stable carbon isotope biogeochemistry of propionate and acetate in methanogenic soils and lake...

29
Accepted Manuscript Stable carbon isotope biogeochemistry of propionate and acetate in methano- genic soils and lake sediments R. Conrad, P. Claus, A. Chidthaisong, Y. Lu, A. Fernandez Scavino, Y. Liu, R. Angel, P.E. Galand, P. Casper, F. Guerin, A. Enrich-Prast PII: S0146-6380(14)00104-1 DOI: http://dx.doi.org/10.1016/j.orggeochem.2014.03.010 Reference: OG 3116 To appear in: Organic Geochemistry Received Date: 22 November 2013 Revised Date: 11 March 2014 Accepted Date: 12 March 2014 Please cite this article as: Conrad, R., Claus, P., Chidthaisong, A., Lu, Y., Fernandez Scavino, A., Liu, Y., Angel, R., Galand, P.E., Casper, P., Guerin, F., Enrich-Prast, A., Stable carbon isotope biogeochemistry of propionate and acetate in methanogenic soils and lake sediments, Organic Geochemistry (2014), doi: http://dx.doi.org/10.1016/ j.orggeochem.2014.03.010 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Upload: a

Post on 30-Dec-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Accepted Manuscript

Stable carbon isotope biogeochemistry of propionate and acetate in methano-

genic soils and lake sediments

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

Angel, P.E. Galand, P. Casper, F. Guerin, A. Enrich-Prast

PII: S0146-6380(14)00104-1

DOI: http://dx.doi.org/10.1016/j.orggeochem.2014.03.010

Reference: OG 3116

To appear in: Organic Geochemistry

Received Date: 22 November 2013

Revised Date: 11 March 2014

Accepted Date: 12 March 2014

Please cite this article as: Conrad, R., Claus, P., Chidthaisong, A., Lu, Y., Fernandez Scavino, A., Liu, Y., Angel,

R., Galand, P.E., Casper, P., Guerin, F., Enrich-Prast, A., Stable carbon isotope biogeochemistry of propionate and

acetate in methanogenic soils and lake sediments, Organic Geochemistry (2014), doi: http://dx.doi.org/10.1016/

j.orggeochem.2014.03.010

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers

we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and

review of the resulting proof before it is published in its final form. Please note that during the production process

errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1

Stable carbon isotope biogeochemistry of propionate and acetate in

methanogenic soils and lake sediments

R. Conrada*, P. Clausa, A. Chidthaisongb, Y. Luc, A. Fernandez Scavinod, Y. Liue, R. Angela, 1, P.E. Galandf, P. Casperg, F. Guerinh, A. Enrich-Prasti

a Max-Planck-Institute for Terrestrial Microbiology, Karl-von-Frisch-Str.10, 35043 Marburg,

Germany b Joint Graduate School of Energy and Environment, King Mongkut´s University of

Technology Thonburi, and Center for Energy Technology and Environment, Ministry of Education, Bangkok, Thailand

c College of Resources and Environmental Sciences, China Agricultural University, Beijing, China

d Departamento de Biosciencias, Facultad de Quimica, Universidad de Republica, General Flores 2124, Montevideo, Uruguay

e Key Laboratory of Tibetan Environment Changes and Land Surface Processes, Institute of Tibetan Plateau Research, Chinese Academy of Science, 100085 Beijing, China

f UPMC, Université Paris 06, and CNRS, UMR 8222, Laboratoire d’Ecogéochimie des Environnements Benthiques-LECOB, Observatoire Océanologique de Banyuls, Banyuls/mer 66650, France

g Leibniz-Institute of Freshwater Ecology and Inland Fisheries, Department of Experimental Limnology, alte Fischerhuette 2, 16775 Stechlin, Germany

h Géosciences Environnement Toulouse (GET), Observatoire Midi Pyréneés, Université de Toulouse, CNRS, IRD, 14 avenue E. Belin, 31400 Toulouse, France

i Biogeochemistry Laboratory, Institute of Biology, Federal University of Rio de Janeiro, Rio de Janeiro, Brazil

1 Present address: University of Vienna, Division of Microbial Ecology, Althanstr.14, 1090

Vienna, Austria ([email protected]). *Corresponding author: Tel.: +49-6421-178801; fax : +49-6421-178809. E mail address: [email protected] (Ralf Conrad)

2

ABSTRACT

In anoxic environments, degradation of organic matter (OM) results in strong fractionation of

carbon isotopes, with formation of 13C-depleted CH4. Propionate and acetate are important

products of OM fermentation. Propionate is further fermented to acetate. Acetate is a direct

precursor of CH4, the remainder usually being produced from H2-mediated CO2 reduction.

There is a paucity of data for the turnover of acetate and, even more so, propionate. We

therefore analyzed the δ13C values of organic carbon, propionate, acetate and the methyl (Me)

group of acetate (acetate-Me) during the production of CH4 in anoxic incubations of various

flooded and non-flooded soils and various lake sediments. Incubation in the presence of

methylfluoride (CH3F), which inhibits CH4 production from acetate, allowed exclusion of

isotope effects during aceticlastic methanogenesis. Despite the variation inherent in the wide

diversity of sample type and origin, the data collectively showed that the δ13C value of acetate

was only marginally different (-2 ± 5‰) from that of OM, while propionate was depleted in

13C relative to total acetate (-6 ± 5‰). Acetate-Me was generally depleted in 13C relative to

total acetate (-8 ± 5‰). Thus, isotopic enrichment factors during the degradation of OM to

total propionate and acetate were much smaller than those during hydrogenotrophic and

aceticlastic methanogenesis or the intramolecular difference in δ13C between the carboxyl

(CO2H) and Me of acetate, so that the δ13C value of OM may be used as a proxy when data for

acetate are not available.

Key words: acetate; propionate; methane; organic matter; fermentation; isotope fractionation;

anoxic soil; lake sediment

3

1. Introduction

The cycling of carbon is a crucial process in ecosystems. The essential processes are

primary production and degradation of organic matter (OM), the latter being catalyzed mainly

by microorganisms. Degradation also occurs in anoxic environments. There, OM is degraded

by a complex microbial community (Conrad and Frenzel, 2002; Megonigal et al., 2003; Zinder,

1993). If there is sufficient supply of oxidants such as SO42- (in marine environments) or Fe3+

(in a few soils and freshwater environments), OM is eventually degraded to CO2 as sole

product. However, when the environment is depleted in SO42- and Fe3+, OM is degraded to

CH4 plus CO2 in most anoxic environments (and to some extent marine ones). The formation

of CH4 is the last step in the degradation, while the initial steps, hydrolysis and fermentation of

the polymeric OM, are mainly identical to those operating when SO42- or Fe3+ are available. In

these initial fermentation steps, acetate is always formed and frequently accumulates, at least

temporarily, in the environment. Besides acetate, propionate is also an important intermediate

in OM breakdown (Boschker et al., 2001; Glissmann and Conrad, 2000; Heuer et al., 2010).

Therefore, the turnover of acetate and the turnover of propionate are important for the

anaerobic degradation process and the formation of CH4 and CO2.

Methanogenic degradation of OM involves several steps that are performed by different

groups of microorganisms (McInerney and Bryant, 1981; Zinder, 1993). Degradation is

initiated by hydrolysis of polymers to oligomers and then to monomers, which become

fermentatively degraded to CO2, H2 and acetate, as well as other simple carbon compounds

such as fatty acids (e.g. propionate) and alcohols. The latter are further fermented (secondary

fermentation) to acetate, CO2 and H2, which are then finally converted to CH4. Furthermore,

acetate is not only a product of these fermentation reactions, but can also be the sole product

of fermentation and can also be formed by chemolithotrophic acetogenesis converting H2 and

CO2 to acetate (Dolfing, 1988; Drake and Küsel, 2003). Overall, acetate often contributes

4

more than two thirds of total CH4 production, the remainder being produced from the

reduction of CO2 (Conrad, 1999).

Each of these steps theoretically exhibits a characteristic stable isotope fractionation.

Production of CH4 is characterized by strong fractionation, such that CH4 is always depleted in

13C relative to the substrate from which it is formed (Whiticar et al., 1986). Methanogenesis

from acetate (aceticlastic) or from CO2 plus H2 (hydrogenotrophic) are the two most common

methanogenic pathways in freshwater sediments and anoxic soils over a wide temperature

range (Conrad et al., 2009c; Nozhevnikova et al., 2007). Since they exhibit quite distinct

carbon isotope fractionation factors, the isotopic signature (δ13C) of CH4 as compared with

those of the substrates (CO2, acetate) allows determination of the relative contribution of

aceticlastic vs. hydrogenotrophic methanogenesis to total CH4 production using mass balance

calculations (Conrad, 2005). Isotopic signatures of CH4 and CO2 in environmental samples and

microbial enrichment cultures have frequently been reported in the literature, with many

different types of OM as primary substrates, including hydrocarbons and long-chain fatty

acids (Feisthauer et al., 2010; Morris et al., 2012). However, while the precise analysis of the

carbon isotopic composition of gaseous CO2 and CH4 and of bulk OM has become routine

through the use of isotope ratio mass spectrometry (IRMS) and/or laser spectrometry, the

corresponding analysis of acetate (or propionate) has been carried out less frequently (Botsch

and Conrad, 2011; Heuer et al., 2009). Moreover, since CH4 is formed from the acetate-Me

during aceticlastic methanogenesis (Ferry, 1992), the δ13C value of this Me is the crucial one

for mass balance calculations. Obtaining such data is cumbersome and therefore relatively rare.

The δ13C values of acetate-Me and acetate-CO2H are often quite different (Conrad et al., 2011).

This intramolecular difference varies between environments and is likely dependent on

whether the acetate was formed mainly fermentatively or chemolithotrophically (Blair et al.,

1985; Gelwicks et al., 1989).

Fractionation factors have been studied in some detail for the final methanogenic steps

(Gelwicks et al., 1994; Goevert and Conrad, 2009; Penning et al., 2005), and to some extent

5

for chemolithotrophic acetogenesis (Blaser et al., 2013; Gelwicks et al., 1989). The factors are

quite substantial (ca. -10 to -90‰). However, data on carbon isotope fractionation during

primary or secondary fermentation are scarce (Blair et al., 1985; Botsch and Conrad, 2011;

Penning and Conrad, 2006). Nevertheless, the few data on fermentation reactions suggest that

fractionation factors are small (< 5‰).

The objectives of the present study were to evaluate the extent to which (i) the δ13C value

of acetate-Me and (ii) of propionate differ from those of total acetate and bulk OM in various

soils and sediments. We collected δ13C data for propionate, acetate and acetate-Me obtained

during anaerobic incubation of soil and sediment samples with different OM δ13C values. The

incubations had primarily been done for studying microbial CH4 production, in particular for

quantifying the methanogenic pathway. However, propionate and acetate had also been

analyzed routinely at the end of incubation. The δ13C values of propionate and acetate were

also analyzed in incubations in which aceticlastic methanogenesis was inhibited by CH3F

(Janssen and Frenzel, 1997), thereby avoiding effects from isotope fractionation during acetate

consumption.

2. Material and methods

Samples of freshwater sediments and soils were from the sites listed in Table 1, as

described in the quoted references. The sites included sediments and soils from different

places around the world, including tropical (Brazil, Thailand, Cameroon), temperate (Germany,

China, Uruguay), arid and semi-arid (USA, Israel, Australia), high altitude (Tibet) and boreal

(Finland) sites. Table 1 lists the δ13C analyses performed on the samples. The sediment and

soil samples had been used for determination of rates and pathways of CH4 production. For

this purpose, δ13C values of CH4, CO2 and acetate were measured, and sometimes also of

acetate-Me. The published δ13C data were collected from the references (Table 1). δ13C values

of propionate having not been published. Measurements from three sample series (Amazon

6

2010 and Lagunes, Brazil) and Cameroon have not been published. The Amazon 2010

sediment samples were from lakes with white water or clear water, as described for Amazon

2009. The lagoon samples were from coastal lagoons in the “Parque Nacional da Restinga de

Jurubatida e do Municipio de Macae” (Furtado et al., 2002). Three soils with a different

degree of moisture content were sampled from sites at Nsimi, a small experimental watershed

in the south of Cameroon (Braun et al., 2012).

The incubations and analysis were carried out as described in the references in Table 1 (e.g.

Conrad et al., 2007). Briefly, samples were placed in glass tubes (28 ml) or glass bottles (60-

120 ml) closed with butyl rubber stoppers and flushed with N2 (usually flushing the headspace

of the vessels or using several rounds of evacuation and gassing). In parallel, vessels were

treated with CH3F at initial concentrations of 0.5, 1, 2 or 3% (v/v) to inhibit aceticlastic

methanogenesis. In many experiments, CH3F was applied at more than one concentration. In

some samples (Lake Dagow sediment) CH4 production was completely inhibited by addition

of bromoethanesulfonate (5, 10, 15, 20 and 30 mM) or CHCl3 (50. 100, 150, 200 and 250 µM).

All treatments were carried out in usually 3, or sometimes 4, replicates. The samples were

incubated until CH4 was produced with a constant rate. The vessels were opened and the

contents centrifuged. The supernatant was filtered through a membrane filter and used for

analysis of acetate and propionate with high-performance liquid chromatography (HPLC) with

refraction index detection. The δ13C values of acetate and propionate was determined after

combustion (C) to CO2 using isotope ratio mass spectrometry (IRMS). The HPLC-C-IRMS

system had a detection limit of ca. 5 µM and a precision of ± 0.3‰. The δ13C values of the

methyl group of acetate were determined after off-line pyrolysis. Details have been given

before (Botsch and Conrad, 2011; Conrad et al., 2007). For δ13C analysis of OM, the samples

were acidified with HCl to remove carbonate, and dried. δ13C values were measured using an

elemental analyzer coupled to an IRMS instrument (Institute for Soil Science and Forest

Nutrition (IBW) at the University of Göttingen, Germany).

7

3. Results

Rice field soils from Italy, China and Thailand were incubated in the presence of either rice

straw or maize straw and exhibited δ13C values characteristic of C3 (-31.0 ± 0.1‰) and C4 (-

13.0 ± 0.1‰) pathways, respectively (Fig. 1). The median values for total acetate were almost

identical to those of the respective straw OM, the mean values being slightly higher (-29‰

and -11‰, respectively), while the upper and lower quartiles were of the order of ± 3-5‰.

In addition to the rice field soils, a wide variety of different soils and sediments (Table 1)

were incubated under methanogenic conditions. The δ13C values of acetate plotted against

those of the OM in the respective soil and sediment samples are summarized in Fig. 2. The

data show a positive correlation between acetate and OM, the average difference (± SD) being

just -1.6 ± 5.3‰ (Fig. 2). However, the rather large variance should be noted.

The acetate was also analyzed with respect to the intramolecular difference of δ13C by

plotting the δ13C value of the Me against that of the total acetate (Fig. 3). The data included

incubations with and without CH3F, which, however, were not different. The data are

separately shown for soil samples (Fig. 3A) and lake sediments (Fig. 3B). The plots revealed

in both sets a linear relationship with a systematic offset showing that Me in soils and

sediments was on average (± SD) depleted in 13C by 6.4 ± 4.2‰ (soils) and by 11.0 ± 4.3‰

(sediments) relative to total acetate, respectively. The offset for all data together was 8.4 ±

4.8‰.

Propionate was analyzed in the incubations of rice field soils from Italy, China and

Thailand in the presence of either rice straw or maize straw (Fig. 4). It accumulated as an

intermediate reaching a maximum concentration of 2.5-9 mM, and then decreased, unless

aceticlastic methanogenesis was inhibited by CH3F. While the δ13C values of propionate

stayed relatively constant in the presence of CH3F, they slowly increased in its absence as

soon as the propionate concentration started to decrease. The δ13C values of propionate in the

non-inhibited samples increased to values that were at maximum of ca. 5-10‰ higher than for

the inhibited samples (Fig. 4). The values for propionate in the inhibited samples were similar

8

to those of the rice straw or maize straw OM. On average (median and mean nearly identical),

propionate was by 2.7 ± 0.2‰ and 5.1 ± 0.3‰ lighter than organic carbon (OC), respectively

(Fig. 1). Together with the data from all the other sediment and soil incubations, the δ13C

values of propionate increased linearly with those of total acetate (Fig. 5). We assume that the

values for propionate also increased linearly with those of OM, but many fewer data pairs

were available than for propionate and acetate. On average, propionate was depleted in 13C by

6 ± 5‰ relative to acetate.

4. Discussion

Analysis of the carbon isotopic composition in numerous samples of methanogenic

freshwater sediments and soils showed that fractionation was small during anaerobic

conversion of complex OM to acetate, propionate and CO2, while the further conversion of

these compounds to CH4 exhibited large fractionation factors (Fig. 6). The samples were

incubated under methanogenic conditions, so that propionate and acetate could only be

degraded by processes coupled to CH4 production. However, aceticlastic methanogenesis was

inhibited by the addition of CH3F (Janssen and Frenzel, 1997). This inhibitor was found to be

quite specific and does normally not inhibit sulfate-reducing, acetogenic or fermenting

bacteria (Janssen and Frenzel, 1997). Hydrogenotrophic methanogens, however, may be

partially inhibited if CH3F concentrations are too high (Conrad and Klose, 1999). This is not

critical for interpretation of the present data as long as aceticlastic methanogenesis is

completely inhibited. Therefore, inhibition experiments were usually conducted at more than

one CH3F concentration to guarantee complete inhibition of aceticlastic methanogenesis. In

methanogenic samples, aceticlastic methanogenesis is normally the sole path for acetate

degradation, unless syntrophic acetate oxidizers exist that are able to couple acetate oxidation

to hydrogenotrophic methanogenesis (Zinder and Koch, 1984). For simplicity, we assumed

that such syntrophs did not play a major role and that acetate consumption was largely

9

impeded by CH3F. Indeed, we always observed net accumulation of acetate under these

conditions (references in Table 1). Therefore, the δ13C value of acetate was probably

unaffected by acetate consumption and should represent the δ13C value of the acetate after

production from precursors derived from OM. However, it is noteworthy that isotopic

fractionation during syntrophic acetate oxidation seems to be smaller than during aceticlastic

methanogenesis and so would have only a small effect on the δ13C value of the acetate pool

(Conrad and Klose, 2011).

The δ13C value of total acetate increased linearly with that of OM, exhibiting a difference

of ca. -6‰ to +4‰ (avg. ca. -2‰; Fig. 6). Hence, total acetate was similar to OC. Acetate

produced from the OC of C4 plants was heavier than that of C3 plants. Such similarity in δ13C

values would be expected if acetate is primarily produced by fermentative production, i.e.

anaerobic heterotrophic metabolism of OM (Blair et al., 1985). In fact, there is a paucity of

isotope fractionation data from pure cultures of fermentative microorganisms (Blair et al.,

1985; Botsch and Conrad, 2011; Penning and Conrad, 2006). The data indicate that acetate

was only slightly enriched in 13C (ca. 3‰; Fig. 6). However, it cannot be excluded that other

fermenting microorganisms would exhibit greater fractionation. Nevertheless, the similarity in

δ13C values of OM and acetate in our data suggest that organisms exhibiting a large

fractionation during fermentation (if they exist) may not be of importance for OM degradation

in anoxic freshwater sediments and soils. In addition, acetate can also be produced from

chemolitotrophic acetogenesis, i.e. the reduction of CO2 to acetate. This process exhibits rather

large fractionation factors, of the order of -56‰ (Blaser et al., 2013; Gelwicks et al., 1989).

The process is dominant in the anoxic hindgut of some termites (Breznak and Switzer, 1986)

and has been suggested to occur in some anoxic soils and sediments (Conrad et al., 1989;

Hoehler et al., 1999; Nozhevnikova et al., 2007). Indeed, we do not rule out the idea that

chemolithotrophic acetogenesis plays a certain role in some of the anoxic environments

analyzed, as suggested by the large variation in the δ13C difference between OC and acetate.

10

On the other hand, our data are such that acetate formation via heterotrophic fermentation of

OM seems to be the more common path in freshwater sediments and soils.

Acetate formation via mainly heterotrophic fermentation is consistent with the observed

intramolecular δ13C difference. Heterotrophic fermentation was reported to result in a rather

large intramolecular difference [δ13Cac-methyl - δ13Cac-carboxyl = 2 (δ13Cac-methyl - δ

13Cac)] of ca.

24‰ (Blair et al., 1985). Chemolithotrophic acetogenesis, on the other hand, exhibits only a

small intramolecular difference, i.e. < 1‰ (Gelwicks et al., 1989). The difference between

acetate-Me and total acetate in the soils (6‰) was on average a little smaller than in the

sediments (11‰). Together the difference was on average -8‰ (Fig. 6), equivalent to an

intramolecular difference of ca. 16‰. These data indicate that acetate was produced mainly by

fermentation, with a minor contribution from chemolithotrophic acetogenesis which, however,

may be more significant in soils than in sediments.

Propionate was produced as an intermediate when freshwater sediments and soils were

incubated under methanogenic conditions. Inhibition of aceticlastic methanogenesis also

resulted in inhibition of propionate degradation, at least during the time of observation, as

explicitly shown for the soils from Italy, China and Thailand (Fig. 4). This observation is

plausible, since propionate is most likely degraded by secondary fermentation, which easily

becomes endergonic when acetate or H2 accumulate (Chidthaisong and Conrad, 2000; Krylova

and Conrad, 1998). When propionate was not degraded, the δ13C values of propionate stayed

constant with time, whereas propionate became slightly heavier when it was consumed. Such

an increase in δ13C indicates isotope fractionation during propionate consumption. The

fractionation was in the range -10 to -5‰, which is relatively large compared with that in a

defined culture of propionate-utilizing Syntrophobacter fumaroxidans, where it was negligible

(Botsch and Conrad, 2011). The reason for this difference is unclear and needs further study.

One reason could be the pathway of propionate utilization. Syntrophobacter utilizes

propionate via the randomizing succinate or methyl malonyl-CoA pathway (CH3CH2COO- +

3H2O � CH3COO- + HCO3- + 3H2 + H+) (Houwen et al., 1990). The same pathway was found

11

to be operative in Italian rice field soil (Krylova et al., 1997). However, in other environments

propionate may be degraded via other pathways, e.g. the non-randomizing pathway

(2CH3CH2COO- + 2H2O � 3CH3COO- + 2H2 + H+) found in Smithella spp., which proceeds

via intermediate butyrate formation (DeBok et al., 2001). Both degradation pathways are

rather complex, so that isotope fractionation during propionate depletion would barely enable

prediction of the isotopic signature of the acetate, which is only one of the possible products of

propionate degradation.

When propionate was not degraded, it was slightly depleted in 13C (ca. 4‰; Fig. 6) relative

to the straw serving as the major substrate. This relatively small fractionation was consistent

with microbial culture studies fermenting pectin, xylan or starch, which showed similar

isotope fractionation (ca. -6‰; Fig. 6) during propionate formation by Opitutus terrae (Botsch

and Conrad, 2011). Overall, fractionation during propionate formation from OM was larger

than during acetate formation, resulting in acetate that was on average ca. 6‰ heavier than

propionate (Figs. 5, 6). However, the range of difference between propionate and acetate was

rather large, indicating that the processes involved in turnover of these compounds are not

well understood.

This notion is further illustrated by the following consideration. In Italian rice field soil,

propionate turnover can account for all of the H2 driving hydrogenotrophic methanogenesis

(Krylova et al., 1997). Hydrogenotrophic methanogenesis accounts for ca. 30% of total CH4

production, the remaining 70% is due to aceticlastic methanogenesis. On a molar basis

propionate turnover produces one acetate and three H2 molecules (succinate pathway).

Therefore, if 30% of the total CH4 is from propionate-derived H2, then 40% of the CH4 must

be from propionate-derived acetate. Thus almost 60% of the total acetate production would be

caused by propionate oxidation. Such a large contribution of propionate turnover to acetate

turnover cannot be generally true, since otherwise the difference between propionate and

acetate should be < 6‰.

12

In conclusion, our study has revealed gaps of knowledge in the detail, but showed a

relatively uniform picture for isotope fractionation during production of acetate and propionate

from OM in soils and sediments sampled from around the world. In general, fractionation

during fermentation was relatively small, of the order of only a few permil (Fig. 6). Even

when taking the large variance into account, the δ13C values of acetate differed from those of

OC on average by only -2‰ (Fig. 6). Likewise, the values of the CO2 produced are not much

different from those of the OM (Conrad et al., 2012b), similar to the situation found for

aerobic CO2 production from soil OM (Werth and Kuzyakov, 2010). The fractionation factors

for the further conversion of acetate and CO2 to CH4 are much larger (Fig. 6). Hence, if δ13C

values for acetate are unknown, those of OC could be used as a proxy for mass balance

calculations. Angel et al. ( 2012) provide an example for determining the fraction of CH4

produced by hydrogenotrophic and aceticlastic methanogenesis. In such a case, the δ13C value

of acetate-Me rather that of total acetate is required. Analysis of our data showed that acetate-

Me was depleted in 13C relative to total acetate, on average by 8 ± 5‰. Again, one might use

this value as a first approximation to correct the δ13C value of total acetate, if a value for

acetate-Me were not available (Conrad et al., 2012a). If only the δ13C value of OM is available,

one might estimate the value for acetate-Me by subtracting ca. 10‰ (Angel et al., 2012). Of

course, all these corrections bear more uncertainty than direct analysis, but may nevertheless

be useful for an initial estimate.

Acknowledgements

The study is part of the ICON project financed by the German Research Foundation. We

thank the Brazilian agencies CNPq, CAPES and FAPERJ for the financial support used to

sample sediments in Brazil. The sampling in Tibet was supported by the National Natural

Science Foundation of China, the Chinese Academy of Science and the German Research

Foundation. The soils from Cameroon were sampled in the Nsimi small experimental

13

watershed (SOERE BVET: Service Observation Recherche en Environnement – Basins

Versants Expérimentaux Tropicaux funded by IRD and INSU/CNRS; http://bvet.omp.obs-

mip.fr) in the framework of the project LIMON funded by EC2CO (INSU/CNRS). P.E.G. is

supported by the Agence Nationale de la Recherche (ANR) project MICADO (ANR-11JSV7-

003-01). We thank two anonymous reviewers for helpful comments.

References Angel, R., Claus, P., Conrad, R., 2012. Methanogenic archaea are globally ubiquitous in

aerated soils and become active under wet anoxic conditions. ISME Journal 6, 847-862.

Blair, N., Leu, A., Munoz, E., Olsen, J., Kwong, E., DesMarais, D., 1985. Carbon isotopic

fractionation in heterotrophic microbial metabolism. Applied and Environmental

Microbiology 50, 996-1001.

Blaser, M. B., Dreisbach, L. K., Conrad, R., 2013. Carbon isotope fractionation of 11

acetogenic strains grown on H2 and CO2. Applied and Environmental Microbiology 79,

1787-1794.

Boschker, H. T. S., DeGraaf, W., Koster, M., Meyer-Reil, L. A., Cappenberg, T. E., 2001.

Bacterial populations and processes involved in acetate and propionate consumption in

anoxic brackish sediment. FEMS Microbiology Ecology 35, 97-103.

Botsch, K. C. and Conrad, R., 2011. Fractionation of stable carbon isotopes during anaerobic

production and degradation of propionate in defined microbial cultures. Organic

Geochemistry 42, 289-295.

Braun, J. J., Marechal, J. C., Riotte, J., Boeglin, J. L., Bedimo Bedimo, J. P., Ndam

Ngoupayou, J. R., Nyeck, B., Robain, H., Sekhar, M., Audry, S., Viers, J., 2012. Elemental

weathering fluxes and saprolite production rate in a Central African lateritic terrain (Nsimi,

South Cameroon). Geochimica et Cosmochimica Acta 99, 243-270.

Breznak, J. A. and Switzer, J. M., 1986. Acetate synthesis from H2 plus CO2 by termite gut

microbes. Applied and Environmental Microbiology 52, 623-630.

14

Chidthaisong, A. and Conrad, R., 2000. Specificity of chloroform, 2-bromoethanesulfonate

and fluoroacetate to inhibit methanogenesis and other anaerobic processes in anoxic rice

field soil. Soil Biology and Biochemistry 32, 977-988.

Conrad, R., 1999. Contribution of hydrogen to methane production and control of hydrogen

concentrations in methanogenic soils and sediments. FEMS Microbiology Ecology 28, 193-

202.

Conrad, R., 2005. Quantification of methanogenic pathways using stable carbon isotopic

signatures: a review and a proposal. Organic Geochemistry 36, 739-752.

Conrad, R., Bak, F., Seitz, H. J., Thebrath, B., Mayer, H. P., Schütz, H., 1989. Hydrogen

turnover by psychrotrophic homoacetogenic and mesophilic methanogenic bacteria in

anoxic paddy soil and lake sediment. FEMS Microbiology Ecology 62, 285-294.

Conrad, R., Chan, O. C., Claus, P., Casper, P., 2007. Characterization of methanogenic

Archaea and stable isotope fractionation during methane production in the profundal

sediment of an oligotrophic lake (Lake Stechlin, Germany). Limnology and Oceanography

52, 1393-1406.

Conrad, R., Claus, P., Casper, P., 2009a. Characterization of stable isotope fractionation

during methane production in the sediment of a eutrophic lake, Lake Dagow, Germany.

Limnology and Oceanography 54, 457-471.

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

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

Fuchskuhle. Limnology and Oceanography 55, 1932-1942.

Conrad, R. and Frenzel, P., 2002, Flooded soils. In: Bitton, G. (Ed.), Encyclopedia of

Environmental Microbiology, John Wiley & Sons, Inc., New York, pp. 1316-1333.

Conrad, R., Ji, Y., Noll, M., Klose, M., Claus, P., Enrich-Prast, A., 2013. Response of the

methanogenic microbial communityies in Amazonian oxbow lake sediments to desiccation

stress. Environmental Microbiology, in press, doi:10.1111/1462-2920.12267.

Conrad, R. and Klose, M., 1999. How specific is the inhibition by methyl fluoride of

acetoclastic methanogenesis in anoxic rice field soil? FEMS Microbiology Ecology 30, 47-

56.

15

Conrad, R. and Klose, M., 2011. Stable carbon isotope discrimination in rice field soil during

acetate turnover by syntrophic acetate oxidation or acetoclastic methanogenesis.

Geochimica et Cosmochimica Acta 75, 1531-1539.

Conrad, R., Klose, M., Claus, P., Dan, J. G., 2009b. Activity and composition of the

methanogenic archaeal community in soil vegetated with wild versus cultivated rice. Soil

Biology and Biochemistry 41, 1390-1395.

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

fractionation, and archaeal community composition in the sediment of two clearwater lakes

of Amazonia. Limnology and Oceanography 55, 689-702.

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

communities in three different anoxic soils amended with rice straw and maize straw.

Frontiers in Microbiology 3, 4, doi: 10.3389/fmicb.2012.00004.

Conrad, R., Klose, M., Noll, M., 2009c. Functional and structural response of the

methanogenic microbial community in rice field soil to temperature change. Environmental

Microbiology 11, 1844-1853.

Conrad, R., Klose, M., Yuan, Q., Lu, Y., Chidthaisong, A., 2012b. Stable carbon isotope

fractionation, carbon flux partitioning and priming effects in anoxic soils during

methanogenic degradation of straw and soil organic matter. Soil Biology and Biochemistry

49, 193-199.

Conrad, R., Noll, M., Claus, P., Klose, M., Bastos, W. R., Enrich-Prast, A., 2011. Stable

carbon isotope discrimination and microbiology of methane formation in tropical anoxic

lake sediments. Biogeosciences 8, 795-814.

DeBok, F. A. M., Stams, A. J. M., Dijkema, C., Boone, D. R., 2001. Pathway of propionate

oxidation by a syntrophic culture of Smithella propionica and Methanospirillum hungatei.

Applied and Environmental Microbiology 67, 1800-1804.

Dolfing, J., 1988, Acetogenesis. In: Zehnder, A. J. B. (Ed.), Biology of Anaerobic

Microorganisms, Wiley, New York, pp. 417-468.

Drake, H. L. and Küsel, K., 2003, How the diverse physiologic potentials of acetogens

determine their in situ realities. In: Ljungdahl, L. G., Adams, M. W., Barton, L. L., Ferry, J.

G., Johnson, M. K. (Eds.), Biochemistry and Physiology of Anaerobic Bacteria, Springer,

New York, pp. 171-190.

16

Feisthauer, S., Siegert, M., Seidel, M., Richnow, H. H., Zengler, K., Gründger, F., Krüger, M..

2010. Isotopic fingerprinting of methane and CO2 formation from aliphatic and aromatic

hydrocarbons. Organic Geochemistry 41, 482-490.Fernandez Scavino, A., Ji, Y., Pump, J.,

Klose, M., Claus, P., Conrad, R., 2013. Structure and function of the methanogenic

microbial communities in Uruguayan soils shifted between pasture and irrigated rice fields.

Environmental Microbiology 15, 2588-2602.

Ferry, J. G., 1992. Methane from acetate [Minireview]. Journal of Bacteriology 174, 5489-

5495.

Furtado, A. L. S., Casper, P., Esteves, F. A., 2002. Methanogenesis in an impacted and two

dystrophic coastal lagoons (Macae, Brazil). Brazilian Archives of Biology and Technology

45, 195-202.

Galand, P. E., Yrjälä, K., Conrad, R., 2010. Stable carbon isotope fractionation during

methanogenesis in three boreal peatland ecosystems. Biogeosciences 7, 3893-3900.

Gelwicks, J. T., Risatti, J. B., Hayes, J. M., 1989. Carbon isotope effects associated with

autotrophic acetogenesis. Organic Geochemistry 14, 441-446.

Gelwicks, J. T., Risatti, J. B., Hayes, J. M., 1994. Carbon isotope effects associated with

aceticlastic methanogenesis. Applied and Environmental Microbiology 60, 467-472.

Glissmann, K. and Conrad, R., 2000. Fermentation pattern of methanogenic degradation of

rice straw in anoxic paddy soil. FEMS Microbiology Ecology 31, 117-126.

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

fractionation during acetoclastic methanogenesis by Methanosarcina barkeri and M.

acetivorans and in rice field soil. Applied and Environmental Microbiology 75, 2605-2612.

Heuer, V. B., Krüger, M., Elvert, M., Hinrichs, K. U., 2010. Experimental studies on the stable

carbon isotope biogeochemistry of acetate in lake sediments. Organic Geochemistry 41, 22-

30.

Heuer, V. B., Pohlman, J. W., Torres, M. E., Elvert, M., Hinrichs, K. U., 2009. The stable

carbon isotope biogeochemistry of acetate and other dissolved carbon species in deep

subseafloor sediments at the northern Cascadia Margin. Geochimica et Cosmochimica Acta

73, 3323-3336.

Hoehler, T. M., Albert, D. B., Alperin, M. J., Martens, C. S., 1999. Acetogenesis from CO2 in

an anoxic marine sediment. Limnology and Oceanography 44, 662-667.

17

Houwen, F. P., Plokker, J., Stams, A. J. M., Zehnder, A. J. B., 1990. Enzymatic evidence for

involvement of the methylmalonyl-CoA pathway in propionate oxidation by

Syntrophobacter wolinii. Archives of Microbiology 155, 52-55.

Janssen, P. H. and Frenzel, P., 1997. Inhibition of methanogenesis by methyl fluoride - studies

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

Environmental Microbiology 63, 4552-4557.

Krylova, N. I. and Conrad, R., 1998. Thermodynamics of propionate degradation in

methanogenic paddy soil. FEMS Microbiology Ecology 26, 281-288.

Krylova, N. I., Janssen, P. H., Conrad, R., 1997. Turnover of propionate in methanogenic

paddy soil. FEMS Microbiology Ecology 23, 107-117.

Liu, Y., Yao, T., Gleixner, G., Claus, P., Conrad, R., 2013. Methanogenic pathways, 13C

isotope fractionation, and archaeal community composition in lake sediments and wetland

soils on the Tibetan Plateau. Journal of Geophysical Research - Biogeosciences 118, in

press, doi:10.1002/jgrg.20055.

McInerney, M. J. and Bryant, M. P., 1981, Basic principles of bioconversions in anaerobic

digestion and methanogenesis. In: Sofer, S. S., Zaborsky, O. R. (Eds.), Biomass Conversion

Processes for Energy and Fuels, Plenum, New York, pp. 277-296.

Megonigal, J. P., Hines, M. E. and Visscher, P. T., 2003, Anaerobic metabolism: linkages to

trace gases and aerobic processes. In: Schlesinger, W. H., Holland, H. D., Turekian, K. K.

(Eds.), Treatise on Geochemistry, vol. 8: Biogeochemistry, Elsevier-Pergamon, Oxford,

UK, pp. 317-424.

Morris, B. E. L., Herbst, F. A., Bastida, F., Seifert, J., VonBergen, M., Richnow, H. H.,

Suflita, J. M., 2012. Microbial interactions during residual oil and n-fatty acid metabolism

by a methanogenic consortium. Environmental Microbiololgy Reports 4, 297-306.

Nozhevnikova, A. N., Nekrasova, V., Ammann, A., Zehnder, A. J. B., Wehrli, B., Holliger, C.,

2007. Influence of temperature and high acetate concentrations on methanogenensis in lake

sediment slurries. FEMS Microbiology Ecology 62, 336-344.

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

fermentation of saccharides by Clostridium papyrosolvens. Geochimica et Cosmochimica

Acta 70, 2283-2297.

18

Penning, H. and Conrad, R., 2007. Quantification of carbon flow from stable isotope

fractionation in rice field soils with different organic matter content. Organic Geochemistry

38, 2058-2069.

Penning, H., Plugge, C. M., Galand, P. E., Conrad, R., 2005. Variation of carbon isotope

fractionation in hydrogenotrophic methanogenic microbial cultures and environmental

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

Werth, M. and Kuzyakov, Y., 2010. 13C fractionation at the root-microorganisms-soil

interface: A review and outlook for partitioning studies. Soil Biology and Biochemistry 42,

1372-1384.

Whiticar, M. J., Faber, E., Schoell, M., 1986. Biogenic methane formation in marine and

freshwater environments: CO2 reduction vs. acetate fermentation - isotopic evidence.

Geochimica et Cosmochimica Acta 50, 693-709.

Zinder, S. H., 1993, Physiological ecology of methanogens. In: Ferry, J. G. (Ed.),

Methanogenesis. Ecology, Physiology, Biochemistry and Genetics, Chapman & Hall, New

York, pp. 128-206.

Zinder, S. H. and Koch, M., 1984. Non-aceticlastic methanogenesis from acetate: acetate

oxidation by a thermophilic syntrophic coculture. Archives of Microbiology 138, 263-272.

19

Table 1: Origin of sediment and soil samples used for determination of δ13C (+, analyzed in all samples; +/-, analyzed in most samples; -, not analyzed; np, unpublished results

a 1, Conrad et al., 2011; 2, Conrad et al., 2013; 3, Galand et al., 2010; 4, Conrad et al., 2010b; 5, Liu et al., 2013; 6, Conrad et al., 2007; 7, Conrad et al., 2009a; 8, Conrad et al., 2010a; 9, Penning and Conrad, 2007; 10, Conrad et al., 2012a; 11, Conrad et al., 2009b; 12, Fernandez Scavino et al., 2013; 13, Angel et al., 2012.

Samples description δ13Corg δ

13Cprop δ13Cac δ

13Cac-methyl Referencea

Lake sediments Pantanal, Brazil Different lakes (16) + + + + 1 Amazon 2009 White water (3), black water (3), clear water lakes (3) + +/- + + 2 Amazon 2010 White water (4), clear water lakes (3) + + + + np Mires, Finland Fens (2), bog (1) + - + - 3 Amazon 2006 Clear water lakes (2; Batata, Mussura) + - + + 4 Tibetan plateau Lake sediments (7), wetlands (2) + - + - 5 Lagoons, Brazil Coastal lagoons (Cabiunas, Imboassica) at Macae, RJ + - + - np Lakes, Germany Lakes Stechlin, Dagow, Grosse Fuchskuhle + +/- + + 6, 7, 8 Soils Vercelli, Italy Rice soil, addition of rice or maize straw, no addition + + + + 9, 10 Fuyang, China Rice soil, addition of rice or maize straw + + + + 10 Suwan, Thailand Rice soil, addition of rice or maize straw + + + - 10 Hainan, China Cultivated rice, wild rice + - + + 11 Uruguay 2009 Drained rice field + + + + 12 Uruguay 2010 Fields within a pasture rice crop rotation (8), pasture (1) + + + - 12 Upland soils Germany, Israel, Cameroon, Australia, Utah + + + +/- 13, np

20

Figure legends

Fig. 1. Box plots of δ13C of total acetate (ac) and propionate (prop) measured at the end

of anaerobic incubations (with 1% CH3F) of rice field soils from Italy, China and

Thailand amended with either rice straw (RS) or maize straw (MS). Horizontal lines

indicate the δ13C of OM in MS (upper line) and RS (lower line); n = 20.

Fig. 2. Values of δ13C of total acetate vs. those of OC at the end of anaerobic incubation

of soil or sediment in the presence of CH3F. Solid and dashed lines indicate no

fractionation and ± 10‰ fractionation, respectively; mean ± SE, n = 3. Also see

supplemental data table.

Fig. 3. Values of δ13C of acetate-Me vs. those of total acetate at the end of anaerobic

incubation of (A) various soils and (B) various freshwater sediments. Incubation was with

and without CH3F. Solid line indicates no fractionation, dashed lines -10‰ and -20‰

fractionation between acetate-Me and total acetate; mean ± SE, n = 3. Also see

supplemental data table.

Fig. 4. Transient accumulation of propionate (upper panel) and δ13C values (lower panel)

of soil from (A,D) Vercelli (Italy), (B,E) Fuyang (China) and (C,F) Suwan (Thailand),

amended with either rice straw or maize straw and incubated under anoxic conditions in

the presence or absence of 1% CH3F.

Fig. 5. δ13C values of propionate vs. those of acetate at the end of anaerobic incubation of

soil or sediment with CH3F. Solid line indicates no fractionation, dashed lines -10‰ and -

20‰ fractionation between acetate-Me and total acetate; mean ± SE, n = 3. Also see

supplemental data table.

21

Fig. 6. Scheme of carbon flow and stable carbon-isotopic enrichment factors (ε) in

methanogenic environments. Bold numbers are from this study, bold black numbers refer

to data in Figs 2, 3 and 5; bold grey numbers refer to data in Fig. 1. Small black numbers

refer to representative literature data, i.e. (from top to bottom): Botsch and Conrad (2011),

Penning and Conrad (2006), Conrad et al. (2012b), Blaser et al. (2013), Conrad et al.

(2011), Goevert and Conrad (2009).

RS.pr.i RS.ac.t.i MS.pr.i MS.ac.t.i

-40

-35

-30

-25

-20

-15

-10

-5

0

5

1

3C

of

pro

pio

na

te o

r a

ce

tate

(p

erm

il)

Straw type, compound

RS, prop RS, ac MS, prop MS, ac

Fig. 1

Figure(s)

-40 -35 -30 -25 -20 -15 -10

-50

-45

-40

-35

-30

-25

-20

-15

-10

-5

13C

of

ace

tate

(p

erm

il)

in t

he

pre

se

nce

of

CH

3F

13

C of organic C (permil)

various lakes

Pantanal

Amazon 2009

Amazon 2010

Tibet Plateau

Uruguay rice

various rice soils

various upland soils

r = 0.64

P < 0.01

Fig. 2

A B

Fig. 3

-55 -50 -45 -40 -35 -30 -25 -20 -15 -10

-55

-50

-45

-40

-35

-30

-25

-20

-15

-10

13C

of

ace

tate

-me

thyl (p

erm

il)

13

C of total acetate (permil)

Pantanal Lakes

with CH3F

Amazon Lakes 2009

with CH3F

Amazon Lakes 2010

with CH3F

other lakes

with CH3F

with BES, CHCl3

-55 -50 -45 -40 -35 -30 -25 -20 -15 -10 -5 0 5 10 15 20 25 30

-55

-50

-45

-40

-35

-30

-25

-20

-15

-10

-5

0

5

10

15

20

25

30

13C

of

ace

tate

-me

thyl (p

erm

il)

13

C of total acetate (permil)

rice soils, rice straw

rice soils, rice straw + CH3F

rice soils, maize straw

rice soils, maize straw + CH3F

other rice soils

other rice soils + CH3F

upland soils + CH3F

0 100 200 300 400 500 600 700 8000.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0 rice straw

rice straw + CH3F

maize straw

maize straw + CH3F

Pro

pio

nate

(m

M)

Time (h)

0 100 200 300 400 500 600 700 800-40

-35

-30

-25

-20

-15

-10

rice straw

rice straw + CH3F

maize straw

maize straw + CH3F

1

3C

of pro

pio

nate

(perm

il)

Time (h)

0 100 200 300 400 500 600 700 800 900

0

1

2

3

4

5

6

7

8

9

10

rice straw

rice straw + CH3F

maize straw

maize straw + CH3F

Pro

pio

na

te (

mM

)

Time (h)

0 100 200 300 400 500 600 700 800 900-40

-35

-30

-25

-20

-15

-10

-5

0

rice straw

rice straw + CH3F

maize straw

maize straw + CH3F

13C

of pro

pio

nate

(p

erm

il)

Time (h)

0 100 200 300 400 500 600 700 8000

2

4

6

8

10

12

rice straw

rice straw + CH3F

maize straw

maize straw + CH3F

Pro

pio

na

te (

mM

)

Time (h)

0 100 200 300 400 500 600 700 800-40

-35

-30

-25

-20

-15

-10

rice straw

rice straw + CH3F

maize straw

maize straw + CH3F

1

3C

of pro

pio

nate

(perm

il)

Time (h)

Vercelli Fuyang Suwan

A B C

D E F

Fig. 4

-45 -40 -35 -30 -25 -20 -15 -10 -5 0 5

-45

-40

-35

-30

-25

-20

-15

-10

13C

of

pro

pio

na

te (

pe

rmil)

13

C of total acetate (permil)

Pantanal

Amazon 2009

Amazon 2010

German Lakes

rice soils, rice straw

rice soils, maize straw

Uruguay rice soil

upland soils

Fig. 5

r = 0.81

P < 0.01

organic C

propionate other compounds

-2‰

-4‰

+6‰

+2‰

-6‰

+1‰

acetate CO2

CH4

acetate-

methyl

-8‰

-56‰

-75‰

-21‰

+3‰

+5‰

Fig. 6

22

Highlights Methanogenic degradation of organic matter in anoxic environments examined. δ

13C of organic C, propionate, acetate and acetate-methyl relate to each other. Smaller fractionation during production of propionate and acetate than of CH4. We suggest δ13C value of in organic C can be used as proxy for that of acetate.