Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/ doi:10.5194/acp-16-195-2016 Author(s) 2016. CC Attribution 3.0 License.
a1,2, M. Krol3,4,1, M. van Weele2, T. van Noije2, P. Le Sager2, and T. Rckmann1
1Institute for Marine and Atmospheric Research Utrecht, Utrecht University, Utrecht, The Netherlands
2Royal Netherlands Meteorological Institute (KNMI), De Bilt, The Netherlands
3Meteorology and Air Quality, Wageningen University and Research Center, Wageningen, The Netherlands
4Netherlands Institute for Space Research (SRON), Utrecht, The Netherlands
Correspondence to: N. Bnda ([email protected])
Received: 29 April 2015 Published in Atmos. Chem. Phys. Discuss.: 10 July 2015 Revised: 9 November 2015 Accepted: 29 November 2015 Published: 18 January 2016
Abstract. The CH4 growth rate in the atmosphere showed large variations after the Pinatubo eruption in June 1991. A decrease of more than 10 ppb yr1 in the growth rate over the course of 1992 was reported, and a partial recovery in the following year. Although several reasons have been proposed to explain the evolution of CH4 after the eruption, their contributions to the observed variations are not yet resolved. CH4 is removed from the atmosphere by the reaction with tropospheric OH, which in turn is produced by O3 photolysis under UV radiation. The CH4 removal after the Pinatubo eruption might have been affected by changes in tropospheric UV levels due to the presence of stratospheric SO2 and sulfate aerosols, and due to enhanced ozone depletion on Pinatubo aerosols. The perturbed climate after the eruption also altered both sources and sinks of atmospheric CH4. Furthermore, CH4 concentrations were inuenced by other factors of natural variability in that period, such as El NioSouthern Oscillation (ENSO) and biomass burning events. Emissions of CO, NOX and non-methane volatile organic compounds (NMVOCs) also affected CH4 concentrations indirectly by inuencing tropospheric OH levels.
Potential drivers of CH4 variability are investigated using the TM5 global chemistry model. The contribution that each driver had to the global CH4 variability during the period 1990 to 1995 is quantied. We nd that a decrease of 8 10 ppb yr1 CH4 is explained by a combination of the above processes. However, the timing of the minimum growth rate is found 69 months later than observed. The long-term decrease in CH4 growth rate over the period 1990 to 1995 is well captured and can be attributed to an increase in OH
Can we explain the observed methane variability after the Mount Pinatubo eruption?
N. Bnd
concentrations over this time period. Potential uncertainties in our modelled CH4 growth rate include emissions of CH4 from wetlands, biomass burning emissions of CH4 and other compounds, biogenic NMVOC and the sensitivity of OH to NMVOC emission changes. Two inventories are used for CH4 emissions from wetlands, ORCHIDEE and LPJ, to investigate the role of uncertainties in these emissions. Although the higher climate sensitivity of ORCHIDEE improves the simulated CH4 growth rate change after Pinatubo, none of the two inventories properly captures the observed CH4 variability in this period.
1 Introduction
Methane (CH4) is the second most important anthropogenic greenhouse gas after carbon dioxide (CO2). Its evolution in the atmosphere since the beginning of the record of continuous atmospheric CH4 measurements in the 1980s is not fully understood, with large discrepancies between bottom-up and top-down estimates of CH4 sources and sinks (Kirschke et al., 2013). One of the events that affected CH4 concentrations was the eruption of Mt Pinatubo on 15 June 1991, the largest eruption in the last century. The eruption caused perturbations to climate and photochemistry for a few years afterwards. We investigate here the sensitivity of CH4 concentrations to these perturbations and our ability to explain the observed CH4 variations in the atmosphere in the early 1990s.
Published by Copernicus Publications on behalf of the European Geosciences Union.
196 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
of natural variability (Telford et al., 2009; Bekki and Pyle, 1994).
Another potential effect of the eruption on the CH4 budget is altered transport between the troposphere and the stratosphere due to the stratospheric warming by volcanic particles (Schaufer and Daniel, 1994). CH4 concentrations in the post-Pinatubo period were also affected by natural variability not related to Pinatubo, such as the 11-year solar cycle, the El NioSouthern Oscillation (ENSO) cycle and biomass burning events. Signicant reductions in the anthropogenic emissions from gas production and distribution in the former Soviet Union (FSU) might have occurred in this period (Law and Nisbet, 1996). However, gaps in reporting make the magnitude and timing of these reductions quite uncertain.
Previous bottom-up studies quantied the impact of individual processes on the CH4 abundance, without attempting to solve the whole CH4 budget. Bekki and Law (1997) found compensating effects of temperature on the CH4 emissions and the CH4 removal by OH on a global scale. Dlugokencky et al. (1996) showed that the increase in both CH4 and CO mixing ratios in late 1991 and early 1992 could be related to a decrease in OH due to UV scattering by stratospheric aerosols and UV absorption by sulfur dioxide (SO2) from the eruption. A more recent estimate from B14 nds a decrease of 11.2 and 6.6 Tg, respectively, in the CH4 removal in the rst and second years following the eruption due to stratospheric sulfur. The above decreases in CH4 removal translate into corresponding increases of 4 and 2 ppb yr1, respectively, in the CH4 growth rate. Here and in the remainder of the paper, we use the ratio of 2.78 Tg ppb1 reported by the IPCC (Denman et al., 2007) to convert from CH4 burden changes to growth rate variations. Using a two-dimensional model, Bekki et al. (1994) attributed half of the decrease in CH4 and CO growth rates in 1992 to stratospheric O3 depletion. Additionally, Dlugokencky et al. (1994) and Law and Nisbet (1996) estimated that a decrease in the gas leak emissions from the FSU could be of the order of 15 Tg yr1, and would also explain part of the observed decrease in the CH4 growth rate in 1992. Using a four-box model, Schaufer and Daniel (1994) found that an increase in the exchange rate between the stratosphere and troposphere due to volcanic aerosol heating can cause lower CH4 concentrations in the troposphere.
Prior attempts to explain the CH4 variability in this period include inverse modelling studies and a column model study.The inverse modelling studies of Bousquet et al. (2006) and Wang et al. (2004) (B06 and W04 from here on, respectively) nd a positive anomaly in the CH4 ux to the atmosphere in 1991, and a negative anomaly in 1992, with a difference of about 30 Tg between the two years. Both studies nd a signicant contribution from wetland emissions to this anomaly and a small contribution from biomass burning. However, the two studies use OH variations of opposite sign. The OH elds in B06 are determined by methyl chloroform inversions, showing a reduced OH sink of CH4
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
After a stable CH4 growth rate of 10 to 13 ppb yr1 in the late 1980s, the growth rate showed large uctuations in the early 1990s. An increased CH4 growth rate of about 17 ppb yr1 was registered by the NOAA network in 1991, followed by a strong decline in the growth rate during the next year, with values reaching nearly zero (Dlugokencky et al., 1994). The growth rate recovered to 6 ppb yr1 in 1993. Processes driving these variations could be related to either the CH4 sources or the CH4 sinks.
CH4 is emitted into the atmosphere from anthropogenic activities (fossil fuel production, agriculture and waste treatment), biomass burning, and from natural sources (wetlands, geological activity, termites). The main sink of atmospheric CH4 is the reaction with the hydroxyl radical (OH) in the troposphere, which removes about 80 % of the CH4. Other removal processes are soil uptake, reactions with chlorine (Cl) in both the troposphere and stratosphere, and reactions with OH and excited oxygen (O(1D)) atoms in the stratosphere. Tropospheric OH is produced by the photolysis of ozone (O3) at wavelengths of 280320 nm, followed by the reaction of O(1D) with water vapour. Therefore the abundance of OH in the troposphere is sensitive both to the amount of incoming UV radiation and to the water vapour abundance. Tropospheric OH also reacts with other atmospheric compounds such as carbon monoxide (CO) and non-methane volatile organic compounds (NMVOCs), and is recycled in the presence of nitrogen oxides (NOX). Thus, changes in the emissions of NOX, CO and NMVOCs also affect the OH abundance (Lelieveld et al., 2002).
Inter-annual variations in both the sources and sinks of CH4 occurred in the early 1990s. In part, these can be related to the Pinatubo eruption, which caused a decrease in tropospheric temperatures of about 0.5 C globally in the two following years (McCormick et al., 1995). The global cooling most likely resulted in a decrease in the CH4 emission rates from wetlands. Methane emissions from wetlands might have also been inhibited by the deposition of volcanic sulfur from the eruption (Gauci et al., 2008).
The removal of CH4 by OH is a temperature-dependent reaction which was also affected by the temperature decrease after the eruption. OH production would have further responded to the decrease in water vapour associated with the temperature reduction (Soden et al., 2002). The climate anomaly after Pinatubo might have affected natural emissions of other species as well, such as NMVOC emissions from vegetation, which may in turn have caused changes in OH and thus the CH4 sink. Changes in UV radiation due to stratospheric aerosols and stratospheric O3 anomalies would have also affected the removal of CH4 by OH in the post-Pinatubo years (Dlugokencky et al., 1996; Bekki et al., 1994; Bnda et al., 2014, from here on referred to as B14). Stratospheric O3 variations in this period were caused by heterogeneous removal on volcanic aerosol particles, by atmospheric transport changes due to the eruption, and by other factors
a et al.: Explaining methane variability after the Pinatubo eruption 197
of about 30 Tg yr1 over the period 1991 to 1993 compared to 1990. To close the CH4 budget, their anthropogenic and wetland CH4 emissions also show pronounced decreases.When using both CH4 and methyl chloroform to constrain OH, Pison et al. (2013) found a more moderate OH variability. In W04, OH is parameterised based on chemical and meteorological parameters with coefcients determined from three-dimensional model simulations. W04 nd an increase of about 5 % in OH from 1991 to 1993 due to stratospheric O3 depletion. Stratospheric aerosols from the eruption are, however, not included as a parameter in their calculation. Our previous column model study (Bnda et al., 2013) supports a decrease in OH in the rst year after the eruption due to aerosols, temperature and water vapour changes, and a subsequent decrease due to stratospheric O3 depletion. In that study we found that the stratospheric O3 depletion together with changes in anthropogenic emissions could explain a 12 ppb yr1 decrease in 1993 compared to 1991. However, the one-dimensional model was unable to capture the strong
CH4 decrease in 1992.
In the present study, we quantify the combined impact of the drivers of CH4 variability described above in the early 1990s using the TM5 global chemistry and transport model, and we identify the potential gaps in our understanding of the CH4 budget. For the rst time, all the known processes that could have signicantly contributed to the CH4 variations in the early 1990s are included in a modelling study.
The paper is structured as follows. A description of the TM5 model is presented in Sect. 2.1 and of the simulation set-up in Sect. 2.2. Section 2.3 describes the drivers that have been considered to explain the observed CH4 variability, and
Sect 3.1 presents their impact on the simulated CH4 budget.
The unexplained CH4 variability is shown in Sect. 3.2, and possible reasons for mismatch between the model and observations are discussed in Sect. 4. Conclusions are drawn in Sect. 5.
2 Method
2.1 Model description
The TM5 global chemistry and transport model (Huijnen et al., 2010; Williams et al., 2012; van Noije et al., 2014) was used to simulate the chemical composition of the atmosphere during the period 1990 to 1995. The model was run on 60 hybrid sigma-pressure vertical levels and at a horizontal resolution of 3 2 (longitude latitude) globally,
except for the polar region, where a reduced grid was used for advection in the zonal direction. The model was driven by meteorological elds from the ECMWF ERA-Interim re-analysis (Dee et al., 2011). The gas-phase chemistry scheme is based on a modied version of the Carbon Bond Mechanism 4 (Houweling et al., 1998). Photolysis frequencies are calculated by the on-line parameterisation scheme based on
N. Bnd
Williams et al. (2006). To account for missing O3 chemistry in the stratosphere, O3 is nudged to the multi-sensor reanalysis data (MSR; van der A et al., 2010) above the 45 hPa level in the tropics (30 S30 N) and above the 90 hPa level in the extra-tropics. CH4 is also nudged above these pressure levels to climatological values from Groo and Russell III (2005) to compensate for possible errors in stratospheric chemistry, specically the loss of CH4 by OH, Cl and O(1D).
The performance of TM5 in simulating atmospheric chemistry is presented in Huijnen et al. (2010) and van Noije et al. (2014). The conclusions of these studies remain valid with the current set-up.
The M7 modal scheme (Vignati et al., 2004; Aan de Brugh et al., 2011) is used to simulate aerosol microphysics. M7 simulates the following aerosol types: sulfate (SO4), black carbon (BC), organic carbon (OC), sea salt and mineral dust.In addition, TM5 calculates ammonium and nitrate partitioning between gas phase and aerosols using the Equilibrium Simplied Aerosol Model (EQSAM; Metzger et al., 2002), and accounts for methyl sulfonic acid (MSA). The aerosols are coupled to the radiative transfer scheme that is used to calculate photolysis frequencies. Some adjustments to the M7 default set-up have been applied in order to realistically simulate stratospheric volcanic aerosols, as described in B14.This set-up has been used in the aforementioned paper to model SO2 and sulfate aerosols from the Pinatubo eruption, and to infer their effect on photolysis frequencies.
Here we use the new massively parallel model version TM5-mp. The new MPI parallelisation does not inuence the model results, but brings a signicant speedup of the model runs by allowing us to use more computing cores more efciently. A few additional updates have been incorporated into the model compared to the version used in B14. Convective uxes used for tracer transport are now read from the ERA-Interim input data, instead of calculating them with the Tiedtke scheme (Tiedtke, 1989). Heterogeneous removal of N2O5 on aerosol and cloud particles was updated according to Huijnen et al. (2014). The CH4 surface nudging timescale has also been adjusted. We now use a nudging timescale of 10 days instead of 3 days in B14. By comparing the budget terms and concentrations of simulations with the different nudging timescales, we nd that both nudging timescales give a similar performance in simulating the CH4 concentrations and inter-annual variability (IAV). Although similar global nudging values are found for the two simulations, the local nudging amount is 2 to 3 times larger using a 3-day nudging. Therefore we nd that a 10-day nudging timescale is more appropriate for inferring CH4 budget mismatches based on monthly observations, allowing for synoptic-scale variability in CH4 concentrations (Dentener et al., 2003). Finally, some of the emission inventories used here differ from the ones used in B14, and are described below. We veried that the model updates did not cause any signicant differences in modelled SO2 and stratospheric aerosols compared to the results presented in B14.
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
a et al.: Explaining methane variability after the Pinatubo eruption
Table 1. Annual mean CH4 sources and sinks, and their inter-annual variability (IAV). All values are in Tg yr1.
Category Inventory 1990 1991 1992 1993 1994 1995 Mean IAV or simulation
Natural wetlands ORCHIDEE 273.9 276.6 251.7 262.1 272.1 260.8 266.2 11.9 LPJ 167.4 163.9 161.9 165.6 167.7 169.3 166.1 2.6
Natural other HYMN 43.0 43.0 43.0 43.0 43.0 43.0 43.0
Anthropogenic EDGAR 256.7 256.2 257.2 256.4 257.3 260.0 257.3 1.2
Biomass burning RETRO 19.5 21.2 23.0 17.8 20.4 16.5 19.7 3.0
Trop CH4+ OH Base1 489.8 489.4 483.7 505.4 512.0 517.8 499.7 8.1
Base2 487.9 487.5 482.8 504.6 511.2 517.1 498.5 7.9
Stratospheric sink Base1 35.1 39.2 39.8 40.3 42.0 41.6 39.7 1.5 Base2 33.0 37.2 38.8 39.4 41.3 40.8 38.4 1.8
Soil sink Base1 29.2 30.4 26.3 28.5 26.4 25.5 27.7 2.1 Base2 27.1 26.8 27.0 27.2 26.9 27.5 27.1 0.3
Nudging Base1 11.3 15.1 18.9 3.5 23.1 11.9 1.1 12.4
Base2 88.8 89.3 68.2 98.5 120.0 100.1 94.1 18.7
Table 1 lists the CH4 source and sink categories, the inventories used, and their yearly global mean and variability over the period 1990 to 1995. Two inventories for CH4 emissions from wetlands are employed to investigate the uncertainty in these emissions. We use the ORCHIDEE (Organising Carbon and Hydrology In Dynamic Ecosystems land surface model) emissions from the WETCHIMP intercomparison project (Wania et al., 2013) and LPJ (Lund-Potsdam-Jena dynamic global model) emissions from the Hydrogen, Methane and Nitrous oxide (HYMN) project (Spahni et al., 2011).Other natural CH4 monthly emissions from wild animals and termites were also used as compiled in HYMN. CH4 soil removal rates from the LPJ inventory were applied. Anthropogenic emissions of CH4, CO, NOX, NMVOC, SO2 and ammonia (NH3) were taken from the EDGAR4.2 inventory, except for the transport sector, for which EDGAR4.1 was used. Decadal anthropogenic emissions of BC and OC were taken from the historical AR5 emission inventory (Lamar-que et al., 2010). Biomass burning emissions of CH4, CO,
NMVOC, NOX, SO2 and NH3 are used from the RETRO inventory (Schultz et al., 2008). Production of NOX by lightning and emissions of dimethyl sulde (DMS), sea salt and dust are calculated online (Huijnen et al., 2010; Vignati et al., 2010; van Noije et al., 2014). Monthly biogenic emissions of CO, NH3 and NMVOC were taken from the MEGAN inventory (Sindelarova et al., 2014). Climatological emissions from the MACC inventory are used for continuous volcanic SO2 (Andres and Kasgnoc, 1998) and biogenic NOX emissions.
198 N. Bnd
2.2 Simulation set-up
We designed a series of simulations with the aim of quantifying the impact of specic processes on the tropospheric CH4 concentrations. The difference between two simulations, one including and one excluding a specic driver, is used to quantify the effect of that driver. For instance, we calculate the effect of IAV in meteorology on tropospheric CH4 as the difference of a simulation that includes varying meteorological elds and a simulation that xes meteorological elds to 1990 values. Two sets of simulations were performed for seven drivers of CH4 variability for the period 1990 to 1995, as outlined in Table 2.
As described in B14, we use a two-step set-up to simulate realistic CH4 concentrations. In a rst simulation of the
Base1 and Base2 scenarios, in which all drivers of CH4 variability are included, we nudge CH4 mixing ratios in the lowest 2 km of the model towards measured mixing ratios from background sites. This allows us to account for uncertainties in sources and sinks of CH4 and to avoid a long-term drift of mixing ratios compared to observations. This set-up is particularly important for the second simulation set using the lower LPJ wetland emissions, where nearly 100 Tg yr1 additional source or reduced sink is needed to close the CH4 budget. The nudging is performed as follows. The north south gradient in the CH4 mixing ratio is reconstructed by interpolating the measured CH4 monthly mean mixing ratios at the background stations South Pole, Cape Grim, Mauna Loa, Niwot Ridge, and Barrow. In this way a realistic zonal mean CH4 distribution is obtained from the observations, which is then compared to the modelled concentrations at the dateline. The concentrations in the whole zonal band are
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
a et al.: Explaining methane variability after the Pinatubo eruption 199
then adjusted based on this comparison. The amount of CH4 needed to compensate for the difference between the model and observation-based northsouth gradient in CH4 is stored on a monthly basis for each 10 latitude band. In the second step, CH4 is no longer constrained by observations, but instead the nudging amount calculated in the rst step is applied as an emission in all scenarios. In B14 we used this nudging to scale the CH4 emissions in each 10 latitude band. In these simulations, we observed slight time shifts in the derived CH4 growth rates in the second step compared to the rst step. To obtain an almost perfect match between the simulated CH4 concentrations of the two steps, in the present study we apply the nudging homogeneously over the latitude band and in the lowest 2 km from the surface. This two-step set-up enables us to model realistic CH4 concentrations, and at the same time allowing them to respond to changes in emissions or photochemistry. The nudging term that is needed to force the Base1 and Base2 simulations to background atmospheric observations indicates in which latitude bands sources and sinks are unbalanced when all drivers of CH4 variability are included. These mismatches will be further analysed in Sect. 3.2. Validation results for this two-step set-up are shown in Appendix A.
The rst simulation set (Set I in Table 2) uses the ORCHIDEE inventory for natural CH4 emissions from wet-lands, while the second set of simulations uses the LPJ inventory. The Base1 simulation from the rst simulation set accounts for all drivers of CH4 IAV in the model, including modelled stratospheric sulfur, varying amounts of stratospheric O3, ERA-Interim elds for temperature and humidity, and emission variations of CH4 and other species. Emissions of SO2 into the stratosphere by the Pinatubo eruption of 18.5 Tg SO2 were considered, and the SO2 subsequently reacts with OH to form aerosols. In the other simulations, we x different variables in the model to investigate their effect on CH4 concentrations. In all simulations except for the
Base1 simulation, we removed the effects of Pinatubo SO2 and sulfate, because using a different meteorology or O3 column might lead to different SO2 and aerosol elds. By using this set-up, we minimise the nonlinearities between drivers of CH4 variability. In FixMeteo1 and FixOzone1, we additionally used the meteorology and, respectively, the stratospheric O3 of the year 1990 for the whole length of the simulation.
In the FixWetl1 simulation we used constant CH4 emissions from wetlands and CH4 soil uptake rates from the year 1990.Simulation FixEmis1 used anthropogenic and biomass burning emissions of CH4 and other species from the year 1990.
Biogenic emissions of CO, NMVOC and NOX are also xed to 1990 in FixEmis. Finally, to infer the combined effect of all seven drivers of CH4 variability, simulation FixAll1 is performed, where all drivers of variability are xed to 1990 values. This allows us to quantify the possible effect of nonlinearities, as discussed in Appendix B.
To test the impact of using a different inventory for CH4 emissions from wetlands, a second set of simulations (Set II
in Table 2) is performed using LPJ emissions. The second set of simulations consists of simulations Base2, NoPinS2 and FixWetl2, which are equivalent to Base1, NoPinS1 and FixWetl1, respectively, of the rst set. The Base2 simulation is similar to the PinS simulation in B14, extended for three more years. The LPJ emissions and soil uptake rates from the year 1990 are used in FixWetl2 throughout the simulation.
2.3 Drivers of CH4 variability
In this section we present the drivers of CH4 variability in our model. The factors that affect the CH4 sinks are stratospheric sulfur, stratospheric O3, changes in meteorology, and emissions of CO and NMVOC. The CH4 emissions for different sectors are also presented. We show model results for sulfate aerosol, while for other drivers of CH4 variability we give a quantitative description of the input data. Their impact on the CH4 concentration is analysed in Sect. 3.1.
The main drivers of CH4 variability in our model for the period 1990 to 1995 are shown in Fig. 1. All variables have been averaged with a 12-month running mean to remove the seasonal cycle. The globally averaged stratospheric aerosol optical depth (AOD) is shown in Fig. 1a. The simulated AOD values were validated against measurements in B14.A global maximum AOD of 0.16 is simulated in early 1992.The aerosols remained in the stratosphere for a few years, with the AOD returning to pre-eruption levels towards the end of 1994.
Figure 1b presents the average stratospheric O3 above 150 hPa between 60 S and 60 N from the MSR data set (van der A et al., 2010). High-latitude O3 anomalies are not taken into account here because the CH4 is mainly oxidised in the tropics and at mid-latitudes during summer. A number of processes caused variations in stratospheric O3 in the period 1990 to 1995. The 11-year cycle in solar radiation determined an overall decrease of 2.5 % (6 Dobson units DU) in the O3 column over this period. The quasi-biennial oscillation (QBO) in stratospheric winds caused an additional IAV of about 1 % (2 DU). In addition to these natural cycles, enhanced O3 depletion occurred in 1992 to 1994 due to dynamical changes and heterogeneous chemistry associated with the presence of Pinatubo sulfate particles in the stratosphere (Telford et al., 2009; Aquila et al., 2013). This caused a 3.5 % (8 DU) decrease in stratospheric O3 from 1991 to early 1993, with the largest perturbations observed at northern mid-latitudes. A smaller decrease of about 4 DU is found in the tropics (Shepherd et al., 2014). The effect of observed O3 variations on CH4 concentrations is investigated in Sect.
3.1. We do not separate in this study the effect of Pinatubo on stratospheric O3 from other factors of O3 variability.
A global-scale decrease in surface temperature was registered in 1992 caused by the decrease in surface shortwave radiation due to the volcanic aerosols. The effect of the eruption was partly counteracted by the 19921993 El Nio (Mc-Cormick et al., 1995). In our ERA-Interim input data we nd
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
Table 2. Set-up of the simulations, including the CH4 wetland inventory used and the drivers of CH4 included in each simulation. The crosses indicate that the variability of a certain driver is included in the simulation. Otherwise the driver is not included (in the case of Pinatubo SO2 and aerosol) or 1990 values are used throughout the simulation.
Drivers of CH4 variability includedSimulation Inventory for Pinatubo SO2 Stratospheric Meteorology CH4 Natural Anthropogenic Biomass burning
name CH4 emissions and aerosol O3 wetland emissions of CO, emissions of CH4 and emissions of CH4 and
from wetlands emissions NMVOC and NH3 other compounds other compoundsSet I
Base1 ORCHIDEE X X X X X X X NoPinS1 ORCHIDEE X X X X X X FixOzone1 ORCHIDEE X X X X X FixMeteo1 ORCHIDEE X X X X X FixWetl1 ORCHIDEE X X X X X FixEmis1 ORCHIDEE X X XFixAll1 ORCHIDEE
Set II
Base2 LPJ X X X X X X X NoPinS2 LPJ X X X X X X FixWetl2 LPJ X X X X X
200 N. Bnd
a global mean temperature decrease of 0.5 C in the free troposphere and of 0.3 C at the surface between 1990 and 1992, followed by a temperature recovery in 1994 (Fig. 1c). The temperature decrease was associated with a 3 % decrease in tropospheric water vapour.
Other meteorological changes in this period, such as dynamics or cloudiness, might have also affected the stratospheretroposphere exchange of CH4 and the CH4 removal by OH. Although these other meteorological parameters are not presented in this section, their potential effects on CH4 are included when comparing the NoPinS and FixMeteo simulations.
Variability in emissions of chemical species might have determined changes in OH concentrations. Global CO and NOX emissions have a similar evolution over the period 1990 to 1996, mainly driven by variations in biomass burning (Fig. 1d). A large biomass burning event in 1992 leads to an increase of 100 Tg CO yr1 in CO emissions and of1.5 Tg NOX yr1 in 1992, and a decrease of similar magnitude 1 year afterwards. The NOX emissions remain relatively constant throughout the rest of the period, while CO emissions decrease by about 50 Tg between 1990 and 1996. Biogenic emissions of NMVOC can also inuence CH4 concentrations indirectly through their reaction with OH. Isoprene, the NMVOC species with the largest emission rate in the atmosphere, is mainly produced by plants. Biogenic isoprene emissions are sensitive to changes in temperature, precipitation and radiation, and were likely affected by the Pinatubo eruption (Telford et al., 2010; Wilton et al., 2011). Global emissions of isoprene, shown in Fig. 1e, given by the MEGAN inventory are 640660 Tg yr1 during 1990
and 1991. Over the course of 1992, the isoprene emissions decrease by about 70 Tg yr1, remaining at about 580
600 Tg yr1 throughout the simulation period.
Changes in the main CH4 emission categories are presented in Fig. 1fh. Inter-annual variability in total CH4 emissions is dominated by emissions from wetlands and biomass burning. Both the LPJ and ORCHIDEE inventories show decreased CH4 emissions from wetlands in 19911993.
However, the exact timing and magnitude vary considerably between the two inventories, as well as the global emission amount and IAV. The ORCHIDEE inventory nds on average 266 Tg yr1 CH4 global emissions from wetlands, with inter-annual variability of 11.9 Tg yr1 (see Table 1). The largest anomaly in this period is found after the Pinatubo eruption, with a decrease of 30 Tg yr1 between the time of the eruption and 1 year afterwards. The LPJ inventory shows global wetland emissions of 166 Tg yr1 with an IAV of 2.6 Tg yr1. A decrease of 8 Tg yr1 in CH4 emissions from wetlands is found by LPJ during the year of the eruption, and a recovery starting from early 1992. Wetland extent has been shown to be important for simulating the IAV in wetland emissions, and might be a reason for the large differences between the two models (Ringeval et al., 2010; Spahni et al., 2011). ORCHIDEE simulates wetland extent interactively through the coupling to the TOPMODEL hydrology, while LPJ uses xed wetland extent. The ORCHIDEE global emissions have been shown to be larger than those of other models from the WETCHIMP intercomparison study, having a high sensitivity to changes in CO2, temperature and precipitation (Melton et al., 2013).
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption 201
data used in the Reg-FIRM regional re model (Schultz et al., 2008). Global anthropogenic emissions were quite stable at 256 to 261 Tg yr1 during 1990 to 1995, according our input data based on EDGAR 4.2 and REAS 1 emission inventories, with an IAV of 1.2 Tg yr1 and an increase towards the end of the period. More signicant changes occurred at a regional level, with an overall increase in emissions in South-East Asia, and a decrease in Europe, North America and the FSU in this time period. Note, however, that inverse modelling studies report signicantly higher total anthropogenic CH4 emissions for this time period than the bottom-up inventories, exceeding 350 Tg yr1 (Kirschke et al., 2013; Pison et al., 2013).
3 Results
3.1 Explained CH4 variability
By taking differences between the budget terms of the simulations in Table 2, we are able to infer the effect of each driver of CH4 variability described above. Figure 2 shows zonally averaged differences in CH4 sources and sinks between the different simulations, vertically integrated over the tropo-sphere. These differences are integrated up to a tropopause level dened here as a function of latitude, as recommended in Lawrence et al. (2001).
Absorption of UV radiation by volcanic SO2 and UV scattering by sulfate aerosol are both included in the Base1 simulation, and lead to a decrease in UV radiation entering the troposphere. As shown in B14, the effect of stratospheric aerosols on global CH4 is dominating and longer lived in comparison to the effect of SO2 absorption. A decrease in the CH4 sink is depicted in Fig. 2a due to stratospheric sulfur, calculated as the difference between the Base1 and NoPinS1 simulations. The impact of stratospheric sulfur is strongest in the months after the eruption in the tropical region, with decreases in the zonal mean CH4 sink of 2 to 5 mg m2 month1. Starting from 1993, the difference in the CH4 sink due to stratospheric sulfur decreases below 1 mg m2 month1 globally.
Stratospheric O3 decreased in the tropical region in 1991 to 1995 compared to 1990. A slight increase in stratospheric O3 was observed in 1991 in the southern mid-latitudes because of an increase in the strength of the BrewerDobson circulation due to stratospheric heating by Pinatubo aerosols (Aquila et al., 2013). Figure 2b shows that these stratospheric O3 changes led to variations in the OH sink of CH4 between
5 and +10 mg m2 month1 in the period 1991 to 1996.
Decreases in the CH4 sink of up to 5 mg m2 month1 are modelled in 1991 in the extra-tropics, compensated by increases of similar magnitude in the equatorial region. From 1992 to 1996, reduced stratospheric O3 levels caused increases in the CH4 sink in the equatorial band, the northern tropics and part of the northern mid-latitudes. An increase in
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
Figure 1. The drivers of CH4 IAV in the early 1990s considered in this study. The black dashed line denotes the timing of the Pinatubo eruption.
We used in our simulations the biomass burning emissions of CH4 reported by the RETRO inventory, which amount to a global mean of 19.7 Tg yr1 over this period, with an IAV of 2.6 Tg yr1 and higher emissions in the years 1991, 1992 and 1994. The IAV in RETRO is determined from national re reports, and from climate, soil moisture and carbon pool
202 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
(a)
(e)
(b)
(f)
(c)
(g)
(d)
Figure 2. Zonal mean differences in the CH4 budget terms caused by the different drivers of CH4 variability, vertically integrated over the troposphere. Changes in the CH4 sink by the reaction with OH are shown for the effects of (a) stratospheric sulfur, (b) stratospheric ozone,(c) meteorology and (d) emissions of CO, NOX and NMVOC. CH4 emission changes are shown for wetlands from (e) ORCHIDEE and(f) LPJ, and (g) for biomass burning and anthropogenic sectors. The years on the x axis in this and later gures refer to the start of the year.
the CH4 sink of more than 5 mg m2 month1 is modelled in the northern tropics in the summers of 1993, 1994 and 1995.
The decrease in temperature and water vapour following the eruption led to a decrease in both OH production and the rate of reaction between OH and CH4. Our model results show that variations in meteorology caused decreases in the CH4 sink of 2 to 10 mg m2 month1 in the northern tropical region during September 1991 to March 1993, and in the northern mid-latitudes during the summers of 1992 and 1993 (Fig. 2c).
Variations in emissions of other species indirectly affect CH4 concentrations through the CH4 sink. On the one hand,
CO and NMVOC emissions decrease OH concentrations because of their reaction with OH. Recycling of OH, on the other hand, increases due to NOX emissions. Anthropogenic activity and biomass burning events lead to emissions of all these compounds. In addition, the difference between NoPinS and FixEmis also includes changes in biogenic emissions of CO and NMVOC that are sensitive to climate variations (Sindelarova et al., 2014). The overall effect on OH is determined by the relative increases in total emissions of
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
a et al.: Explaining methane variability after the Pinatubo eruption 203
CO and NMVOC compared to total NOX emissions (Dalsren and Isaksen, 2006). The OH variability due to changes in emissions is shown in Fig. 2d. Particularly large decreases in emissions of CO from biomass burning and in biogenic NMVOC emissions in our input data occurred between 1992 and 1993, leading to an increase of 3 to 5 mg m2 month1 in the CH4 sink.
Figure 2e and f show differences in wetland emission strength with respect to the year 1990 for the ORCHIDEE and LPJ emission inventories. These emission differences are often larger that those found for the CH4 sink, but they are of shorter duration and more localised. The most striking difference between the CH4 emission variations in the LPJ and in the ORCHIDEE inventories is their magnitude. While the magnitudes of LPJ differences compared to 1990 have values mostly between 10 and 10 mg m2 month1, the
ORCHIDEE differences often exceed 20 mg m2 month1.
Both inventories show a decrease in emissions in the tropics and in the Northern Hemisphere in late 1991 and early 1992.The strength and the duration of the decrease differs between the two inventories. In the equatorial region, ORCHIDEE gives decreases that often surpass 20 mg m2 month1 from 1992 to 1995. LPJ reports decreased emissions in the equatorial region in the second half of 1991 and the rst half of 1992. This period is followed by alternating short periods of increased and decreased emissions near the Equator until 1995. Both inventories give some increases in emissions in the subtropics throughout the period and a strong increase in emissions throughout the tropics in the second half of 1995 and beginning of 1996. In the northern mid-latitudes and high latitudes, decreases of 2 to 5 mg m2 month1 are maintained from the second half of 1991 to mid-1993 in LPJ. In ORCHIDEE, CH4 emissions decrease by up to 20 mg m2 month1 in the rst half of 1992, but are compensated by increases of similar magnitude in the second half of 1992. Increased emissions are reported by ORCHIDEE in this region every summer from 1993 to 1995.
Variations in emissions other than wetlands also cause signicant changes in the CH4 budget. The CH4 emissions are affected by variations in biomass burning and anthropogenic activity (Fig. 2g). Anthropogenic emissions in the northern mid-latitudes show a gradual decrease of up to 20 mg m2 month1 from 1990 to 1996. This decrease is compensated by an increase in anthropogenic emissions in the tropical region, with values reaching 5 to 10 mg m2 month1 between 15 and 30 N. Enhanced biomass burning emissions are also found close to the Equator in the autumn of the years 1991, 1992, and 1994.
The modelled growth rate variations caused by the changes in CH4 sources and sinks described above are depicted in Fig. 3. A 12-month running mean was applied to all the growth rate variations, to remove seasonal effects. The black line shows the cumulative effect of these processes on the CH4 growth rate, representing the combined effect of all considered drivers of variability on the global CH4 growth
N. Bnd
rate, which we will refer to as the explained CH4 growth rate. We assume here additivity between the changes in CH4 growth rate caused by the different drivers. The effect of nonlinearities is discussed in Appendix B. Figure 3a and b show the two cases in which the ORCHIDEE and LPJ inventory, respectively, are used to represent the effect of IAV on wet-land CH4 emissions. In early 1991 the CH4 growth rate remains within 2 ppb yr1 of that in 1990. Values of more than 10 ppb yr1 are found in 19901991 (see Fig. A1) due to relatively high emissions from all source categories of CH4.The relatively large ozone column values in these years, together with large emissions of CO and NMVOC, lead to a reduced CH4 sink in 1990 and 1991. During 1991 the explained CH4 growth rate remains stable or shows a moderate increase depending on whether the ORCHIDEE or, respectively, LPJ inventory is used. The explained CH4 growth rate decreases afterwards by 910 ppb yr1 in both cases, partly due to a decrease in wetland emissions. While CH4 emissions from wetlands gradually recover, a combination of other processes leads to a continued decrease in CH4 growth rate until 1993. The processes contributing to the decrease in the explained CH4 growth rate in the second half of 1992 and early 1993 are stratospheric O3 depletion, a recovery of stratospheric aerosols towards background levels, and changes in emissions other than CH4 emissions from wetlands. The former effect includes a decrease in CH4 emissions from biomass burning of 45 Tg yr1 between 1992 and 1993, and a 2 Tg yr1 decrease in anthropogenic emissions. This would lead to a combined decrease of 67 Tg yr1, or 2 ppb yr1. A decrease in isoprene emissions of 60 Tg yr1 and a decrease of 50 Tg yr1 in CO emissions also occur over this period.
They cause an increase in the CH4 removal and lead to a decrease of 6 ppb yr1 in FixEmis1 compared to NoPinS1. The overall effect remains stable at 6 to 7 ppb yr1 from mid
1993 to spring 1994.
Although the evolution of the explained CH4 growth rate shows similar features when using the two wetland emission inventories, some signicant differences occur. First, the more pronounced decrease in CH4 emissions in ORCHIDEE in 1991 compensates the concurrent decrease in the CH4 sink due to stratospheric aerosols and meteorological changes, leading to a stable CH4 growth rate during 1991. The decrease in LPJ emissions in 1991 is much smaller, resulting in a 3.5 ppb yr1 increase in CH4 growth rate in the second half of 1991. The decrease in the explained CH4 growth rate that follows in 1992 starts about half a year later when using LPJ emissions than when using ORCHIDEE. The 8 ppb yr1 decrease in the growth rate obtained due to ORCHIDEE in late 1991 and early 1992 causes an earlier decrease in the overall growth rate. The minimum CH4 growth rate also oc-curs about 3 months earlier with ORCHIDEE than with LPJ.Finally, an additional 3 ppb yr1 growth rate decrease occurs during 1994 with ORCHIDEE, while LPJ gives a stable CH4 growth rate between mid-1993 and mid-1995.
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
204 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
(a)
(b)
Figure 3. Global CH4 growth rate variations with respect to the year 1990 induced by the different drivers, using (a) ORCHIDEE and (b)
LPJ to represent variability in CH4 emissions from wetlands. Simulations from Set I are used in both plots to infer the effect of stratospheric sulfur, stratospheric ozone, meteorology and emissions other than wetlands. Simulations from Set II are used only to infer the effect of LPJ emission variability. The overall variability (All) is calculated as the sum of the individual variations.
Background5, is taken from the CH4 background data used for nudging the model, based on measurements from ve stations (see Sect. 2.2). The other CH4 growth rate curves in Fig. 4 are the same as the black lines in Fig. 3, representing the explained growth rate differences with respect to 1990. In order to obtain the variability with respect to the year 1990 from the observations, the 1990 CH4 growth rates have been subtracted from both observation time series (see absolute growth rate values in Fig. A1). A running mean of 12 months was further applied to remove seasonal effects.The explained CH4 growth rate variability shows differences of 1.0 to 2.5 ppb yr1 in 1991 with respect to 1990,
which falls in between the two observational estimates of 3 ppb yr1 for GLOBALVIEW-CH4, and 1.5 ppb yr1 for
Background5. The model gives an increase of 3.5 ppb yr1 over the course of 1991 when using LPJ, while using OR
CHIDEE, we nd a decrease of 1 ppb yr1. The observations from GLOBALVIEW show an increase in the CH4 growth rate in the rst half of 1991 and a decrease in the second half, while the Background5 growth rate decreases throughout the year 1991. A continued decrease in the CH4 growth rate is found in observations in 1992, reaching 10 ppb yr1
in autumn 1992. The explained growth rate also decreases compared to 1990 when using ORCHIDEE emissions, reach-
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
Figure 4. The explained CH4 growth rate variability in the model using two wetland CH4 emission inventories, and global deseasonalised CH4 growth rate differences with respect to 1990 from the
GLOBALVIEW observations, and from observations at ve background stations.
Our explained CH4 growth rates are compared to observations in Fig. 4. Two estimates for the observed global mean growth rate are shown. The rst estimate is calculated from the NOAA GLOBALVIEW-CH4 (2009) marine boundary layer zonal mean CH4 concentrations. The second estimate,
N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption 205
Figure 5. Global deseasonalised CH4 nudging term for the Base1 (left y axis) and Base2 (right y axis) simulations, using CH4 wet-land emissions from ORCHIDEE and LPJ, respectively.
ing 8 ppb yr1 about half a year later than the observa
tions. With LPJ emissions, the decrease in the CH4 growth rate occurs between spring 1992 and summer 1993, reaching 6 ppb yr1. The observations show a further recovery
at the end of 1992 and the rst half of 1993 to 5 ppb yr1,
remaining relatively constant for the rest of the period. The recovery is not captured in the model, irrespective of the wet-land emissions used. The explained growth rate in the model remains stable in the second half of 1993 and rst half of 1994 and, in the case of ORCHIDEE emissions, decreases again in the second half of 1994.
3.2 Unexplained CH4 variability
As explained in Sect. 2.2, CH4 observations are used in a rst step to quantify the mismatch between the model and observations. This nudging amount is analysed further to better understand the possible reasons for this mismatch.
The global deseasonalised nudging amounts for Base1 and Base2 are shown in Fig. 5. Due to the difference of about 100 Tg yr1 in the global emissions from the two wetland inventories, the global nudging amount also shows an offset of similar magnitude. The global nudging over the period 1990 to 1995 is close to 0 with ORCHIDEE, and about 100 Tg yr1 with LPJ. However, there is quite some variability for both simulations within a range of about 50 Tg yr1.
In the rst period of the simulation, the nudging term for the ORCHIDEE simulations varies between 0 and 20 Tg yr1,
and between 75 and 105 Tg yr1 for LPJ. The nudging increases by 2530 Tg yr1 during the rst half of 1993 for both Base1 and Base2 because the increase in the observed
CH4 growth rate in this period is not reproduced by the explained growth rate. Between the second half of 1993 and the end of 1995, the nudging term for the two simulations shows similar variations, between 5 and 30 Tg yr1 for OR
CHIDEE, and between 100 and 125 Tg yr1 for LPJ.
To better quantify the nudging term, the zonal mean nudging amounts for the two simulations as a function of time
are given in Fig. 6c and d. The zonal mean emission and sink strengths for the Base1 simulation are also shown for comparison in Fig. 6a and b. In both simulations, nudging is predominantly needed in the Northern Hemisphere, where most CH4 is emitted. This suggests that the uncertainties related to the CH4 sources dominate the ones related to the sink or to transport, which occur in both hemispheres. Positive nudging is needed in the tropics in both hemispheres with a maximum during autumn, and following the position of the inter-tropical convergence zone (ITCZ). This suggests either missing emissions from tropical wetlands or biomass burning, or an overestimated strength of the tropical OH sink.In the northern mid-latitudes, the model overestimates CH4 concentrations in the winter, and underestimates concentrations in the summer. This suggests a missing seasonality in one of the CH4 budget terms. Positive nudging amounts at northern high latitudes are needed throughout the two simulations, except for Base1 during autumn, where negative nudging is required. These might be related to underestimated anthropogenic emissions, as well as to errors in transport or vertical mixing, since both natural emissions and the CH4 sink are almost negligible at high latitudes during winter.
In Fig. 6e and f the nudging for the year 1990 was subtracted from the whole time series, and a 12-month running mean was applied. These plots thus highlight the potential missing IAV in the model compared to observations. For both simulations we nd a missing decrease in the CH4 burden in the northern tropics during 1991 to 1993. This points to either a missing decrease in both wetland emission inventories in this region or to potential uncertainties in tropical biomass burning. At northern mid-latitudes the CH4 emissions are dominated by anthropogenic emissions; therefore, the missing CH4 increase revealed in this region suggests that the decrease in anthropogenic emissions over Europe, North America and the FSU in this time period might be overestimated. When using ORCHIDEE emissions we nd an additional missing increase in the southern tropics from 1992 onwards, in particular in 1993 and 1995.
4 Discussion
We have quantied for the rst time all known major drivers of CH4 variability in the early 1990s in a global chemistry and transport model. Previous attempts to explain the CH4 evolution in this period were the inverse modelling studies of W04 and B06. The sensitivity of the CH4 growth rate to most of the drivers presented here was also investigated in our previous study (Bnda et al., 2013), using a column chemistry model. Other bottom-up studies have focused on only one of the potential causes of CH4 variability during the post-
Pinatubo period. In the following sections we compare our results to previous top-down and bottom-up studies, and discuss potential sources of uncertainty in our results.
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
206 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
Figure 6. Zonal mean CH4 (a) emission and (b) sink by reaction with OH in the Base1 simulation, zonal mean CH4 nudging term when using CH4 wetland emissions from (c) ORCHIDEE (Base1 simulation) and (d) LPJ (Base2 simulation), and (e) and (f) the deseasonalised nudging anomaly compared to 1990.
4.1 Comparison to inverse modelling studies
The CH4 inverse modelling studies of W04 and B06 included the post-Pinatubo period. The OH elds used were derived by different methods. W04 used parameterised OH elds based on chemistry model results. Changes in meteorology, O3 column, concentrations of CO and NMVOC were included in their parameterisation. However, stratospheric aerosols from the Pinatubo eruption were not included. The OH elds in B06 were determined from an inversion of methyl chloroform observations, and might be affected by uncertainties in emissions of methyl chloroform.
With regard to the either applied or inferred OH sink variations, the inversion studies for this time period are only partly in line with our simulated variations in the OH sink, including the combined effect of stratospheric aerosols, stratospheric O3 depletion and tropospheric emissions of CO and
NMVOCs. We will now discuss the comparison between our results with these earlier studies for the different time periods: (i) the sharp decrease in the CH4 growth rate in the second half of 1991 and throughout 1992, (ii) the subsequent increase in the CH4 growth rate in 1993, (iii) the moderate
growth rates over the years 1994 and 1995, and (iv) the decrease in the CH4 growth rate over the period 1990 to 1995.
19911992 W04 inferred a slight decrease in the OH sink of CH4 in 1991 compared to 1990, and an increase of about 10 Tg yr1 in 1992. In B06, a decrease of about25 Tg yr1 in the OH sink was found during 1990 to 1992. In our study we nd smaller OH variations than both studies, with a decrease in the OH sink of CH4 of about 5 Tg yr1 in 1992 compared to 1990 and 1991.
This reduction in OH is due to sulfate aerosols and atmospheric cooling, compensated by the OH increase caused by ozone depletion (see Table 1, Fig. 3).
Based on their assumed OH concentrations, W04 obtained a 2025 Tg decrease in wetland emissions in 1992 compared to 1991 and 1990. To explain the decrease in the CH4 growth rate in 1992, B06 found a 3540 Tg yr1 decrease in wetland emissions in the rst half of 1992. To compensate for their simultaneous decrease in OH sink, they found 510 Tg lower biomass burning emissions in 1992 than 1991 and 1993, and a20 Tg decrease in anthropogenic emissions, which is
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
a et al.: Explaining methane variability after the Pinatubo eruption 207
sustained for the rest of the 1990s. With ORCHIDEE we applied a 25 Tg yr1 emission decrease in 1992.
Our nudging suggests that this decrease could reach29 Tg yr1, which is similar in magnitude to the other studies.
1993 The observed CH4 growth rate partially recovers during 1993 from the strong decrease in the year 1992. In our study we capture the decrease in CH4 growth rate in 1993 rather than 1992 due to a 22 Tg increase in the OH sink of CH4 because of reduced ozone columns, biomass burning emissions and a recovery of sulfate aerosols. A partial recovery in wetland emissions is found in both inventories. This recovery might be underestimated, since an increase in the nudging term is needed to explain the observed growth rate. Both W04 and B06 found increases in the CH4 sink of about 10 Tg yr1 during 1993. Similarly to our study,
W06 simulate a continued decrease in CH4 growth rate rather than a recovery. In B06, increases in wetland and biomass burning emissions in 1993 are found to overwhelm the increase in the CH4 sink, leading to an increase in the CH4 growth rate.
19941995 The OH sink of CH4 showed variations of 3 to 5 Tg yr1 in W06 in the years 1993 to 1995, while an increase of about 15 Tg yr1 in the years 1993 to 1995 was obtained in B06. In agreement with B06, we nd a 12 Tg yr1 increase in the OH sink of CH4 between 1993 and 1995 due to the recovery of the anomaly in temperature, water vapour and stratospheric aerosols caused by the eruption.
Similar to the study of W06, we nd that the decrease in the CH4 growth rate over the period 1990 to 1995 can be entirely explained by an increase in the OH sink of CH4, rather than by changes in emissions. In B06, however, this decrease in CH4 growth rate is explained by a decrease in anthropogenic emissions.
The more recent study of Pison et al. (2013) extended the study of B06 by using both methyl chloroform and CH4 observations to constrain OH concentrations in their INVARR inversion. Smaller OH variability is found in this case compared to B06, and their derived global emission changes are more in line with W04 and with the emissions applied in our rst simulation set.
The 1.6 % IAV we nd for the CH4 loss by reaction with OH supports the conclusion of Montzka et al. (2011) that OH concentrations are buffered against atmospheric perturbations, having an IAV of about 2 %. The large OH inter-annual variations, often exceeding 10 %, previously found for the 1990s using methyl chloroform observations are not reproduced in our chemistry-transport simulations (Prinn et al., 2005; Bousquet et al., 2005).
N. Bnd
4.2 Comparison to previous bottom-up studies
In Bnda et al. (2013), we analysed the CH4 growth rate variability after the Pinatubo eruption using a column chemistry model. The timing of the minimum CH4 growth rate reported in that study is similar to the one found in this three-dimensional study, implying that the delay compared to observations is a result of uncertainties in model input rather than model set-up. However, some differences between the two studies occur in the magnitude and contribution of the different processes to the CH4 growth rate decrease in 1991 to 1993. In Bnda et al. (2013), the overall explained growth rate decrease was found to be 12 ppb yr1, while here we nd only 8 to 10 ppb yr1. A 5 ppb yr1 decrease was found due to CH4, NOX and CO anthropogenic emission changes.This is similar to the decrease of 6 ppb yr1 due to changes in non-wetland emissions obtained in this study. However, our current estimate also includes variations of emissions from biomass burning, natural emissions of CO and NMVOC.The sulfate aerosol and O3 column effects also differ by 2 to 3 ppb yr1 from the estimates presented in Bnda et al.
(2013), probably because the regional distribution of these effects could not be taken into account in the simplied column model approach. Furthermore, our previous study showed that CH4 concentrations are affected for a long time period after a perturbation is applied due to the CH4 lifetime of about 10 years. This delayed effect can be seen here for stratospheric sulfur, where a small negative difference in the CH4 growth rate is found towards the end of the simulation period (Fig. 3). The delayed effect that a perturbation in a driver of CH4 variability has on the CH4 growth rate also oc-curs in our other simulations. However, it is in general overwhelmed by the instantaneous effect of variability in the CH4 driver.
Other studies have focused on only one of the drivers of CH4 variability after the Pinatubo eruption. Bekki and Pyle (1994) found a decrease in the CH4 growth rate of 7 ppb yr1 globally due to stratospheric O3 using a two-dimensional model between spring 1991 and autumn 1992. Here we obtain a comparable estimate of 5 ppb yr1 decrease over the period 1991 to autumn 1993 due to the pronounced stratospheric O3 depletion in the tropics and northern mid-latitudes in 1993.
Using a two-dimensional chemistry and transport model, Bekki and Law (1997) investigated the effect of temperature on both chemistry and wetland emissions in 19911992.They found that the temperature decrease after Pinatubo led to a 4 ppb yr1 increase in the global CH4 growth rate between mid-1991 and mid-1992, similar to our meteorological effect of 5 ppb yr1. By applying a Q10 = 2 temperature sen
sitivity of CH4 emissions (Duneld et al., 1993), they found that CH4 emissions from wetlands would decrease the CH4 growth rate by 2 ppb yr1 in the same period. This is similar to our result using LPJ wetland emissions. The ORCHIDEE inventory gives a much larger decrease in the CH4 growth
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
208 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
by the eruption have a short lifetime of a few days (Guo et al., 2004; Niemeier et al., 2009), and were found to have a negligible effect on global radiation. Changes in water vapour are included through ERA-Interim reanalysis, and might contain some uncertainties (Dessler et al., 2014). Sulfur deposition has also been proposed to affect CH4 emissions from wetlands (Gauci and Chapman, 2006). This effect is not included in our input data. In Bnda et al. (2013) we made a rough estimate of this effect, and found it to be of the order of 1 Tg CH4 yr1 for the Pinatubo eruption.
The zonal mean nudging term in our model points to either underestimated emissions or overestimated sinks in the tropics and during summer at mid-latitudes. An overestimated source is identied at mid-latitude during winter. Uncertainties in the OH sink of CH4 might relate to uncertainties in the chemistry. CO concentrations in TM5 are underestimated at the northern mid-latitudes (Huijnen et al., 2010; van Noije et al., 2014), which might lead to overestimated OH concentrations in this region. An overestimate in OH concentrations at northern mid-latitudes and an overestimated north south gradient in global chemistry models is also found from methyl chloroform observations in Patra et al. (2014). In the tropics, OH uncertainties might arise due to NMVOC chemistry, and are discussed in more detail below. Other uncertainties in tropical OH might relate to NOX and CO emission factors from biomass burning. CO decreases OH concentrations, while NOX increases OH recycling; therefore, the overall effect of biomass burning emissions on OH is strongly dependent on the emission factors of these species, which are uncertain in the tropical region in the early 1990s (Schultz et al., 2008).
Our results show that the decrease in CH4 growth rate observed in 1992 is reasonably well explained by the processes considered here. However, the exact timing of the minimum growth rate is captured 6 to 9 months later than in the observations. Since missing processes are estimated to have a small impact on CH4 variability in the early 1990s, the mismatch between modelled and measured CH4 concentrations can only be related to uncertainties in either input data or modelled CH4 processes. Measurement uncertainty might also contribute to the mismatch. The differences given by the two measurement-based estimates of the global CH4 growth rate in Fig. 4 show that uncertainties can be of the order of 23 ppb yr1. Our input data and chemical processes related to aerosols, O3 and meteorological effects are fairly well studied and understood. The uncertainty related to these processes is of the order of 1020 %, and cannot explain the different timing of the decrease in CH4 growth rate between the model and observations.
Larger uncertainties are related to the CH4 emissions.
The differences between the CH4 emissions from the ORCHIDEE and LPJ inventories, both in terms of magnitude and IAV (Table 1), show that there are still many unknowns in the processes governing emissions from wetlands.One of the most important differences between the two mod-
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
rate of 9 ppb yr1, which overwhelms the meteorological effect on the CH4 sink. This shows that the climate sensitivity of wetland emissions is larger in ORCHIDEE, where changes in wetland extent are taken into account.
Stratospheric aerosols were found to enhance the Brewer Dobson circulation after the Pinatubo eruption (Aquila et al., 2013). This change in the dynamics of the atmosphere might also affect CH4 concentrations. Schaufer and Daniel (1994)
hypothesised that increased exchange between the stratosphere and troposphere might be responsible for the decrease in CH4 growth rate observed in 1992. By performing an additional simulation where only temperature and humidity were xed to 1990 values, we found that the meteorological effect is dominated by the effect of these two variables (results not shown). The global impact of changes in ERA-Interim wind elds is marginal. The wind elds in ERA-Interim have some uncertainty for the rst weeks after the eruption related to the fact that Pinatubo aerosols are not explicitly accounted for. However, the longer-term effect on temperature and the corresponding dynamical effect are included in ERA-Interim through the assimilation of satellite radiances.
Dlugokencky et al. (1994) and Law and Nisbet (1996) hypothesised that the emission decline in the FSU could have had a signicant contribution to the decrease in CH4 concentrations during 1992, because the decrease is primarily found in the Northern Hemisphere. However, our results indicate a missing CH4 burden decrease in 1992 which originates in the northern tropics (Fig. 6). Furthermore, a missing increase in CH4 concentrations is found in the Northern Hemisphere extratropics, pointing to a potential missing source in this region rather than reduced emissions due to gas leak xes. Furthermore, the overall decrease in growth rate between 1990 and 1996 is captured by our model, and can be attributed to stratospheric O3 decrease over this period, and decreases in biomass burning and biogenic emissions of NMVOC and CO. We acknowledge, however, that the nudging procedure used here introduces some uncertainty in providing the location of missing emissions. The procedure attributes the sourcesink mismatch at the dateline to sources or sinks in the same 10 latitude band. Potential sub-monthly transport of emissions from other latitudes is not taken into account.
To further constrain the sources of modelmeasurement mismatch, an inverse modelling study should be performed to estimate the variability of the CH4 sources using modelled
OH variability.
4.3 Potential sources of uncertainty
In this study all known major drivers of CH4 variability have been included. We estimate that potential missing processes had a minor effect on CH4 concentrations, and would therefore not signicantly affect our results. Such processes are the radiative effects of ash and water vapour injected into the stratosphere by the eruption, and the effect of sulfur deposition on CH4 emissions from wetlands. Ash particles emitted
a et al.: Explaining methane variability after the Pinatubo eruption 209
els is the fact that ORCHIDEE simulates changes in wet-land extent, while LPJ uses xed wetland extent. This might be the cause of the larger IAV in ORCHIDEE, and of the higher climate sensitivity. Large differences in the response of ten wetland emission models to a temperature perturbation were also found in the WETCHIMP model intercomparison project, showing that a better understanding of wetland processes is needed (Melton et al., 2013).
Biomass burning emission uncertainties could also contribute to the mismatch between model and observations.However, given the IAV in biomass burning emissions in both the RETRO and the more recent GFED emission inventory of about 3 Tg CH4 yr1, it is unlikely that uncertainties in biomass burning CH4 emissions could be the sole reason for the mismatch. As explained above, it is also unlikely that anthropogenic emission changes due to gas leak xes within the FSU contributed to the mismatch.
We nd a signicant impact of CO and NMVOC emission changes on the CH4 removal by OH. Some uncertainty is associated with NMVOC emission changes and their effect on CH4 chemistry in the period after Pinatubo. Natural emissions of isoprene respond to both changes in climate and in solar radiation (Telford et al., 2010; Wilton et al., 2011).A decrease in surface temperatures would lead to a reduction in isoprene emissions. Concerning the effects of radiation, the increase in diffuse radiation after the eruption, leading to deeper penetration into canopies, has been shown to have overwhelmed the effect of decreased direct radiation in terms of plant growth (Mercado et al., 2009). Therefore the increase in diffuse radiation would have increased isoprene emissions. However, in this study only the effect of climate change and the effect of decrease in total shortwave radiation after the eruption are included through the MEGAN inventory. The isoprene emissions in MEGAN show a decline of50 Tg yr1 globally during 1992, likely due to the decrease in surface temperature and global shortwave radiation. The CH4 growth rate decreases by about 4 ppb yr1, or 11 Tg yr1 due to changes in emissions of other species than CH4 including isoprene. These changes are of similar magnitude as found in Telford et al. (2010), where a 40 Tg decrease in isoprene emissions between 1990 and 1992 resulted in a 5 Tg increase in CH4 removal by OH. The estimated effect of NMVOC emissions on CH4 concentrations has several sources of uncertainty. Firstly, including the effect of Pinatubo on diffuse radiation might have led to increased NMVOC emissions in 19911992, and an even stronger decrease in 1993, when the aerosols were removed from the atmosphere. Secondly, recent studies have shown that the sensitivity of OH concentrations to NMVOC is smaller than previously thought (Stone et al., 2011; Rohrer et al., 2014). The CBM4 chemistry scheme used here does not include an updated isoprene chemistry mechanism, and might exhibit too high an OH sensitivity to isoprene. Telford et al. (2010) used a chemistry scheme that includes the Mainz isoprene mechanism, a parameterisation based on the Master Chemical Mechanism
(MCM), which was also shown to misrepresent OH recycling in VOC-rich environments (Pschl et al., 2000; Stone et al., 2011). Nevertheless, it is important to take NMVOC emission changes into account for evaluating CH4 variability. Due to the potentially overestimated OH sensitivity, our calculated effect on CH4 can be seen as an upper estimate.
The latitudinal nudging term needed to correct the mismatch between modelled and observed CH4 mixing ratios and presented in Sect. 3.2 is calculated from measurements at ve remote stations. Some uncertainty exists in these terms due to possible observational uncertainty and shortcomings of the nudging procedure. An indication of the observational uncertainty is given in Fig. 4, where two observation-based estimates of the global mean CH4 growth rate variations are shown. The GLOBALVIEW data use a more complete set of stations, but might contain measurements affected by nearby emissions. Furthermore, additional processing is done to gap-ll and homogenise the station data. The two estimates are in good agreement except for the year 1991, where they differ by about 4 ppb yr1. Some uncertainty also exists in the timing and location of the missing emission variations given by the nudging term. Our nudging procedure is in general able to capture the global growth rate variations.However, because the nudging corrects the amount of CH4 in the zonal band where the mismatch occurs, this procedure does not account for sub-monthly transport between zonal bands. An inverse modelling set-up would be needed to exploit all available measurements to better resolve the sources of mismatch.
5 Conclusions
The processes responsible for CH4 variability in the early 1990s have been investigated in the TM5 global chemistry and transport model. Known drivers of CH4 variations include (i) photochemical effects of stratospheric sulfur from Pinatubo and (ii) of stratospheric O3 changes, (iii) temperature and humidity perturbations, and their effect on CH4 chemistry, (iv) variations in CH4 emissions from wetlands,(v) biomass burning and (vi) anthropogenic sources, and (vii) changes in emissions of other compounds and their effect on OH. We nd that all these processes contributed in a signi-cant way to the CH4 growth rate variations in the early 1990s.
The explained growth rate evolution falls within the observational range during 1991. However, the increase in growth rate modelled at the end of 1991 using wetlands CH4 emissions from LPJ is not found in the observations. The observed decrease of about 10 ppb yr1 in CH4 growth rate during the year 1992 is captured by the model with a delay of half a year to 1 year. We have used two inventories for CH4 emission from wetlands to explore the potential role of uncertainties in this emission sector. Although they have a signicantly different variability, the two inventories give a similar performance in capturing the global CH4 variations.
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
N. Bnd
210 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
When using ORCHIDEE, the global CH4 growth rate is better captured in 1990 to 1993, while using LPJ we are able to reproduce better the CH4 growth rate at the end of the analysed period. The increase in CH4 in 1993 is not captured by either of the two scenarios. According to our breakdown in individual causes for CH4 growth rate changes, the overall decrease in the CH4 growth rate of 5 ppb yr1 during 1990 to 1995 is explained by the observed decrease in O3 column due to the 11-year cycle in solar activity, and by the estimated decrease in CO and NMVOC emissions in this period.
By analysing the nudging term, we nd that the mismatch most likely originates in the northern tropical region. Since the effects of UV changes and temperature changes on OH are considered to be robust, the most likely source of missing variability are natural CH4 emissions from wetlands.
Uncertainties in tropical biomass burning emissions, and in biogenic NMVOC emissions and their effect on OH, might have also contributed to the mismatch between the modelled and observed CH4 concentrations. Modelling CH4 emissions from wetlands is a challenging topic, due to the large spatial and temporal variability of these ecosystems. The large differences between the two emission data sets used here
in terms of CH4 emission amount and variability show that further research is needed to understand the factors driving emissions from wetlands and their response to environmental factors. Furthermore, the effect of changes in diffuse radiation and sulfur deposition after the eruption are not taken into account in the inventories used in this study. Another source of uncertainty are changes in NMVOC emissions, as well as the impact of NMVOC changes on OH. Further study is recommended using an updated isoprene chemistry scheme that considers the OH recycling by NMVOC, and using NMVOC emission models that take into account the effect of both direct and diffuse radiation. Finally, some uncertainty exists in our nudging procedure. An inverse modelling approach using OH elds from this study and all available CH4 station measurement data could better resolve the sources and timing of modelmeasurement mismatch. This might, almost 25 years after the Pinatubo eruption, further improve our knowledge of the drivers of CH4 growth rate variations.
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
a et al.: Explaining methane variability after the Pinatubo eruption 211
Appendix A: Validation of the two-step nudging set-up
N. Bnd
Figure B1. The combined effect of the seven drivers of CH4 variability on the global growth rate using ORCHIDEE emissions, assuming additivity (blue) and from the difference between Base1 and FixAll1.
In Fig. 3 we have shown the combined effect of the drivers for CH4 variability assuming additivity between the different drivers. We verify this assumption by using results from the FixAll1 simulation, where all drivers of CH4 were xed to 1990 values. Figure B1 shows the combined effect of the seven drivers for CH4 variability found from the sum of individual drivers for the rst simulation set (also shown in Fig. 5.4 and labelled All Orchidee) and the combined effect found from the difference between Base1 and FixAll1. The two global CH4 growth rate curves nearly overlap each other, with differences less than 0.2 ppb yr1. This shows that, indeed, the assumption of additivity is valid.
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
Figure A1. Surface mean CH4 (a) concentrations and (b) growth rate for the two runs of the Base1 simulation at the dateline (solid lines) and global mean (dashed lines). The black line shows the background mean concentrations based on ve background stations.
We use a two-step method to simulate realistic CH4 concentrations. In a rst step, Base1 and Base2 scenarios are run with near-surface CH4 mixing ratios nudged towards the zonal mean background mixing ratios inferred from measurements at the ve stations of South Pole, Cape Grim, Mauna Loa, Niwot Ridge, and Barrow. The nudging amount is stored on a monthly basis for each 10 latitude band, and used in the second step. In this second step, CH4 is no longer constrained by observations, but instead the nudging amount calculated in the rst step is applied as an emission in all scenarios in the lower 2 km.
Figure A1 presents the global and dateline monthly mean CH4 concentrations and the deseasonalised growth rates obtained in the two runs of Base1, and their comparison with Background5 observations. Please note that we show here absolute CH4 growth rates, and not variations with respect to 1990 as shown in Fig. 4. Also note that the actual CH4 growth rates in Base1 are plotted, unlike in Figs. 3 and 4 and where differences between simulations were presented. CH4 concentrations at the dateline were nudged in the rst step to the concentration values indicated by Background5. Therefore, the fact that the CH4 dateline mean concentrations follow reasonably well the observations is a result of the
nudging procedure. The global surface mean concentrations in the model show similar variations as the dateline mean, but with concentrations of 10 to 15 ppb higher. This is due to the fact that the dateline crosses the Pacic Ocean, and is remote from CH4 sources, while in the global mean both remote and polluted areas are included. The observed deseasonalised CH4 growth rate is well reproduced in the model both by the dateline mean and by the surface mean. The modelled dateline growth rate is within 1 ppb yr1 from the observed one, while the model global mean is within 2 ppb yr1. Both the surface mean and the dateline mean of the two steps from the model follow well the observed growth rate variations, with a slightly better performance of the dateline towards the end of the simulation.
The global mean and dateline mean concentrations and growth rates of the second step are nearly identical to those of the rst step. This shows that the procedure in which the correction is applied does not signicantly inuence the results, and that the simulation results of the two steps give a similar performance in capturing the observed CH4 growth rates.
Appendix B: Additivity between drivers of CH4 variability
212 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
Acknowledgements. We thank Bruno Ringeval and Joe Melton for providing the ORCHIDEE emission data. This work was supported by the Netherlands Organisation for Scientic Research (NWO). We thank the SURF Foundation (http://www.surfsara.nl
Web End =www.surfsara.nl ) for their support in using the Dutch national e-infrastructure.
Edited by: P. Jckel
References
Aan de Brugh, J. M. J., Schaap, M., Vignati, E., Dentener, F., Kahnert, M., Soev, M., Huijnen, V., and Krol, M. C.: The European aerosol budget in 2006, Atmos. Chem. Phys., 11, 11171139, doi:http://dx.doi.org/10.5194/acp-11-1117-2011
Web End =10.5194/acp-11-1117-2011 http://dx.doi.org/10.5194/acp-11-1117-2011
Web End = , 2011.
Andres, R. J. and Kasgnoc, A. D.: A time-averaged inventory of subaerial volcanic, J. Geophys. Res.h, 103, 2525125261, doi:http://dx.doi.org/10.1029/98JD02091
Web End =10.1029/98JD02091 http://dx.doi.org/10.1029/98JD02091
Web End = , 1998.
Aquila, V., Oman, L. D., Stolarski, R., Douglass, A. R., and Newman, P. A.: The Response of Ozone and Nitrogen Dioxide to the Eruption of Mt. Pinatubo at Southern and Northern Mid-latitudes, J. Atmos. Sci., 70, 894900, doi:http://dx.doi.org/10.1175/JAS-D-12-0143.1
Web End =10.1175/JAS-D-12- http://dx.doi.org/10.1175/JAS-D-12-0143.1
Web End =0143.1 , 2013.
Bnda, N., Krol, M., van Weele, M., van Noije, T., and Rck-mann, T.: Analysis of global methane changes after the 1991 Pinatubo volcanic eruption, Atmos. Chem. Phys., 13, 2267 2281, doi:http://dx.doi.org/10.5194/acp-13-2267-2013
Web End =10.5194/acp-13-2267-2013 http://dx.doi.org/10.5194/acp-13-2267-2013
Web End = , 2013.
Bnda, N., Krol, M., van Noije, T., van Weele, M., Williams, J. E.,
Le Sager, P., Niemeier, U., Thomason, L., and Rckmann, T.: The effect of stratospheric sulfur from Mount Pinatubo on tropospheric oxidizing capacity and methane, J. Geophys. Res.-Atmos., 120, 2014JD022137, doi:http://dx.doi.org/10.1002/2014JD022137
Web End =10.1002/2014JD022137 http://dx.doi.org/10.1002/2014JD022137
Web End = , 2014. Bekki, S. and Law, K. S.: Sensitivity of the atmospheric CH4 growth rate to global temperature changes observed from 1980 to 1992, Tellus, 49B, 409416, 1997.
Bekki, S. and Pyle, J. A.: A two-dimensional modeling study of the volcanic eruption of Mount Pinatubo, J. Geophys. Res., 99, 1886118869, 1994.
Bekki, S., Law, K. S., and Pyle, J. A.: Effect of ozone depletion on atmospheric CH4 and CO concentrations, Nature, 371, 595597, 1994.
Bousquet, P., Hauglustaine, D. A., Peylin, P., Carouge, C., and Ciais, P.: Two decades of OH variability as inferred by an inversion of atmospheric transport and chemistry of methyl chloroform, Atmos. Chem. Phys., 5, 26352656, doi:http://dx.doi.org/10.5194/acp-5-2635-2005
Web End =10.5194/acp-5- http://dx.doi.org/10.5194/acp-5-2635-2005
Web End =2635-2005 , 2005.
Bousquet, P., Ciais, P., Miller, J. B., Dlugokencky, E. J., Hauglustaine, D. A., Prigent, C., Van der Werf, G. R., Peylin, P., Brunke, E.-G., Carouge, C., Langenfelds, R. L., Lathire, J., Papa, F., Ramonet, M., Schmidt, M., Steele, L. P., Tyler,S. C., and White, J.: Contribution of anthropogenic and natural sources to atmospheric methane variability, Nature, 443, 43943, doi:http://dx.doi.org/10.1038/nature05132
Web End =10.1038/nature05132 http://dx.doi.org/10.1038/nature05132
Web End = , 2006.
Dalsren, S. B. and Isaksen, I. S. A.: CTM study of changes in tropospheric hydroxyl distribution 19902001 and its impact on methane, Geophys. Res. Lett., 33, L23811, doi:http://dx.doi.org/10.1029/2006GL027295
Web End =10.1029/2006GL027295 http://dx.doi.org/10.1029/2006GL027295
Web End = , 2006.
Dee, D. P., Uppala, S. M., Simmons, A. J., Berrisford, P., Poli,P., Kobayashi, S., Andrae, U., Balmaseda, M. A., Balsamo, G.,
Bauer, P., Bechtold, P., Beljaars, A. C. M., Berg, L. V. D., Bidlot, J., Bormann, N., Delsol, C., Dragani, R., Fuentes, M., Geer,A. J., Haimberger, L., Healy, S. B., Hersbach, H., Holm, E. V., Isaksen, L., and Kallberg, P.: The ERA-Interim reanalysis : conguration and performance of the data assimilation system, Q. J. Roy. Meteorol. Soc., 137, 553597, doi:http://dx.doi.org/10.1002/qj.828
Web End =10.1002/qj.828 http://dx.doi.org/10.1002/qj.828
Web End = , 2011. Denman, K., Brasseur, G., Chidthaisong, A., Ciais, P., Cox, P.,
Dickinson, R., Hauglustaine, D., Heinze, C., Holland, E., Jacob,D., Lohmann, U., Ramachandran, S., Dias, P. D. S., Wofsy, S., and Zhang, X.: Couplings Between Changes in the Climate System and Biogeochemistry, in: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, Tech. rep., Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 2007.
Dentener, F., Peters, W., Krol, M., Weele, M. V., and Bergamaschi,P.: Interannual variability and trend of CH4 lifetime as a measure for OH changes in the 19791993 time period, J. Geophys. Res., 108, D15, doi:http://dx.doi.org/10.1029/2002JD002916
Web End =10.1029/2002JD002916 http://dx.doi.org/10.1029/2002JD002916
Web End = , 2003.
Dessler, A. E., Schoeberl, M. R., Wang, T., Davis, S. M., Rosenlof,K. H., and Vernier, J.-P.: Variations of stratospheric water vapor over the past three decades, J. Geophys. Res.-Atmos., 119, 1258812598, doi:http://dx.doi.org/10.1002/2014JD021712
Web End =10.1002/2014JD021712 http://dx.doi.org/10.1002/2014JD021712
Web End = , 2014. Dlugokencky, E. J., Masaire, K. A., Lang, P. M., Tans, P. P., Steele,L. P., and Nisbet, E. G.: A dramatic decrease in the growth rate of atmospheric methane in the northern hemisphere during 1992, Geophys. Res. Lett., 21, 4548, 1994.
Dlugokencky, E. J., Dutton, E. G., Novelli, P. C., Tans, P. P.,
Masarie, K. A., Lantz, K. O., and Madronich, S.: Changes in CH4 and CO growth rates after the eruption of Mt. Pinatubo and their link with changes in tropical tropospheric UV ux, Geophys. Res. Lett., 23, 27612764, 1996.
Duneld, P., Knowles, R., Dumont, R., and Moore, T. R.: Methane production and consumption in temperate and subarctic peat soils: Response to temperature and pH, Soil Biol. Biochem., 25, 321326, doi:http://dx.doi.org/10.1016/0038-0717(93)90130-4
Web End =10.1016/0038-0717(93)90130-4 http://dx.doi.org/10.1016/0038-0717(93)90130-4
Web End = , 1993.
Gauci, V. and Chapman, S. J.: Simultaneous inhibition of CH4 efux and stimulation of sulphate reduction in peat subject to simulated acid rain, Soil Biol. Biochem., 38, 35063510, doi:http://dx.doi.org/10.1016/j.soilbio.2006.05.011
Web End =10.1016/j.soilbio.2006.05.011 http://dx.doi.org/10.1016/j.soilbio.2006.05.011
Web End = , 2006.
Gauci, V., Blake, S., Stevenson, D. S., and Highwood, E. J.:
Halving of the northern wetland CH 4 source by a large Icelandic volcanic eruption, J. Geophys. Res., 113, 18, doi:http://dx.doi.org/10.1029/2007JG000499
Web End =10.1029/2007JG000499 http://dx.doi.org/10.1029/2007JG000499
Web End = , 2008.
GLOBALVIEW-CH4: Cooperative Atmospheric Data Integration
Project Methane, CD-ROM, NOAA ESRL, Boulder, Colorado, also available via anonymous FTP to ftp://ftp.cmdl.noaa.gov
Web End =ftp://ftp.cmdl.noaa.gov , Path: ccg/ch4/GLOBALVIEW, 2009.
Groo, J.-U. and Russell III, James M.: Technical note: A stratospheric climatology for O3, H2O, CH4, NOx, HCl and HF derived from HALOE measurements, Atmos. Chem. Phys., 5, 27972807, doi:http://dx.doi.org/10.5194/acp-5-2797-2005
Web End =10.5194/acp-5-2797-2005 http://dx.doi.org/10.5194/acp-5-2797-2005
Web End = , 2005.
Guo, S., Bluth, G. J. S., Rose, W. I., Watson, I. M., and Prata, A. J.:
Re-evaluation of SO2 release of the 15 June 1991 Pinatubo eruption using ultraviolet and infrared satellite sensors, Geochem. Geophys. Geosyst., 5, Q04001, doi:http://dx.doi.org/10.1029/2003GC000654
Web End =10.1029/2003GC000654 http://dx.doi.org/10.1029/2003GC000654
Web End = , 2004.
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
a et al.: Explaining methane variability after the Pinatubo eruption 213
Houweling, S., Dentener, F., and Lelieveld, J.: The impact of non-methane hydrocarbon compounds on tropospheric photochemistry radical, J. Geophys. Res., 103, 1067310696, 1998.
Huijnen, V., Williams, J., van Weele, M., van Noije, T., Krol, M.,
Dentener, F., Segers, A., Houweling, S., Peters, W., de Laat, J., Boersma, F., Bergamaschi, P., van Velthoven, P., Le Sager, P., Eskes, H., Alkemade, F., Scheele, R., Ndlec, P., and Ptz, H.-W.: The global chemistry transport model TM5: description and evaluation of the tropospheric chemistry version 3.0, Geosci. Model Dev., 3, 445473, doi:http://dx.doi.org/10.5194/gmd-3-445-2010
Web End =10.5194/gmd-3-445-2010 http://dx.doi.org/10.5194/gmd-3-445-2010
Web End = , 2010.Huijnen, V., Williams, J. E., and Flemming, J.: Modeling global impacts of heterogeneous loss of HO2 on cloud droplets, ice particles and aerosols, Atmos. Chem. Phys. Discuss., 14, 85758632, doi:http://dx.doi.org/10.5194/acpd-14-8575-2014
Web End =10.5194/acpd-14-8575-2014 http://dx.doi.org/10.5194/acpd-14-8575-2014
Web End = , 2014.
Kirschke, S., Bousquet, P., Ciais, P., Saunois, M., Canadell, J. G., Dlugokencky, E. J., Bergamaschi, P., Bergmann, D., Blake, D. R., Bruhwiler, L., Cameron-Smith, P., Castaldi, S., Chevallier, F., Feng, L., Fraser, A., Heimann, M., Hodson, E. L., Houweling,S., Josse, B., Fraser, P. J., Krummel, P. B., Lamarque, J.-F., Langenfelds, R. L., Le Qur, C., Naik, V., ODoherty, S., Palmer,P. I., Pison, I., Plummer, D., Poulter, B., Prinn, R. G., Rigby,M., Ringeval, B., Santini, M., Schmidt, M., Shindell, D. T., Simpson, I. J., Spahni, R., Steele, L. P., Strode, S. a., Sudo,K., Szopa, S., van der Werf, G. R., Voulgarakis, A., van Weele,M., Weiss, R. F., Williams, J. E., and Zeng, G.: Three decades of global methane sources and sinks, Na. Geosci., 6, 813823, doi:http://dx.doi.org/10.1038/ngeo1955
Web End =10.1038/ngeo1955 http://dx.doi.org/10.1038/ngeo1955
Web End = , 2013.
Lamarque, J.-F., Bond, T. C., Eyring, V., Granier, C., Heil, A., Klimont, Z., Lee, D., Liousse, C., Mieville, A., Owen, B., Schultz, M. G., Shindell, D., Smith, S. J., Stehfest, E., Van Aardenne, J., Cooper, O. R., Kainuma, M., Mahowald, N., Mc-Connell, J. R., Naik, V., Riahi, K., and van Vuuren, D. P.: Historical (18502000) gridded anthropogenic and biomass burning emissions of reactive gases and aerosols: methodology and application, Atmos. Chem. Phys., 10, 70177039, doi:http://dx.doi.org/10.5194/acp-10-7017-2010
Web End =10.5194/acp- http://dx.doi.org/10.5194/acp-10-7017-2010
Web End =10-7017-2010 , 2010.
Law, K. S. and Nisbet, E. G.: Sensitivity of the CH4 growth rate to changes in CH4 emissions from natural gas and coal, J. Geophys.
Res., 101, 1438714397, 1996.
Lawrence, M. G., Jckel, P., and von Kuhlmann, R.: What does the global mean OH concentration tell us?, Atmos. Chem. Phys., 1, 3749, doi:http://dx.doi.org/10.5194/acp-1-37-2001
Web End =10.5194/acp-1-37-2001 http://dx.doi.org/10.5194/acp-1-37-2001
Web End = , 2001.
Lelieveld, J., Peters, W., Dentener, F. J., and Krol, M. C.: Stability of tropospheric hydroxyl chemistry, J. Geophys. Res., 107, 4715, doi:http://dx.doi.org/10.1029/2002JD002272
Web End =10.1029/2002JD002272 http://dx.doi.org/10.1029/2002JD002272
Web End = , 2002.
McCormick, M. P., Thomason, L. W., and Trepte, C. R.: Atmospheric effects of the Mt. Pinatubo eruption, Nature, 373, 399 404, 1995.
Melton, J. R., Wania, R., Hodson, E. L., Poulter, B., Ringeval, B., Spahni, R., Bohn, T., Avis, C. A., Beerling, D. J., Chen, G., Eliseev, A. V., Denisov, S. N., Hopcroft, P. O., Lettenmaier, D.P., Riley, W. J., Singarayer, J. S., Subin, Z. M., Tian, H., Zrcher,S., Brovkin, V., van Bodegom, P. M., Kleinen, T., Yu, Z. C., and Kaplan, J. O.: Present state of global wetland extent and wetland methane modelling: conclusions from a model inter-comparison project (WETCHIMP), Biogeosciences, 10, 753 788, doi:http://dx.doi.org/10.5194/bg-10-753-2013
Web End =10.5194/bg-10-753-2013 http://dx.doi.org/10.5194/bg-10-753-2013
Web End = , 2013.
Mercado, L. M., Bellouin, N., Sitch, S., Boucher, O., Huntingford,C., Wild, M., and Cox, P. M.: Impact of changes in diffuse ra-
diation on the global land carbon sink, Nature, 458, 10141018, doi:http://dx.doi.org/10.1038/nature07949
Web End =10.1038/nature07949 http://dx.doi.org/10.1038/nature07949
Web End = , 2009.
Metzger, S., Dentener, F., Pandis, S., and Lelieveld, J.: Gas/aerosol partitioning: 1. A computationally efcient model, Journal of Geophysical Research: Atmospheres, 107, ACH 161ACH 16 24, doi:http://dx.doi.org/10.1029/2001JD001102
Web End =10.1029/2001JD001102 http://dx.doi.org/10.1029/2001JD001102
Web End = , 2002.
Montzka, S. A., Dlugokencky, E. J., and Butler, J. H.: Non-CO2 greenhouse gases and climate change, Nature, 476, 4350, doi:http://dx.doi.org/10.1038/nature10322
Web End =10.1038/nature10322 http://dx.doi.org/10.1038/nature10322
Web End = , 2011.
Niemeier, U., Timmreck, C., Graf, H.-F., Kinne, S., Rast, S., and Self, S.: Initial fate of ne ash and sulfur from large volcanic eruptions, Atmos. Chem. Phys., 9, 90439057, doi:http://dx.doi.org/10.5194/acp-9-9043-2009
Web End =10.5194/acp- http://dx.doi.org/10.5194/acp-9-9043-2009
Web End =9-9043-2009 , 2009.
Patra, P. K., Krol, M. C., Montzka, S. A., Arnold, T., Atlas, E. L., Lintner, B. R., Stephens, B. B., Xiang, B., Elkins, J. W., Fraser,P. J., Ghosh, A., Hintsa, E. J., Hurst, D. F., Ishijima, K., Krummel, P. B., Miller, B. R., Miyazaki, K., Moore, F. L., ODoherty,S., Prinn, R. G., Steele, L. P., Takigawa, M., Wang, H. J., Weiss,R. F., Wofsy, S. C., and Young, D.: Observational evidence for interhemispheric hydroxyl-radical parity, Nature, 513, 219223, doi:http://dx.doi.org/10.1038/nature13721
Web End =10.1038/nature13721 http://dx.doi.org/10.1038/nature13721
Web End = , 2014.
Pison, I., Ringeval, B., Bousquet, P., Prigent, C., and Papa, F.: Stable atmospheric methane in the 2000s: key-role of emissions from natural wetlands, Atmos. Chem. Phys., 13, 1160911623, doi:http://dx.doi.org/10.5194/acp-13-11609-2013
Web End =10.5194/acp-13-11609-2013 http://dx.doi.org/10.5194/acp-13-11609-2013
Web End = , 2013.
Pschl, U., von Kuhlmann, R., Poisson, N., and Crutzen, P.: Development and Intercomparison of Condensed Isoprene Oxidation Mechanisms for Global Atmospheric Modeling, J. Atmos. Chem., 37, 2952, doi:http://dx.doi.org/10.1023/A:1006391009798
Web End =10.1023/A:1006391009798 http://dx.doi.org/10.1023/A:1006391009798
Web End = , 2000. Prinn, R. G., Huang, J., Weiss, R. F., Cunnold, D. M., Fraser,P. J., Simmonds, P. G., McCulloch, A., Harth, C., Reimann, S., Salameh, P., ODoherty, S., Wang, R. H. J., Porter, L. W., Miller,B. R., and Krummel, P. B.: Evidence for variability of atmospheric hydroxyl radicals over the past quarter century, Geophys. Res. Lett., 32, L07809, doi:http://dx.doi.org/10.1029/2004GL022228
Web End =10.1029/2004GL022228 http://dx.doi.org/10.1029/2004GL022228
Web End = , 2005. Ringeval, B., de Noblet-Ducoudr, N., Ciais, P., Bousquet, P., Pri-gent, C., Papa, F., and Rossow, W. B.: An attempt to quantify the impact of changes in wetland extent on methane emissions on the seasonal and interannual time scales, Global Biogeochem. Cy., 24, GB2003, doi:http://dx.doi.org/10.1029/2008GB003354
Web End =10.1029/2008GB003354 http://dx.doi.org/10.1029/2008GB003354
Web End = , 2010.
Rohrer, F., Lu, K., Hofzumahaus, A., Bohn, B., Brauers, T., Chang,C.-C., Fuchs, H., Haseler, R., Holland, F., Hu, M., Kita, K., Kondo, Y., Li, X., Lou, S., Oebel, A., Shao, M., Zeng, L., Zhu, T., Zhang, Y., and Wahner, A.: Maximum efciency in the hydroxyl-radical-based self-cleansing of the troposphere, Nat. Geosci., 7, 559563, 2014.
Schaufer, S. M. and Daniel, J. S.: On the effects of stratospheric circulation changes on trace gas trends, J. Geophys. Res., 99, 2574725754, 1994.
Schultz, M. G., Heil, A., Hoelzemann, J. J., Spessa, A., Thonicke, K., Goldammer, J. G., Held, A. C., Pereira, J. M. C., and van het Bolscher, M.: Global wildland re emissions from 1960 to 2000, Global Biogeochem. Cy., 22, GB2002, doi:http://dx.doi.org/10.1029/2007GB003031
Web End =10.1029/2007GB003031 http://dx.doi.org/10.1029/2007GB003031
Web End = , 2008.
Shepherd, T. G., Plummer, D. A., Scinocca, J. F., Hegglin, M. I., Fioletov, V. E., Reader, M. C., Remsberg, E., von Clarmann, T., and Wang, H. J.: Reconciliation of halogen-induced ozone loss with the total-column ozone record, Nat. Geosci., 7, 443449, doi:http://dx.doi.org/10.1038/ngeo2155
Web End =10.1038/ngeo2155 http://dx.doi.org/10.1038/ngeo2155
Web End = , 2014.
N. Bnd
www.atmos-chem-phys.net/16/195/2016/ Atmos. Chem. Phys., 16, 195214, 2016
214 N. Bnd
a et al.: Explaining methane variability after the Pinatubo eruption
Sindelarova, K., Granier, C., Bouarar, I., Guenther, A., Tilmes, S., Stavrakou, T., Mller, J.-F., Kuhn, U., Stefani, P., and Knorr, W.: Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 years, Atmos. Chem. Phys., 14, 93179341, doi:http://dx.doi.org/10.5194/acp-14-9317-2014
Web End =10.5194/acp-14-9317-2014 http://dx.doi.org/10.5194/acp-14-9317-2014
Web End = , 2014.
Soden, B. J., Wetherald, R. T., Stenchikov, G. L., and Robock, A.: Global Cooling After the Eruption of Mount Pinatubo: A Test of Climate Feedback by Water Vapor, Science, 296, 727730, doi:http://dx.doi.org/10.1126/science.296.5568.727
Web End =10.1126/science.296.5568.727 http://dx.doi.org/10.1126/science.296.5568.727
Web End = , 2002.
Spahni, R., Wania, R., Neef, L., van Weele, M., Pison, I., Bousquet,P., Frankenberg, C., Foster, P. N., Joos, F., Prentice, I. C., and van Velthoven, P.: Constraining global methane emissions and uptake by ecosystems, Biogeosciences, 8, 16431665, doi:http://dx.doi.org/10.5194/bg-8-1643-2011
Web End =10.5194/bg- http://dx.doi.org/10.5194/bg-8-1643-2011
Web End =8-1643-2011 , 2011.
Stone, D., Evans, M. J., Edwards, P. M., Commane, R., Ingham,T., Rickard, A. R., Brookes, D. M., Hopkins, J., Leigh, R. J., Lewis, A. C., Monks, P. S., Oram, D., Reeves, C. E., Stewart, D., and Heard, D. E.: Isoprene oxidation mechanisms: measurements and modelling of OH and HO2 over a South-East Asian tropical rainforest during the OP3 eld campaign, Atmos. Chem. Phys., 11, 67496771, doi:http://dx.doi.org/10.5194/acp-11-6749-2011
Web End =10.5194/acp-11-6749-2011 http://dx.doi.org/10.5194/acp-11-6749-2011
Web End = , 2011.
Telford, P., Braesicke, P., Morgenstern, O., and Pyle, J.: Reassessment of causes of ozone column variability following the eruption of Mount Pinatubo using a nudged CCM, Atmos. Chem. Phys., 9, 42514260, doi:http://dx.doi.org/10.5194/acp-9-4251-2009
Web End =10.5194/acp-9-4251-2009 http://dx.doi.org/10.5194/acp-9-4251-2009
Web End = , 2009. Telford, P. J., Lathire, J., Abraham, N. L., Archibald, A. T.,
Braesicke, P., Johnson, C. E., Morgenstern, O., OConnor, F. M., Pike, R. C., Wild, O., Young, P. J., Beerling, D. J., Hewitt, C.N., and Pyle, J.: Effects of climate-induced changes in isoprene emissions after the eruption of Mount Pinatubo, Atmos. Chem. Phys., 10, 71177125, doi:http://dx.doi.org/10.5194/acp-10-7117-2010
Web End =10.5194/acp-10-7117-2010 http://dx.doi.org/10.5194/acp-10-7117-2010
Web End = , 2010. Tiedtke, M.: A comprehensive mass ux scheme for cumulus parameterization in large-scale models, Mon. Weather Rev., 117, 17791800, 1989.van der A, R. J., Allaart, M. A. F., and Eskes, H. J.: Multi sensor re-analysis of total ozone, Atmos. Chem. Phys., 10, 1127711294, doi:http://dx.doi.org/10.5194/acp-10-11277-2010
Web End =10.5194/acp-10-11277-2010 http://dx.doi.org/10.5194/acp-10-11277-2010
Web End = , 2010.van Noije, T. P. C., Le Sager, P., Segers, A. J., van Velthoven, P. F.J., Krol, M. C., Hazeleger, W., Williams, A. G., and Chambers,S. D.: Simulation of tropospheric chemistry and aerosols with the climate model EC-Earth, Geosci. Model Dev., 7, 24352475, doi:http://dx.doi.org/10.5194/gmd-7-2435-2014
Web End =10.5194/gmd-7-2435-2014 http://dx.doi.org/10.5194/gmd-7-2435-2014
Web End = , 2014.
Vignati, E., Wilson, J., and Stier, P.: M7: An efcient size-resolved aerosol microphysics module for large-scale aerosol transport models, J. Geophys. Res.-Atmos., 109, D22202, doi:http://dx.doi.org/10.1029/2003JD004485
Web End =10.1029/2003JD004485 http://dx.doi.org/10.1029/2003JD004485
Web End = , 2004.
Vignati, E., Facchini, M., Rinaldi, M., Scannell, C., Ceburnis, D., Sciare, J., Kanakidou, M., Myriokefalitakis, S., Dentener, F., and ODowd, C.: Global scale emission and distribution of sea-spray aerosol: Sea-salt and organic enrichment, Atmos. Environ., 44, 670677, doi:http://dx.doi.org/10.1016/j.atmosenv.2009.11.013
Web End =10.1016/j.atmosenv.2009.11.013 http://dx.doi.org/10.1016/j.atmosenv.2009.11.013
Web End = , 2010.
Wang, J. S., Logan, J. A., Mcelroy, M. B., Duncan, B. N., Megretskaia, I. A., and Yantosca, R. M.: A 3-D model analysis of the slowdown and interannual variability in the methane growth rate from 1988 to 1997, Global Biogeochem. Cy., 18, GB3011, doi:http://dx.doi.org/10.1029/2003GB002180
Web End =10.1029/2003GB002180 http://dx.doi.org/10.1029/2003GB002180
Web End = , 2004.
Wania, R., Melton, J. R., Hodson, E. L., Poulter, B., Ringeval,B., Spahni, R., Bohn, T., Avis, C. A., Chen, G., Eliseev, A. V., Hopcroft, P. O., Riley, W. J., Subin, Z. M., Tian, H., van Bode-gom, P. M., Kleinen, T., Yu, Z. C., Singarayer, J. S., Zrcher, S., Lettenmaier, D. P., Beerling, D. J., Denisov, S. N., Prigent, C., Papa, F., and Kaplan, J. O.: Present state of global wetland extent and wetland methane modelling: methodology of a model inter-comparison project (WETCHIMP), Geosci. Model Dev., 6, 617641, doi:http://dx.doi.org/10.5194/gmd-6-617-2013
Web End =10.5194/gmd-6-617-2013 http://dx.doi.org/10.5194/gmd-6-617-2013
Web End = , 2013.
Williams, J. E., Landgraf, J., Bregman, A., and Walter, H. H.: A modied band approach for the accurate calculation of online photolysis rates in stratospheric-tropospheric Chemical Transport Models, Atmos. Chem. Phys., 6, 41374161, doi:http://dx.doi.org/10.5194/acp-6-4137-2006
Web End =10.5194/acp-6-4137-2006 http://dx.doi.org/10.5194/acp-6-4137-2006
Web End = , 2006.
Williams, J. E., Strunk, A., Huijnen, V., and van Weele, M.: The application of the Modied Band Approach for the calculation of on-line photodissociation rate constants in TM5: implications for oxidative capacity, Geosci. Model Dev., 5, 1535, doi:http://dx.doi.org/10.5194/gmd-5-15-2012
Web End =10.5194/gmd-5-15-2012 http://dx.doi.org/10.5194/gmd-5-15-2012
Web End = , 2012.
Wilton, D. J., Hewitt, C. N., and Beerling, D. J.: Simulated effects of changes in direct and diffuse radiation on canopy scale isoprene emissions from vegetation following volcanic eruptions, Atmos. Chem. Phys., 11, 1172311731, doi:http://dx.doi.org/10.5194/acp-11-11723-2011
Web End =10.5194/acp- http://dx.doi.org/10.5194/acp-11-11723-2011
Web End =11-11723-2011 , 2011.
Atmos. Chem. Phys., 16, 195214, 2016 www.atmos-chem-phys.net/16/195/2016/
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
Copyright Copernicus GmbH 2016
Abstract
The CH<sub>4</sub> growth rate in the atmosphere showed large variations after the Pinatubo eruption in June 1991. A decrease of more than 10ppbyr<sup>-1</sup> in the growth rate over the course of 1992 was reported, and a partial recovery in the following year. Although several reasons have been proposed to explain the evolution of CH<sub>4</sub> after the eruption, their contributions to the observed variations are not yet resolved. CH<sub>4</sub> is removed from the atmosphere by the reaction with tropospheric OH, which in turn is produced by O<sub>3</sub> photolysis under UV radiation. The CH<sub>4</sub> removal after the Pinatubo eruption might have been affected by changes in tropospheric UV levels due to the presence of stratospheric SO<sub>2</sub> and sulfate aerosols, and due to enhanced ozone depletion on Pinatubo aerosols. The perturbed climate after the eruption also altered both sources and sinks of atmospheric CH<sub>4</sub>. Furthermore, CH<sub>4</sub> concentrations were influenced by other factors of natural variability in that period, such as El Niño-Southern Oscillation (ENSO) and biomass burning events. Emissions of CO, NO<sub><i>X</i></sub> and non-methane volatile organic compounds (NMVOCs) also affected CH<sub>4</sub> concentrations indirectly by influencing tropospheric OH levels.</p><p class="p">Potential drivers of CH<sub>4</sub> variability are investigated using the TM5 global chemistry model. The contribution that each driver had to the global CH<sub>4</sub> variability during the period 1990 to 1995 is quantified. We find that a decrease of 8-10ppbyr<sup>-1</sup> CH<sub>4</sub> is explained by a combination of the above processes. However, the timing of the minimum growth rate is found 6 to 9 months later than observed. The long-term decrease in CH<sub>4</sub> growth rate over the period 1990 to 1995 is well captured and can be attributed to an increase in OH concentrations over this time period. Potential uncertainties in our modelled CH<sub>4</sub> growth rate include emissions of CH<sub>4</sub> from wetlands, biomass burning emissions of CH<sub>4</sub> and other compounds, biogenic NMVOC and the sensitivity of OH to NMVOC emission changes. Two inventories are used for CH<sub>4</sub> emissions from wetlands, ORCHIDEE and LPJ, to investigate the role of uncertainties in these emissions. Although the higher climate sensitivity of ORCHIDEE improves the simulated CH<sub>4</sub> growth rate change after Pinatubo, none of the two inventories properly captures the observed CH<sub>4</sub> variability in this period.
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer