<<
Home , Oil

Geochimica et Cosmochimica Acta 71 (2007) 271–283 www.elsevier.com/locate/gca

Carbon and isotope fractionation associated with the aerobic microbial oxidation of methane, , propane and butane

Franklin S. Kinnaman a, David L. Valentine b,*, Stanley C. Tyler c

a Graduate Program in Marine Sciences, University of California, Santa Barbara, CA 93106, USA b Department of Earth Science and Marine Sciences Institute, University of California, Santa Barbara, CA 93106, USA c Department of Earth System Science, University of California, Irvine, CA 92697, USA

Received 1 February 2006; accepted in revised form 6 September 2006

Abstract

Carbon isotope fractionation factors associated with the aerobic consumption of methane (C1), ethane (C2), propane (C3), and n- butane (C4) were determined from incubations of marine sediment collected from the Coal Oil Point seep field, located offshore Santa Barbara, CA. Hydrogen isotope fractionation factors for C1, C2 and C3 were determined concurrently. Fresh sediment samples from two seep areas were each slurried with sea and treated with C1, C2, C3 or C4, or with mixtures of all four gases. Triplicate samples were incubated aerobically at 15 C, and the stable isotope composition and headspace levels of C1–C4 were moni- tored over the course of the experiment. Oxidation was observed for all C1–C4 gases, with an apparent preference for C3 and C4 over C1 and C2 in the mixed-gas treatments. Fractionation factors were calculated using a Rayleigh model by comparing the d13C and dD of the residual C1–C4 gases to their headspace levels. Carbon isotope fractionation factors (reported in e or (a 1) · 1000 notation) were con- sistent between seep areas and were 26.5& ± 3.9 for C1, 8.0& ± 1.7 for C2, 4.8& ± 0.9 for C3 and 2.9& ± 0.9 for C4. Fraction- ation factors determined from mixed gas incubations were similar to those determined from individual gas incubations, though greater variability was observed during C1 consumption. In the case of C1 and C3 consumption, carbon isotope fractionation appears to decrease as substrate becomes limiting. Hydrogen isotope fractionation factors determined from the two seep areas differed for C1 oxi- dation but were similar for C2 and C3. Hydrogen isotope fractionation factors ranged from 319.9& to 156.4& for C1 incubations, and averaged 61.9& ± 8.3 for C2 incubations and 15.1& ± 1.9 for C3 incubations. The fractionation factors presented here may be applied to estimate the extent of C1–C4 oxidation in natural gas samples, and should prove useful in further studying the microbial oxi- dation of these compounds in the natural environment. 2006 Elsevier Inc. All rights reserved.

1. Introduction higher are present only at trace levels (Scho- ell, 1988). When gas production is due to thermal decom- Hydrocarbon gases including methane (CH4; C1), eth- position (thermogenic gas) C2–C4 are more abundant, ane (C2H6; C2), propane (C3H8, C3), and n-butane (n- frequently comprising 10% or more of the gas. Natural C4H10, n-C4, also C4 hereafter) are produced naturally in gas is widely distributed in the subsurface due to its high deep subsurface environments and represent the major mobility and is mined from the subsurface in areas where components of natural gas. The hydrocarbons composing geologic traps have led to gas accumulation (Hunt, 1996). natural gas are produced primarily from the decomposition The conventional or recoverable global reservoir of natural of organic material. When decomposition is biologically gas is estimated at 100,000 Teragrams (Tg = 1012 g) and is mediated C1 is the dominant product (biogenic gas), and currently being used at a rate of 1500 Tg yr1 (Energy Information Administration, 2004). Global unconvention- * Corresponding author. Fax: +1 805 893 2314. al gas resources in basement rock, shales, coal and hydrates E-mail addresses: [email protected] (F.S. Kinnaman), are not well quantified but are estimated to greatly exceed [email protected] (D.L. Valentine). conventional resources (Hunt, 1996). C1 is the principle

0016-7037/$ - see front matter 2006 Elsevier Inc. All rights reserved. doi:10.1016/j.gca.2006.09.007 272 F.S. Kinnaman et al. 71 (2007) 271–283 component of natural gas, and has received significantly and Hanson, 1996). C1 oxidation occurs in both oxic and more attention than C2–C4, particularly because of its anoxic environments; the focus of this study is on aerobic importance to energy, climate change and atmospheric oxidation. The enzymes responsible for this activity are chemistry (Cicerone and Oremland, 1988; Fung et al., the two primary forms of methane monooxygenase 1991; Prather and Ehhalt, 2001). (MMO): the particulate (pMMO) and soluble (sMMO) The atmospheric biogeochemical cycles of C2–C4 forms (Hanson and Hanson, 1996). Both MMO enzymes hydrocarbons are similar to that of C1 in that emissions are able to activate a range of multi-carbon substrates are primarily terrestrial and loss is primarily atmospheric. including short chain and alkenes, and sMMO is As with C1, the atmospheric levels of C2–C4 hydrocarbons also known to activate ethers, alicyclics, aromatics and het- are elevated in the Northern Hemisphere and display erocyclics (Colby et al., 1977). Some aerobic methano- annual cycles related to their removal by OH radicals trophs have recently been shown to grow facultatively (Singh and Zimmerman, 1992; Clarkson et al., 1997). with concurrent consumption of both C1 and multi-carbon However, there are several important differences. The organic substrates. Dedysh et al. (2005) report that the mixing ratios of C2–C4 are as much as three orders of addition of acetate to methanotrophic cultures may serve magnitude lower than C1 in unpolluted air (Kanakidou as a long term stimulus for C1 oxidation. In this case, the et al., 1991), and typical atmospheric residence times in addition of acetate also lowered the concentration thresh- the tropics are 20 days for C2, 5.3 days for C3 and 2.1 days old for C1 oxidation. However, the effect of alternative for n-C4 (Singh and Zimmerman, 1992), compared to 8.4 substrates on the uptake of C1 is not well established, espe- years for C1 (Prather and Ehhalt, 2001). Sources of cially with regard to the presence of higher hydrocarbons in atmospheric C2 include biomass burning (6.4 Tg yr1; environmental systems. Greenberg and Zimmerman, 1984), soils (0.4 Tg yr1; Ru- The microbial consumption of C2–C4 hydrocarbons has dolph, 1995), vegetation (flux unknown; Warneck, 1988), been investigated previously, though not to the same extent and oceans (0.2–0.5 Tg yr1; Rudolph, 1995). However, as methanotrophy. C2–C4 oxidation seems to be two major sources of atmospheric C1, natural wetlands limited to a group of gram-positive organisms in the and rice cultivation, are not significant sources of C2, C3 Corynebacterium–Nocardia–Mycobacterium–Rhodococcus or C4 (Rudolph, 1995; Matthews, 2000; Prather and group (Ashraf et al., 1994), along with some gram-negative Ehhalt, 2001). Leakage of C2 from subsurface reservoirs Pseudomonas species (Takahashi, 1980; Hamamura et al., due to natural gas loss during gas and oil drilling, coal 1999). Some C3 consuming bacteria utilize monooxygenase mining, gas venting, gas transmission, and transportation enzymes to oxidize alkanes to alcohols by attacking the is substantial and has been estimated at 2.6–11 Tg yr1 subterminal and terminal carbon positions (Perry, 1980; with a best estimate of 6 Tg yr1 (Rudolph, 1995). Sources Stephens and Dalton, 1986; Ashraf et al., 1994). A terminal of C3 and C4 alkanes are not as well constrained but oxidation pathway was determined for the C4 oxidizing generally correspond with those of C2 and include biomass Pseudomonas butanovora (Arp, 1999). Alkane monoxygen- burning (Clarkson et al., 1997; Scholes et al., 2003), oceans ases also demonstrate broad specificity; it has been shown (Saito et al., 2000), and natural gas leakage (Wang et al., that C2–C4 grown bacteria can degrade chlorinated hydro- 2000). The amount of radiocarbon-free C1 in the (Hamamura et al., 1997), ammonia (Hamamura atmosphere is higher than the amount of C1 estimated et al., 1999; Berthe-Corti and Fetzner, 2002), and ethylene from global anthropogenic natural gas loss (Lacroix, (Hamamura et al., 1999). The C4 monooxygenase of 1993; Etiope and Klusman, 2002), raising the possibility Nocardioides sp. strain CF8 was shown to be similar to of significant geologic emission of C1–C4 hydrocarbons copper containing monooxygenases such as ammonia into the atmosphere through gas permeable faults and monooxygenase (AMO) and particulate methane monoox- fractured rock. Alternatively, release of C1 from older peat ygenase (pMMO), while other C4 monoxygenases found in deposits may also account for this observation (Zimov P. butanovora contained functional heme groups (Hamam- et al., 1997). Natural and anthropogenic emissions of ura and Arp, 2000). Pseudomonas putida was recently iden- C2–C4 contribute to atmospheric pollution at the regional tified as being capable of C3 and C4 oxidation using an level as a result of photochemical reactions involving alkane hydroxylase enzyme (Johnson and Hyman, 2006). the hydroxyl radical and nitrogen oxides (Singh and The distribution of stable isotopes (d13C and dD) in Zimmerman, 1992). hydrocarbon gases provides information on the sources Microbes are known to affect the distributions of hydro- and sinks for these gases. For example, the isotopic distri- carbons, primarily through their capacity to consume butions associated with C1 sources and isotopic fraction- hydrocarbons in the presence of a suitable oxidant (Atlas, ation associated with sinks have been used to quantify 1995; Hanson and Hanson, 1996; Widdel and Rabus, the global C1 budget (Stevens and Rust, 1982; Stevens, 2001). The microbial consumption of C1 is referred to as 1988; Quay et al., 1991, 1999; Lassey et al., 1993; Gupta methanotrophy. This process has been the subject of et al., 1996; Tyler et al., 1999; Miller et al., 2002) and are numerous investigations due to its importance in the C1 cy- used routinely for regional and process-oriented investiga- cle and because of the unique involved (Cice- tions (Reeburgh et al., 1991; Bilek et al., 1999; Valentine rone and Oremland, 1988; Reeburgh et al., 1993; Hanson et al., 2001; Grant and Whiticar, 2002; von Fischer and Aerobic isotope fractionation of C1–C4 hydrocarbons 273

Hedin, 2002; Heeschen et al., 2004; Kessler et al., 2005; 119W, 49.9170) and is located at 10 m water depth about Heeschen et al., 2005). In deep subsurface hydrocarbon res- 7 km removed from the center of the Coal Oil Point seep ervoirs, the production of 13C depleted, biogenic C1 is a field. Gas flux from this seep area has not been quantified, major process mediated by anaerobic methanogens, most but the estimated flux is 2–3 orders of magnitude lower likely involving syntrophic bacteria that also utilize higher than from Shane Seep. The composition of gas emitted hydrocarbons (Zengler et al., 1999). The isotopic composi- from the seafloor at Brian Seep is 91.2% C1, 7.5% CO2, tion in C2–C4 hydrocarbon gases is routinely used to infer 1.1% C2, 0.2% C3, with less than 0.05% C4. Both locations the conditions at which the gas was formed (James, 1983; represent minor seeps in the Coal Oil Point seep field, Whiticar, 1990). James and Burns (1984) noted significant which is estimated to emit 80 ± 12 tons of C1 day1 and carbon isotopic enrichments in C3 and C4 from deep sub- 26 ± 12 tons of gaseous C2+ hydrocarbons day1 (Hornaf- surface hydrocarbon reservoirs, which they used to infer ius et al., 1999). The sediment at Brian Seep is composed microbial consumption of these gases. Preference for C3 primarily of sand, and little tar is apparent in the sedi- degradation followed by C4 and then C2 has been noted ments, whereas sediment from Shane Seep is silty and con- as well (Boreham et al., 2001), with particularly high levels tains greater amounts of residual tar. Samples were of biodegradation in lower temperature (<60 C) reservoirs collected by scuba diver, using a hand shovel to scrape (Wenger et al., 2002). Boreham et al. (2004) found biodeg- the top 2–3 cm of sediment into a sealable 1 L con- radation of C2–C4 hydrocarbons in Australian natural tainer. Samples were collected January 19th, 2005 at Brian gases resulting in hydrogen isotope enrichments of over Seep and January 25th, 2005 at Shane Seep, and were re- an order of magnitude greater than corresponding carbon turned to the lab within 1–2 h. isotope enrichments. Here, the C2 and C3 hydrocarbons showed over 200& enrichments in D compared to their 2.2. Sample processing equivalent non-biodegraded natural gases. Such processes in these deep subsurface hydrocarbon reservoirs are more Once sediment samples were returned to the laboratory, likely to involve anaerobic microorganisms (Head et al., sterile seawater was added to the sediment in a 1:1 ratio 2003). Formolo et al. (2004) used the isotopic composition (volume to volume) and the resulting slurry was homoge- of carbonate-associated gas-hydrates in the Gulf of Mexico nized. Samples were prepared in triplicate as follows. Fifty to argue for the importance of C2+ hydrocarbons to car- milliliters of the sediment slurry was added to a 150 mL bon cycling around thermogenic gas hydrates. Few other serum bottle, which was crimp-sealed with a gas-tight butyl investigations have used the isotopic composition of C2– rubber stopper. The headspace in the serum bottle was sub- C4 to investigate the biogeochemical cycling of these gases, sequently purged with ultra-high purity (UHP) N2 gas for and no fractionation factors are currently available for 5 min at approximately 150 mL min1. Each sample was their microbial consumption. then amended with the addition of and hydrocar- In this work, we investigate the aerobic microbial con- bon gases to the headspace. The proportions of hydrocar- sumption of C1–C4 hydrocarbons in marine sediment sam- bon and oxygen were chosen based on the stoichiometry ples from natural seeps of thermogenic gas. Specifically, we for complete hydrocarbon oxidation, such that the hydro- seek to determine whether C1–C4 oxidation occurs in sim- carbon gas would be the limiting reactant but that the ini- ulated seep environments, how the C1–C4 composition im- tial O2 would not exceed 20% of the headspace gas. The pacts patterns of gas consumption, the magnitude of expected reaction stoichiometries for the complete aerobic carbon isotope fractionation during microbial oxidation oxidation of C1–C4 hydrocarbons, exclusive of biomass of C1–C4, and the magnitude of hydrogen isotope fraction- generation, are as follows: ation associated with the microbial oxidation of C1–C3. CH4 þ 2O2 ! CO2 þ 2H2O ð1Þ

2. Methods C2H6 þ 3:5O2 ! 2CO2 þ 3H2O ð2Þ

C3H8 þ 5O2 ! 3CO2 þ 4H2O ð3Þ 2.1. Sediment source and collection C4H10 þ 6:5O2 ! 4CO2 þ 5H2O ð4Þ

The sediment samples used in this study were collected After these additions a small quantity of UHP N2 was added from the surface sediment layer at two distinct seep envi- as an overpressure such that subsequent sampling did not ronments located offshore Santa Barbara, CA. The first site draw a vacuum. Two types of incubations were performed: is informally referred to as Shane Seep (34N, 24.3700; those with only one hydrocarbon gas and those with a mix 119W, 53.4280) and is located at 21 m water depth in the of C1–C4. The average (±SD) initial mixing ratios for C1– Coal Oil Point seep field (Allen et al., 1970). The composi- C4 in the individual gas incubations were 7.0% ± 0.8 for tion of gas emitted from the seafloor at Shane Seep is C1 treatments, 3.8% ± 0.3 for C2 treatments, 2.1% ± 0.4 79.7% C1, 15.2% CO2, 0.87% C2, 0.72% C3, and 0.23% for C3 treatments, and 0.9% ± 0.3 for C4 treatments. The C4. Shane Seep is estimated to emit approximately 1.9 tons average (±SD) initial mixing ratios for the mixed gas incuba- C1 per day to the atmosphere (Washburn et al., 2005). The tions were 0.91% ± 0.03 for C1, 0.87% ± 0.03 for C2, second site is referred to as Brian Seep (34N, 24.1090; 0.71% ± 0.1 for C3, and 0.43% ± 0.05 for C4. 274 F.S. Kinnaman et al. 71 (2007) 271–283

Incubations for all four gases were performed in tripli- quantitatively converted to CO2 before passing into the cate for sediment from each seep, with duplicate killed con- mass spectrometer. 13 12 trols and blanks, totaling 19 treatments per seep. Killed C/ C ratios were measured relative to CO2 reference controls were heat sterilized twice (121 C for 15 min) gas, which was calibrated against NBS19 (NIST SRM and contained headspace gases in similar proportion to #8544), and are expressed in per mil notation (&) relative the mixed gas incubations. Blank treatments consisted of to the international standard V-PDB (Pee Dee belemnite sediment slurry without any exogenous hydrocarbon, and carbonate, as established by the International Atomic served as a control against possible hydrocarbon produc- Energy Agency [IAEA] in Vienna, Austria; Coplen, tion. All experimental and control treatments were incubat- 1995). The CO2 reference gas was introduced into the mass ed at 15 C in the dark on an orbital shaker set at 35 rpm. spectrometer at the beginning and end of each GC run. Overall precision of the GC–IRMS system was ±0.30 per 2.3. Mixing ratio and stable isotope analyses mil for samples with >0.1% hydrocarbon in the sample headspace, based on replicate injections of standard. Preci- The gas headspace of each experimental treatment was sion was lower for samples ranging from 0.01% to 0.05% sampled at intervals of 5–7 days, and samples were stored and are displayed as error bars in Figs 1 and 2. This preci- for later analysis of gas composition as well as hydrocar- sion was determined by comparing the amplitude of a given bon d13C and dD. Gas samples were removed by syringe peak with a standard of similar amplitude, and applying and transferred to 5 mL crimp-sealed serum bottles. In or- the 1r error determined from replicate injections of the der to avoid diluting the sample during transfer, the storage latter. bottles were filled completely with water, which was then Ratios of D/H were measured in all samples using a displaced with the incoming gas sample using a dual syr- Thermo Finnigan model Delta XL-Plus isotope ratio mass inge displacement technique. Once the gas samples were spectrometer (IRMS) at UC Irvine. The basic methodology collected, they were stored in the dark at room temperature (Hilkert et al., 1999; Rice et al., 2001) is similar to 13C/12C prior to analysis. analyses, as described above. Differences include a helium Hydrocarbon mixing ratios were determined by gas carrier flow rate of 2.6 mL min1, the use of an Agilent chromatography (GC), using a Shimadzu GC-14A model 6890 GC, and the use of a pyrolysis reactor (set at equipped with a flame ionization detector (FID). C1–C4 1450 C) in place of a unit. Multiple aliquots compounds were resolved on the GC–FID with isothermal of each sample were injected with a gas-tight syringe (ali- 1 60 C runs using N2 as the carrier gas (20 mL min flow quot size ca. 100–300 lL at STP depending on concentra- 0 100 rate) through a 12 8 packed column (n-octane on Res- tion of the hydrocarbon) into the inlet port of the GC. Sil C). Headspace levels are reported as % of headspace Chromatography was performed isothermally using a (v/v). Standard error in the mixing ratios is ±3% based 25 m PoraPLOT Q capillary column at 30 C for both C1 on replicate injections of a standard. Samples were re- and C2 samples and at 100 C for C3 samples. moved by syringe from the storage bottles using a dual syr- A working reference gas cylinder of UHP H2 from Oxy- inge technique by displacing the gas headspace with water gen Services Co. (Costa Mesa, CA) was assigned a value by and subsequently passing 0.7 mL through a 100 lL sample comparison with H2 gas purchased from Oztech Gas Co. loop. O2 levels were monitored using a 3000A microGC with an assigned value of 165.3& versus v-SMOW. Our (Agilent) equipped with a thermal conductivity detector. working reference gas for the cf-GC/IRMS instrument 13 12 Stable carbon isotope ratios ( C/ C) of C1–C4 hydro- was assigned the value of 169.4& for H2 versus SMOW. carbons were measured for selected time points by contin- Precision of measurement for clean dry calibrated stan- uous flow isotope ratio mass spectrometry (IRMS). This dards of CH4 in air are ±1.1& for dD. Greater error was was achieved using a Delta Plus XP IRMS coupled with noted for higher hydrocarbons, and standard deviations a Trace Gas chromatograph via a GC/Combustion III (1r) for multiple determinations of each sample aliquot interface (system components all from Thermo Finnigan, are provided as error bars in Fig. 3. now ThermoElectron), at the UCSB Marine Science Insti- tute’s Analytical Laboratory. A 30 m · 0.32 mm ID Car- 2.4. Determination of fractionation factors bonPlot capillary column (Agilent Technologies) was used with helium as the carrier gas at a constant flow rate Fractionation factors were determined using the follow- of 1.0 mL/min. The size of the injected gas sample was ing equation, which describes the fractionation that occurs 0.05–1.0 mL, depending on the mixing ratio. Individual in a single step unidirectional reaction in a closed isotopic hydrocarbons were resolved using an isothermal tempera- system (Mariotti et al., 1981). ture setting: 60 C for C1, 110 C for C2, 160 C for C3 3 3 10 df þ1 and 210 C for C4. For samples with C1–C4 hydrocarbon 10 ln 3 e ¼ 10 diþ1 ð5Þ mixtures, a temperature program starting at 60 C for two lnð1 f Þ minutes and ramping to 250 Cat20C min1 was used. 13 The eluent from the GC passes into the 950 C Pt/Ni/Cu where di is the d C of the introduced hydrocarbon in the ini- 13 combustion reactor, in which the organic components are tial headspace, df is the d C of the hydrocarbon in the head- Aerobic isotope fractionation of C1–C4 hydrocarbons 275 space at subsequent time points, e is the fractionation factor halt in oxidation for C1–C3 in samples collected from or ‘per mil enrichment factor’ of the product (the aggregate Shane Seep corresponds to depletion of O2 in the experi- of biomass and CO2) relative to the hydrocarbon substrate, mental bottles to levels below our detection limit of and f is the fractional yield or fraction of hydrocarbon con- 0.015% (data not shown). O2 depletion was presumably sumed at the given time. The fractionation factor can be cal- related to the microbial oxidation of native organic materi- culated by plotting the numerator of the expression on the al from the sediment, as oxygen was added in stoichiome- right side of Eq. (5) against the denominator (ln(1 f)) tric excess. for a series of time points and performing a linear regression Fig. 2A–D displays the time course changes in head- to solve for the slope (Chu et al., 2004). e is most accurate at space C1 (Fig. 2A), C2 (Fig. 2B), C3 (Fig. 2C), and C4 d values close to 0& and is related to a, another notation (Fig. 2D) levels for the treatments involving a mixture of commonly used for reporting fractionation factors, by the C1, C2, C3 and C4 hydrocarbon gases. Results indicate following equation: (Mariotti et al., 1981) that C3 and C4 are oxidized more rapidly than C1 and C2, and the rate of C1 and C2 oxidation increases after e ¼ða 1Þ1000 ð6Þ C3 and C4 are consumed. The staggered timing of C1 For ease of use, e is used throughout this paper. The isoto- and C2 oxidation indicates possible preference of C3 and pic fractionation associated with dissolution of hydrocar- C4 by hydrocarbon oxidizing bacteria. Premature O2 bons in water is expected to be small and is not depletion was again noted in the mixed gas experimental considered in these experiments. treatments originating from Shane Seep. The derivation of Eq. (5) assumes a closed isotopic sys- tem. In these experiments gas samples were removed period- 3.2. Fractionation of carbon isotopes ically, thus the system was not closed. In order to eliminate the systematic error caused by gas removal, f was calculated A trend toward carbon isotopic enrichment in the resid- as a ratio of mixing ratios and is thus independent of gas ual hydrocarbon gas is observed in all experimental treat- removal. However, the use of mixing ratios in place of par- ments concurrent with hydrocarbon consumption. The 13 tial pressure causes an error in f associated with changes in range of observed d C values was from 22.0 to the quantity of bulk gas relative to hydrocarbon substrate. +76.4& for C1, 32.1 to +4.2& for C2, 33.7 to In this case, the consumption of hydrocarbons and oxygen 14.6& for C3, and 30.8 to 18.3& for C4. This behav- and the production of CO2 all change the quantity of bulk ior is typical of a closed isotopic system and can be treated gas and thus the mixing ratio of residual hydrocarbons. similar to a Rayleigh distillation (Coleman et al., 1981). The consumption of oxygen and hydrocarbon are partially For the incubations involving only one hydrocarbon gas, 13 offset by the production of carbon dioxide. This gives rise Fig. 1E–H displays the change in d C of headspace C1 to slight systematic error in the reported concentrations. (Fig. 1E), C2 (Fig. 1F), C3 (Fig. 1G) and C4 (Fig. 1H) For C1 the error is 2% for f = 0.1 and 16% for f = 0.9, as a function (ln(1 f)) of the fraction consumed. Sets of and decreases with increasing chain length from C1 to C4. triplicate samples from each of the two seep environments 13 This systematic error is incorporated into the results by cal- are displayed on each panel. The d C of each hydrocarbon culating the error for a given amount of consumption and was analyzed only at select intervals, for a total of 3–10 iso- adjusting the reported fraction of hydrocarbon consumed topic measurements per bottle (average = 6.0). (f), according to the stoichiometry given in Eqs. (1)–(4). Fractionation factors determined from each experimen- tal bottle are provided in Table 1, along with r2 values and 3. Results overall averages. The initial d13C was similar for all six treatments for individual hydrocarbon gases, to within 3.1. Hydrocarbon oxidation ±5.5& for C1, ±0.5& for C2, ±0.3& for C3 and ±1.4& for C4; therefore the graphical representation of the frac- Hydrocarbon oxidation was noted in all experimental tionation factors given in Fig. 1 includes all d13C values treatments when O2 was present. No hydrocarbon oxida- for each hydrocarbon gas. Averages of the carbon isotopic tion was noted in killed controls, and no hydrocarbon pro- fractionation factors (ec) for the individual substrate incu- duction was noted in the un-amended sediment (data not bations are 25.8& ± 1.6 for C1, 9.3& ± 1.1 for C2, shown). Fig. 1A–D displays the time course changes in 4.9& ± 1.1 for C3, and 2.4& ± 1.1 for C4. The slope headspace levels of C1 (Fig. 1A), C2 (Fig. 1B), C3 of the aggregate determined from the regression lines in (Fig. 1C) and C4 (Fig. 1D) for the incubations involving Fig. 1 are slightly different from these values; the difference only one hydrocarbon gas. Some variability is apparent is 1.1& for C1, 0.9& for C2, 0.0& for C3, and 0.4& for in the rates of C1–C4 oxidation, but the general trend to- C4, and is within the standard error for each gas. In the ward consumption is clear. In the case of C4 consumption case of C1 and C3 there is an apparent discontinuity in was nearly quantitative for each replicate and each seep the fractionation as substrate levels become low (at environment. In the case of C1–C3, consumption occurred f 0.87 or below 2.0 on the ln (1 f) scale, equivalent to a greater extent in the samples collected from Brian to 0.4% C1 and 0.03% C3 in the headspace). These values Seep, than in those collected from Shane Seep. The abrupt are treated separately and are indicated by the dashed lines 276 F.S. Kinnaman et al. 71 (2007) 271–283

Fig. 1. Incubation of seep sediments with only one hydrocarbon species. (A–D) Display the time course change in C1–C4, respectively, and plots (E–H) display the d13C of the corresponding gas as a function (ln (1 f)) of the fraction consumed. Sets of triplicate samples from each of the two seep environments are displayed on each panel. Solid symbols indicate samples from Shane Seep and open symbols indicate samples from Brian Seep. The d13C of each hydrocarbon was analyzed only at select intervals. Dashed lines indicate an apparent change in fractionation. Error bars are smaller than the corresponding symbol, except as noted. Sets of triplicate samples from each of the two seep environments are displayed on each panel. in Fig. 1E and G. The apparent fractionation factors asso- bon gas individually. The slopes of the regression lines for ciated with the lower hydrocarbon levels are 19.5 each experimental treatment are detailed in Table 1, along (r2 = 0.911) for C1 and 3.7 for C3 (r2 = 0.999). with r2 values and averages from both seep locations. Aver- 13 Fig. 2E–H displays the d C of headspace C1 (Fig. 1E), ages of the carbon isotopic fractionation factors (ec) for the C2 (Fig. 1F), C3 (Fig. 1G) and C4 (Fig. 1H) as a function mixed C1–C4 substrate incubations (Table 1) differ some- (ln(1 f)) of the fraction consumed—for the incubations what compared with ec determined from incubations with involving a mixture of all four hydrocarbon gases. Sets of only one hydrocarbon; the difference is 7.4& for C1, triplicate samples from each of the two seep environments 2.6& for C2, 0.2& for C3, and 0.8& for C4. In the case are displayed on each panel and all treatments are repre- of C1 there is again an apparent discontinuity in the frac- sented on each panel. As with the pure gas treatments, tionation as substrate levels decrease (at f 0.25; equiva- the initial d13C was similar for each individual gas in all lent to 0.29 on the ln (1 f) scale or 0.75% headspace six treatments, to within ±3.0& for C1, ±1.0& for C2 C1), as indicated by the dashed line in Fig. 2E. The appar- and C3, and ±1.6& for C4 and linear regressions given ent fractionation factor associated with the lower C1 level in Fig. 2 include all relevant d13C values for each hydrocar- is 11.9 (r2 = 0.726). This fractionation change also corre- Aerobic isotope fractionation of C1–C4 hydrocarbons 277

Fig. 2. Incubation of seep sediment with a mixture of hydrocarbon species. (A–D) Display time course changes in headspace C1, C2, C3 and C4, respectively, and plots E–H display the change in d13C of the corresponding gas as a function (ln (1 f)) of the fraction consumed. Sets of triplicate samples from each of the two seep environments are displayed on each panel. Solid symbols indicate samples from Shane Seep and open symbols indicate samples from Brian Seep. The d13C of each hydrocarbon was analyzed only at select intervals. The dashed line in (E) corresponds to an apparent change in fractionation. Error bars are smaller than the corresponding symbol, except as noted. sponds to exhaustion of the C2+ hydrocarbons and is not- enrichments, fractionation factors were determined in a ed only in samples from Brian Seep O2 depletion pre- manner similar to a Rayleigh distillation. eH values aver- vented C1 from being consumed to low levels in samples aged from both Shane Seep and Brian Seep incubations from Shane Seep. In contrast to the incubations involving were 238.2& ± 115.6 for C1, 61.9& ± 8.3 for C2, only one hydrocarbon in the headspace, there is no appar- and 15.1& ± 1.9 for C3. Of the samples analyzed, one ent discontinuity in fractionation associated with C3. C3 measurement was discarded as it was anomalously depleted relative to all other samples in the time series. 3.3. Fractionation of hydrogen isotopes Fractionation factors for C1 differed between sediment from the two locations, giving rise to a large standard er- A trend toward deuterium enrichment in the residual ror in the reported average. The eH value for incubations hydrocarbon gas was observed in experimental treatments of sediment taken from Shane Seep was 156.4&, while concurrent with hydrocarbon consumption. Due to analyt- that from Brian Seep was 319.9&. The dD values of ical limitations only results from select C1–C3 treatments residual hydrocarbons were found to correlate with the are available (Table 2). d13C of the corresponding sample. Fig. 3 presents a plot The range of observed dD values is 86.2& to of d13C versus dD (C–D plot), inclusive of all samples +317.5& for C1, 347.0& to 252.9& for C2, and for which both analyses were performed. The slopes of 143.7& to 120.5& for C3. As with the carbon isotopic the linear regressions are indicative of the simultaneous 278 F.S. Kinnaman et al. 71 (2007) 271–283

Fig. 3. Simultaneous enrichment in dD and d13C (C/D plot) for C1–C3 during microbial consumption. Solid symbols denote incubations originating from Shane Seep, while open symbols denote incubations originating from Brian Seep. Temporal trajectories trend towards the lower right corner. Short dashed lines indicate the approximate trajectory predicted for methanotrophy by Whiticar (1999), and long dashed lines indicate the trajectory predicted by Coleman et al. (1981). For C1 and C2 the solid black lines and corresponding equations incorporate data from both seep locations. For C3 separate regressions were performed for the two seep locations due to differing initial dD values, and the average of the two regressions is provided. Error bars are smaller than the corresponding symbol, except as noted.

Table 1 a Carbon isotope fractionation factors (ec) for the microbial oxidation of C1–C4 in units of & Individual gas incubationsb Gas mixture incubationsc Shane seep Brian seep Shane seep Brian seep e r2 e r2 e r2 e r2

CH4 25.1 0.948 25.8 0.931 33.3 0.622 19.7 0.721 26.5 0.816 25.4 0.989 25.8 0.935 32.0 0.702 28.3 0.869 23.6 0.924 50.5 0.902 31.7 0.858 Average 26.6 ± 1.6 24.9 ± 1.2 36.5 ± 12.7 27.8 ± 7.0 epure average: 25.8 ± 1.6 emix average: 32.2 ± 10.3 C2H6 9.8 0.983 8.9 0.996 8.1 0.800 6.2 0.989 11.2 0.992 8.2 0.998 6.6 0.919 7.9 0.990 9.3 0.992 8.4 0.999 5.9 0.841 5.6 0.990 Average 10.1 ± 1.0 8.5 ± 0.4 6.9 ± 1.1 6.6 ± 1.2 epure avg. = 9.3 ± 1.1 emix avg. = 6.7 ± 1.0 C3H8 4.0 0.993 5.4 0.976 4.8 0.948 3.9 0.983 5.9 0.999 5.6 0.971 5.8 0.994 5.1 0.997 3.0 1 5.2 0.957 4.2 0.986 5.0 0.998 Average 4.3 ± 1.5 5.4 ± 0.2 4.9 ± 0.8 4.7 ± 0.7 epure avg. = 4.9 ± 1.1 emix avg. = 4.8 ± 0.7 C4H10 1.3 0.844 3.4 0.988 2.2 0.634 3.5 0.974 1.6 0.880 3.5 0.999 3.4 0.996 3.3 0.965 2.8 0.992 3.6 0.997 Average 2.4 ± 1.0 3.4 2.8 ± 0.7 3.5 ± 0.1 epure avg. = 2.4 ± 1.1 emix avg. = 3.2 ± 0.6 a ec for C1–C4 are reported in the abstract as averages of the epure and emix values (excluding Shane Seep for C1) and are: 26.5& ± 3.9 for C1, 8.0& ± 1.7 for C2, 4.8& ± 0.9 for C3, and 2.9& ± 0.9 for C4. b Refers to fractionation factors from incubations involving only one hydrocarbon gas (epure). c Refers to fractionation factors from incubations containing mixtures of C1–C4 (emix).

13 13 enrichment in d C and dD associated with aerobic micro- theoretical shifts in d CCH4 and dDCH4 as a result of C1 bial C1–C3 oxidation. These C/D enrichment factors (in oxidation and projected a C and H isotope ratio trajecto- units of &d13C/&dD) are calculated to be 0.114 for C1, ry of approximately 0.165 (dotted line on Fig. 3). Cole- 0.084 for C2, and 0.384 for C3. Whiticar (1999) discussed man et al. (1981) measured trajectories for two cultures Aerobic isotope fractionation of C1–C4 hydrocarbons 279

Table 2 reaction is approximately 175 times faster than the C1 reac- Hydrogen isotope fractionation factors (eh) for C1–C3 in units of & tion at 25 C(Bryukov et al., 2004). Shane seep Brian seep Details regarding the microbes responsible for oxidation 2 2 eh r eh r Average eh of C2–C4 in seep environments are lacking, though microbes capable of growth on C2 (Dworkin and Foster, 1958), C3 CH4 156.4 0.936 319.9 0.991 238.1 ± 115.6 C2H6 67.8 0.796 56.0 0.686 61.9 ± 8.3 (Perry, 1980; Stephens and Dalton, 1986), and C4 (Takah- C3H8 13.8 1.0 16.5 0.928 15.1 ± 1.9 ashi, 1980; Hou et al., 1983; Hamamura et al., 1999) have been isolated from other environments. The identity and activity of the organisms responsible for C2–C4 oxidation at different temperatures which ranged from 0.13 to .07 in seep environments is a subject of ongoing research in (dashed lines on Fig. 3). our lab, and several strains have been isolated from the incu- bations described here (Redmond and Valentine, 2005). 4. Discussion 4.2. Fractionation of stable isotopes 4.1. Microbial consumption of C1–C4 hydrocarbons 4.2.1. Methane Results from the experiments presented here clearly Numerous studies have reported carbon isotope fraction- demonstrate the aerobic consumption of C1–C4 hydrocar- ation factors for C1. Those reported here are slightly larger bons by natural microbial communities inhabiting shallow than many previously reported values (compiled in Ree- marine gas seeps. While the microbial consumption of C1, burgh, 2003; see also Grant and Whiticar, 2002 and Temple- or methanotrophy, has been the subject of numerous inves- ton et al., 2006). These results agree best with those obtained tigations, the microbial oxidation of C2 (ethanotrophy), C3 in incubations performed 10 C warmer (Barker and Fritz, (propanotrophy) and C4 (butanotrophy) has received little 1981; Coleman et al., 1981). No fractionation factors have attention—particularly with regard to their occurrence in been previously reported for sediment gathered from hydro- nature. The study sites in this work were chosen because carbon cold seep environments. The high level of fraction- 13 the ebullition of thermogenic hydrocarbon gas from the ation observed in this study is consistent with d C–CH4 sediments into oxic led us to hypothesize C1–C4 enrichments of as much as 60& found in sediment sur- oxidation was occurring. The consumption of C1–C4 com- rounding Shane Seep (Kinnaman and Valentine, 2004). pounds in fresh sediment slurries supports this hypothesis. Carbon isotope fractionation was inconsistent for C1 Methanotrophs are known to be active around marine during early stages of the mixed gas incubations. Table 1 gas seeps, and several previous reports detail aspects of displays notably high values of ec for two of the incuba- their distribution and activity (Bidle et al., 1999; Lanoil tions originating from Shane Seep. Methane consumption et al., 2001; Valentine et al., 2001; Joye et al., 2004). Many ceased prematurely in these incubations due to limitation methanotrophs are able to cometabolize higher alkanes, by O2, resulting in only a small change in f over the course but require the presence of C1 for growth (Berthe-Corti of the incubation. Large changes in f are needed to accu- and Fetzner, 2002). Results presented here (Figs. 1 and 2) rately determine ec, and may explain the high values report- clearly demonstrate the sustained consumption of C2–C4 ed in Table 1. For this case, considering the aggregate data independent of C1. Results from mixed-hydrocarbon treat- (Fig. 2E) seems a more accurate approach. Therefore ec for ments further indicate that C3 and C4 are largely con- C1 calculated from the mixed C1–C4 incubations originat- sumed prior to C1 and C2 being consumed, with some ing from Shane Seep were not assimilated into an overall indication that C2 is also consumed prior to C1. This pref- average as was done for C2–C4 incubations. Fractionation erential oxidation of C2–C4 hydrocarbons was noted in factors determined from other individual treatments (Table treatments from both seep environments. The reason for 1) were otherwise consistent with the fractionation deter- this apparent preference is not clear, and may be related mined from the aggregate data. to any of several factors including the substrate preferences The deuterium fractionation factors for C1 reported for different members of the microbial community, the in Table 2 fall within the range of previously reported expression of different enzyme systems within an organism, fractionation factors. Coleman et al. (1981) report e val- enzymatic affinity for different gases, competition for sub- ues of 103 ± 6 and 297 ± 2 for incubations at 11.5 strates and nutrients, or to the inherent metabolic capacity and 26 C, respectively, for an aerobic culture taken of the sediment samples. On purely thermodynamic from a freshwater well. Snover and Quay (2000) report grounds, the oxidation of C3 and C4 is favored over C1 hydrogen isotope effects of 99 ± 30 for oxidation of and C2. While alkanes are among the least chemically reac- atmospheric levels of C1 in grasslands and 66 ± 7 for tive organic compounds, C3 and C4 contain C–H bonds at forest soils. the secondary position (at the non-terminal carbon site) which are more reactive than C–H bonds at the primary 4.2.2. Ethane, propane and butane position (e.g., C1 and C2). In non-biological reactions, No carbon or hydrogen isotope fractionation factors have such as gas phase reactions of alkanes with OH, the C3 been reported for microbial consumption of C2–C4 alkanes 280 F.S. Kinnaman et al. 71 (2007) 271–283 to our knowledge. Previous investigations of hydrocarbon considered (Rees, 1973; Gelwicks et al., 1994; Hall et al., biodegradation in natural environments have focused on 1999; Hayes, 2001; Valentine et al., 2004; Templeton tracing organic contaminant sources and fates in geochemi- et al., 2006). Different pure cultures can show great varia- cal systems and have resulted in reports of fractionation fac- tions in fractionation for similar biological processes. Mor- tors for other multi-carbon substrates. Examples of these asch et al. (2002) found varying levels of carbon isotope studies include the anaerobic biodegradation of toluene by fractionation during aerobic toluene degradation for bacte- microbes from soil (ec = 1.5 ± 0.2; Meckenstock et al., rial strains with different enzyme mechanisms. Hall et al. 1999), the aerobic biodegradation of by Acinetobac- (1999) concluded that the observed variation in fraction- ter and Burkholderia sp. (ec = 1.46 ± 0.06 and eh = ation during aerobic degradation of phenol and benzoate 12.8 ± 0.7 for Acinetobacter; ec = 3.53 ± 0.26 and was due to differences in bacterial metabolic pathways eh = 11.2 ± 1.8 for Burkholderia: Hunkeler et al., 2001), and growth rates. Conflicting reports of the amount of frac- and the aerobic biodegradation of vinyl chloride and trichlo- tionation occurring for polycyclic aromatic hydrocarbons roethylene (ec = 5.7 ± 1.1 and 1.1 ± 0.3, respectively; (PAH’s), summarized by Abrajano et al. (2004) imply that Chu et al., 2004). Stable carbon and hydrogen isotopic frac- enzymes catalyzing similar reactions in different organisms tionation of n-alkanes (C15–C18) has only been observed dur- display dissimilar fractionation factors. This is comparable ing heavy biodegradation (Sun et al., 2005). The to photosynthesis, where plants with C3,C4 and crassula- fractionation factors for C2–C4 from this work are generally cean acid (CAM) pathways exhibit different 13 greater than for the larger molecules as outlined above, levels of C fractionation during CO2 fixation. For the aer- though the intrinsic carbon isotope effect (see Morasch obic oxidation of C1–C4, the first step in the biochemical et al., 2004) appears similar for C3 (14.4 ± 2.7), C4 pathway is presumably a uni-directional reaction (e.g., an (11.6 ± 3.6), toluene (10.5 ± 1.4), benzene (8.76 ± irreversible oxidation; (Stephens and Dalton, 1986; Hanson 0.36—for Burkholderia) and trichloroethylene (11.4 ± 2.2). and Hanson, 1996; Hamamura et al., 1999; Johnson and Hyman, 2006), and thus the isotopic fractionation must 4.2.3. Variable fractionation be determined at or prior to this step. The discontinuity in Variability in the extent of isotope fractionation was ob- ec observed in this study is consistent with a transport-lim- served both between treatments originating from different ited situation. The extreme case of transport-limitation oc- locations and within individual treatments. Three of the curs when a cell goes from having bi-directional substrate experimental treatments (one involving only C1, one flux (e.g., substrate exchanges rapidly in and out of the cell involving only C3, and one involving C1 in a C1–C4 mix- relative to biochemical activation) to having uni-directional ture) exhibit apparent discontinuities in isotope fraction- substrate flux (e.g., all substrate entering the cell is con- ation near the end of the incubation (Figs. 1, 2E). In the sumed). In the case of bi-directional substrate flux the con- single hydrocarbon incubations (Fig. 1) these features were centration and isotopic composition of the substrate is interpreted from samples which were below 2.0 on the ln influenced by the fractionation associated with the trans- (1 f) scale, equivalent to 0.4% headspace for C1 (5 port into and out of the cell and the isotopic fractionation samples) or to 0.03% for C3 (3 samples), and also corre- associated with the biological process, while in uni-direc- spond to slower rates of oxidation and decreased oxygen tional flow the biological isotopic fractionation is zero levels. For the mixed gas incubation this feature was inter- and the only fractionation expressed is during transport preted from 9 samples which were below 0.7% C1 head- into the cell. In actuality these two extreme scenarios repre- space. These fractionation factors are illustrated with a sent end-members of a continuum, with substrate concen- dashed line in Figs. 1 and 2, and each is smaller than the tration, transport rates and cell-specific catabolism as the corresponding fractionation determined using data collect- primary controls. While this biochemical explanation is ed earlier in the experiment. This discontinuity was not consistent with the experimental observations, other factors consistent across all experimental treatments. For example, may also be important including oxygen levels and shifts in no variation in isotope fractionation was observed for C3 the structure of the microbial community. For example, dif- in the mixed gas incubations. Incubations involving C2 ferences in C3 fractionation between the C3-only incuba- and C4 did not display a discontinuity. tions and mixed gas incubations may be due to differences Hydrogen isotope fractionation was found to differ for C1 in the microbial populations, or to the impact of cometab- between sediments from the two seep environments. The sed- olism. Differences in the rate of hydrocarbon oxidation (as iment originating from Brian Seep displayed an eh of 319.9 interpreted in the time course changes in C1–C4 hydrocar- while that from Shane Seep displayed an eh of 156.4. The bons in Figs. 1 and 2) between replicate incubations do not underlying cause for this difference is not apparent from seem to cause variations in fractionation in these results. the experimental results. This magnitude of variability in hydrogen isotope fractionation has been observed previous- 4.2.4. Application of fractionation factors to natural systems ly, associated with both the consumption (Coleman et al., Distributions of d13CanddD in hydrocarbon gases have 1981) and production (Valentine et al., 2004) of C1. previously been applied to approximate the extent of Variable isotope fractionation has been observed in hydrocarbon oxidation. This approach is well established other systems, and a variety of potential causes have been for C1 (Schoell, 1988; Whiticar, 1999), but less so for Aerobic isotope fractionation of C1–C4 hydrocarbons 281

C2–C4 (James, 1983; Boreham et al., 2001; Plummer et al., Bilek, R.S., Tyler, S.C., Sass, R.L., Fisher, F.M., 1999. Differences in CH4 2005). The fractionation factors presented in this work are oxidation and pathways of production between rice cultivars deduced specific to oxic marine sediments at 15 C, and care must be from measurements of CH4 flux and delta C-13 of CH4 and CO2. Global Biogeochem. Cycles 13 (4), 1029–1044. taken in applying these values to other environmental sys- Boreham, C.J., Hope, J.M., Hartung-Kagi, B., 2001. Understanding tems. Specifically, temperature, hydrocarbon distributions source distribution and preservation of Australian natural gas: a and dominant terminal electron-accepting processes may geochemical perspective. APPEA J. 41 (1 200), 523–547. each impact the magnitude and expression of isotopic frac- Boreham, C.J., Hope, J.M., Edwards, D.S., Logan, G.A., 2004. D/H tionation in a given environment. The reported variability ratios of Australian natural gases. In: McIntyre, C. (Ed.) : Interfaces and Interactions. Combined National Conference of in carbon isotope fractionation associated with C1 oxida- the Australian Organic Geochemists and the International Humic tion (Reeburgh, 2003) in different environments serves as Substances Society, 16–19th February 2004, Blue Mountains, NSW, such an example. The fact that fractionation factors pre- Australia. Conference Program and Abstracts, p. 43. Published by sented here for C2–C4 were similar between two distinct CSIRO , ISBN Number: 0-643-09036-3. seep environments is promising, but does not guarantee Bryukov, M.G., Knyazev, V.D., Lomnicki, S.M., McFerrin, C.A., Dellinger, B., 2004. Temperature-dependent kinetics of the gas-phase uniformity in anoxic, terrestrial or subsurface environ- reactions of OH with Cl2,CH4, and C3H8. J. Phys. Chem. A 108 (47), ments. Despite these shortcomings we encourage the appli- 10464–10472. cation of these fractionation factors to estimate the extent Chu, K.H., Mahendra, S., Song, D.L., Conrad, M.E., Alvarez-Cohen, L., of C2–C4 oxidation in different systems, with an under- 2004. Stable carbon isotope fractionation during aerobic biodegrada- standing of the associated caveats, and the realization that tion of chlorinated ethenes. Environ. Sci. Technol. 38 (11), 3126–3130. Cicerone, R.J., Oremland, R.S., 1988. Biogeochemical aspects of atmo- future investigations may refine the relevant fractionation spheric methane. Global Biogeochem. Cycles 2 (4), 299–327. factors for different settings. Clarkson, T.S., Martin, R.J., Rudolph, J., 1997. Ethane and propane in the Southern marine troposphere. Atmos. Environ. 31 (22), 3763–3771. Colby, J., Stirling, D.I., Dalton, H., 1977. Soluble methane mono- Acknowledgments oxygenase of Methylococcus–Capsulatus-(Bath)—ability to oxygenate normal-alkanes, normal-alkenes, ethers, and alicyclic, aromatic and We wish to thank Molly Redmond, Chris Boreham and heterocyclic-compounds. Biochem. J. 165 (2), 395–402. two anonymous reviewers for their input during the writing Coleman, D.D., Risatti, J.B., Schoell, M., 1981. Fractionation of carbon of this paper. We also wish to thank Luis Busso, Justine and hydrogen isotopes by methane-oxidizing bacteria. Geochim. Cosmochim. Acta 45 (7), 1033–1037. Kimball and Bob Petty for laboratory assistance. Funding Coplen, T.B., 1995. New IUPAC guidelines for the reporting of stable for this work was provided by the NSF through a CA- hydrogen, carbon, and oxygen isotope-ratio data. J. Res. Natl. Ins. REER award in Chemical Oceanography (OCE- Stand. Technol. 100 (3), 285. 0447395), and by the American Chemical Society’s Petro- Dedysh, S.N., Knief, C., Dunfield, P.F., 2005. Methylocella species are leum Research Fund through a type G starter grant facultatively methanotrophic. J. Bacteriol. 187 (13), 4665–4670. Dworkin, M., Foster, J.W., 1958. Experiments with some microorgan- (40643-G2). This work has also benefited from a major re- isms which utilize ethane and hydrogen. J. Bacteriol. 75 (5), 592– search instrumentation award by the National Science 603. Foundation (ATM-9871077). Energy Information Administration, 2004. International Energy Outlook 2004. Office of Integrated Analysis and Forecasting, U.S. Department of Energy. Associate editor: David R. Cole Etiope, G., Klusman, R.W., 2002. Geologic emissions of methane to the atmosphere. Chemosphere 49 (8), 777–789. References Formolo, M.J., Lyons, T.W., Zhang, C.L., Kelley, C., Sassen, R., Horita, J., Cole, D.R., 2004. Quantifying carbon sources in the formation of Allen, A.A., Schleuter, R.S., Mikolaj, P.G., 1970. Natural oil seepage at authigenic carbonates at gas hydrate sites in the Gulf of Mexico. Chem. Coal Oil Point, Santa Barbara, California. Science 170, 974–977. Geol. 205 (3-4), 253–264. Abrajano, T., Yan, B., O’Malley, V., 2004. High molecular weight Fung, I., John, J., Lerner, J., Matthews, E., Prather, M., Steele, L.P., petrogenic and pyrogenic hydrocarbons in aquatic environments. In: Fraser, P.J., 1991. 3-Dimensional model synthesis of the global Sherwood Lollar (Ed.), Treatise on , vol. 9. Elsevier- methane cycle. J. Geophys. Res. Atmos. 96 (D7), 13033–13065. Pergamon, pp. 475–509. Gelwicks, J.T., Risatti, J.B., Hayes, J.M., 1994. Carbon-isotope effects Arp, D.J., 1999. Butane metabolism by butane-grown ‘Pseudomonas associated with aceticlastic methanogenesis. Appl. Environ. Microbiol. butanovora’. Microbiology-UK 145, 1173–1180. 60 (2), 467–472. Ashraf, W., Mihdhir, A., Murrell, J.C., 1994. Bacterial oxidation of Grant, N.J., Whiticar, M.J., 2002. Stable carbon isotopic evidence for propane. FEMS Microbiol. Lett. 122 (1–2), 1–6. methane oxidation in plumes above Hydrate Ridge, Cascadia Oregon Atlas, R.M., 1995. Bioremediation of petroleum pollutants. Int. Biodete- Margin. Global Biogeochem. Cycles 16 (4), 71-1–71-13. rior. Biodegrad. 35 (1-3), 317–327. Greenberg, J.P., Zimmerman, P.R., 1984. Nonmethane hydrocarbons in Barker, J.F., Fritz, P., 1981. Carbon isotope fractionation during remote tropical, continental, and marine atmospheres. J. Geophys. Res. microbial methane oxidation. Nature 293 (5830), 289–291. Atmos. 89 (ND3), 4767–4778. Berthe-Corti, L., Fetzner, S., 2002. Bacterial metabolism of n-alkanes and Gupta, M., Tyler, S., Cicerone, R., 1996. Modeling atmospheric delta 13C– ammonia under oxic, suboxic and anoxic conditions. Acta Biotechnol. CH4 and the causes of recent changes in atmospheric CH4 amounts.J. 22 (3-4), 299–336. Geophys. Res. Atmos. 101 (D17), 22923–22932. Bidle, K.A., Kastner, M., Bartlett, D.H., 1999. A phylogenetic analysis of Hall, J.A., Kalin, R.M., Larkin, M.J., Allen, C.C.R., Harper, D.B., 1999. microbial communities associated with methane hydrate containing Variation in stable carbon isotope fractionation during aerobic marine fluids and sediments in the Cascadia margin (ODP site 892B). degradation of phenol and benzoate by contaminant degrading FEMS Microbiol. Lett. 177 (1), 101–108. bacteria. Org. Geochem. 30, 801–811. 282 F.S. Kinnaman et al. 71 (2007) 271–283

Hamamura, N., Arp, D.J., 2000. Isolation and characterization of alkane- Lanoil, B.D., Sassen, R., La Duc, M.T., Sweet, S.T., Nealson, K.H., 2001. utilizing Nocardioides sp. strain CF8. Fems Microbiol. Lett. 186 (1), Bacteria and Archaea physically associated with Gulf of Mexico gas 21–26. hydrates. Appl. Environ. Microbiol. 67 (11), 5143–5153. Hamamura, N., Page, C., Long, T., Semprini, L., Arp, D.J., 1997. Lassey, K.R., Lowe, D.C., Brenninkmeijer, C.A.M., Gomez, A.J., 1993. cometabolism by butane-grown CF8, Pseudomonas but- Atmospheric methane and its carbon isotopes in the southern- anovora, and Mycobacterium vaccae JOB5 and methane-grown Meth- hemisphere—their time-series and an instructive model. Chemosphere ylosinus trichosporium OB3b. Appl. Environ. Microbiol. 63 (9), 3607– 26 (1–4), 95–109. 3613. Mariotti, A., Germon, J.C., Hubert, P., Kaiser, P., Letolle, R., Tardieux, Hamamura, N., Storfa, R.T., Semprini, L., Arp, D.J., 1999. Diversity in A., Tardieux, P., 1981. Experimental determination of nitrogen kinetic butane monooxygenases among butane-grown bacteria. Appl. Environ. isotope fractionation: some principles; illustration for the denitrifica- Microbiol. 65 (10), 4586–4593. tion and nitrification processes. Plant Soil 62 (3), 413–430. Hanson, R.S., Hanson, T.E., 1996. Methanotrophic bacteria. Microbiol. Matthews, E., 2000. Wetlands. In: Khalil, M. (Ed.), Atmospheric Methane. Rev. 60 (2), 439–471. Springer-Verlag, pp. 202–233. Hayes, J.M. 2001. Fractionation of carbon and hydrogen isotopes in Meckenstock, R.U., Morasch, B., Warthmann, R., Schink, B., Annweiler, biosynthetic processes. In: Organic Geochemistry of Contemporaneous E., Michaelis, W., Richnow, H.H., 1999. C-13/C-12 isotope fraction- and Ancient Sediments, Rev Min Geochem vol. 43, pp. 225–277. ation of aromatic hydrocarbons during microbial degradation. Envi- Head, I.M., Jones, D.M., Larter, S.R., 2003. Biological activity in the deep ron. Microbiol. 1 (5), 409–414. subsurface and the origin of heavy oil. Nature 426 (6969), 344–352. Miller, J.B., Mack, K.A., Dissly, R., White, J.W.C., Dlugokencky, E.J., Heeschen, K.U., Collier, R.W., de Angelis, M.A., Suess, E., Rehder, G., Tans, P.P., 2002. Development of analytical methods and measure- Linke, P., Klinkhammer, G.P., 2005. Methane sources, distributions, ments of C-13/C-12 in atmospheric CH4 from the NOAA Climate and fluxes from cold vent sites at Hydrate Ridge, Cascadia Margin. Monitoring and Diagnostics Laboratory global air sampling network. Global Biogeochem. Cycles 19 (2). J. Geophys. Res. Atmos. 107 (D13), 11-1–11-15. Heeschen, K.U., Keir, R.S., Rehder, G., Klatt, O., Suess, E., 2004. Morasch, B., Richnow, H.H., Schink, B., Vieth, A., Meckenstock, R.U., Methane dynamics in the Weddell Sea determined via stable isotope 2002. Carbon and hydrogen stable isotope fractionation during ratios and CFC-11. Global Biogeochem. Cycles 18 (2), 1–18, GB2012. aerobic bacterial degradation of aromatic hydrocarbons. Appl. Envi- Hilkert, A.W., Douthitt, C.B., Schluter, H.J., Brand, W.A., 1999. Isotope ron. Microbiol. 68, 5191–5194. ratio monitoring gas chromatography mass spectrometry of D/H by Morasch, B., Richnow, H.H., Vieth, A., Schink, B., Meckenstock, R.U., high temperature conversion isotope ratio mass spectrometry. Rapid 2004. Stable isotope fractionation caused by glycyl radical enzymes Commun. Mass Spectrom. 13 (13), 1226–1230. during bacterial degradation of aromatic compounds. Appl. Environ. Hornafius, J.S., Quigley, D., Luyendyk, B.P., 1999. The world’s most Microbiol. 70 (5), 2935–2940. spectacular marine hydrocarbon seeps (Coal Oil Point, Santa Barbara Perry, J., 1980. Propane utilization by microorganisms. Adv. Appl. Channel, California): quantification of emissions. J. Geophys. Res. Microbiol. 26, 89–115. Oceans 104 (C9), 20703–20711. Plummer, R.E., Pohlman, J.W., Coffin, R.B., 2005. Compound-specific Hou, C.T., Patel, R., Laskin, A.I., Barnabe, N., Barist, I., 1983. stable carbon isotope analysis of low concentration complex hydro- Production of methyl ketones from secondary alcohols by cell- carbon mixtures from natural gas hydrate systems. Eos Trans. AGU 86 suspensions of C-2 to C-4 n-alkane-grown bacteria. Appl. Environ. (52) (Fall Meet. Suppl.), Abstract OS43A-0608. Microbiol. 46 (1), 178–184. Prather, M., Ehhalt, D., 2001. Atmospheric chemistry and greenhouse Hunkeler, D., Anderson, N., Aravena, R., Bernasconi, S.M., Butler, B.J., gases. In: Houghton, J. (Ed.), Climate Change 2001: The Scientific 2001. Hydrogen and carbon isotope fractionation during aerobic Basis. Cambridge University Press, Cambridge, pp. 239–287. biodegradation of benzene. Environ. Sci. Technol. 35 (17), 3462–3467. Quay, P., Stutsman, J., Wilbur, D., Snover, A., Dlugokencky, E., Brown, Hunt, J.M., 1996. Petroleum Geochemistry and Geology. W.H Freeman & T., 1999. The isotopic composition of atmospheric methane. Global Co.. Biogeochem. Cycles 13 (2), 445–461. James, A., 1983. Correlation of natural gas by use of carbon isotopic Quay, P.D., King, S.L., Stutsman, J., Wilbur, D.O., Steele, L.P., Fung, I., distribution between hydrocarbon components. Am. Assoc. Petrol. Gammon, R.H., Brown, T.A., Farwell, G.W., Grootes, P.M., Geol. Bull. 67 (7), 1176–1191. Schmidt, F.H., 1991. Carbon isotopic composition of atmospheric

James, A.T., Burns, B.J., 1984. Microbial alteration of subsurface CH4: fossil and biomass burning source strengths. Global Biogeochem. natural-gas accumulations. Am. Assoc. Petrol. Geol. Bull. 68 (8), Cycles 5, 25–47. 957–960. Redmond, M. Valentine, D.L., 2005. Isolation and characterization of Johnson, E.L., Hyman, M.R., 2006. Propane and n-butane oxidation by ethane, propane, and butane consuming bacteria from marine hydro- Pseudomonas putida GPo1. Appl. Environ. Microbiol. 72 (1), 950– carbon seeps. Eos Trans. AGU 86 (52) (Fall Meet. Suppl.), Abstract 952. B21A-1013. Joye, S.B., Boetius, A., Orcutt, B.N., Montoya, J.P., Schulz, H.N., Reeburgh, W.S., 2003. Global methane biogeochemistry. In: Keeling, R. Erickson, M.J., Lugo, S.K., 2004. The anaerobic oxidation of methane (Ed.), The Atmosphere, vol. 4. Elsevier-Pergamon, pp. 65–90. and sulfate reduction in sediments from Gulf of Mexico cold seeps. Reeburgh, W.S., Ward, B.B., Whalen, S.C., Sandbeck, K.A., Kilpatrick, Chem. Geol. 205 (3-4), 219–238. K.A., Kerkhof, L.J., 1991. Black-sea methane geochemistry. Deep-Sea Kanakidou, M., Singh, H.B., Valentin, K.M., Crutzen, P.J., 1991. A 2- Res. A Oceanogr. Res. Pap. 38, S1189–S1210. dimensional study of ethane and propane oxidation in the troposphere. Reeburgh W. S., Whalen S. C., and Alperin M. J. (1993) The role of J. Geophys. Res. Atmos. 96 (D8), 15395–15413. methanotrophy in the global methane budget. In Microbial Growth on Kessler, J.D., Reeburgh, W.S., Southon, J., Varela, R., 2005. Fossil C1 Compounds (ed. J. C. Murrell and D. P. Kelly), pp. 1-14. Intercept methane source dominates Cariaco Basin water column methane Ltd. geochemistry. Geophys. Res. Lett. 32 (12). Rees, C.E., 1973. Steady-state model for sulfur isotope fractionation in Kinnaman, F., Valentine D.L., 2004. Microbial methane oxidation in bacterial reduction processes. Geochim. Cosmochim. Acta 37 (5), 1141– marine gas seeps interpreted from two dimensional distributions of 1162. 13 d C-CH4. Eos Trans. AGU 85 (47) (Fall Meet. Suppl.), Abstract Rice, A.L., Gotoh, A.A., Ajie, H.O., Tyler, S.C., 2001. High-precision B43B-0164. continuous-flow measurement of delta C-13 and delta D of atmo-

Lacroix, A.V., 1993. Unaccounted for sources of fossil and isotopically spheric CH4. Anal. Chem. 73 (17), 4104–4110. enriched methane and their contribution to the emissions inventory: a Rudolph, J., 1995. The tropospheric distribution and budget of ethane.J. review and synthesis. Chemosphere 26, 507–557. Geophys. Res. Atmos. 100 (D6), 11369–11381. Aerobic isotope fractionation of C1–C4 hydrocarbons 283

Saito, T., Yokouchi, Y., Kawamura, K., 2000. Distributions of C2–C6 Valentine, D.L., Blanton, D.C., Reeburgh, W.S., Kastner, M., 2001. hydrocarbons over the western North Pacific and eastern Indian Water column methane oxidation adjacent to an area of active hydrate Ocean. Atmos. Environ. 34 (25), 4373–4381. dissociation, Eel River Basin. Geochim. Cosmochim. Acta 65 (16), Schoell, M., 1988. Multiple origins of methane in the earth. Chem. Geol. 71 2633–2640. (1-3), 1–10. Valentine, D.L., Chidthaisong, A., Rice, A., Reeburgh, W.S., Tyler, S.C., Scholes, M., Matrai, P., Andreae, M., Smith, K., Manning, M., 2003. 2004. Carbon and hydrogen isotope fractionation by moderately Biosphere–atmosphere interactions. In: Brasseur, G.P., Prinn, R.G., thermophilic methanogens. Geochim. Cosmochim. Acta 68 (7), 1571– Pszenny, A.P. (Eds.), Atmospheric Chemistry in a Changing World: An 1590. Integration and Synthesis of a Decade of Tropospheric Chemistry von Fischer, J.C., Hedin, L.O., 2002. Separating methane production and Research. Springer, pp. 48–52. consumption with a field-based isotope pool dilution technique. Global Singh, H., Zimmerman, P., 1992. Atmospheric distribution and sources of Biogeochem. Cycles 16 (3). nonmethane hydrocarbons. In: Nriagu, J. (Ed.), Gaseous Pollutants: Wang, J., Wang-Hsien, D., Chen, T., 2000. Source determination of light Characterization and Cycling. John Wiley & Sons, Inc., pp. 177–235. non-methane hydrocarbons by simultaneous multi-site sampling in a Snover, A.K., Quay, P.D., 2000. Hydrogen and carbon kinetic isotope metropolitian area. Chemos. Global Change Sci. 2, 11–22. effects during soil uptake of atmospheric methane. Global Biogeochem. Warneck, P. 1988. Chemistry of the Natural Atmosphere. Int’l Geophys. Cycles 14 (1), 25–39. Series vol. 41, Academic Press, San Diego. Stephens, G.M., Dalton, H., 1986. The role of the terminal and Washburn, L., Clark, J.F., Kyriakidis, P., 2005. The spatial scales, subterminal oxidation pathways in propane metabolism by bacteria. distribution, and intensity of natural marine hydrocarbon seeps near J. Gen. Microbiol. 132, 2453–2462. Coal Oil Point, California. Mar. Petrol. Geol. 22 (4), 569–578. Stevens, C.M., 1988. Atmospheric methane. Chem. Geol. 71, 11–21. Wenger, L.M., Davis, C.L., Isaksen, G.H., 2002. Multiple controls on Stevens, C.M., Rust, F.E., 1982. The carbon isotopic composition of petroleum biodegradation and impact on oil quality. SPE 71450, Society atmospheric methane. Journal of Geophysical Research-Oceans and of Petroleum Engineers. Atmospheres 87 (NC7), 4879–4882. Whiticar, M.J., 1990. A geochemical perspective of natural gas and Sun, Y.G., Chen, Z.Y., Xu, S.P., Cai, P.X., 2005. Stable carbon and atmospheric methane. Org. Geochem. 16 (1–3), 531–547. hydrogen isotopic fractionation of individual n-alkanes accompanying Whiticar, M.J., 1999. Carbon and hydrogen isotope systematics of biodegradation: evidence from a group of progressively biodegraded bacterial formation and oxidation of methane. Chem. Geol. 161 (1– oils. Org. Geochem. 36 (2), 225–238. 3), 291–314. Takahashi, J., 1980. Production of intracellular from n-butane Widdel, F., Rabus, R., 2001. Anaerobic biodegradation of saturated by Pseudomonas butanovora sp. nov. Adv. Appl. Microbiol. 26, 117– and aromatic hydrocarbons. Curr. Opin. Biotechnol. 12 (3), 259– 127. 276. Templeton, A.S., Chu, K., Alvarez-Cohen, L., Conrad, M.E., 2006. Zimov, S.A., Voropaev, Y.V., Semiletov, I.P., Davidov, S.P., Prosianni-

Variable carbon isotope fractionation expressed by aerobic CH4- kov, S.F., Chapin, F.S., Chapin, M.C., Trumbore, S., Tyler, S., 1997. oxidizing bacteria. Geochim. Cosmochim. Acta 70, 1739–1752. North Siberian lakes: a methane source fueled by Pleistocene carbon. Tyler, S.C., Ajie, H.O., Gupta, M.L., Cicerone, R.J., Blake, D.R., Science 277 (5327), 800–802. Dlugokencky, E.J., 1999. Stable carbon isotopic composition of Zengler, K., Richnow, H.H., Rossello´-Mora, R., Michaelis, W., Widdel, atmospheric methane: a comparison of surface level and free tropo- F., 1999. Methane formation from long-chain alkanes by anaerobic spheric air. J. Geophys. Res. Atmos. 104 (D11), 13895–13910. microorganisms. Nature 401, 266–269.