AEM Accepted Manuscript Posted Online 26 February 2016 Appl. Environ. Microbiol. doi:10.1128/AEM.03502-15 Copyright © 2016, American Society for Microbiology. All Rights Reserved.

1

Carbon isotope fractionation during catabolism and anabolism in

2

acetogenic bacteria growing on different substrates

3 4

Christoph Freude, Martin Blaser

5 6

Department of Biogeochemistry, Max-Planck-Institute for Terrestrial Microbiology, Karl-

7

von-Frisch-Str. 10, 35043 Marburg, Germany

8 9 10

Keywords:

acetogen,

homoacetogen,

chemolithotroph,

fractionation into biomass.

11 12

Corresponding author:

13

Martin B. Blaser

14

Max Planck Institute for Terrestrial Microbiology

15

Karl-von-Frisch-Str. 10,

16

35043 Marburg, Germany

17

Tel:

+49-6421-178 870

18

Fax:

+49-6421-178 999

19

Email: [email protected]

20

acetate,

formate,

methanol,

21

Abstract

22

Homoacetogenic bacteria are versatile microbes, which use the acetyl-CoA pathway to

23

synthesize acetate from CO2 and hydrogen. Likewise the acetyl-CoA pathway may be used to

24

incorporate other one carbon substrates (e.g. methanol or formate) into acetate or to

25

homoferment monosaccharides completely to acetate. In this study we analyzed the

26

fractionation of acetogenic pure cultures grown on different carbon substrates: While the

27

fractionation of Sporomusa sphaeroides grown on C1 compounds was strong εC1 = -49 to -64

28

‰, the fractionation of Moorella thermoacetica and Thermoanaerobacter kivui using glucose

29

(εGlu = -14.1 ‰) was roughly 1/3 of these values suggesting a contribution of less depleted

30

acetate from fermentative processes. For M. thermoacetica this could indeed be validated by

31

the addition of nitrate, which inhibited the acetyl-CoA pathway resulting in a fractionation

32

during fermentation of εferm = -0.4‰. In addition we determined the fractionation into the

33

microbial biomass of T. kivui grown on H2/CO2 (εanabol. = -28.6 ‰) as well as glucose (εanabol.

34

= +2.9 ‰).

35

36

1. Introduction

37

Acetogenic bacteria are a widespread microbial group unified by their ability to use the

38

reductive acetyl-CoA pathway which allows chemolithoautotrophic growth on hydrogen and

39

carbon dioxide and is the only known pathway that combines carbon dioxide fixation with

40

energy conservation (1-6). Most acetogens can use the reductive acetyl-CoA pathway for

41

autotrophic growth reducing two carbon dioxide molecules with four hydrogen to liberate one

42

acetate. Under these conditions CO2 reduction with H2 yields a ΔG0’ of – 95 kJ/mol. This is

43

barely enough to generate one ATP using either a sodium or proton-dependent electron

44

motive force (7). On the other hand, the acetyl-CoA pathway can be accompanied by

45

fermentation of carbon substrates. Under these autotrophic conditions the energy yield for e.g.

46

glucose fermentation to three acetate is ΔG0’ = - 310.9 kJ/mol. In contrast to most other

47

anaerobe fermentations of glucose which yield only one to three ATP via substrate level

48

phosphorylation during glycolysis, acetogens can redirect the reducing equivalents to the

49

reductive acectyl-CoA pathway, which allows the generation of an additional ATP as

50

described above (3, 8).

51

One approach to evaluate the contribution of different bacterial pathway to certain

52

compound pools in environmental systems is to use the natural abundance of stable carbon

53

isotopes to track their metabolic activity in the environment(9, 10): In principle each

54

metabolic pathway is characterized by a specific preference for a certain carbon isotope

55

(usually the lighter 12C) during catalysis, resulting in a characteristic depletion of 13C between

56

substrate and end product. Mathematically this depletion is characterized by the so called

57

apparent fractionation factor (α) or enrichment factor (ε = 10³[1-α]). For a reaction A → B,

58

the fractionation factor is defined as αA,B = (δ13CA + 103)/(δ13CB +103) (11).

59

There are two types of isotope effects: the equilibrium isotope effect and the kinetic

60

isotope effect. In the first case equilibrium is attained in a closed, mixed system in which a

61

bidirectional reaction occurs. The associated isotope effects are typically small e.g. the

62

dissolution of inorganic carbon (DIC) in seawater results in unequal isotope distribution

63

between the atmospheric CO2 being -7.9 ‰ depleted in 13C compared to the dissolved

64

bicarbonate (12, 13). In contrast biological reactions usually display a kinetic isotope effect,

65

which generally discriminates against the heavy isotope, so that the isotopic value of the

66

product is lower than that of the substrate. Biochemical pathways usually consist of a

67

sequence of irreversible or unidirectional reactions which influence the overall fractionation

68

factor to various extents (14-16). Most biochemical reactions that interconvert C1 compounds

69

have large fractionation effects; in contrast, most heterotrophic reactions involving complex

70

organic substrates have small fractionations (see below). Usually the fractionation factor are

71

determined in microbial pure cultures and can be derived from either substrate depletion or

72

product accumulation (17). Well documented example for carbon isotope fractionations

73

comprise e.g. methanogenesis, showing fractionation factors between -25 to -69‰ for

74

methanogenesis from CO2 (18-21), -73 to -83‰ from methanol (22, 23), and -7 to -35‰ for

75

methanogenesis from acetate (21, 24-27), respectively. In environmental samples

76

fractionation factors can be used to discriminate different pathways, if the difference in

77

fractionation is strong enough e.g. E.g., the fractionation factor for hydrogenotrophic

78

methanogenesis can be determined by incubation of soil samples in the presence of CH3F (28-

79

31), which specifically inhibits aceticlastic methanogenesis (32, 33).

80

The acetyl-CoA pathway imprints a strong isotopic enrichment of -55 to -60 ‰ during

81

acetate formation using H2/CO2 as substrate (34-36). In contrast the fractionation associated

82

with acetate fermentation by microbial glycolytic fermentation is small ( 1 CH3COOH + 2 CO2 + 2 H2O

(2)

148

1 HCOOH + 3 H2 + CO2=> 1 CH3COOH + 2 H2O

(3)

149

4 CH3OH + 2 CO2 => 3 CH3COOH + 2 H2O

(4)

150

1 C6H12O6 => 3 CH3COOH

(5)

151

152

Cultures were grown in 120 mL serum bottles, which contained 50 mL medium. At the

153

times indicated 100 µL gas samples were taken with a gas tight syringe (VICI, Baton Rouge,

154

LA USA) for analysis of concentration and δ13C of both CO2 and CH4; 2 mL of liquid

155

samples were taken for analysis of concentration and δ13C of glucose and short chained fatty

156

acids (SCFA).

157

2.2. Chemical and isotopic analyses: CH4 was analyzed by gas chromatography (GC)

158

using a flame ionization detector (Shimadzu, Kyoto, Japan). CO2 was analyzed after

159

conversion to CH4 with a methanizer (Ni-catalyst at 350 °C, Chrompack, Middelburg,

160

Netherlands). Isotope measurements of 13C/12C in gas samples were performed on a gas

161

chromatograph combustion isotope ratio mass spectrometer (GC-C–IRMS) system (Thermo

162

Fisher Scientific, Bremen, Germany). The principle operation was described by Brand (51).

163

The gaseous compounds were first separated in a Hewlett Packard 6890 GC using a Pora Plot

164

Q column (27.5 m length, 0.32 mm internal diameter, and 10 µm film thickness; Chromopack

165

Frankfurt, Germany) at 30 °C and He (99.996 % purity; 2.6 ml/min) as carrier gas. The

166

sample was run through the Finnigan Standard GC Combustion Interface III and the isotope

167

ratio of 13C/12C was analyzed in the IRMS (Finnigan MAT Deltaplus). The reference gas was

168

CO2 (99.998 % purity) (Air Liquide, Düsseldorf, Germany), calibrated with the working

169

standard methylstearate (Merck). The latter was intercalibrated at the Max Planck Institute for

170

Biogeochemistry, Jena, Germany (courtesy of W. A. Brand) against the NBS-22 and USGS-

171

24 standards and reported in the delta notation versus Vienna Pee Dee Belemnite.

172

δ 13C = 103(Rsample/Rstandard-1)

173

with R = 13C/12C of sample and standard, respectively.

174

Isotopic analysis and quantification of glucose and SCFA were performed on a high

(6)

175

pressure liquid chromatography (HPLC) system (Spectra System P1000, Thermo Fisher

176

Scientific, San Jose, CA, USA; Mistral, Spark, Emmen, the Netherlands) equipped with an

177

ion-exclusion column (Aminex HPX-87-H, BioRad, München, Germany) and coupled to

178

Finnigan LC IsoLink (Thermo Fisher Scientific, Bremen, Germany) as described (52). Isotope

179

ratios were detected on an Isotope Ratio Mass Spectrometer (IRMS) (Finnigan MAT

180

Deltaplus Advantage). The precision of the GC-IRMS was ±0.2 ‰, and that of the HPLC-

181

IRMS was ±0.3 ‰. For further details on determination of acetate concentration via HPLC-

182

IRMS compare the supplementary material (Fig. S9, Tab. S2).

183

The isotopic fractionation into the biomass was evaluated by harvesting three replicate

184

samples (120 mL serum bottles) each containing 50 mL culture for each time point.The cells

185

were harvested by centrifugation (10 minutes at 6.000 rpm (9.700g)). The resulting pellet was

186

dried in tin capsules and sent to the isotope centre in Göttingen. There the isotopic values

187

were determined using an elemental analyser coupled to an IRMS (Kompetenzzentrum

188

Stabile Isotope at the University of Göttingen, Germany). Pure acetate was used as a reference

189

to compare the results with our HPLC-IRMS measurements.

190

2.3. Calculations:

191

The apparent fractionation factor α for a reaction A → B is defined after (11) as:

192

αA/B =(δA + 1000)/(δB + 1000)

193

also expressed as ε ≡ 103 (1 – α). The isotope enrichment factor ε associated with

194

(7)

homoacetogenesis was determined as described by Mariotti (17) from the residual reactant

195

δr = δri + ε[ln(1− f)]

(8)

196

and from the product formed

197

δp = δri − ε(1− f)[ln(1− f)]/ f

198

where δri is the isotope composition of the reactant (CO2) at the beginning, δr and δp are the

(9)

199

isotope compositions of the residual CO2 and the pooled acetate, respectively, at the instant

200

when f was determined. Linear regression of δr against ln(1 – f) and of δp against (1 – f)[ln(1 –

201

f )]/f gives ε as the slope of best-fit lines. The fractional yield f of a reaction is usually based

202

on the consumption of the substrate CO2 (0 < f < 1); in a closed system the amount of used

203

substrate (f) can be derived from the measured isotopic composition (35).

204

fdelta= (δri − δr)/( δp – δr)

205

In contrast to the isotope enrichment factor ε derived from microbial pure cultures as

206

(10)

described above, many environmental studies express the apparent fractionation Δ as:

207

Δ = δprod – δsubst

208

Where δsubst is the δ13C value of the substrate and the δprod is the δ13C value of the product.

209

While this estimated fractionation is valid for open systems like environmental samples, it

210

is only valid at the very beginning of a reaction in a closed system with limited substrate

211

concentrations (17, 53, 54). Nevertheless we used this estimation to calculate the apparent

212

fractionation for the mixotroph experiment to estimate the fractionation into acetate vs.

213

biomass for cells grown on formate or H2/CO2.

214

(11)

215

Results

216

Growth on C1 compounds

217

In order to evaluate the fractionation factor in the acetyl-CoA pathway, we measured the

218

production of acetate during the consumption of different C1 compounds which are either

219

direct intermediates (CO2, formate) or can be easily incorporated into the acetyl-CoA pathway

220

(methanol) (compare Fig. 1). These experiments were conducted using Sporomusa

221

sphaeroides. The substrate depletion and product formation was analyzed over time and the

222

delta 13C of all carbon containing compounds was measured during the reaction (Fig. 2).

223

Despite different 13C values of the initial substrates methanol δmethanol = -38 ‰, formate δformate

224

= -31 ‰ or H2/CO2 δCO2 = -17 ‰ the released acetate eventually reached the same value

225

δacetate = -67 ‰. The delta 13C of all three substrates became enriched in heavy carbon with

226

roughly the same rate. When the apparent fractionation factors were calculated based on the

227

substrate data the following values were obtained: methanol εmethanol = -56.9 ‰, formate

228

εformate = -56.5 ‰ and H2/CO2 εCO2 = -63.8 ‰. The delta 13C of CO2 was not changing much

229

when methanol or formate served as substrates due to the large background of bicarbonate

230

(4.5 mmol/bottle NaHCO3) in the culture medium. Details on the acetate formation for the

231

individual substrates are given in Fig. S2. Details and further discussion on the isotope values

232

of acetate are given in Fig. S3 and supplementary discussion.

233

In a different approach the influence of different C1 compounds was tested using the

234

thermophilic acetogen Thermoanaerobacter kivui grown under a headspace of N2/CO2 on

235

formate (1 mmol per bottle). In parallel T. kivui was inoculated under a headspace of H2/CO2

236

(2.5 mmol and 0.6 mmol per bottle (70 mL headspace in 120 mL)) either alone or

237

mixotrophic in the presence of low (0.25 mmol per bottle (50 ml medium per 120 mL bottle))

238

or high (1 mmol per bottle (50 ml medium per 120 mL)) formate concentrations (Fig. 3, Tab.

239

1). The bicarbonate concentration was 4.5 mmol per bottle in all incubations. Details on the

240

carbon turnover are given in Fig. S4. Using hydrogen (equ. 1) and formate (equ. 3)

241

consumption as proxies for the substrate turnover into acetate an almost complete turnover

242

could be documented for all experiments (Tab. 1). This was used to calculate the relative

243

contribution of both substrates under mixotrophic conditions and to estimate the apparent

244

fractionation Δ between substrate (t0) and product (acetate, tend) or substrate (t0) and biomass

245

(tend). A certain statistically not significant trend towards 13C depleted acetate as well as

246

biomass could be observed for increasing contributions of formate vs. CO2 (Tab. 1). The

247

calculated enrichment factor using the linear regression given in equ. 13 resulted in εCO2 = -

248

48.7 ‰ for the incubations under H2/CO2 and εform = -51.9 ‰ for the formate incubations

249

under N2/CO2.

250 251

Growth on glucose and nitrate inhibition

252

In a further set of experiments the fractionation was determined when acetogenic pure

253

cultures were grown on complex carbon substrates. As an example Moorella thermoacetica

254

grown on glucose (200 µmol), glucose (200 µmol) plus nitrate (300 µmol), or H2/CO2 (2.5

255

mmol / 625 µmol) was used. Indeed our isotopic measurements revealed almost no

256

fractionation with glucose and nitrate εGlu+NO3 = -0.4 ‰, strong fractionation with H2/CO2

257

εH2/CO2 = -55.8 ‰ and an intermediate fractionation with glucose εGlu = -18.5 ‰ reflecting the

258

mixing of both acetate generating pathways (Fig. 4). Additional data on the nitrate depletion

259

are given in Fig. S5. Similar results could be obtained for T. kivui where the fractionation

260

under glucose (εGlu = -14.1 ‰) was roughly 1/3 of the fractionation under H2/CO2 (εH2/CO2 = -

261

53.0 ‰) (compare Fig. S6) and also for Acetobacterium carbinolium (εGlu = -18.8 ‰ vs.

262

εH2/CO2 = -54.0 ‰) (data not shown).

263 264

Fractionation into biomass

265

In contrast to aerobe bacteria anaerobic bacteria can only use a small amount of the

266

available energy for anabolism (usually around 10 %) and hence only a small portion of the

267

available carbon is fixed into biomass. However it was observed that T. kivui grew to dense

268

cultures when grown on glucose but only faint turbidity was observed for cells grown on

269

H2/CO2. Therefore it was questioned how substrate usage affects the delta 13C values of the

270

microbial biomass. Therefore, Thermoanaerobacter kivui was grown on glucose (200 µmol

271

under N2/CO2) or on H2/CO2 (2.5 mmol / 625 µmol) in the described bicarbonate buffered

272

minimal medium. Hence we expected that under both conditions roughly the same amount of

273

acetate would be produced. The delta 13C of substrate and product as well as biomass was

274

followed over time (Fig. 5). The fractionation during catabolism was calculated using the 13C

275

values of the substrate together with the 13C values of the acetate using eq. 8; the fractionation

276

during anabolism was calculated respectively using the 13C values of the substrate together

277

with the 13C values of the biomass. While growth on glucose resulted in a small positive

278

fractionation during catabolism εcatabol. = + 4.2 ‰ as well as anabolism εanabol. = + 2.9 ‰,

279

growth on H2/CO2 was accompanied by a stronger fractionation caused by catabolism εcatabol.

280

= -48.6 ‰ than anabolism εanabol. = -28.6 ‰.

281

282

Discussion

283

Growth on C1 compounds

284

The data presented in this manuscript confirmed the overall strong fractionation of

285

acetogens using the acetyl-CoA pathway when H2/CO2 was the substrate (34, 35); in addition

286

it could be shown, that the fractionation of cells grown on different C1-compounds likewise

287

resulted in a very strong fractionation and was largely independent from substrate usage

288

(different C1 compounds). This is in contrast to several published examples where the rate

289

limiting step (usually the initial step of a reaction cascade) is primarily determining the

290

overall fractionation.

291

One prominent example that the initial step of a reaction cascade is determining the overall

292

fractionation can be found in the plant system, where the initial CO2 fixing reaction of C3 and

293

C4 plants (RubisCO vs. phosphoenolpyruvate carboxylase) dictates strong differences in the

294

resulting plant material (55) which is largely unchanged throughout the food chain (56, 57). In

295

the microbial world a similar behavior is well documented for methanogenic pure cultures on

296

different substrates, where each substrate is accompanied by a unique fractionation factor: e.g.

297

methanogenic pure cultures of M. barkeri grown on different carbon substrates show highly

298

distinct carbon isotope fractionations from substrate to product when grown on H2/CO2 (-45.4

299

to -46.3 ‰), methanol (-72.5 to -83.4 ‰) or acetate (-21.2 to -34.8 ‰) (22, 58). These results

300

have been independently validated for other methanogenic pure cultures on H2/CO2 (20, 21),

301

methanol (23) and acetate (25, 27). Further on, it was shown for hydrogenotrophic

302

methanogens (using H2/CO2) that the fractionation is highly affected by the hydrogen

303

(energy) level of the incubations resulting in stronger fractionation under H2 limiting

304

conditions (20-22). For a sulfate reducing Desulfovibrio the fractionation of sulfur isotopes

305

was also affected by the electron donors and negatively correlated to the sell-specific sulfate

306

reduction rate (59). Recent evidence suggest that also for sulfate reducing bacteria the overall

307

isotope fractionation during sulfate reduction is influenced by all steps in the dissimilatory

308

pathway (16, 60).

309

By contrast, differences in the substrate ratios of H2 and CO2 using acetogenic cultures did

310

not affect the fractionation (61). Furthermore, it has been shown that there is no

311

intramolecular fractionation during acetate formation: isotope values of the methyl-group

312

were similar to the delta 13C values of the overall acetate signal (35). The range of

313

fractionation for different C1 compounds reported in this study is well within the previous

314

reported range (-38 to -68 ‰) for acetogenic cultures grown on H2/CO2 (34, 35). And now

315

our data on the conversion of different C1 compound to acetate suggest a similar

316

discrimination against 13C irrespective of the individual C1 compound.

317

Our results of the S. sphaeroides incubations can be interpreted in two ways: 1.) One

318

explanation is that the initial fractionation of all possible initial enzymes is highly similar. 2.)

319

Alternatively our results can be explained if we assume that the rate limiting step determining

320

the overall fractionation is the C-C bond formation by the CODH/ACS complex. We will

321

discuss both possibilities:

322

1.) The individual entry points for the different C1 compounds are schematically given in

323

Fig. 1. In detail they involve the following four (a-d) enzymes: methanol is incorporated using

324

a methanol cobalamin methyltransferase (a) (62). While one molecule of CO2 is reduced via

325

CO to the carboxyl-group of acetate by the enzyme-complex CO dehydrogenase / acetyl-CoA

326

synthase (b) a second molecule of CO2 is reduced to formate by the formate dehydrogenase

327

(c). Hence formate is a direct intermediate and can either be oxidized to CO2 or further

328

reduced in an ATP-dependent reaction catalyzed by the formyl-tetrahydrofolate synthetase (d)

329

(3). In principle it is possible that all these four entry points exhibit the same fractionation.

330

However, fractionation of methanol and CO2 in methanogenic pure cultures renders

331

completely different fractionation factors using homologous enzymes for the initial

332

incorporation. Therefore we think that in acetogenic bacteria the fractionation is mainly

333

controlled by a later step during the reductive acetyl-CoA pathway; presumably the

334

condensation of the two carbons in the CODH / ACS system:

335

2.) All the substrates we used were incorporated into the methyl branch of the acetyl-CoA

336

pathway (Fig. 1) yielding the methyl-group of the released acetate. However, the reduction of

337

formate is accompanied by the release of CO2 (eq. 2); while one part of the methanol is

338

oxidized to CO2 to provide reducing equivalents three molecules of methanol and CO2 are

339

reduced to acetate resulting in a net consumption of CO2 (eq. 4). In any case our experimental

340

set up does not allow the discrimination whether the carboxyl-group is directly formed from

341

external CO2 reduction (e.g. carbonate buffer) or originates from an intracellular CO2 pool

342

generated through oxidation of methanol or formate.

343

The acetyl-CoA pathway can operate in both directions; while most acetogens use it for

344

carbon fixation from CO2 and energy conservation via an Ech of Rnf system (7) others can

345

operate in the reverse direction together with an syntrophic partner to perform syntrophic

346

acetate oxidation (63, 64). As a result most reactions of the acetyl-CoA pathway can be seen

347

as highly reversible and hence can easily equilibrate different isotopic δ13C values. Therefore

348

it is plausible that not the initial step but the first condensing step is the critical point for the

349

fractionation. Indeed this explanation would be in agreement with the results obtained for the

350

C3 and C4 plants where the initial step is a C-C forming condensation reaction.

351

When T. kivui is grown on the expense of a methylated aromatic compound (vanillic acid)

352

the initial step (O-demethylation of vanillic acid) may be less reversible. Indeed initial trials

353

resulted in less depleted acetate for T. kivui grown on the expense of vanillic acid (δac =-50

354

‰; compare Fig. S7) compared to δac =-67 ‰ derived from cultures grown on C1 compounds

355

(compare Fig. 1). This data could however not be further evaluated since the 13C value of the

356

substrate vanillic acid could not be measured using our HPLC-IRMS approach.

357 358

Growth on glucose and nitrate inhibition

359

Both investigated cultures (M. thermoacetica and T. kivui) showed a strong fractionation

360

(ca. -50 to -55 ‰) during acetate formation when the acetyl-CoA pathway was the only

361

operative pathway (34). As soon as glucose was used as a substrate the apparent fractionation

362

was a mixture of fermentation and acetyl-Co pathway resulting in a fractionation of less than

363

< -20 ‰.

364

These findings have environmental implications: The strong fractionation during acetate

365

production via the acetyl-CoA pathway can easily be tracked in natural environments. Ideally

366

it can be used to estimate the contribution of acetogens to the acetate pool, when the

367

fractionation of all other acetate generating and depleting reactions is known (9). The mixed

368

fractionation of acetogens grown on larger molecules like glucose (less than < -20 ‰) may be

369

difficult to differentiate in environmental systems where the acetate signal deviates by ± 10 ‰

370

from the delta13C value of the total organic carbon (65). Part of this variation may originate

371

from the contribution of the acetogenically formed 13C depleted acetate. Indeed incubations of

372

lake sediments only resulted in 13C depleted acetate under H2/CO2 addition and not in control

373

experiments under N2 (66), suggesting that the acetogens in the later samples are either less

374

active or grow on complex substrates.

375 376

Fractionation into biomass

377

Many studies use published pure culture fractionation factors to deduce the importance of a

378

certain reaction in environmental settings. Most of these fractionation factors have been

379

calculated using a mass balance equation based on the substrate to product conversion but

380

neglecting the buildup of biomass. Hence we aimed to validate if this approach is reasonable.

381

Comparing the fractionation during catabolism and anabolism shows that fractionation into

382

the biomass is usually lower than the discrimination during catabolism. E.g. the fractionation

383

into the biomass is roughly 1/3 of the fractionation into methane for M. barkeri grown on

384

H2/CO2 (-13.9 vs. -45.3 ‰), acetate (-7.3 vs. -34.4 ‰), methanol (-31.1 vs. -83.4 ‰) and

385

trimethylamine (-24.8 vs. -66.5 ‰) (22).

386

So far the fractionation into biomass of bacterial pure cultures has mainly been documented

387

using endpoint measurements of the microbial biomass (compare Table S1). In contrast we

388

wanted to calculate the apparent fractionation into the biomass following a time series

389

experiment. Indeed we found that the fractionation during anabolism was roughly 2/3 of the

390

fractionation during catabolism. We discuss some possible reasons for this ration in the

391

supplementary discussion together with Fig. S8.

392

In general we would conclude that the fractionation into biomass is usually smaller than

393

the fractionation during catabolism. While oxidation of glucose yields ΔG0’ = -2.870 kJ mol-1

394

while anaerobic fermentation to CO2 and CH4 liberates only ΔG0’ = -390 kJ mol-1 (67). Hence

395

it can be assumed that under anaerobe conditions most of the available carbon is used for

396

catabolism and only a small portion can be used for anabolism. Therefore it is unlikely that

397

the bias caused by the fractionation into the biomass will impact on the apparent fractionation

398

of the catabolic reactions.

399

Another way to rationalize the fractionation into the biomass becomes apparent when we

400

compare different substrates (Table 2). The strongest fractionation can be found for C1

401

compounds (CO2, methanol) (-12.9 to -23 ‰). In comparison the biomass of organisms using

402

acetate ranges from 1.4 to -6.5 ‰; while the organisms utilizing larger molecules like glucose

403

have almost no discriminatory power regarding the different carbon isotopes (-0.4 to -3-3 ‰).

404

The isotope signal of organic carbon in environmental samples is rather constant; as is the

405

signal in important carbon pools (65). Usually the small differences observed can be

406

explained the small fractionation factors reported for fermentative processes (catabolic as well

407

as anabolic). On the other hand a strong fractionation into biomass has been reported for

408

methanogens using end point measurements (22), and was observed for acetogens grown on

409

H2/CO2 (this study). Indeed our results suggest that fractionation of autotrophic organisms

410

into their respective biomass is relatively strong and may in principle affect the organic

411

carbon pool in environmental samples. However the microbial communities responsible for

412

e.g. methanogenesis usually comprise only a few percent of the total microbial community

413

and hence it may be difficult to follow the isotopic differences imprinted by C1 metabolisms

414

in environmental settings.

415 416

Acknowledgements

417

The authors thank Ralf Conrad (MPI Marburg) for critically reading the manuscript. We

418

would like to thank three anonymous referees for valuable comments. We thank Rolf Thauer

419

(MPI Marburg) and Wolfgang Buckel (MPI Marburg) for helpful discussions on the isotopic

420

fractionation into biomass.

421

References

422

1.

423 424

Drake HL, Gößner AS, Daniel SL. 2008. Old acetogens, new light. Annals of the New York Academy of Sciences 1125:100-128.

2.

Drake HL, Kuesel K, Matthies C. 2002. Ecological consequences of the

425

phylogenetic and physiological diversities of acetogens. Antonie Van Leeuwenhoek

426

International Journal of General and Molecular Microbiology 81:203-213.

427

3.

Drake HL, Küsel K, Matthies C. 2006. Acetogenic prokaryotes, p. 354-420. In

428

Dworkin M, Falkow S, Rosenberg E, Schleifer K-H, Stackebrandt E (ed.), The

429

Prokaryotes; An Evolving Electronic Resource for Microbial Community, 3 ed, vol. 2.

430

Springer.

431

4.

432 433

fixation pathways. Applied and Environmental Microbiology 77:1925-1936. 5.

434 435

Huegler M, Sievert SM. 2011. Beyond the calvin cycle: autotrophic carbon fixation in the ocean. Annual Review of Marine Science 3:261-289.

6.

436 437

Berg IA. 2011. Ecological aspects of the distribution of different autotrophic CO2

Wood HG. 1952. A study of carbon dioxide fixation by mass determination of the types of C-13-acetate. Journal of Biological Chemistry 194:905-931.

7.

Schuchmann K, Muller V. 2014. Autotrophy at the thermodynamic limit of life: a

438

model for energy conservation in acetogenic bacteria. Nature Reviews Microbiology

439

12:809-821.

440

8.

Ragsdale SW, Pierce E. 2008. Acetogenesis and the Wood-Ljungdahl pathway of

441

CO2 fixation. Biochimica Et Biophysica Acta-Proteins and Proteomics 1784:1873-

442

1898.

443 444

9.

Conrad R. 2005. Quantification of methanogenic pathways using stable carbon isotopic signatures: a review and a proposal. Organic Geochemistry 36:739-752.

445

10.

Elsner M, Zwank L, Hunkeler D, Schwarzenbach RP. 2005. A new concept

446

linking observable stable isotope fractionation to transformation pathways of organic

447

pollutants. Environmental Science & Technology 39:6896-6916.

448

11.

449 450

Hayes JM. 1993. Factors controlling C-13 contents of sedimentary organiccompounds - principles and evidence. Marine Geology 113:111-125.

12.

Mook WG, Bommerso JC, Staverma WH. 1974. Carbon isotope fractionation

451

between dissolved bicarbonate and gaseous carbon-dioxide. Earth and Planetary

452

Science Letters 22:169-176.

453

13.

Myrttinen A, Becker V, Barth J. 2012. A review of methods used for equilibrium

454

isotope fractionation investigations between dissolved inorganic carbon and CO2.

455

Earth-Science Reviews 115:192-199.

456

14.

457 458

Northrop DB. 1981. The expression of isotope effects on enzyme-catalyzed reactions. Annual Review of Biochemistry 50:103-131.

15.

Ohkouchi NO, N.O. Chikaraishi, Y.; Tanaka, H.; Wada, E. 2015. Biochemical and

459

physiological bases for the use of carbon and nitrogen isotopes in environmental and

460

ecological studies. Progress in Earth and Planetary Science 2:2-17.

461

16.

Wing BA, Halevy I. 2014. Intracellular metabolite levels shape sulfur isotope

462

fractionation during microbial sulfate respiration. Proceedings of the National

463

Academy of Sciences of the United States of America 111:18116-18125.

464

17.

Mariotti A, Germon JC, Hubert P, Kaiser P, Letolle R, Tardieux A, Tardieux P.

465

1981. Experimental-determination of nitrogen kinetic isotope fractionation - some

466

principles - illustration for the denitrification and nitrification processes. Plant and Soil

467

62:413-430.

468

18.

Botz R, Pokojski HD, Schmitt M, Thomm M. 1996. Carbon isotope fractionation

469

during bacterial methanogenesis by CO2 reduction. Organic Geochemistry 25:255-

470

262.

471

19.

Games LM, Hayes JM, Gunsalus P. 1978. Methane-producing bacteria: natural

472

fractionations of the stable carbon isotopes. Geochimica et Cosmochimica Acta

473

42:1295-1297.

474

20.

Penning H, Plugge CM, Galand PE, Conrad R. 2005. Variation of carbon isotope

475

fractionation in hydrogenotrophic methanogenic microbial cultures and environmental

476

samples at different energy status. Global Change Biology 11:2103-2113.

477

21.

Valentine DL, Chidthaisong A, Rice A, Reeburgh WS, Tyler SC. 2004. Carbon

478

and hydrogen isotope fractionation by moderately thermophilic methanogens.

479

Geochimica et Cosmochimica Acta 68:1571-1590.

480

22.

Londry KL, Dawson KG, Grover HD, Summons RE, Bradley AS. 2008. Stable

481

carbon isotope fractionation between substrates and products of Methanosarcina

482

barkeri. Organic Geochemistry 39:608-621.

483

23.

Penger J, Conrad R, Blaser M. 2012. Stable carbon isotope fractionation by

484

methylotrophic methanogenic archaea. Applied and Environmental Microbiology

485

78:7596-7602.

486

24.

487 488

Gelwicks JT, Risatti JB, Hayes JM. 1994. Carbon isotope effects associated with aceticlastic methanogenesis. Applied and Environmental Microbiology 60:467-472.

25.

Goevert D, Conrad R. 2009. Effect of substrate concentration on carbon isotope

489

fractionation during acetoclastic methanogenesis by Methanosarcina barkeri and M.

490

acetivorans and in rice field soil. Applied and Environmental Microbiology 75:2605-

491

2612.

492

26.

493 494

House CH, Schopf JW, Stetter KO. 2003. Carbon isotopic fractionation by Archaeans and other thermophilic prokaryotes. Organic Geochemistry 34:345-356.

27.

Penning H, Claus P, Casper P, Conrad R. 2006. Carbon isotope fractionation

495

during acetoclastic methanogenesis by Methanosaeta concilii in culture and a lake

496

sediment. Applied and Environmental Microbiology 72:5648-5652.

497

28.

Conrad R, Claus P, Casper P. 2010. Stable isotope fractionation during the

498

methanogenic degradation of organic matter in the sediment of an acidic bog lake,

499

Lake Grosse Fuchskuhle. Limnology and Oceanography 55:1932-1942.

500

29.

Conrad R, Ji Y, Noll M, Klose M, Claus P, Enrich-Prast A. 2014. Response of the

501

methanogenic microbial communities in Amazonian oxbow lake sediments to

502

desiccation stress. Environmental Microbiology 16:1682-1694.

503

30.

Conrad R, Klose M, Claus P, Enrich-Prast A. 2010. Methanogenic pathway, (13)C

504

isotope fractionation, and archaeal community composition in the sediment of two

505

clear-water lakes of Amazonia. Limnology and Oceanography 55:689-702.

506

31.

Conrad R, Klose M, Lu Y, Chidthaisong A. 2012. Methanogenic pathway and

507

archaeal communities in three different anoxic soils amended with rice straw and

508

maize straw. Frontiers in microbiology 3:1-12 DOI 10.3389/fmicb.2012.00004.

509

32.

Daebeler A, Gansen M, Frenzel P. 2013. Methyl fluoride affects methanogenesis

510

rather than community composition of methanogenic archaea in a rice field soil. Plos

511

One 8:e53656.

512

33.

Janssen PH, Frenzel P. 1997. Inhibition of methanogenesis by methyl fluoride:

513

Studies of pure and defined mixed cultures of anaerobic bacteria and archaea. Applied

514

and Environmental Microbiology 63:4552-4557.

515

34.

Blaser MB, Dreisbach LK, Conrad R. 2013. Carbon isotope fractionation of 11

516

acetogenic strains grown on H2 and CO2. Applied and Environmental Microbiology

517

79:1787-1794.

518

35.

519 520

Gelwicks JT, Risatti JB, Hayes JM. 1989. Carbon isotope effects associated with autotrophic acetogenesis. Organic Geochemistry 14:441-446.

36.

Preuss A, Schauder R, Fuchs G, Stichler W. 1989. Carbon isotope fractionation by

521

autotrophic bacteria with 3 different CO2 fixation pathways. Zeitschrift fur

522

Naturforschung C-A Journal of Biosciences 44:397-402.

523

37.

Blair N, Leu A, Munoz E, Olsen J, Kwong E, Des Marais DJ. 1985. Carbon

524

isotopic fractionation in heterotrophic microbial-metabolism. Applied and

525

Environmental Microbiology 50:996-1001.

526

38.

Botsch KC, Conrad R. 2011. Fractionation of stable carbon isotopes during

527

anaerobic production and degradation of propionate in defined microbial cultures.

528

Organic Geochemistry 42:289-295.

529

39.

Penning H, Conrad R. 2006. Carbon isotope effects associated with mixed-acid

530

fermentation of saccharides by Clostridium papyrosolvens. Geochimica et

531

Cosmochimica Acta 70:2283-2297.

532

40.

Zhang CLL, Ye Q, Reysenbach AL, Gotz D, Peacock A, White DC, Horita J,

533

Cole DR, Fong J, Pratt L, Fang JS, Huang YS. 2002. Carbon isotopic fractionations

534

associated with thermophilic bacteria Thermotoga maritima and Persephonella

535

marina. Environmental Microbiology 4:58-64.

536

41.

537 538

enzyme from Clostridium thermoaceticum. Methods Enzymol 53:360-372. 42.

539 540

Ljungdahl LG, Andreesen JR. 1978. Formate dehydrogenase, a selenium-tungsten

Ragsdale SW. 2008. Enzymology of the Wood-Ljungdahl pathway of acetogenesis. Annals of the New York Academy of Sciences 1125:129-136.

43.

Fontaine FE, Peterson WH, Mccoy E, Johnson MJ, Ritter GJ. 1942. A new type

541

of glucose fermentation by Clostridium thermoaceticum n sp. Journal of Bacteriology

542

43:701-715.

543

44.

Frostl JM, Seifritz C, Drake HL. 1996. Effect of nitrate on the autotrophic

544

metabolism of the acetogens Clostridium thermoautotrophicum and Clostridium

545

thermoaceticum. Journal of Bacteriology 178:4597-4603.

546

45.

Arendsen AF, Soliman MQ, Ragsdale SW. 1999. Nitrate-dependent regulation of

547

acetate biosynthesis and nitrate respiration by Clostridium thermoaceticum. Journal of

548

Bacteriology 181:1489-1495.

549

46.

Seifritz C, Daniel SL, Gossner A, Drake HL. 1993. Nitrate as a preferred electron

550

sink for the acetogen Clostridium thermoaceticum. Journal of Bacteriology 175:8008-

551

8013.

552

47.

Pearson A. 2010. Pathways of carbon assimilation and their impact on organic matter

553

values δ13C, p. 143-156. In Timmis KN (ed.), Handbook of Hydrocarbon and Lipid

554

Microbiology.

555

48.

Zyakun AM, Kochetkov VV, Baskunov BP, Zakharchenko VN, Peshenko VP,

556

Laurinavichius KS, Anokhina TO, Siunova TV, Sizova OI, Boronin AM. 2013.

557

Use of glucose and carbon isotope fractionation by microbial cells immobilized on

558

solid-phase surface. Microbiology 82:280-289.

559

49.

Hunkeler D, Aravena R. 2000. Evidence of substantial carbon isotope fractionation

560

among substrate, inorganic carbon, and biomass during aerobic mineralization of 1,2-

561

dichloroethane by Xanthobacter autotrophicus. Applied and Environmental

562

Microbiology 66:4870-4876.

563

50.

564 565

Journal of Biological Chemistry 238:2882-2886. 51.

566 567

Wolin EA, Wolin MJ, Wolfe RS. 1963. Formation of methane by bacterial extracts.

Brand WA. 1996. High precision isotope ratio monitoring techniques in mass spectrometry. Journal of Mass Spectrometry 31:225-235.

52.

Krummen M, Hilkert AW, Juchelka D, Duhr A, Schluter HJ, Pesch R. 2004. A

568

new concept for isotope ratio monitoring liquid chromatography/mass spectrometry.

569

Rapid Communications in Mass Spectrometry 18:2260-2266.

570

53.

571 572

Hayes JM. 2002. Practice and principles of isotopic measurements in organic geochemistry. web.gps.caltech.edu/~als/research_articles/other_stuff/hayespnp.pdf.

54.

Hayes JM. 2004. Isotopic order, biogeochemical processes, and earth history -

573

Goldschmidt Lecture, Davos, Switzerland, August 2002. Geochimica et

574

Cosmochimica Acta 68:1691-1700.

575

55.

576 577

567. 56.

578 579

Oleary MH. 1981. Carbon isotope fractionation in plants. Phytochemistry 20:553-

Bowling DR, Pataki DE, Randerson JT. 2008. Carbon isotopes in terrestrial ecosystem pools and CO2 fluxes. New Phytologist 178:24-40.

57.

Schoeller DA, Klein PD, Watkins JB, Heim T, Maclean WC. 1980. C-13

580

abundances of nutrients and the effect of variations in C-13 isotopic abundances of test

581

meals formulated for (CO2)-C-13 breath tests. American Journal of Clinical Nutrition

582

33:2375-2385.

583

58.

Krzycki JA, Kenealy WR, Deniro MJ, Zeikus JG. 1987. Stable carbon isotope

584

fractionation by Methanosarcina barkeri during methanogenesis from acetate,

585

methanol, or carbon dioxide-hydrogen. Applied and Environmental Microbiology

586

53:2597-2599.

587

59.

Sim MS, Ono S, Donovan K, Templer SP, Bosak T. 2011. Effect of electron donors

588

on the fractionation of sulfur isotopes by a marine Desulfovibrio sp. Geochimica et

589

Cosmochimica Acta 75:4244-4259.

590

60.

Ono S, Sim MS, Bosak T. 2014. Predictive isotope model connects microbes in

591

culture and nature. Proceedings of the National Academy of Sciences of the United

592

States of America 111:18102-18103.

593

61.

Blaser MB, Dreisbach LK, Conrad R. 2015. Carbon isotope fractionation of

594

Thermoanaerobacter kivui in different growth media and at different total inorganic

595

carbon concentration. Organic Geochemistry 81:45-52.

596

62.

597 598 599

Vandermeijden P, Vanderdrift C, Vogels GD. 1984. Methanol conversion in Eubacterium limosum. Archives of Microbiology 138:360-364.

63.

Hattori S. 2008. Syntrophic acetate-oxidizing microbes in methanogenic environments. Microbes and Environments 23:118-127.

600

64.

Hattori S, Kamagata Y, Hanada S, Shoun H. 2000. Thermacetogenium phaeum

601

gen. nov., sp nov., a strictly anaerobic, thermophilic, syntrophic acetate-oxidizing

602

bacterium. International Journal of Systematic and Evolutionary Microbiology

603

50:1601-1609.

604

65.

Conrad R, Claus P, Chidthaisong A, Lu Y, Fernandez Scavino A, Liu Y, Angel R,

605

Galand PE, Casper P, Guerin F, Enrich-Prast A. 2014. Stable carbon isotope

606

biogeochemistry of propionate and acetate in methanogenic soils and lake sediments.

607

Organic Geochemistry 73:1 - 7

608

66.

Heuer VB, Krueger M, Elvert M, Hinrichs KU. 2010. Experimental studies on the

609

stable carbon isotope biogeochemistry of acetate in lake sediments. Organic

610

Geochemistry 41:22-30.

611 612 613 614

67.

Schink B. 1997. Energetics of syntrophic cooperation in methanogenic degradation. Microbiology and Molecular Biology Reviews 61:262-280.

615 616

617 618

Table 2 average fractionation into biomass for different organismic groups and substrates. Given are average values ± DS. Details can be found in Table S1. organismic group Acetogen Methanogen Sulfate reducer Phototroph Methanogen Methanogen Sulfate reducer Fermentating bacterium Acetogen

619

Substrate CO2 CO2 CO2 CO2 methanol acetate acetate glucose glucose

Δsubstrate-biomass -12.9 ± 5.0 -18.5 ± 6.3 -20.4 ± 13.9 -23.0 ± 8.4 -20.5 ± 11.9 1.4 ± 6.7 -6.5 ± 1.3 -0.4 ± 1.9 -3.3 ± 3.0

n = 18 n = 23 n=5 n=8 n=5 n=4 n=2 n=6 n=6

620 621 622

Figure 1. Schematic representation of the reductive acetyl CoA pathway and entering points

623

of possible substrates. In black the reduction of CO2 to either the methyl or the carboxyl-

624

group is highlighted; the reverse reaction (oxidation) is given in grey arrows. For simplicity,

625

reducing-equivalents are given as H2. The entrance points of different C1 substrates is given

626

in dark grey. Note that the carboxyl-group can be formed by reduction of CO2 or indirectly by

627

oxidation of the respective C1-compound. The enzymes acting on the different C1 compounds

628

are listed: Methanol-cobalamin methyltransferase (1); CO dehydrogenase / acetyl-CoA

629

synthase (2); formate dehydrogenase (3); formyl-tetrahydrofolate synthetase (4). Details on

630

the reactions are given in equations 1 – 6 and in the text.

631 632

10 0 -10

delta 13C

-20 -30 -40 -50 -60 -70 -80 0

633

5

10

15

20 25 time (days)

30

35

40

CO2 (H2/CO2)

acetate (H2/CO2)

CO2 (methanol)

acetate (methanol)

methanol (methanol)

CO2 (formate)

acetate (formate)

formate (formate)

45

634

Figure 2. Sporomusa sphaeroides DSM 2875 grown on different C1 compounds (substrates

635

are given in brackets behind the measured analyte as: H2/CO2, formate or methanol). Depicted

636

is the 13C value of substrate (black filled symbols) and product (acetate; open symbols); CO2

637

for incubations with methanol or formate are given as open grey symbols. Values are given as

638

average of three independent replicates ± standard deviation. Individual plots for the different

639

substrates are given as Fig. S.1

640

20 10 0

delta 13C [‰]

-10 -20 -30 -40 -50 -60 -70 0

5

10

15

20

25

30

time (days) H₂/CO₂ (CO₂)

641

H₂/CO₂ (Ac)

low mix (CO₂)

high mix (CO₂)

formate (CO₂)

low mix (Form)

high mix (Form)

formate (Form)

low mix (Ac)

high mix (Ac)

formate (Ac)

642

Figure 3. Mixotrophic growth of Thermoanaerobacter kivui DSM 2030 on formate and

643

H2/CO2 leads to incorporation of both substrates. The measured analytes CO2, formate (Form)

644

or acetate (Ac) are given in brackets. H2/CO2 concentrations were 2.5 mmol/bottle (H2) and

645

0.6 mmol/bottle (CO2) in the H2/CO2 treatment as well as in the low and high mix treatment.

646

Formate concentrations were 0.25 mmol/bottle in the low mix and 1 mmol/bottle in the high

647

mix and formate treatment respectively. The δ13C value of the released acetate is a direct

648

result from the mixing of the δ13C values of the substrates while the fractionation of the

649

individual substrates is quite constant. Values are given as average of three independent

650

replicates ± SD.

651

10

A

0

delta 13C (‰)

-10 -20 -30 -40 -50 -60

CO₂ (H₂/CO₂)

Acetate (H₂/CO₂)

Glucose (Glu)

Acetate (Glu)

Glucose (Glu + NO₃)

Acetate (Glu+NO₃)

-70 0

5

10

15

20

25

time (days)

652 20

B

delta 13C of Substrate (‰)

15 10

H2/CO2

ɛ=-55.8‰

Glu

ɛ=-18.5‰

Glu + NO3 ɛ=-0.4‰

5 0 -5 -10 -15 -0.6

653

-0.5

-0.4

-0.3

-0.2

-0.1

0.0

ln(1-f)

654

Figure 4. Moorella thermoacetica DSM 2955 grown on H2/CO2 () or glucose () or

655

glucose with nitrate (). A) δ13C values of substrate (CO2 and glucose) and product (acetate)

656

data. The different treatments H2/CO2, glucose (Glu) and glucose plus nitrate (Glu + NO3) are

657

given in brackets. B) regression analysis of the respective incubations. The δ13C value of the

658

acetate reflects the stoichiometry of the producing pathway: e.g. on glucose 2/3 of the

659

released acetate originate from glycolysis and decarboxylation (glucose + nitrate ɛ = -0.4 ‰)

660

and 1/3 from the acetyl-CoA pathway (H2/CO2 ɛ = -55.8 ‰); the coordinated action of both

661

pathways results in a mixed fractionation reflecting the contribution of both pathways: ɛ = -

662

18.5 ‰.

20

A

10

delta 13C (‰)

0 -10 -20 -30 -40 -50 -60 0

5

10

15

20

25

30

35

time (days)

663 10

glucose (Glu)

acetate (Glu)

cells (Glu)

CO₂ (H₂/CO₂)

acetate (H₂/CO₂)

cells (H₂/CO₂)

B

y = -48.569x - 14.03 R² = 0.9959

5 0 delta 13C (‰)

-5

y = -28.559x - 14.459 R² = 0.9599

-10 -15 -20 y = 4.2214x - 26.207 R² = 0.9755

-25 -30

y = 2.8678x - 26.153 R² = 0.9732

-35 -1.2

-1

-0.8

-0.6

-0.4

-0.2

0

ln(1-f)

664 665

Figure 5. Fractionation into biomass: A) Delta 13C of Thermoanaerobacter kivui grown on

666

either H2/CO2 or glucose n = 3. Substrate is given in brackets. B) Regression analysis of CO2

667

conversion into acetate  or biomass  for cells grown on H2/CO2 and regression analysis of

668

glucose into acetate  or biomass for cells grown on glucose.

Carbon Isotope Fractionation during Catabolism and Anabolism in Acetogenic Bacteria Growing on Different Substrates.

Homoacetogenic bacteria are versatile microbes that use the acetyl coenzyme A (acetyl-CoA) pathway to synthesize acetate from CO2 and hydrogen. Likewi...
1MB Sizes 0 Downloads 11 Views