Comparison of the Chemical Composition of Dissolved Organic Matter

Jan 24, 2018 - (21) Although the general chemistry of DOM in these two contrasting lakes has been reported in terms of DOC concentration, ultraviolet ...
0 downloads 6 Views 2MB Size
Subscriber access provided by READING UNIV

Article

Comparison of the chemical composition of dissolved organic matter in three lakes in Minnesota, USA Xiaoyan Cao, George R. Aiken, Kenna Butler, Jingdong Mao, and Klaus Schmidt-Rohr Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b04076 • Publication Date (Web): 24 Jan 2018 Downloaded from http://pubs.acs.org on January 25, 2018

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

Comparison of the chemical composition of dissolved organic matter in three lakes in

2

Minnesota, USA

3

Xiaoyan Cao1,2, George R. Aiken3,+, Kenna D. Butler3, Jingdong Mao1,*, Klaus Schmidt-Rohr2

4 5

1

6

Norfolk, Virginia 23529, USA

7

2

8

02453, USA

9

3

Department of Chemistry and Biochemistry, Old Dominion University, 4541 Hampton Blvd,

Department of Chemistry, Brandeis University, 415 South Street, Waltham, Massachusetts

United States Geological Survey, 3215 Marine Street, Boulder, Colorado 80303, USA

10 11 12

Submitted to Special issue of Environmental Science & Technology honoring the

13

contributions of George Aiken

14 15 16

*Corresponding author:

17

Jingdong Mao, e-mail: [email protected]; phone: 757-683-6874; fax: 757-683-4628

18

+

Deceased.

19

1 ACS Paragon Plus Environment

Environmental Science & Technology

20

Abstract

21

New information on the chemical composition of dissolved organic matter (DOM) in three lakes

22

in Minnesota has been gained from spectral editing and two-dimensional nuclear magnetic

23

resonance (NMR) methods, indicating the effects of lake hydrological settings on DOM

24

composition. Williams Lake (WL), Shingobee Lake (SL) and Manganika Lake (ML) have

25

different source inputs, and the lake water residence time (WRT) of WL is markedly longer than

26

that of SL and ML. The hydrophobic organic acid (HPOA) and transphilic organic acid (TPIA)

27

fractions combined composed > 50% of total DOM in these lakes, and contained carboxyl-rich

28

alicyclic molecules (CRAM), aromatics, carbohydrates, and N-containing compounds. The

29

previously understudied TPIA fractions contained fewer aromatics, more O-rich CRAM, and

30

more N-containing compounds compared to the corresponding HPOA. CRAM represented the

31

predominant component in DOM from all lakes studied, and more so in WL than in SL and ML.

32

Aromatics including lignin residues and phenols decreased in relative abundances from ML to SL

33

and WL. Carbohydrates and N-containing compounds were minor components in both HPOA

34

and TPIA and did not show large variations among the three lakes. The increased relative

35

abundances of CRAM in DOM from ML, SL to WL suggested the selective preservation of

36

CRAM with increased residence time.

37 38

2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Environmental Science & Technology

39

1. INTRODUCTION

40

Substantial evidence has accumulated that lakes are hotspots of carbon cycling even though they

41

comprise only a small fraction of the Earth’s surface.1-3 Dissolved organic matter (DOM) is the

42

largest pool of organic carbon (OC) in most lake water, and contains OC both exported from

43

land (allochthonous) and fixed by indigenous primary production (autochthonous). With the

44

exception of eutrophic lakes, lake DOM is strongly dominated by allochthonous material

45

imported from the catchment.4 A large share of lake DOM is altered and lost by in-lake

46

processes including microbial respiration/mineralization, photochemical degradation, and

47

flocculation.5-7 The magnitude of these in-lake processes appears to depend heavily on water

48

residence time (WRT), i.e., the amount of time that water spends in the lake.8-12 For instance,

49

Algesten et al.12 found that OC (mostly DOC) loss can be readily predicted from logWRT and

50

increases rapidly with increasing WRT up to 2–3 years. Catalán et al.13 reported a negative

51

relationship between the decay rate of OC in inland waters and WRT. Therefore, the molecular

52

composition of lake DOM can be highly variable across landscapes, hydrologies, and climates,

53

reflecting the net influences of DOM source, reactivity, and all the transformation processes

54

occurring within the lake system.8, 9, 14-16

55

The Shingobee River headwaters area, located in north-central Minnesota, provides a unique

56

opportunity for comparing and contrasting two lakes (Williams Lake and Shingobee Lake) with

57

similar biologic, geologic, and climatic settings, but different hydrologies.17 The hydrologically

58

closed Williams Lake has no surface inlet or outlet. Groundwater inflow represents 58-76% of

59

the annual water input with the rest being from precipitation.18 The hydrologically open

60

Shingobee Lake, about 5 kilometers from Williams Lake, has the Shingobee River flowing

3 ACS Paragon Plus Environment

Environmental Science & Technology

61

through it and dominating the annual water flux to and from the lake. As a result, Shingobee

62

Lake has a ten-fold shorter WRT (0.3-0.5 years) than Williams Lake (3-4 years).19

63

A previous 13C nuclear magnetic resonance (NMR) study14 has shown that Williams Lake

64

fulvic acid (FA; an operationally defined fraction of DOM isolated by XAD adsorption20) is

65

more aliphatic and less aromatic than Shingobee Lake FA. The fluorescence characteristics of

66

these FA isolates fell between the microbially-derived and terrestrially-derived end members,

67

suggesting DOM sources from both microbially and terrestrially derived organic material.21

68

Although the general chemistry of DOM in these two contrasting lakes has been reported in

69

terms of DOC concentration, ultraviolet (UV) and fluorescence characteristics, and carbon

70

functional group composition,14, 21 more detailed molecular-level structural information is

71

lacking. Advanced NMR techniques, such as spectral editing and two-dimensional correlation,

72

yield greater structural information beyond that which can be obtained from simply integrating

73

the broad and overlapping NMR resonances. In addition, previous work has focused only on the

74

relatively hydrophobic FA fraction of DOM (30-40% of total DOC),14, 21 whereas the more

75

hydrophilic fraction (termed transphilic acid, TPIA), which represents an important piece of the

76

DOC pie (~20% of total DOC), has received little attention. To extend this comparison, DOM

77

samples from Manganika Lake, a hypereutrophic lake in northeastern Minnesota with

78

anthropogenic influences from two major water inputs (the Virginia wastewater treatment plant

79

and United Taconite mine waters) were also included. The residence time of water in Manganika

80

Lake is estimated to be somewhat shorter than that of water in Shingobee Lake.

81

The present study aimed to (1) provide detailed chemical-composition characterization of

82

both hydrophobic acid and transphilic acid fractions of DOM from lakes with distinct

83

hydrological conditions, using one- and two-dimensional solid-state NMR spectroscopy; and to 4 ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Environmental Science & Technology

84

(2) examine a previous conclusion8 that terrestrially derived DOM is selectively lost as residence

85

time increases.

86 87

2. MATERIALS AND METHODS

88

2.1. Water Sampling and DOM Isolation. Sampling was conducted in June of 2012 in

89

Manganika Lake (ML) in northeastern Minnesota, and in September of 2013 in Williams Lake

90

(WL) and Shingobee Lake (SL) in north-central Minnesota. More detailed site descriptions can be

91

found elsewhere.22, 23 Large volume (155-415 L) water samples were filtered (0.45 µm) in the

92

field and shipped on ice to the U.S. Geological Survey (USGS) laboratory in Boulder (Colorado,

93

United States) for DOC and UV absorbance analyses.

94

Water samples were then processed by XAD isolation and concentration as described by

95

Aiken et al.24 Briefly, samples were acidified to pH 2 with hydrochloric acid (HCl) and passed

96

first through a column of XAD-8 resin, followed by a column of XAD-4 resin. Each column was

97

then eluted with 0.1 M sodium hydroxide (NaOH) to obtain the XAD-8 (hydrophobic organic

98

acids, HPOA) and XAD-4 (transphilic organic acids, TPIA) fractions. The eluates were

99

immediately acidified to minimize sample alteration at high pH, desalted, lyophilized, and stored

100

at room temperature.

101 102

2.2. Elemental, UV-Visible Absorbance, and Carbon Isotopic Measurements. Elemental

103

analyses (C, H, O, N, S and ash) of DOM isolates were performed by Huffman Laboratories

104

(Golden, Colorado) using the method described by Huffman and Stuber.25 Specific UV

105

absorbance (SUVA254) was determined by dividing the UV-visible absorbance at λ = 254 nm by

106

DOC concentration, and is correlated to DOM aromaticity.26 Stable carbon isotope ratios (δ13C) 5 ACS Paragon Plus Environment

Environmental Science & Technology

107

were determined by isotope ratio mass spectrometry on dried DOM fractions following vapor

108

phase acidification, and are expressed relative to the Pee Dee Belemnite (PDB) standard. The

109

radiocarbon ratios of the HPOA isolates were measured by accelerator mass spectrometry at the

110

Center for Accelerator Mass Spectrometry at Lawrence Livermore National Laboratory

111

(California, United States). ∆14C data (in ‰) were corrected for isotopic fraction using measured

112

δ13C values. The ∆14C and radiocarbon age were determined from percent modern carbon using

113

the year of sample analysis according to Stuiver and Polach.27 Ages were presented as Modern

114

when the fraction modern exceeded 1.

115 116

2.3. NMR Analysis. All NMR experiments were performed at 100 MHz for 13C and 400

117

MHz for 1H using a Bruker Avance 400 spectrometer equipped with a 4-mm double-resonance

118

probe head. The 13C chemical shifts were referenced to tetramethylsilane, using the COO

119

resonance of glycine in the α-modification at 176.49 ppm as a secondary reference. NMR

120

experiments included quantitative multiple-cross polarization (multiCP),28 multiCP plus dipolar

121

dephasing, cross-polarization with total suppression of sidebands (CP/TOSS), 13C chemical-

122

shift-anisotropy (CSA) filter, 13C CSA filter plus dipolar dephasing, 1H–13C 2D heteronuclear

123

correlation (HETCOR), 2D HECTOR with dipolar dephasing, and 2D HETCOR with 1H spin

124

diffusion. Experimental details are described in the Supporting Information (SI) and a summary

125

of these NMR methods and their purposes is provided in Table S1. The uncertainties of the NMR

126

integrals (Table S2) were reported based on the propagation of analytical uncertainty, and

127

estimated from the signal-to-noise ratio.

128 129

3. RESULTS AND DISCUSSION 6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

130

Environmental Science & Technology

3.1. Basic Chemical Properties. Table S3 lists DOC concentrations, SUVA254, elemental and

131

carbon isotopic data of DOM from three lakes. DOC concentrations were greater in WL (7.2 mg

132

C L-1) than in SL (5.2 mg C L-1), but lower than in ML (10.5 mg C L-1). The hydrophobic acids

133

(HPOA) accounted for about 30-40% of DOC in all lakes, and were less abundant in WL (30% of

134

DOC) than in SL (37%) and ML (40%). The fraction of transphilic acids (TPIA) was higher in

135

WL (21% of DOC) and SL (22%) than in ML (17% of DOC). These observations are consistent

136

with McElmurry et al.,29 who suggested that hydrologic “short-circuiting” may influence the

137

hydrophobic/hydrophilic fractions of DOM. The SUVA254 (indicating aromaticity) associated

138

with HPOA and TPIA isolates showed the same trend, increasing in values from WL to SL to ML;

139

data generated during this study are available at https://doi.org/10.5066/F77M06VP.30 This result

140

is consistent with the previously reported inverse relationship between DOC color (measured as

141

absorbance) and WRT.31 The TPIA isolate always had a lower SUVA254 value than the

142

corresponding HPOA from the same lake. Elemental analyses showed that N contents were low

143

(< 2%) in all HPOA isolates, resulting in atomic C/N values of 36-37, but relatively higher (2-

144

3%) in the TPIA fractions leading to atomic C/N values of 18-22. The δ13C values were less

145

negative for WL HPOA (-26.1‰) than for SL HPOA (-29.1‰), but both fell within the range

146

reported for DOC in lakes located in Sweden32 and Northern U.S.33, 34 The SL HPOA had more

147

depleted ∆14C values than WL HPOA, which translated to 14C ages of 315 ybp for SL HPOA and

148

modern for WL HPOA.

149 150

3.2. Specific Functional Group Compositions of DOM in Different Lakes. A previous 13C

151

NMR analysis of DOM in the Williams and Shingobee lake systems14 identified six types of

152

functional groups, including aliphatic I carbon (0-62 ppm), aliphatic II carbon (62-90 ppm), 7 ACS Paragon Plus Environment

Environmental Science & Technology

153

acetal carbon (90-110 ppm), aromatic carbon (110-160 ppm), carboxyl carbon (160-190 ppm),

154

and ketone C (190-230 ppm). The current study applied spectral-editing techniques to identify

155

more specific functional groups and make more accurate assignments. For instance, multiCP

156

spectra obtained after dipolar dephasing (Figure 1, red lines), showed signals of nonprotonated

157

and mobile carbons, such as CH3, quaternary C (Cq), nonprotonated OC (OCnp), ketal C (OCnpO),

158

nonprotonated aromatic C-C, aromatic C-O (144-160 ppm), COO/NC=O, and ketone C.

159

Although both aldehyde and ketone C resonate in the 190-220 ppm region, only the signals of

160

nonprotonated ketone C can survive dipolar dephasing. These dipolar-dephased spectra therefore

161

provided supporting evidence for the previous assignment of the 190-230 ppm region to

162

ketones;14 the nearly unchanged intensity of the ketone peak (in red lines) relative to that of the

163

aldehyde/ketone peak (in black lines) indicated that the carbonyls were present as ketones, not

164

aldehydes in all samples. In addition, there was a substantial peak at ~56 ppm in the dipolar-

165

dephased spectrum of ML HPOA (Figure 1(c), red line), characteristic of methoxyl groups in

166

lignin residues, but this sharp peak was missing in the spectra of all other samples. Moreover,

167

signals of Cq and nonprotonated OC, indicative of highly branched structures, were prominent in

168

the spectra of all lake isolates (Figure 1, red lines). Notably, the CSA-filtered spectra (Figure 1,

169

bold black lines) resolved signals of di-oxygenated aliphatic carbons (O–C–O, see black arrows),

170

which span the range of 100-123 ppm. The spectra obtained from the CSA filter with dipolar

171

dephasing (Figure 1, blue lines) further indicated that nonprotonated di-oxygenated alkyl (OCnpO,

172

see black arrows), i.e., ketal carbons, largely contributed to the NMR signals in the 100-123 ppm

173

region. Therefore the previous assignment of 90-110 ppm only to acetal carbons (O–CH–O) was

174

not appropriate.14

8 ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

175

Environmental Science & Technology

Table 1 shows the distribution of the 13 specific functional groups in all HPOA/TPIA samples.

176

The alkyl C (0-64 ppm) accounted for 34-47% and 33-41% of the C in the HPOA and TPIA

177

isolates, respectively, and decreased in the order WL > SL > ML. The total aromatic C fraction

178

(aromaticity; including protonated aromatic C-H, nonprotonated aromatic C-C, and oxygen

179

substituted aromatic C-O) ranged from 12-27% and 8-16% for HPOA and TPIA isolates,

180

respectively, and aromatic C-C dominated the aromatic C pool. The fractions of these three types

181

of aromatic C generally displayed the same trend as aromaticity: WL < SL < ML. The abundances

182

of O-alkyl C (sum of OC and OCO) remained rather constant among three lakes, accounting for

183

19-20% of HPOA and 27-28% of TPIA. There were relatively more protonated OC (OCHn) than

184

nonprotonated OC (OCnp), but fewer protonated OCO (OCHO, ~1%) than nonprotonated OCO

185

(OCnpO, 2-3%). This suggested that these O-alkyl moieties were unlikely to be present in pure

186

carbohydrate environments, where protonated OCH and OCHO are predominant. The

187

COO/NC=O constituted 17-18% of HPOA and 20-21% of TPIA, with NC=O being at most 2-

188

3% for HPOA and 4-5% for TPIA. Overall, HPOA isolates were more enriched in aromatics but

189

depleted in O-alkyl carbons than corresponding TPIA isolates. This is consistent with the lower

190

H/C and O/C atomic ratios of HPOA relative to the respective TPIA isolates, and with NMR

191

measurements on HPOA/TPIA isolates from rivers and groundwater.24 The higher aromatic

192

signature of HPOA isolates also agreed with previous findings that ~90% of the lignin phenols

193

were recovered in the HPOA fractions.35, 36 Notably, all HPOA and TPIA samples contained

194

abundant nonprotonated aliphatic C (Cq + OCnp + OCnpO) (11-13% for HPOA and 15-16% for

195

TPIA) as well as CH3 (9-13%), indicating the presence of highly branched or cyclized aliphatic

196

structures.

197

9 ACS Paragon Plus Environment

Environmental Science & Technology

198

3.3. Chemical Structures of Functional Groups in DOM from Different Lakes. Two-

199

dimensional 1H-13C HETCOR NMR probes through-space 1H-13C correlations, identifies the

200

chemical structure corresponding to a given carbon peak, and thus provides more information

201

than 13C or 1H NMR alone.37 Moreover, the correlation of nonprotonated carbons with nearby

202

(non-bonded) protons becomes possible, which enables characterization of the environment of

203

nonprotonated groups. Particular emphasis is given to the chemical environment of O-alkyl C,

204

COO, and abundant nonprotonated aliphatic carbons.

205

Figure 2 presents the 2D HETCOR spectra of HPOA isolates from WL, SL and ML. For WL

206

and SL, the 1H spectra associated with OC carbons contained signals from both O-alkyl protons

207

and alkyl protons of similar intensities (Figure 2(d, e)). The OCO carbons showed cross peaks

208

primarily with alkyl protons, indicating that both OC and OCO carbons were in close proximity

209

to alkyl protons, and therefore, alkyl carbons. For ML HPOA (Figure 2(f)), both OC and OCO

210

carbons correlated mainly with O-alkyl protons, but also more weakly with alkyl protons. The 1H

211

spectra extracted at aromatic C (131 ppm) showed signals from both aromatic and alkyl protons

212

(Figure 2(d-f)), indicating close association of aromatic and alkyl components. But the relative

213

intensities of aromatic protons increased relative to those of alkyl protons from WL HPOA to SL

214

HPOA and WL HPOA (Figure 2(d-f)). For all HPOA isolates, quaternary carbons (Cq) showed

215

cross peaks mainly to alkyl protons (~2 ppm) (Figure 3(a-c)). The 1H spectra associated with

216

OCnp (86 ppm) showed signals mainly from alkyl protons, but contributions from O-alkyl

217

protons were also observed, and more pronounced for ML HPOA (Figure 3(c)) than WL and SL

218

HPOA (Figure 3(a, b)). The ketal OCnpO (108 ppm) carbons appeared to associate primarily with

219

alkyl protons for WL and SL HPOA. For ML HPOA, there were also contributions from O-alkyl

220

protons and aromatic protons. These results suggested that nonprotonated OC and ketal carbons 10 ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

Environmental Science & Technology

221

did not occur in carbohydrates where they would be surrounded by O-alkyl protons. The

222

COO/NC=O carbons showed cross peaks primarily with alkyl protons near 2-3 ppm, with

223

additional contribution from O-alkyl protons, indicating that they were attached mainly to alkyl

224

and O-alkyl carbons, consistent with structures of carboxyl-rich alicyclic molecules (CRAM).

225

The prominent cross peaks of COO carbons with the acidic COOH protons (~12 ppm) (Figure

226

S1(a-c)) confirmed the presence of carboxylic acids rather than esters. Ketone C mainly

227

correlated with aliphatic protons, indicating that they were bonded to aliphatic carbons. The 2D

228

spectrum of ML HPOA also showed two distinct cross peaks of OCH3 (Figure S1(c)), red boxes):

229

one of OCH3 carbon with its directly bonded protons, and the other of OCH3 carbon with

230

aromatic protons. They confirmed the presence of lignin residues in ML HPOA.

231

Unlike those of HPOA isolates, the 2D HETCOR spectra of TPIA isolates (Figure S2(a-c))

232

showed pronounced cross peaks of OCO (~100 ppm) and OC (~72 ppm) with their directly

233

attached protons, indicating the presence of sugar rings. The correlation of OC/OCO carbons

234

with alkyl protons were much weaker for the TPIA (Figure S2(d-f)) than for the corresponding

235

HPOA. With a 1-ms 1H spin diffusion time, the 1H spectra associated with OC carbons showed

236

increasing signals from alkyl protons while those associated with alkyl C showed increasing

237

contributions from O-alkyl protons (Figure S3(d-f)), compared to 1H spectra obtained without

238

mixing time (Figure S2(d-f)). These trends indicated that alkyl and O-alkyl components were in

239

close proximity and that sugar-ring structures did not form large carbohydrate domains in TPIA.

240

There seemed to be two overlapping cross peaks of quaternary C near 45-60 ppm: one of Cq

241

carbons with alkyl protons as similarly observed in spectra of HPOA samples (Figure S1), and

242

the other of Cq carbons with OCH/NCH protons (Figure 4). The 1H slices associated with OCnp

243

(~86 ppm) all showed major contributions from alkyl protons, but proximity to O-alkyl protons 11 ACS Paragon Plus Environment

Environmental Science & Technology

244

was also observed. The OCnpO carbons appeared to correlate primarily with alkyl protons in

245

TPIA isolates from WL and SL (Figure S4(a,b)). For ML TPIA, the contribution from O-alkyl

246

protons was equally, if not more important (Figure S4(c)). The 1H slices extracted at

247

COO/NC=O (Figure S4) all showed a major band centered at 2-3 ppm and correlations with

248

O/N-alkyl protons, suggesting that COO/NC=O groups are mostly attached to aliphatic carbons.

249

The 1H slices extracted at ketone C for all TPIA isolates demonstrated that these carbons were

250

attached to alkyl protons.

251 252

3.4. New Structural Information on DOM in Lakes from Spectral Editing and 2D NMR.

253

The main conclusion from previous NMR analysis of the Shingobee watershed is that “FA in WL

254

is more aliphatic (0-62 ppm) and less aromatic (110-160 ppm) than FA samples from Shingobee

255

Lake”.14 With spectral editing and two-dimensional NMR, the present study obtained new

256

information and more structural details such as the makeup of aliphatic moieties. The structural

257

components present in two major DOM fractions (HPOA and TPIA), which composed > 50% of

258

total DOM in WL, SL and ML, included CRAM, aromatics, carbohydrates, and N-containing

259

compounds. Their relative carbon percentages in HPOA/TPIA isolates were estimated following

260

Cao et al.38 and are shown in Figure 5.

261

Carbohydrates were a very minor component, contributing to ~ 5% in HPOA and 8% or less

262

in TPIA isolates, respectively. This may seem surprising given that O-alkyl carbons (OC and

263

OCO) accounted for ~20% of HPOA and ~28% of TPIA isolates, which would be traditionally

264

attributed to carbohydrates. However, spectral editing and 2D NMR data showed strong evidence

265

that approximately 1/3 to 1/2 of O-alkyl sites occurred as nonprotonated carbons close to alkyl

266

protons (Figures 1, 3 and 4), indicating that they were unlikely present in carbohydrate 12 ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28

Environmental Science & Technology

267

environments. Even some protonated O-alkyl C (OCHn + OCHO) may not be completely

268

associated with carbohydrates because they were also in close proximity to alkyl protons (Table

269

1, Figures S3 and S5), which would not be expected in typical sugar environments. These results

270

therefore caution against the routine assignment of 13C NMR resonances in 62-110 ppm to

271

carbohydrates for DOM samples,39 in particular those isolated with solid-phase (e.g. XAD)

272

extraction, which are known not to retain large carbohydrates.38

273

CRAM represented the predominant structural component in both HPOA (61-76%) and TPIA

274

(67-74%) isolates, but CRAM structures in TPIA are more oxygen-rich than CRAM in HPOA. A

275

striking finding from our 2D NMR data is that abundant nonprotonated aliphatics (OCnp, OCnpO,

276

and Cq, 11–13% of HPOA and 15–16% of TPIA) and isolated O–CH carbons were associated

277

with CRAM structures in DOM from the three lakes. This characteristic of CRAM is also

278

documented for DOM collected from the Yukon River, but has not been realized in other studies,

279

where the analytical techniques employed were not capable of identifying these structural

280

moieties.40-43

281

Aromatics increased in abundance for both HPOA and TPIA from WL to SL and to ML

282

(Figure 5), and their carbon percentages were linearly correlated to SUVA254 (r2 = 0.95, Figure

283

S6). Characteristic peaks associated with lignin were evident only in ML HPOA. Furthermore,

284

aromatic C abundances were lower in TPIA than in HPOA isolates. Based on C/N atomic ratios

285

and the assumption that each N atom was attached to two C atoms, N-containing compounds (i.e.,

286

carbons bonded to nitrogen) can account for up to 6% of HPOA isolates and 11% of TPIA

287

isolates. There was little variation among the three lakes, but TPIA isolates contained more N

288

than corresponding HPOA isolates.

13 ACS Paragon Plus Environment

Environmental Science & Technology

289

The more hydrophilic fraction, TPIA, which represents an important slice of the DOM pie

290

(~20% of total DOC), has received little attention. Major differences between HPOA and TPIA

291

fractions are that TPIA contains fewer aromatics, but more O-containing functional groups such

292

as COO and O-alkyl moieties.24 More specifically from this study, these O-containing functional

293

groups were mainly associated with CRAM structures, making CRAM in TPIA more oxygen-

294

rich than CRAM in HPOA. Another difference was that N-containing compounds were more

295

abundant in TPIA than in HPOA samples. The relatively higher aromatic content in HPOA than

296

in TPIA fractions has been related to the 2−3 fold higher mercury methylation of the HPOA

297

fraction compared to the TPIA fraction.44 On the other hand, molecules in the TPIA fractions,

298

with their greater heteroatom and carboxyl contents, may exhibit considerable geochemical

299

significance in processes such as mineral weathering and water acidification.

300 301

3.5. Effects of Hydrology on DOM Structure in Lakes. Recent studies have explored the

302

chemical composition of DOM from hundreds of Swedish lakes across a wide range of land-use,

303

hydrology, and climate gradients, using fluorescence spectroscopy9 and mass spectrometry.8, 45

304

The advanced NMR methods used in the present study can capture aspects of DOM composition

305

that are missed by UV-Vis, fluorescence, or mass spectrometry, and they offer an important

306

addition to current knowledge on DOM chemistry in lakes driven by hydrology. This work

307

studied DOM from lakes in a small headwaters watershed in Minnesota on more localized scales,

308

and allowed us to decipher the importance of lake hydrology in driving DOM chemistry.

309

Lake hydrological settings influence the DOM source inputs, terrestrial/watershed processes

310

that occur during DOM transport, and its in-lake processing,14, 29 which leads to the observed

311

differences in DOM structure among the three lakes. DOM aromaticity decreased from 14 ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Environmental Science & Technology

312

Manganika Lake to Shingobee Lake and to Williams Lake. DOM from Manganika Lake was

313

most enriched in aromatics, which were at least partially derived from lignin residues based on

314

1D and 2D NMR data (Figures 1 and S1), while characteristic signals of lignin were not detected

315

in Williams and Shingobee lakes. This agrees with the previous claim8 that terrestrially-derived

316

polyphenols were selectively lost as WRT increases. Yet WRT may not be the decisive factor in

317

determining the relative contribution of terrestrially-derived DOM. Although longer WRT may

318

facilitate the accumulation of autochthonous DOM9, 10 and in-lake photochemical processing of

319

terrestrially-derived aromatics,14 the influence of source materials cannot be neglected.14

320

Shingobee Lake receives terrestrial inputs from the Shingobee River, which supply more

321

aromatic DOM than groundwater entering Williams Lake.14 Manganika Lake receives yet more

322

terrestrial organic matter, from the Virginia wastewater treatment plant and United Taconite

323

mine waters. Therefore the lowest abundance of aromatics in Williams Lake among the three

324

lakes may also be attributed to the source input; WRT may be a correlate, and not fundamentally

325

associated with the specific process (e.g. photochemical transformation) responsible for the

326

variations in DOM observed. Furthermore, the influence of terrestrial processes that occur during

327

DOM transport can be also important. For instance, the absence of lignin observed in WL and SL

328

may also result from the removal of hydrophobic fractions during terrestrial/watershed processes

329

(rather than in-lake/autochthonous processes). Meier et al.46 reported the removal of more

330

aromatic and more hydrophobic components in DOM by sorption to soil mineral phases. Yano et

331

al.47 suggested DOM removal in soil occurs mostly via abiotic sorption. Kawahigashi et al.48

332

found that sorptive interactions of DOM with the soil mineral phase generally increase with

333

depth, and thus the depth of the active layer likely controls the quantity and quality of DOM

334

exported to aquatic systems. 15 ACS Paragon Plus Environment

Environmental Science & Technology

335

In addition to supporting the previous conclusion that terrestrially-derived lignin is selectively

336

lost with increasing residence time,8 our study has also revealed DOM structures that may be

337

selectively preserved as residence time increases. Aliphatic components are most enriched in

338

DOM from Williams Lake, which could be attributed to the in-lake production of aliphatic

339

compounds,10 or an indirect consequence of the depletion of aromatic components. NMR data

340

clearly identified these aliphatic moieties to be mainly associated with CRAM. Note that neither

341

fluorescence nor mass spectrometry can differentiate CRAM from interfering molecules, such as

342

lignin, with similar elemental ratios.45 In addition to its photo-resistance, CRAM also contains

343

quaternary and nonprotonated O-alkyl carbons, which represent more humified components than

344

carbohydrates.49, 50 CRAM were a predominant component in DOM from all three lakes, and was

345

relatively more abundant in WL than in SL and ML (Figure 5). Therefore our results suggest the

346

selective preservation of CRAM with increasing water residence time.

347 348

3.6. Environmental Implications. Variations in the chemical composition of DOM in

349

aquatic systems, as observed from this study, are significant factors controlling DOM chemical

350

reactivity. For instance, the aromatic carbon content in DOM influences the strength of DOM

351

interactions with organic pollutants. Graham et al.44 reported that highly aromatic DOM strongly

352

enhanced mercury methylation, consistent with previous identification of environments with

353

high concentrations or fluxes of highly aromatic DOM (e.g., wetlands) as hot spots for

354

methylmercury production.51 Aromatic moieties are photoreactive and play mechanistic roles in

355

DOM photoreactions, including oxidation/reduction reactions that control the speciation and

356

chemistry of metals such as Fe and Hg.52 The globally ubiquitous, quantitatively important

357

component of DOM, CRAM, has been predicted to be microbially and photochemically 16 ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Environmental Science & Technology

358

refractory. This stable nature is further supported by our observation that CRAM are selectively

359

preserved with increased lake WRT, suggesting their critical role in the global carbon cycle.

360

Nevertheless, whether CRAM are closely tied to biogeochemical processes such as interactions

361

with inorganic and organic pollutants remains unknown and warrants further investigations.

362 363 364

Acknowledgments This work was funded by USDA-NIFA Capacity Building Grant Program (grant 2010-38821-

365

21558), National Aeronautics and Space Administration (grants NNX09AU89G and

366

NNH04AA62I), the National Science Foundation (grants CBET-0853950 and CBET-0853682)

367

and the U.S Geological Survey National Research Program. Any use of trade, firm, or product

368

names is for descriptive purposes only and does not imply endorsement by the U.S. Government.

369 370

Supporting Information

371

The Supporting Information is available free of charge on the ACS Publications website.

372

More NMR experimental details. Tables summarizing the NMR methods and their purposes, as

373

well as the uncertainties of the NMR integrals. A table containing DOC concentrations, SUVA254,

374

elemental and carbon isotopic data. Figures showing the 2D 1H-13C HETCOR NMR spectra and

375

the 1H slices extracted from the 2D spectra. The relationship between percent aromaticity

376

determined by 13C NMR and SUVA254.

377

17 ACS Paragon Plus Environment

Environmental Science & Technology

378

References

379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420

1. Sobek, S.; Söderbäck, B.; Karlsson, S.; Andersson, E.; Brunberg, A. K., A carbon budget of a small humic lake: an example of the importance of lakes for organic matter cycling in boreal catchments. AMBIO: A Journal of the Human Environment 2006, 35, (8), 469-475. 2. Cole, J. J.; Prairie, Y. T.; Caraco, N. F.; McDowell, W. H.; Tranvik, L. J.; Striegl, R. G.; Duarte, C. M.; Kortelainen, P.; Downing, J. A.; Middelburg, J. J., Plumbing the global carbon cycle: integrating inland waters into the terrestrial carbon budget. Ecosystems 2007, 10, (1), 172185. 3. Tranvik, L. J.; Downing, J. A.; Cotner, J. B.; Loiselle, S. A.; Striegl, R. G.; Ballatore, T. J.; Dillon, P.; Finlay, K.; Fortino, K.; Knoll, L. B., Lakes and reservoirs as regulators of carbon cycling and climate. Limnology and Oceanography 2009, 54, (6), 2298-2314. 4. Wilkinson, G. M.; Pace, M. L.; Cole, J. J., Terrestrial dominance of organic matter in north temperate lakes. Global Biogeochemical Cycles 2013, 27, (1), 1-9. 5. Bertilsson, S.; Tranvik, L. J., Photochemical transformation of dissolved organic matter in lakes. Limnology and Oceanography 2000, 45, (4), 753-762. 6. von Wachenfeldt, E.; Sobek, S.; Bastviken, D.; Tranvik, L. J., Linking allochthonous dissolved organic matter and boreal lake sediment carbon sequestration: the role of lightmediated flocculation. Limnology and Oceanography 2008, 53, (6), 2416. 7. Koehler, B.; von Wachenfeldt, E.; Kothawala, D.; Tranvik, L. J., Reactivity continuum of dissolved organic carbon decomposition in lake water. Journal of Geophysical Research: Biogeosciences 2012, 117, G01024. 8. Kellerman, A. M.; Dittmar, T.; Kothawala, D. N.; Tranvik, L. J., Chemodiversity of dissolved organic matter in lakes driven by climate and hydrology. Nature Communications 2014, 5, 3804. 9. Kothawala, D. N.; Stedmon, C. A.; Müller, R. A.; Weyhenmeyer, G. A.; Köhler, S. J.; Tranvik, L. J., Controls of dissolved organic matter quality: evidence from a large‐scale boreal lake survey. Global change biology 2014, 20, (4), 1101-1114. 10. Köhler, S. J.; Kothawala, D.; Futter, M. N.; Liungman, O.; Tranvik, L., In-lake processes offset increased terrestrial inputs of dissolved organic carbon and color to lakes. PloS one 2013, 8, (8), e70598. 11. Hanson, P. C.; Hamilton, D. P.; Stanley, E. H.; Preston, N.; Langman, O. C.; Kara, E. L., Fate of allochthonous dissolved organic carbon in lakes: a quantitative approach. PLoS One 2011, 6, (7), e21884. 12. Algesten, G.; Sobek, S.; Bergström, A. K.; Ågren, A.; Tranvik, L. J.; Jansson, M., Role of lakes for organic carbon cycling in the boreal zone. Global change biology 2004, 10, (1), 141147. 13. Catalán, N.; Marcé, R.; Kothawala, D. N.; Tranvik, L. J., Organic carbon decomposition rates controlled by water retention time across inland waters. Nature Geoscience 2016, 9, (7), 501-504. 14. Aiken, G. R.; McKnight, D. M.; Winter, T., The influence of hydrological factors on the nature of organic matter in the Williams and Shingobee Lake Systems. Interdisciplinary research initiative: hydrological and biogeochemical research in the Shingobee River headwaters area, north-central Minnesota. Document 1997, 96-4215.

18 ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464

Environmental Science & Technology

15. Cory, R. M.; McKnight, D. M.; Chin, Y. P.; Miller, P.; Jaros, C. L., Chemical characteristics of fulvic acids from Arctic surface waters: microbial contributions and photochemical transformations. Journal of Geophysical Research: Biogeosciences 2007, 112, (G4), G04S51. 16. Berggren, M.; Laudon, H.; Jansson, M., Landscape regulation of bacterial growth efficiency in boreal freshwaters. Global Biogeochemical Cycles 2007, 21, (4), GB4002. 17. Winter, T. C. Hydrological and biogeochemical research in the Shingobee River headwaters area, north-central Minnesota; US Dept. of the Interior, US Geological Survey: Information Services [distributor]: 1997. 18. LaBaugh, J. W.; Rosenberry, D. O.; Winter, T. C., Groundwater contribution to the water and chemical budgets of Williams Lake, Minnesota, 1980-1991. Canadian Journal of Fisheries and Aquatic Sciences 1995, 52, (4), 754-767. 19. Stets, E. G.; Striegl, R. G.; Aiken, G. R.; Rosenberry, D. O.; Winter, T. C., Hydrologic support of carbon dioxide flux revealed by whole‐lake carbon budgets. Journal of Geophysical Research: Biogeosciences 2009, 114, (G1), G01008. 20. Aiken, G. R., Isolation and concentration techniqes for aquatic humic substances. In Humic Substanceses in Soil, Sediment and Water: Geochemistry, Isolation and Characterization, Aiken G.R.; McKnight D.M.; Wershaw R.L.; P., M., Eds. John Wiley & Sons: New York, 1985; pp 363-385. 21. McKnight, D. M.; Boyer, E. W.; Westerhoff, P. K.; Doran, P. T.; Kulbe, T.; Andersen, D. T., Spectrofluorometric characterization of dissolved organic matter for indication of precursor organic material and aromaticity. Limnology and Oceanography 2001, 46, (1), 38-48. 22. Winter, T. C.; Rosenberry, D. O., Physiographic and geologic characteristics of the Shingobee River headwaters area. Interdisciplinary Research Initiative: Hydrological and Biogeochemical Research in the Shingobee River Headwaters Area, North-Central Minnesota: US Geological Survey, Water Resources Investigations Research 1997, 96-4215. 23. Berndt, M. E.; Bavin, T. K. Sulfate and Mercury Cycling in Five Wetlands and a Lake Receiving Sulfate from Taconite Mines in Northeastern Minnesota; Minnesota Department of Natural Resources, Division of Lands and Minerals, St. Paul, MN.: 2011. 24. Aiken, G. R.; McKnight, D. M.; Thorn, K. A.; Thurman, E. M., Isolation of hydrophilic organic acids from water using nonionic macroporous resins. Organic Geochemistry 1992, 18, (4), 567-573. 25. Huffman, E.; Stuber, H., Analytical methodology for elemental analysis of humic substances. In Humic Substances in Soil, Sediment and Water: Geochemistry, Isolation, and Characterization., Aiken G.R.; McKnight D.M.; Wershaw R.L.; P., M., Eds. John Wiley & Sons: New York, 1985; pp 433-455. 26. Weishaar, J. L.; Aiken, G. R.; Bergamaschi, B. A.; Fram, M. S.; Fujii, R.; Mopper, K., Evaluation of specific ultraviolet absorbance as an indicator of the chemical composition and reactivity of dissolved organic carbon. Environmental Science & Technology 2003, 37, (20), 4702-4708. 27. Stuiver, M.; Polach, H. A., Discussion: reporting of 14C data. Radiocarbon 1977, 19, (3), 355-363. 28. Johnson, R. L.; Schmidt-Rohr, K., Quantitative solid-state 13C NMR with signal enhancement by multiple cross polarization. Journal of Magnetic Resonance 2014, 239, 44-49.

19 ACS Paragon Plus Environment

Environmental Science & Technology

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

29. McElmurry, S. P.; Long, D. T.; Voice, T. C., Stormwater dissolved organic matter: influence of land cover and environmental factors. Environmental science & technology 2013, 48, (1), 45-53. 30. Breitmeyer, S. E.; Butler, K. D.; Aiken, G. R., Dissolved organic matter data in surface water samples from Minnesota Lakes, 2012 to 2013: U.S. Geological Survey data release. https://doi.org/10.5066/F77M06VP 2017. 31. Curtis, P. J.; Schindler, D. W., Hydrologic control of dissolved organic matter in loworder Precambrian Shield lakes. Biogeochemistry 1997, 36, (1), 125-138. 32. Karlsson, J.; Jonsson, A.; Meili, M.; Jansson, M., Control of zooplankton dependence on allochthonous organic carbon in humic and clear‐water lakes in northern Sweden. Limnology and Oceanography 2003, 48, (1), 269-276. 33. Zigah, P. K.; Minor, E. C.; Werne, J. P., Radiocarbon and stable-isotope geochemistry of organic and inorganic carbon in Lake Superior. Global Biogeochemical Cycles 2012, 26, (1), GB1023. 34. Bade, D. L.; Carpenter, S. R.; Cole, J. J.; Pace, M. L.; Kritzberg, E.; Van de Bogert, M. C.; Cory, R. M.; McKnight, D. M., Sources and fates of dissolved organic carbon in lakes as determined by whole-lake carbon isotope additions. Biogeochemistry 2007, 84, (2), 115-129. 35. Spencer, R. G. M.; Aiken, G. R.; Dyda, R. Y.; Butler, K. D.; Bergamaschi, B. A.; Hernes, P. J., Comparison of XAD with other dissolved lignin isolation techniques and a compilation of analytical improvements for the analysis of lignin in aquatic settings. Organic Geochemistry 2010, 41, (5), 445-453. 36. Spencer, R. G. M.; Aiken, G. R.; Wickland, K. P.; Striegl, R. G.; Hernes, P. J., Seasonal and spatial variability in dissolved organic matter quantity and composition from the Yukon River basin, Alaska. Global Biogeochemical Cycles 2008, 22, (4), GB4002. 37. Mao, J. D.; Xing, B. S.; Schmidt-Rohr, K., New structural information on a humic acid from two-dimensional 1H-13C correlation solid-state nuclear magnetic resonance. Environ. Sci. Technol. 2001, 35, (10), 1928-1934. 38. Cao, X.; Aiken, G. R.; Spencer, R. G.; Butler, K.; Mao, J.; Schmidt-Rohr, K., Novel insights from NMR spectroscopy into seasonal changes in the composition of dissolved organic matter exported to the Bering Sea by the Yukon River. Geochimica et Cosmochimica Acta 2016, 181, 72-88. 39. Schwede-Thomas, S. B.; Chin, Y.-P.; Dria, K. J.; Hatcher, P.; Kaiser, E.; Sulzberger, B., Characterizing the properties of dissolved organic matter isolated by XAD and C-18 solid phase extraction and ultrafiltration. Aquatic Sciences-Research Across Boundaries 2005, 67, (1), 61-71. 40. Hertkorn, N.; Benner, R.; Frommberger, M.; Schmitt-Kopplin, P.; Witt, M.; Kaiser, K.; Kettrup, A.; Hedges, J. I., Characterization of a major refractory component of marine dissolved organic matter. Geochimica Et Cosmochimica Acta 2006, 70, (12), 2990-3010. 41. McCaul, M. V.; Sutton, D.; Simpson, A. J.; Spence, A.; McNally, D. J.; Moran, B. W.; Goel, A.; O’Connor, B.; Hart, K.; Kelleher, B. P., Composition of dissolved organic matter within a lacustrine environment. Environmental Chemistry 2011, 8, (2), 146-154. 42. Lam, B.; Baer, A.; Alaee, M.; Lefebvre, B.; Moser, A.; Williams, A.; Simpson, A. J., Major structural components in freshwater dissolved organic matter. Environmental Science & Technology 2007, 41, (24), 8240-8247.

20 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

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

Environmental Science & Technology

43. Zigah, P. K.; Minor, E. C.; Abdulla, H. A.; Werne, J. P.; Hatcher, P. G., An investigation of size-fractionated organic matter from Lake Superior and a tributary stream using radiocarbon, stable isotopes and NMR. Geochim. Cosmochim. Acta 2014, 127, 264-284. 44. Graham, A. M.; Aiken, G. R.; Gilmour, C. C., Effect of dissolved organic matter source and character on microbial Hg methylation in Hg–S–DOM solutions. Environmental Science & Technology 2013, 47, (11), 5746-5754. 45. Kellerman, A. M.; Kothawala, D. N.; Dittmar, T.; Tranvik, L. J., Persistence of dissolved organic matter in lakes related to its molecular characteristics. Nature Geoscience 2015, 8, (6), 454-457. 46. Meier, M.; Chin, Y.-P.; Maurice, P., Variations in the composition and adsorption behavior of dissolved organic matter at a small, forested watershed. Biogeochemistry 2004, 67, (1), 39-56. 47. Yano, Y.; Lajtha, K.; Sollins, P.; Caldwell, B. A., Chemistry and dynamics of dissolved organic matter in a temperate coniferous forest on andic soils: effects of litter quality. Ecosystems 2005, 8, (3), 286-300. 48. Kawahigashi, M.; Kaiser, K.; Kalbitz, K.; Rodionov, A.; Guggenberger, G., Dissolved organic matter in small streams along a gradient from discontinuous to continuous permafrost. Global Change Biology 2004, 10, (9), 1576-1586. 49. Mao, J. D.; Kong, X. Q.; Schmidt-Rohr, K.; Pignatello, J. J.; Perdue, E. M., Advanced solid-state NMR characterization of marine dissolved organic matter isolated using the coupled reverse osmosis/electrodialysis method. Environmental Science & Technology 2012, 46, (11), 5806-5814. 50. Helms, J. R.; Mao, J.; Stubbins, A.; Schmidt-Rohr, K.; Spencer, R. G.; Hernes, P. J.; Mopper, K., Loss of optical and molecular indicators of terrigenous dissolved organic matter during long-term photobleaching. Aquatic Sciences 2014, 353-373. 51. Aiken, G. R.; Gilmour, C. C.; Krabbenhoft, D. P.; Orem, W., Dissolved organic matter in the Florida Everglades: implications for ecosystem restoration. Critical Reviews in Environmental Science and Technology 2011, 41, (S1), 217-248. 52. Aiken, G., 1.11 Dissolved Organic Matter in Aquatic Systems. In Comprehensive Water Quality and Purification, Ahuja, S., Ed. Elsevier: Waltham, Massachusetts, 2014; Vol. 1, pp 205-220.

539

21 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 28

TPIA

HPOA

Table 1. Peak areas (in %) in 13C multiCP NMR spectra of HPOA and TPIA isolates from different lakes, and the assigned structural moieties associated with the spectral regions. ppm 190160220-190 143-100 123-100 (OCO) 100-64 (OC) 64-0 (Alkyl C) Samples 160 143 Ketone COO/ Arom. Arom. Arom. CH2/CH/ OCnpOb OCHOb OCnpa OCHn Cqa OCH3a CH3a a C NC=O C-O NCH C-C C-H

a b

Williams

1.9

17.4

3.0

5.6

4.1

1.8

0.9

6.4

11.0

5.2

0.1

29.0

13.5

Shingobee

2.6

18.3

4.5

10.0

5.0

2.1

0.8

7.0

9.8

4.4

0.9

23.1

11.5

Manganika

2.5

16.9

6.7

13.2

7.2

2.1

0.8

6.1

10.3

3.1

1.6

20.4

9.3

Williams

1.3

20.8

2.4

3.6

2.4

2.4

1.1

8.9

15.9

4.0

< 0.1

26.8

10.4

Shingobee

2.4

21.3

3.2

5.9

3.2

2.9

0.9

9.3

14.9

3.8

< 0.1

23.3

8.9

Manganika

2.6

20.0

4.2

8.6

3.2

2.8

1.4

9.4

14.2

3.6

< 0.1

21.2

8.8

Based on a multiCP spectrum with 40-µs dipolar dephasing. Based on a CSA-filtered CP/TOSS spectrum and a CSA-filtered CP/TOSS spectrum with dipolar dephasing.

ACS Paragon Plus Environment

Page 23 of 28

Environmental Science & Technology

Figure Captions Figure 1. Solid-state 13C multiCP NMR spectra and spectral editing for identification of specific functional groups of HPOA and TPIA isolates from Williams Lake (a and d), Shingobee Lake (b and e) and Manganika Lake (c and f). Thin black lines: MultiCP spectra showing signals of all C. Red lines: MultiCP with dipolar dephasing showing nonprotonated C and mobile segments such as CH3. Bold black lines: selection of sp3-hybridized C signals by a 13C CSA filter for the separation (see arrows) of O-C-O from aromatic C. Blue lines: selection of nonprotonated sp3hybridized C signals and mobile segments by CSA filter with dipolar dephasing for the separation of nonprotonated O-C-O (shaded area) from nonprotonated aromatic C. The multiCP spectra were scaled to give the same intensity of the COO/NC=O band. Figure 2. 2D 1H-13C HETCOR spectra with 0.5-ms HH-CP of HPOA isolates from (a) Williams Lake, (b) Shingobee Lake and (c) Manganika Lake. 1H slices extracted from the 2D spectra: (d) refers to 1H slices of spectrum (a), (e) 1H slices of spectrum (b), and (f) 1H slices of spectrum (c). Figure 3. 1H slices extracted from the 2D 1H-13C HETCOR spectrum with 0.5-ms HH-CP and 40-µs dipolar dephasing of HPOA isolates from (a) Williams Lake, (b) Shingobee Lake, and (c) Manganika Lake. Figure 4. 2D 1H-13C HETCOR spectrum with 0.5-ms HH-CP and 40-µs dipolar dephasing of TPIA isolates from (a) Williams Lake, (b) Shingobee Lake, and (c) Manganika Lake. Figure 5. Relative carbon percentages of the four compound classes (carbohydrates, CRAM, aromatics, and N-containing) within HPOA and TPIA isolates in Williams Lake, Shingobee Lake, and Manganika Lake.

23 ACS Paragon Plus Environment

Environmental Science & Technology

Figure 1.

24 ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

Environmental Science & Technology

Figure 2.

25 ACS Paragon Plus Environment

Environmental Science & Technology

Figure 3.

Figure 4.

26 ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

Environmental Science & Technology

Figure 5.

27 ACS Paragon Plus Environment

Environmental Science & Technology

ACS Paragon Plus Environment

Page 28 of 28