Abstract
The observed rise in atmospheric methane (CH4) from 375 ppbv during the Last Glacial Maximum (LGM: 21,000 years ago) to 680 ppbv during the late preindustrial era is not well understood. Atmospheric chemistry considerations implicate an increase in CH4 sources, but process‐based estimates fail to reproduce the required amplitude. CH4 stable isotopes provide complementary information that can help constrain the underlying causes of the increase. We combine Earth System model simulations of the late preindustrial and LGM CH4 cycles, including process‐based estimates of the isotopic discrimination of vegetation, in a box model of atmospheric CH4 and its isotopes. Using a Bayesian approach, we show how model‐based constraints and ice core observations may be combined in a consistent probabilistic framework. The resultant posterior distributions point to a strong reduction in wetland and other biogenic CH4 emissions during the LGM, with a modest increase in the geological source, or potentially natural or anthropogenic fires, accounting for the observed enrichment of δ 13CH4.
Keywords: wetlands, Last Glacial Maximum, methane, 13CH4, greenhouse gas, isotopic discrimination
Key Points
For the first time we combine process‐based models of the CH4 cycle as priors, with ice core observations in a Bayesian framework
For the glacial we infer unchanged or increased geological sources, but a halved wetland flux, which is not captured by process modeling
This suggests missing processes in methane models, with potentially important consequences for confidence in future methane projections
1. Introduction
Atmospheric methane (CH4) is an important greenhouse gas. Its concentration has risen sharply over the past two centuries, reaching 1,799 ppbv by Common Era (CE) 2010 (Kirschke et al., 2013). This is estimated to have contributed around 25% of the anthropogenic greenhouse gas effect since CE 1750 (Myhre et al., 2013). CH4 is also reactive with a lifetime of about 9 years (Prather et al., 2012), and so its abundance indirectly affects the concentrations of other trace gases such as nitrous oxide and ozone.
Observed variations in the growth rate over the past decades are not well understood (e.g., Bousquet et al., 2011; Kirschke et al., 2013). Different studies have implicated changes in biogenic emissions (e.g., Nisbet et al., 2016; Schaefer et al., 2016), lifetime (Rigby et al., 2017) or biomass burning emissions (Worden et al., 2018) in the recent growth rate. Prior to the observational era, changes in atmospheric methane can be accurately reconstructed from gas bubbles recovered in ice cores. During the late Quaternary, CH4 mixing ratios are correlated with climate (e.g., Loulergue et al., 2008). Explaining the magnitude of these variations remains a challenge (Hopcroft et al., 2017; Levine, Wolff, Jones, Sime, Valdes, et al., 2011; Murray et al., 2014), reflecting incomplete understanding of the source and sink processes.
The increase in atmospheric CH4 concentration from the LGM to the beginning of the Industrial Revolution (hereafter the late preindustrial) is among the largest such changes, with an increase from 375 ppbv to 680 ppbv (Loulergue et al., 2008; Mitchell et al., 2013; WAIS Divide Project Members, 2015). This concentration increase is likely the result of a near doubling of the source of methane (Levine, Wolff, Jones, Sime, Valdes, et al., 2011; Murray et al., 2014). Considerable debate remains over how this occurred (Kaplan et al., 2006; Levine, Wolff, Jones, Sime, Valdes, et al., 2011; Murray et al., 2014; Valdes et al., 2005), because of uncertainties in the response of methane sources and sinks to climate, and in terms of the relative importance of different sources in the late preindustrial era (Hopcroft et al., 2017).
Concurrent changes in the stable isotopes of methane can provide additional constraints, because different sources and sinks of methane have distinct isotopic signatures and fractionation rates (e.g., Bock et al., 2017). Between the LGM and preindustrial, δ 13CH4 became depleted from −43 ± 0.3‰ to −48 ± 0.3‰, with a similar depletion of deuterium of CH4 (δDCH4) from −79 ± 4‰ to −98 ± 3‰ (Fischer et al., 2008; Möller et al., 2013; Sowers, 2006, 2010).
Fischer et al. (2008) used a box model of atmospheric CH4 and CH4 isotopes and adjusted the relative contributions of sources and sinks to explain the observed glacial‐interglacial changes. They inferred an increase in wildfire CH4 emissions (+18%) at the LGM and a reduction in wetland emissions (−41%) and CH4 lifetime (−32%). The latter is in conflict with most model‐based estimates.
More recently, Möller et al. (2013) extended ice core records to 160,000 years Before Present and found that the δ 13CH4 is positively correlated with the atmospheric concentration of CO2 and not CH4. This suggests that the observed variations of δ 13CH4 likely arise through the effect of CO2 and climatic changes on the isotopic fractionation rates of CH4 sources or sinks (Whiticar & Schaefer, 2007), rather than only by changes in the relative importance of the different source types, as previously assumed (Fischer et al., 2008). Möller et al. (2013) suggest that shifts in the isotopic signature of tropical wetland CH4 are the main contributor.
Here we use process‐based Earth System model simulations of the late preindustrial and LGM methane cycles (Hopcroft et al., 2017, H17 hereafter), to better understand the change in CH4. We calculate the implied isotopic composition of methane, including the environmentally driven change in the isotopic discrimination of vegetation and hence the isotopic signature of wetland and biomass burning emissions. We combine these results in a Bayesian framework, which allows us to incorporate uncertainty estimates and include prior information.
2. Incorporating Process‐Based Estimates of Changes in Methane Sources and Lifetime Into a Model of Atmospheric Methane Stable Isotopes
We use the CH4 sources as incorporated in the Earth System model simulations of the late preindustrial and LGM with HadGEM2‐ES (Collins et al., 2011; HadGEM2 Development Team, 2011, and see supporting information), configured with glacial boundary conditions as described previously (Hopcroft & Valdes, 2015; Singarayer & Valdes, 2010; H17). HadGEM2‐ES is a widely used, coupled Earth System model (e.g., Booth et al., 2012; Caesar et al., 2013; Hopcroft & Valdes, 2015; Jones et al., 2011; Kandlbauer et al., 2013). It includes wetlands (Gedney et al., 2004; Marthews et al., 2015) and tropospheric chemistry (O'Connor et al., 2014). Separate models of peatlands (Wania et al., 2010), biomass burning (Kaplan et al., 2016; Pfeiffer et al., 2013), oceans, and termites (Kaplan et al., 2003; Sanderson, 1996) were also used, see the supporting information.
Using the standard wetland scheme, the high‐latitude emissions are likely underestimated, as shown by a comparison of CH4 concentrations for the present day (Hayman et al., 2014). Including the peatland flux leads to a stronger high‐latitude source, as well as a larger reduction in total wetland/peatland emissions at the LGM of 42% versus 30% without.
H17 presented three alternative scenarios of LGM fire emissions: standard‐fire (simulated with a process‐based dynamic vegetation model), standard+LGM humans, which additionally incorporates an empirically based estimate of hunter‐gatherer fire activities during the LGM (Kaplan et al., 2016), and low fire, in which LGM fire emissions are arbitrarily set to 10% of late preindustrial values. Here preindustrial fire emissions are scaled based on ice core evidence and other modeling studies (Ferretti et al., 2005; Thonicke et al., 2005), as described in the supporting information.
We introduce a geological source which comprises mud volcanoes, marine seeps, microseepage, and geothermal methane, which are commonly missing from methane inventories (e.g., Etiope et al., 2008). We reduce the ocean source to 1 TgCH4/year in light of recent observations (Kirschke et al., 2013) and set both the hydrate and geological terms to 10 TgCH4/year (Petrenko et al., 2017; H17). We incorporate OH, soil, and stratospheric sink terms as modeled by H17. We also here include a boundary layer atomic chlorine (Cl) sink, which we set to 3% of the total CH4 sink (Allan et al., 2007; Platt et al., 2004). The assumed isotopic signature of each source and the fractionation factors for the four methane sinks are listed in Table 1, along with estimated prior and posterior values.
Table 1.
LGM | |||||||
---|---|---|---|---|---|---|---|
Prior | Posterior | ||||||
Sources (TgCH4/year) | Late Preindustrial | Mean | ±1 s.d. | ΔLGM (%) | Mean | ±1 s.d. | ΔLGM (%) |
N extratropical wetland | 55 | 18 | 4.7 | −67 | 12.0 | 4.0 | −78 |
Tropical wetland | 73 | 55 | 14.3 | −25 | 32.0 | 4.7 | −56 |
S extratropical wetlanda | 11 | 6.2 | — | −44 | — | — | |
Biomass burning | 21 | 13.5 | 4.8 | −36 | 16.6 | 2.0 | −21 |
Termites | 20 | 12 | 3.1 | −40 | 10.8 | 3.0 | −46 |
Hydrates | 10 | 10 | 3.6 | 0 | 6.9 | 2.2 | −31 |
Other geological | 10 | 10 | 4.6 | 0 | 10.5 | 2.5 | 5 |
Oceansa | 1 | 0.8 | − | −20 | − | − | − |
Sum | 201 | 126 | 34 | −38 | 96 | 18 | −52 |
Total lifetime (yr) | 9.7 | 10.3b | 0.2 | 6 | 10.4c | 0.1 | 7 |
Note. The LGM means and standard deviations are derived from the PDFs shown in the blue curves in Figure 3. The individual prior sink terms are given in Table S4 in the supporting information. Refer to Tables S2 and S3 for isotopic signatures of individual sources and sinks.
Indicates sources not varied as part of the Bayesian algorithm.
Calculated using the prior mean LGM fire emissions.
Calculated using the posterior mean LGM fire emissions. LGM = Last Glacial Maximum.
We combine these estimates using a three‐box model of atmospheric CH4 and stable isotopes as described in the supporting information (Lassey et al., 2000; Miller, 2005). The results show a preindustrial CH4 concentration, δ 13C and δD of 683 ppb, −49.0‰ and −98.4‰ respectively. These are in agreement with ice core measurements, as compared in Figure 1. Using LGM fluxes (with standard‐fire) results in LGM CH4, δ 13C and δD(CH4) values of 447 ppbv, −48.6‰, and −93.7‰, respectively, also shown in Figure 1. For both time periods, the box model values are extracted at either the northern or southern box depending on the relevant ice core location, as listed in Table S1. This LGM simulation underestimates both the reduced LGM CH4 concentration and shift in isotopes. The low‐fire scenario in which all fire emissions are set to 10% of the late preindustrial results in a better prediction of the LGM CH4 change, but at the expense of the δ 13CH4 and δD(CH4), which both shift to more isotopically depleted values. The standard‐fire+LGM humans has the opposite effect but is closer to the standard‐fire simulation. As a result, to simplify the presentation we only use the standard‐fire scenario (filled circles in Figure 1) in the following.
We evaluated the potential influence of changes in the Cl sink following the approach of Levine, Wolff, Jones, and Sime (2011). Using monthly fields from late preindustrial and LGM simulations with HadGEM2‐ES (Hopcroft & Valdes, 2015; H17), this results in a small enrichment of δ 13C at the LGM of 0.12‰, shown in Figure 1.
3. Changes in the Isotopic Discrimination by Vegetation and Potential Influence on CH4 Emissions
During the LGM the δ 13C of atmospheric CO2 was enriched by 0.1‰ (Schmitt et al., 2012), and thus, this has relatively little impact. More significant are climate‐induced variations in the isotopic discrimination by plants (e.g., Kaplan, Prentice, & Buchmann, 2002). Three natural sources of CH4 (wetlands, biomass burning, and termites) are potentially influenced by the isotopic signature of leaf carbon. For example, C3 and C4 plants exhibit very different leaf carbon isotope ratios. Termites are not observed to show any relationship between emitted δ 13CH4 and the proportion of C3 versus C4 plants (Tyler et al., 1988), so we do not further consider this source term.
Leaf δ 13C is a function of the isotopic discrimination at each stage of gas transfer from the ambient environment to the chloroplasts within the leaf where photosynthesis occurs (Lloyd & Farquhar, 1994). Changes in moisture stress and water use efficiency of a plant will influence the amount of time that stomata are open and hence the overall ratio of intercellular and ambient CO2. During the LGM, the lower CO2 significantly reduced plant water use efficiency and is thought to be responsible for around 15% of the total change in δ 13C between the two time periods (Kaplan, Prentice, Knorr, et al., 2002). The lower temperatures and hence potential evaporation and generally drier environment will have competing impacts on the plant available moisture (e.g., Scheff et al., 2017).
We performed late preindustrial and LGM atmosphere‐only simulations with HadGEM2‐ES similar to those reported previously (Hopcroft & Valdes, 2015; H17) but with new intercellular leaf CO2 diagnostics (implemented within the dynamic vegetation scheme of Cox (2001), see supporting information, to calculate the isotopic discrimination by vegetation for the two time periods (following Kaplan, Prentice, & Buchmann, 2002; Lloyd & Farquhar, 1994). The resultant leaf to ambient CO2 ratio and anomalies are shown in Figure 2 (weighted by plant functional type fractional coverage). There is a global shift to less negative values, from a global average of −26.3‰ to −25.9‰. This is dominated by positive changes in semiarid regions. A reduction in δ 13C during the LGM relative to the late preindustrial in South Africa is in agreement with the model‐data comparison of Bragg et al. (2013).
To decompose the drivers of the leaf δ 13C change, we used an updated version of the land surface component of HadGEM2‐ES (JULES v4.1: Harper et al., 2016), to quantify the relative influence from changes in climate, atmospheric CO2, and vegetation, shown in Figure S1. Compared to the globally averaged leaf δ 13C increase of 0.36‰ (which is close to the global average change simulated with HadGEM2), CO2, climate, and dynamic vegetation separately cause changes of −0.29‰, −0.11‰, and 0.21‰, respectively (and averaged over preindustrial land points only). The CO2 forced change dominates in the tropics, whereas climate is more significant in midlatitudes. Both of these factors lead to negative excursions for the LGM relative to the preindustrial. The vegetation distribution change has a more widespread and mixed influence, with regions of negative shifts (LGM relative to preindustrial) in leaf δ 13C in southern Africa, western north America, and Eurasia, and positive changes in South East Asia, South America, and Australia.
The isotopic signature of CH4 emissions was calculated from the weighted sum of the simulated monthly leaf δ 13C values, using either wetland CH4 emissions or monthly burned area (H17), see Figure 2. We retain the 4‰ difference in δ 13CH4 between tropical and extratropical wetlands and assumed no offset between δ 13 C leaf and that of biomass burning (Chanton et al., 2000). We are unable to model this offset for wetland emissions, because, like most global models, JULES does not represent the production, transport, and oxidation of CH4 isotopes. The δ 13C shift is averaged regionally corresponding with the three‐box model.
The global mean wetland source increases by 0.7‰ and the biomass burning source increases by 1.6‰. For wetlands, the change in the distribution of emissions alone, mostly a relocation to the southern extratropics and tropics, causes a shift of 0.9‰. The leaf 13CH4 change alone induces a change of 3.3‰. The influence from the assumed offset between the tropical and extratropical wetlands was quantified by setting this difference to zero. It has a strong influence on the result: without this effect the overall signature change is negligible. We find that these separate changes do not combine linearly. For the biomass burning, the change in the global distribution of burning causes a 1.4‰ increase, while the leaf 13CH4 change alone causes a 1.3‰ increase.
The isotopic fractionation of CH4 shows a very small effect from the temperature dependence of fractionation during methanogenesis and methane consumption (following empirical evidence from Tyler et al., 1994) and as described in the supporting information (Blair et al., 1993; Conrad, 2005; Moosavi & Crill, 1998; Schaefer & Whiticar, 2008; Whiticar, 1999). The changes of −0.3‰ and −2‰ for the wetland signature and soil uptake fractionation factor respectively have a negligible impact on the δ 13CH4.
The calculated changes in the isotopic signature of wetland δ 13C averaged in three latitude bands (−0.1, 0.1, and 0.23‰ in the northern, tropical, and southern boxes, respectively) and similarly for biomass burning (2.6, −0.1, and 4.5‰) were included along with the small change in fractionation factors due to the temperature dependence of methanogenesis and uptake in the box model. The zonally averaged values can be summed to give the global total when weighted by changing sources strengths in the three bands, given in Table 1. The results are also shown in Figure 1. The simulated influence of leaf δ 13C on the atmospheric isotopic signature is relatively limited at 0.6‰. This does not capture a substantial fraction of the observed change and implies that changes to both the source mixture and the individual source signatures are required.
4. Inferring Source Changes From Ice Core Observations
To infer the CH4 sources that are consistent with the ice core observations, we employ a Bayesian framework (e.g., Denison et al., 2002). In this, information from ice cores is combined with the model simulations in a probabilistic formulation which accounts for the estimated uncertainties. We use the model simulations of H17 as prior information and condition the posterior probability density functions (PDFs) for the source strengths using the ice core observations. This approach differs from that of Fischer et al. (2008), because we incorporate prior information rather than sampling unconstrained.
We use a Metropolis‐Hastings Markov chain Monte Carlo algorithm (MCMC, see Gilks et al., 1995) to sample the posterior, conditioned on how well the model reproduces the LGM CH4 concentration and the preindustrial and LGM δ 13C and δD(CH4) values. Assigned observational uncertainties are 2 ppbv, 0.3‰, and 4.0‰ for CH4, δ 13CH4, and δDCH4, respectively (Mitchell et al., 2013; Möller et al., 2013; Sowers, 2006, 2010; WAIS Divide Project Members, 2015), see Table S1. These could be modified to account for representational uncertainty associated with the use of a coarse box model, but we have not done this here.
Uncertainty surrounding the isotopic signatures or fractionation factors for sources and sinks are taken into account by sampling from prior distributions. The mean values are detailed in supporting information Table S2 for sources (Etiope et al., 2008; Fisher et al., 2017; Miller, 2005; Sherwood et al., 2017; Snover et al., 2000; Thornton et al., 2016; Whiticar & Schaefer, 2007) and sinks (DeMore, 1993; Gierczak et al., 1997; Saueressig et al., 2001). The uncertainty estimates are given in Table S3 for sources (Etiope et al., 2008; Fischer et al., 2008; Quay et al., 1999; Sherwood et al., 2016; Snover et al., 2000) and sinks (Röckmann et al., 2011; Saueressig et al., 1995, 1996, 2001; Snover & Quay, 2000).
The prior information (summarized in Table 1) on the five methane source terms are normal distributions, with mean values equal to the LGM simulated values from H17, except for hydrate and geological sources for which the preindustrial values are used. The standard deviation of these prior distributions is set to the average of top‐down estimates of Kirschke et al. (2013). Assuming the quoted ranges are representative of the 95% limits, these equal 13%. We double this value to account for additional uncertainty for past time periods. Since modern biomass burning rates are influenced by anthropogenic activity (Bowman et al., 2009), we take a range from the literature of 14.3 TgCH4/year (Lamarque et al., 2010) and 37 TgCH4/year (Thonicke et al., 2005) and derive a normal distribution standard deviation of approximately 4 TgCH4/year. The geological and hydrates terms are also subject to wider uncertainties (Etiope et al., 2008; Petrenko et al., 2017. We assume that these terms have a similar relative uncertainty as biomass burning, that is, a standard deviation of 2.8 TgCH4/year.
Any change in biomass burning CH4 would be accompanied by a change in wildfire emission of nitrogen oxides and carbon monoxide which would impact on the CH4 lifetime. In H17 varying the LGM fire source between 10% and 84% of the late preindustrial source total, led to 2% and 8% relative increase in CH4 lifetime. Thus, we include this co‐variation.
The MCMC algorithm was run for one million iterations, with the initial 250,000 iterations discarded (Gilks et al., 1995). The four cases in Figure 3 are using (i) only CH4 concentration observations (green), (ii) both the concentration and δ 13CH4 or (iii) δDCH4, and (iv) the concentration and both isotopes (blue). The top three panels show the fit to the LGM ice core observations in each case. In the lower panels, the distributions of the model parameters are shown (i.e., inferred source strengths at the LGM). Prior distributions are shown by dashed lines.
The results show that the MCMC algorithm reproduces the observed LGM methane concentration, δ 13CH4 and δD(CH4) well. All four cases point to a reduced tropical wetland CH4 flux at the LGM. The posterior mean derived by considering all observations is 32.0 ± 5.0 TgCH4/year (mean ±1 standard deviation), and this is 59% of the prior mean (54.7 ± 15 TgCH4/year), or 44% of the preindustrial value. Northern wetlands are inferred to decrease to 22% of the preindustrial value when all observations are included, a reduction of 33% relative to the prior distribution. The prior and posterior are less divergent for termite, geological, and hydrate source strengths. The geological term is the only source with an increase at the LGM (posterior mean is 10.5 ± 2.5 TgCH4/yr), prior (10.0 ± 2.8 TgCH4/year). This emerges in both the all data and the concentration with δ 13C cases. It is driven by the high δ 13CH4 of geological sources (−33‰) and the model‐informed prior distribution (H17) on biomass burning (the other δ 13CH4‐enriched source), which points to a reduction in biomass burning at the LGM.
In the remaining cases, the solutions that stem from only considering the CH4 concentration alone or with one isotope type, are caused by the interplay between satisfying the prior information and the observations. For example, in the CH4‐only case, both the biomass burning and geological terms are reduced, because the isotopic constraint is absent and because this allows the wetland source to stay nearer to its prior distribution.
The inferences drawn separately from the δ 13CH4 and δD(CH4) are not entirely consistent. Including the concentration with deuterium leads to an increased hydrate flux and a reduced biomass burning term, while δ 13CH4 leads to the opposite behavior. Overall the reduced tropical wetlands are required to satisfy the concentration constraint, while relatively high biomass burning and geological sources are required by the δ 13CH4. The main constraint from δD is more difficult to isolate. It acts in a compensatory manner between the biomass burning and hydrate terms.
Doubling the observational or prior uncertainty estimates has very little impact on the posterior distributions (Figures S2 and S3). With uniform priors (Figure S4), the posterior distributions show even stronger reductions in both extratropical and tropical wetlands (for example, the posterior mean for tropical wetlands is 61% lower than the prior mean). The termite term is significantly higher than assumed with the original prior, but this is inconsistent with the best estimate for the LGM emission strength (H17). This demonstrates the potential value of introducing prior information. We also tested the approach by using the prior for the LGM but with the CH4 concentration and isotope observations replaced with the late preindustrial values (see Figure S5). Overall, the sensitivity tests compared in Figures S2–S5, support the main conclusions above.
5. Discussion
Using recent process‐based Earth Model simulations of the late preindustrial and LGM methane budget, we calculate the likely impact on atmospheric CH4 isotopes. This shows that current estimates of the methane cycle underestimate both the concentration change and the isotopic response, even accounting for uncertainty in LGM biomass burning emissions.
We extend previous work (Kaplan, 2002; Ringeval et al., 2013; H17) to calculate the potential influence of leaf δ 13C of terrestrial vegetation on the natural sources of CH4 (i.e., wetlands and biomass burning). Though the results are likely model dependent (because they are sensitive to the spatial distribution of emissions), they suggest a limited role for changes in the isotopic discrimination of vegetation in the glacial‐interglacial changes in δ 13CH4. This raises the question of why there is such a clear correlation between δ 13CH4 and CO2 in the ice core record, which Möller et al. (2013) have attributed in part to the influence of leaf 13C on the emissions of CH4 from tropical wetlands.
We also found that climatically driven changes in the boundary layer Cl sink contributed only 0.12‰ to the δ 13CH4 at the LGM relative to the preindustrial, consistent with past work (Levine, Wolff, Jones, & Sime, 2011). A substantial increase in sea salt aerosols at high latitudes during the LGM (Fischer et al., 2007) may have enhanced the Cl sink (Levine, Wolff, Jones, & Sime, 2011), but this is yet to be included in models (Levine et al., 2014). We have also not included any terrestrial sources of CH3Cl.
We used a probabilistic approach to show that both biomass burning and geologic sources of methane may have played a role in setting the observed enrichment of atmospheric CH4 (Bock et al., 2017). The latter is supported by the proposed sea level control on marine seepage of CH4 (Etiope et al., 2008; Luyendyk et al., 2005). Hence, our modeling results are potentially consistent with the Möller et al. (2013) observations because of the high degree of covariance between CO2, sea level, and other climatically relevant variables on a glacial‐interglacial timescale.
Further, the estimated modern day natural geological CH4 source of around 54 TgCH4/year (Kirschke et al., 2013) cannot be reconciled with the required reduction in total methane sources at the LGM (H17). This is because the geological source likely increased at the LGM and because 54 TgCH4/year is a substantial fraction of the natural source. Recent mass balance calculations based on late glacial 14CH4 measurements (Petrenko et al., 2017) also support this.
The inferred large reduction in wetland emissions is not simulated by H17 and is at the extreme limit of the uncertainty range of recent model‐based estimates of the LGM‐PI wetland change, (29% to 67%: Hopcroft et al., 2014; Ringeval et al., 2013; Weber et al., 2010). However, several processes (e.g., wetland carbon cycle processes, nutrient status, and tree‐mediated transport) that are not properly represented (Melton et al., 2013; Pangala et al., 2017; Wania et al., 2013; H17) need to be evaluated.
6. Conclusions
Current process‐based estimates suggest that a 50% reduction in sources is required to explain the CH4 concentration during the LGM. Bottom‐up estimates of emissions fail to replicate this. A comprehensive Earth System model study (H17) also underestimates the changes in CH4 stable isotopes. Accounting for changes in isotopic discrimination of vegetation does not explain the observed LGM‐PI δ 13CH4 shift.
We applied a Bayesian framework to resolve the potentially conflicting information from models and observations. The results suggest that the concentration change was predominantly driven by wetlands, but the inferred emissions response is at the extreme end of model predictions. The isotopic change is either driven by a relatively limited reduction in biomass burning at the LGM, perhaps brought about by human activities (Kaplan et al., 2016), or an increase in nonhydrate geological CH4 emissions. Future work needs to better understand the climatic sensitivity of natural CH4 sources, including wetland and geological terms, and this study suggests that incorporation of CH4 stable isotopes into Earth System model simulations is one avenue for further progress.
Supporting information
Acknowledgments
P. O. H. was funded by the NERC projects NE/I010912/1 and NE/P002536/1, by a Past Earth Network Feasibility Study (EP/M008363/1), and presently by a University of Birmingham Fellowship. J. O. K. was supported by the European Research Council (313797, COEVOLVE). HadGEM2‐ES simulations were performed using UK national computing facility ARCHER (http://www.archer.ac.uk). We thank CMS at NCAS for maintaining HadGEM2‐ES on this facility. HadGEM2‐ES output is available for further analysis from http://www.bridge.bris.ac.uk/resources/simulations.
Hopcroft, P. O. , Valdes, P. J. , & Kaplan, J. O. (2018). Bayesian analysis of the glacial‐interglacial methane increase constrained by stable isotopes and Earth System modeling. Geophysical Research Letters, 45, 3653–3663. https://doi.org/10.1002/2018GL077382
References
- Allan, W. , Struthers, H. , & Lowe, D. (2007). Methane carbon isotope effects caused by atomic chlorine in the marine boundary layer: Global model results compared with Southern Hemisphere measurements. Journal of Geophysical Research, 112, D04306 https://doi.org/10.1029/2006JD007369 [Google Scholar]
- Blair, N. , Boehme, S. , & Carter, W., Jr. (1993). The carbon isotope biogeochemistry of methane production in anoxic sediments: 1. Field observations In Oremland R. S. (Ed.), Biogeochemistry of Global Change (pp. 574–593). New York: Chapman and Hall. [Google Scholar]
- Bock, M. , Schmitt, J. , Beck, J. , Seth, B. , Chappellaz, J. , & Fischer, H. (2017). Glacial/interglacial wetland, biomass burning, and geologic methane emissions constrained by dual stable isotopic CH4 ice core records. Proceedings of the National Academy of Sciences of the United States of America, 114, E5778–E5786. https://doi.org/10.1073/pnas.1613883114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Booth, B. , Dunstone, N. , Halloran, P. , Andrews, T. , & Bellouin, N. (2012). Aerosols implicated as a prime driver of twentieth‐century North Atlantic climate variability. Nature, 484(7393), 228–232. https://doi.org/10.1038/nature10946 [DOI] [PubMed] [Google Scholar]
- Bousquet, P. , Ringeval, B. , Pison, I. , Dlugokencky, E. , Brunke, E.‐G. , Carouge, C. , et al. (2011). Source attribution of the changes in atmospheric methane for 2006–2008. Atmospheric Chemistry and Physics, 11, 3689–3700. [Google Scholar]
- Bowman, D. , Balch, J. , Artaxo, P. , Bond, W. , Carlson, J. M. , Cochrane, M. A. , et al. (2009). Fire in the Earth System. Science, 324, 481–484. [DOI] [PubMed] [Google Scholar]
- Bragg, F. , Prentice, I. , Harrison, S. , Eglinton, G. , & Foster, P. (2013). Stable isotope and modelling evidence for CO2 as a driver of glacial–interglacial vegetation shifts in southern Africa. Biogeosciences, 10, 2001–2010. https://doi.org/10.5194/bg-10-2001-2013 [Google Scholar]
- Caesar, J. , Palin, E. , Liddicoat, S. , Lowe, J. , Burke, E. , Pardaens, A. , et al. (2013). Response of the HadGEM2 Earth System model to future greenhouse gas emissions pathways to the year 2300. Journal of Climate, 26, 3275–3284. [Google Scholar]
- Chanton, J. , Rutkowski, C. , Schwartz, C. , Ward, D. , & Boring, L. (2000). Factors influencing the stable carbon isotopic signature of methane from combustion and biomass burning. Journal of Geophysical Research, 105(D2), 1867–1877. [Google Scholar]
- Collins, W. , Bellouin, N. , Doutriaux‐Boucher, M. , Gedney, N. , Halloran, P. , Hinton, T. , et al. (2011). Development and evaluation of an Earth‐system model—HadGEM2. Geoscientific Model Development, 4, 1051–1075. https://doi.org/10.5194/gmd-4-1051-2011 [Google Scholar]
- Conrad, R. (2005). Quantification of methanogenic pathways using stable carbon isotopic signatures: A review and a proposal. Organic Geochemistry, 36, 739–752. [Google Scholar]
- Cox, P. (2001). Description of the TRIFFID dynamic global vegetation model, Hadley Centre, Met Office, Technical Note 24, UK. [Google Scholar]
- DeMore, W. (1993). Rate constant ratio for the reaction of OH with CH3D and CH4. Journal of Physical Chemistry, 97, 8564–8566. [Google Scholar]
- Denison, D. , Holmes, C. , Mallick, B. , & Smith, A. (2002). Bayesian methods for nonlinear classification and regression. Chichester: John Wiley. [Google Scholar]
- Etiope, G. , Milkov, A. , & Derbyshire, E. (2008). Did geologic emissions of methane play any role in Quaternary climate change? Global and Planetary Change, 61, 79–88. [Google Scholar]
- Ferretti, D. , Miller, J. , White, J. , Etheridge, D. , Lassey, K. , Lowe, D. , et al. (2005). Unexpected changes to the global methane budget over the past 2000 years. Science, 309, 1714–1717. [DOI] [PubMed] [Google Scholar]
- Fischer, H. , Siggaard‐Andersen, M.‐L. , Ruth, U. , Röthlisberger, R. , & Wolff, E. (2007). Glacial/interglacial changes in mineral dust and sea‐salt records in polar ice cores: Sources, transport, and deposition. Reviews of Geophysics, 45, RG1002 https://doi.org/10.1029/2005RG000192 [Google Scholar]
- Fischer, H. , Behrens, M. , Bock, M. , Richter, U. , Schmitt, J. , Loulergue, L. , et al. (2008). Changing boreal methane sources and constant biomass burning during the last termination. Nature, 452, 864–867. https://doi.org/10.1038/nature06825 [DOI] [PubMed] [Google Scholar]
- Fisher, R. , France, J. , Lowry, D. , Lanoisellé, M. , Brownlow, R. , Pyle, J. , et al. (2017). Measurement of the 13C isotopic signature of methane emissions from Northern European wetlands. Global Biogeochemical Cycles, 31, 605–623. https://doi.org/10.1002/2016GB005504 [Google Scholar]
- Gedney, N. , Cox, P. , & Huntingford, C. (2004). Climate feedback from wetland methane emissions. Geophysical Research Letters, 31, L20503 https://doi.org/10.1029/2004GL020919 [Google Scholar]
- Gierczak, T. , Talukdar, R. , Herdndon, S. , Vaghjiani, G. L. , & Ravishankara, A. R. (1997). Rate coefficients for the reactions of hydroxyl radicals with methane and deuterated methanes. Journal of Physical Chemistry, 101, 3125–2134. [Google Scholar]
- Gilks, W. , Richardson, S. , & Spiegelhalter, D. (1995). Introducing Markov chain Monte Carlo methods In Gilks W. R., Richardson S., & Spiegelhalter D. J. (Eds.), Markov chain Monte Carlo in practice. London: Chapman and Hall. [Google Scholar]
- HadGEM2 Development Team (2011). The HadGEM2 family of Met Office unified model climate configurations. Geoscientific Model Development, 4, 7233–757. https://doi.org/10.5194/gmd-4-723-2011 [Google Scholar]
- Harper, A. , Cox, P. , Friedlingstein, P. , Wiltshire, A. , Jones, C. , Sitch, S. , et al. (2016). Improved representation of plant functional types and physiology in the Joint UK Land Environment Simulator (JULES v4.2) using plant trait information. Geoscientific Model Development, 9, 2415–2440. https://doi.org/10.5194/gmd-9-2415-2016 [Google Scholar]
- Hayman, G. , O'Connor, F. , Dalvi, M. , Clark, D. , Gedney, N. , Huntingford, C. , et al. (2014). Comparison of the HadGEM2 climate‐chemistry model against in situ and SCIAMACHY atmospheric methane data. Atmospheric Chemistry and Physics, 14, 13,257–13,280. [Google Scholar]
- Hopcroft, P. , & Valdes, P. (2015). Last Glacial Maximum constraints on the Earth System model HadGEM2‐ES. Climate Dynamics, 45(5), 1657–1672. https://doi.org/10.1007/s00382-014-2421-0 [Google Scholar]
- Hopcroft, P. , Valdes, P. , Wania, R. , & Beerling, D. (2014). Limited response of peatland CH4 emissions to abrupt changes in Atlantic ocean circulation in glacial states. Climate of the Past, 10, 137–154. https://doi.org/10.5194/cp-10-137-2014 [Google Scholar]
- Hopcroft, P. , Valdes, P. , O'Connor, F. , Kaplan, J. , & Beerling, D. (2017). Understanding the glacial methane cycle. Nature Communications, 8, 14383 https://doi.org/10.1038/ncomms14383 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones, C. , Hughes, J. , Bellouin, N. , Hardiman, S. , Jones, G. , Knight, J. , et al. (2011). The HadGEM2‐ES implementation of CMIP5 centennial simulations. Geoscientific Model Development, 4, 543–570. [Google Scholar]
- Kandlbauer, J. , Hopcroft, P. , Valdes, P. , & Sparks, R. (2013). Climate and carbon cycle response to the 1815 Tambora volcanic eruption. Journal of Geophysical Research: Atmospheres, 118, 12,497–12,507. https://doi.org/10.1002/2013JG019767 [Google Scholar]
- Kaplan, J. (2002). Wetlands at the Last Glacial Maximum: Distribution and methane emissions. Geophysical Research Letters, 29(6), 1079 https://doi.org/10.1029/2001GL013366 [Google Scholar]
- Kaplan, J. , Prentice, I. , & Buchmann, N. (2002). The stable carbon isotope composition of the terrestrial biosphere: Modeling at scales from the leaf to the globe. Global Biogeochemical Cycles, 16(4), 1060 https://doi.org/10.1029/2001GB001403 [Google Scholar]
- Kaplan, J. , Prentice, I. , Knorr, W. , & Valdes, P. (2002). Modeling the dynamics of terrestrial carbon storage since the Last Glacial Maximum. Geophysical Research Letters, 29(22), 2074 https://doi.org/10.1029/2002GL015230 [Google Scholar]
- Kaplan, J. , Folberth, G. , & Hauglustaine, D. (2006). Role of methane and biogenic volatile organic compound sources in the late glacial and Holocene fluctuations of atmospheric methane concentrations. Global Biogeochemical Cycles, 20, GB2016 https://doi.org/10.1029/2005GB002590 [Google Scholar]
- Kaplan, J. , Bigelow, N. H. , Prentice, I. C. , Harrison, S. P. , Bartlein, P. J. , Christensen, T. R. , et al. (2003). Climate change and Arctic ecosystems: 2. Modeling, paleodata‐model comparisons, and future projections. Journal of Geophysical Research, 108(D19), 8171 https://doi.org/10.1029/2002JD002559 [Google Scholar]
- Kaplan, J. , Pfeiffer, M. , Kolen, J. , & Davis, B. (2016). Large scale anthropogenic reduction of forest cover in Last Glacial Maximum Europe. PLoS One, 11(11), e0166726 https://doi.org/10.1371/journal.pone.0166726 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kirschke, S. , Bousquet, P. , Ciais, P. , Saunois, M. , Canadell, J. G. , Dlugokencky, E. J. , et al. (2013). Three decades of global methane sources and sinks. Nature Geoscience, 6, 813–822. https://doi.org/10.1038/NGEO1955 [Google Scholar]
- Lamarque, J.‐F. , Bond, T. C. , Eyring, V. , Heil, A. , Klimont, Z. , Lee, D. , et al. (2010). Historical (1850–2000) gridded anthropogenic and biomass burning emissions of reactive gases and aerosols: Methodology and application. Atmospheric Chemistry and Physics, 10, 7017–7039. [Google Scholar]
- Lassey, K. , Lowe, D. , & Manning, M. (2000). The trend in atmospheric methane δ 13C and implications for isotopic constraints on the global methane budget. Global Biogeochemical Cycles, 14(1), 41–49. [Google Scholar]
- Levine, J. , Wolff, E. , Jones, A. , Sime, L. , Valdes, P. , Archibald, A. , et al. (2011). Reconciling the changes in atmospheric methane sources and sinks between the Last Glacial Maximum and the pre‐industrial era. Geophysical Research Letters, 38, L23804 https://doi.org/10.1029/2011GL049545 [Google Scholar]
- Levine, J. , Wolff, E. , Jones, A. , & Sime, L. (2011). The role of atomic chlorine in glacial‐interglacial changes in the carbon‐13 content of atmospheric methane. Geophysical Research Letters, 38, L04801 https://doi.org/10.1029/2010GL046122 [Google Scholar]
- Levine, J. , Yang, X. , Jones, A. E. , & Wolff, E. W. (2014). Sea salt as an ice core proxy for past sea ice extent: A process‐based model study. Journal of Geophysical Research: Atmospheres, 119, 5737–5756. https://doi.org/10.1002/2013JD020925 [Google Scholar]
- Lloyd, J. , & Farquhar, G. (1994). 13C discrimination during CO2 assimilation by the terrestrial biosphere. Oecologia, 99, 201–215. [DOI] [PubMed] [Google Scholar]
- Loulergue, L. , Schilt, A. , Spahni, R. , Masson‐Delmotte, V. , Blunier, T. , Lemieux, B. , et al. (2008). Orbital and millennial‐scale features of atmospheric CH4 over the past 800,000 years. Nature, 453, 383–386. [DOI] [PubMed] [Google Scholar]
- Luyendyk, B. , Kennett, J. , & Clark, J. (2005). Hypothesis for increased atmospheric methane input from hydrocarbon seeps on exposed continental shelves during glacial low sea level. Marine and Petroleum Geology, 22, 591–596. [Google Scholar]
- Marthews, T. , Dadson, S. , Lehner, B. , Abele, S. , & Gedney, N. (2015). A high‐resolution global dataset of topographic index values for use in large‐scale hydrological modelling. Hydrology and Earth System Sciences, 19, 91–104. [Google Scholar]
- Melton, J. , Wania, R. , Hodson, E. , Poulter, B. , Ringeval, B. , Spahni, R. , et al. (2013). Present state of global wetland and wetland CH4 modelling: Conclusions from a model intercomparison project (WETCHIMP). Biogeosciences, 10, 753–788. https://doi.org/10.5194/bg-10-753-2013 [Google Scholar]
- Miller, J. (2005). chap The carbon isotopic composition of atmospheric methane and its constraint on the global methane budget In Flanagan L., Ehleringer J., & Pataki D. (Eds.), Stable isotopes and biosphere‐atmosphere interactions: Processes and Biological Controls (Vol. 16, pp. 288–312). San Diego, CA: Elsevier Academic Press. [Google Scholar]
- Mitchell, L. , Brook, E. , Lee, J. E. , Buizert, C. , & Sowers, T. (2013). Constraints on the Late Holocene Anthropogenic Contribution to the Atmospheric Methane Budget. Science, 342, 964–966. [DOI] [PubMed] [Google Scholar]
- Möller, L. , Sowers, T. , Bock, M. , Spahni, R. , Behrens, M. , Schmitt, J. , et al. (2013). Independent variations of CH4 emissions and isotopic composition over the past 160,000 years. Nature Geoscience, 6, 885–891. https://doi.org/10.1038/NGEO1922 [Google Scholar]
- Moosavi, S. , & Crill, P. (1998). CH4 oxidation by tundra wetlands as measured by a selective inhibitor technique. Journal of Geophysical Research, 103(D22), 29,093–29,106. [Google Scholar]
- Murray, L. , Mickley, L. J. , Kaplan, J. O. , Sofen, E. D. , Pfeiffer, M. , Alexander, B. , et al. (2014). Factors controlling variability in the oxidative capacity of the troposphere since the Last Glacial Maximum. Atmospheric Chemistry and Physics, 14, 3589–3622. https://doi.org/10.5194/acp-14-3589-2014 [Google Scholar]
- Myhre, G. , Shindell, D. , Breon, F.‐M. , Collins, W. , Fuglestvedt, J. , Huang, J. , et al. (2013). Anthropogenic and natural radiative forcing In Stocker T., et al. (Eds.), Climate Change 2013: The physical science basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, UK and New York: Cambridge University Press. [Google Scholar]
- Nisbet, E. , Dlugokencky, E. J. , Manning, M. R. , Lowry, D. , Fisher, R. E. , France, J. L. , et al. (2016). Rising atmospheric methane: 2007–2014 growth and isotopic shift. Global Biogeochemical Cycles, 30, 1356–1370. https://doi.org/10.1002/2016GB005406 [Google Scholar]
- O'Connor, F. , Johnson, C. , Morgenstern, O. , Abraham, N. , Braesicke, P. , Dalvi, M. , et al. (2014). Evaluation of the new UKCA climate‐composition model—Part 2: The troposphere. Geoscientific Model Development, 7, 41–91. https://doi.org/10.5194/gmd-7-41-2014 [Google Scholar]
- Pangala, S. , Enrich‐Prast, A. , Basso, L. , Peixoto, R. B. , Bastviken, D. , Hornibrook, E. R. C. , et al. (2017). Large emissions from floodplain trees close the Amazon methane budget. Nature, 552, 230–234. https://doi.org/10.1038/nature24639 [DOI] [PubMed] [Google Scholar]
- Petrenko, V. , Smith, A. , Schaefer, H. , Riedel, K. , Brook, E. , Baggenstos, D. , et al. (2017). Minimal geological methane emissions during the Younger Dryas‐Preboreal abrupt warming event. Nature, 548, 443–446. https://doi.org/10.1038/nature23316 [DOI] [PubMed] [Google Scholar]
- Pfeiffer, M. , Spessa, A. , & Kaplan, J. (2013). A model for global biomass burning in preindustrial time: LPJ‐LMfire (v1.0). Geoscientific Model Development, 6, 643–685. https://doi.org/10.5194/gmd-6-643-2013 [Google Scholar]
- Platt, U. , Allan, W. , & Lowe, D. (2004). Hemispheric average Cl atom concentration from 13C/12C ratios in atmospheric methane. Atmospheric Chemistry and Physics, 4, 2393–2399. [Google Scholar]
- Prather, M. , Holmes, C. , & Hsu, J. (2012). Reactive greenhouse gas scenarios: Systematic exploration of uncertainties and the role of atmospheric chemistry. Geophysical Research Letters, 39, L09803 https://doi.org/10.1029/2012GL051440 [Google Scholar]
- Quay, P. , Statesman, J. , Wilbur, D. , Snover, A. , Dlugokencky, E. , & Brown, T. (1999). The isotopic composition of atmospheric methane. Global Biogeochemical Cycles, 13(2), 445–461. [Google Scholar]
- Rigby, M. , Montzka, S. , Prinn, R. , White, J. W. C. , Young, D. , OŠDoherty, S. , et al. (2017). Role of atmospheric oxidation in recent methane growth. Proceedings of the National Academy of Sciences of the United States of America, 114(21), 5373–5377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ringeval, B. , Hopcroft, P. , Valdes, P. , Ciais, P. , Ramstein, G. , Dolman, A. , & Kageyama, M. (2013). Response of methane emissions from wetlands to the Last Glacial Maximum and an idealised Dansgaard‐Oeschger event: Insights from two models of different complexity. Climate of the Past, 9, 149–171. https://doi.org/10.5194/cp-0-149-2013 [Google Scholar]
- Röckmann, T. , Brass, M. , Borchers, R. , & Engel, A. (2011). The isotopic composition of methane in the stratosphere: high‐altitude balloon sample measurements. Atmospheric Chemistry and Physics, 11, 13,287–13,304. https://doi.org/10.5194/acp-11-13287-2011 [Google Scholar]
- Sanderson, M. (1996). Biomass of termites and their emission of methane and carbon dioxide: A global database. Global Biogeochemical Cycles, 10(4), 543–557. [Google Scholar]
- Saueressig, G. , Bergamaschi, P. , Crowley, J. , Fischer, H. , & Harris, G. (1995). Carbon kinetic isotope effect in the reaction CH4 with Cl atoms. Geophysical Research Letters, 22(10), 1225–1228. [Google Scholar]
- Saueressig, G. , Bergamaschi, P. , Crowley, J. , Fischer, H. , & Harris, G. (1996). D/H kinetic isotope effect in the reaction CH4 + Cl. Geophysical Research Letters, 23(24), 3619–3622. [Google Scholar]
- Saueressig, G. , Bergamaschi, P. , Crowley, J. , Fischer, H. , & Harris, G. (2001). Carbon 13 and D kinetic isotope effects in the reactions of CH4 with O(1D) and OH: New laboratory measurements and their implications for the isotopic composition of stratospheric methane. Journal of Geophysical Research, D19, 23,127–23,138. [Google Scholar]
- Schaefer, H. , & Whiticar, M. (2008). Potential glacial‐interglacial changes in stable carbon isotope ratios of methane sources and sink fractionation. Global Biogeochemical Cycles, 22, GB1001 https://doi.org/10.1029/2006GB002889 [Google Scholar]
- Schaefer, H. , Mikaloff Fletcher, S. , Veidt, C. , Lassey, K. , Brailsford, G. , Bromley, T. , et al. (2016). A 21st century shift from fossil‐fuel to biogenic methane emissions indicated by 13CH4 . Science, 352, 80–84. https://doi.org/10.1126/science.aad2705 [DOI] [PubMed] [Google Scholar]
- Scheff, J. , Seager, R. , Liu, H. , & Coats, S. (2017). Are glacials dry? Consequences for paleoclimatology and for greenhouse warming. Journal of Climate, 30, 6593–6609. https://doi.org/10.1175/JCLI-D-16-0854.1 [Google Scholar]
- Schmitt, J. , Schneider, R. , Elsig, E. , Leuenberger, D. , Lourantou, A. , Chappellaz, J. , et al. (2012). Carbon isotope constraints on the deglacial CO2 rise from ice cores. Science, 336, 711–714. [DOI] [PubMed] [Google Scholar]
- Sherwood, O. , Schwietzke, S. , Arlin, V. , & Etiope, G. (2016). Global inventory of fossil and non‐fossil methane δ 13C source signature measurements for improved atmospheric modeling (Tech. Rep.). U.S. National Oceanic and Atmospheric Administration. https://doi.org/10.15138/G37P4D
- Sherwood, O. , Schwietzke, S. , Arling, V. , & Etiope, G. (2017). Global inventory of gas geochemistry data from fossil fuel, microbial and burning sources, version 2017. Earth System Science Data, 9, 639–656. https://doi.org/10.5194/essd-9-639-2017 [Google Scholar]
- Singarayer, J. , & Valdes, P. (2010). High‐latitude climate sensitivity to ice‐sheet forcing over the last 120 kyr. Quaternary Science Reviews, 29(1–2), 43–55. https://doi.org/10.1016/j.quascirev.2009.10.011 [Google Scholar]
- Snover, A. , & Quay, P. (2000). Hydrogen and carbon kinetic isotope effects during soil uptake of atmospheric methane. Global Biogeochemical Cycles, 14(1), 25–39. [Google Scholar]
- Snover, A. , Quay, P. , & Has, W. (2000). The D/H content of methane emitted from biomass burning. Global Biogeochemical Cycles, 14(1), 11–24. [Google Scholar]
- Sowers, T. (2006). Late Quaternary atmospheric CH4 isotope record suggests marine clathrates are stable. Science, 311, 838–840. [DOI] [PubMed] [Google Scholar]
- Sowers, T. (2010). Atmospheric methane isotope records covering the Holocene period. Quaternary Science Reviews, 29, 213–221. [Google Scholar]
- Thonicke, K. , Prenctice, I. , & Hewitt, C. (2005). Modeling glacial‐interglacial changes in global fire regimes and trace gas emissions. Global Biogeochemical Cycles, 19, GB3008 https://doi.org/10.1029/2004GB002278 [Google Scholar]
- Thornton, B. , Wik, M. , & Crill, P. (2016). Double‐counting challenges the accuracy of high‐latitude methane inventories. Geophysical Research Letters, 43, 12,569–12,577. https://doi.org/10.1002/2016GL071772 [Google Scholar]
- Tyler, S. , Zimmerman, P. , Cumberbatch, C. , Greenberg, J. , Westberg, C. , & Darlington, J. (1988). Measurements and interpretation of δ 13C of methane from termites, rice paddies, and wetlands in Kenya. Global Biogeochemical Cycles, 2(4), 341–355. [Google Scholar]
- Tyler, S. , Crill, P. , & Brailsford, G. (1994). 13C/12C fractionation of methane during oxidation in a temperate forested soil. Geochimica et Cosmochinica Acta, 58(6), 1625–1633. [Google Scholar]
- Valdes, P. , Beerling, D. , & Johnson, C. (2005). The ice age methane budget. Geophysical Research Letters, 32, L02704 https://doi.org/10.1029/2004GL021004 [Google Scholar]
- WAIS Divide Project Members (2015). Precise interpolar phasing of abrupt climate change during the last ice age. Nature, 520, 661–665. [DOI] [PubMed] [Google Scholar]
- Wania, R. , Ross, I. , & Prentice, I. C. (2010). Implementation and evaluation of a new methane model within a dynamic global vegetation model: LPJ‐WHyMe v1.3.1. Geoscientific Model Development, 3, 565–584. https://doi.org/10.5194/gmd-3-565-2010 [Google Scholar]
- Wania, R. , Melton, J. , Hodson, E. , Poulter, B. , Ringeval, B. , Spahni, R. , et al. (2013). Present state of global wetland extent and wetland methane modelling: Methodology of a model intercomparison project (WETCHIMP). Geoscientific Model Development, 6, 617–641. [Google Scholar]
- Weber, S. , Drury, A. , Toonen, W. , & van Weele, M. (2010). Wetland methane emissions during the Last Glacial Maximum estimated from PMIP2 simulations: Climate, vegetation, and geographic controls. Journal of Geophysical Research, 115, D06111 https://doi.org/10.1029/2009JD012110 [Google Scholar]
- Whiticar, M. (1999). Carbon and hydrogen isotope systematics of bacterial formation and oxidation of methane. Chemical Geology, 161, 291–314. [Google Scholar]
- Whiticar, M. , & Schaefer, H. (2007). Constraining past global tropospheric methane budgets with carbon and hydrogen isotope ratios in ice. Global Biogeochemical Cycles, 365, 1793 https://doi.org/10.1098/rsta.2007.2048 [DOI] [PubMed] [Google Scholar]
- Worden, J. , Bloom, A. , Pandey, S. , Jiang, Z. , Worden, H. M. , Walker, T. W. , et al. (2018). Reduced biomass burning emissions reconcile conflicting estimates of the post‐2006 atmospheric methane budget. Nature Communications, 8, 2227 https://doi.org/10.1038/s41467-017-02246-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.