Stable carbon isotope fractionation as tracer of carbon cycling in anoxic soil ecosystems

Stable carbon isotope fractionation as tracer of carbon cycling in anoxic soil ecosystems

Available online at www.sciencedirect.com ScienceDirect Stable carbon isotope fractionation as tracer of carbon cycling in anoxic soil ecosystems Mar...

692KB Sizes 0 Downloads 100 Views

Available online at www.sciencedirect.com

ScienceDirect Stable carbon isotope fractionation as tracer of carbon cycling in anoxic soil ecosystems Martin Blaser and Ralf Conrad While the structure of microbial communities can nowadays be determined by applying molecular analytical tools to soil samples, microbial function can usually only be determined by physiological experiments requiring incubation of samples. However, analysis of stable isotope fractionation might be able to analyse microbial function without incubation in soil samples. We describe the limitations of diagnosing and quantifying carbon flux pathways in soil by using the determination of stable carbon isotope composition in soil compounds and emphasize the importance of determining stable isotope fractionation factors for defined biochemical pathways. Fractionation factors are sufficiently different for some central biochemical pathways in anaerobic degradation of organic carbon. Thus, it is possible to quantify the relative contribution of CH4 production by hydrogenotrophic or aceticlastic methanogenic pathways, and of acetate formation by chemolithotrophic (acetyl-CoA synthase) or heterotrophic (fermentation) pathways. In addition, stable isotope analysis may allow the differentiation between different organic substrates used for degradation, for example, the relative contribution of root exudation versus soil organic matter degradation, provided the different substrates are sufficiently distinct in their isotopic compositions (e.g., mixture of C3 and C4 plants) and the carbon conversion pathways display only small fractionation factors or are identical for the different substrates. Address Max Planck Institute for Terrestrial Microbiology, Karl-von-Frisch-Str. 10, 35043 Marburg, Germany Corresponding author: Blaser, Martin ([email protected])

Current Opinion in Biotechnology 2016, 41:122–129 This review comes from a themed issue on Analytical biotechnology Edited by Hans-Hermann Richnow and Tillmann Lueders

http://dx.doi.org/10.1016/j.copbio.2016.07.001 0958-1669/# 2016 Elsevier Ltd. All rights reserved.

Introduction Carbon cycling in soil ecosystems is mainly the result of input of carbon substrates by plants and degradation by soil microorganisms. The input by plants is mostly in the form of litter, dead plant material or root exudates. The Current Opinion in Biotechnology 2016, 41:122–129

degradation, however, is achieved by a complex soil microbial community comprising many different biochemical pathways, which depend on the chemical nature of the organic substrate and the environmental conditions (Figure 1). In general, the anaerobic degradation of soil organic matter can be separated in the hydrolysis and fermentation of polymeric substrates to short-chain fatty acids, alcohols, CO2 and H2. Further fermentation results in degradation to acetate, CO2 and H2, which are the dominant precursors of methanogenesis. In anaerobic systems, CH4 and CO2 are stable end products. Oxidation of CH4 usually requires the presence of O2 and methane oxidizing bacteria. Understanding the carbon cycling in soil ecosystems therefore requires knowledge of both the structure and the function of the microbial communities. There has recently been much progress in elucidating the structure of microbial communities by applying molecular analytical tools, in particular sequencing of microbial genes and gene products [1]. These analyses just require sampling of soil and — the sometimes technically demanding — extraction of nucleic acids or proteins. While some information on microbial functioning can be obtained by combining genomic and metaproteomic approaches [2–4] or using stable isotope probing techniques [5,6], the in situ functions of the microbial communities usually can only be analyzed by incubation and measurement of the temporal change of analytically accessible variables. However, analysis of stable isotope signatures in soil samples might overcome this problem, since the isotopic signatures partially reflect the microbial functioning [7]. Our review intends to explore the limits of using stable carbon isotopic signatures for elucidating the microbial functional pathways of organic carbon degradation in soil, anaerobic degradation of small organic molecules originating from hydrolysis and fermentation, and methane production in particular.

Isotopic signatures, isotope effects, and fractionation factors The 13C isotopic signature of a particular carbon compound is given by its ratio R = 13C/12C and is usually denoted relative to a standard (st) as d13C = 103 (R/ Rst  1) [8,9]. The reactions in a biochemical pathway, especially those involving the cleavage of carbon bonds, often display characteristic fractionation factors (a) or enrichment factors (e = 103[1  a]), which define the extent to which the carbon atoms have been fractionated during the conversion of a substrate to a product. For a reaction (or pathway) A ! B, the fractionation factor is defined as aA,B = (d13CA + 103)/(d13CB + 103) [10]. The www.sciencedirect.com

Stable isotope fractionation in soil carbon cycling Blaser and Conrad 123

Figure 1

C4 plant (mais)

C3 plant (rice) δ13CCO2= –8‰

CO2

assimilation

assimilation

δ13CC3= –25‰

δ13CCH4= –47‰

CO2

δ13CC4= –15‰

CH4

Water

Oxic zone Anoxic zone

acetate, H2/CO2

CO2 fermentation

Soil organic matter

respiration

Current Opinion in Biotechnology

Carbon cycling in soil shown for a flooded ecosystem planted with a C3 plant (i.e. rice) for a non-flooded ecosystem planted with a C4 plant (e.g., maize). The CO2 of the atmosphere is fixed by either RUBISCO (C3) or PEP-carboxylase (C4) resulting in different d13C values of the plant material. Dead plant material, roots and root exudates are the major sources of soil organic matter, which can either be respired to CO2 in oxic soils or fermented to CH4 and CO2 in anoxic soils. Details on the carbon isotope fractionation during anaerobic degradation to CH4 can be found in Figure 2. Delta13C are taken from Refs. [15,90,91].

fractionation factors thus quantify how much a given biochemical reaction (or pathway) discriminates against the substrate molecules containing the heavy isotope 13C. The discrimination itself is caused by the isotope effect. This is either a kinetic isotope effect (KIE) caused by reaction rate constants being larger for substrate with 12C than 13C, or an equilibrium isotope effect (EIE) that is involved in the establishment of chemical equilibria, such as dissociations, protonations and may either prefer 13C or 12 C [11–13]. Biochemical reactions involving carbon often display a KIE, which usually discriminates against the heavy 13C isotope, so that the d13C of the product is always lower than that of the substrate. Biochemical pathways usually consist of a sequence of reactions which may influence the overall fractionation factor to various extents. In many cases the initial reaction has great implications for the overall fractionation in the pathway [14–17]. A special case is branched biochemical pathways leading to more than one product (e.g., A ! B + C). Here, the fractionation factor aA,B is often different from aA,C. Moreover, one branch may discriminate against www.sciencedirect.com

12

C while the other discriminates against 13C, although the KIE of all the individual reactions are generally positive [18,19]. Hence, branched pathways may (but not obligatorily) result in a product with d13C larger than that of the substrate, if the other branch results in a product with correspondingly lower d13C fulfilling isotopic mass balance. When using the isotopic signature in microbial biomass for elucidating the pathway by which it has been formed, one is always confronted with the problem of a branched pathway with a catabolic and an anabolic branch, respectively. In summary, stable isotopic fractionation is a general feature of (bio)chemical reactions. If two pathways display sufficiently different fractionation factors, reflected in the difference of d13C between substrate and product these pathways can be differentiated using stable carbon isotope measurements.

Determination of fractionation factors for microbial pathways The compound specific stable isotope analysis (CSIA) of d13C usually requires an isotope ratio mass spectrometer (IRMS), which measures the d13C in CO2. If a particular Current Opinion in Biotechnology 2016, 41:122–129

124 Analytical biotechnology

carbon compound can be isolated (e.g., by chromatography) and converted to CO2 (e.g., by combustion), the average d13C of all its C atoms can be determined usually using a GC–IRMS or HPLC–IRMS system. For some simple compounds (e.g., CO2, CH4), the d13C can also be determined using laser-based absorption spectrometry, for example, cavity ringdown spectrometers, which allow continuous, non-destructive operation in contrast to IRMS, which can only analyze discrete samples. For reviews on recent instrumental developments see [20,21]. As the fractionation factors are of pivotal importance for the detection of different pathways, knowledge of their magnitudes is mandatory. However, determination of the fractionation factor for a particular pathway requires that there is no interference by other pathways, no restriction by small-scale mass transfer processes or by the so-called commitment to catalysis favoring a non-equilibrium mass flow according to the direction of the pathway taking place inside the microbial cell [16,19,22]. This requirement is usually not met when incubating environmental samples in which an a priori unknown number of pathways may contribute to a particular substrate–product conversion. Fractionation factors have to be determined under well defined conditions, which are usually only met by assaying biochemical reactions or defined microbial cultures in which the desired pathway operates. Then, fractionation factors can be determined either in a closed system (e.g., batch culture) during the course of substrate depletion and product accumulation using the Rayleigh equation for the fractional ( f) turnover of the substrate into the product [23,24]; for the substrate: dr = dri + e[ln(1  f)] and the product: dp = dri  e(1  f)[ln(1  f)]/f, where dri is the initial isotope composition of the reactant, dr and dp the isotope composition of the substrate and product at the instant when f was determined. In an open system (e.g., chemostat) under steady state conditions the isotope fractionation can be directly deduced from the difference in d13C of substrate and product [18,24]; for the substrate: dr = dri + e(1  f) and for the product: dp = dri  ef. This can be simplified to: e = dr  dp. Even in a defined microbial culture the catabolic path of substrate conversion is not completely isolated, as anabolic substrate conversion is taking place simultaneously (i.e., branched pathway). Since the energy yield is usually low (<2 ATP per substrate consumed) in anaerobic metabolism, anabolism contributes accordingly little (usually less than 10%) to total anaerobic substrate consumption. Nevertheless, fractionation during anaerobic anabolism is happening: In methanogenic archaea, fractionation factors during biomass production range from 31 to +7 % [25–28]. Catabolic CH4 production, on the other hand, shows fractionation factors between 25 and 69% for methanogenesis from CO2 [29–32], 73 to Current Opinion in Biotechnology 2016, 41:122–129

83% from methanol [27,28], and 7 to 35% for methanogenesis from acetate [25,27,32–34], respectively. The bias caused by interfering anabolism can be overcome by incubating microbial cells under defined nongrowing conditions (e.g., absence of a nitrogen source). There are many conceptual and practical problems when determining fractionation factors in microbial cultures. While the strongest fractionation reported for hydrogenotrophic methanogenic archaea was achieved under energy limitation, weaker fractionation factors were determined for energy rich conditions [31,32]; temperature had no effect on the fractionation [35]. By contrast the fractionation of acetogenic bacteria was not influenced by changing the energy supply (H2:CO2 ratios) but was influenced by media composition [36] and substrate usage [37]. Additional constraints on the fractionation due to substrate transport across membranes [14,38], cell density [39,40], availability of co-substrates and electron acceptors [22,41– 43] or commitment to catalysis [15,20] have been described in the literature. Likewise the fractionation factors of microbial cultures show strain-specific variability resulting in slightly different fractionation factors for each strain [25,27,28,32,34,44]. For the anaerobic degradation of acetate by methanogenic archaea, sulfur and sulfate reducers it could be shown that not only the pathway usage (acetylCoA pathway or reversed TCA) but likewise the enzyme responsible for acetate activation (acetate kinase vs. acetyl-CoA synthase) impact the overall fractionation of acetate to the respective products [45,46]. For example, the acetoclastic methanogens can be differentiated into members of the Methanosaetaceae which use the acetyl-CoA synthase, as an activating enzyme and showed a relative small isotope fractionation of 7 to 9% [32,34]; while the second group of acetoclastic methanogens the Methanosarcinaceae usually use the acetate kinase to activate the acetate and can use several substrates in addition to acetate. They have a stronger fractionation of 12 to 35% [25,27,33]. As a consequence the fractionation factors reported for microbial pure cultures have to be interpreted in the context of their detailed biochemical pathways and also in the context of the experimental conditions. Even identical conditions may give a relatively large variability of the fractionation factors. For example, 21 independent replicates of the acetogen T. kivui covered a range of eTIC = 63 to 47% (average, 54%) [44]. Such uncertainties have several consequences: For application to environmental samples a range of possible fractionation factors rather than a constant value should be used. Likewise the interpretation of fractionation factors deduced from environmental samples is uncertain in accordance to the ranges observed in pure cultures. Currently, stable isotope fractionation can safely be used to differentiate very strongly fractionating pathways. While investigating more microbial pure cultures [28,35,36, 44,47] and further constraining the environmental variables www.sciencedirect.com

Stable isotope fractionation in soil carbon cycling Blaser and Conrad 125

influencing the fractionation factors may help to make differentiation of pathways more accurate, another possibility is applying the fractionation of several stable isotopes together, for example, 13C with 2H or 18O [48–50]. These techniques are very valuable in resolving, for example, the anaerobic degradation of organic pollutants where the apparent fractionation factors of the individual isotopes are comparatively small or may be masked by other effects, for example, commitment to catalysis [16,17,22,32,50,51]. Recent application of these combined (clumped) carbon and hydrogen isotope analysis on environmental methane, could show that the microbially derived methane does not follow the thermodynamic equilibrium of abiotically formed methane [52,53,54,55]. However, so far no clumped isotope measurements for microbial pure cultures are available. The data of hydrogen isotope analysis in methanogenic cultures is very limited and restricted to thermophilic hydrogenotrophic methanogens [32,56,57]. The isotopic values of microbial pure cultures can be used to interpret environmental signals. Under certain circumstances, fractionation factors may also be determined directly in soil incubations. However, such determination requires incubation experiments and works only if the operation of microbial pathways can be isolated in situ, for example, by inhibiting alternative pathways. For example, the fractionation factor for hydrogenotrophic methanogenesis can be determined by incubation of soil samples in the presence of CH3F, which specifically inhibits aceticlastic methanogenesis [58,59]. However, this method falls short if CH4 is in addition produced by methylotrophic methanogenesis, which is not inhibited by CH3F [28].

Diagnosis and quantification of carbon flux pathways in soil The measurement of d13C of individual soil carbon compounds may allow the diagnosis of particular biochemical pathways that operated until the time of sampling [60–62]. The usual approach is defining theoretically possible C flux paths that can be differentiated by their fractionation factors and testing them by d13C analysis of their substrates and products. The crucial point is that the pathways to be discriminated must display sufficiently different fractionation factors, reflected in the difference of d13C between substrate and product. Ecological studies sometimes have used tiny differences in the carbon and nitrogen isotopic values for interpretation of underlying food chains using context knowledge (reviewed in [63,64]). In microbial ecology, on the other hand, the underlying biochemical pathways are usually much harder to resolve. We consider two aspects of carbon flux in soil, (i) diagnosis and quantification of the relative contribution of different substrates to the formation of a product; www.sciencedirect.com

and (ii) diagnosis and quantification of the relative contribution of different biochemical pathways to substrate conversion and product formation. (i) Soil organic matter usually originates from biomass that is produced by CO2 fixation. Isotope fractionation is relatively strong, depending on the pathway of CO2 fixation. For example, the pathway of CO2 fixation by C4 plants has a smaller fractionation factor than that of C3 plants resulting in organic carbon that is less depleted in 13C [15,65] (Figures 1 and 2). The isotopic signature of biomass is relatively well preserved in organic matter, since subsequent degradation by respiratory processes displays only small fractionation [66–69]. The same is true for microbial fermentation, which displays only a small fractionation factor [67,70,71,72,73,74]. Therefore, different respiration or fermentation pathways cannot be distinguished using carbon isotope signals alone. However, it is possible to determine the contribution of different substrates to CO2 production provided the d13C of the substrates is sufficiently different. For example, the contribution of the anaerobic degradation of root versus soil organic matter to CO2 production can be quantified, if crops are changed from C3 to C4 plants or vice versa [66,68,69]. Then, soil organic matter has still the typical 13C signature while plant material has a C4 signature. Similarly, this approach can also be used for quantifying the contribution of root exudation versus degradation of soil organic matter (or straw carbon) to production of other compounds, such as microbial biomass, acetate or CH4 [66,75–78,79]. For example, it was shown that most of the CH4 produced in flooded rice field soil is derived from root exudation [77,78]. In case of aerobic degradation to CO2 with negligible isotope fractionation the fraction of C3 ( fC3,CO2) and C4 ( fC4,CO2) organic carbon contributing to total CO2 production is calculated by the following mass balance: d13 CCO2 ¼ d13 C3 f C3 ;CO2 þ ð1f C3 ;CO2 Þd13 CC4 : For the anaerobic degradation to CH4, we have to assume that the overall enrichment factor ðeorg;CH4 Þ for the conversion of organic carbon to CH4 is the same for C3 and C4 organic carbon. This can be written as the following equation: d13 CCH4 ¼ d13 C3 f C3 ;CO2 þ ð1f C3 ;CO2 Þd13 CC4 þ eorg;CH4 : The enrichment factor eorg;CH4 can be calculated from factors in the literature enrichment ði:e:; eCO2 ;CH4 ; eac;CH4 Þ if the path of CH4 production is known: eorg;CH4 ¼ eCO2 ;CH4 f CO2 ; CH4 þ ð1f CO2 ; CH4 Þeac;CH4 : Current Opinion in Biotechnology 2016, 41:122–129

126 Analytical biotechnology

The exact quantification of the relative contribution of different substrates to the degradation process requires that the degradation process has a small fractionation factor or is identical for each substrate. Therefore, application to anaerobic degradation processes is challenging, since large fractionation factors are involved in the conversion of CO2 to acetate or CH4. Furthermore, it can a priori not be excluded that the degradation process is (at least to some extent) different for different substrates (e.g., root exudates versus soil organic matter). (ii) While many respiration or fermentation pathways have only a small discrimination against 13C; there are several anaerobic degradation processes, which display large fractionation factors (up to e = 80% reported for methylotrophic methanogens [27,28]) that can be used for diagnosis and quantification of biochemical pathways (Figure 2). Examples are acetate formation by chemolithotrophic acetogenic bacteria using CO2 as carbon substrate. This biochemical pathway exhibits a very large carbon fractionation (e = 68 to 38%) factor [44] allowing to determine whether acetate has been formed from CO2 or from organic compounds [80–82]. Similarly, CH4 production from CO2 displays a much larger fractionation factor than from acetate (see above), so that the path of CH4 production in anoxic environments can be differentiated [75,83–85]. If soil organic matter is anaerobically degraded to CH4 and CO2, fermentation first produces CO2 and acetate displaying little isotopic fractionation, so that these intermediary products have a similar d13C than the organic matter itself [70,72,74]. Their further conversion, however, results in fractionation with CH4 produced from CO2 being much more depleted in 13C than that from acetate [25,27,32,34]. The residual CO2 and

acetate, on the other hand, become then correspondingly enriched in 13C. These processes cannot only be used to diagnose the occurrence of different pathways, but can even be used to quantify their contributions, provided their fractionation factors are known and the d13C in the substrates (CO2, acetate) and products (CH4) can be measured [75,83–85]. For example, for calculation of the fractions of hydrogenotrophic ðf CO2 ;CH4 Þ and aceticlastic ðf ac;CH4 Þ methanogenesis of total CH4 production it is necessary to determine the d13C in the CH4, CO2 and the methyl group of acetate. With these data the following mass balance equations can be used for calculation:

Figure 2

d13 Cac ¼ d13 CCO2 þ eCO2 ;ac

–8‰ CO2

C3 C4

–25‰ 25‰ Org. g C

15‰ –15‰

<3‰ <–5‰

–55‰

–50‰2

CO2

CH4

–80‰

methanol Current Opinion in Biotechnology

Scheme of carbon flow and stable carbon isotope enrichment factors (e) in methanogenic environments. In grey typical values for C3 and C4 plants as well as atmospheric CO2 are given. The isotope enrichment factors give the rounded average of pure culture experiments; details on the variability are given in the text. 1Fractionation factor is given for pure cultures of Methanosaetacea; the fractionation of Methanosarcinacea is on average 24%. 2The fractionation of hydrogenotrophic methanogenesis is highly flexible and best determined for the respective environment using CH3F as inhibitor for acetoclastic methanogenesis. Current Opinion in Biotechnology 2016, 41:122–129

d13 Cmc ¼ d13 CCO2 þ eCO2 ;CH4 d13 Cma ¼ d13 Cacmethyl þ eac;CH4 f CO2 ;CH4 þ f ac;CH4 ¼ 1 eCO2 ;CH4 is the isotopic enrichment factor for the hydrogenotrophic formation of CH4 from CO2 (e.g., 50%), and eac;CH4 that for the aceticlastic formation of CH4 from the methyl group of acetate (e.g., 8%). If d13Cac-methyl cannot be determined, it may be approximated from either the d13C of total acetate (d13Cac): d13Cac-methyl  d13Cac  8% or of the d13C of organic C (d13Corg) d13Cac  d13Corg  2% [72]. Analogously, fractions of acetate produced from chemolithotrophic ðf CO2 ;ac Þ and fermentative ( forg,ac) acetogenesis can be calculated by the following mass balance, if d13C of acetate, CO2 and organic C have been measured. d13 Cac ¼ d13 Cac f CO2 ;ac þ ð1f CO2 ;ac Þd13 Cao

d13 Cao ¼ d13 Corg þ eorg;ac

–8‰1

acetate

d13 CCH4 ¼ d13 Cmc f CO2 ;CH4 þ ð1f CO2 ;CH4 Þd13 Cma

f CO2 ;ac þ f org;ac ¼ 1 where eCO2 ;ac is the isotopic enrichment factor for the chemolithotrophic formation of acetate from CO2 via the acetyl-CoA pathway (e.g., 55%), and eac,CH4 that for the fermentative formation of acetate from organic substrate (e.g., 2%).

Conclusions and future perspectives Stable isotope fractionation can be useful for distinguishing different carbon flux pathways provided the ranges covered by the fractionation factors associated with the individual pathways do not overlap. This is nicely the case for anaerobic processes involved in the degradation of organic matter, thus allowing the differentiation and even quantification of methanogenesis by hydrogenotrophic or aceticlastic www.sciencedirect.com

Stable isotope fractionation in soil carbon cycling Blaser and Conrad 127

pathways, and of acetogenesis by chemolithotrophic (acetyl-CoA synthase) or heterotrophic (fermentation) pathways. So far many pathways incorporating C1 compounds have the tendency to strongly discriminate against 13C. Hence other pathways involving C1 compounds are the interest of current (e.g., aerobic or anaerobic methane oxidation [86,87]) and future research projects. In addition, stable isotope analysis allows distinguishing between different substrates used for degradation, for example, quantifying the relative contribution of one versus the other substrate in the degradation process. However, such distinction is only possible if the different substrates are sufficiently distinct in their isotopic compositions and if the carbon conversion pathways display only small fractionation factors or are identical for the different substrates. For both applications the knowledge of the fractionation factor is of crucial importance. More studies on effectors influencing the range of fractionation factors in microbial pure cultures are therefore of great importance. To further resolve the underlying principles and pinpoint the fractionating steps of a pathway and possible influencing factors more biochemical investigations on isolated enzyme systems would also be welcome. In addition new analytical methods, such as cavity ring down spectroscopy, which increase the speed and the ease of the analysis, or clumped isotope techniques, which increase the resolution of isotope effects, will shape the future of carbon isotope studies in soil environments. At the time being, isotopic composition of microbial metabolites and microbial biomass is only measured in relatively bulky samples. Hence, interpretation of the isotope values is integrating over relatively large scales. This is unfortunate, since the microbial populations and the individual microbial cells operate on a microscale. Ideally, one would like to trace carbon utilization on a cellular scale. NanoSims technology allows such measurements, but only after labeling with stable isotopes [88]. The natural abundance can presently not be assessed on a cellular scale with great precision [89]. Less ambitious, but still very desirable, would be the analysis of isotopic values along small gradients, for example, vertical profiles of in CH4, TOC, and acetate in a sediment or along a plant root. Gas extraction for analysis of d13C has to make sure that the sampling procedure itself does not cause fractionation.

References and recommended reading Papers of particular interest, published within the period of review, have been highlighted as:  of special interest  of outstanding interest 1.

Nesme J, Achouak W, Agathos SN, Bailey M, Baldrian P, Brunel D, Frostegard A, Heulin T, Jansson JK, Jurkevitch E et al.: Back to the future of soil metagenomics. Front Microbiol 2016, 7:73.

www.sciencedirect.com

2.

Ram RJ, Verberkmoes NC, Thelen MP, Tyson GW, Baker BJ, Blake RC 2nd, Shah M, Hettich RL, Banfield JF: Community proteomics of a natural microbial biofilm. Science 2005, 308:1915-1920.

3.

Ravin NV, Mardanova AV, Skryabin KG: Metagenomics as a tool for the investigation of uncultured microorganisms. Genetika 2015, 51:519-528.

4.

Simon C, Daniel R: Metagenomic analyses: past and future trends. Appl Environ Microbiol 2011, 77:1153-1161.

5.

Blagodatskaya E, Kuzyakov Y: Active microorganisms in soil: critical review of estimation criteria and approaches. Soil Biol Biochem 2013, 67:192-211.

6.

Uhlik O, Leewis MC, Strejcek M, Musilova L, Mackova M, Leigh MB, Macek T: Stable isotope probing in the metagenomics era: a bridge towards improved bioremediation. Biotechnol Adv 2013, 31:154-165.

7.

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

8.

Brand WA, Huang L, Mukai H, Chivulescu A, Richter JM, Rothe M: How well do we know VPDB? Variability of delta C-13 and delta O-18 in CO2 generated from NBS19-calcite. Rapid Commun Mass Spectrom 2009, 23:915-926.

9.

Hayes JM: Isotopic order, biogeochemical processes, and earth history — Goldschmidt Lecture, Davos, Switzerland, August 2002. Geochim Cosmochim Acta 2004, 68:1691-1700.

10. Hayes JM: Factors controlling C-13 contents of sedimentary organic-compounds — principles and evidence. Marine Geol 1993, 113:111-125. 11. Galimov E: Isotope organic geochemistry. Org Geochem 2006, 37:1200-1262. 12. Nijenhuis I, Richnow HH: Stable isotope fractionation concepts for characterizing biotransformation of organohalides. Curr Opin Biotechnol 2016, 41 this issue. 13. Wong KY, Xu YQ, Xu L: Review of computer simulations of isotope effects on biochemical reactions: from the Bigeleisen equation to Feynman’s path integral. Biochim Biophy ActaProteins Proteomics 2015, 1854:1782-1794. 14. Nijenhuis I, Andert J, Beck K, Kastner M, Diekert G, Richnow HH: Stable isotope fractionation of tetrachloroethene during reductive dechlorination by Sulfurospirillum multivorans and Desulfitobacterium sp strain PCE-S and abiotic reactions with cyanocobalamin. Appl Environ Microbiol 2005, 71:3413-3419. 15. Oleary MH: Carbon isotope fractionation in plants. Phytochemistry 1981, 20:553-567. 16. Thullner M, Fischer A, Richnow HH, Wick LY: Influence of mass transfer on stable isotope fractionation. Appl Microbiol Biotechnol 2013, 97:441-452. 17. Vogt C, Cyrus E, Herklotz I, Schlosser D, Bahr A, Herrmann S, Richnow HH, Fischer A: Evaluation of toluene degradation pathways by two-dimensional stable isotope fractionation. Environ Sci Technol 2008, 42:7793-7800. 18. Hayes JM: Fractionation of carbon and hydrogen isotopes in biosynthetic processes. Rev Miner Geochem 2001, 43:225-277. 19. Northrop DB: The expression of isotope effects on enzymecatalyzed reactions. Ann Rev Biochem 1981, 50:103-131. 20. Eiler JM, Bergquist B, Bourg I, Cartigny P, Farquhar J, Gagnon A,  Guo WF, Halevy I, Hofmann A, Larson TE et al.: Frontiers of stable isotope geoscience. Chem Geol 2014, 372:119-143. Comprehensive review on recent advances in isotope techinques and applications. 21. Griffis TJ: Tracing the flow of carbon dioxide and water vapor between the biosphere and atmosphere: a review of optical isotope techniques and their application. Agric Forest Meteorol 2013, 174:85-109. 22. Thullner M, Centler F, Richnow HH, Fischer A: Quantification of organic pollutant degradation in contaminated aquifers using Current Opinion in Biotechnology 2016, 41:122–129

128 Analytical biotechnology

compound specific stable isotope analysis — review of recent developments. Org Geochem 2012, 42:1440-1460. 23. Mariotti A, Germon JC, Hubert P, Kaiser P, Letolle R, Tardieux A, Tardieux P: Experimental-determination of nitrogen kinetic isotope fractionation — some principles — illustration for the denitrification and nitrification processes. Plant Soil 1981, 62:413-430. 24. Fry B: Stable Isotope Ecology. Springer; 2006. 25. Goevert D, Conrad R: Effect of substrate concentration on carbon isotope fractionation during acetoclastic methanogenesis by Methanosarcina barkeri and M. acetivorans and in rice field soil. Appl Environ Microbiol 2009, 75:2605-2612. 26. House CH, Schopf JW, Stetter KO: Carbon isotopic fractionation by Archaeans and other thermophilic prokaryotes. Org Geochem 2003, 34:345-356. 27. Londry KL, Dawson KG, Grover HD, Summons RE, Bradley AS: Stable carbon isotope fractionation between substrates and products of Methanosarcina barkeri. Org Geochem 2008, 39:608-621. 28. Penger J, Conrad R, Blaser M: Stable carbon isotope fractionation by methylotrophic methanogenic archaea. Appl Environ Microbiol 2012, 78:7596-7602. 29. Botz R, Pokojski HD, Schmitt M, Thomm M: Carbon isotope fractionation during bacterial methanogenesis by CO2 reduction. Org Geochem 1996, 25:255-262. 30. Games LM, Hayes JM, Gunsalus P: Methane-producing bacteria: natural fractionations of the stable carbon isotopes. Geochim Cosmochim Acta 1978, 42:1295-1297. 31. Penning H, Plugge CM, Galand PE, Conrad R: Variation of carbon isotope fractionation in hydrogenotrophic methanogenic microbial cultures and environmental samples at different energy status. Global Change Biol 2005, 11:2103-2113. 32. Valentine DL, Chidthaisong A, Rice A, Reeburgh WS, Tyler SC: Carbon and hydrogen isotope fractionation by moderately thermophilic methanogens. Geochim Cosmochim Acta 2004, 68:1571-1590. 33. Gelwicks JT, Risatti JB, Hayes JM: Carbon isotope effects associated with aceticlastic methanogenesis. Appl Environ Microbiol 1994, 60:467-472. 34. Penning H, Claus P, Casper P, Conrad R: Carbon isotope fractionation during acetoclastic methanogenesis by Methanosaeta concilii in culture and a lake sediment. Appl Environ Microbiol 2006, 72:5648-5652. 35. Penger J, Conrad R, Blaser M: Stable carbon isotope fractionation of six strongly fractionating microorganisms is not affected by growth temperature under laboratory conditions. Geochim Cosmochim Acta 2014, 140:95-105. 36. Blaser MB, Dreisbach LK, Conrad R: Carbon isotope  fractionation of Thermoanaerobacter kivui in different growth media and at different total inorganic carbon concentration. Org Geochem 2015, 81:45-52. This study highlights different factors influencing carbon isotope fractionation in microbial pure cultures and assesses the variability in replicated growth conditions. 37. Freude C, Blaser M: Carbon isotope fractionation during catabolism and anabolism in acetogenic bacteria growing on different substrates. Appl Environ Microbiol 2016, 82:2728-2737. 38. Cichocka D, Siegert M, Imfeld G, Andert J, Beck K, Diekert G, Richnow HH, Nijenhuis I: Factors controlling the carbon isotope fractionation of tetra- and trichloroethene during reductive dechlorination by Sulfurospirillum ssp and Desulfitobacterium sp strain PCE-S. FEMS Microbiol Ecol 2007, 62:98-107. 39. Kampara M, Thullner M, Harms H, Wick LY: Impact of cell density on microbially induced stable isotope fractionation. Appl Microbiol Biotechnol 2009, 81:977-985. 40. Templeton AS, Chu KH, Alvarez-Cohen L, Conrad ME: Variable carbon isotope fractionation expressed by aerobic Current Opinion in Biotechnology 2016, 41:122–129

CH4-oxidizing bacteria. Geochim Cosmochim Acta 2006, 70:1739-1752. 41. Kampara M, Thullner M, Richnow HH, Harms H, Wick LY: Impact of bioavailability restrictions on microbially induced stable isotope fractionation. 2. Experimental evidence. Environ Sci Technol 2008, 42:6552-6558. 42. Mancini SA, Hirschorn SK, Elsner M, Lacrampe-Couloume G, Sleep BE, Edwards EA, Lollar BS: Effects of trace element concentration on enzyme controlled stable isotope fractionation during aerobic biodegradation of toluene. Environ Sci Technol 2006, 40:7675-7681. 43. Rosell M, Finsterbusch S, Jechalke S, Hubschmann T, Vogt C, Richnow HH: Evaluation of the effects of low oxygen concentration on stable isotope fractionation during aerobic MTBE biodegradation. Environ Sci Technol 2010, 44:309-315. 44. Blaser MB, Dreisbach LK, Conrad R: Carbon isotope fractionation of 11 acetogenic strains grown on H2 and CO2. Appl Environ Microbiol 2013, 79:1787-1794. 45. Goevert D, Conrad R: Carbon isotope fractionation by sulfatereducing bacteria using different pathways for the oxidation of acetate. Environ Sci Technol 2008, 42:7813-7817. 46. Goevert D, Conrad R: Stable carbon isotope fractionation by acetotrophic sulfur-reducing bacteria. FEMS Microbiol Ecol 2010, 71:218-225. 47. Zyakun AM, Kochetkov VV, Baskunov BP, Zakharchenko VN, Peshenko VP, Laurinavichius KS, Anokhina TO, Siunova TV, Sizova OI, Boronin AM: Use of glucose and carbon isotope fractionation by microbial cells immobilized on solid-phase surface. Microbiology 2013, 82:280-289. 48. Gilevska T, Gehre M, Richnow HH: Multidimensional isotope analysis of carbon, hydrogen and oxygen as tool for identification of the origin of ibuprofen. J Pharm Biomed Anal 2015, 115:410-417. 49. Gilevska T, Ivdra N, Bonifacie M, Richnow HH: Improvement of analytical method for chlorine dual-inlet isotope ratio mass spectrometry of organochlorines. Rapid Commun Mass Spectrom 2015, 29:1343-1350. 50. Zhang N, Geronimo I, Paneth P, Schindelka J, Schaefer T, Herrmann H, Vogt C, Richnow HH: Analyzing sites of OH radical attack (ring vs. side chain) in oxidation of substituted benzenes via dual stable isotope analysis (delta(13)C and delta(2)H). Sci Total Environ 2016, 542:484-494. 51. Elsner M, Zwank L, Hunkeler D, Schwarzenbach RP: A new concept linking observable stable isotope fractionation to transformation pathways of organic pollutants. Environ Sci Technol 2005, 39:6896-6916. 52. Stolper DA, Lawson M, Davis CL, Ferreira AA, Neto EVS, Ellis GS, Lewan MD, Martini AM, Tang Y, Schoell M et al.: Formation temperatures of thermogenic and biogenic methane. Science 2014, 344:1500-1503. 53. Stolper DA, Martini AM, Clog M, Douglas PM, Shusta SS, Valentine DL, Sessions AL, Eiler JM: Distinguishing and  understanding thermogenic and biogenic sources of methane using multiply substituted isotopologues. Geochim Cosmochim Acta 2015, 161:219-247. Measurement of multiply substituted methane isotopologues ((CH3D)-C13 and (CH2D2)-C-12) improve the resolution of thermogenic versus biogenic methane. 54. Stolper DA, Sessions AL, Ferreira AA, Neto EVS, Schimmelmann A, Shusta SS, Valentine DL, Eiler JM: Combined C-13-D and D–D clumping in methane: methods and preliminary results. Geochim Cosmochim Acta 2014, 126:169191. 55. Wang DT, Gruen DS, Lollar BS, Hinrichs KU, Stewart LC, Holden JF, Hristov AN, Pohlman JW, Morrill PL, Konneke M et al.: Methane cycling. Nonequilibrium clumped isotope signals in microbial methane. Science 2015, 348:428-431. 56. Kawagucci S, Kobayashi M, Hattori S, Yamada K, Ueno Y, Takai K, Yoshida N: Hydrogen isotope systematics among H-2–H2O– CH4 during the growth of the hydrogenotrophic methanogen www.sciencedirect.com

Stable isotope fractionation in soil carbon cycling Blaser and Conrad 129

Methanothermobacter thermautotrophicus strain Delta H. Geochim Cosmochim Acta 2014, 142:601-614. 57. Yoshioka H, Sakata S, Kamagata Y: Hydrogen isotope fractionation by Methanothermobacter thermoautotrophicus in coculture and pure culture conditions. Geochim Cosmochim Acta 2008, 72:2687-2694. 58. Daebeler A, Gansen M, Frenzel P: Methyl fluoride affects methanogenesis rather than community composition of methanogenic archaea in a rice field soil. PLoS One 2013:8. 59. Janssen PH, Frenzel P: Inhibition of methanogenesis by methyl fluoride: studies of pure and defined mixed cultures of anaerobic bacteria and archaea. Appl Environ Microbiol 1997, 63:4552-4557. 60. Dumig A, Rumpel C, Dignac MF, Kogel-Knabner I: The role of lignin for the delta C-13 signature in C-4 grassland and C-3 forest soils. Soil Biol Biochem 2013, 57:1-13. 61. Lerch TZ, Nunan N, Dignac MF, Chenu C, Mariotti A: Variations in microbial isotopic fractionation during soil organic matter decomposition. Biogeochemistry 2011, 106:5-21. 62. Kodina LA: Carbon isotope fractionation in various forms of biogenic organic matter: I. Partitioning of carbon isotopes between the main polymers of higher plant biomass. Geochem Int 2010, 48:1157-1165. 63. Perkins MJ, McDonald RA, van Veen FJ, Kelly SD, Rees G, Bearhop S: Application of nitrogen and carbon stable isotopes (delta(15)N and delta(13)C) to quantify food chain length and trophic structure. PLoS One 2014, 9:e93281. 64. Ohkouchi NONO, Chikaraishi Y, Tanaka H, Wada E: Biochemical  and physiological bases for the use of carbon and nitrogen isotopes in environmental and ecological studies. Prog Earth Planetary Sci 2015, 2:2-17. Up to date review on carbon and nitrogen isotopes in environmental and ecological studies. 65. Farquhar GD, Ehleringer JR, Hubick KT: Carbon isotope discrimination and photosynthesis. Ann Rev Plant Physiol Plant Mol Biol 1989, 40:503-537. 66. Werth M, Kuzyakov Y: Root-derived carbon in soil respiration and microbial biomass determined by C-14 and C-13. Soil Biol Biochem 2008, 40:625-637. 67. Deniro MJ, Epstein S: Influence of diet on distribution of carbon isotopes in animals. Geochim Cosmochim Acta 1978, 42:495506. 68. Mueller CW, Gutsch M, Kothieringer K, Leifeld J, Rethemeyer J, Brueggemann N, Koegel-Knabner I: Bioavailability and isotopic composition of CO2 released from incubated soil organic matter fractions. Soil Biol Biochem 2014, 69:168-178.

Thermotoga maritima and Persephonella marina. Environ Microbiol 2002, 4:58-64. 75. Conrad R, Klose M, Lu Y, Chidthaisong A: Methanogenic pathway and archaeal communities in three different anoxic soils amended with rice straw and maize straw. Frontiers Microbiol 2012, 3:1-12 http://dx.doi.org/10.3389/ fmicb.2012.00004. 76. Conrad R, Klose M, Yuan Q, Lu Y, Chidthaisong A: Stable carbon isotope fractionation, carbon flux partitioning and priming effects in anoxic soils during methanogenic degradation of straw and soil organic matter. Soil Biol Biochem 2012, 49:193-199. 77. Tokida T, Adachi M, Cheng WG, Nakajima Y, Fumoto T, Matsushima M, Nakamura H, Okada M, Sameshima R, Hasegawa T: Methane and soil CO2 production from currentseason photosynthates in a rice paddy exposed to elevated CO2 concentration and soil temperature. Global Change Biol 2011, 17:3327-3337. 78. Yuan Q, Pump J, Conrad R: Partitioning of CH4 and CO2 production originating from rice straw, soil and root organic carbon in rice microcosms. PLoS One 2012, 7 e49073-e49073. 79. Yuan Q, Pump J, Conrad R: Straw application in paddy soil  enhances methane production also from other carbon sources. Biogeosciences 2014, 11:237-246. Recent application using carbon isotope values to determine the contribution of soil organic carbon, straw and root exudates to the production of methane and carbon dioxide. And finding a prinming effect of straw addition on the degradation of root exudates and soil organic carbon. 80. Haedrich A, Heuer VB, Herrmann M, Hinrichs KU, Kuesel K: Origin and fate of acetate in an acidic fen. FEMS Microbiol Ecol 2012, 81:339-354. 81. Heuer VB, Krueger M, Elvert M, Hinrichs KU: Experimental studies on the stable carbon isotope biogeochemistry of acetate in lake sediments. Org Geochem 2010, 41:22-30. 82. Heuer VB, Pohlman JW, Torres ME, Elvert M, Hinrichs KU: The stable carbon isotope biogeochemistry of acetate and other dissolved carbon species in deep subseafloor sediments at the northern Cascadia Margin. Geochim Cosmochim Acta 2009, 73:3323-3336. 83. Conrad R: Quantification of methanogenic pathways using stable carbon isotopic signatures: a review and a proposal. Organic Geochemistry Stable Isotope Applications in Methane Cycle Studies. Elsevier; 2005:. pp. 739–752. 84. Mach V, Blaser MB, Claus P, Chaudhary PP, Rulik M: Methane production potentials, pathways, and communities of methanogens in vertical sediment profiles of river Sitka. Front Microbiol 2015, 6:506.

69. Werth M, Kuzyakov Y: C-13 fractionation at the root– microorganisms–soil interface: a review and outlook for partitioning studies. Soil Biol Biochem 2010, 42:1372-1384.

85. Ji Y, Scavino AF, Klose M, Claus P, Conrad R: Functional and structural responses of methanogenic microbial communities in Uruguayan soils to intermittent drainage. Soil Biol Biochem 2015, 89:238-247.

70. Blair N, Leu A, Munoz E, Olsen J, Kwong E, Des Marais DJ: Carbon isotopic fractionation in heterotrophic microbial-metabolism. Appl Environ Microbiol 1985, 50:996-1001.

86. Milucka J, Kirf M, Lu L, Krupke A, Lam P, Littmann S, Kuypers MM, Schubert CJ: Methane oxidation coupled to oxygenic photosynthesis in anoxic waters. ISME J 2015, 9:1991-2002.

71. Botsch KC, Conrad R: Fractionation of stable carbon isotopes during anaerobic production and degradation of propionate in defined microbial cultures. Org Geochem 2011, 42:289-295.

87. Musat F, Vogt C, Richnow HH: Carbon and hydrogen stable isotope fractionation associated with the aerobic and anaerobic degradation of saturated and alkylated aromatic hydrocarbons. J Mol Microbiol Biotechnol 2016, 26:211-226.

72. Conrad R, Claus P, Chidthaisong A, Lu Y, Fernandez Scavino A,  Liu Y, Angel R, Galand PE, Casper P, Guerin F et al.: Stable carbon isotope biogeochemistry of propionate and acetate in methanogenic soils and lake sediments. Org Geochem 2014, 73:1-7. Extensive database of isotope signals in organic C, propionate, acetate and acetate-methyl observed in environmental samples. 73. Penning H, Conrad R: Carbon isotope effects associated with mixed-acid fermentation of saccharides by Clostridium papyrosolvens. Geochim Cosmochim Acta 2006, 70:2283-2297. 74. Zhang CLL, Ye Q, Reysenbach AL, Gotz D, Peacock A, White DC, Horita J, Cole DR, Fong J, Pratt L et al.: Carbon isotopic fractionations associated with thermophilic bacteria

www.sciencedirect.com

88. Musat N, Musat F, Weber PK, Pett-Ridge J: Tracking microbial interactions with NanoSIMS. Curr Opin Biotechnol 2016, 41 this issue. 89. Orphan VJ, House CH, Hinrichs KU, McKeegan KD, DeLong EF: Methane-consuming archaea revealed by directly coupled isotopic and phylogenetic analysis. Science 2001, 293:484-487. 90. Cuntz M: Carbon cycle a dent in carbon’s gold standard. Nature 2011, 477:547-548. 91. Quay P, Stutsman J, Wilbur D, Snover A, Dlugokencky E, Brown T: The isotopic composition of atmospheric methane. Global Biogeochem Cycles 1999, 13:445-461.

Current Opinion in Biotechnology 2016, 41:122–129