You are on page 1of 9

GLOBAL BIOGEOCHEMICAL CYCLES, VOL. 21, GB4013, doi:10.

1029/2006GB002734, 2007

A global inventory of the soil CH4 sink


Laure Dutaur1 and Louis V. Verchot1
Received 6 April 2006; revised 16 February 2007; accepted 8 March 2007; published 15 November 2007.

[1] Methane uptake by soils is a small but important flux in the global budget of
atmospheric methane, and could be susceptible to changes in land use and climate.
Estimates of this sink vary between 20 and 45 Tg yr 1. We propose to develop a better
constrained estimate using a mechanistic understanding of the biogeochemical controls of
soil CH4 uptake. We reviewed over 120 published papers reporting field measurements
of CH4 uptake and made over 318 annual estimates of CH4 uptake for various types of
ecosystems. We collected data from these papers for a number of parameters that are
known to influence the magnitude of the sink including climatic zone, ecosystem,
latitude, annual mean rainfall, annual mean temperature, and the soil texture. Regression
analyses with the continuous variables (latitude, rainfall, and temperature) yielded results
with poor predictive ability and no significant relationship. Stratification by class
variables such as climatic zone, ecosystem type and soil texture provided better predictive
ability (R2 = 0.29, P < 0.0001). The mean largest uptake rates were observed in temperate
forests with coarse soil texture, but the variance within this stratum was also large.
Without any stratification, we estimate that the global soil CH4 sink is 36 ± 23 Tg yr 1.
With stratification, the best current estimate of the global soil uptake of CH4 is 22 ±
12 Tg yr 1. The ecosystem type accounted for the largest part of the variation in the
global data set. This inventory showed that ecosystem type, geographic zone, and soil
texture strongly control CH4 uptake. Inventory methods that take into account underlying
factors that control the process provide better estimates of sink strength.
Citation: Dutaur, L., and L. V. Verchot (2007), A global inventory of the soil CH4 sink, Global Biogeochem. Cycles, 21, GB4013,
doi:10.1029/2006GB002734.

1. Introduction and therefore, requires molecular O2. Methanogenesis is the


process of microbial production of CH4 in anaerobic con-
[2] Understanding the sources and sinks of atmospheric
ditions. Methanogenesis is an important process in wetland
trace gas constituents is important for understanding anthro-
soils and rice paddies and these systems are usually sources
pogenic impacts on the radiative balance of the atmosphere.
of CH4 for the atmosphere. However, methanogenesis can
Many of the budgets of these gases are poorly constrained,
also occur in upland soils, inside soil aggregates where
which limits our ability to understand the dynamics of these
anaerobic ‘microsites’ occur. Methanotrophy is the domi-
gases. The atmospheric concentration of methane (CH4) has
nant process in upland soils, where oxidation generally
doubled since pre-industrial times. The current concentration
exceeds production and there is a net uptake by the soil
is 1780 ppbv and, until recently, the concentration was
of CH4 from the atmosphere. Methanotrophy is also an
increasing. The rate of increase in the concentration of
important process in wetland soils at the aerobic soil-water
atmospheric CH4 slowed from about 15 ppbv yr 1 in the
interface. One estimate of the importance of soil methano-
1980s to near zero in 1999 [Dlugokencky, 2001]. Since 1990,
trophy suggests that as much as 50% of the CH4 produced
the annual rate of CH4 increase in the atmosphere has varied
in soils and sediments is consumed therein [Reeburgh et al.,
between less than zero and 15 ppbv. The reasons for this
1993].
change are not known.
[4] In the global CH4 budget soils are the largest biotic
[3] Soils both produce and consume CH4. The net soil –
sink for atmospheric CH4, consuming 15– 45 Tg annually, a
atmosphere CH4 flux is the result of the balance between the
rate that is of similar magnitude as the rate of CH4
two offsetting processes of methanogenesis (microbial pro-
accumulation in the atmosphere during the 1990s [Potter
duction) and methanotrophy (microbial consumption).
et al., 1996; Ridgwell et al., 1999]. Thus any significant
Methanotrophs and methyltrophs are all obligate aerobes;
change made to the soil CH4 sink could alter the net
the biochemical process requires a monooxygenase enzyme
biosphere – atmosphere flux and alter the atmospheric accu-
mulation rate of this potent greenhouse gas. Ojima et al.
1
[1993] suggested that there has been a significant increase
International Centre for Research in Agroforestry, Nairobi, Kenya.
in the soil sink as atmospheric mixing ratios of CH4 have
Copyright 2007 by the American Geophysical Union.
increased. At the same time, land use change and expansion
0886-6236/07/2006GB002734 of agriculture has significantly reduced the strength of the

GB4013 1 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

soil sink [Potter et al., 1996; Smith et al., 2000]. Ojima et variability in the CH4 uptake. Boeckx and Van Cleemput
al. [1993] calculated that the temperate forest and grassland [2001] noted that CH4 uptake was higher in forest soils than
sink for atmospheric CH4 has been reduced by 1.5 to 6 Tg in grasslands and arable soils.
CH4 yr 1. The changes in sink strength associated with land [8] We used the relationships developed from this analy-
use change are of an order of magnitude such that they are sis to explore different stratification schemes for estimating
affecting the atmospheric accumulation rate of this impor- the global soil CH4 sink. We judged the success of this
tant greenhouse gas. approach through the reduction in variation within strata
[5] Several general reviews have recently discussed the compared to more aggregated stratification schemes.
role of soils as source and sink for atmospheric CH4
[Conrad, 1989; Nedwell, 1996; Topp and Pattey, 1997; Le 2. Methods
Mer and Roger, 2001]. During the past fifteen years, the
number of published papers reporting on soil CH4 uptake [9] We compiled a database from 120 studies that reported
has grown. This work has shown that variability in the CH4 fluxes from different terrestrial biomes (auxiliary
methane uptake results from the interactions of a number of material1). We included data from studies that reported soil
factors that regulate methane oxidation in the soil, for texture, annual rainfall, mean temperature and latitude coor-
example climate, soil physical properties like water content dinates. Few studies were designed to estimate annual
and porosity, and biological variables like the nature of the uptake of methane by soils. Some reported fluxes during
soil microbial community. Among these factors, soil phys- more than one season, but several reported fluxes only
ical properties are likely to have the greatest influence on during a brief period, which was often during the summer
the soil CH4 sink at the regional to global scales [Potter et in temperate regions. Whenever the authors provided their
al., 1996; Striegl et al., 1992; Verchot et al., 2000; Del own extrapolation to an annual estimate, we use their
Grosso et al., 2000; Borken and Beese, 2006]. Low CH4 estimate. Where authors provided no annual extrapolations,
oxidation rates are consistently associated with fine textured we used our best judgment to make an estimate of the annual
soils that have low porosity and high water retention, which total.
combine to restrict diffusion of CH4 into the soil profile. [10] The method for extrapolation varied with the climatic
Coarse texture soils tend to drain rapidly and have high zone, as we took into account seasonal variations. General-
porosity and therefore high diffusion rates of CH4 into the ly, these studies presented data that were complete enough
soil profile. Several authors [Verchot et al., 2000; Kiese et to calculate the total flux from the site for the frost-free
al., 2003] noted that in environments with seasonal rainfall season. However, Mosier et al. [1997] working in alpine
patterns, CH4 uptake closely followed the pattern of rainfall, tundra ecosystems showed that even in frozen soils CH4
increasing rapidly during the dry season as soil water fluxes were maintained, albeit at lower rates than during
content declines. Other factors also affect uptake of CH4 frost-free seasons. Thus, when studies reported fluxes only
by soils. For example, cold winter temperatures inhibit for the frost-free part of the year, we assumed that these
microbial activity over the large landmasses of the northern fluxes represented 70% of the annual flux. Therefore,
hemisphere. However, Mosier et al. [1997] have shown that depending on the latitude, we calculated the CH4 sink for
frozen soils continue to consume CH4 at a reduced rate. the measurement period between 100 and 365 days, and for
[6] The first objective of this paper is to use this new data ecosystems where soils froze, we assumed that this value
to reevaluate the global inventory of CH4 uptake by soils. represented 70% of the annual flux.
Extrapolation estimates of greenhouse gas fluxes are often [11] All CH4 uptake fluxes were expressed in terms of kg
done by surveying the literature, averaging means from a CH4 ha 1yr 1. When the annual mean temperature, annual
number of studies, and multiplying by the area covered by mean rainfall or latitude data were missing, we used
the biome or ecosystem [Potter et al., 1996; Mosier et al., globally gridded data sets available on the worldwide
1998; Smith et al., 2000]. However, there is a tremendous web: www.weatherbase.com and www.mappoint.msn.com,
amount of variation within a biome and we understand according to the site description provided in the paper.
enough about the controlling factors that we ought to be [12] We stratified the data into different biomes according
able to use them reduce the variance of the means within the to the Leemans’ classification [Leemans, 1990]. Many
extrapolation units. Thus our second objective is to examine papers did not give the type of forest, so we grouped several
the potential for explaining variability across studies in the of Leemans’ categories together: tundra forest with boreal
literature using parameters that are often measured and that forest, cold temperate forests with warm temperate forests,
have mechanistic significance. and tropical seasonal forests with rain forests. We also
[7] In this study, we looked at a number of factors that are grouped cold and hot deserts together because of their
likely to explain large-scale variability in CH4 uptake by low biological activity and because of the lack of data.
soils. Latitude is a good proxy for the length of the growing The soil texture data were used to classify soils into in four
season or of the frost-free season. Seasonal variations in texture classes: organic, fine, medium and coarse, according
precipitation also correspond to variations in temperature to the USDA classification system [Brady, 1974].
and these two factors work in concert to affect fluxes. Thus [13] We made over 310 estimates of annual CH4 uptake
parameters such as mean annual precipitation and mean by soils (auxiliary material). This number differs from the
temperature can be used as indicators of climate controls on
microbial processes as well as on diffusion across the soil—
1
atmosphere boundary. Ecosystem type also affects the Auxiliary materials are available at ftp://ftp.agu.org/apend/gb/
2006gb002734.

2 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

Figure 1. Location of study sites used for this analysis (GIS Unit, ICRAF).

number of published papers because several papers reported tionally, measurement density in the taiga, tundra and
fluxes for more than one type of ecosystem or for more than deserts was low.
one crop. In a few cases, we combined estimates from more [15] According to classification scheme of Leemans
than one published paper for the same study area to estimate [1990], which was also used by Potter et al. [1996] for
one single annual flux for the site. We analyzed the data to their global extrapolation, the land area of the globe can be
explore the relationships between CH4 flux and environ- divided into 14 ecosystem types. We present these ecosys-
mental or physical variables through both regression anal- tem types in Table 1 along with the number sites for which
ysis and analysis of variance using SAS statistical software we made estimates of annual CH4 uptake and the mean
[SAS Institute, 2005]. The data set was not normally uptake rates. We note that the sum of the area data in Table 1
distributed and even with a logarithmic transformation, is superior to those used in other CH4 sink estimates in this
the data set deviated significantly from a normal distribu- paper. In later estimates, we removed lithosols from the area
tion, albeit to a lesser degree than untransformed data. Thus, as we expect that the CH4 uptake in these soils is nil.
to reduce heteroskedasticity and to improve the distribution, Additionally the FAO classification lists 30% of the soils as
we conducted all statistical analyses with transformed data. undefined for tundra, so stratification based upon texture is
All extrapolations were based upon the arithmetic mean impossible. Therefore the tundra area that we used for
rather than the geometric mean, as the arithmetic mean is our estimate was only 0.05  106 km2 rather than 9.41 
the center of probability of a distribution and is therefore 106 km2.
parameter of choice for extrapolation [Parkin and Robinson, [16] Forests, with 205 sites, represented more than 65% of
1994]. the study sites. Over 29% of these sites were from temperate
forests; 19% were from tropical forest; and 16% were from
3. Results boreal forests. Among the remaining sites, 42% were culti-
vated. Desert, tundra, tropical steppes, savannas, and chap-
3.1. Adequacy of the Data Set arral were all under represented in the data set. The variance
[14] The data set was extensive and covered the major in the temperate forest was very high, but in the other
ecosystems of the world. We note however, that the distri- ecosystems, they were of similar order of magnitude as the
bution of the sites was uneven, with high concentration of mean. Temperate forests were very well represented in the
sites in eastern North America, Western Europe and north- literature and very well distributed around the world, with
ern South America (Figure 1). Large areas of Africa, Asia measurements at 92 different sites by 40 different authors.
and southern South America had no measurements. Addi- [17] CH4 uptake rates varied between 0.04 and 49.93 kg
(CH4) ha 1yr 1. The highest consumption rates were

3 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

Table 1. Summary of the Data Set Used for the Analyses in This Paper by Biome
Total Land Area,b Number of Mean,
Ecosystem Typea  106 km2 Study Sites kgCH4 ha 1 yr 1
Variance
Tundra 9.41 11 1.49 5.52
Forest 51 2.64 8.06
Forest tundra 8.18
Boreal forest 15.31
Desert 5 1.10 1.03
Cool desert 2.63
Hot desert 19.29
Grasslands 2.61 29 2.32 2.92
Cultivationc 24.79 47 1.23 1.22
Chaparral 2.17 3 2.25 0.54
Forest 92 5.70 31.50
Cool temperate forest 5.07
Warm temperate forest 1.69
Tropical steppe/savanna 18 1.49 1.54
Tropical semiarid steppe 7.98
Tropical savanna 11.50
Tropical forest 62 3.33 4.82
Tropical seasonal forest 12.83
Tropical rain forest 7.17
Total 130.6 318 3.29 14.80
a
Leemans [1990].
b
Potter et al. [1996].
c
Breakdown of sites: boreal, 3; tropical, 6; temperate, 38.

measured by Singh et al. [1997] in savanna and tropical of outliers. The coefficient of variation for the temperate
forest ecosystems. Excluding these very high rates, the observed zone was 120%, compared to 115% for the boreal zone and
fluxes range 0.04 to 27.74 kg (CH4) ha 1yr 1. There is 77% for the tropical zone.
no obvious explanation for such high average fluxes, as [20] We then stratified the global data set by ecosystem
compared with other CH4 fluxes from same ecosystems type. Analysis of variance indicated that there was a major
(auxiliary material) even if they are representative of par- ecosystem effect, so we stratified the data set into forest and
ticular situations. Thus they were not considered in the nonforest (Figure 3). This stratification scheme explained
statistical analysis. 12% of the global variation and was highly significant (P <
0.0001). The average uptake for forests in the global data set
3.2. Analysis of Drivers at the Global Scale was 4.22 kg (CH4) ha 1yr 1, while that of the other
[18] Among the explanatory variables that we collected ecosystems was 1.60 kg (CH4) ha 1yr 1. The forest had
from the publications, several were continuous variables,
which lend themselves to regression analysis; others were
categorical variables, which lend themselves to analysis of
variance. For regression analyses on the continuous varia-
bles, we found a marginally significant relationship between
CH4 and latitude (P = 0.0679) and significant relationships
with annual rainfall (P = 0.0016) and mean annual temper-
ature (P = 0.0226). However, these relationships had low
predictive power: rainfall explained about 3% of the vari-
ation in the global data set; annual temperature explained
2%; and latitude explained 1%.
[19] For the analysis of the categorical variables, we
began by stratifying the global data set into three zones:
boreal, temperate and tropical. The stratification by zone
only explained 2% of the variation, but was significant (P =
0.0296) and the boreal, temperate and tropical soils had a
mean uptake of 2.38, 3.96 and 2.74 kg (CH4) ha 1yr 1,
respectively (Figure 2). The boreal zone was significantly
different from the temperate and tropical zones (P < 0.05), Figure 2. Data stratified by climatic zone. Box and
while the tropical and temperate zones were not different whisker plot of CH4 emissions in different climatological
from each other. The temperate zone had the highest zones. Boxes show 25th and 75th percentiles, and the
number of observations and the highest variability. All of median is indicated as a solid line within the boxes. The
the zones have several outliers, but the temperate zone whiskers show the 10th and 90th percentiles. Solid circles
showed the greatest variability with the greatest number indicate outliers; the dotted lines indicate the means.

4 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

further light on these results (Table 2). This scheme


explained 17% of the variation (P < 0.0001) and further
reduced the global estimate to 25 Tg yr 1. The temperate
forest class captured the highest degree of variability.
[23] We ran an ANOVA with a full model for these three
variables with interaction terms and the only significant
main effect was ecosystem (forest versus other). Two-way
interactions terms were all significant: climatic zone 
ecosystem type (P = 0.0707); ecosystem type  texture
(P = 0.0397); and climatic zone  texture (P = 0.0350). The
three-way interaction term was not significant. We then
tested a number of reduced models and found that a model
based on the ecosystem type plus the three two-way
interaction terms, provided the best alternative model,
where all of the effects in the model were significant at
the P < 0.05 level. This model explained 29% of the variation
and was highly significant (P < 0.0001). The presence of
Figure 3. Data stratified by ecosystem type. Box and significant interaction terms suggested that the best global
whisker plot of CH4 emissions in different ecosystems. estimate from this data set could be obtained from a full
Boxes show 25th and 75th percentiles, and the median is stratification by the three main categorical variables. On the
indicated as a solid line within the boxes. The whiskers basis of this analysis, we developed a global estimate of the
show the 10th and 90th percentiles. Solid circles indicate soil CH4 sink using a full stratification (Table 3).
outliers; the dotted lines indicate the means. [24] Our best estimate of the global soil sink is 22 Tg yr 1.
It is impossible to calculate properly the uncertainty around
this value because we do not know the uncertainty associ-
the highest number of observations (205) compared to the ated with the area estimates. If we account only for the
other ecosystems (113) and the highest variability. The uncertainty associated with the flux to calculate a standard
coefficients of variation were 104% in the forest and 93% error, the range of uncertainty would be between 10 and
for all other ecosystems combined. 34 Tg yr 1 for the full stratification.
[21] Finally, we stratified the global data set by soil
texture. Coarse textured soils had greater uptake than the 4. Discussion
other texture classes (Figure 4). The stratification by texture
4.1. Global Drivers of CH4 Sink Strength
explained 1.5% of the variation, but was not significant (P =
0.2043). Coarse texture soils had mean uptake of 4.19 kg [25] The single factor analyses indicated that the most
(CH4) ha 1yr 1, while fine and medium texture soils had important global driver of CH4 uptake rates is the type of
uptake of 1.98 and 3.31 kg (CH4) ha 1yr 1, respectively. ecosystem. Other authors have noted similar differences in
Organic soils, which occurred primarily in the boreal zone, temperate and tropical soils [Del Grosso et al., 2000;
had a mean flux of 3.05 kg (CH4) ha 1yr 1. The coefficients
of variation were between 69 and 128% for the texture classes
and increased with the number of observations.

3.3. Estimates of the Global Sink


[22] We calculated the effects of using each of these
variables to stratify the global data set on the estimated
global soil CH4 sink (Table 2). For comparison with other
stratification schemes, we also developed an estimate based
upon the unstratified data set (Table 2). The unstratified
estimate of 36 Tg yr 1 was greater than the stratified
estimates. Stratification by climatic zone or soil texture
gave a negligible reduction in the estimate, while stratifi-
cation by ecosystem reduced the global estimated sink
strength. In each one of the stratification schemes, the
largest part of the variation was captured in one stratum.
Stratification by climatic zone indicated that the highest
variability was in the temperate zone, while stratification by Figure 4. Data stratified by soil texture class. Box and
ecosystem showed high variability in forest ecosystems. whisker plot of CH4 emissions in soils of different texture.
Medium texture soils had the highest variability in the Boxes show 25th and 75th percentiles, and the median is
texture class stratification. For each stratification scheme, indicated as a solid line within the boxes. The whiskers
the highest variability was associated with the category of show the 10th and 90th percentiles. Solid circles indicate
the highest flux rate. A double stratification scheme shed outliers; the dotted lines indicate the means.

5 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

Table 2. Estimates of CH4 Uptake by Soils With Different Stratification Schemesa


Number of Mean, Area, Total Flux, SE of
Stratification Level Measures kgCH4 ha 1 yr 1
Variance  106 km2 Tg yr 1 Total Flux
None 318 3.3 14.8 108.20 35.6 23.3
Climatic zone
Boreal 65 2.4 7.5 20.83 5.0 7.1
Temperate 163 4.0 22.6 50.07 19.8 18.6
Tropical 90 2.7 4.4 37.30 10.2 8.3
Total 35.0 21.6
Ecosystem
Forest 205 4.2 19.3 42.01 17.7 12.9
Other 113 1.6 2.2 66.19 10.6 9.3
Total 28.3 15.9
Soil texture
Organic 54 3.1 9.2 2.14 0.7 0.9
Coarse 85 4.2 21.1 22.73 9.5 11.3
Medium 121 3.3 18.0 63.94 21.1 24.6
Fine 46 2.0 1.8 19.39 3.8 3.9
Total 35.2 27.4
Climatic zone + ecosystem
Boreal Forest 51 2.6 8.0 16.87 4.5 6.7
Other 14 1.4 4.3 3.96 0.6 2.2
Subtotal (zone) 5.1
Temperate Forest 92 5.7 31.5 5.82 3.3 3.4
Other 71 1.7 2.3 44.25 7.6 7.9
Subtotal (zone) 10.9
Tropical Forest 62 3.3 4.8 19.32 6.4 5.4
Other 28 1.4 1.2 17.98 2.6 3.7
Subtotal (zone) 9.0
Total 25.0 12.9
a
The first estimate shows mean and standard error estimates from the data set without stratification. The global data set is
then stratified by zone or ecosystem or by soil texture individually, and then by zone + ecosystem.

Verchot et al., 2000]. Uptake rates were consistently higher within soils, Verchot et al. [2000] concluded that texture class
in forests compared to all other ecosystems across all was more important than other factors at the biome scale.
climatic zones. Forest soils differ from the soil of other [27] Coarse texture soils make up about 21% of the
ecosystems by the presence of a forest floor. As a result, the world’s soils and contribute to only 22 percent of the global
upper layers of the soil profile have high organic matter sink. Thus, despite high CH4 uptake rates, their contribution
content, high microbial biomass, low bulk density and high to the global sink is proportionately low. Medium texture
porosity. Thus one would expect high consumption rates in soils make up about 59% of the world’s soils and contribute
these soil profiles. Biological activity is high, and with low 60% of the global sink. Organic soils are currently under
bulk density, diffusion of atmospheric CH4 into the forest represented in the global data set. The available data suggest
soil is rapid. This data set is not adequate to provide more that these soils make an important contribution only in the
mechanistic answers, but perhaps the high variances in boreal zone. Given the small area that these soils occupy in
forest ecosystems are due to factors associated with the other climatic zones, it is unlikely that this view will change
nature of the forest floor, which is largely determined by the with more data from these zones.
floristic composition of the forest ecosystem and by other [28] For the full stratification, temperate forests with
biotic factors such as earthworm activity [Finzi et al., 1998; coarse soils had the highest uptake CH4 values (7.5 kg
Finzi and Schlesinger, 2002; Carreiro et al., 2000; Verchot CH4 ha 1 yr 1), and a high variance. This value is higher
et al., 2001]. than the median range of 1.6 6.4 kg CH4 ha 1 yr 1
[26] The fact that soil texture accounted for an important suggested by Smith et al. [2000], although our value was
part of the variation in the global data set is consistent with based upon the mean rather than the median. Interestingly,
current thinking on the biogeochemical controls of CH4 the soil texture effect seems to be important only in temperate
oxidation in soils [Striegl, 1993; Dörr et al., 1993; Del forests, which explains the interaction terms in the full rank
Grosso et al., 2000]. Verchot et al. [2000] summarized the model. In other ecosystems, CH4 uptake on coarse texture
data for forest sites in the neotropics and noted that fine- soils is similar to uptake rates in other texture classes.
textured soils generally consumed between 1.5 and 2.0 kg [29] Seasonal patterns of soil water content closely follow
(CH4) ha 1yr 1, while medium and coarse textured soils the pattern of rainfall and several authors have noted that
consumed >4.0 kg (CH4) ha 1yr 1. Other authors have CH4 uptake also follows this pattern [Verchot et al., 2000;
shown that coarse textured soils have higher oxidizing Kiese et al., 2003]. A simplistic extension of this observa-
capacity than fine textured soils [see, e.g., Boeckx et al., tion suggests that annual rainfall might be a global predictor
1997]. As Striegl [1993] noted, because gas transport of O2 of CH4 uptake because soils are drier for longer periods in
and CH4 are the most important factors affecting CH4 low rainfall areas. This of course ignores the complex
fluxes, and soil texture strongly affects diffusivity of gases interaction between soil, vegetation type, and water avail-
ability on the size and nature of the soil microbial commu-

6 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

Table 3. Estimates of CH4 Uptake by Soils With a Zone Plus Ecosystem Plus Soil Texture Stratification
Scheme
Number of Mean, Area, Total Flux, SE of
Stratification Level Measures kgCH4 ha 1 yr 1
Variance  106 km2 Tg yr 1 Total Flux
Climatic zone + ecosystem + soil texture
Boreal Forest Organic 28 3.6 10.0 1.76 0.6 1.1
Coarse 6 3.2 7.6 3.06 1.0 3.4
Medium 14 0.5 0.2 11.18 0.6 1.4
Fine 1 2.0 – 0.87 0.2 –
Sub-subtotal (zone + ecosystem) 2.4
Boreal Other Organic 8 1.7 7.7 0.05 0.0 0.1
Coarse 3 1.0 0.1 1.08 0.1 0.2
Medium 3 1.0 0.2 2.77 0.3 0.7
Fine – – – 0.06 – –
Sub-subtotal (zone + ecosystem) 0.4
Subtotal (zone) 2.8
Temperate Forest Organic 7 4.6 10.3 0.08 0.0 0.1
Coarse 30 7.5 34.0 1.00 0.7 1.1
Medium 43 5.6 37.1 3.60 2.0 3.4
Fine 12 2.3 1.5 1.14 0.3 0.4
Sub-subtotal (zone + ecosystem) 3.0
Temperate Other Organic 9 1.4 1.5 0.10 0.0 0.0
Coarse 18 1.7 1.7 10.35 1.8 3.2
Medium 36 1.7 2.1 28.24 4.9 6.8
Fine 6 1.2 4.6 5.56 0.7 4.9
Sub-subtotal (zone + ecosystem) 7.4
Subtotal (zone) 10.4
Tropical Forest Organic 2 3.6 – 0.12 0.0 –
Coarse 16 4.3 6.8 2.43 1.0 1.6
Medium 19 3.8 3.8 9.32 3.6 4.2
Fine 21 2.1 1.5 7.45 1.6 2.0
Sub-subtotal (zone + ecosystem) 6.2
Tropical Other Organic 0 0.03
Coarse 12 0.9 1.4 4.81 0.4 1.6
Medium 6 2.3 0.6 8.83 2.0 2.7
Fine 6 1.5 1.1 4.31 0.6 1.9
Sub-subtotal (zone + ecosystem) 3.0
Subtotal (zone) 9.2
Total for the full stratification 22.4 12.1

nity. We did find a significant linear relationship between some of the variation among biomes. In Table 3, for
CH4 uptake rates and total annual rainfall, but the relation- example, subtotals for the total temperate zone flux was
ship had low predictive value and was positive. It appears 10.4 Tg yr 1, which was much lower than the estimate of
that this relationship is not explaining the effects of physical 19.8 Tg yr 1 derived from the simple climatic zone strat-
restriction of water in soils on diffusion, but is an indicator ification (Table 2). When there was no large source of
of the biological controls of the soil microbial community variation within a climatic stratum, the nature of the sub-
on uptake. stratification had little bearing on the estimate. For example,
the total flux estimated for the tropical zone in the simple
4.2. Success of Stratification climatic zone stratification is similar to that estimated by the
[30] The question that remains is whether the stratification climatic zone + ecosystem + soil texture stratification.
that we used was successful in providing a better estimate of
the sink strength. Stratification reduced the variance in the 4.3. Comparison With Other Estimates
majority of strata each time. In the stratification by climate [32] Estimates of the global soil sink for atmospheric CH4
zones, the variances in the boreal and tropical strata were have been made by a number of authors and fall between 17
lower than the global variance, but variance in the temperate and 44 Tg yr 1 (Table 4). These efforts have involved both
stratum was larger than the global variance. In the stratifi- extrapolation of field studies and modeling. The modeling
cation by climatic zone + ecosystem, the variances were efforts have attempted to capture the effects of key biogeo-
lower in all strata except for the temperate forest. chemical drivers on soil CH4 consumption, while most of
[31] Thus stratification had a significant effect on the the stratify-and-multiply approaches have focused on cli-
estimate of the magnitude of the terrestrial sink strength matic zone and ecosystem effects. The one exception to this
by allowing us to compartmentalize the largest source of is the estimate of Dörr et al. [1993], who based their
variation. More importantly, stratification compartmental- extrapolation on soil texture stratification. These authors
ized the variance and provided narrower standard error used the results of intensive measurements made at five
ranges. In this sense, the stratification has been successful sites in southern Germany and spot measurements at a wide
in producing a more realistic estimate by accounting for range of sites in Central Europe and in different parts of the

7 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

Table 4. Estimates of the Global Soil Methane Sink From predicted that the highest uptake rates would occur in dry
Different Sources Compared With Our Estimation Obtained tropical ecosystems, whereas the experimental data summa-
Without and With Stratification rized here suggests that forest ecosystems have the highest
CH4 Sink, uptake rates.
1
Reference Tg CH4 yr
From modeled estimates 5. Conclusion
Potter et al. [1996] 17
Ridgwell et al. [1999] 38 [36] The increase in available data from around the world
From extrapolation estimates over the past several years has allowed us to produce a more
Dörr et al. [1993] 29 well constrained estimate of the global soil CH4 sink than
Mosier et al. [1998] 44
Potter et al. [1996] 21 has been possible previously. We used 318 data represen-
Smith et al. [2000] 29 tative of a wide variety of ecosystems, climates and soil
Intergovernmental Panel on Climate Change [2001] 30 types, and produced a number of estimates based upon
Our estimate without stratification 36 different stratification schemes that cover the lower to
Our estimate with full stratification 22
middle of the range of the other estimates. Our estimate
of 22 ± 12 Tg yr 1 CH4 is toward the lower end of the
accepted range for the soil sink.
[37] Just as for process based modeling estimates, this
world to calibrate the relationship between texture and CH4 study shows that incorporation of mechanistic elements into
consumption. They produced an estimate of the global soil the stratification scheme improves the global estimate and
sink assuming that this parameterization was valid for soils reduces uncertainty in extrapolations from field data. With-
across the globe. out stratification, the range of the global estimate was
[33] Other authors have produced estimates of the global between 12 and 60 and the best simple stratification scheme
soil sink. Mosier et al. [1998] derived global CH4 uptake produced a range of 14 to 57. The standard errors of the
from the references given by Reeburgh et al. [1993] and the different strata point to the opportunities to improve further
land use data of Bouwman [1990]. They estimated CH4 the estimates of the soil sink. In the boreal zone, the
uptake rates for 10 land cover types. A major difference standard error is highest in forests with coarse texture soils
between their estimates and ours is the difference in land and the number of observations on these soils was low. In
cover definitions between the sources of land cover data. the temperate and tropical forests, highest standard errors
Smith et al. [2000] derived an estimate based upon data were associated with medium texture soils. Both of these
from 109 sites and estimated that while the mean was 29 Tg strata have high flux rates and there were a large number of
yr1, the range of uncertainty was very large, between 7 and observations, which points to the need for more targeted
109 Tg yr 1. These authors used the standard deviation of mechanistic work. Nonforest ecosystems are not well rep-
the transformed data to estimate the range of uncertainty resented in the data set and, at least in the temperate zone,
associated with the estimate rather than standard error of the the standard errors associated with these estimates are
mean, which explains the wide range in their reported relatively high. Clearly more observations in these situa-
uncertainty compared to our narrower range. The standard tions are warranted. Thus efforts to better account for the
error would have provided a more realistic uncertainty variability within these strata will make the greatest contri-
estimate. bution to improving the global sink estimate.
[34] Potter et al. [1996] made both a modeling estimate [38] This study points to a number of gaps in the global
and an extrapolation estimate. For the extrapolation esti- data set. One of the key findings is that a number of
mate, forest tundra was not included because it had been important ecosystems continue to be underrepresented in
characterized in the literature as a variably wet environment the global data set. In particular, grasslands in both the
where high methane production can occur during thaw temperate and tropical zones are underrepresented, as are
season, resulting in net annual CH4 emissions to the chaparral ecosystems. Desert ecosystems are also underrep-
atmosphere. Data availability was also a problem for this resented, and although fluxes are low in these ecosystems,
extrapolation estimate. For six of the 14 major ecosystems they represent a large portion of the earth’s surface. Thus
where they produced an estimate, three or fewer site data efforts to fill these gaps should be a high priority. Perhaps
sets could be used for the extrapolation to annual fluxes. highest priority should be given to grasslands. Any assess-
Over one half of the measured flux data sets came from ment of the global soil sink may be seriously in error until
temperate sites. Making an estimate for tropical grasslands more representative data for the whole land surface are
and savannas was also problematic, owing to variability in available.
soil condition associated with fires and human management, [39] Finally, this study raises questions about the sources
and the fact that several authors reported that CH4 emission of variability in temperate forests soils. The effects of soil
was present at some times during the year. texture were most pronounced in temperate forests and high
[35] The model described by Ridgwell et al. [1999] also uptake rates were associated with high variances. We do not
takes a mechanistic approach to determine the global understand the source of the large degree of variability in
consumption of CH4 by soils incorporating descriptions of these soils, and more mechanistic work to elucidate the
the most important controlling factors related to both underlying processes will be helpful.
gaseous diffusion and microbial activity. Ridgwell’s model

8 of 9
GB4013 DUTAUR AND VERCHOT: A GLOBAL INVENTORY OF THE SOIL CH4 SINK GB4013

[40] Acknowledgments. Support for this research was provided by Mosier, A. R., W. J. Parton, D. W. Valentine, D. S. Ojima, D. S. Schimel,
the International Centre for Research in Agroforestry. We would also like to and O. Heinemeyer (1997), CH4 and N2O fluxes in Colorado shortgrass
thank two anonymous reviewers who provided comments on an earlier draft steppe: 2. Long-term impact of land use change, Global Biogeochem.
of this paper. Cycles, 11, 29 – 42.
Mosier, A. R., J. M. Duxbury, J. R. Freney, O. Heinemeyer, K. Minami, and
References D. E. Johnson (1998), Mitigating agricultural emissions of methane,
Clim. Change, 40, 39 – 80.
Boeckx, P., and O. Van Cleemput (2001), Estimates of N2O and CH4 fluxes Nedwell, D. B. (1996), Methane production and oxidation in soils and
from agricultural lands in various regions in Europe, Nutr. Cycling Agroe- sediments, in Microbiology of Atmospheric Trace Gases, edited by J. C.
cosyst., 60, 35 – 47. Murell and D. P. Kelly pp. 33 – 50, Springer, New York.
Boeckx, P., O. Van Cleemput, and I. Villaralvo (1997), Methane oxidation Ojima, D. S., D. W. Valentine, A. R. Mosier, W. J. Parton, and D. S.
in soils with different textures and land use, Nutr. Cycling Agroecosyst., Schimel (1993), Effect of land use change on methane oxidation in
49, 91 – 95. temperate forest and grasslands, Chemosphere, 26, 675 – 685.
Borken, W., and F. Beese (2006), Methane and nitrous oxide fluxes of soils Parkin, T. B., and J. A. Robinson (1994), Statistical treatment of microbial
in pure and mixed stands of European beech and Norway spruce, Eur. data, in Methods of Soil Analysis: Part 2. Microbial and Biochemical
J. Soil Sci., 57, 617 – 625. Properties, edited by R. W. Weaver et al., pp. 15 – 39, Soil Sci. Soc. of
Bouwman, A. F. (1990), Exchange of greenhouse gases between terrestrial Am., Madison, Wis.
ecosystems and the atmosphere, Soils and the Greenhouse Effect, edited Potter, C. S., E. A. Davidson, and L. V. Verchot (1996), Estimation of
by A. F. Bouwman, pp. 61 – 127, John Wiley, Hoboken, N. J. global biogeochemical controls and seasonality in soil methane consump-
Brady, N. C. (1974), The Nature and Property of Soils, Macmillan, New tion, Chemosphere, 32, 2219 – 2246.
York. Reeburgh, W. S., S. C. Whalen, and M. J. Alperin (1993), The role of
Carreiro, M. M., R. L. Sinsabaugh, D. A. Repert, and D. F. Parkhurst methyltrophy in the global CH4 budget, in Microbial Growth on C1
(2000), Microbial enzyme shifts explain litter decay responses to simu- Compounds, edited by J. C. Murell and D. P. Kelley, pp. 1 – 14, Intercept,
lated nitrogen deposition, Ecology, 81(9), 2359 – 2365. Andover, U. K.
Conrad, R. (1989), Control of methane production in terrestrial ecosystems, Ridgwell, A., S. J. Marshall, and W. F. J. Gregson (1999), Consumption of
Exchange of Trace Gases Between Terrestrial Ecosystems and the Atmo- atmospheric methane by soils: a process-based model, Global Biogeo-
sphere: Dahlem Workshop Reports, Life Sci. Res. Rep., vol. 47, edited by chem. Cycles, 13, 59 – 70.
M. O. Andreae and D. S. Schimel, pp. 39 – 58, John Wiley, Hoboken, N. J. SAS Institute (2005), SAS/STAT Users Guide, Cary, N. C.
Del Grosso, S. J., et al. (2000), General CH4 oxidation model and compar- Singh, J. S., S. Singh, A. S. Raghubanhi, S. Singh, A. K. Kashyap, and V. S.
isons of CH4 oxidation in natural and managed systems, Global Biogeo- Reddy (1997), Effects of soil nitrogen, carbon and moisture on methane
chem. Cycles, 14, 999 – 1019. uptake by dry tropical forest soils, Plant Soil, 196, 115 – 121.
Dlugokencky, E. (2001), NOAA CMDL Carbon Cycle Greenhouse Gases: Smith, K. A., et al. (2000), Oxidation of atmospheric methane in northern
Global average atmospheric methane mixing ratios, NOAA CMDL co- European soils, comparison with other ecosystems, and uncertainties in
operative air sampling network, Natl. Oceanic and Atmos. Admin., Silver the global terrestrial sink, Global Change Biol., 6, 791 – 803.
Spring, Md. (Available at http://www.cmdl.noaa.gov/ccg/figures/ Striegl, R. G. (1993), Diffusional limits to the consumption of atmospheric
ch4trend_global.gif) methane by soils, Chemosphere, 26, 715 – 720.
Dörr, H., L. Katruff, and I. Levin (1993), Soil texture parameterization of Striegl, R. G., T. A. McConnaughey, D. C. Thorstenson, E. P. Weeks, and
the methane uptake in aerated soils, Chemosphere, 26, 697 – 713. J. C. Woodward (1992), Consumption of atmospheric methane by desert
Finzi, A. C., and W. H. Schlesinger (2002), Species control variation in soils, Nature, 357, 145 – 147.
litter decomposition in a pine forest exposed to elevated CO2, Global Topp, E., and E. Pattey (1997), Soils as sources and sinks for atmospheric
Change Biol., 8, 1217 – 1229. methane, Can. J. Soil Sci., 77, 167 – 178.
Finzi, A. C., N. Van Breemen, and C. D. Canham (1998), Canopy tree-soil Verchot, L. V., E. A. Davidson, J. H. Cattânio, and I. L. Ackerman (2000),
interactions within temperate forests: Species effects on soil carbon and Land-use change and biogeochemical controls of methane fluxes in soils
nitrogen, Ecol. Appl., 8(2), 440 – 446. of eastern Amazonia, Ecosystems, 3, 41 – 56.
Intergovernmental Panel on Climate Change (2001), Climate Change 2001: Verchot, L. V., Z. Holmes, L. Mulon, P. M. Groffman, and G. M. Lovett
The scientific Basis—Contribution of Working Group I to the Third (2001), Gross versus net rates of N mineralization and nitrification as
Assessment Report of the Intergovernmental Panel on Climate Change, indicators of functional differences between forest types, Soil Biol. Bio-
edited by J. T. Houghton et al., 881 pp., Cambridge Univ. Press, chem., 33, 1889 – 1901.
New York.
Kiese, R., B. Hewett, A. Graham, and K. Butterbach-Bahl (2003), Seasonal
variability of N2O emissions and CH4 uptake by tropical rainforest soils
of Queensland, Australia, Global Biogeochem. Cycles, 17(2), 1043, L. Dutaur and L. V. Verchot, International Centre for Research in
doi:10.1029/2002GB002014. Agroforestry, United Nations Avenue, P.O. Box 30677-00100, Nairobi,
Leemans, R. (1990), Possible change in natural vegetation patterns due to a Kenya. (l.verchot@cgiar.org)
global warming, Laxenburg Working Pap. WP-08, 22 pp., Int. Inst. for
Appl. Syst. Anal., Laxenburg, Austria.
Le Mer, J., and P. Roger (2001), Production, oxidation, emission and con-
sumption of methane by soils: A review, Eur. J. Soil Biol., 37, 25 – 50.

9 of 9

You might also like