI 10.1002/mrc.4209. 51. Sudasinghe, N.; Cort, J. R.; Hallen, R.; Olarte, M.; Schmidt, A.; Schaub, T., Hydrothermal liquefaction oil and hydrotreated product from pine feedstock characterized by heteronuclear twodimensional NMR spectroscopy and FT-ICR mass spectrometry. Fuel 2014, 137, 60-69; DOI 10.1016/j.fuel.2014.07.069. 52. Meckenstock, R. U.; von Netzer, F.; Stumpp, C.; Lueders, T.; Himmelberg, A. M.; Hertkorn, N.; Schmitt-Kopplin, P.; Harir, M.; Hosein, R.; Haque, S.; Schulze-Makuch, D., Water droplets in oil are microhabitats for microbial life. Science 2014, 345 (6197), 673-676; DOI 10.1126/science.1252215. 53. Kim, D.; Jin, J. M.; Choa, Y.; Kim, E.-H.; Cheong, H.-K.; Kim, Y. H.; Kim, S., Combination of ring type HPLC separation, ultrahigh-resolution mass spectrometry, and high field NMR for comprehensive characterization of crude oil compositions. Fuel 2015, 157, 48–55; DOI 10.1016/j.fuel.2015.04.061. 54. Stookey, L. L., Ferrozine - a new spectrophotometric reagent for iron. Anal. Chem. 1970, 42 (7), 779–781; DOI 10.1021/ac60289a016. 55. Cline, J. D., Spectrometric Determination of Hydrogen Sulfide in Natural Waters. Limnol. Oceanogr. 1969, 14 (3), 454-458; DOI 10.4319/lo.1969.14.3.0454. 56. Dittmar, T.; Koch, B.; Hertkorn, N.; Kattner, G., A simple and efficient method for the solidphase extraction of dissolved organic matter (SPE-DOM) from seawater. Limnology and Oceanography: Methods 2008, 6, 230-235; DOI 10.4319/lom.2008.6.230. 57. Anneser, B.; Richters, L.; Griebler, C. Application of High-Resolution Groundwater Sampling in a Tar Oil-Contaminated Sandy Aquifer, Studies on Small-Scale Abiotic Gradients. In Advances in Subsurface Pollution of Porous Media - Indicators, Processes and Modelling: IAH selected papers, volume 14; Candelad, J. L.; Vadillo, I.; Elorza, F. J., Eds.; CRC Press/Balkema: Leiden 2008; pp 107-122. 58. Winderl, C.; Anneser, B.; Griebler, C.; Meckenstock, R. U.; Lueders, T., Depth-Resolved Quantification of Anaerobic Toluene Degraders and Aquifer Microbial Community Patterns in Distinct Redox Zones of a Tar Oil Contaminant Plume. Appl. Environ. Microbiol. 2008, 74 (3), 792801; DOI 10.1128/aem.01951-07. 59. Casagrande, D. J.; Idowu, G.; Friedman, A.; Rickert, P.; Siefert, K.; Schlenz, D., H2S incorporation in coal precursors: origins of organic sulphur in coal. Nature 1979, 282 (5739), 599600; DOI 10.1038/282599a0. 60. Casagrande, D. J.; Gronli, K.; Sutton, N., The distribution of sulfur and organic matter in various fractions of peat: origins of sulfur in coal. Geochim. Cosmochim. Acta 1980, 44 (1), 25-32; DOI 10.1016/0016-7037(80)90174-X. 61. Henneke, E.; Luther III, G. W.; De Lange, G. J.; Hoefs, J., Sulphur speciation in anoxic hypersaline sediments from the eastern Mediterranean Sea. Geochim. Cosmochim. Acta 1997, 61 (2), 307-321; DOI 10.1016/S0016-7037(96)00355-9. 62. Heitmann, T.; Goldhammer, T.; Beer, J.; Blodau, C., Electron transfer of dissolved organic matter and its potential significance for anaerobic respiration in a northern bog. Global Change Biology 2007, 13 (8), 1771-1785; DOI 10.1111/j.1365-2486.2007.01382.x.

Page 27 of 28

ACS Paragon Plus Environment

Environmental Science & Technology

672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722

Page 28 of 28

63. Vairavamurthy, A.; Mopper, K., Geochemical formation of organosulphur compounds (thiols) by addition of H2S to sedimentary organic matter. Nature 1987, 329 (6140), 623-625; DOI 10.1038/329623a0. 64. Perlinger, J. A.; Kalluri, V. M.; Venkatapathy, R.; Angst, W., Addition of Hydrogen Sulfide to Juglone. Environ. Sci. Techol. 2002, 36 (12), 2663-2669; DOI 10.1021/es015602c. 65. Heitmann, T.; Blodau, C., Oxidation and incorporation of hydrogen sulfide by dissolved organic matter. Chemical Geology 2006, 235 (1-2), 12-20; DOI 10.1016/j.chemgeo.2006.05.011. 66. Yu, Z. G.; Peiffer, S.; Gottlicher, J.; Knorr, K. H., Electron transfer budgets and kinetics of abiotic oxidation and incorporation of aqueous sulfide by dissolved organic matter. Environ Sci Technol 2015, 49 (9), 5441-9; DOI 10.1021/es505531u. 67. Einsiedl, F.; Pilloni, G.; Ruth-Anneser, B.; Lueders, T.; Griebler, C., Spatial distributions of sulphur species and sulphate-reducing bacteria provide insights into sulphur redox cycling and biodegradation hot-spots in a hydrocarbon-contaminated aquifer. Geochim. Cosmochim. Acta 2015, 156, 207–221; DOI 10.1016/j.gca.2015.01.014. 68. Longnecker, K.; Kujawinski, E. B., Composition of dissolved organic matter in groundwater. Geochim. Cosmochim. Acta 2011, 75 (10), 2752-2761; DOI 10.1016/j.gca.2011.02.020. 69. Einsiedl, F.; Hertkorn, N.; Wolf, M.; Frommberger, M.; Schmitt-Kopplin, P.; Koch, B. P., Rapid biotic molecular transformation of fulvic acids in a karst aquifer. Geochim. Cosmochim. Acta 2007, 71 (22), 5474-5482; DOI 10.1016/j.gca.2007.09.024. 70. Griebler, C.; Safinowski, M.; Vieth, A.; Richow, H.; Meckenstock, R. U., Combined Application of Stable Carbon Isotope Analysis and Specific Metabolites Determination for Assessing In Situ Degradation of Aromatic Hydrocarbons in a Tar Oil-Contaminated Aquifer. Environ. Ecol. Stat. 2004, 38 (2), 617–631; DOI 10.1021/es0344516. 71. Annweiler, E.; Michaelis, W.; Meckenstock, R. U., Anaerobic cometabolic conversion of benzothiophene by a sulfate-reducing enrichment culture and in a tar-oil-contaminated aquifer. Appl. Environ. Microbiol. 2001, 67 (11), 5077-5083; DOI 10.1128/AEM.67.11.5077-5083.2001. 72. Jarling, R.; Kuhner, S.; Basilio Janke, E.; Gruner, A.; Drozdowska, M.; Golding, B. T.; Rabus, R.; Wilkes, H., Versatile transformations of hydrocarbons in anaerobic bacteria: substrate ranges and regio- and stereo-chemistry of activation reactions. Front. Microbiol. 2015, 6 (880), 1-14; DOI 10.3389/fmicb.2015.00880. 73. Safinowski, M.; Griebler, C.; Meckenstock, R. U., Anaerobic Cometabolic Transformation of Polycyclic and Heterocyclic Aromatic Hydrocarbons: Evidence from Laboratory and Field Studies. Environ. Sci. Techol. 2006, 40 (13), 4165–4173; DOI 10.1021/es0525410. 74. Martus, P.; Puttmann, W., Formation of alkylated aromatic acids in groundwater by anaerobic degradation of alkylbenzenes. Sci. Total Environ. 2003, 307 (1-3), 19-33; DOI 10.1016/s00489697(02)00512-0. 75. Coble, P. G., Characterization of marine and terrestrial DOM in seawater using excitationemission matrix spectroscopy Mar. Chem. 1996, 51 (4), 325-346; DOI 10.1016/0304-4203(95)000623. 76. Ishii, S. K. L.; Boyer, T. H., Behavior of Reoccurring PARAFAC Components in Fluorescent Dissolved Organic Matter in Natural and Engineered Systems: A Critical Review. Environ. Sci. Techol. 2012, 46 (4), 2006–2017; DOI 10.1021/es2043504. 77. Yang, L.; Hur, J.; Zhuang, W., Occurrence and behaviors of fluorescence EEM-PARAFAC components in drinking water and wastewater treatment systems and their applications: a review. Environmental Science and Pollution Research 2015, 22 (9), 6500-6510; DOI 10.1007/s11356-0154214-3. 78. Wang, Z.; Cao, J.; Meng, F., Interactions between protein-like and humic-like components in dissolved organic matter revealed by fluorescence quenching. Water Res. 2015, 68, 404-413; DOI 10.1016/j.watres.2014.10.024. 79. Muyzer, G.; Stams, A. J., The ecology and biotechnology of sulphate-reducing bacteria. Nature Rev. Microbiol. 2008, 6, 441-454; DOI 10.1038/nrmicro1892. Page 28 of 28

ACS Paragon Plus Environment