<<

Supplementary Information for

Isotopic evidence for the roles of quasi-equilibrium in natural gas formation

Authors: Nivedita Thiagarajan1*, Hao Xie1, Camilo Ponton1, Nami Kitchen1, Brian Peterson2 Michael Lawson3, Michael Formolo4, Yitian Xiao3, John Eiler1

Affiliations: 1Department of Geological and Planetary Sciences. California Institute of Technology. 2ExxonMobil Corporate Strategic Research, Annandale, NJ, USA 3ExxonMobil Exploration Company, Spring, TX, USA. 4ExxonMobil Upstream Research Company. *Correspondence to: [email protected]

This PDF file includes:

Supplementary text Figures S1 to S11 Datasets S1 to S5 SI References

1 www.pnas.org/cgi/doi/10.1073/pnas.1906507117

Supplementary Information Text

Sample Collection: In all but two instances, gas samples were collected either directly from the well head or at test separators of producing wells. At each well location, gases were collected into two separate industry standard 300 cm3 stainless steel cylinders connected by polytetrafluoroethylene sealed NPT pipe fittings to the well head/separator. Cylinders were pre-evacuated, baked and leak-tested prior to being dispatched to the field for sampling. Cylinders were purged for 5 minutes after being connected to the well head to flush any dead space between the cylinder and the sampling valve. One gas cylinder from each pair was shipped to either Isotech Laboratories or GeoMark Research, USA to determine the abundance and compound specific stable carbon and hydrogen isotopic composition of volatile hydrocarbon compounds (methane, ethane, propane, n- and i- butane and n- and i-pentane) in addition to major non-hydrocarbon gases (N2 and CO2) using standard procedures described later in the methods. The second cylinder was shipped to the California Institute of Technology for methane clumped isotope analysis. Samples from Keathley Canyon and Walker Ridge were collected downhole in MDT chambers during the testing of the reservoir formation post-drill. These MDT chambers were rocked and re-equilibrated at reservoir and temperature at Core Laboratories. Upon equilibration of the gas and oil, a small aliquot of oil was pushed from the cylinder, and the solution gas was then flashed at atmospheric conditions and collected in the cylinders described above.

Methane purification Methane (CH4) was purified from mixed gas samples using previously described methods (1). Approximately 60 µmol of natural gas was sampled. Gas samples were introduced to a vacuum glass line and exposed to liquid nitrogen to trap H2O, CO2, and H2S. The gases in the headspace (including CH4, O2, and N2) were then exposed and transferred to a 20 K cold trap. Residual gases in the headspace (mostly He and H2) were pumped away. The cold trap was then sealed, heated to 80 K, cooled to 45 K, and opened to vacuum to remove N2 and O2. This step was repeated until <2.67 Pa of gas remained in the cold trap at 45 K, corresponding to a purity of CH4 of 99.8%. The cryostat was then heated to 70 K, and CH4 was transferred to a PyrexTM breakseal containing molecular o sieve (EM Science; type 5A) immersed in liquid N2. Samples were heated to 135 C for 2 hours prior to introduction into the Thermo Finnegan MAT 253 Ultra (2).

Vibrational Frequency Calculations for C1-C5 species We calculated 13C/12C and D/H isotope equilibria in unlabeled and singly- substituted hydrocarbons using the Urey-Bigeleisen approach (3) with harmonic potential energy surfaces and frequencies calculated with density functional theory in the Gaussian 09 (Revision D) program (4) (SI Dataset 3). Calculations on all used the same methods and parameters. The B3LYP density functional was used with the Dunning correlation-consistent triple-zeta basis set with added diffuse functions: aug-cc-pvtz. Strict grid and convergence tolerances were used (opt=verytight, SCF=verytight, int=ultrafine) to ensure precision and reproducibility.

2

High-precision isotope equilibria from calculations beyond the harmonic approach are largely unexplored for hydrocarbons larger than propane and so the effects of anharmonicity and the low-energy torsional modes have not been quantified. Webb and Miller (5) found that, for propane at least, harmonic and path integral Monte Carlo results were very similar for calculations performed on the same potential energy surface and so we used harmonic calculations which were feasible for all of the molecules of interest. To address errors in the DFT model calculation, we conducted a detailed study of the external errors associated with the accuracy and reproducibility of the first-principles theoretical models that are used to predict isotopic signatures of equilibrium and disequilibrium chemistry. The carbon isotope calculations have not been experimentally validated for methane, ethane, propane, butane and pentane. Therefore, we assigned a range in model predictions that we regard as typical based on the variances among models and differences between models and experimental constraints from several recent studies of relevant systems. First, we evaluated the range in predictions of the site specific carbon isotope fractionation of propane at 400K and found that three different model predictions ranged by 0.9‰ at 400K (Webb and Miller, 2014; Piasecki et al., 2016; this study). We also note that Xie et al 2018 (7), experimentally tested the accuracy of these three models, demonstrating that their variance is generally similar to their average accuracy. Second, a recent study evaluated the range in predictions of the bulk carbon isotope fractionation between CO2 and methane found that the models fell within 2‰ of the measurements from 300-1200°C (8). Finally, a recent study (9), has evaluated the factors that affect accuracy of calculated equilibrium isotope fractionations for carbon and hydrogen isotopes among a diverse range of organic compounds. The authors found that the deviation of the calculated isotopic value from the measured equilibrium value for B3LYP DFTs (which is also the DFT that we use in our study) averaged 4‰ for carbon isotopes and 30‰ for hydrogen isotopes at 25°C. However, Iron and Gropp (2019) modeled a set of organic molecules that include a variety of complex functional groups including –SH, -OH and hydrocarbon rings. Therefore, their analysis over- estimates the inaccuracies that would arise from modeling fractionations among more closely similar species like the small . For these reasons, we assign representative DFT model errors of 2‰ for carbon isotopes and 30‰ for hydrogen isotopes, and evaluate the extent of overlap of data with these ranges, also given measurement errors of 0.5‰ for carbon isotopes and 13‰ for hydrogen isotopes (10).

Isotope Nomenclature We refer to the fractionation between two species using epsilon notation where ε is calculated in the following manner:

1000 + 4 1 = 1 1000 1000 + 13 13 𝛿𝛿 𝐶𝐶𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏 𝜀𝜀 𝐶𝐶 𝐶𝐶 − 𝐶𝐶 � 13 − � ∗ and 𝛿𝛿 𝐶𝐶𝑚𝑚𝑚𝑚𝑚𝑚ℎ𝑎𝑎𝑎𝑎𝑎𝑎 1000 + 4 1 = 1 1000 1000 + 𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿 𝜀𝜀𝜀𝜀 𝐶𝐶 − 𝐶𝐶 � − � ∗ 𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿ℎ𝑎𝑎𝑎𝑎𝑎𝑎 3

Gas Concentration and Isotope Fractionation during Sampling Acquisition of a natural gas sample representative of the concentrations of gases in the reservoir is difficult to obtain. For a well to produce, a pressure gradient must exist from the surface to the bottom of the well. As the pressure decreases in the reservoir, the upper boundary of the phase loop can be breached. This condition allows for different flows of gas and liquid which are not consistent with the gas to oil ratio in the reservoir itself. Additionally, as the pressure in the reservoir changes, gas in the solution may exsolve, further changing the gas composition. These complications can be addressed by sealing the well and sampling with a PVT device. However, this approach is cost prohibitive. We therefore focus in this study on the isotopic compositions of the gases rather than their concentrations. Previous experiments have shown that the vapor pressure isotope effect (VPIE) for hydrocarbons at reservoir conditions is minor as discussed in the following section.

Vapor Pressure Isotope Effects for Light Hydrocarbons

During the industrial extraction of a petroleum reservoir, isotope fractionation due to degassing could potentially alter the final isotopic composition of natural gas samples. The isotopic fractionation between the gas phase and condensed phase is known as the vapor pressure isotope effects (VPIE) (11, 12). In 1967, the existing VPIE data for methane from 91-112oK was compiled and corrected for molar volume and gas nonideality. If the calculated equation is extrapolated to 0-400oC, the δ13C and δD for methane VPIEs are -0.47 to -0.46‰ and -7.1 to-3.8‰ respectively. However, extrapolating VPIE to higher temperatures is generally questionable because the slope of VPIE vs temperature is known to be variable (12). Therefore extrapolating, especially for hundreds of degrees can lead to erroneous values. Previous work (13) has presented an experimental test to examine the carbon isotope fractionation related to petroleum degassing. A confined petroleum container was depressurized from 89 atm to 1atm and the headspace gas was sequentially sampled and analyzed for carbon isotopes. There was no systematic change (<1‰) in carbon isotope composition for C1-C4 species, implying the VPIE of carbon isotopes has an insignificant impact on the isotope ratios of the resulting gas. We are not aware of any measured or calculated VPIEs for single deuterium substitution in C2-C5 alkanes in our interested temperature range. Future work will have to determine the VPIE of hydrogen isotopes for these alkanes.

Homolytic and β-Scission Models We determine isotope fractionation of thermogenic natural gas formation from scission of a long chain hydrocarbon compounds into C1-C5 alkanes using previously determined kinetic parameters for homolytic scission of carbon and hydrogen (14) (SI Dataset 4 and 5). We used these data to calculate six different isotope fractionations related to cracking, an instantaneous, cumulative and Rayleigh fractionation factor for homolytic scission and an instantaneous, cumulative and Rayleigh fractionation factor for β-scission. For both types of cracking, the instantaneous fractionation factor at any given

4

temperature assumes that all the gas is generated at that temperature. The cumulative and Raleigh fractionation factors are both calculated using a temperature ramp of 0-400oC with a constant heating of 1oC/Mya and assumes that the gas at any given temperature is a cumulative mixture of all the gas that was generated from the beginning of the temperature ramp to that temperature. However the cumulative model assumes that the precursor does not change in isotopic composition as the reaction progresses while the Rayleigh fractionation model assumes that the precursor is depleted as the products are being formed.

Primary and Secondary Cracking Model We developed a multi-component gas generation kinetic model to simulate isotope fractionation of thermogenic natural gas generation from the cracking of oil compounds. In this model, oil (n-alkanes) breaks down into C1-C4 alkanes at temperatures between 50-200oC with a constant heating rate of 2oC/Mya. We address two types of chemical reactions in this system: oil cracking into each gas component, and the secondary cracking of C2-C4 species. We assume that both of these reactions happen through a simple homolytic C-C radical cleavage mechanism and that in secondary cracking reactions, the C2-C4 will only decompose to methane. We programmed the reactions using temperature-dependent kinetic parameters (frequency factor, activation energy and variance of activation energy) for both primary and secondary cracking reactions (15) and temperature-dependent kinetic isotope effects (KIE) for modeled position-specific cleavage KIE of n-octane (16). We set up the differential equations array that includes the mass/carbon transfer of every mentioned above and their associated carbon isotopic fractionations and solve using ode15s in MATLAB. We use this model to investigate the increasing magnitude of fractionation between gas species with increasing temperature (SI Figure 8). We find that as the temperature increases, more of the precursor molecule is destroyed and there is an increase in secondary cracking which increases the isotope fractionation between the gases (see SI Figure 11). This increase in secondary cracking could explain the increase in fractionation between gas species in the cracking experiments and would not necessarily discount the cracking experiments as an analog for natural gas formation.

Radical Reaction Model We built a model with 79 free radical reaction schemes and 24 species including the 12C and 13C version of butane, butene, propane, propene, ethane, ethane, methane, as well as hydrogen and hydrogen . A singly 13C-substituted version is created for each carbon-containing species and the kinetic isotope effect (KIE) is assumed to be 1 for all reactions. This is a nonchemical labelling study and because KIEs are ignored, this exercise is just to show that C-C bonds are broken and formed sufficiently to establish near equilibrium at the temperature of the simulation. We chose cleavage, H-transfer and beta elimination reactions (and their reverse reactions), and acquired the kinetic parameters for these reactions from literature (17–19). The temperature of the system is set at 500K and we solve this differential equation set of the species using a numeric

5

method (ode 15s from MATLAB, at a relative tolerance of 1e-9). We show two tests using this model to show that a 13C label can recirculate among the different species through our programmed radical reactions. The first simulation has a 10% of 13C enrichment in butane while the second simulation has a 10% of 13C enrichment in methane (see SI Figure 10). In both simulations (and in every other simulation where a 13C spike was added, the 24 species reached the same steady state value, suggesting the system had come to equilibrium.

Reactions included in the Radical Reaction Model: Equation 1 = [C2H6]→2[CH3]* Equation 2 = [C3H8]→[C2H5]*+CH3* Equation 3 = [C4H10]→2*[C2H5]* Equation 4 = [C2H6]+[CH3]*→[C2H5]*+[CH4] Equation 5 = [C3H8]+[CH3]*→[C3H7]*+[CH4] Equation 6 = [C4H10]+[CH3]*→[C4H9]*+ [CH4] Equation 7 = [C3H8]+[C2H5]*→[C3H7]*+ [C2H6] Equation 8 = [C4H10]+[C2H5]*→[C4H9]*+ [C2H6] Equation 9 = [C4H10]+[C3H7]*→[C4H9]*+[C3H8] Equation 10 = [C2H5]*+[H]*→[C2H6] Equation 11 = [C3H7]*+[H]*→[C3H8] Equation 12 = [C4H9]*+[H]*→[C4H10] Equation 13 = [CH3]*+[CH3]*→[C2H6] Equation 14 = [C2H5]*+[CH3]*→[C3H8] Equation 15 = [C3H7]*+[CH3]*→[C4H10] Equation 16 =[CH4] → [CH3]*+[H]* 13 12 13 Equation 17 =[ C CH6] → [CH3]*+[ CH3]* 13 12 13 12 Equation 18 =[ C C2H8] → [CH3]*+[ C CH5]* 13 12 13 Equation 19 =[ C C2H8] → [C H3]*+[C2H5]* 13 12 13 12 Equation 20 =[ C C3H10] → [C2H5]*+[ C CH5]* 13 12 13 12 Equation 21 =[ C CH6]+[CH3]* → [ C CH5]*+[CH4] 13 13 Equation 22 =[C2H6]+[ CH3]* → [C2H5]*+[C H4] 13 12 13 12 Equation 23 =[ C C2H8]+[CH3]* → [ C C2H7]*+[CH4] 13 13 Equation 24 =[C3H8]+[ CH3]* → [C3H7]*+[C H4] 13 12 13 12 Equation 25 = [ C C3H10]+[CH3]* → [ C C3H9]*+[CH4] 13 13 Equation 26 = [C4H10]+[ CH3]* → [C4H9]*+[C H4] 13 12 13 12 Equation 27 = [ C C2H8]+[C2H5]* → [ C C2H7]*+[C2H6] 13 12 13 12 Equation 28 = [C3H8]+[ C CH5]* → [C3H7]*+[ C CH6] 13 12 13 12 Equation 29 =[ C C3H10]+[C2H5]* → [ C C3H9]*+[C2H6] 13 12 13 12 Equation 30 = [C4H10]+[ C CH5]* → [C4H9]*+[ C CH6] 13 12 13 12 Equation 31 = [ C C3H10]+[C3H7]*→ [ C C3H9]*+[C3H8] 13 12 13 12 Equation 32 = [C4H10]+[ C C2H7]* → [C4H9]*+[ C C2H8] 13 12 13 12 Equation 33 =[ C CH5]*+[H]* → [ C CH6] 13 12 13 12 Equation 34 =[ C C2H7]*+[H]* → [ C C2H8] 13 12 13 12 Equation 35 =[ C C3H9]*+[H]* → [ C C3H10] 13 13 12 Equation 36 =[CH3]*+[ CH3]* → [ C CH6] 13 12 13 12 Equation 37 =[ C CH5]*+[CH3]* → [ C C2H8]

6

13 13 12 Equation 38 =[C2H5]*+[ CH3]* → [ C C2H8] 13 12 13 12 Equation 39 =[ C C2H7]*+[CH3]* → [ C C3H10] 13 13 12 Equation 40 =[C3H7]*+[ CH3]* → [ C C3H10] 13 13 Equation 41 =[ CH4] → [ CH4]*+[H]* Equation 42 =[C4H10] → [CH3]*+[C3H7] 13 12 13 12 Equation 43 =[ C C3H10] → [CH3]*+[ C C2H7]* 13 12 13 Equation 44 =[ C C3H10] → [ CH3]*+[C3H7]* Equation 45 =[H]*+[H]* → [H2] Equation 46 =[H2] → [H]*+[H]* Equation 47 =[C3H8]+[H]* → [C3H7]*+[H2] 13 12 13 12 Equation 48 =[ C C2H8]+[H]* → [ C C2H7]*+[H2] Equation 49 =[C4H10]+[H]* → [C4H9]*+[H2] 13 12 13 12 Equation 50 =[ C C3H10]+[H]* → [ C C3H9]*+[H2] Equation 51 =[CH3]*+[H]* → [CH4] 13 13 Equation 52 =[ CH3]*+[H]* → [ CH4] Equation 53 =[C2H6]+[H]* → [C2H5]*+[H2] 13 12 13 12 Equation 54 =[ C CH6]+[H]* → [ C CH5]*+[H2] Equation 55 =[C2H5]* → [C2H4]+[H]* 13 12 13 12 Equation 56 =[ C CH5]* → [ C CH4]+[H]* Equation 57 =[C3H7]*→ [C3H6]+[H]* 13 12 13 12 Equation 58 =[ C C2H7]* → [ C C2H6]+[H]* Equation 59 = [C4H9]* → [C4H8]+[H]* 13 12 13 12 Equation 60 = [ C C3H9]* →[ C C3H6]+[H]* Equation 61 =[C4H9]* → [C3H6]+[CH3]* 13 12 13 12 Equation 62 =[ C C3H9]* → [ C C2H6]+[CH3]* 13 12 13 Equation 63 =[ C C3H9]* → [C3H6]+[ CH3]* Equation 64 =[C3H6]+[CH3]* → [C4H9]* 13 12 13 12 Equation 65 =[ C C2H6]+[CH3]* → [ C C3H9]* 13 13 12 Equation 66 =[C3H6]+[ CH3]* → [ C C3H9]* Equation 67 =[C2H4]+[H2] → [C2H6] 13 12 13 12 Equation 68 =[ C CH4]+[H2] → [ C CH6] Equation 69 = [C4H8]+[H]* → [C4H9]* 13 12 13 12 Equation 70 = [ C C3H8]+[H]* → [ C C3H9]* Equation 71 =[C2H4]+[CH3]* → [C3H7]* 13 12 13 12 Equation 72 =[ C CH4]+[CH3]* → [ C C2H7]* 13 13 12 Equation 73 =[C2H4]+[ CH3]* → [ C C2H7]* Equation 74 =[C2H4]+[H]* → [C2H5]* 13 12 13 12 Equation 75 =[ C CH4]+[H]* → [ C CH5]* Equation 76 =[C3H6]+[H]* → [C3H7]* 13 12 13 12 Equation 77 =[ C C2H6]+[H]* → [ C C2H7]* Equation 78 =[CH4]+[H]* → [CH3]*+[H2] 13 13 Equation 79 =[ CH4]+[H]* → [ CH3]*+[H2]

Implications for intramolecular carbon isotope distribution in ethane and propane

7

Previous measurements have been made on clumped 13C-13C isotopologues of ethane (20) and site-specific 13C of propane (21–23). The maximum range resulting from internal isotopic equilibrium of 13C-13C clumping in ethane is 0.5‰ between 0 to 1000oC. However, the full range of observed variation in 13C-13C clumping in ethane is 4.8‰ implying that 13C-13C clumping in ethane does not reflect equilibrium processes. Clog et al, 2018 finds evidence that values of 13C-13C clumping in ethane decreases with ethane abundance and increasing δ13C values. This correspondence suggests that the 13C-13C clumping in ethane is controlled by kinetic effects associated with the production and destruction of ethane, both in experiments and natural gas basins. This finding is in contrast to our radical reaction network model predicting that small n-alkanes should approach equilibrium, especially high maturity fields, such as Haynesville, and Marcellus. A similar discrepancy is also seen in site-specific 13C in propane (21). There are several possible explanations for these discrepancies: 1) None of the ethane and propane measurements are anchored to a stochastic or thermodynamic reference frame, due to the thermodynamic properties of ethane and propane. Therefore, it is difficult to know which if any samples are in equilibrium. 2) Our radical reaction network model only includes singly substituted species. It is possible that clumped isotopologue species take a substantially longer time to reach internal equilibrium and the measurements reflect that longer timescale, or achieve steady states that are so disturbed by mixing effects that they depart noticeably from equilibrium. 3) The clumped isotopologue measurements were made in very mature basins, where there is extensive secondary cracking. Secondary cracking could be another process similar to biodegradation which drives clumped isotopes away from equilibrium values. (See (24) for further discussion) 4) Both the ethane and propane measurements were made using older mass spectrometers, and they were made with substantial effort to correct for contaminants + (e.g. CH3OH for ethane clumped isotope measurement). We expect that newer high- resolution mass spectrometers (e.g. (25, 26)) should have sufficiently high mass- resolving power to avoid isobaric interferences. It is possible that with more improved technologies that are available, the 13C-13C clumping values measured might change. Future work could address these issues.

SI Dataset 1. Methane stable and clumped isotope data for natural gas basins in this study.

SI Dataset 2. Stable isotope data for C2+ species in natural gas basins in this study.

SI Dataset 3. Carbon and Hydrogen Fractionation Factors for Thermodynamic Equilibrium

SI Dataset 4. Carbon and Hydrogen Fractionation Factors for Homolytic C-C Cleavage

SI Dataset 5. Carbon Fractionation Factors for Beta Scission

8

Fig. S1. (A) ε13C (i-butane-ethane) vs ε13C (propane-ethane) and (B) ε13C (i-pentane- 13 propane) vs ε C (butane-ethane) for carbon isotopes of gases in the study. 76% of the gases in panel A and 88% of the gases in panel B are consistent with thermodynamic equilibrium (red solid line) considering model and measurement uncertainties. A portion 13 of the gases are shaded and have ε C (C3-C2) values greater than 5. These gases are 13 associated with biodegradation and have low δ C-CH4, δD-CH4 and Methane 13 13 temperatures, tend to have high δ C-CO2 values and approach the high ε C (propane- ethane) values seen in the biodegraded gases of Antrim Shale which range from 10-25‰.

9

Fig. S2. (A) εD (butane-ethane) vs εD (propane-ethane) and (B) εD (pentane-propane) vs εD (butane-ethane) for gases considered in this study. 72% of gases in panel A and 77% of gases in panel B are consistent with thermodynamic equilibrium (red solid line) rather than the predictions for homolytic cleavage (yellow and blue lines). The equilibration of hydrogen isotopes in n-alkanes has been seen previously (27). The shaded gases are gases associated with biodegradation.

10

Fig. S3. ε13C (butane-ethane) vs ε13C (propane-ethane) of a global compilation (28) of coal (n=73) (A), conventional (n=973) (B) and shale gases (n=48) (C). A similar pattern to our dataset (Figure 2) is seen in the global compilation. 59% of the coal gases, 84% of the conventional gases and 100% of the shale gases are consistent with being in 13 thermodynamic equilibrium. Mature thermogenic gases (high δ C of C2) tend to be close to the equilibrium lines, while immature thermogenic gases tend to be farther away from equilibrium values.

11

Fig. S4. (A) εD (i-butane-ethane) vs εD (propane-ethane) and (B) εD (i-pentane-propane) vs εD (i-butane-ethane) for hydrogen isotopes of gases in the study. Only 34% of the gases in panel A and 22% of the gases in panel B are consistent with thermodynamic equilibrium (red solid line). The shaded gases are gases associated with biodegradation 13 13 and have low δ C-CH4, δD-CH4 and Methane temperatures and high ε C (propane- ethane) values. Measurement errors are 13‰ 1σ.

12

13 Fig. S5. ε C (CO2-ethane) of carbon isotopes vs Methane temperature for gases in the study subdivided by basins. None of the gases in the study fall on the prediction for thermodynamic equilibrium fractionation (red solid line). The measurement errors for carbon isotopes of CO2 and ethane are 0.5‰ and 1σ error bars are plotted for methane temperatures.

13

Fig. S6. εD (propane-methane) vs Methane clumped isotope apparent temperature for gases considered in this study. 88% of the gases are consistent with predictions for thermodynamic equilibrium fractionation (red solid lines) rather than predictions for homolytic cleavage (yellow and blue lines). The observation that hydrogen isotopes of n- alkanes are in equilibrium has been seen before (27). The shaded gases are gases associated with biodegradation. The measurement errors for hydrogen isotopes of propane and methane are 13‰ and 1σ error bars are plotted for methane temperatures.

14

\

Fig. S7. (A) ε13C (butane-ethane) vs ε13C (propane-ethane) and (B) ε13C (butane-ethane) vs Temperature for the catalytic metathetic experiments performed by Mango et al 1999 (29). Most of the recovered gases tend to reflect thermodynamic equilibrium values.

15

Fig. S8. (A) ε13C (butane-ethane) vs ε13C (propane-ethane) and (B) ε13C (propane-ethane) vs Temperature for the cracking experiments (30–33). The recovered gases below 400°C tend to reflect thermodynamic equilibrium values in C2+ isotopic compositions, however closer inspection of these experiments reveal that the isotopic fractionation increases with temperature (400-700°C). This feature of increasing ε13C(propane-ethane) with increasing temperature is not seen in the natural gas data (SI Figure 9) or in the thermodynamic fractionation factor or various cracking models.

16

Fig. S9. ε13C (butane-ethane) vs Temperature for the natural gas data. 88% of the gases are consistent with thermodynamic equilibrium (red solid line). The biodegraded gases are shaded. The measurement errors for carbon isotopes of butane and ethane are 0.5‰ and 1σ error bars are plotted for methane temperatures.

17

Fig. S10. Results of two different simulations in the Radical Reaction Model. The blue line shows the result of a 10% spike of 13C in butane, while the orange line shows the result of a 10% spike of 13C in methane. Both simulations reach the same steady state value, implying the system has reached an equilibrium value.

18

Fig. S11. (A) The fraction of each component over the modelled reaction and (B) The fractionation between butane and methane, ethane and propane over the course of the model. As the model run increases to higher temperatures and the oil is depleted, secondary cracking of butane (and propane and ethane) begins and the fractionation between butane and the other n-alkanes begins to increase. This effect is similar to what is seen in the cracking gas experiments, but not in the natural gas data.

19

References 1. D. A. Stolper, et al., Combined 13C–D and D–D clumping in methane: Methods and preliminary results. Geochimica et Cosmochimica Acta 126, 169–191 (2014).

2. J. M. Eiler, et al., A high-resolution gas-source isotope ratio mass spectrometer. International Journal of 335, 45–56 (2013).

3. H. C. Urey, The thermodynamic properties of isotopic substances. J. Chem. Soc. 1947, 562–581 (1947).

4. T. H. Dunning, Gaussian basis sets for use in correlated molecular calculations. I. The boron through neon and hydrogen. J. Chem. Phys. 90, 1007–1023 (1989).

5. M. A. Webb, T. F. Miller, Position-Specific and Clumped Stable Isotope Studies: Comparison of the Urey and Path-Integral Approaches for Carbon Dioxide, Nitrous Oxide, Methane, and Propane. J. Phys. Chem. A 118, 467–474 (2014).

6. A. Piasecki, A. Sessions, B. Peterson, J. Eiler, Prediction of equilibrium distributions of isotopologues for methane, ethane and propane using density functional theory. Geochimica et Cosmochimica Acta 190, 1–12 (2016).

7. H. Xie, et al., Position-specific hydrogen isotope equilibrium in propane. Geochimica et Cosmochimica Acta 238, 193–207 (2018).

8. N. Kueter, M. W. Schmidt, M. D. Lilley, S. M. Bernasconi, Experimental determination of equilibrium CH4–CO2–CO carbon isotope fractionation factors (300–1200°C). Earth and Planetary Science Letters 506, 64–75 (2019).

9. M. A. Iron, J. Gropp, Cost-effective density functional theory (DFT) calculations of equilibrium isotopic fractionation in large organic molecules. Phys. Chem. Chem. Phys. 21, 17555–17570 (2019).

10. T. Umezawa, et al., Interlaboratory comparison of δ13C and δD measurements of atmospheric CH4 for combined use of data sets from different laboratories. Atmos. Meas. Tech. 11, 1207–1231 (2018).

11. J. Bigeleisen, Vapor of isotopic molecules*. J. Chem. Phys. 60, 35–43 (1963).

12. G. Jancso, W. A. Van Hook, Condensed phase isotope effects. Chem. Rev. 74, 689– 750 (1974).

13. E. M. Galimov, Carbon isotope fractionation as a function of temperature in the CH4-C2H6-C3-H8-C4H10 system. International (1972).

20

14. Y. Ni, et al., Fundamental studies on kinetic isotope effect (KIE) of hydrogen isotope fractionation in natural gas systems. Geochimica et Cosmochimica Acta 75, 2696–2707 (5).

15. H. Tian, X. Xiao, R. Wilkins, Y. Tang, New insights into the volume and pressure changes during the thermal cracking of oil to gas in reservoirs: Implications for the in-situ accumulation of gas cracked from oils. AAPG Bulletin 92 (2008).

16. Y. Ni, et al., Fundamental studies on kinetic isotope effect (KIE) of hydrogen isotope fractionation in natural gas systems. Geochimica et Cosmochimica Acta 75, 2696–2707 (5).

17. K. M. Sundaram, G. F. Froment, Modeling of Thermal Cracking Kinetics. 3. Radical Mechanisms for the Pyrolysis of Simple Paraffins, Olefins, and Their Mixtures. Ind. Eng. Chem. Fund. 17, 174–182 (1978).

18. J. R. Fincke, et al., Plasma Thermal Conversion of Methane to Acetylene. Plasma Chemistry and Plasma Processing 22, 105–136 (2002).

19. H. Wang, M. Frenklach, A detailed kinetic modeling study of aromatics formation in laminar premixed acetylene and ethylene flames. and Flame 110, 173–221 (1997).

20. M. Clog, et al., A reconnaissance study of 13C–13C clumping in ethane from natural gas. Geochimica et Cosmochimica Acta 223, 229–244 (2018).

21. A. Piasecki, et al., Position-specific 13C distributions within propane from experiments and natural gas samples. Geochimica et Cosmochimica Acta 220, 110– 124 (2018).

22. A. Gilbert, K. Yamada, K. Suda, Y. Ueno, N. Yoshida, Measurement of position- specific 13C isotopic composition of propane at the nanomole level. Geochimica et Cosmochimica Acta 177, 205–216 (2016).

23. L. Gao, et al., Determination of position-specific carbon isotope ratios in propane from hydrocarbon gas mixtures. Chemical Geology 435, 1–9 (2016).

24. B. K. Peterson, M. J. Formolo, M. Lawson, Molecular and detailed isotopic structures of petroleum: Kinetic Monte Carlo analysis of cracking. Geochimica et Cosmochimica Acta 243, 169–185 (2018).

25. E. D. Young, et al., The relative abundances of resolved l2CH2D2 and 13CH3D and mechanisms controlling isotopic bond ordering in abiotic and biotic methane gases. Geochimica et Cosmochimica Acta 203, 235–264 (2017).

26. J. M. Eiler, et al., The isotopic structures of geological organic compounds. Geological Society, London, Special Publications 468, 53 (2018).

21

27. E. P. Reeves, J. S. Seewald, S. P. Sylva, Hydrogen isotope exchange between n- alkanes and water under hydrothermal conditions. Geochimica et Cosmochimica Acta 77, 582–599 (2012).

28. O. Sherwood, S. Scwietzke, V. Arling, G. Etiope, Global Inventory of Gas Geochemistry Data from Fossil Fuel, Microbial and Burning Sources. Earth System Science Data 9 (2017).

29. F. D. Mango, L. W. Elrod, The carbon isotopic composition of catalytic gas: A comparative analysis with natural gas. Geochimica et Cosmochimica Acta 63 (1999).

30. F. Lorant, A. Prinzhofer, F. Behar, A.-Y. Huc, Carbon isotopic and molecular constraints on the formation and the expulsion of thermogenic hydrocarbon gases. Chemical Geology 147, 249–264 (1998).

31. C. Pan, L. Jiang, J. Liu, S. Zhang, G. Zhu, The effects of calcite and montmorillonite on oil cracking in confined pyrolysis experiments. Organic Geochemistry 41, 611–626 (2010).

32. B. Andresen, T. Throndsen, A. Råheim, J. Bolstad, A comparison of pyrolysis products with models for natural gas generation. Chemical Geology 126, 261–280 (1995).

33. J. Du, Z. Jin, H. Xie, L. Bai, W. Liu, Stable carbon isotope compositions of gaseous hydrocarbons produced from high pressure and high temperature pyrolysis of lignite. Organic Geochemistry 34, 97–104 (2003).

22