You are on page 1of 8

Environmental Microbiology Reports (2009) 1(5), 285–292 doi:10.1111/j.1758-2229.2009.00038.

Minireview

The global methane cycle: recent advances in


understanding the microbial processes involved emi4_038 285..292

Ralf Conrad* concentration (Bousquet et al., 2006). The global budget


Max-Planck-Institute for Terrestrial Microbiology, of atmospheric methane is on the order of 500–600 Tg
Karl-von-Frisch-Str.8, 35043 Marburg, Germany. CH4 year-1 (Lelieveld et al., 1998; Wang et al., 2004). The
individual sources contributing to this budget are shown
as a pie chart in Fig. 1.
Summary
About 25% of all CH4 sources are associated with
The global budget of atmospheric CH4, which is on mining and combustion of fossil fuels or with burning of
the order of 500–600 Tg CH4 per year, is mainly the biomass. About 69% of the sources is the result of micro-
result of environmental microbial processes, such bial processes and another 6% is due to chemical pro-
as archaeal methanogenesis in wetlands, rice fields, duction of CH4 from plant material. This overview shows
ruminant and termite digestive systems and of micro- that most of the atmospheric CH4 originates from micro-
bial methane oxidation under anoxic and oxic condi- bial metabolism, more precisely from CH4 production by
tions. This review highlights recent progress in the methanogenic archaea. Methanogenic archaea and CH4
research of anaerobic CH4 oxidation, of CH4 pro- production is typically found at those sites where organic
duction in the plant rhizosphere, of CH4 serving as matter is decomposed in the absence of oxygen or of
substrate for the aquatic trophic food chain and other oxidants, such as nitrate, sulfate or ferric iron. Wet-
the discovery of novel aerobic methanotrophs. It also lands are the largest individual source of microbial CH4.
emphasizes progress and deficiencies in our knowl- There is comparatively little direct influence of humans
edge of microbial utilization of low atmospheric CH4 on the CH4 source strength of wetlands, at least little in
concentrations in soil, CH4 production in the plant comparison with flooded rice fields, which are anthro-
canopy, intestinal methanogenesis and CH4 produc- pogenic wetlands. Rumen fermentation in cattle, sheep
tion in pelagic water. and other ruminants are another important individual CH4
source that is under human management. Further anthro-
pogenic sources are anaerobic digestion plants and land-
The global methane cycle
fills, in which waste materials are degraded to produce
Methane is the second most important anthropogenic large quantities of CH4. Intestinal fermentation in termites,
greenhouse gas after CO2, contributing about 30% to CH4 production in the water column of ocean and CH4
the total net anthropogenic radiative forcing of 1.6 W m-2 release from gas hydrates are comparatively minor but
(IPCC WG I fourth assessment 2008; http://www.ipcc. significant sources of atmospheric CH4, all without much
ch/pdf/assessment-report/ar4/syr/ar4_syr_spm.pdf). The direct human impact.
concentration of CH4 in the atmosphere has been increas- The global CH4 sources are balanced by sinks of similar
ing from pre-industrial values of about 715 ppbv to cur- magnitude. In fact, the total sink strength has for long time
rently about 1770 ppbv. The growth rate of atmospheric been a bit smaller than the total source strength, thus
CH4, which is determined by the balance of sources and causing the steady growth in atmospheric CH4 concen-
sinks, was about 12 ppb year-1 in the 1980s, but has tration since pre-industrial times. Fortunately, the CH4 sink
decreased since the early 1990s and with a value of about strength increases proportionally with the increasing CH4
4 ppb year-1 since 1999 (Bousquet et al., 2006). However, concentration in the atmosphere (quasi first-order reac-
there are indications of large year-to-year fluctuations tion) thus neutralizing changes in the source strength at
of sources and sinks, which eventually could result least partially. The lifetime of a CH4 molecule in the atmo-
in resumption of the growth of the atmospheric CH4 sphere is only on the order of 8 years (Lelieveld et al.,
1998; Wang et al., 2004). The largest sink (> 80% of the
Received 1 April, 2009; accepted 18 May, 2009. *For correspon-
dence. E-mail conrad@mpi-marburg.mpg.de; Tel. (+49) 6421 total) is the photochemical oxidation of CH4 initiated by
178801; Fax (+49) 6421 178809. the reaction with OH radicals. The other two sinks are
© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd
286 R. Conrad

plants Anaerobic methane oxidation


6% wetlands
18%
termites natural Existence of anaerobic oxidation of CH4 has been postu-
23% ocean lated for a long time in order to explain concentration
7% gas hydrates
microbial gradients and isotopic signatures of CH4 in marine sedi-
rice fields
4% ruminants
ments. Furthermore, it has been possible for long time to
7% 3% landfills measure the conversion of 14CH4 to 14CO2 in environ-
3% sewage treatment mental samples, which usually were of marine origin. It
17% 2% biomass burning has been postulated that CH4 is oxidized by reduction
10% fossil fuel of sulfate and that this would possibly be carried out
by CH4-oxidizing H2-producing archaea syntrophically
Fig. 1. Global methane sources in per cent of the total budget of
associated with H2-consuming sulfate-reducing bacteria
about 500–600 Tg CH4 per year.
(Valentine and Reeburgh, 2000; Reeburgh, 2007). Such
a process was shown to be consistent with the ther-
diffusion into the stratosphere and microbial CH4 oxidation modynamic conditions in marine sediments, in particular
in soil. with the H2 partial pressures (Hoehler et al., 1994). The
It should be noted that the strengths of these sources or microbes involved in such a process were for the first time
sinks represent net fluxes between the planetary surface visualized by applying fluorescent in situ hybridization
and the atmosphere. Gross fluxes can be much larger. (FISH) with nucleic acid probes specific for archaea and
In rice fields, for example, much (about 20%) of the pro- for sulfate-reducing bacteria (Boetius et al., 2000). Using
duced CH4 is consumed by aerobic methanotrophs living FISH coupled to secondary ion mass spectrometry
in the rhizosphere of the rice plants, thus attenuating the (SIMS) it was possible to demonstrate that the archaea
net flux into the atmosphere (Schütz et al., 1989; Frenzel, indeed were highly depleted in 13C as expected when
2000; Groot et al., 2003). For most CH4 sources shown in utilizing CH4 as carbon source (Orphan et al., 2001).
Fig. 1, rates of CH4 production are usually much larger It was shown that these consortia consist of archaea,
than rates of CH4 emission since a large fraction of the which form the phylogenetically distinct cluster ANME-2
initially produced CH4 is consumed by microorganisms within the methanogenic order Methanosarcinales, and
before it can enter the atmosphere (Frenzel, 2000; of sulfate-reducing bacteria from the Desulfosarcina–
Reeburgh, 2003). Desulfococcus branch of Deltaproteobacteria. Subse-
The quantitatively most important example is CH4 quently similar consortia were found with archaea of the
emission from gas hydrates, which presently constitutes clusters ANME-1, which is distantly related to Methanosa-
only about 1% of the total atmospheric CH4 budget or rcinales and Methanomicrobiales (Orphan et al., 2002;
about 5–6 Tg CH4 year-1. However, the total reservoir of Knittel et al., 2005), and ANME-3, which is related to
gas hydrates in the marine sediments of the continental the genera Methanococoides and Methanolobus within
margins is huge, on the order of 500 000–10 000 000 Tg the Methanosarcinales (Lösekann et al., 2007). ANME
CH4, or about 150–3000 the amount of CH4 in the atmo- archaea contain gene homologues of the methyl-
sphere (Reeburgh, 2007). In fact, mobilization of CH4 from coenzyme M reductase, the key enzyme of CH4 synthe-
the gas hydrates is much larger than the net emission sis, suggesting that a similar enzyme may be involved
into the atmosphere, but most of the mobilized CH4 is in anaerobic oxidation of CH4 to CO2 (Hallam et al., 2004).
fortunately consumed by anaerobic methane oxidizers, Indeed, two nickel proteins resembling cofactor 430 from
which globally account for a consumption rate of about the methyl-coenzyme M reductase were isolated in rela-
70–300 Tg CH4 year-1 (Reeburgh, 2007). Without this tively high abundance (10% of total protein) from micro-
consumption activity, the atmospheric CH4 budget would bial mats exhibiting anaerobic CH4 oxidation (Krüger
be higher by 10–60% resulting in dramatically higher et al., 2003). However, the biochemical mechanism of
concentrations of atmospheric CH4. the enzymatic oxidation reaction is still unclear (Thauer
There are several comprehensive reviews of the global and Shima, 2008).
CH4 cycle and the environmental microbiology of metha- The CH4-oxidizing archaea of the ANME clusters have
nogenic and methanotrophic microorganisms (Reeburgh, so far only been grown in enrichment cultures with very
2003; Conrad, 2007; Dunfield, 2007; Reeburgh, 2007; Liu low growth rates (doubling times of weeks to months) and
and Whitman, 2008; Thauer and Shima, 2008; Trotsenko low growth yields (only 1% of the CH4 oxidized is used
and Murrell, 2008). In this minireview, I will therefore con- for biomass production), but have not yet been isolated
centrate on some recent advances in the microbiology (Girguis et al., 2005; Nauhaus et al., 2007). It is also
of the CH4 cycle and discuss some problems that are still debatable whether these archaea really need sulfate
unsolved. reducers as syntrophic partners or can reduce sulfate by

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292
Global methane cycle 287

themselves (Thauer and Shima, 2008). The idea of CH4 2.0–3.5, temperature optimum 55–60°C) and have
oxidation within one sole microbial species stems from recently been placed into the new genus Methylacidi-
another recent discovery, the anaerobic oxidation of CH4 philum (OpdenCamp et al., 2009). This was the first dis-
with nitrate as electron acceptor (Raghoebarsing et al., covery that methanotrophic growth is not restricted to
2006). Methane oxidation coupled to denitrification was the Proteobacteria, but also exists in another bacterial
discovered in an anaerobic enrichment culture from a phylum. Future research will show how broadly distributed
canal sediment. This enrichment culture consisted of 10% methanotrophy really is among prokaryotes.
archaea of the order Methanosarcinales distantly related It has been a paradigm that methanotrophic bacteria
to ANME-2, the rest being bacteria. Hence, it was plau- are obligately methylotrophic, i.e. can only utilize C1 com-
sible to assume that anoxic CH4 oxidation with nitrate pounds. This paradigm was recently disproved by the
would operate similarly as with sulfate, i.e. the archaea discovery that Methylocella sp. are facultative hetero-
oxidize CH4 to CO2 and transfer the electrons to the deni- trophs, which can also utilize acetate as energy source
trifiers reducing nitrate or nitrite to N2. However, it turned and actually prefer this substrate over CH4 (Dedysh et al.,
out that the oxidation of CH4 is catalysed exclusively by 2005). Methylocella is the only methanotrophic genus,
the denitrifiers without participation of the archaea (Ettwig which contains sMMO but no pMMO. Acetate was found
et al., 2008). Hence, we may expect an interesting bio- to suppress transcription of sMMO and thus CH4 oxidation
chemical mechanism for CH4 oxidation with oxidized nitro- (Theisen et al., 2005). However, since acetate stimulates
gen compounds. Interestingly, the analogous oxidation growth of Methylocella, acetate may have a positive effect
of ammonia can also be initiated with NO2 instead of O2 on the longer term through increase of the population size
(Schmidt and Bock, 1997; Schmidt et al., 2004). However, of methanotrophs. In fact, such effect has been observed
it is unclear how NO2 could be formed in the absence of with atmospheric CH4 oxidation in alpine tundra soil,
O2. The mechanism of CH4 oxidation in the denitrifiers but has been attributed to indirect stimulation through
is presently investigated by Jetten and co-workers. CH4 production by acetoclastic methanogens (West and
Schmidt, 1999). It would be interesting to see whether
there exist more facultatively heterotrophic methanotro-
Novel aerobic methanotrophs
phs in nature and which effect this has on CH4 cycling.
Aerobic methanotrophic bacteria are widely distributed
in nature, typically at anoxic–oxic interfaces, such as sedi-
Utilization of low methane concentrations
ments, where they attenuate the flux of CH4 into the
atmosphere. Canonical aerobic methanotrophs belong to Soils act as a sink for atmospheric CH4. The atmospheric
either the Gammaproteobacteria or type I methanotrophs concentrations of CH4 typically are 1.7 ppmv, after disso-
(10 different genera) or the Alphaproteobacteria or type lution in water about 2 nM CH4. This is a very poor energy
II methanotrophs (the genera Methylocystis, Methylo- source for methanotrophic bacteria and demands that
sinus, Methylocella, Methylocapsa). Type I methano- the enzymes of these bacteria have a high affinity for
trophs assimilate C via the ribulose monophosphate the substrate (Conrad, 1984). Indeed, such high-affinity
pathway, while type II methanotrophs use the serine methanotrophs actually exist in soil (Bender and Conrad,
pathway (Trotsenko and Murrell, 2008). A few years ago, 1992), but their phylogenetic affiliation is not exactly
the filamentous bacterium Crenothrix polyspora was known. Uptake of CH4 from the atmosphere into the soil
found to be methanotrophic (Stoecker et al., 2006). Simi- is usually associated with the detection of particular
larly, Clonothrix fusca was described as a methanotroph sequences of the pmoA gene (coding for a subunit of the
(Vigliotta et al., 2007). These were surprising discoveries, pMMO) (Holmes et al., 1999; Henckel et al., 2000; Knief
since these filamentous bacteria had not been known to et al., 2003). Many of these sequences, e.g. those of
be able to consume CH4. However, both species belong to USCa, are closely related to the pmoA of Methylocapsa
the Gammaproteobacteria and are closely related to the acidiphila (Ricke et al., 2005), but isolates of these high-
type I methanotrophs. Molecular techniques useful to affinity methanotrophs still do not exist.
detect these methanotrophs in the environment are based Attempts to isolate high-affinity methanotrophs so far
on characteristic sequences of the 16S rRNA gene and resulted only in the retrieval of Methylocystis sp. (Dunfield
of functional genes coding for subunits of the particulate et al., 1999), which express high affinity for CH4 under
(pMMO) or soluble methane monooxygenase (sMMO), starvation conditions (Dunfield and Conrad, 2000). Many
and have recently been reviewed (McDonald et al., of the Methylocystis species contain two types of pMMO,
2008). Very recently, methanotrophs within the phylum the conventional enzyme and a second one (Ricke et al.,
Verrucomicrobia were discovered in hot environments 2004). It has now been shown that the second pMMO
(Dunfield et al., 2007; Pol et al., 2007; Islam et al., 2008). is specifically required for growth and maintenance
These isolates are thermoacidophilic (pH optimum on low CH4 concentrations (10–100 ppmv), while the

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292
288 R. Conrad

conventional pMMO is only active at high CH4 concentra- plant microcosms (Keppler et al., 2006). The resulting
tions (> 600 ppmv) (Baani and Liesack, 2008). Since the emission of CH4 into the atmosphere, when extrapolated
apparent Km values (i.e. 0.11 mM) for CH4 of these Methy- to the global scale, is significant (Fig. 1) and could be
locystis species are at the higher end of those found in soil the additional CH4 source in tropical regions that had
environments, and since they have been found in hydro- been predicted from satellite measurements and model
morphic soils in high abundance (Knief et al., 2006), it calculations (Frankenberg et al., 2005). Keppler and
is hypothesized that they may be specifically adapted to co-workers claimed that CH4 is produced by aerobic
environments where CH4 concentrations change between chemical reactions in the plant leaves, for example by
high and low values. degradation of pectin. The potential of plants to aerobi-
cally produce and emit CH4 has been discussed contro-
versially, in particular since some researchers were
Methane production in the rice rhizosphere
unable to reproduce the original results (Dueck et al.,
Rice fields and natural wetlands are an important source 2007; Beerling et al., 2008). However, the crucial factor
in the global CH4 cycle. In rice fields, up to 60% of CH4 seems to be illumination of plant leaves with UV instead
emission is due to decomposition of root exudates or of photosynthetically active radiation. UV radiation indeed
dead roots both originating from recently assimilated resulted in substantial CH4 production from leaf material
CO2 (Watanabe et al., 1999). This was shown using pulse and also from pectin (McLeod et al., 2008; Vigano et al.,
labelling of rice plants with 13C-CO2 which resulted in the 2008).
production and emission of a proportional amount of 13C- Nevertheless, there are additional processes in
CH4. The same approach was used to label the metha- plants that can result in the emission of CH4 through the
nogenic archaea in the rhizosphere that incorporated plant canopy. The best characterized is the transport
the photosynthetically fixed carbon into their cells using of CH4 through the plant gas vascular (aerenchyma)
the stable isotope probing technique (Radajewski et al., system, which accounts for most of the flux from anoxic
2000). As a result, Rice Cluster I methanogens were soils vegetated with aquatic plants (Schütz et al., 1989;
identified as the major group responsible for CH4 produc- Frenzel, 2000). In addition, CH4 can be transported in the
tion in the rice rhizosphere (Lu and Conrad, 2005). These dissolved state from the anoxic soil via the transpiration
methanogens, which produce CH4 by reduction of CO2, stream into canopy. Such CH4 transport indeed exists, as
seem to be specifically adapted to this environment, since has been shown by studies on rice plants (Nouchi et al.,
replacement of the rhizospheric population with Metha- 1990) and trees (Rusch and Rennenberg, 1998; Ter-
nomicrobiales resulted in lower population densities on azawa et al., 2007), but CH4 flux by transpiration is always
the roots and in much lower rates of CH4 production and orders of magnitude lower than the CH4 transport through
emission (Conrad et al., 2008). Rice Cluster I metha- the plant aerenchyma system. Finally, CH4 can be pro-
nogens have been known as a phylogenetic group of duced inside of living tree stems by anaerobic decom-
Archaea since about 10 years ago (Grosskopf et al., position of wetwood (Zeikus and Ward, 1974), which
1998), but attempts to isolate methanogens from this involves a methanogenic community that degrades wood,
cluster have not been successful until recently. Prior to including pectin (Schink et al., 1981). The CH4 accumu-
isolation, the entire genome sequence of a Rice Cluster I lating inside of the stems is often under high pressure
methanogen has been retrieved from an enrichment (Zeikus and Ward, 1974), but it is not known to what
culture and reassembled by metagenomics (Erkel et al., extent this CH4 can be emitted through the plant canopy.
2006). The sequence data indicated that these methano- The emission from plants by the transpiration stream or by
gens may be specifically adapted to the environmental release from wetwood has never been studied explicitly
conditions in the rhizosphere, especially to the presence with respect to the role in the global CH4 budget.
of O2, by having a unique set of antioxidant and
O2-insensitive enzymes. Now, a first isolate of Rice
Methane production in intestinal systems
Cluster I methanogens exists (Sakai et al., 2007), which is
described as Methanocella paludicola representing the Methane emission by ruminants and termites is a large
new order of Methanocellales (Sakai et al., 2008). source in the budget of atmospheric CH4 (Fig. 1). The
microbiology and methanogenic processes in these habi-
tats have recently been reviewed in detail (Purdy, 2007;
Methane production and emission by plants
Janssen and Kirs, 2008; Liu and Whitman, 2008; Brune,
While microbial CH4 production in the rhizosphere of 2009). In both rumen and hindgut of insects (termites,
aquatic plants is a common process, production within cockroaches, scarab beetles), CH4 seems to be exclu-
the plant canopy is unexpected. Nevertheless, exactly this sively produced from H2 + CO2 and its activity seems to
was found recently during incubation experiments using be limited by the supply of H2. Methane production from

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292
Global methane cycle 289

acetate has never been demonstrated and acetoclastic littoral zone, where CH4 is produced in the anoxic sedi-
methanogens are usually not detected in these environ- ment (Schmidt and Conrad, 1993). In the open ocean, on
ments. In the rumen, CH4 production largely occurs in the the other hand, it can only be explained by in situ produc-
gut lumen where it decreases H2 to concentrations that tion, albeit the conditions (high O2 and sulfate) are not
are not permissive for acetogenesis. In the insect hindgut, conducive for CH4 production. Therefore, the processes
however, methanogenesis is restricted to the gut wall or discussed generally assume CH4 production in anoxic
other gut structures, or is associated with gut flagellates, microniches and from non-competitive substrates, such
while in the gut lumen acetogenesis is often the main as methylated compounds. Hence it was discussed that
H2-consuming process. The reason for this difference in CH4 is released from fish intestine, from zooplankton
localization of methanogens in the rumen versus the or from organic particles (Oremland, 1979; Sieburth et al.,
insect hindgut is presently unknown. Why are methano- 1993; DeAngelis and Lee, 1994; Karl and Tilbrook, 1994;
gens in the insect gut restricted to the gut wall, where O2 Marty et al., 1997; Holmes et al., 2000; Damm et al.,
partial pressures are uncomfortably high? Why do rumi- 2008). However, anoxic aggregates may be only an
nants not allow H2-dependent acetogenesis instead of ephemeral phenomenon in the oxic water column (Ploug
methanogenesis, although they would gain a significant et al., 1997), too short-lived to allow development of a
amount of additional energy [about 6–10% of the energy methanogenic population. It is also doubtful whether small
value of food (Liu and Whitman, 2008)] which is lost as zooplankton can provide a permanently anoxic environ-
CH4. The differences are possibly the result of an evolu- ment, since these animals are too small to prevent O2
tionary process, as methanogenesis is restricted to a few diffusion into their interior. Hence, it is not really clear how
animal taxa (Hackstein et al., 1996). In this context it CH4 is produced in pelagic water.
is important to ask in which respect animals benefit from
the methanogens in their gut system (Brune, 2009). Are
insects able to control their methanogenic populations Summary of key issues
better than ruminants do? Several key issues can be identified that need to be
addressed for better understanding of the ecology of CH4-
Methane in the aquatic environment producing or -consuming microorganisms and for their
role in CH4 cycling:
Methane production in aquatic sediments is a common
feature, since these are anoxic environments. It was rec- i. Elucidation of the biochemical mechanisms respon-
ognized early on that most of the CH4 produced is oxi- sible for activation and oxidation of CH4 in the absence
dized at the sediment surface or in the water column by of O2. This issue would be helped by isolation of
aerobic methanotrophic bacteria so that little CH4 escapes anaerobic methanotrophs into pure culture.
ii. Identification, isolation and biochemical characteri-
into the atmosphere (Rudd and Taylor, 1980). However,
zation of high-affinity methanotrophs able to oxidize
since a substantial part (about 50%) of the primary pro-
atmospheric CH4 in soil.
duction of a lake can be returned to the water column in
the form of CH4 (Fallon et al., 1980), CH4 can provide an iii. Identification, isolation and biochemical characteri-
zation of aero-tolerant methanogens thriving in the
important source of energy for the methanotrophic bacte-
pelagic zone of lakes and ocean.
rioplankton and the trophic food chain depending on these
iv. Ecological and evolutionary aspects of heterotrophic
bacteria. There is growing evidence that methanotrophic
methanotrophs. Why are most of the known methan-
bacteria indeed provide an important food source for
otrophs obligately methylotrophic?
zooplankton and insect larvae (Bastviken et al., 2002;
v. Ecological and evolutionary aspects of gut symbio-
Deines et al., 2007; Jones et al., 2008). Methanotrophs
ses with methanogens, reasons for existence and
apparently can also provide a food source for Myxo-
non-existence.
bacteria, amoebae, flagellates and ciliates (Murase and
Frenzel, 2007). However, a quantitative evaluation of the
CH4 flux through the trophic food chain is not yet avail-
able. It is worth mentioning that methanotrophs also serve Acknowledgments
as symbionts in marine invertebrates (Cavanaugh, 1993) I thank A. Brune and W. Liesack for helpful discussion and the
and apparently also in Sphagnum moss (Raghoebarsing Fonds der Chemischen Industrie for financial support.
et al., 2005).
Methane production in the oxic water column is a
paradox, but nevertheless exists (Reeburgh, 2007). In References
lakes, supersaturation of oxic water bodies with CH4 can Baani, M., and Liesack, W. (2008) Two isozymes of par-
sometimes be explained by lateral transport from the ticulate methane monooxygenase with different methane

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292
290 R. Conrad
oxidation kinetics are found in Methylocystis sp. strain Dunfield, P.F., and Conrad, R. (2000) Starvation alters the
SC2. Proc Natl Acad Sci USA 105: 10203–10208. apparent half-saturation constant for methane in the type
Bastviken, D., Ejlertsson, J., and Tranvik, L. (2002) Measure- II methanotroph Methylocystis strain LR1. Appl Environ
ment of methane oxidation in lakes: a comparison of Microbiol 66: 4136–4138.
methods. Environ Sci Technol 36: 3354–3361. Dunfield, P.F., Liesack, W., Henckel, T., Knowles, R., and
Beerling, D.J., Gardiner, T., Leggett, G., McLeod, A., and Conrad, R. (1999) High-affinity methane oxidation by a soil
Quick, W.P. (2008) Missing methane emissions from enrichment culture containing a type II methanotroph. Appl
leaves of terrestrial plants. Global Change Biol 14: 1821– Environ Microbiol 65: 1009–1014.
1826. Dunfield, P.F., Yuryev, A., Senin, P., Smirnova, A.V., Stott,
Bender, M., and Conrad, R. (1992) Kinetics of CH4 oxidation M.B., Hou, S.B., et al. (2007) Methane oxidation by an
in oxic soils exposed to ambient air or high CH4 mixing extremely acidophilic bacterium of the phylum Verrucomi-
ratios. FEMS Microbiol Ecol 101: 261–270. crobia. Nature 450: 879–882.
Boetius, A., Ravenschlag, K., Schubert, C.J., Rickert, D., Erkel, C., Kube, M., Reinhardt, R., and Liesack, W. (2006)
Widdel, F., Gieseke, A., et al. (2000) A marine microbial Genome of Rice Cluster I archaea – the key methane
consortium apparently mediating anaerobic oxidation of producers in the rice rhizosphere. Science 313: 370–372.
methane. Nature 407: 623–626. Ettwig, K.F., Shima, S., VandePas-Schoonen, K.T., Kahnt, J.,
Bousquet, P., Ciais, P., Miller, J.B., Dlugokencky, E.J., Medema, M.H., OpdenCamp, H.J.M., et al. (2008) Deni-
Hauglustaine, D.A., Prigent, C., et al. (2006) Contribution trifying bacteria anaerobically oxidize methane in the
of anthropogenic and natural sources to atmospheric absence of Archaea. Environ Microbiol 10: 3164–3173.
methane variability. Nature 443: 439–443. Fallon, R.D., Harrits, S., Hanson, R.S., and Brock, T.D.
Brune, A. (2009) Methanogenesis in the digestive tracts of (1980) The role of methane in internal carbon cycling
insects. In Microbiology of Hydrocarbons. Timmis, K. (ed.). in Lake Mendota during summer stratification. Limnol
Heidelberg, Germany: Springer (in press). Oceanogr 25: 357–360.
Cavanaugh, C.M. (1993) Methanotroph-invertebrate symbio- Frankenberg, C., Meirink, J.F., VanWeele, M., Platt, U., and
ses in marine environment: ultrastructural, biochemical and Wagner, T. (2005) Assessing methane emissions from
molecular studies. In Microbial Growth on C1 Compounds. global space-borne observations. Science 308: 1010–
Murrell, J.C., and Kelly, D.P. (eds). Andover, MA, USA: 1014.
Intercept, pp. 315–328. Frenzel, P. (2000) Plant-associated methane oxidation in rice
Conrad, R. (1984) Capacity of aerobic microorganisms to fields and wetlands [Review]. Adv Microb Ecol 16: 85–114.
utilize and grow on atmospheric trace gases. In Current Girguis, P.R., Cozen, A.E., and DeLong, E.F. (2005) Growth
Perspectives in Microbial Ecology. Klug, M.G., and Reddy, and population dynamics of anaerobic methane-oxidizing
C.A. (eds). Washington, DC, USA: American Society for archaea and sulfate-reducing bacteria in a continuous-flow
Microbiology, pp. 461–467. bioreactor. Appl Environ Microbiol 71: 3725–3733.
Conrad, R. (2007) Microbial ecology of methanogens and Groot, T.T., VanBodegom, P.M., Harren, F.J.M., and Meijer,
methanotrophs. Adv Agron 96: 1–63. H.A.J. (2003) Quantification of methane oxidation in the
Conrad, R., Klose, M., Noll, M., Kemnitz, D., and Bodelier, rice rhizosphere using 13C-labelled methane. Biogeochem
P.L.E. (2008) Soil type links microbial colonization of rice 64: 355–372.
roots to methane emission. Global Change Biol 14: 657– Grosskopf, R., Stubner, S., and Liesack, W. (1998) Novel
669. euryarchaeotal lineages detected on rice roots and in the
Damm, E., Kiene, R.P., Schwarz, J., Falck, E., and Dieck- anoxic bulk soil of flooded rice microcosms. Appl Environ
mann, G. (2008) Methane cycling in Arctic shelf water and Microbiol 64: 4983–4989.
its relationship with phytoplankton biomass and DMSP. Hackstein, J.H.P., Langer, P., and Rosenberg, J. (1996)
Mar Chem 109: 45–59. Genetic and evolutionary constraints for the symbiosis
DeAngelis, M.A., and Lee, C. (1994) Methane production between animals and methanogenic bacteria. Environ
during zooplankton grazing on marine phytoplankton. Monit Assess 42: 39–56.
Limnol Oceanogr 39: 1298–1308. Hallam, S.J., Putnam, N., Preston, C.M., Detter, J.C.,
Dedysh, S.N., Knief, C., and Dunfield, P.F. (2005) Methylo- Rokhsar, D., Richardson, P.M., and DeLong, E.F. (2004)
cella species are facultatively methanotrophic. J Bacteriol Reverse methanogenesis: testing the hypothesis with
187: 4665–4670. environmental genomics. Science 305: 1457–1462.
Deines, P., Bodelier, P.L.E., and Eller, G. (2007) Methane- Henckel, T., Jäckel, U., Schnell, S., and Conrad, R. (2000)
derived carbon flows through methane-oxidizing bacteria Molecular analyses of novel methanotrophic communi-
to higher trophic levels in aquatic systems. Environ Micro- ties in forest soil that oxidize atmospheric methane. Appl
biol 9: 1126–1134. Environ Microbiol 66: 1801–1808.
Dueck, T.A., DeVisser, R., Poorter, H., Persijn, S., Gorissen, Hoehler, T.M., Alperin, M.J., Albert, D.B., and Martens, C.S.
A., DeVisser, W., et al. (2007) No evidence for substantial (1994) Field and laboratory studies of methane oxidation in
aerobic methane emission by terrestrial plants: a 13C- an anoxic marine sediment: evidence for a methanogen–
labelling approach. New Phytol 175: 29–35. sulfate reducer consortium. Global Biogeochem Cycles
Dunfield, P.F. (2007) The soil methane sink. In Greenhouse 8: 451–463.
Gas Sinks. Reay, D.S., Hewitt, N., Smith, K.A., and Grace, Holmes, A.J., Roslev, P., McDonald, I.R., Iversen, N.,
J. (eds). Wallingford, UK: CABI Publishing, pp. 152– Henriksen, K., and Murrell, J.C. (1999) Characterization
170. of methanotrophic bacterial populations in soils showing

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292
Global methane cycle 291
atmospheric methane uptake. Appl Environ Microbiol 65: Marty, D., Nival, P., and Yoon, W.D. (1997) Methanoarchaea
3312–3318. associated with sinking particles and zooplankton collected
Holmes, M.E., Sansone, F.J., Rust, T.M., and Popp, B.N. in the northeastern tropical Atlantic. Oceanol Acta 20:
(2000) Methane production, consumption, and air–sea 863–869.
exchange in the open ocean: an evaluation based on Murase, J., and Frenzel, P. (2007) A methane-driven micro-
carbon isotopic ratios. Global Biogeochem Cycles 14: bial food web in a wetland rice soil. Environ Microbiol
1–10. 9: 3025–3034.
Islam, T., Jensen, S., Reigstad, L.J., Larsen, O., and Birke- Nauhaus, K., Albrecht, M., Elvert, M., Boetius, A., and
land, N.K. (2008). Methane oxidation at 55°C and pH 2 by Widdel, F. (2007) In vitro cell growth of marine archaeal–
a thermoacidophilic bacterium belonging to the Verrucomi- bacterial consortia during anaerobic oxidation of methane
crobia phylum. Proc Natl Acad Sci USA 105: 300–304. with sulfate. Environ Microbiol 9: 187–196.
Janssen, P.H., and Kirs, M. (2008) Structure of the archaeal Nouchi, I., Mariko, S., and Aoki, K. (1990) Mechanisms of
community of the rumen [Review]. Appl Environ Microbiol methane transport from the rhizosphere to the atmosphere
74: 3619–3625. through rice plants. Plant Physiol 94: 59–66.
Jones, R.I., Carter, C.E., Kelly, A., Ward, S., Kelly, D.J., and OpdenCamp, H.J.M., Islam, T., Stott, M.B., Harhangi, H.R.,
Grey, J. (2008) Widespread contribution of methane-cycle Hynes, A., Schouten, S., et al. (2009) Environmental,
bacteria to the diets of lake profundal chironomid larvae. genomic and taxonomic perspectives on methanotrophic
Ecology 89: 857–864. Verrucomicrobia. Environ Microbiol Rep (in press).
Karl, D.M., and Tilbrook, B.D. (1994) Production and trans- doi:10.1111/j.1758-2229.2009.00022.x.
port of methane in oceanic particulate organic matter. Oremland, R.S. (1979) Methanogenic activity in plankton
Nature 368: 732–734. samples and fish intestines: a mechanism for in situ metha-
Keppler, F., Hamilton, J.T.G., Brass, M., and Röckmann, T. nogenesis in oceanic surface waters. Limnol Oceanogr
(2006) Methane emissions from terrestrial plants under 24: 1136–1141.
aerobic conditions. Nature 439: 187–191. Orphan, V.J., House, C.H., Hinrichs, K.U., McKeegan, K.D.,
Knief, C., Lipski, A., and Dunfield, P.F. (2003) Diversity and and DeLong, E.F. (2001) Methane-consuming archaea
activity of methanotrophic bacteria in different upland soils. revealed by directly coupled isotopic and phylogenetic
Appl Environ Microbiol 69: 6703–6714. analysis. Science 293: 484–487.
Knief, C., Kolb, S., Bodelier, P.L.E., Lipski, A., and Dunfield, Orphan, V.J., House, C.H., Hinrichs, K.U., McKeegan, K.D.,
P.F. (2006) The active methanotrophic community in hydro- and DeLong, E.F. (2002) Multiple archaeal groups mediate
morphic soils changes in response to changing methane methane oxidation in anoxic cold seep sediments. Proc
concentration. Environ Microbiol 8: 321–333. Natl Acad Sci USA 99: 7663–7668.
Knittel, K., Losekann, T., Boetius, A., Kort, R., and Amann, R. Ploug, H., Kuehl, M., Buchholz-Cleven, B., and Joergensen,
(2005) Diversity and distribution of methanotrophic B.B. (1997) Anoxic aggregates – an ephemeral phenom-
archaea at cold seeps. Appl Environ Microbiol 71: 467– enon in the pelagic environment? Aquat Microb Ecol 13:
479. 285–294.
Krüger, M., Meyerdierks, A., Glöckner, F.O., Amann, R., Pol, A., Heijmans, K., Harhangi, H.R., Tedesco, D., Jetten,
Widdel, F., Kube, M., et al. (2003) A conspicuous nickel M.S.M., and OpdenCamp, H.J.M. (2007) Methanotrophy
protein in microbial mats that oxidize methane anaerobi- below pH 1 by a new Verrucomicrobia species. Nature 450:
cally. Nature 426: 878–881. 874-878.
Lelieveld, J., Crutzen, P.J., and Dentener, F.J. (1998) Purdy, K.J. (2007) The distribution and diversity of euryar-
Changing concentrations, lifetime and climate forcing of chaeota in termite guts. Adv Appl Microbiol 62: 63–80.
atmospheric methane. Tellus 50B 128–150. Radajewski, S., Ineson, P., Parekh, N.R., and Murrell, J.C.
Liu, Y., and Whitman, W.B. (2008) Metabolic, phylogenetic, (2000) Stable-isotope probing as a tool in microbial
and ecological diversity of the methanogenic archaea. Ann ecology. Nature 403: 646–649.
N Y Acad Sci 1125: 171–189. Raghoebarsing, A.A., Smolders, A.J.P., Schmid, M.C.,
Lu, Y.H., and Conrad, R. (2005) In situ stable isotope probing Rijpstra, W.I.C., Wolters-Arts, M., Derksen, J., et al. (2005)
of methanogenic archaea in the rice rhizosphere. Science Methanotrophic symbionts provide carbon for photosyn-
309: 1088–1090. thesis in peat bogs. Nature 436: 1153–1156.
Lösekann, T., Knittel, K., Nadalig, T., Fuchs, B., Niemann, H., Raghoebarsing, A.A., Pol, A., van de Pas-Schoonen, K.T.,
Boetius, A., and Amann, R. (2007) Diversity and abun- Smolders, A.J.P., Ettwig, K.F., Rijpstra, W.I.C., et al. (2006)
dance of aerobic and anaerobic methane oxidizers at the A microbial consortium couples anaerobic methane oxida-
Haakon Mosby mud volcano, Barents Sea. Appl Environ tion to denitrification. Nature 440: 918–921.
Microbiol 73: 3348–3362. Reeburgh, W.S. (2003) Global methane biogeochemistry.
McDonald, I.R., Bodrossy, L., Chen, Y., and Murrell, J.C. In Treatise on Geochemistry, Vol. 4: The Atmosphere.
(2008) Molecular ecology techniques for the study of Keeling, R.F., Holland, H.D., and Turekian, K.K. (eds).
aerobic methanotrophs [Review]. Appl Environ Microbiol Oxford, UK: Elsevier-Pergamon, pp. 65–89.
74: 1305–1315. Reeburgh, W.S. (2007) Oceanic methane biogeochemistry
McLeod, A.R., Fry, S.C., Loake, G.J., Messenger, D.J., [Review]. Chem Rev 107: 486–513.
Reay, D.S., Smith, K.A., and Yun, B.W. (2008) Ultraviolet Ricke, P., Erkel, C., Kube, M., Reinhardt, R., and Liesack, W.
radiation drives methane emissions from terrestrial plant (2004) Comparative analysis of the conventional and novel
pectins. New Phytol 180: 124–132. pmo (particulate methane monooxygenase) operons from

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292
292 R. Conrad
Methylocystis strain SC2. Appl Environ Microbiol 70: are all new immunotypes and probably include new taxa.
3055–3063. Mar Ecol Prog Ser 95: 81–89.
Ricke, P., Kube, M., Nakagawa, S., Erkel, C., Reinhardt, R., Stoecker, K., Bendinger, B., Schöning, B., Nielsen, P.H.,
and Liesack, W. (2005) First genome data from uncultured Nielsen, J.L., Baranyi, C., et al. (2006) Cohn’s Crenothrix is
upland soil cluster alpha methanotrophs provide further a filamentous methane oxidizer with an unusual methane
evidence for a close phylogenetic relationship to Methylo- monooxygenase. Proc Natl Acad Sci USA 103: 2363–
capsa acidiphila B2 and for high-affinity methanotrophy 2367.
involving particulate methane monooxygenase. Appl Terazawa, K., Ishizuka, S., Sakatac, T., Yamada, K., and
Environ Microbiol 71: 7472–7482. Takahashi, M. (2007) Methane emissions from stems of
Rudd, J.W.M., and Taylor, C.D. (1980) Methane cycling in Fraxinus mandshurica var. japonica trees in a floodplain
aquatic environments. Adv Aquat Microbiol 2: 77–150. forest. Soil Biol Biochem 39: 2689–2692.
Rusch, H., and Rennenberg, H. (1998) Black alder [Alnus Thauer, R.K., and Shima, S. (2008) Methane as fuel for
glutinosa (L.) Gaertn.] trees mediate methane and nitrous anaerobic microorganisms [Review]. Ann N Y Acad Sci
oxide emission from the soil to the atmosphere. Plant Soil 1125: 158–170.
201: 1–7. Theisen, A.R., Ali, M.H., Radajewski, S., Dumont, M.G.,
Sakai, S., Imachi, H., Sekiguchi, Y., Ohashi, A., Harada, H., Dunfield, P.F., McDonald, I.R., et al. (2005) Regulation
and Kamagata, Y. (2007) Isolation of key methanogens for of methane oxidation in the facultative methanotroph
global methane emission from rice paddy fields: a novel Methylocella silvestris BL2. Mol Microbiol 58: 682–692.
isolate affiliated with the clone cluster rice cluster I. Appl Trotsenko, Y.A., and Murrell, J.C. (2008) Methabolic aspects
Environ Microbiol 73: 4326–4331. of aerobic obligate methanotrophy. Adv Appl Microbiol 63:
Sakai, S., Imachi, H., Hanada, S., Ohashi, A., Harada, H., 183–229.
and Kamagata, Y. (2008) Methanocella paludicola gen. Valentine, D.L., and Reeburgh, W.S. (2000) New perspec-
nov., sp. nov., a methane-producing archaeon, the first tives on anaerobic methane oxidation [Review]. Environ
isolate of the lineage ‘Rice Cluster I’, and proposal of the Microbiol 2: 477–484.
new archaeal order Methanocellales ord. nov. Int J Syst Vigano, I., VanWeelden, H., Holzinger, R., Keppler, F.,
Evol Microbiol 58: 929–936. McLeod, A., and Röckmann, T. (2008) Effect of UV radia-
Schink, B., Ward, J.C., and Zeikus, J.G. (1981) Microbiology tion and temperature on the emission of methane from
of wetwood: importance of pectin degradation and plant biomass and structural components. Biogeosciences
Clostridium species in living trees. Appl Environ Microbiol 5: 937–947.
42: 526–532. Vigliotta, G., Nutricati, E., Carata, E., Tredici, S.M.,
Schmidt, I., and Bock, E. (1997) Anaerobic ammonia oxida- DeStefano, M., Pontieri, P., et al. (2007) Clonothrix fusca
tion with nitrogen dioxide by Nitrosomonas eutropha. Arch Roze 1896, a filamentous, sheathed, methanotrophic
Microbiol 167: 106–111. gamma-proteobacterium. Appl Environ Microbiol 73:
Schmidt, I., VanSpanning, R.J.M., and Jetten, M.S.M. (2004) 3556–3565.
Denitrification and ammonia oxidation by Nitrosomonas Wang, J.S., Logan, J.A., McElroy, M.B., Duncan, B.N.,
europaea wild-type, and NirK- and NorB-deficient mutants. Megretskaia, I.A., and Yantosca, R.M. (2004) A 3-D model
Microbiology – UK 150: 4107–4114. analysis of the slowdown and interannual variability in
Schmidt, U., and Conrad, R. (1993) Hydrogen, carbon mon- the methane growth rate from 1988 to 1997 [Review].
oxide, and methane dynamics in Lake Constance. Limnol Global Biogeochem Cycles 18: B3011. doi:10.1029/
Oceanogr 38: 1214–1226. 2003GB002180.
Schütz, H., Seiler, W., and Conrad, R. (1989) Processes Watanabe, A., Takeda, T., and Kimura, M. (1999) Evaluation
involved in formation and emission of methane in rice of origins of CH4 carbon emitted from rice paddies.
paddies. Biogeochem 7: 33–53. J Geophys Res 104: 23623–23629.
Sieburth, J.M., Johnson, P.W., Macario, A.J.L., and Conway- West, A.E., and Schmidt, S.K. (1999) Acetate stimulates
DeMacario, E. (1993) C1 bacteria in the water column of atmospheric CH4 oxidation by an alpine tundra soil. Soil
Chesapeake Bay, USA. 2. The dominant O2-tolerant and Biol Biochem 31: 1649–1655.
H2S-tolerant methylotrophic methanogens, coenriched with Zeikus, J.G., and Ward, J.C. (1974) Methane formation in
their oxidative and sulphate reducing bacterial consorts, living trees: a microbial origin. Science 184: 1181–1183.

© 2009 Society for Applied Microbiology and Blackwell Publishing Ltd, Environmental Microbiology Reports, 1, 285–292

You might also like