Fractionation of Mercury Stable Isotopes during Microbial

Jul 8, 2016 - The biological production of monomethylmercury (MeHg) in soils and sediments is an important factor controlling mercury (Hg) accumulatio...
0 downloads 0 Views 627KB Size
Subscriber access provided by La Trobe University Library

Article

Fractionation of mercury stable isotopes during microbial methylmercury production by iron- and sulfate- reducing bacteria Sarah E. Janssen, Jeffra K Schaefer, Tamar Barkay, and John R Reinfelder Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00854 • Publication Date (Web): 08 Jul 2016 Downloaded from http://pubs.acs.org on July 9, 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 25

Environmental Science & Technology

Fractionation of mercury stable isotopes during microbial methylmercury

1

production by iron- and sulfate- reducing bacteria

2 3

Sarah E. Janssena*, Jeffra K. Schaefera, Tamar Barkayb, and John R. Reinfeldera

4

5

a

6

New Brunswick, New Jersey 08901, United States

7

b

8

New Brunswick, New Jersey 08901, United States

Department of Environmental Sciences, Rutgers University, 14 College Farm Road,

Department of Biochemistry and Microbiology, Rutgers University, 76 Lipman Drive,

9 10 11

*[email protected]

12

Phone (856) 308-7795

1 ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 25

13

ABSTRACT

14

The biological production of monomethylmercury (MeHg) in soils and sediments is an

15

important factor controlling mercury (Hg) accumulation in aquatic and terrestrial food

16

webs. In this study we examined the fractionation of Hg stable isotopes during Hg

17

methylation in non-growing cultures of the anaerobic bacteria Geobacter sulfurreducens

18

PCA and Desulfovibrio desulfuricans ND132. Both organisms showed mass-dependent,

19

but no mass-independent fractionation of Hg stable isotopes during Hg methylation.

20

Despite differences in methylation rates, the two bacteria had similar Hg fractionation

21

factors (αr/p = 1.0009 and 1.0011, respectively). Unexpectedly, δ202Hg values of MeHg

22

for both organisms were 0.4‰ higher than the value of initial inorganic Hg after about

23

35% of inorganic Hg had been methylated. These results indicate that a 202Hg-enriched

24

pool of inorganic Hg was preferentially utilized as a substrate for methylation by these

25

organisms, but that multiple intra- and/or extracellular pools supplied inorganic Hg for

26

biological methylation. Understanding the controls of the Hg stable isotopic composition

27

of microbially-produced MeHg is important to identifying bioavailable Hg in natural

28

systems and the interpretation of Hg stable isotopes in aquatic food webs.

29

2 ACS Paragon Plus Environment

Page 3 of 25

30

Environmental Science & Technology

INTRODUCTION

31

Methylmercury (MeHg) is a neurotoxic form of mercury (Hg) that readily

32

bioaccumulates in food webs which can lead to elevated risk of human exposure. The

33

process of mercury methylation, which converts inorganic Hg(II) into MeHg, is mediated

34

by anaerobic microorganisms in a variety of different ecological niches including

35

saturated soils, wetlands, and sediments. Within these matrices, mercury methylators

36

with a variety of metabolisms have been identified in pure culture studies and in

37

environmental samples by molecular techniques and incubations with specific inhibitors

38

and stimulators: sulfate reducing bacteria (SRB); iron reducing bacteria (IRB);

39

dehalogenating bacteria; fermentative bacteria; and methanogenic archaea.1-5 Despite

40

their phylogenetic variation, all of the known methylators contain two specific genes,

41

hgcA and hgcB, which are mandatories for the production of MeHg.5 The production of

42

MeHg in aquatic systems is connected to SRB based on correlations with bulk sediment

43

parameters such as acid volatile sulfide (AVS), total sulfur, and sulfate reduction rates.

44

2, 6, 7

45

sediments and freshwater systems has also been shown.8 Pure culture work with IRB

46

has also shown that they can methylate mercury at comparable rates to SRB.4, 9 The

47

rate of methylation is dependent on the cellular metabolism, rate of Hg uptake, and

48

bioavailable pools which can vary greatly between strains and higher order taxonomic

49

groups of microorganisms.4, 10 Despite current knowledge, the exact contribution of

50

different microbial taxa to MeHg pools as well as methylation rates in the environment

51

remains unclear.

However, the importance of other Deltaproteobacteria, such as IRB in low sulfate

3 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 25

52

Mercury stable isotope geochemistry has allowed for the identification of

53

isotopically distinct pools of Hg, such as those in fish tissue and sediment, and limited

54

source apportionment.11-17 In addition, Hg isotope fractionation, the partitioning of Hg

55

isotopes between reactant and product pools during kinetic or equilibrium reactions, has

56

been examined.17-19 Investigations of mercury isotope fractionation aim to develop

57

applications of Hg stable isotopes as biogeochemical proxies and have examined a

58

variety of Hg transformations such as photochemical reduction/demethylation18-20,

59

sorption18, 21, abiotic methylation22,24, microbial transformations25, 26, and volatilization31.

60

Previous Hg isotope studies examined microbially-mediated reduction of Hg(II)

61

and reductive demethylation in resistant organisms with the mer operon.25, 26 Only a

62

few studies have examined the fractionation of Hg stable isotopes during biological Hg

63

methylation32,33 despite its central role in Hg biogeochemistry, mostly due to analytical

64

challenges. Microbial studies are particularly challenging since, unlike animal tissues,

65

the majority of Hg present in microbial cultures is in the inorganic form, which must be

66

separated from MeHg for isotopic analysis.

67 68

Online techniques, utilizing transient GC signals for multicollector inductively

69

coupled plasma mass spectrometry (MC-ICP/MS)34,35, have been used in the past to

70

estimate Hg isotope fractionation factors for Hg methylation by SRB.32,33 Important

71

insight can be obtained from this work, but there are some caveats associated with this

72

approach. Online separation and analysis of MeHg produces a lower external precision

73

in the delta values in comparison with preconcentration and continuous sample

4 ACS Paragon Plus Environment

Page 5 of 25

Environmental Science & Technology

74

introduction.12 In addition to analytical differences, low MeHg yields in previous studies

75

of microbial Hg methylation complicated the approximation of fractionation factors.36

76

Given the wide variety of organisms capable of producing MeHg, it is critical to

77

future systematic isotope studies to understand whether the fractionation of Hg isotopes

78

is strictly due to the reaction catalyzed by the cobalamin-binding protein HgcA, which is

79

common to all Hg methylating organisms, or if Hg isotope fractionation varies across

80

phylogenetic groups and different metabolisms. It has been shown that metabolic

81

differences within the same species of SRB do not influence the fractionation of Hg

82

stable isotopes during methylation32, but it is unknown whether MeHg produced by

83

different organisms such as IRB show a significant shift in isotope ratios in comparison

84

to SRB due to phylogenetic differences or differences in methylation rates.

85

Our work examines the kinetic fractionation of Hg stable isotopes during Hg

86

methylation by pure cultures of the widely studied SRB and IRB strains, Desulfovibrio

87

desulfuricans ND132 and Geobacter sulfurreducens PCA, respectively. These

88

organisms represent two distinct, yet environmentally relevant groups of organisms

89

associated with the biogeochemical cycling of Hg. If different types of methylating

90

microbes produce different MeHg isotopic signatures, Hg isotope fractionation may

91

allow future differentiation of pathways leading to formation of environmental pools of

92

MeHg, specifically MeHg found in sediment and fish tissue.11 Microbially driven Hg

93

isotope fractionation may also prove useful for deciphering multi-step Hg

94

transformations within cells and the bioavailability of various forms of Hg(II) in the

95

environment.

96

5 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 25

97 98

METHODS

99

Microorganisms and culture conditions

100

Strain Geobacter sulfurreducens PCA (ATCC 51573) was obtained from the

101

American Type Culture Collection, and Desulfovibrio desulfuricans ND132 was obtained

102

from C. Gilmour. G. sulfurreducens PCA was cultured under fumarate-reducing (40 mM)

103

conditions with acetate (10 mM) as an electron donor and carbon source at 30°C.9

104

Strain ND132 was initially grown in a modified medium for D. vulgaris with lactate (60

105

mM) as the electron donor and sulfate (30 mM) as the acceptor at 30°C.30 Cultures of D.

106

desulfuricans ND132 were later directly inoculated into a low sulfate media amended

107

with pyruvate/fumarate one week prior to methylation experiments. G. sulfurreducens

108

PCA and D. desulfuricans ND132 were grown under a nitrogen atmosphere.

109

Cell biomass was quantified using the Bradford protein assay (Bio-Rad

110

Laboratories) in order to calculate specific methylation rates for the cultures. Aliquots of

111

cell cultures (1 mL) were taken from methylation assay vials during sampling. The

112

cultures were centrifuged for 10 mins at 9000 x g to remove the supernatant and pellets

113

were stored frozen at -20 °C until analysis. Cells were resuspended in 1 mL of 10 mM

114

TRIS buffer and sonicated for 3 mins in 10 sec intervals. Samples and BSA (Bovine

115

Serum Albumin) standards were mixed with the dye reagent and measured via UV-Vis

116

spectrophotometry (Thermo Scientific Genesys 10S).

117 118

Mercury Methylation Experiments

6 ACS Paragon Plus Environment

Page 7 of 25

119

Environmental Science & Technology

Mercury methylation experiments were performed using washed cells under

120

fermentative conditions. All manipulations of the strains during the washing procedure

121

were performed in an anaerobic chamber (Coy Laboratory Products Inc.) under a gas

122

mixture of 97% N2 and 3% H2. Cells were harvested at exponential phase, centrifuged

123

for 7 min. at 7400 x g, and washed three times in fresh assay buffer and then

124

resuspended in the same assay buffer to an OD660 of 1.5. Assay buffer for G.

125

sulfurreducens PCA and D.desulfuricans ND132 consisted of 10 mM MOPS (pH = 6.8)

126

amended with acetate/fumarate or pyruvate/fumarate, respectively. Prior to the start of

127

the methylation experiments, 10 µM cysteine was equilibrated with 50 nM mercuric

128

nitrate (NIST 3133) in 100 mL of assay buffer in a stoppered serum bottle for one hour

129

at 30°C. Aliquots of washed cells (2.5 mL to 3.5 mL) were then added to the assay vials

130

to a final density (OD660) of 0.05 for G. sulfurreducens PCA and 0.07 for D.

131

desulfuricans ND132 (2.7 ± 0.16 and 5.9 ± 0.64 mg protein mL-1, n=3, respectively).

132

Cells were incubated with Hg(II) for 1-24 hours at 30°C. Methylation rates were

133

calculated using concentrations of MeHg from 0 to 6 h and protein values for each

134

species. At designated time points, Hg methylation was stopped by freezing the entire

135

contents of individual bottles. Bottles from all experiments remained frozen until MeHg

136

analysis.

137

The Hg isotopic composition of cellular and dissolved phase Hg was examined in

138

duplicate incubations of G. sulfurreducens PCA. Cell partitioning was not examined in

139

D. desulfuricans ND132, since previous work showed that with 10 µM cysteine, 95-

140

100% of inorganic mercury is associated with the cells.10 Samples were filtered using a

141

Teflon syringe filter holder with 0.2 µm polycarbonate filters which collected the sum of

7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 25

142

intra and extra cellular Hg. Both filtrate (dissolved Hg) and filters (cell associated Hg)

143

were stored in Teflon vials and digested with bromine chloride to a final concentration of

144

0.06 N.

145

The extent of demethylation was determined in separate experiments that

146

employed similar protocols to methylation assays except that 10 nM of

147

monomethylmercury chloride (Brooks Rand Labs) was added to assays with 10 µM

148

cysteine in place of Hg(NO3)2 (NIST 3133).

149 150 151

Mercury Analysis Samples were distilled prior to methylmercury analysis using a Tekran 2750 gas

152

manifold and heating unit according to EPA method 1630. The only modification made

153

to the distillation procedure was the addition of cupric sulfate (1 M) in place of 1% APDC

154

to mitigate interferences from remnant sulfide in the samples.31 Distillation blanks were

155

all below 0.4 pM and MeHg concentration spike recoveries were within EPA guidelines

156

(112.7 ± 5.5 %, n = 8).32 To assess the extent of abiotic methylation during distillation

157

assay bottles containing media, heat inactivated cell cultures, 10 µM cysteine, and 50

158

nM mercuric nitrate were acidified and also distilled. For both types of media, abiotic

159

methylation produced 1.4 to 2.6 pM which is negligible in comparison to the

160

concentrations of MeHg produced by live cultures and collected for isotopes (2 to 36

161

nM). Distilled samples were derivatized using aqueous phase ethylation and analyzed

162

by gas chromatography coupled to cold vapor atomic fluorescence spectroscopy (GC-

163

CVAFS) on a Tekran 2700.33

8 ACS Paragon Plus Environment

Page 9 of 25

164

Environmental Science & Technology

For total mercury analysis, 5-10 mL aliquots were removed from the original

165

assay or growth bottle and placed into acid cleaned Teflon digestion vials. Samples

166

were oxidized using bromine monochloride (BrCl), to a final concentration of 0.01 N.

167

Total mercury analysis was performed at least 24 hours after BrCl addition. Samples

168

were diluted in ultra-pure water and excess BrCl was neutralized using 2 M

169

hydroxylamine hydrochloride prior to sample introduction. Mercury analysis was

170

performed using tin chloride (0.45 M) reduction followed by cold vapor atomic

171

absorption spectroscopy (CVAAS) on a Hydra AA (Teledyne-Leeman Labs). Procedural

172

blanks were all below 0.1 nM.

173 174 175

Mercury Isotope Analysis Methylmercury isotope samples were distilled and separated from the inorganic

176

matrix using a modified GC separation technique.12 Preconcentration onto gold traps

177

was performed prior to desorption into a 0.06 M potassium permanganate solution

178

according to previously described methods.12 To ensure that no fractionation occurred

179

during processing, MeHg standards (Brooks Rand Methylmercury Chloride) of known

180

isotopic composition were processed and collected during each batch. No significant

181

fractionation was observed for separated MeHg standard (δ202Hg = -0.94 ± 0.18 ‰, n =

182

7) in comparison to directly analyzed MeHg standard (δ202Hg = -0.97 ± 0.12 ‰) (Table

183

S1). Total mercury isotope samples were diluted to appropriate concentrations (2.5 to 5

184

ng g-1) prior to mass spectrometry.

185 186

All samples for mercury isotope analysis were analyzed at Rutgers University using a Thermo-Fisher Neptune multicollector inductively coupled plasma mass

9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 25

187

spectrometer (MC-ICP-MS) according to previously published protocols.18, 34 Bracketing

188

standards (NIST 3133) were concentration and matrix matched prior to analysis.

189

Standards for preconcentrated MeHg samples were diluted in a KMnO4- H2SO4 matrix

190

and standards for THg samples were diluted into a BrCl matrix ( 0.05) was observed in abiotic controls or in live incubations of either

219

bacterium when exposed to 10 nM MeHg (Table 1). Thus, within the analytical

220

uncertainty of MeHg concentration measurements, we observed unidirectional Hg

221

methylation and no demethylation. This indicates that differences in rates and extents

222

of Hg methylation between the two organisms were not due to different abilities to

223

degrade MeHg.

224 225 226

Mercury Stable Isotopic Composition of Microbially Produced MeHg Inorganic Hg added to the methylation assays had an initial isotopic composition

227

(δ202Hg) of 0.00 ± 0.06‰ (n= 10). The concentration of total Hg (MeHg plus Hg(II))

228

varied between bottles by < 20% due to differences in Hg addition, but the isotopic

229

composition of total Hg (δ202Hg = 0.01 ± 0.16‰) remained similar to that of the initial

230

value (Tables S2 and S3) confirming Hg isotope mass balance.

231

For both bacterial strains, δ202Hg values of directly analyzed MeHg were initially

232

lower than reactant Hg(II) by 0.5‰ to 0.7‰, consistent with previous observations that

11 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 25

233

light Hg isotopes preferentially react during kinetically-controlled processes (Fig 2)17-19

234

Minor mass independent fractionation (MIF) of Hg stable isotopes was observed in

235

some THg samples from G. sulfurreducens incubations, but no MIF (∆199Hg >0.13 ‰)

236

was observed in MeHg samples during Hg methylation by either G. sulfurreducens PCA

237

or D. desulfuricans ND132 (Tables S2 and S3). As Hg methylation proceeded, δ202Hg

238

values of MeHg increased as expected with the depletion of 198Hg in reactant Hg(II).

239

After 20% to 25% of Hg(II) was methylated over the first 4 to 6 h, δ202Hg values of

240

MeHg increased above that of the initially added inorganic Hg(II) eventually exceeding

241

the starting value by nearly 0.4‰. In G. sulfurreducens PCA, Hg methylation stopped

242

after 36% of added Hg(II) was methylated (at 48 h; Table S2) and the δ202Hg of MeHg

243

had reached its highest value (0.38‰). In D. desulfuricans ND132, however,

244

methylation continued until about 60% of added Hg(II) was methylated in 24 h (40%

245

remaining as inorganic Hg(II)). As the fraction of unmethylated Hg(II) remaining in

246

incubations of D. desulfuricans ND132 decreased from 66 to 40% between 6 and 24 h,

247

the δ202Hg of MeHg produced by this strain decreased from 0.35‰ to approximately

248

0‰.

249

The production of MeHg with higher δ202Hg values than the Hg(II) initially added

250

indicates that the intracellular pool of substrate Hg(II) available for methylation in both

251

strains was enriched in 202Hg relative to total Hg(II) in the incubations. This was

252

confirmed in short-term experiments with G. sulfurreducens PCA in which cell

253

associated Hg became enriched in 202Hg by 0.13 to 0.28‰ relative to initial Hg(II) (Fig.

254

S2). In G. sulfurreducens PCA, the proportion of Hg that was methylated (36%)

255

exceeded the fraction of total Hg associated with cells (11 to 23%, Fig. S1). Thus, the

12 ACS Paragon Plus Environment

Page 13 of 25

Environmental Science & Technology

256

intracellular pool of substrate Hg(II) for methylation in G. sulfurreducens PCA must have

257

been supplied with Hg(II) from outside the cell. At the concentrations of Hg(II) used in

258

these incubations and in the presence of excess cysteine, the speciation of extracellular

259

Hg(II) was most likely overwhelmingly dominated by cysteine complexes.39 The isotopic

260

enrichment of intracellular Hg(II) may therefore have resulted from the fractionation of

261

Hg isotopes during the exchange of Hg(II) among its various cysteine complexes or

262

among intracellular and extracellular pools during Hg(II) transport.

263

In contrast to G. sulfurreducens PCA, D. desulfuricans ND132 cells accumulate

264

(in the cytoplasm and on the cell surface) most (>90%) of the Hg(II) to which they are

265

exposed.10 Thus, the partitioning of Hg(II) among subcellular compartments was likely

266

responsible for the isotopic enrichment of the substrate pool of Hg(II) in this bacterium.

267

In this case, the observed decrease in δ202Hg values of MeHg after D. desulfuricans

268

ND132 had methylated 34% of added Hg(II) indicates that a different pool of cellular Hg

269

that was not enriched in 202Hg relative to Hg(II) at the start of the experiment supplied

270

substrate Hg(II) once the pool enriched in the heavier isotope was consumed.

271 272 273

Mercury Isotope Enrichment Factors Instantaneous mercury isotope enrichment factors (ε) for bacterial Hg

274

methylation were estimated using δ202Hg values for microbially produced MeHg and the

275

fraction of remaining (unmethylated) Hg(II). Results were fit to a linear fractionation

276

model for a closed system with accumulated product40, 41

277

δ p = δ i + (f r ) × ε p / r

(1)

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 25

278

in which δp corresponds to the δ202Hg values of accumulated MeHg, δi is the initial

279

δ202Hg of substrate Hg(II), fr is the fraction of remaining Hg(II), and εp/r is the product-

280

reactant mercury isotope enrichment factor. Enrichment factors were estimated with

281

respect to the intracellular pool of Hg(II) that was a substrate for methylation by

282

rescaling this "bioreactive fraction" of remaining Hg(II) to 0%. For G. sulfurreducens

283

PCA, the bioreactive fraction of Hg(II) was estimated as the fraction of initially added

284

Hg(II) consumed when methylation stopped (36%) (Fig. S5). Since Hg isotope

285

fractionation may have continued, had Hg methylation by G. sulfurreducens PCA not

286

stopped at 36%, the size of the bioreactive pool of Hg(II) and the estimated enrichment

287

factor for G. sulfurreducens PCA, should be considered minimum values. For D.

288

desulfuricans ND132, the bioreactive fraction of Hg(II) was estimated as the fraction of

289

initially added Hg(II) consumed by the time the δ202Hg of MeHg had stopped increasing

290

(34%).

291

Product-reactant mercury isotope enrichment factors (εp/r) calculated using Eq. 1

292

were -0.92‰ for G. sulfurreducens PCA and -1.1‰ for D. desulfuricans ND132, which

293

correspond to reactant/product fractionation factors (αr/p) of 1.0009 and 1.0011,

294

respectively (Table S5). Our results gave similar enrichment factors for both linear and

295

non-linear (Rayleigh distillation) fractionation models, although no curvature was

296

observed in the δ202Hg vs. fr plots as would be expected for a distillation, and slightly

297

higher coefficients of determination were obtained for the linear model. Linear

298

fractionation of stable isotopes is observed during reversible reactions in which products

299

and reactants reach a state of pseudo-equilibrium (e.g. acid-base or mineral dissolution

300

reactions 41), or during irreversible processes when a proximate pool of substrate 14 ACS Paragon Plus Environment

Page 15 of 25

Environmental Science & Technology

301

exchanges with and is partially buffered by a secondary and often larger pool of

302

reactant (e.g. sulfate reduction in open systems 42) such that isotopic enrichment of the

303

reactant pool proceeds more slowly than would occur if the substrate pool were isolated

304

and finite. Assuming that the final step of Hg methylation in G. sulfurreducens PCA and

305

D. desulfuricans ND132 is irreversible (Table 1), the apparent linear fractionation of Hg

306

stable isotopes during methylation we observed indicates that there were at least two

307

pools of bioreactive Hg(II). The secondary pool may have exchanged with, and partially

308

buffered, the proximate pool of substrate Hg(II) in these bacteria, consistent with the

309

patterns of δ202Hg values in produced MeHg described above.

310

The Hg isotope fractionation factors we estimated for Hg methylation by iron and

311

sulfate-reducing bacteria are higher than those for abiotic Hg methylation (αr/p = 1.0006-

312

1.0007) 20 although Jiménez-Moreno et al.19 recently measured an αr/p of 1.004 for Hg

313

methylation by reaction with methylcobalamin. They are, however, lower than previously

314

reported fractionation factors for two other sulfate-reducing bacteria, Desulfobulbus

315

propionicus, grown under fumarate reducing conditions (1.0026 ± 0.0004)26, and

316

Desulfovibrio dechloracetivorans grown under fumarate- (1.0044 ±1.0011) or sulfate-

317

reducing (1.0025 ± 1.0011) conditions.25 Biologically-catalyzed transformations of Hg,

318

including demethylation of MeHg and Hg(II) reduction were previously shown to have

319

fractionation factors in the range of 1.0003 to 1.0015.23, 24, 29 Differences between the

320

fractionation factors determined in this study and those measured previously may arise

321

from differences in the extents or rates of Hg(II) methylation. For example, the low rates

322

of Hg methylation observed in previous studies25 may have resulted in higher Hg

323

isotope fractionation than observed here. In addition, previously reported fractionation

15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 25

324

factors25 were based on the isotopic composition of Hg(II), not MeHg, and assumed that

325

all added Hg(II) was bioreactive, which our results suggest may not be the case (Fig. 2).

326

Our and prior studies highlight the challenges associated with measuring and modelling

327

Hg isotope fractionation in complex biological media with multiple and exchangeable

328

intra- and extracellular forms of Hg(II).

329

Our results indicate that different taxonomic groups of Hg methylating bacteria

330

have similar Hg isotope fractionation factors. The effects of metabolism on Hg isotope

331

fractionation during Hg methylation by Fe-reducing bacteria has not been examined,

332

and in this study, G. sulfurreducens was grown and incubated under fermentative rather

333

than Fe-reducing conditions. However, as described above, Hg isotope fractionation

334

factors for Hg methylation were not significantly different in Desulfovibrio

335

dechloracetivorans incubated under fumarate or sulfate-reducing conditions.25 This

336

indicates that the fractionation of Hg isotopes is likely dominated by an enzymatic step

337

of the methylation process that is common among diverse anaerobic microorganisms,

338

and is presumably catalyzed by the gene products of hgcA and hgcB.4 However, all

339

strains tested to date are gram-negative bacteria that belong to Deltaproteobacteria.

340

With the discovery that gram-positive bacteria 4 and methanogenic archaea also

341

methylate Hg, further experiments with these organisms may change this picture.

342

Similarities in fractionation factors notwithstanding, major differences in the cellular

343

accumulation of Hg(II) by G.sulfurreducens PCA and D. desulfuricans ND132 indicate

344

that these organisms access distinct intra- and extracellular pools of Hg(II) prior to

345

methylation.

16 ACS Paragon Plus Environment

Page 17 of 25

Environmental Science & Technology

346

In addition to biological effects, the extracellular speciation of Hg(II) will also likely

347

affect the isotopic composition of biologically produced MeHg. For example, it has been

348

shown that Hg(II) bound to thiol16 or iron oxyhydroxide particles22 has lower δ202Hg

349

values compared with aqueous chloro- or hydroxide complexes. How the formation of

350

these and other species of Hg(II) in natural systems affect Hg bioavailability and isotope

351

fractionation in Hg methylating microorganisms has yet to be examined.

352

In our experiments, cells were provided with a relatively high concentration of

353

bioavailable Hg(II) so that the rate of biological Hg methylation would not be limited by

354

the supply of Hg(II) and that the maximum extent of isotope fractionation associated

355

with this transformation would be observed. Apparent product-reactant 202Hg

356

enrichment factors (εp/r) for Hg methylation estimated as the y-intercepts of un-rescaled

357

δ202Hg-fr relationships were -0.58‰ for G. sulfurreducens PCA and -0.87‰ for D.

358

desulfuricans ND132 (Fig. S4). If these values are representative, they would represent

359

the maximum isotopic fractionation between MeHg and total dissolved Hg(II) that may

360

be observed in natural environments, such as sediment porewaters. However, in the

361

environment, the limitation of Hg methylation by the low bioavailability of Hg(II)43-45

362

could result in lower extents of kinetic fractionation and the production of 202Hg-enriched

363

MeHg, as was recently observed in a tidal estuary.12

364

In summary, our results show that 1) the intracellular pool of substrate Hg(II) in

365

Hg methylating anaerobic bacteria is enriched in 202Hg relative to total Hg(II), 2) multiple

366

pools of intra- and extracellular Hg(II) contribute bioreactive Hg for methylation, and 3)

367

Hg stable isotopes provide a useful tool to study the intracellular compartmentalization

368

of Hg(II) and the link between biochemical Hg methylation and Hg(II) bioavailability.

17 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 25

369 370

Supporting Information

371

Concentrations (Fig. S1) and isotopic compositions (Fig. S2) of filtered cell assay

372

experiments, demethylation assays (Fig. S3), isotopic compositions of MeHg check

373

standards to ensure no fractionation (Table S1), concentration and isotopic

374

compositions during experiment for G. sulfurreducens (Table S2) and D. desulfuricans

375

(Table S3), percentage recoveris of separated MeHg samples (Table S4) calculations

376

for fractionation factor (Table S5), and figures for fractionation factor (Fig. S4).

377 378

Acknowledgments

379

This work was supported by grants from NSF's Geobiology and Low-Temperature

380

Geochemistry program (EAR-0952291), the U.S. Department of Energy, Office of

381

Science (BER, DE-SC0007051), the Department of Education's Graduate Assistance in

382

Areas of National Need program, and a Hatch/McIntyre-Stennis grant through the New

383

Jersey Agricultural Experiment Station.

384 385

18 ACS Paragon Plus Environment

Page 19 of 25

Environmental Science & Technology

REFERENCES 1.Yu, R.Q.; Reinfelder, J. R.; Hines, M. E.; Barkay, T. Mercury methylation by the methanogen Methanospirillum hungatei. Appl. Environ. Microbiol. 2013, 79, (20), 6325-6330. 2.Gilmour, C. C.; Henry, E. A.; Mitchell, R. Sulfate stimulation of mercury methylation in freshwater sediments. Environ. Sci. Technol. 1992, 26, 2281-2287. 3.Compeau, G. C.; Bartha, R. Sulfate-reducing bacteria: Principle methylators of ercury in anoxic estuarine sediment. Appl. Environ. Microbiol. 1985, 50, (2), 498-502. 4.Gilmour, C. C.; Podar, M.; Bullock, A. L.; Graham, A. M.; Brown, S. D.; Somenahally, A. C.; Johs, A.; Hurt, R. A.; Bailey, K. L.; Elias, D. A. Mercury methylation by novel microorganisms from new environments. Environ. Sci. Technol. 2013, 47, (20), 11810-11820. 5.Parks, J. M.; Johs, A.; Podar, M.; Bridou, R.; Hurt, R. A.; Smith, S. D.; Tomanicek, S. J.; Qian, Y.; Brown, S. D.; Brandt, C. C.; Palumbo, A. V.; Smith, J. C.; Wall, J. D.; Elias, D. A.; Liang, L. The genetic basis for bacterial mercury methylation. Science 2013, 339, (6125), 1332-1335. 6.Schartup, A. T.; Balcom, P. H.; Mason, R. P. Sediment-porewater partitioning, total sulfur, and methylmercury production in estuaries. Environ. Sci. Technol. 2014, 48, (2), 954-960. 7.King, J. K.; Saunders, F. M.; Lee, R. F.; Jahnke, R. A. Coupling mercury methylation rates to sulfate reduction rates in marine sediments. Environ. Toxicol. Chem. 1999, 18, (7), 1362-1369. 8.Fleming, E. J.; Mack, E. E.; Green, P. G.; Nelson, D. C. Mercury methylation from unexpected sources: molybdate-inhibited freshwater sediments and an iron-reducing bacterium. Appl. Environ. Microbiol. 2006, 72, (1), 457-64. 9.Schaefer, J. K.; Morel, F. M. M. High methylation rates of mercury bound to cysteine by Geobacter sulfurreducens. Nat. Geosci. 2009, 2, (2), 123-126. 10.Schaefer, J. K.; Rocks, S. S.; Zheng, W.; Liang, L.; Gu, B.; Morel, F. M. M. Active transport, substrate specificity, and methylation of Hg(II) in anaerobic bacteria. Proc. Natl. Acad. Sci. USA 2011, 108, (21), 8714-8719. 11.Masbou, J.; Point, D.; Sonke, J. E. Application of a selective extraction method for methylmercury compound specific stable isotope analysis (MeHg-CSIA) in biological materials. J. Anal. At. Spectrom. 2013, 28, (10), 1620-1628. 12.Janssen, S. E.; Johnson, M. W.; Blum, J. D.; Barkay, T.; Reinfelder, J. R. Separation of monomethylmercury from estuarine sediments for mercury isotope analysis. Chem. Geol. 2015, 411, (0), 19-25. 13.Gehrke, G. E.; Blum, J. D.; Marvin-DiPasquale, M. Sources of mercury to San Francisco Bay surface sediment as revealed by mercury stable isotopes. Geochim. Cosmochim. Acta 2011, 75, (3), 691705. 14.Sherman, L. S.; Blum, J. D. Mercury stable isotopes in sediments and largemouth bass from Florida lakes, USA. Sci. Total Environ. 2013, 448, 163-175. 15.Ma, J.; Hintelmann, H.; Kirk, J. L.; Muir, D. C. G. Mercury concentrations and mercury isotope composition in lake sediment cores from the vicinity of a metal smelting facility in Flin Flon, Manitoba. Chem. Geol. 2013, 336, 96-102. 16.Blum, J. D.; Sherman, L. S.; Johnson, M. W. Mercury isotopes in earth and environmental sciences. Annu. Rev.of Earth Planet. Sci. 2014, 42, (1), 249-269. 17.Yin, R.; Feng, X.; Li, X.; Yu, B.; Du, B., Trends and advances in mercury stable isotopes as a geochemical tracer. Trends Environ.Anal. Chem. 2014, 2, 1-10. 18.Wiederhold, J. G.; Cramer, C. J.; Daniel, K.; Infante, I.; Bourdon, B.; Kretzschmar, R. Equilibrium mercury isotope fractionation between dissolved Hg (II) and thiol-bound Hg. Environ. Sci. Technol. 2010, 44, (11), 4191-4197.

19 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 25

19.Bergquist, B. A.; Blum, J. D. The odds and evens of mercury isotopes: Applications of mass-dependent and mass-independent isotope fractionation. Elements 2009, 5, (6), 353-357. 20.Bergquist, B. A.; Blum, J. D. Mass-dependent and -independent fractionation of Hg isotopes by photoreduction in aquatic systems. Science 2007, 318, (5849), 417-20. 21.Jiskra, M.; Wiederhold, J. G.; Bourdon, B.; Kretzschmar, R. Solution speciation controls mercury isotope fractionation of Hg(II) sorption to goethite. Environ. Sci. Technol. 2012, 46, (12), 665462. 22.Jiménez-Moreno, M.; Perrot, V.; Epov, V. N.; Monperrus, M.; Amouroux, D. Chemical kinetic isotope fractionation of mercury during abiotic methylation of Hg(II) by methylcobalamin in aqueous chloride media. Chem.Geol.2013, 336, 26-36. 23.Perrot, V.; Jimenez-Moreno, M.; Berail, S.; Epov, V. N.; Monperrus, M.; Amouroux, D. Successive methylation and demethylation of methylated mercury species (MeHg and DMeHg) induce mass dependent fractionation of mercury isotopes. Chem. Geol. 2013, 355, 153-162. 24.Malinovsky, D.; Vanhaecke, F. Mercury isotope fractionation during abiotic transmethylation reactions. International Journal of Mass Spectrometry 2011, 307, (1-3), 214-224. 25.Kritee, K.; Barkay, T.; Blum, J. D. Mass dependent stable isotope fractionation of mercury during mer mediated microbial degradation of monomethylmercury. Geochim Cosmochim Acta 2009, 73, (5), 1285-1296. 26.Kritee, K.; Blum, J. D.; Johnson, M. W.; Bergquist, B. A.; Barkay, T. Mercury stable isotope fractionation during reduction of Hg(II) to Hg(0) by mercury resistant microorganisms. Environ. Sci. Technol. 2007, 41, (6), 1889-1895. 27.Zheng, W.; Hintelmann, H. Nuclear Field Shift Effect in Isotope fractionation of mercury during abiotic reduction in the absence of light. J Phys. Chem. A 2010, 114, 4238-4245. 28.Zheng, W.; Hintelmann, H. Isotope fractionation of mercury during Its photochemical reduction by low-molecular-weight organic compounds. J Phys.Chemistry A 2010, 114, (12), 4246-4253. 29.Koster van Groos, P. G.; Esser, B. K.; Williams, R. W.; Hunt, J. R. Isotope effect of mercury diffusion in air. Environ. Sci. Technol. 2014, 48, (1), 227-233. 30.Yin, R.; Feng, X.; Wang, J.; Bao, Z.; Yu, B.; Chen, J. Mercury isotope variations between bioavailable mercury fractions and total mercury in mercury contaminated soil in Wanshan Mercury Mine, SW China. Chem. Geol. 2013, 336, 80-86. 31.Ghosh, S.; Schauble, E. A.; Lacrampe Couloume, G.; Blum, J. D.; Bergquist, B. A. Estimation of nuclear volume dependent fractionation of mercury isotopes in equilibrium liquid–vapor evaporation experiments. Chem.Geol. 2013, 336, 5-12. 32.Perrot, V.; Bridou, R.; Pedrero, Z.; Guyoneaud, R.; Monperrus, M.; Amouroux, D. Identical Hg isotope mass dependent fractionation signature during methylation by sulfate-reducing bacteria in sulfate and sulfate-free environment. Environ. Sci. Technol. 2015, 49, (3), 1365-1373. 33.Rodriguez-Gonzalez, P.; Epov, V. N.; Bridou, R.; Tessier, E.; Guyoneaud, R.; Monperrus, M.; Amouroux, D. Species-specific stable isotope fractionation of mercury during Hg(II) methylation by an anaerobic bacteria (Desulfobulbus propionicus) under dark conditions. Environ. Sci. Technol. 2009, 43, (24), 9183-9188. 34.Epov, V. N.; Rodriguez-Gonzalez, P.; Sonke, J. E.; Tessier, E.; Amouroux, D.; Bourgoin, L. M.; Donard, O. F. X. Simultaneous determination of species-specific isotopic composition of Hg by gas chromatography coupled to multicollector ICPMS. J Anal.Chem. 2008, 80, (10), 3530-3538. 35.Dzurko, M.; Foucher, D.; Hintelmann, H. Determination of compound-specific Hg isotope ratios from transient signals using gas chromatography coupled to multicollector inductively coupled plasma mass spectrometry (MC-ICP/MS). Anal.Bioanal. Chem. 2009, 393, (1), 345-55. 36.Kritee, K.; Blum, J. D.; Reinfelder, J. R.; Barkay, T. Microbial stable isotope fractionation of mercury: A synthesis of present understanding and future directions. Chem.Geol. 2013, 336, 13-25.

20 ACS Paragon Plus Environment

Page 21 of 25

Environmental Science & Technology

37.Zane, G. M.; Yen, H.-c. B.; Wall, J. D. Effect of the deletion of qmoABC and the promoter-distal gene encoding a hypothetical protein on sulfate reduction in Desulfovibrio vulgaris Hildenborough. Appl. Environ. Microbiol. 2010, 76, (16), 5500-5509. 38.Olson, M. L.; Cleckner, L. B.; Hurley, J. P.; Krabbenhoft, D. P.; Heelan, T. W. Resolution of matrix effects on analysis of total and methyl mercury in aqueous samples from the Florida Everglades. Fresen. J. Anal. Chem. 1997, 358, (3), 392-398. 39.Methyl mercury in water by distillation, qqueous ethylation, purge and trap, and CVAFS; EPA 1630; U.S. Environmental Protection Agency, U.S. Government Printing Office: Washington D.C., 2001. 40.Liang, L.; Horvat, M.; Bloom, N. S. An improved speciation method for mercury GC/CVAFS after aqueous phase ethylation and room temperature precollection. Talanta 1994, 41, (3), 371-379. 41.Biswas, A.; Blum, J. D.; Bergquist, B. A.; Keeler, G. J.; Xie, Z. Natural mercury isotope variation in coal deposits and organic soils. Environ. Sci. Technol.2008, 42, (22), 8303-8309. 42.Blum, J. D.; Bergquist, B. A. Reporting of Variations in the natural isotopic composition of mercury. Anal.bioanal.chem.2007, 388, 353-359. 43.Gilmour, C. C.; Elias, D. A.; Kucken, A. M.; Brown, S. D.; Palumbo, A. V.; Schadt, C. W.; Wall, J. D. Sulfate-reducing bacterium Desulfovibrio desulfuricans ND132 as a model for understanding bacterial mercury methylation. Appl. Environ. Microbiol. 2011, 77, (12), 3938-3951. 44.Bridou, R.; Monperrus, M.; Gonzalez, P. R.; Guyoneaud, R.; Amouroux, D. Simultaneous determination of mercury methylation and demethylation capacities of various sulfate-reducing bacteria using species-specific isotopic tracers. Environ. Toxicol. Chem. 2011, 30, (2), 337-344. 45.Graham, A. M.; Bullock, A. L.; Maizel, A. C.; Elias, D. A.; Gilmour, C. C. Detailed assessment of the kinetics of Hg-cell association, Hg methylation, and methylmercury degradation in several desulfovibrio species. Appl. Environ. Microbiol. 2012, 78, (20), 7337-7346. 46.Cardiano, P.; Falcone, G.; Foti, C.; Sammartano, S. Sequestration of Hg2+ by some biologically important thiols. J. Chem. Eng. Data 2011, 56, (12), 4741-4750. 47.Hoefs, J. Stable isotope geochemistry Springer-Verlag: Berlin Heidelberg, 2004. 48.Valley, J. W. Stable isotope geochemistry of metamorphic rocks. Rev. Mineral. Geochem. 1986, 16, (1), 445-489. 49.Canfield, D. E. Biogeochemistry of Sulfur Isotopes. Rev. Mineral. Geochem. 2001, 43, (1), 607-636. 50.Benoit, J. M.; Gilmour, C. C.; Mason, R. P. Aspects of bioavailability of mercury for methylation in pure cultures of Desulfobulbus propionicus (1pr3). Appl. Environ. Microbiol. 2001, 67, (1), 51-58. 51.Benoit, J. M.; Gilmour, C. C.; Mason, R. P.; Heyes, A. Sulfide controls on mercury speciation and bioavailability to methylating bacteria in sediment pore waters. Environ. Sci. Technol.1999, 33, (6), 951-957. 52.Drott, A.; Lambertsson, L.; Bjorn, E.; Skyllberg, U. Importance of dissolved neutral mercury sulfides for methyl mercury production in contaminated sediments. Environ. Sci. Technol. 2007, 41, (7), 2270-2276.

21 ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 25

Table 1: MeHg concentrations in cultures of G. sulfurreducens and D. desulfuricans incubated with 10 nM MeHg and 10 µM cysteine. Abiotic controls containedmedia plus MeHg and cysteine, but no cells.

G. sulfurreducens (PCA)

D. desulfuricans (ND132)

Time, hours

MeHg Conc, nM

MeHg Conc, nM

0

9.2 ± 0.6

9.7 ± 0.6

1

8.1 ± 0.5

11.0 ± 2.0

2

8.4 ± 1.6

8.6 ± 0.7

3

8.7 ± 0.4

11.1 ± 1.4

6

8.1 ± 2.0

10.9 ± 0.1

24

8.2 ± 1.8

9.6 ± 2.2

Abiotic (0)

8.0 ± 0.6

9.8 ± 0.5

Abiotic (24)

8.2 ± 0.5

10.2 ± 0.2

22 ACS Paragon Plus Environment

Page 23 of 25

Environmental Science & Technology

Fig. 1: MeHg production by G. sulfurreducens (open symbols) and D. desulfuricans (solid symbols) in washed cell assays with 10 µM cysteine and 50-60 nM mercuric nitrate (NIST 3133). Each point represents a separate assay bottle and symbol shapes correspond to biological replicates performed on different days. The average analytical uncertainty (2SD) for MeHg measurements was 2.6 nM.

23 ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 25

A

B

Fig. 2: Mass-dependent fractionation of 202Hg in MeHg produced by (A) G. sulfurreducens and (B) D. desulfuricans in washed cell assays with 10 µM cysteine and 50-60 nM mercuric nitrate (NIST 3133). δ202Hg values for MeHg are plotted (symbols) with respect to the fraction of remaining total inorganic Hg(II). Gray bars represent the analytical precision associated with the isotopic composition of inorganic Hg (NIST 3133) supplied to the organisms at the start of the experiment. Values are for individual incubation bottles and symbol shapes correspond to biological replicates performed on different days. Uncertainty for MeHg samples were estimated from a separated MeHg standard (Brooks Rand Labs) 2SD = 0.18 ‰. Fitted lines are shown in Fig. S4 and fitting parameters in Table S5.

24 ACS Paragon Plus Environment

Page 25 of 25

Environmental Science & Technology

84x47mm (300 x 300 DPI)

ACS Paragon Plus Environment