Anaerobic Dehalogenation of Chloroanilines by Dehalococcoides

Feb 24, 2017 - Whereas CBDB1 preferentially abstracts doubly flanked Cl, 14DCB1 removes Cl ortho to a carbon-attached H. Arrow coding: Bold: main path...
2 downloads 14 Views 2MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Anaerobic Dehalogenation of Chloroanilines by Dehalococcoides mccartyi Strain CBDB1 and Dehalobacter Strain 14DCB1 via Different Pathways as Related to Molecular Electronic Structure Shangwei Zhang, Dominik Wondrousch, Myriel Cooper, Stephen H. Zinder, Gerrit Schüürmann, and Lorenz Adrian Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b05730 • Publication Date (Web): 24 Feb 2017 Downloaded from http://pubs.acs.org on February 24, 2017

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 31

Environmental Science & Technology

1

1 2

Anaerobic Dehalogenation of Chloroanilines by

3

Dehalococcoides mccartyi Strain CBDB1 and Dehalobacter

4

Strain 14DCB1 via Different Pathways as Related to

5

Molecular Electronic Structure

6 7

Shangwei Zhang,†‡§ Dominik Wondrousch,†§ Myriel Cooper,‡ Stephen H. Zinderǁ,

8

Gerrit Schüürmann⃰ †§ and Lorenz Adrian ⃰ ‡┴

9 10

†UFZ Department of Ecological Chemistry, Helmholtz Centre for Environmental Research,

11

Permoserstraße 15, 04318 Leipzig, Germany

12

‡UFZ Department of Isotope Biogeochemistry, Helmholtz Centre for Environmental

13

Research, Permoserstraße 15, 04318 Leipzig, Germany

14

§Technical University Bergakademie Freiberg, Institute for Organic Chemistry,

15

Leipziger Straße 29, 09596 Freiberg, Germany

16

┴Technische Universität Berlin, Fachgebiet Applied Biochemistry,

17

Gustav-Meyer-Allee 25, 13355 Berlin, Germany

18

ǁ

Department of Microbiology, Wing Hall, Cornell University,

19

Ithaca, New York 14853, United States

20

*Corresponding authors: Gerrit Schüürmann, Tel +49-341-235-1262, Fax +49-341-235-45-

21

1262, Email [email protected], and Lorenz Adrian, Tel +49-341-235-1435, Fax +49-

22

341-235-1443, Email [email protected].

23

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 31

2

24

TOC

25 26

ACS Paragon Plus Environment

Page 3 of 31

Environmental Science & Technology

3

27 28

ABSTRACT

29

Dehalococcoides mccartyi strain CBDB1 and Dehalobacter strain 14DCB1 are or-

30

ganohalide-respiring microbes of the phyla Chloroflexi and Firmicutes, respectively.

31

Here we report the transformation of chloroanilines by these two bacterial strains via

32

dissimilar dehalogenation pathways, and discuss the underlying mechanism with

33

quantum chemically calculated net atomic charges of the substrate Cl, H and C at-

34

oms. Strain CBDB1 preferentially removed Cl doubly flanked by two Cl or by one Cl

35

and NH2, whereas strain 14DCB1 preferentially dechlorinated Cl that has an ortho H.

36

For the CBDB1-mediated dechlorination, comparative analysis with Hirshfeld charges

37

shows that the least negative Cl discriminates active from non-active substrates in 14

38

out of 15 cases, and may represent the preferred site of primary attack through

39

cob(I)alamin. For the latter trend, 3 of 7 active substrates provide strong evidence,

40

with partial support from 3 of the remaining 4 substrates. Regarding strain 14DCB1,

41

the most positive carbon-attached H atom discriminates active from non-active

42

chloroanilines in again 14 out of 15 cases. Here, regioselectivity is governed for 10 of

43

the 11 active substrates by the most positive H attached to the highest-charge (most

44

positive or least negative) aromatic C carrying the Cl to be removed. These findings

45

suggest the aromatic ring H as primary site of attack through the supernucleophile

46

Co(I), converting an initial H-bond to a full electron transfer as start of the reductive

47

dehalogenation. For both mechanisms, one- and two-electron transfer to Cl (strain

48

CBDB1) or H (strain 14DCB1) are compatible with the presently available data.

49

Computational chemistry research into reaction intermediates and pathways may fur-

50

ther add in understanding the bacterial reductive dehalogenation at the molecular

51

level.

52 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 31

4

53

INTRODUCTION

54

Chloroanilines are intermediates in the synthesis of dyes, plastics, pharma-

55

ceuticals and pesticides.1, 2 Aromatic pesticides with amino groups such as diuron,

56

linuron, anilide, propanil, and triclocarban can be transformed to chloroanilines in na-

57

tural environments.3,

58

corresponding chloroanilines with the fungicide pentachloronitribenzene being a pro-

59

minent example, which contributes to the widespread occurrence and accumulation

60

of chloroanilines in groundwater, soil, sludge, household waste, rainwater and crops.2

61

Chloroanilines may enter the food chain and therefore represent a potential health

62

risk for animals and humans. Some chloroaniline congeners are recognized as neph-

63

rotoxic, cytotoxic or carcinogenic pollutants5, 6 and may bind to crops, yielding non-

64

extractable residues.7 Adsorption to metal oxide surfaces such as birnessite (MnO2)

65

and iron(III) oxide may trigger chloroanilino radical formation through electron trans-

66

fer, resulting in various dimerization products. 8

4

Moreover, chlorinated nitrobenzenes can be reduced to the

67

Whereas in oxic environments lower chlorinated anilines can be transformed

68

to chlorocatechols and subsequently used as carbon and nitrogen source by a va-

69

riety of microorganisms,9, 10 little is known about their transformation in anoxic envi-

70

ronments. Anaerobic transformation was observed in methanogenic aquifers and

71

aquifer slurries,1, 11-13 flooded soils14 and estuarine sediments;15 the responsible orga-

72

nisms, however, were not identified. In mixed consortia, different dehalogenation

73

pathways were observed: In anaerobic enrichment cultures with estuarine sediments

74

as an inoculum, pentachloroaniline (PeCA) was converted to 2,3,4,5-tetrachloroani-

75

line (2,3,4,5-TeCA),15 i.e. via ortho-dechlorination. A methanogenic enrichment cul-

76

ture transformed PeCA to mainly 2,3,5,6-TeCA with minor amounts of 2,3,4,5-TeCA

77

via dehalogenation of the chlorine substituents in para- and ortho-position.14, 15

ACS Paragon Plus Environment

Page 5 of 31

Environmental Science & Technology

5

78

Recently, the transformation of 2,3-dichloroaniline (2,3-DCA) to 3-monochloro-

79

aniline (3-MCA) by Dehalococcoides mccartyi strain CBDB1 was reported.16 Mem-

80

bers of the genus Dehalococcoides such as strain CBDB1 depend on organohalides

81

as terminal electron acceptor for respiration, energy production and growth. Dehalo-

82

coccoides can transform a large range of different organohalides, and thus are con-

83

sidered as key organisms for natural attenuation and bioremediation of halogenated

84

xenobiotics.17 Besides 2,3-DCA, Dehalococcoides strain CBDB1 has been shown to

85

transform halogenated ethenes, benzenes, phenols, dioxins, benzonitriles and biphe-

86

nyls.16 However, strain CBDB1 was not able to transform 2,4- and 2,6-DCA. Reduct-

87

ive dehalogenation of higher chlorinated anilines has not been investigated with

88

strain CBDB1 or other pure strains. When using 2,3-DCA as terminal electron accep-

89

tor, strain CBDB1 dehalogenated the chlorine substituent in ortho-position of the ami-

90

no group, but abstraction of the chlorine in meta-position was not observed.16 A simi-

91

lar dehalogenation pattern was observed for 1,2,3-trichlorobenzene. Strain CBDB1

92

dehalogenated the doubly flanked chlorine, whereas the singly flanked chlorines

93

were not removed,18 indicating that the amino group in chloroanilines supports deha-

94

logenation in a similar manner as the chlorine substituents in chlorobenzenes. Elec-

95

tron density based natural population analysis (NPA) revealed that the most positive

96

halogen partial charge could predict with a success rate of 96% the regioselective

97

dehalogenation by strain CBDB1 in different compound classes.16 However, chloro-

98

anilines were the few exceptions whose dehalogenation pathway was not predicted

99

correctly using NPA and Mulliken charges. In addition, the prediction of dehalogena-

100

tion pathways based on the most positive halogen partial charge was not applicable

101

to organohalide-respiring bacterial strains from bacterial classes other than the Deha-

102

lococcoides.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 31

6

103

Like Dehalococcoides strains, Dehalobacter strains are specialized on respira-

104

tion with organohalides as terminal electron acceptors and cannot grow by fermenta-

105

tion.19, 20 Strains of both genera use hydrogen as electron donor. The enzymes that

106

catalyze the dehalogenation step in both genera are called reductive dehalogenases.

107

They contain two iron sulfur clusters and a corrinoid ring in the active site, show sig-

108

nificant phylogenic relatedness,20 and thus seem to be very similar.21 Nevertheless,

109

very different electron acceptor ranges and catalyzed reaction pathways were ob-

110

served.20 Pure cultures, defined co-cultures or enrichments containing Dehalobacter

111

spp. have been described to reductively dehalogenate chlorinated methanes,

112

ethanes, ethenes, cyclohexanes, benzenes, phenols, biphenyls and phthalides.19, 20,

113

22, 23

114

transform singly flanked or isolated chlorine substituents. For instance, Dehalobacter

115

strains 12DCB1, 13DCB1 and 14DCB1 dehalogenated 1,2-dichlorobenzene, 1,3-di-

116

chlorobenzene and 1,4-dichlorobenzene, respectively, yielding monochlorobenzene

117

or even some benzene.19 Therefore, Dehalobacter and Dehalococcoides spp. are

118

complementary regarding organohalogen transformation. The potential for reductive

119

dehalogenation of chlorinated anilines by members of the genus Dehalobacter has

120

not been explored.

In addition, several Dehalobacter strains have been described to preferentially

121

While dehalogenation may render organohalides less toxic and more suscepti-

122

ble for subsequent mineralization under aerobic conditions, reductive dehalogenation

123

may also result in the accumulation of more toxic compounds.24 The deeper under-

124

standing of dehalogenation pathways and mechanisms of different dehalogenating

125

strains may allow the enhanced fate prediction of organohalogens from anthropoge-

126

nic sources in the environment and a targeted application of such organisms in bio-

127

stimulation and bio-augmentation processes.

ACS Paragon Plus Environment

Page 7 of 31

Environmental Science & Technology

7

128

In this study, two representative bacterial species, Dehalococcoides mccartyi

129

CBDB1 and Dehalobacter sp. 14DCB1 with different dehalogenation pathways des-

130

cribed for chlorobenzenes,18, 19, 25 were selected to investigate the anaerobic transfor-

131

mation of chlorinated anilines. The objectives of this study were (1) to investigate the

132

pathways and features of dehalogenation of chloroanilines by both stains via cultiva-

133

tion and whole-cell enzymatic activity tests, and (2) to explore the dehalogenation

134

mechanisms through complementary computational chemistry analyses of the elec-

135

tron-acceptor substrates.

136 137 138

MATERIALS AND METHODS

139

Chemicals. Aniline (C6H5NH2) and fifteen of the nineteen possible chloroani-

140

line congeners were obtained from ABCR GmbH, Alfa-Aesar, or Merck Sigma-Aldrich

141

at the highest available purity (Table S1, Supporting Information). The trichloroani-

142

lines 2,3,5-TrCA, 2,3,6-TrCA and the tetrachloroanilines 2,3,4,5-TeCA, 2,3,4,6-TeCA

143

were not available for the experiments, but included in the electron density calcula-

144

tions.

145

Cultivation of Dehalococcoides mccartyi Strain CBDB1 and Dehalobac-

146

ter Strain 14DCB1. The two strains were cultivated as described previously with mi-

147

nor modification.19, 24 Briefly, both strains were grown in vitamin amended, titanium

148

(III) citrate reduced, bicarbonate-buffered mineral medium with 5 mM acetate as car-

149

bon source and hydrogen as electron donor (7.5 mM nominal concentration). All 15

150

chloroanilines were individually used for cultivation of strain CBDB1 and seven (2,3-

151

DCA, 3,4-DCA, 2,3,4-TrCA, 2,4,5-TrCA, 3,4,5-TrCA, 2,3,5,6-TeCA and PeCA) for

152

strain 14DCB1. Chloroanilines were added to the medium at concentrations of 40

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 31

8

153

µM. The cultures were first pressurized to 0.35 bar overpressure of N2:CO2 mixture

154

(80:20%) and then to 0.45 bar overpressure of hydrogen. Both strains were inocu-

155

lated to a starting cell number of about 2.0 × 106 cells/mL. Cultures were set up in

156

triplicates and were statically incubated in the dark at 30 °C. Cells were quantified by

157

direct epifluorescence microscopic cell counting on agarose-coated slides after stain-

158

ing with SybrGreen as described previously.26 Chemical controls without cells, nega-

159

tive controls without terminal electron acceptors, and positive controls with hexa-

160

chlorobenzene, or with a mixture of 1,2,3- and 1,2,4-trichlorobenzene were set up for

161

the strains CBDB1 and 14DCB1, respectively. Cultures were re-fed with 40 µM of the

162

respective electron acceptor and re-gassed every seven days.

163

Enzymatic Activity Assay. A whole-cell enzymatic activity assay using re-

164

duced methyl viologen as artificial electron donor and chloroanilines as electron

165

acceptor was used to quantify reductive dehalogenation. All fifteen available

166

chloroanilines were employed individually for both species. Activity was either as-

167

sessed by gas chromatography with flame ionization detection of dehalogenated

168

products or by photometric quantification of the decolorization of reduced methyl vio-

169

logen during dehalogenation as described previously.27,

170

components of reactions mixtures, which contained 732.5 µL anaerobic water, 125

171

µL methyl viologen (10 mM), 125 µL potassium acetate buffer (1 M), 5.0 µL titanium

172

(III) citrate (15%), and 12.5 µL electron acceptor (50 mM) in 1.0 mL reaction mixture.

173

Whole-cell assays were set-up in an anaerobic glove box in 2 ml HPLC crimp glass

174

vials with 1.0 mL reaction mixture and 150 µL whole cells from parent cultures.

175

Chemical controls and negative controls without electron acceptor were prepared in

176

each test. The cell number of Dehalococcoides mccartyi CBDB1 and Dehalobacter

177

sp. 14DCB1 in parent cultures used for activity assays were 1.9 × 107 and 1.2 × 107

178

cells/mL, respectively. After incubating for 24 h at 30 °C, all samples were sacrificed ACS Paragon Plus Environment

28

Here we optimized the

Page 9 of 31

Environmental Science & Technology

9

179

for analysis of chlorinated anilines by GC-FID and GC-MS.

180

Sample Preparation and Analysis. Samples (1.0 mL from cultures, or 1.15

181

mL from enzymatic activity assays) were sacrificed at planned sampling points. Each

182

sample was mixed with 400 µL hexane and extracted in a shaker with 800 rpm for 24

183

h. Then the hexane extract was analyzed with a Hewlett-Packard GC 6890 coupled

184

to a flame ionization detector, equipped with an HP-5 capillary column (30 m length,

185

0.25 mm i.d., 0.25 µm film thickness). The column temperature was kept at 70 °C for

186

1 min, and then increased to 120 °C at 4 °C/min, 126 °C at 2 °C/min, 136 °C at 0.5

187

°C/min, 160 °C at 2 °C/min, and 300 °C at 20 °C/min. Nitrogen was used as carrier

188

gas at a rate of 1.0 mL/min. Auto injection with 1.0 µL samples was done in the split-

189

less mode. The temperature of the injector and detector were 240 and 300 °C,

190

respectively. Chloroanilines for which no standards were commercially available were

191

analyzed by GC-MS (Agilent 7890A GC-5975 MSD) with an HP-5 column (30 m

192

length, 0.25 mm i.d., 0.25 µm film thickness). The initial temperature of 70 °C was

193

held for 1 min, after which the temperature was increased to 100 °C at 10 °C/min,

194

and finally raised to 200 °C at 4 °C/min. Helium was used as carrier gas with a rate of

195

1.0 mL/min. Auto injection with 1.0 µL samples was done in splitless mode. The

196

temperatures of the injector, MSD source, quadrupole, and transfer line were 280,

197

230, 150 and 280 °C, respectively.

198

Computational Details. Quantum chemical calculations including optimiza-

199

tion of molecular geometries were performed at the B3LYP and BLYP density func-

200

tional theory (DFT) level of theory employing the basis sets Def2-SVP and Def2-

201

TZVP as implemented in Gaussian 09 Revision C.01.29 All ground-state geometries

202

were verified through the absence of imaginary frequencies of the associated Hes-

203

sian matrices. Besides gas-phase calculation, aqueous solution was simulated

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

10

204

through the COSMO30 continuum-solvation model augmented by non-electrostatic

205

contributions as implemented in the CPCM module of Gaussian. Net atomic charges

206

were quantified using the Hirshfeld method31 through Multiwfn 3.3.8,32 and by em-

207

ploying natural population analysis (NPA) of the natural bond orbital (NBO) scheme33

208

calculated with NBO 5.9.34

209 210 211

RESULTS

212

Biotransformation of Chloroanilines by Bacterial Strains CBDB1 and

213

14DCB1. Both Dehalococcoides mccartyi strain CBDB1 and Dehalobacter sp. strain

214

14DCB1 dehalogenated chloroanilines in cultures with hydrogen as electron donor.

215

Seven chloroanilines (2,3-DCA, 3,4-DCA, 2,3,4-TrCA, 2,4,5-TrCA, 3,4,5-TrCA,

216

2,3,5,6-TeCA and PeCA) were transformed by strain CBDB1 (Figure 1, Figure S1 for

217

2,3,5,6-TeCA). All these chloroanilines except PeCA were also dechlorinated in the

218

culture of strain 14DCB1, however, with different rates and to different products (Fig-

219

ure 2). In negative controls without cells, the concentrations of the added chloroani-

220

lines were almost constant over time, and no dehalogenated products were formed.

221

With strain CBDB1, the first products were observed after 7 days of incubation

222

(Figure 1, Figure S1). Both 2,3-DCA and 3,4-DCA were transformed to 3-MCA, how-

223

ever, the transformation of 2,3-DCA was faster than the transformation of 3,4-DCA.

224

For TrCAs, chlorine substituents flanked by two other chlorine substituents were

225

more reactive than chlorine substituents ortho to one Cl and to the amino group. Still

226

slower was the removal of Cl if that was ortho to one other Cl and to H. Taking 2,3,4-

227

TCA as an example, the C3-attached Cl was abstracted preferably with 2,4-DCA as

228

main product. As a minor reaction, removal of C2-attached Cl occurred yielding 3,4-

ACS Paragon Plus Environment

Page 11 of 31

Environmental Science & Technology

11

229

DCA, but no 2,3-DCA was produced, indicating the resistance of C4-attached Cl to

230

microbially mediated dehalogenation. Long-term incubations of PeCA yielded 2,4,6-

231

TrCA, 2,4,5-TrCA, 2,5-DCA and 3,5-DCA as ultimate metabolites. To identify which

232

tetrachloroanilines were formed from PeCA, a CBDB1 culture grown with hexachloro-

233

benzene as electron acceptor was incubated for 12 h at 30 °C with PeCA. Product

234

analysis with GC-MS yielded two peaks at retention times 18.58 min and 18.76 min

235

associated with an m/z value of 230.8 qualifying them as tetrachloroanilines. One

236

peak (18.58 min) had almost the same retention time as a neat standard of 2,3,5,6-

237

TeCA (18.57 min), but it might also be 2,3,4,5-TeCA for which we could not obtain a

238

neat standard. The second peak at 18.76 min was likely to represent 2,3,4,6-TeCA

239

as the other tetrachloroaniline. Production of 2,4,6-TrCA from PeCA indicated further

240

that 2,3,4,6-TeCA was an intermediate in the transformation of PeCA, because the

241

microbial dehalogenation proceeds step by step. No metabolites were found when

242

strain CBDB1 was incubated with 2-MCA, 3-MCA, 4-MCA, 2,4-DCA, 2,5-DCA, 2,6-

243

DCA or 2,4,6-DCA for 55 days (data not shown). A common characteristic of all

244

these non-transformed chlorinated anilines is that they do not contain vicinal chlorine

245

substituents.

246

Dehalobacter sp. strain 14DCB1 dehalogenated three of the seven tested

247

chloroanilines quantitatively (2,3,4-TrCA, 2,4,5-TrCA, 2,3,5,6-TeCA). For three other

248

compounds (2,3-DCA, 3,4-DCA, 3,4,5-TrCA) minor amounts of products were formed

249

(Figure 2), and PeCA was not transformed. Thus strain 14DCB1 removed Cl only if

250

that was ortho to H and to either Cl or NH2. This pattern found for strain 14DCB1 is

251

termed lateral dechlorination because only the two Cl substituents at the beginning

252

and end of a Cl substituent series are removed. Thus, for 3,4,5-TrCA Cl at C3 and at

253

C5 are removed, whereas for 2,3,4-TrCA C2-Cl and C3-Cl without ortho H are not ab-

254

stracted. Interestingly, with the isomer 2,4,5-TrCA as substrate both C2-Cl and C4-Cl ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 31

12

255

are cleaved, but C5-Cl resists 14DCB1-mediated dehalogenation. The lack of in-

256

crease in dehalogenation rate of some substrates in the last period of incubation

257

(Figures 1 and 2) could result from time-dependent variations in chloroaniline toxicity

258

or reflect different bacterial growth stages, and may be subject to future investigati-

259

ons (day 28 had not been included in sampling).

260

As a result of the strain-specific dehalogenation selectivities, the metabolite

261

patterns differ between Dehalococcoides mccartyi strain CBDB1 and Dehalobacter

262

sp. strain 14DCB1 (Figure 3). Strain CBDB1 only dehalogenated chloroanilines with

263

vicinal chlorines, preferring the abstraction of doubly-flanked substituents. By con-

264

trast, strain 14DCB1 dehalogenated chloroanilines with at least one unsubstituted

265

aromatic carbon, removing Cl ortho to H and one other substituent (Figures 2 and 3).

266

Here, the presence of at least one H at a vicinal aromatic carbon is mandatory to

267

make the substrate active for dehalogenation. Accordingly, PeCA resists 14DCB1-

268

mediated Cl abstraction, which holds also for the earlier demonstrated respective sta-

269

bility of hexachlorobenzene.19

270

Enzymatic Activity Assessment. An in vitro test was employed using methyl

271

viologen as artificial electron donor to measure the whole-cell activity of strains

272

CBDB1 and 14DCB1 with the 15 chloroanilines independently from growth, employ-

273

ing hexachlorobenzene and 1,2,3-trichlorobenzene as electron acceptors, respect-

274

ively (Table 1). The results confirm the dehalogenation patterns observed during cul-

275

tivation. Chloroanilines with neighboring Cl atoms were transformed by strain CBDB1

276

cells, and Cl flanked by two other Cl was dechlorinated with the highest specific acti-

277

vity. By contrast, strain CBDB1 did not dehalogenate chloroanilines without neighbor-

278

ing Cl atoms. The aniline group ortho to Cl provided only weak support for CBDB1-

279

mediated dechlorination, converting for instance 2,3-DCA to 3-MCA. Deamination

ACS Paragon Plus Environment

Page 13 of 31

Environmental Science & Technology

13

280

was never observed.

281

As seen in the cultivation, cells of Dehalobacter sp. strain 14DCB1 trans-

282

formed chlorinated anilines with at least one unsubstituted aromatic carbon except

283

for the three monochloroanilines (Table 1). Moreover, for both bacterial strains the

284

metabolites found with the enzymatic activity assay are in line with the pathways ob-

285

served during cultivation. For example, 2,3,4-TrCA was dechlorinated to 2,4-DCA

286

and 3,4-DCA at a ratio of 9:1 by CBDB1, whereas 2,3,4-TrCA was exclusively trans-

287

formed to 2,3-DCA by 14DCB1.

288

Hirshfeld Charge vs CBDB1-Mediated Dechlorination. To explore potential

289

relationships between the electronic structure of the substrates and their activity for

290

dechlorination, Hirshfeld31 net atomic charges have been calculated because of their

291

respective suitability in a previous study.16 Figures 4 and S2 show these charges for

292

all Cl and associated aromatic C atoms as well as for all H atoms of the 15 experi-

293

mentally tested chloroanilines and of four commercially unavailable derivatives; the

294

corresponding NPA33 net atomic charges are available in Figure S3. Here, the color

295

code employed for the Cl atoms is as follows: red indicates Cl abstraction only

296

through strain CBDB1, green Cl abstraction only by strain 14DCB1, and gold Cl abs-

297

traction with both bacterial strains.

298

Regarding strain CBDB1, detailed inspection of Figure 4 unravels links bet-

299

ween the calculated net atomic charge of Cl, qCl, and the experimentally observed

300

dechlorination activity that can be summarized in the two following rules, considering

301

charges as identical if their difference is less or equal 0.003.

302 303

Strain CBDB1 Dechlorination Rule 1:

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 31

14

304

The chloroaniline is active as a strain CBDB1 dehalogenation substrate if there is at

305

least one Cl with qCl ≥ –0.042, ignoring qCl differences up to 0.003.

306 307

Strain CBDB1 Dechlorination Rule 2:

308

Regarding strain CBDB1 regioselectivity, the least negative Cl atom is cleaved pre-

309

ferably, again ignoring qCl differences up to 0.003.

310 311

Rule 1 applies for 14 out of 15 cases corresponding to a success rate of 93%,

312

and thus appears to represent a powerful means for discriminating between active

313

and non-active chloroaniline substrates with strain CBDB1. Taking 2-chloroaniline

314

(A1 in Figure 4) as example, qCl is –0.046 and thus (with the 0.003 range of assumed

315

uncertainty) considered to be below –0.042, qualifying the Cl atom as resistant

316

against CBDB1-mediated dechlorination. For C3 (3,5-DCA), the Hirshfeld charge is

317

–0.052 for both (symmetrical) Cl substituents and thus even more below the thresh-

318

old of –0.042 that would trigger abstraction by strain CBDB1. By contrast, E3 (PeCA)

319

features qCl values of –0.021 and –0.022, respectively, indicating that this congener

320

is active as CBDB1 substrate. The only counter example is D3 (2,4,6-TrCA) with qCl

321

= –0.038 and thus clearly above the threshold, but nevertheless dehalogenated only

322

by strain 14DCB1 (and not by strain CBDB1).

323

Rule 2 reflects a minority trend rather than a majority pattern. As can be seen

324

from Figure 4, B1, D1 and E2 making up 3/7 = 43% follow this regioselectivity rule.

325

By contrast, for the remaining four CBDB1-active substrates (C2, D2, E1, E3) a

326

systematic pattern could not be identified, although D2, E1 and E3 include metabo-

327

lites with the least negative Cl atom being removed. For D1 (2,3,4-TrCA) as example,

ACS Paragon Plus Environment

Page 15 of 31

Environmental Science & Technology

15

328

the Cl atoms with qCl values of –0.030 and –0.033 are abstracted, whereas the Cl

329

atom attached to C4 with qCl = –0.050 resists strain CBDB1, but is removed by

330

14DCB1. The most pronounced counter example is C2 (3,4-DCA), where only the Cl

331

with the lower qCl (–0.056 vs –0.045) is cleaved by CBDB1 (and in fact also by strain

332

14DCB1).

333

Both rules refer to the electronic charge associated with the Cl substituents,

334

and thus suggest an initial attack of the CBDB1-dehalogenating enzyme at Cl, which

335

is discussed in more detail in the next section.

336

Hirshfeld charge vs 14DCB1-Mediated Dechlorination. As outlined above,

337

Dehalobacter sp. strain 14DCB1 yields a metabolite pattern significantly different

338

from the one obtained with Dehalococcoides mccartyi strain CBDB1. In Figure 4, the

339

Cl substituents experimentally active with 14DCB1 are marked through green or gold

340

color, with the latter being also cleaved by CBDB1. Comparative analysis of these

341

experimental results with the Hirshfeld charges obtained for H, Cl and chlorinated

342

carbon yield the following two rules:

343 344

Strain 14DCB1 Dechlorination Rule 1:

345

The chloroaniline is active as a strain 14DCB1 dehalogenation substrate if there is at

346

least one carbon-attached H with qH ≥ 0.047, ignoring qH differences up to 0.003.

347 348

Strain 14DCB1 Dechlorination Rule 2:

349

Regarding strain 14DCB1 regioselectivity, abstraction occurs preferably for Cl ortho

350

to the most positive H attached to the highest-charge (most positive or least nega-

351

tive) carbon, ignoring qCl, qH and qC differences up to 0.003.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 31

16

352 353

Examples for rule 1 are the 14DCB1-active substrates B1 (maximum qH of

354

0.044 equivalent within 0.003 uncertainty to the qH threshold of 0.047), and B2 (one

355

carbon-attached H with qH = 0.048), and A1 as chloroaniline resistant to 14DCB1

356

(maximum qH = 0.043). Here, the only counter example is D3 (2,4,6-TrCA) that is not

357

metabolized by 14DCB1 despite two H atoms with qH = 0.050. Accordingly, the over-

358

all success rate of this rule is 14/15 = 93%.

359

B1 (2,3-DCA) is a simple example for rule 2, because only one of the two Cl

360

atoms has an ortho H atom. In B2 (2,4-DCA), the most positive H atom is ortho to

361

both Cl atoms, but the one actually abstracted is attached to the carbon with the

362

higher Hirshfeld charge (qC 0.010 vs 0.005). Among the four Cl substituents of E2

363

(2,3,5,6-TeCA), only two are ortho to carbon-attached H and thus abstracted through

364

14DCB1, contrasting with the removal of the other two Cl substituents (red color in

365

Figure 4) upon exposure to CBDB1. This rule applies in 10 of 11 cases (all 14DCB1-

366

active substrates except D3 where the Cl attached to the most positive carbon resists

367

abstraction), corresponding to a success rate of 91%.

368

Interestingly, both rules involve the net atomic charge of the H atoms. In addi-

369

tion, the presence of at least one carbon-attached H atom has turned out as prere-

370

quisite for a possible 14DCB1 dechlorination activity of chloroanilines (see above).

371

These findings suggest that the 14DCB1 enzyme catalyzing the Cl abstraction may

372

proceed through initially attacking an H atom of the aromatic substrate, which con-

373

trasts with the presumed dehalogenation mechanism discussed so far in the litera-

374

ture35 as well as with the Cl charge as trigger of the dechlorination activity presently

375

observed with CBDB1.

376

For the presently analyzed chloroanilines, the Hirshfeld charges show only mi-

ACS Paragon Plus Environment

Page 17 of 31

Environmental Science & Technology

17

377

nor variations across different functionals, basis sets and when including simulated

378

aqueous solvation (Figures S5-S7), confirming their relatively low degree of depen-

379

dence on the particular choice of the quantum chemical method.

380 381

382

DISCUSSION

383

Dechlorination Substrates. This study demonstrates that chloroanilines are

384

reductively transformed by Dehalococcoides mccartyi strain CBDB1 and Dehalobac-

385

ter sp. strain 14DCB1. However, the associated dehalogenation pathways differ, re-

386

sulting in different overall dehalogenation patterns. Chloroaniline congeners with

387

neighboring chlorines were transformed by Dehalococcoides mccartyi strain CBDB1,

388

while isolated chlorines were not removed (Figure 1, Figure S1 and Table 1). This

389

dechlorination pattern is in agreement with the ones of strain CBDB1 observed pre-

390

viously with chlorobenzenes,24, 36 and polychlorinated biphenyls37. The dehalogena-

391

tion of most chlorophenols26 and polychlorinated dibenzo-p-dioxins38 by this strain al-

392

so follow this pattern. Especially the case of the chlorinated dioxins emphasizes the

393

importance of the regioselectivity of dehalogenation: strain CBDB1 mainly removed

394

from 1,2,3,7,8-pentachlorodibenzo-p-dioxin the peripheral chlorine substituent form-

395

ing the most toxic dioxin congener 2,3,7,8-tetrachlorodibenzo-p-dioxin.38 In the pre-

396

sent study we confirm the finding of an earlier study16 with three chloroanilines that

397

the amino group in chloroanilines facilitates the dehalogenation of neighboring Cl

398

substituents in a way similar to the supporting function of the oxygen atoms in dioxin.

399

In addition, we certify that Cl substituents provide stronger support for dehalogena-

400

tion than amino substituents with strain CBDB1.

401

By contrast, Dehalobacter sp. strain 14DCB1 removed chlorine substituents ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 31

18

402

that are flanked at least by one hydrogen, i.e. doubly flanked chlorine substituents

403

were not removed. Translated to dioxins, this could prevent the formation of highly to-

404

xic intermediates, raising biotechnological interest in this dehalogenation pattern.

405

This dechlorination pattern was observed previously for the transformation of six dif-

406

ferent chlorobenzenes by strain 14DCB1.19 However, pentachlorobenzene, 1,2,3,5-

407

tetrachlorobenzene, 1,3,5-trichlorobenzene and 1,3-dichlorobenzene with hydrogen

408

in the aromatic ring were not metabolized by Dehalobacter sp. strain 14DCB1.19 The

409

persistence of these three chlorobenzenes against microbial transformation by strain

410

14DCB1 may be induced by the spatial structure of the relevant dehalogenase or

411

dehalogenases, which we have not clarified yet.

412

Reductive Dehalogenases. Both Dehalococcoides mccartyi strain CBDB1

413

and Dehalobacter sp. strain 14DCB1 are organohalide respiring microbes. Strain

414

CBDB1 cannot grow and transform organohalides without exogenous cobalamin.39

415

Similarly, Dehalobacter sp. E1 was not able to grow and dechlorinate β-hexachloro-

416

cyclohexane without the cobalamin offered by Sedimentibacter,22 suggesting that co-

417

balamin is an essential cofactor in the reductive dehalogenase of Dehalobacter spp.,

418

similar to other dehalogenating microorganisms.40-44 In addition, recent reports on the

419

structure of two dehalogenases provided direct evidence for the position of corrin co-

420

factors in reductive dehalogenases, suggesting electron transfer from cob(I)alamin to

421

organohalides as mechanism initiating the dehalogenation.35, 41 However, the struc-

422

ture of the dehalogenases from Dehalococcoides mccartyi strain CBDB1 and Deha-

423

lobacter sp. strain 14DCB1 have not been elucidated yet, which complicates the

424

comparison of mechanisms driving different dehalogenation pathways and patterns.

425

Alternatively, the electronic structure of terminal electron acceptors together with the

426

experimental determination of dehalogenation pathways can give insights into the

427

mechanism of dehalogenation. ACS Paragon Plus Environment

Page 19 of 31

Environmental Science & Technology

19

428

Strain CBDB1 Substrates and Mechanism. Focusing on the substrates, the

429

quantum chemically calculated Hirshfeld31 net atomic charges provide a rationale for

430

the observed dehalogenation pathways of the 15 chloroanilines tested experimentally.

431

Corresponding solution-phase results yield a similar charge pattern (Figure S7),

432

keeping in mind that catalytic sites in protein pockets are usually not exposed to

433

aqueous solvation. Regarding strain CBDB1, the least negative Cl Hirshfeld charge

434

qCl discriminates active from non-active substrates in 14 out of 15 cases (93%). How-

435

ever, only for a subset of three active substrates the observed regioselectivity favors

436

the least negative Cl as primary site of attack (43%), although for some other sub-

437

strates removal of that Cl is included in one of several metabolic routes (see Figure

438

4). Moreover, the Cl net atomic charges are the only ones significantly related to the

439

CBDB1-mediated dehalogenation, confirming previous findings regarding homo- and

440

heterocyclic aromatics as CBDB1 substrates.16, 45 Accordingly and in line with a re-

441

cent study16 we hypothesize that for this bacterial strain, initial electron transfer from

442

cob(I)alamin to Cl initiates reductive abstraction of the latter, contrasting with a long-

443

held view about the aromatic carbon as primary site of enzyme attack.46

444

The present hypothesis is supported by the recently determined structure of

445

the heterologously expressed dehalogenase from Nitratireductor pacificus.35 Regard-

446

ing the electron transfer, at least two mechanistic variants would be compatible with

447

the experimental findings available so far: First, single electron transfer from cob(I)al-

448

amin to Cl could induce its abstraction as Cl–, and the aromatic radical intermediate

449

may receive a further electron from the intermediate cob(II)alamin and a proton to

450

form the reduced product (A in Scheme 1).16 Second, transfer to Cl of two electrons

451

from the supernucleophile cob(I)alamin leads to Cl– cleavage, and the resultant aro-

452

matic anion becomes protonated to yield the dechlorinated metabolite (B in Scheme

453

1). ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 31

20

454

Strain 14DCB1 Substrates and Mechanism. Various dechlorinated metabo-

455

lites generated by strain 14DCB1 differ from the ones obtained through strain CBDB1.

456

This can be traced back to the finding that with 14DCB1, the dehalogenation activity

457

appears to be triggered by the most positive carbon-attached H atom. This holds for

458

14 out of 15 cases (93%), and implies the presence of at least one unsubstituted aro-

459

matic carbon as necessary condition for the substrate activity. Moreover, the Hirsh-

460

feld charges can even predict the regioselectivity in 10 of 11 cases (91%), now in-

461

volving both the H atoms and the aromatic carbon bonded to the Cl that is being re-

462

moved. Again, corresponding solution-phase results as well as respective NPA char-

463

ges are available in Figure S3.

464

So far, the H atoms of unsubstituted aromatic carbons did not play a role in

465

hypotheses about the mechanism underlying the reductive dehalogenation of aroma-

466

tic halides. By contrast, literature considered the aromatic carbon of the substrate as

467

initial site of attack,46 which would make it difficult to rationalize the dechlorination re-

468

sistance of pentachloroaniline. The demonstrated impact of the presence and elec-

469

tronic charge of carbon-attached hydrogen on the activity and regioselectivity of de-

470

chlorination through strain 14DCB1, however, suggests the dehalogenase enzyme to

471

proceed through a primary attack at respective H atoms.

472

Associated possible one- and two-electron transfer mechanisms are outlined

473

in the lower half of Scheme 1. Transfer of one electron from Co(I) of cob(I)alamin to

474

the positively charged H atom could be transmitted along the H–CarCar–Cl unit to

475

cleave Cl–, and the intermediately formed aromatic radical may receive a further elec-

476

tron and a proton to form the dechlorinated product (C in Scheme 1). Alternatively, a

477

two-electron transfer to Cl with cleavage of Cl– could form an aromatic anion that be-

478

comes protonated to yield the reduced metabolite (D in Scheme 1).

479

In any case, the present findings suggest the carbon-attached H atom as priACS Paragon Plus Environment

Page 21 of 31

Environmental Science & Technology

21

480

mary site of attack, which is compatible with previous reports about favorable inter-

481

actions between transition metals and hydrogen of aromatic rings.47,

482

viewpoint, supernucleophilic Co(I) could initially form an H-bond to a carbon-attached

483

H-atom, and then proceed with a full transfer of one or two electrons as outlined in

484

Scheme 1 (C and D).

48

From this

485

The two complementary modes of reductive dehalogenation catalyzed by De-

486

halococcoides mccartyi strain CBDB1 and Dehalobacter sp. strain 14DCB1, respect-

487

ively, provide opportunity to develop targeted strategies for biotechnological conver-

488

sion and biological remediation. Regarding a molecular-level understanding of the

489

underlying mechanisms, computational chemistry research into reaction intermedi-

490

ates and pathways may provide further insight as was demonstrated for the case of

491

cytochrome P450 catalysis.49-51

492 493

ASSOCIATED CONTENT

494

Supporting information

495 496

The Supporting Information is available free of charge on ACS Publications website at DOIW. Additional information as noticed in the text (PDF).

497 498

499

Notes The authors declare no competing financial interest.

500 501

ACKNOWLEDGEMENTS

502

L.A. is supported by the Deutsche Forschungsgemeinschaft in the frame of

503

the Forschergruppe FOR1530. Shangwei Zhang thanks Prof. Dr. Frank-Dieter KopinACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 31

22

504

ke from the UFZ Department of Environmental Engineering for helpful discussions.

505

We thank Benjamin Scheer from the UFZ Department of Isotope Biogeochemistry for

506

technical chemical analyses.

507 508

ACS Paragon Plus Environment

Page 23 of 31

Environmental Science & Technology

23

509

REFERENCES

510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561

(1) Kuhn, E. P.; Suflita, J. M. Sequential reductive dehalogenation of chloroanilines by microorganisms from a methanogenic aquifer. Environ. Sci. Technol. 1989, 23 (7), 848-852. (2) Tas, D. O.; Pavlostathis, S. G. Occurrence, toxicity, and biotransformation of pentachloronitrobenzene and chloroanilines. Crit. Rev. Environ. Sci. Technol. 2014, 44 (5), 473518. (3) Silar, P.; Dairou, J.; Cocaign, A.; Busi, F.; Rodrigues-Lima, F.; Dupret, J.-M. Fungi as a promising tool for bioremediation of soils contaminated with aromatic amines, a major class of pollutants. Nat. Rev. Microbiol. 2011, 9 (6), 477-477. (4) Miller, T. R.; Heidler, J.; Chillrud, S. N.; DeLaquil, A.; Ritchie, J. C.; Mihalic, J. N.; Bopp, R.; Halden, R. U. Fate of triclosan and evidence for reductive dechlorination of triclocarban in estuarine sediments. Environ. Sci. Technol. 2008, 42 (12), 4570-4576. (5) Ito, Y.; Matsuda, Y.; Suzuki, T. Effects of 3, 4-dichloroaniline on expression of AHR2 and CYP1A1 in zebrafish adults and embryos. Comp. Biochem. Physiol. C: Toxicol. Pharmacol. 2010, 152 (2), 189-194. (6) Lee, J.-B.; Sohn, H.-Y.; Shin, K.-S.; Kim, J.-S.; Jo, M.-S.; Jeon, C.-P.; Jang, J.-O.; Kim, J.-E.; Kwon, G.-S. Microbial biodegradation and toxicity of vinclozolin and its toxic metabolite 3,5-dichloroaniline. J. Microbiol. Biotechnol. 2008, 18 (2), 343-349. (7) Lange, B. M.; Hertkorn, N.; Sandermann, H. Chloroaniline/lignin conjugates as model system for nonextractable pesticide residues in crop plants. Environ. Sci. Technol. 1998, 32 (14), 2113-2118. (8) Pizzigallo, M. D. R.; Ruggiero, P.; Crecchio, C.; Mascolo, G. Oxidation of Chloroanilines at Metal Oxide Surfaces. J. Agric. Food Chem. 1998, 46(5), 2049-2054. (9) Boon, N.; Goris, J.; De Vos, P.; Verstraete, W.; Top, E. M. Bioaugmentation of activated sludge by an indigenous 3-chloroaniline-degrading Comamonas testosteroni strain, I2GFP. Appl. Environ. Microbiol. 2000, 66 (7), 2906-2913. (10) Hongsawat, P.; Vangnai, A. S. Biodegradation pathways of chloroanilines by Acinetobacter baylyi strain GFJ2. J. Hazard. Mater. 2011, 186 (2), 1300-1307. (11) Kuhn, E. P.; Townsend, G. T.; Suflita, J. M. Effect of sulfate and organic carbon supplements on reductive dehalogenation of chloroanilines in anaerobic aquifer slurries. Appl. Environ. Microbiol. 1990, 56 (9), 2630-2637. (12) Okutman Tas, D.; Thomson, I. N.; Löffler, F. E.; Pavlostathis, S. G. Kinetics of the microbial reductive dechlorination of pentachloroaniline. Environ. Sci. Technol. 2006, 40 (14), 4467-4472. (13) Okutman Tas, D.; Pavlostathis, S. G. Microbial reductive transformation of pentachloronitrobenzene under methanogenic conditions. Environ. Sci. Technol. 2005, 39 (21), 82648272. (14) Pardue, J. H.; Kongara, S.; Jones, J. W. Effect of cadmium on reductive dechlorination of trichloroaniline. Environ. Toxicol. Chem. 1996, 15 (7), 1083-1088. (15) Susarla, S.; Yonezawa, Y.; Masunaga, S. Reductive dehalogenation of chloroanilines in anaerobic estuarine sediment. Environ. Technol. 1997, 18 (1), 75-83. (16) Cooper, M.; Wagner, A.; Wondrousch, D.; Sonntag, F.; Sonnabend, A.; Brehm, M.; Schüürmann, G.; Adrian, L. Anaerobic microbial transformation of halogenated aromatics and fate prediction using electron density modeling. Environ. Sci. Technol. 2015, 49 (10), 6018-6028. (17) Leys, D.; Adrian, L.; Smidt, H. Organohalide respiration: microbes breathing chlorinated molecules. Philos. Trans. R. Soc., B 2013, 368 (1616), 20120316. (18) Hölscher, T.; Görisch, H.; Adrian, L. Reductive dehalogenation of chlorobenzene congeners in cell extracts of Dehalococcoides sp. strain CBDB1. Appl. Environ. Microbiol. 2003, 69 (5), 2999-3001. (19) Nelson, J. L.; Jiang, J.; Zinder, S. H. Dehalogenation of chlorobenzenes, dichlorotoluenes, and tetrachloroethene by three Dehalobacter spp. Environ. Sci. Technol. 2014, 48 (7), 3776-3782.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 31

24

562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617

(20) Hug, L. A.; Maphosa, F.; Leys, D.; Löffler, F. E.; Smidt, H.; Edwards, E. A.; Adrian, L. Overview of organohalide-respiring bacteria and a proposal for a classification system for reductive dehalogenases. Philos. Trans. R. Soc., B 2013, 368 (1616), 20120322. (21) Nijenhuis, I.; Kuntze, K. Anaerobic microbial dehalogenation of organohalides – state of the art and remediation strategies. Curr. Opin. Biotechnol. 2016, 38, 33-38. (22) Maphosa, F.; Passel, M. W.; Vos, W. M.; Smidt, H. Metagenome analysis reveals yet unexplored reductive dechlorinating potential of Dehalobacter sp. E1 growing in co‐culture with Sedimentibacter sp. Environ. Microbiol. Rep. 2012, 4 (6), 604-616. (23) Justicia-Leon, S. D.; Ritalahti, K. M.; Mack, E. E.; Löffler, F. E. Dichloromethane fermentation by a Dehalobacter sp. in an enrichment culture derived from pristine river sediment. Appl. Environ. Microbiol. 2012, 78 (4), 1288-1291. (24) Adrian, L.; Szewzyk, U.; Wecke, J.; Görisch, H. Bacterial dehalorespiration with chlorinated benzenes. Nature 2000, 408 (6812), 580-583. (25) Jayachandran, G.; Görisch, H.; Adrian, L. Dehalorespiration with hexachlorobenzene and pentachlorobenzene by Dehalococcoides sp. strain CBDB1. Arch. Microbiol. 2003, 180 (6), 411-416. (26) Adrian, L.; Hansen, S. K.; Fung, J. M.; Görisch, H.; Zinder, S. H. Growth of Dehalococcoides strains with chlorophenols as electron acceptors. Environ. Sci. Technol. 2007, 41 (7), 2318-2323. (27) Wagner, A.; Cooper, M.; Ferdi, S.; Seifert, J.; Adrian, L., Growth of Dehalococcoides mccartyi strain CBDB1 by reductive dehalogenation of brominated benzenes to benzene. Environ. Sci. Technol. 2012, 46 (16), 8960-8968. (28) Kublik, A.; Deobald, D.; Hartwig, S.; Schiffmann, C. L.; Andrades, A.; Bergen, M.; Sawers, R. G.; Adrian, L. Identification of a multi‐protein reductive dehalogenase complex in Dehalococcoides mccartyi strain CBDB1 suggests a protein‐dependent respiratory electron transport chain obviating quinone involvement. Environ. Microbiol. 2016, 18, 3044-3056. (29) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision C.01, Gaussian, Inc.: Wallingford, CT, USA, 2011. (30) Klamt, A.; Schüürmann, G. COSMO: A new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient. J. Chem. Soc. Perkin Trans. 2 1993 (5), 799-805. (31) Hirshfeld, F. L. Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acta 1977, 44 (2), 129-138. (32) Lu, T.; Chen, F. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33 (5), 580-592. (33) Reed, A. E.; Curtiss, L. A.; Weinhold, F. Intermolecular interactions from a natural bond orbital, donor-acceptor viewpoint. Chem. Rev. 1988, 88 (6), 899-926. (34) Glendening, E.; Badenhoop, J.; Reed, A.; Carpenter, J.; Bohmann, J.; Morales, C.; Weinhold, F. NBO 5.9. Theoretical Chemistry Institute, University of Wisconsin, Madison 2009. (35) Payne, K. A.; Quezada, C. P.; Fisher, K.; Dunstan, M. S.; Collins, F. A.; Sjuts, H.; Levy, C.; Hay, S.; Rigby, S. E.; Leys, D. Reductive dehalogenase structure suggests a mechanism for B12-dependent dehalogenation. Nature 2015, 517 (7535), 513-516. (36) Adrian, L.; Szewzyk, U.; Görisch, H. Bacterial growth based on reductive dechlorination of trichlorobenzenes. Biodegradation 2000, 11 (1), 73-81. ACS Paragon Plus Environment

Page 25 of 31

Environmental Science & Technology

25

618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662

(37) Adrian, L.; Dudková, V.; Demnerová, K.; Bedard, D. L. “Dehalococcoides” sp. strain CBDB1 extensively dechlorinates the commercial polychlorinated biphenyl mixture Aroclor 1260. Appl. Environ. Microbiol. 2009, 75 (13), 4516-4524. (38) Bunge, M.; Adrian, L.; Kraus, A.; Opel, M.; Lorenz, W. G.; Andreesen, J. R.; Görisch, H.; Lechner, U. Reductive dehalogenation of chlorinated dioxins by an anaerobic bacterium. Nature 2003, 421 (6921), 357-360. (39) Schipp, C. J.; Marco-Urrea, E.; Kublik, A.; Seifert, J.; Adrian, L. Organic cofactors in the metabolism of Dehalococcoides mccartyi strains. Phil. Trans. R. Soc. B 2013, 368 (1616), 20120321. (40) Parthasarathy, A.; Stich, T. A.; Lohner, S. T.; Lesnefsky, A.; Britt, R. D.; Spormann, A. M. Biochemical and EPR-spectroscopic investigation into heterologously expressed vinyl chloride reductive dehalogenase (VcrA) from Dehalococcoides mccartyi strain VS. J. Am. Chem. Soc. 2015, 137 (10), 3525-3532. (41) Bommer, M.; Kunze, C.; Fesseler, J.; Schubert, T.; Diekert, G.; Dobbek, H. Structural basis for organohalide respiration. Science 2014, 346 (6208), 455-458. (42) Kräutler, B.; Fieber, W.; Ostermann, S.; Fasching, M.; Ongania, K. H.; Gruber, K.; Kratky, C.; Mikl, C.; Siebert, A.; Diekert, G. The Cofactor of Tetrachloroethene Reductive Dehalogenase of Dehalospirillum multivorans Is Norpseudo‐B12, a New Type of a Natural Corrinoid. Helv. Chim. Acta 2003, 86 (11), 3698-3716. (43) Krasotkina, J.; Walters, T.; Maruya, K. A.; Ragsdale, S. W. Characterization of the B12- and Iron-Sulfur-containing Reductive Dehalogenase from Desulfitobacterium chlororespirans. J. Biol. Chem. 2001, 276 (44), 40991-40997. (44) Maillard, J.; Schumacher, W.; Vazquez, F.; Regeard, C.; Hagen, W. R.; Holliger, C. Characterization of the corrinoid iron-sulfur protein tetrachloroethene reductive dehalogenase of Dehalobacter restrictus. Appl. Environ. Microbiol. 2003, 69 (8), 4628-4638. (45) Lu, G.-N.; Tao, X.-Q.; Huang, W.; Dang, Z.; Li, Z.; Liu, C.-Q. Dechlorination pathways of diverse chlorinated aromatic pollutants conducted by Dehalococcoides sp. strain CBDB1. Sci. Total Environ. 2010, 408 (12), 2549-2554. (46) Banerjee, R.; Ragsdale, S. W. The Many faces of Vitamin B12: Catalysis by Cobalamin-Dependent Enzymes. Annu. Rev. Biochem. 2003, 72, 209-247. (47) Ribas, X.; Calle, C.; Poater, A.; Casitas, A.; Gómez, L.; Xifra, R. l.; Parella, T.; BenetBuchholz, J.; Schweiger, A.; Mitrikas, G. Facile C− H Bond Cleavage via a Proton-Coupled Electron Transfer Involving a C−H———CuII Interaction. J. Am. Chem. Soc. 2010, 132 (35), 12299-12306. (48) Begum, R. A.; Day, V. W.; Kumar, M.; Gonzalez, J.; Jackson, T. A.; Bowman-James, K. M⋯H–C interaction – Agostic or not: A comparison of phenyl- versus pyridyl-bridged transition metal dimers. Inorg. Chim. Acta 2014, 417, 287-293. (49) Ji, L.; Schüürmann, G. Computational evidence for α-nitrosamino radical as initial metabolite for both the P450 dealkylation and denitrosation of carcinogenic nitrosamines. J. Phys. Chem. B 2012, 116 (2), 903-912. (50) Ji, L.; Schüürmann, G. Model and Mechanism: N‐Hydroxylation of Primary Aromatic Amines by Cytochrome P450. Angew. Chem. Int. Ed. 2013, 52 (2), 744-748; Angew. Chem. 125 (2), 772-776. (51) Ji, L.; Schüürmann, G. Computational biotransformation profile of paracetamol catalyzed by cytochrome P450. Chem. Res. Toxicol. 2015, 28 (4), 585-596.

663 664

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 31

26

665

Tables

666 667

Table 1. Transformation of chloroanilines by Dehalococcoides mccartyi strain CBDB1 and

668

Dehalobacter sp. 14DCB1 measured with a whole-cell activity test. The cells of strain CBDB1

669

and 14DCB1 were grown in a medium with hexachlorobenzene and 1,2,3-trichlorobenzene,

670

respectively.a Dehalococcoides mccartyi strain CBDB1 Electron acceptor used

Products and their relative amount (%)

Specific enzymatic activity

(nkat/mg protein)

Dehalobacter strain. 14DCB1 Products and their relative amount (%)

Specific enzymatic activity

(nkat/mg protein)

2-MCA

nd

0

nd

0

3-MCA

nd

0

nd

0

4-MCA

nd

0

nd

0

2,3-DCA

3-MCA

2,4-DCA

6.7 ± 1.0

2-MCA

0.12 ± 0.01

nd

0

2-MCA

0.71 ± 0.02

2,5-DCA

nd

0

3-MCA

2.2 ± 0.1

2,6-DCA

nd

0

2-MCA

0.65 ± 0.03

3,4-DCA

3-MCA

3.7 ± 0.6

3-MCA

1.4 ± 0.2

3,5-DCA

nd

0

3-MCA

2.4 ± 0.2

2,3,4-TrCA

2,4-DCA (88.9),

37.9 ± 3.7

2,3-DCA

2.7 ± 0.2

24.6 ± 1.1

3,4-DCA (73.2) 2,5-DCA (26.8)

6.9 ± 0.5

3,4-DCA (11.1) 2,4,5-TrCA

2,5-DCA

2,4,6-TrCA

nd

3,4,5-TrCA

3,5-DCA (95.5),

0

2,4-DCA

0.48 ± 0.05

39.2 ± 7.4

3,4-DCA

0.29 ± 0.04

3,4-DCA (4.5) 2,3,5,6-TeCA

2,3,5-TrCA

nq

2,3,6-TrCA 2,3-DCA, 2,6-DCA

nq

PeCA

2,3,4,6-TeCA 2,4,6-TrCA

nq

nd

0

671

a

7

7

672

cells/mL for bacterial strain 14DCB1. A value of 30 fg protein per cell as described previously

673

late the protein amount from counted cell numbers. Abbreviations: nd, no product detected; nq, not quantified be-

674

cause no neat standard was available.

The cell density of the culture was about 1.9×10 cells/mL for bacterial strain CBDB1, and about 1.2×10 27

ACS Paragon Plus Environment

was used to calcu-

Page 27 of 31

Environmental Science & Technology

27

675

Figures

676 677

Figure 1. Formation of dehalogenation products from diverse chlorinated anilines in cultures

678

of Dehalococcoides mccartyi strain CBDB1. Used terminal electron acceptors: (A) 2,3-DCA,

679

(B) 3,4-DCA, (C) 2,3,4-TrCA, (D) 2,4,5-TrCA, (E) 3,4,5-TrCA, (F) PeCA. All cultures initially

680

contained 40 µM, and were re-fed with 40 µM of the respective electron acceptor every

681

seven days.

682 683

ACS Paragon Plus Environment

Environmental Science & Technology

Page 28 of 31

28

684 685

Figure 2. Formation of dehalogenation products from chlorinated anilines in cultures of

686

Dehalobacter sp. strain 14DCB1. Used electron acceptors: (A) 2,3-DCA, (B) 3,4-DCA, (C)

687

2,3,4-TrCA, (D) 2,4,5-TrCA, (E) 3,4,5-TrCA, (F) 2,3,5,6-TeCA. Inlays are magnifications of

688

the curves in cultures with low activity. Cultures initially contained 40 µM and were re-fed

689

with 40 µM of the respective electron acceptor every seven days. No products were found in

690

cultures with PeCA. The concentration of 2,3,6-TrCA in panel F was quantified using 2,4,5-

691

TrCA as standard.

692

ACS Paragon Plus Environment

Page 29 of 31

Environmental Science & Technology

29

693 694

Figure 3. Dehalogenation pathways of chloroanilines in cultures with Dehalococcoides

695

mccartyi strain CBDB1 (A, green arrows) and Dehalobacter sp. strain 14DCB1 (B, blue ar-

696

rows), respectively. Whereas CBDB1 preferentially abstracts doubly-flanked Cl, 14DCB1 re-

697

moves Cl ortho to a carbon-attached H. Arrow coding: Bold = main pathway, ordinary = side

698

pathway, dashed = pathway to GC-unresolved products. Chemical structures in parentheses

699

are hypothesized as intermediates, but not commercially available for experimental confirma-

700

tion.

701

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 31

30

702 703

Figure 4. Chloroaniline test compounds with Hirshfeld net atomic charges (B3LYP/Def2SVP)

704

and color-coded Cl atoms indicating abstraction only by Dehalococcoides mccartyi strain

705

CBDB1 (red), only by Dehalobacter sp. strain 14DCB1 (green), and by both CBDB1 and

706

14DCB1 (gold). Annotation “major” indicates the major site of dechlorination as observed

707

experimentally. The rows and columns are labeled by capital letters (A-E) and numbers (1, 2,

708

3), respectively, to facilitate compound identification (e.g. compound A2 is located in the top

709

middle position of the figure). ACS Paragon Plus Environment

Page 31 of 31

Environmental Science & Technology

31

710

Schemes

711

712

713

Scheme 1. One- and two electron transfer mechanisms for the reductive dehalogenation of

714

chloroanilines catalyzed by Dehalococcoides mccartyi strain CBDB1 (A, B) and Dehalobacter

715

sp. strain 14DCB1 (C, D), respectively, employing 2,3-DCA as example substrate. The pre-

716

sumed electron donor is the supernucleophile Co(I) from cob(I)alamin.

717 718

ACS Paragon Plus Environment