Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/300816406

Organic Matter Mineralization and Decomposition

Chapter · January 2013


DOI: 10.2136/sssabookser10.c20

CITATIONS READS

18 2,198

6 authors, including:

Scott Bridgham Rongzhong Ye


University of Oregon Clemson University
182 PUBLICATIONS   10,298 CITATIONS    50 PUBLICATIONS   722 CITATIONS   

SEE PROFILE SEE PROFILE

Ronald D Delaune Konda Ramesh Reddy


Louisiana State University University of Florida
500 PUBLICATIONS   19,025 CITATIONS    409 PUBLICATIONS   20,594 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Swamp Restoration Project View project

Mesopotamian Marsh Research View project

All content following this page was uploaded by Scott Bridgham on 27 August 2016.

The user has requested enhancement of the downloaded file.


SE VEN

Carbon Dynamics and Ecosystem Processes


SCOT T D. BR IDGHAM

THE BASIC CARBON CYCLE: METRICS OF PRODUCTIVITY ANAEROBIC CARBON CYCLING


AND CARBON BALANCE
WHY DO WETLANDS ACCUMULATE SOIL CARBON?
BREAKDOWN AND DECOMPOSITION OF ORGANIC MATTER
DISSOLVED ORGANIC MATTER FLUXES FROM WETLANDS
Short-Term Breakdown and Decomposition
Litter Quality Controls over Decomposition THE BIG PICTURE
Exogenous Controls over Decomposition
Methane Fluxes
Nutrient Mineralization-Immobilization during Decomposition
Global Soil Carbon Pools and Accumulation

One of the primary functions of wetlands is their role in The Basic Carbon Cycle:
the global carbon cycle. This topic has been of increas- Metrics of Productivity and Carbon Balance
ing interest because of the potential for globally signifi-
cant feedbacks between various aspects of wetland carbon The basic carbon cycle is largely generic to any wetland
dynamics and global change (Bridgham et al. 1995; Lim- or terrestrial ecosystem, although the sizes of the various
pens et al. 2008; Dise 2009; DeLaune and White 2011; Kir- pools and rates vary significantly among different types
wan and Mudd 2012; Bridgham et al. 2013; see also Chap- of ecosystems (Fig. 7.1). Wetlands, in particular, vary dra-
ter 10). Wetland carbon dynamics are also relevant at the matically in their degree of openness to the fluxes of dis-
scale of watersheds because of their importance in the solved inorganic carbon (DIC), dissolved organic matter (DOM),
export of dissolved organic matter to surface waters, with and particulate organic matter (POM) that enter and leave an
its myriad effects on the chemistry and biology of surface ecosystem. DOM and POM are also often termed dissolved
waters, which are discussed later in this chapter. Carbon organic carbon (DOC) and particulate organic carbon (POC),
and nutrient cycles are intimately linked in all ecosystems, depending on whether the carbon fraction is being empha-
so any understanding of nutrient cycling must entail a con- sized or not.
sideration of associated aspects of carbon dynamics. Per- A number of different carbon balance and productiv-
haps most fundamentally, carbon dynamics are intimately ity terms are used, and often misused. Net primary produc-
linked to the formation and development of wetlands; tivity (NPP) is the difference between plant carbon diox-
this link is particularly dramatic in the hydrogeochemis- ide uptake in photosynthesis (or gross primary productivity,
try of peatlands (Moore and Bellamy 1974; Bridgham et al. GPP) and plant respiration; NPP can also be thought of as
1996; Belyea and Baird 2006) and salt marshes (Kirwan and the production of organic compounds by plants over some
Mudd 2012). This chapter primarily focuses on soil carbon period of time. This can be measured at the leaf level with
dynamics and the dissolved and gaseous exports that result a portable photosynthesis analyzer, or at the whole plant or
from those dynamics, although the basic carbon cycle and plot level with destructive techniques such as peak stand-
overall net carbon accumulation or loss are also considered. ing biomass. Although many studies have measured NPP

185

53488p151-286.indd 185 8/19/14 9:06 AM


Net Ecosystem Producvity or Net Ecosystem Exchange
CO2 + CH4 CO2 CO2 CO2 CO2, CO
Fire

Respiraon Respiraon Photosynthesis

Consumers Herbivory Harvesng


Plant NPP

Predaon Death, Lier


Excreon
DIC, DOM, POM DIC, DOM, POM
Decompo- Available Soil Preservaon
Decomposers sion Organic Unavailable
Processes
Maer Soil Organic Maer
Detrivore
Death
Δ C = Net Ecosystem Carbon Balance

Leaching of DIC and DOM

FIGURE 7.1. The basic carbon cycle in ecosystems.

in wetlands (see Chapter 5), the vast majority focus only The net ecosystem carbon balance (NECB) is the net gain or
on determining the accumulation of new plant biomass loss of carbon by an ecosystem over time:
aboveground, although belowground NPP can be equal
to or greater than aboveground NPP (Weltzin et al. 2000). NECB = dC / dt = (gaseous inputs –​losses) + (dissolved
This is because measuring belowground NPP is difficult, inputs –​losses) + (particulate inputs –​losses)
can be very time consuming, and all methods have impor-
tant assumptions that are probably often not met (Lauen- = (NEE + FCH4 + F VOC) + (FDOC + FDIC) + FPOC, (7.2)
roth 2000). NPP may be underestimated by measuring just
peak standing crop because herbivory and mortality of where F refers to a flux. NECB includes all net fluxes of car-
individual plant parts can remove plant biomass before it bon in and out of an ecosystem, including those resulting
is sampled, and not all NPP results in the production of from disturbances such as fire and harvesting of trees.
new plant tissue—​e.g., root exudates, carbon allocation to Overall carbon dynamics at an ecosystem scale have been
mycorrhizae, and emissions of volatile organic compounds most extensively studied in northern peatlands, so that is
(VOC) (Chapin III et al. 2011). used as an example of these concepts here. NECB in north-
Net ecosystem production (NEP) is the difference between ern peatlands averages between about 20 to 70 g C m–​2 yr–​1
gross primary productivity and ecosystem respiration (R E), in individual peatlands, based on multiple years of data (Fig.
which is the sum of plant respiration and heterotrophic res- 7.2). The largest component of this is the net flux of carbon
piration (R H): dioxide in and out of the ecosystem (i.e., NEE), which, for
most northern peatlands, ranges between 20 and 60 g m–​2
NEP = GPP –​R E = NPP –​R H. (7.1) yr–​1, with large interannual variability (Blodau 2002; Lim-
pens et al. 2008), which in turn causes large interannual
Actually measuring the various components of NEP is variability in NECB (Fig. 7.3). Leaching of DOC is a larger
someplace between very difficult and impossible, so it is loss term than methane emissions in these peatlands.
often approximated by net ecosystem exchange (NEE), which The net carbon balance (NCB) is derived from modeling
is the net gaseous exchange of carbon between an ecosys- long-term carbon accumulation rates in peat and is compa-
tem and the atmosphere. NEE is measured with chamber rable to the NECB derived from the annual measurements
or eddy covariance techniques. NEE is often construed to of the terms in Equation 7.2 (Yu 2012). The NCB in north-
include only the net exchange of carbon dioxide, although ern peatlands averaged over the Holocene was 19 g m–​2 yr–​1,
methane (CH4) fluxes can be a component of heterotrophic but rates during the last millennium were only 10.4 g m–​2
respiration in wetlands. To add to the confusion, a net posi- yr–​1 (Fig. 7.3). Thus, carbon accumulation rates in north-
tive accumulation of carbon into an ecosystem is treated as ern peatlands determined by NCB are somewhat lower than
either a positive or a negative NEE flux, depending on the NECB, but they are perhaps surprisingly close given the dif-
study. In this chapter, a positive value for NEE indicates a ferent methods employed, time frames examined, and lim-
net flux of carbon into the ecosystem. ited number of sites in which NECB has been measured.

186  CHAP TER Seven

53488p151-286.indd 186 8/29/14 4:35 AM


FIGURE 7.2. Net ecosystem exchange (NEE), dissolved organic carbon
flux (DOC), and methane flux (CH4) in five peatlands averaged over
multiple years. The numbers above each bar are the net ecosystem
carbon balance (±1 standard deviation). Positive numbers indicate
carbon flux into the ecosystem, and negative numbers indicate
carbon flux from the ecosystem. Data are taken from Yu (2012),
but the original references are Roulet et al. (2007) for Mer Bleue, an
ombrotrophic bog in Canada with seven years of data; Nilsson et
al. (2008) for Degerö Stormyr, a minerotrophic fen in Sweden with
two years of data; Dinsmore et al. (2010) for Auchencorth Moss, an
ombrotrophic bog in Scotland with two years of data; Koehler et al.
(2011) for Glencar, a blanket bog in Ireland with six years of data;
and Olefeldt et al. (2012) for Stordalen, a permafrost palsa mire in
Sweden with two years of data. Auchencorth Moss includes other
fluxes such as stream evasion of carbon dioxide and methane, so the
components do not add up to the NECB.

FIGURE 7.3. (A) Interannual variability in the net ecosystem carbon balance from the sites in Figure 7.2. The box is the average + standard
error. (B) Average net ecosystem carbon balance over the Holocene in northern peatlands. Adapted from Yu (2012).

Breakdown and Decomposition of


by converting them into more oxidized forms—​ w ith
Organic Matter
CO2 being the most oxidized endpoint—​a nd using that
The breakdown of organic matter couples physical factors energy for growth and maintenance requirements. Also,
and biological decomposition processes. For wetlands this as with many ecological processes, it can be understood
was recently reviewed by Bridgham and Lamberti (2009)—​ fundamentally as a stoichiometric process whereby organ-
this portion of this chapter is an update of that review. isms obtain elements such as carbon, nitrogen, and phos-
Methods used to determine rates of anaerobic carbon phorus in relative proportion to the constraints imposed
cycling are described in Bridgham and Ye (2013). (Roles of by the elemental makeup of their tissues. Because of the
microbes in decomposition are covered in detail in Chap- basic importance of decomposition and the relative ease
ter 4.) and low cost of many of the techniques used to quantify
Decomposition—​a lso referred to as mineralization—​is the it (Bridgham and Ye 2013), it has been widely studied in
biological transformation of organic compounds into inor- wetlands and other ecosystems for many years. It was esti-
ganic forms of nutrients and carbon—​largely CO2 and CH4 mated that, in the 20 years prior to 1997, over 1,000 papers
gases—​and simpler organic compounds. As with many bio- had been published on this topic (Heal et al. 1997). The
logical processes, this can be understood fundamentally four aspects of this process that make decomposition in
as an oxidation-reduction process in which organisms wetlands different from other ecosystems are (1) the tissue
release the energy stored in reduced organic compounds chemistry of Sphagnum spp. mosses of peatlands with their

Carbon Dy namics and Ecosystem Processes   187

53488p151-286.indd 187 8/19/14 9:06 AM


A B

C D

FIGURE 7.4. Stages of leaf decomposition in aquatic ecosystems. Maple (Acer rubrum) leaf after (A) 1 week, (B) 2 weeks, (C) 4 weeks, and Fig. 4
(D) 6 weeks of decomposition in a Michigan, USA, stream. Note holes in leaf caused by detritivore feeding at 2 and 4 weeks, and residual
veins and ribs after 6 weeks. Photos courtesy of A. Bobeldyk (from Bridgham and Lamberti 2009).

very low rates of decomposition; (2) waterlogging and the carbon dioxide, methane, and inorganic nutrients. Leach-
resultant anaerobiosis; (3) the imbalance between decom- ing is the abiotic removal of soluble compounds by water
position and production in many wetlands, leading to the and is particularly important in the initial stages of the
buildup of large soil carbon stores; and (4) the low density breakdown of fresh litterfall. Fragmentation is the physical
of mineral surfaces for organic matter to react with, at least breakdown of detritus into smaller particles, largely by the
in peatlands. feeding activities of detritivorous invertebrates (Fig. 7.4).
Dead organic matter is termed detritus, and the input of Catabolism is the metabolic decomposition of detritus into
freshly senescent detritus is termed litterfall—​despite its heat, metabolic energy, and simpler organic and inorganic
name, however, this should include belowground inputs. compounds. Soil organisms then use these compounds
The controls over rates of breakdown and decomposition and metabolic energy for their growth and reproduction in
of relatively fresh litterfall and soil organic matter may be the process of anabolism. Traditionally it has been thought
quite different, and thus they are considered separately that the accumulation of soil organic matter is the result of
below. The breakdown and decomposition of organic detri- the process of humification wherein highly complex, recal-
tus involves a number of basic processes that occur in a citrant organic compounds that have very long turnover
foodweb cascade where detritus is repeatedly processed times are formed in situ. This traditional concept has been
by the decomposer community with the progressive loss sharply challenged in recent years, as discussed further
of mass; the loss of potential chemical energy in reduced below. Soil organic matter may also be stabilized by associ-
organic carbon bonds; the turnover of decomposer biomass ations with mineral particles that appear to provide physi-
by death and predation; and ultimately the production of cal protection from microbial attack.

188  CHAP TER Seven

53488p151-286.indd 188 8/19/14 9:06 AM


FIGURE 7.5. Decomposition rates in Canadian
peatlands. Adapted from Moore and Basiliko (2006).

Short-Term Breakdown and Decomposition


of decomposition studies shows more overlap in the decay
The short-term dynamics of breakdown and decomposi- rates among vascular plant groups in wetlands, although
tion are normally defined over a number of years, when similar trends are evident (Table 7.1). However, aquatic spe-
the original plant litter is identifiable and, operationally, cies have very high rates, and Sphagnum mosses, which are
within the timeframe of most research studies. The break- a dominant component of many peatland plant communi-
down and decomposition of plant litter is most often quan- ties, have very low rates. The few studies that have exam-
tified by mass loss with the litter bag technique (Bridgham ined the decay of roots and rhizomes have indicated that
and Ye 2013). The mass loss of litter over a period of years they decay at slower rates than corresponding leaf mass
often follows a simple exponential relationship: (Bragazza et al. 2009). The reasons for these differences in
decay rates are explored in the next section.
Xt
= e –kt (7.3) Interestingly, with the exception of aquatic species and
Xo
Sphagnum mosses, wetland plants do not appear to have
where Xt is the mass of litter remaining at time t, Xo is the lower average decay rates than vascular plants in terrestrial
original mass of litter, and k is the decay rate—​usually with ecosystems (Table 7.1). Similar results were found in a more
units of year–​1. The mean residence time is then the inverse controlled comparison in Canadian peatlands and nearby
of k. Much of the decomposition literature reports results in terrestrial ecosystems (Moore and Basiliko 2006; Moore et
terms of k, so it is important to be familiar with this simple al. 2008). Thus it appears that the short-term decay of plant
mathematical relationship. Alternatively, decomposition litter is a poor predictor of the ability of wetlands to seques-
rates can be determined over shorter periods by measuring ter soil carbon, with the exception of those containing
rates of the production of CO2 and CH4, but this approach Sphagnum mosses. To put the difference in decay rates for
is more typically followed for soil organic matter (Bridgham fresh plant litter and older soil organic matter in perspec-
and Ye 2013). tive, the time-averaged decay rate (k) of the catotelm—​t he
Typical exponential decay relationships are shown in Fig- permanently waterlogged portion of a peat—​is 8.6 × 10 –​6
ure 7.5. In general, decay rates follow the following pattern: yr–​1 for northern peatlands and 2.4 × 10 –​4 yr–​1 for tropical
woody material < hummock Sphagnum species < hollow/ peatlands (Yu 2011). Thus long-term decay of soil organic
lawn Sphagnum species < tree leaves < shrub leaves < sedge matter in wetlands is many orders of magnitude slower
leaves, and evergreen leaves < deciduous leaves (Moore than the decomposition of fresh plant litter (Table 7.1), and
and Basiliko 2006; Moore et al. 2007). A broader survey it almost certainly has a different set of controls.

Carbon Dy namics and Ecosystem Processes   189

53488p151-286.indd 189 8/19/14 9:06 AM


table 7.1
Decomposition Rates in Wetlands and Other Freshwater Aquatic Habitats

Mean k ± SD t50
Plant Type Na (yr-1) (yr) Reference

Wetlands
Emergent nonwoody aquatic 280 3.0 (7.8) 0.23 Chimney and Pietro (2006)
macrophytesb
Floating aquatic speciesb 80 13.9 (12.3) 0.05 Chimney and Pietro (2006)
Submerged aquatic speciesb 107 17.3 (19.2) 0.04 Chimney and Pietro (2006)
Peatland evergreen shrubs 18 0.30 (0.15) 2.3 Aerts et al. (1999)
and treesc
Peatland deciduous shrubs 4 0.39 (0.08) 1.8 Aerts et al. (1999)
and treesc
Peatland graminoidsc 32 0.40 (0.23) 1.7 Aerts et al. (1999)
Peatland vascular plantsd 69 0.23 (0.16) 3.8 Bragazza et al. (2009)
Peatland Sphagnum spp. mossesc 17 0.16 (0.07) 4.3 Aerts et al. (1999)
Peatland Sphagnum spp. mossesd 22 0.08 (0.05) 8.5 Bragazza et al. (2009)
Marsh plantsd 4 0.31 (0.06) 2.2 Bragazza et al. (2009)
Northern peatland catotelm NA 0.0000086 81,070 Yu (2011)
peate
Southern peatland catolelm NA 0.00024 2,937 Yu (2011)
peate

Terrestrial Ecosystems
Forestsd 34 0.42 (0.22) 1.6 Bragazza et al. (2009)
Mediterranean shrublandsd 20 0.29 (0.18) 2.4 Bragazza et al. (2009)
Grasslandsd 6 0.74 (0.18) 0.94 Bragazza et al. (2009)
Desertsd 2 0.30 (0.04) 2.4 Bragazza et al. (2009)
Tundrad 22 0.32 (0.20) 2.2 Bragazza et al. (2009)
a Number of published studies considered.
b Fields studies using a variety of techniques and duration (10 to 1052 d; median = 137 d) from tropics to arctic.
c Northern peatlands using litter bags based on first year of decomposition.

d Studies used litter bag technique for more than one year of field deployment (most 2–​5 yr).

e Time-averaged value for the entire catotelm.

Litter Quality Controls over Decomposition


cellulose and hemicellulose are entwined into strands or
The initial chemistry (often termed the “quality”) of plant fibers to form cell walls and require exoenzymes secreted by
litter is one of the most important determinants of the rate microbes to cleave them into smaller subunits before they
of short-term decomposition (Berg and Laskowski 2006). can be taken into cells. However, these macromolecules still
Plant litter is actually a diverse assemblage of organic com- decompose relatively quickly and are considered part of
pounds, each with its own characteristic decay rate under Phase 2 here. Lignin is a complex, decay-resistant, hetero-
a defined set of environmental conditions. Highly soluble polymer with many aromatic rings that is a structural com-
cell compounds such as some carbohydrates and polyphe- pound in vascular plants and that occurs with particular
nolics are leached from the plant litter in a matter of days high concentrations in wood, although it is also often proxi-
to weeks (Phase 1 in Fig. 7.6). Relatively simple compounds mally characterized as an acid-insoluble fraction of detritus.
such as many sugars, proteins, and some polysaccharides are Lignin typically has a turnover time of many years (Phase
rapidly catabolized (Phase 2). Large macromolecules such as 3). It can be closely bound with celluloses in cell walls and

190  CHAP TER Seven

53488p151-286.indd 190 8/19/14 9:06 AM


Phase 1

100 Phase 2 Phase 3 Phase 4

Soluble cell products FIGURE 7.6. Representative time


course of decomposition of litter.
The time scale varies by climate,
pH, degree of waterlogging, etc.
Leaching predominates during Phase
Mass remaining (% of original)

Non-lignified cellulose &


hemicellulose 1. During Phase 2, relatively labile
cell products are mineralized and
decomposition is regulated by the
amount of labile carbon and nutrients
in the substrate and exogenous soil
Microbial products
nutrient concentrations. During
Phase 3, decomposition is determined
Physically protected
soil organic matter by the amount of remaining lignin,
lignified carbohydrates, other
Lignified cellulose & hemicellulose
aromatic cell constituents, and fats
and waxes. Phase 4 represents long-
term soil stabilization mechanisms,
Lignin, other aromatic cell wall constituents (e.g.,
Sphagnum), fats and waxes which are currently in great
Humus dispute. Extensively adapted from
0 Berg and Laskowski (2006).
Time (years)

impede their decomposition. Fats and waxes, such as cutin ratio and total nitrogen content of litter were the primar-
(which forms the outer waxy layer on leaves), also have very ily controls over decomposition rates at the global level,
slow decomposition rates. Mineral protection of soil organic with climatic factors (temperature, precipitation, and lati-
matter has been shown to be very important in terrestrial tude) being of secondary importance through both direct
ecosystems, and there is no reason to assume that it is not effects and indirect effects through litter quality (Zhang et
equally important in wetlands with mineral soil. The forma- al. 2008). These explanatory variables described 87.5% of
tion of chemically recalcitrant humus compounds during the variance in k rates for this large global database.
Phase 4 is currently under great debate (see below). Similar factors control decomposition in wetlands. For
The relative proportion of these different plant com- example, in a review of numerous decomposition studies in
pounds are a primary determinant of the rate of decompo- peatlands, Bragazza et al. (2009) found that decomposition
sition (Fig. 7.6). For typical non-woody plant biomass, dur- rates were best correlated with the acid insoluble reside (~
ing the early stages of decomposition (Phase 2) there will lignin):N ratio for both vascular peatland plant litter (r = –​
be abundant labile carbon compounds, so the microbes that 0.46, p <0.05, n = 27) and for Sphagnum litter (r = –​0.66, p <
are responsible for decomposition will be primarily limited 0.05, n = 11). k values for peatland vascular plant litter were
by available nutrients, particularly nitrogen. During the later also positively correlated with mean annual temperature (r
stages of decomposition, most of the initial labile plant com- = 0.37, p <0.001) and mean total annual precipitation (r =
pounds have been degraded so the remaining lignin concen- 0.39, p <0.001), but k values for Sphagnum were not corre-
tration becomes increasingly important (Taylor et al. 1989). lated with climatic factors, suggesting that litter quality is
However, plant litter with very low initial concentrations of of paramount importance.
labile compounds, as is often found in low-nutrient envi- Sphagnum mosses have very low decomposition rates
ronments, may forego any period of nutrient limitation of (Table 7.1) and are responsible for a substantial fraction of
decomposition and immediately enter into Phase 3 (Fig. 7.6). the world’s peat accumulation, but they lack true lignin in
The initial carbon:nitrogen and lignin:nitrogen ratios of their tissues (Williams et al. 1998). The acid-insoluble frac-
plant litter are often the best predictors of mass loss rates tion in Sphagnum is comprised of cell wall polymers domi-
over the first few years of decomposition within a particu- nated by phenolic compounds, including “sphagnum acid”
lar study. The C:N ratio reflects the ratio of cytoplasm to (p-hydroxy-β[carboxymethyl]-cinnamic acid) (Williams et
cell wall material, so it is a measure of carbon quality as al. 1998), as well as the cell wall polysaccharide sphagnan,
well as the availability of nitrogen for decomposition. Simi- which acidifies the environment (Stalheim et al. 2009). The
larly, the lignin:N ratio is an integrated measure of carbon fructose:pentose carbohydrate ratio was found to be a good
quality and nitrogen concentration. For example, a recent predictor of moss decomposition in peatlands, with fruc-
summary of terrestrial decomposition studies in 110 sites, tosans being metabolic carbohydrates and pentoses being
including those in wetlands, indicated that the initial C:N structural carbohydrates (Turetsky et al. 2008). The former

Carbon Dy namics and Ecosystem Processes   191

53488p151-286.indd 191 8/19/14 9:06 AM


are more common in Sphagnum species inhabiting hollows alternating flooded and drained conditions may enhance
and the latter in hummock-inhabiting species, potentially decomposition because of factors such as increased leach-
explaining the slower decomposition of hummock species. ing, pulses of microbial activity, and amelioration of mois-
A number of other studies have also noted that C:P and C:N ture limitation under drier conditions (Lockaby et al. 1996a,
ratios significantly explain Sphagnum decomposition rates. 1996b; Moore et al. 2007; see also Chapter 4). Leaf litter typ-
Bragazza et al. (2009) suggested that increased atmospheric ically falls on the surface of wetlands, which is often not
nitrogen inputs into peatlands should increase Sphagnum inundated; this may explain why decomposition rates are
decomposition through an increase in both their tissue similar in wetlands and uplands in general (Table 7.1).
nitrogen content and also a decrease in soluble phenolic con- If nutrients are limiting for decomposition, it is reason-
tent, as phenolics can inhibit microbial activity (see below). able to assume that fertilization would lead to faster decay
Sphagnum has long been known to have important anti- rates. However, fertilization studies have found contradic-
biotic and inhibitory properties (van Breemen 1995; Ver- tory results regarding this hypothesis (see Bridgham and
hoeven and Toth 1995; Stalheim et al. 2009). Sphagnum- Lamberti 2009). The phosphorus eutrophication gradient
derived compounds may be particularly inhibitory to from agricultural run-off in the Everglades is one of the few
methanogens, with important implications for the often clear examples of exogenous nutrient limitation of decom-
low methane fluxes observed from bogs and for the poten- position in wetlands (DeBusk and Reddy 1998; Qualls and
tial of these peatlands to become more dominated by vas- Richardson 2000). Even in the Everglades, the positive
cular species with climate change (Bridgham et al. 2013). effect of eutrophication on NPP far outweighs its positive
effect on decomposition, so the net effect is a large increase
in peat accretion (Craft and Richardson 1993).
Exogenous Controls over Decomposition
The effects of nutrients on decomposition are complex,
Exogenous factors also control decomposition rates at a and understanding them requires taking a whole-ecosys-
local scale in wetlands; these factors include temperature tem approach. Figure 7.7 combines a number of lines of evi-
(determined both by climate and position in the soil pro- dence to propose very different sets of nutrient controls on
file), soil moisture availability, anaerobiosis, pH, and soil decomposition in high versus low nutrient environments.
nutrient availability. The effects of temperature and precip- Plants adapted to low nutrient environments have foliage
itation on decomposition rates in litterbag studies at the with low nutrient concentrations and low carbon quality
global scale in syntheses of the published literature were (i.e., high concentrations of the structural compounds lig-
described above. Decomposition of both fresh plant litter nin and cellulose and inhibitory compounds such as phe-
and older soil organic matter typically have a Q10 (increase nolics). These plants also have low inherent potential to
in rate for a 10 °C increase in temperature) that ranges respond positively to nutrient additions. In contrast, plants
from 2 to 4 (Brinson et al. 1981; Heal et al. 1981; Raich and adapted to high nutrient environments have the opposite
Schlesinger 1992). set of traits. In initially low nutrient environments, fertil-
High soil acidity (i.e., low pH) provides a strong environ- ization may cause an increase in nutrient uptake into plants
mental constraint on many microbial groups (Fierer and accompanied by increased growth (limiting nutrient), an
Jackson 2006). Numerous studies have shown that micro- increase in nutrient uptake into plants with no increase in
bial respiratory activity is inhibited at low pH in wetlands growth (luxury uptake), or no increase in nutrient uptake
(e.g., Benner et al. 1985; Kok et al. 1990; Ye et al. 2012). (non-limiting). In a fertilization study in a bog, interme-
However, studies that have manipulated pH in the field and diate fen, and rich fen, all three responses were observed
laboratory have not always found an inhibitory effect on (Iversen et al. 2010). Because plants from low nutrient
decomposition and nutrient cycling (Bridgham and Rich- environments have litter with low amounts of labile car-
ardson 1992, 2003; Chapin et al. 2004), suggesting that bon, decomposition rates may be unaffected by either high
some microbial groups and associated processes are adapted endogenous or exogenous nutrient availability until spe-
to low pH, or other controlling factors are more important cies that are adapted to higher nutrient availability become
than pH in some circumstances. established. For these plants, fertilization will result in
Decomposition rates decrease with depth in the soil pro- both increased litter production and increased endogenous
file because of greater duration of waterlogging, cooler tem- nutrient concentrations, which should increase decomposi-
peratures, and decreased carbon quality (Updegraff et al. tion rates because their foliage has high carbon quality and,
1995; Moore et al. 2007). Numerous studies have shown thus, the initial stages of decomposition are nutrient rather
that prolonged anaerobic conditions lead to substantially than carbon limited (Fig. 7.6).
reduced rates of decomposition (e.g., Godshalk and Wetzel
1978; Updegraff et al. 1995; Bridgham et al. 1998). The sub-
Nutrient Mineralization-Immobilization
stantial oxidation of soil organic matter in peatlands upon
during Decomposition
their drainage (Armentano and Menges 1986; Bridgham et
al. 2006) is compelling evidence of the importance of water- Carbon and nutrient cycling are intimately linked dur-
logging in soil carbon sequestration in wetlands. However, ing decomposition. The heterotrophic microbes that are

192  CHAP TER Seven

53488p151-286.indd 192 8/19/14 9:06 AM


availability

Nutrient
immobilization

A. Low Nutrient Net B. High Nutrient Net


Primary Primary
Environment production
Environment production
Increased Increased
0 (+) litter
0 (+) + litter
0 (+)
nutrient nutrient
content content
+ +
Increased Increased
Litter carbon Litter carbon
plant plant
quality (low) quality (high)
nutrient nutrient
(limiting or
content 0 (limiting content +
luxury uptake) — uptake)
+ (0) +
— +
Increasing Increasing
0 exogenous
+
exogenous Decomposition Decomposition
nutrient rate nutrient rate
availability availability

+ +

Nutrient Nutrient
immobilization immobilization

FIGURE 7.7. Conceptual model


Net of how changes in exogenous nutrient availability affect decomposition rates and nutrient immobilization in
B. High Nutrient
low-nutrient Primary
(A) and high-nutrient environments (B). Positive effects are shown by +, negative effects by −, and neutral effects by 0. Effects in
Environment production
parentheses indicate alternative possibilities based upon inconsistent results among previous studies. From Bridgham and Richardson (2003).
Increased
+ litter
0 (+)
nutrient
content
C:P
> 300 300 + 200 < 200
C:P > 300 300
Increased 200 < 200 Litter carbon
C:N > 30 plant 30 20 < 20 quality (high)
C:N > 30 30
nutrient 20 < 20
(limiting content +
uptake)
+
+ FIGURE 7.8. Time course of gross and

Increasing % Original N or P net immobilization and mineraliza-


exogenous
+ % Original N or P tion of nitrogen and phosphorus in lit-
Decomposition
nutrient Maximumrate net
Maximum net ter as a function of C:N and C:P ratios.
availability immobilization
immobilization ⇒

immobilization 100 If gross immobilization is greater than


Rate of gross immobilization ⇒

100
⇐ Percent of original C, N, or P
Maximum net gross mineralization, net immobi-
or gross mineralization

⇐ Percent of original C, N, or P
Maximum net
mineralization

+ mineralization lization occurs (area with horizon-


% Original C mineralization tal hatching); if gross mineralization
% Original C Nutrient is greater than gross immobiliza-
immobilization tion, gross mineralization occurs (area
with vertical hatching). The distance
of gross

Gross mineralization
Gross mineralization between the gross immobilization and
or gross

Net mineralization curves determines the


Net
immobilization Net mineralization relative magnitude of net immobiliza-
Rate

immobilization Net mineralization


tion or mineralization. The percentage
Gross immobilization of the original mass of nitrogen and
Gross immobilization
phosphorus in the litter reflects the
cumulative effect of these processes.
From Bridgham and Lamberti (2009).

Time ⇒
Time ⇒

responsible for decomposition are continually both min- opposite is true. Although both gross mineralization and
eralizing nutrients and carbon, i.e., converting them from immobilization are always simultaneously occurring, their
organic to inorganic forms, and immobilizing nutrients, relative proportions vary dramatically through the course
i.e., taking up nutrients as they build biomass (Fig. 7.8). of decomposition, with important implications for nutrient
The gross rates of mineralization and immobilization are cycling and availability.
the actual rates of both processes, but these rates are dif- During decomposition, the microbial community
ficult to measure (requiring isotopic tracers) because both respires organic carbon to carbon dioxide and retains essen-
are happening simultaneously. Thus, what is actually mea- tial nutrients for growth, particularly nitrogen and phos-
sured often is the net change in nutrient concentrations phorus. Thus, the C:N and C:P ratios of detritus (which
over time in plant litter and soil organic matter, or net min- operationally includes the plant material plus the associated
eralization or immobilization. Net mineralization occurs if microbial biomass that cannot be easily separated) decrease
the rate of gross mineralization is greater than the rate of through time. However, all organisms have stoichiometric
gross immobilization, and net immobilization occurs if the constraints on the C:nutrient ratios that are imposed by the

Carbon Dy namics and Ecosystem Processes   193

53488p151-286.indd 193 8/19/14 9:39 AM


cellular components of their biomass (Tezuka 1990; Elser (Megonigal et al. 2004; Bridgham et al. 2013). There are four
et al. 1996). Because of the large amount of holocellulose main processes in anaerobic heterotrophic carbon cycling:
and lignin in the cell walls of plant cells, the C:N and C:P the breakdown of complex biopolymers, fermentation, res-
ratios of plant litter are initially much higher than they piration, and acetoclastic methane production. There are
are in microbial biomass (Swift et al. 1979). Consequently, also three chemoautotrophic processes (i.e., processes that use
when plant litter with high initial amounts of labile car- the energy from reduced inorganic chemical compounds to
bon is decomposed, the microbial community actually fix carbon dioxide into biomass) that are an essential part
takes in nutrients from the surrounding environment (i.e., of anaerobic carbon cycling: methane production from car-
net immobilization occurs) and there is an increase in the bon dioxide and hydrogen, homoacetogenesis, and meth-
absolute amount of nutrients in the detritus (remember that, ane oxidation.
operationally, it is the plant litter plus the associated micro- The first step in anaerobic carbon cycling is the break-
bial biomass), and there is a decrease in soil nutrient avail- down of complex biopolymers by microbial exoenzymes
ability (Fig. 7.8). A switch to net mineralization occurs when into smaller compounds that can be transported across cell
the C:nutrient ratio of the detritus approaches that of the membranes (Fig. 7.9). There is an extensive literature on
microbes (i.e., the critical ratio). The critical ratios in bacte- these exoenzymes that is not discussed further here. The
rial cultures are about 15:1 for C:N and 60:1 for C:P (Tezuka second step is a series of fermentation reactions that form
1990), but substrates with more complex structural com- low-molecular-weight organic acids and alcohols, hydrogen
pounds, such as plant litter, have higher effective ratios. Ste- gas (H2), and carbon dioxide. The third step is heterotro-
venson (1986) suggested critical ratios in soil organic matter phic microbial respiration, where various inorganic elec-
of 20:1 to 30:1 for C:N and 200:1 to 300:1 for C:P. tron acceptors are coupled to the oxidation of organic car-
The immobilization phase of decomposition can be bon. The simplest way to understand the role of electron
thought of as a source-sink relationship between the lit- acceptors in respiration is to realize that, at their essence,
ter-microbe complex (the sink) and external or exogenous all heterotrophic reactions involve the oxidation of reduced
nutrient availability (the source). Litter that has lower ini- organic carbon compounds and the accompanying release
tial endogenous concentrations of a limiting nutrient has of energy for microbial growth and maintenance (see Chap-
a higher sink capacity, whereas higher soil nutrient avail- ter 4). However, all oxidation reactions must be coupled
ability provides greater source strength (Fig. 7.7) (Bridgham to a reduction reaction to proceed because there must be
and Richardson 2003; Moore et al. 2008). Moreover, the net a sink for the electrons produced in oxidation reactions
mineralization dynamics of phosphorus are complicated (see Chapters 2 and 4). In the case of aerobic respiration,
further by geochemical sorption of mineralized phospho- the terminal electron acceptor is oxygen, forming carbon
rus, which varies dramatically among different soil types dioxide, but oxygen is quickly consumed in the soil under
(Bridgham et al. 1998). waterlogged conditions, and thus microbes must use other
In contrast to nitrogen and phosphorus, calcium and electron acceptors (Fig. 7.9). In anaerobic respiration, the
magnesium are found in similar concentrations in micro- organic carbon is oxidized to carbon dioxide using the fol-
bial biomass and litter (Swift et al. 1979), so these com- lowing inorganic electron acceptors in order of thermody-
pounds tend to be lost at approximately the same rate as namic favorability: nitrate, Fe(III), manganese, and sulfate
organic carbon (Day Jr 1983; Bridgham and Richardson (see Chapter 2, Table 2.5). There is also mounting evidence
2003). In contrast to all four previously mentioned ele- that organic substances with aromatic structures can act as
ments, potassium is quite soluble and is rapidly removed electron acceptors, and these substances may be the domi-
through leaching. nant acceptors in peatlands (Bridgham et al. 2013; Keller
and Takagi 2013). Because the pathways for these terminal
electron acceptors are more thermodynamically favorable
Anaerobic Carbon Cycling
than pathways of methanogenesis, methane production
All wetlands experience periods of prolonged inundation is usually low in wetlands with large abundances of these
by definition, so anaerobic carbon cycling is a core wetland acceptors (Fig. 7.9). The availabilities of electron acceptors in
function. The two end products of anaerobic carbon miner- wetland soils are controlled by a complex set of factors, but
alization are carbon dioxide (CO2) and methane (CH4), both particularly important are the degree of mineral content of
of which are important greenhouse gases. Moreover, wet- the soil, oxidation events in the soil caused by mechanisms
lands are the largest natural source of methane, and there such as water table drawdown or oxygen leakage through
is great concern that with anthropogenic climate change, the rhizosphere, and the degree of oceanic influence. Sea-
wetlands may emit more methane in a positive feedback water has high sulfate concentrations, so sulfate reduction
cycle (see below and Chapter 10). is the major pathway of anaerobic carbon mineralization in
Organic carbon is broken down into its gaseous end prod- many salt marshes and mangroves (Bartlett et al. 1987; Pof-
ucts by a complicated series of steps that involve numerous fenbarger et al. 2011). High rates of atmospheric deposition
microbial functional groups that can be either competitive of sulfate can also reduce methane production in freshwater
or syntrophic (requiring each other for efficient growth) wetlands (Gauci et al. 2004).

194  CHAP TER Seven

53488p151-286.indd 194 8/19/14 9:06 AM


CH4 Flux

Aerenchyma

Ebullition

Diffusion
Litter
Inputs
Aerobic Zone Aerobic Methane CO2
Production CH4 Methane
Biopolymers Oxidation
(e.g., soil organic matter,
Anaerobic Zone cellulose)

Exo-cellular
Enzymes CH4 CO2
Anaerobic Methane
Monomers Oxidation
(e.g., glucose and other
simple sugars)

Microbial
Fermentation Alternative Terminal Electron Acceptors (TEAs)
Denitrification
LMW Fatty Acids and TEA Respiration NO3- NH4+, N2
Alcohols
(e.g., ethanol, Manganese Reduction
Mn (III, IV) Mn (II)
Root propionate, acetate)
Iron Reduction
Exudates Secondary Fermentation / Fe(III) Fe(II)
Acetogenesis
Humic Reduction
HSox HSRED
TEA Respiration
Sulfate Reduction
Acetate H2 + CO2 SO42- S2-, S0
Homoacetogenesis

Acetoclastic
Hydrogenotrophic
Methanogenesis
Methanogenesis

CH4

FIGURE 7.9. Anaerobic cycling in wetland ecosystems. Pools of carbon are shown in white boxes and solid arrows show the progressive
mineralization of these carbon pools by the identified microbial processes . Dotted lines illustrate carbon inputs from the plant community.
Dashed lines represent the flux of the gaseous end products of these processes (CH4 and CO2) to the atmosphere. TEA = terminal electron
acceptor. Adapted from Bridgham et al. (2013).

Methane is produced through the acetoclastic pathway in wetlands. However, acetate pooling is often observed in
where acetate is degraded into carbon dioxide and methane, bogs (Ye et al. 2012; Bridgham et al. 2013), so homoace-
and also through the hydrogenotrophic pathway where car- togenesis may provide a poorly understood limitation to
bon dioxide and hydrogen gas are converted into methane methane production in these wetlands.
(Fig. 7.9). The acetoclastic pathway is a heterotrophic reac- Methane oxidation to carbon dioxide in aerobic envi-
tion, whereas the hydrogenotrophic pathway is a chemoau- ronments is not an anaerobic process, but given the impor-
totrophic reaction. The acetoclastic pathway is dominant in tance of methane fluxes in wetlands, it has been extensively
most freshwater environments, but the hydrogenotrophic studied. It has been estimated that methane oxidation con-
pathway is important in many peatlands (Bridgham et al. sumes between 40% and 70% of gross methane production
2013). Under saline conditions, methanogens can also use (Megonigal et al. 2004). The pathway of methane transport
a variety of methylated compounds as substrates (Bridgham from the soil to the atmosphere is very important in terms
et al. 2013). of the proportion of methane that is oxidized. Methane is
Homoacetogenesis (i.e., the conversion of CO2 + H2  ace- transported by diffusion, ebullition (i.e., bubble release),
tate; Fig. 7.9) has not been well studied in wetlands because and plant-mediated transport. Diffusion results in maximal
it is thought to be thermodynamically unfavorable, but it methane oxidation, whereas ebullition results in almost no
may be more important than previously thought (Drake methane oxidation. In plant-mediated transport, methane
et al. 2009; Hädrich et al. 2012). Homoacetogens compete can be oxidized in the rhizosphere, but once the methane
with hydrogenotrophic methanogens for hydrogen but pro- is in the aerenchyma of the plant, it undergoes little addi-
vide acetate for acetoclastic methanogens, which may cause tional oxidation. Plant-mediated methane transport con-
little net change in the total amount of methane produced tributes from about 30% to 100% of the methane flux in

Carbon Dy namics and Ecosystem Processes   195

53488p151-286.indd 195 8/19/14 9:06 AM


wetlands (Bridgham et al. 2013). Consequently, important + CO2
- CH4
controls over methane emissions from wetlands are the Electron Plant
Acceptors Community
abundance of vascular plant with aerenchyma, which con- Oxygen + +
trols plant-mediated transport; and the water-table level, + + +/-
which controls diffusive flux. C and N + Microbes
+ Phenol
Oxidase
- Phenolics,
Anaerobic methane oxidation (AOM) has been docu- Availability Quinones
- -
mented for decades in marine systems and is thought to - + -
consume greater than 90% of the methane produced in Low pH + Hydrolases
+
these systems (Bridgham et al. 2013). Although the mech- +
anisms for AOM in marine systems are still under debate, +
electron acceptors other than oxygen—​particularly sul- Inorganic
Nutrients -
fate—​are involved. There is intriguing evidence that AOM CO2, CH4 DOM
Mineralization
may also be important in freshwater wetlands, but it is very
poorly studied and the mechanisms are unclear (Smemo
and Yavitt 2011; Blazewicz et al. 2012).
FIGURE 7.10. Multiple roles of aromatic substances in carbon
cycling in wetlands. Phenol oxidase activity is thought to be an
Why Do Wetlands Accumulate Soil Carbon? “enzymatic latch” over carbon storage in peatlands because of the
widespread negative effects of polyphenols on decomposition and
There are three potential, nonexclusive, reasons why many
microbial activity. This enzymatic latch mechanism is of unknown
wetlands, particularly peatlands, accumulate such large importance in other types of wetlands. Other aromatic compounds,
amounts of soil organic matter. The first is that wetland e.g., quinones, are important organic terminal electron acceptors.
plants have inherently low carbon quality and thus form As with other electron acceptors, they cause preferential anaerobic
low-quality soil organic matter. To understand this first mineralization to carbon dioxide versus methane. Enzymatic latch
mechanisms are shown in green, organic terminal electron acceptor
reason, it is instructive to compare wetland soils to terres-
mechanisms are shown in blue, controls over both mechanisms are
trial soils because of the much greater body of research on shown in pink, and generalized microbial controls are uncolored.
the latter. Humic substances are traditionally described as Extensively adapted from Freeman et al. (2012).
a heterogeneous group of high-molecular-weight, aromatic,
refractory organic compounds of secondary origin in soils
(Sposito 2008). Humic substances in both the dissolved and late soil carbon even though they experience long periods
solid phases are considered to occur at very high concentra- with a water table far from the surface during the growing
tions in wetlands (Kracht and Gleixner 2000; Collins and season (e.g., Bridgham and Richardson 1992). Thus, there
Kuehl 2001). These substances have been thought to be fun- is evidence that wetland soils may have unique properties
damentally chemically recalcitrant to microbial attack and of their soil organic matter that promote carbon accumu-
to be responsible for much of the “old” soil carbon in ter- lation, although almost all wetlands experience large oxi-
restrial and wetland soils. However, the very existence of dative losses of soil carbon upon drainage, suggesting that
humic substances as traditionally envisioned and the idea anaerobiosis is essential for them to maintain and grow
of inherent chemical recalcitrance of soil organic matter their soil carbon stores.
have been resoundingly attacked recently in the context of The second hypothesis to explain the large amounts of
terrestrial soils (e.g., Kleber and Johnson 2010; Schmidt et soil organic matter accumulated in many wetlands is that
al. 2011). Organic-mineral soil interactions increasingly are the low energy yields of the anaerobic pathways of carbon
being recognized as the primary mechanism of the stabili- mineralization reduce microbial activity enough to allow
zation of soil organic matter in terrestrial soils. This may be for soil carbon accumulation. As discussed above, abundant
the case in wetlands with mineral soils, although it has yet evidence shows that anaerobic carbon mineralization is
to be studied, but it cannot be the case for peatlands, which several-fold slower than aerobic carbon mineralization, but
by definition have very low amounts of mineral matter. this seems unlikely to be a large enough effect to account
As discussed above, research on Sphagnum mosses shows for the large amounts of soil carbon accumulation seen in
that they have a unique cell wall chemistry and suggests many wetlands.
that they form organic matter with a high degree of aro- The third hypothesis is what has been termed the enzyme
matic structures; they may also have direct inhibitory latch mechanism (Fig. 7.10) (Freeman et al. 2001; Limpens
effects on soil organisms. High amounts of phenolic com- et al. 2008). As discussed above, polyphenols bind many
pounds in some wetland plants may also impede decom- enzymes and make them inactive. Phenol oxidase and per-
position of plant litter (see, e.g., Bridgham and Richardson oxidase exoenzymes are responsible for the degradation
2003). Polyphenols inhibit carbon mineralization by inhib- of polyphenolic and aromatic compounds and are pro-
iting microorganisms, binding proteins and polysaccha- duced by a range of bacteria and fungi. However, the activ-
rides, and inactivating enzymes (Harborne 1997; Freeman ity of these enzymes is constrained by low oxygen avail-
et al. 2012). Moreover, some peatlands continue to accumu- ability, low pH, and low temperature, all of which are very

196  CHAP TER Seven

53488p151-286.indd 196 8/19/14 9:06 AM


common in many wetland soils. The basic premise of the conditions, its recalcitrant chemical nature slows this pro-
enzyme latch hypothesis is that the inactivity of phenol cess. DOM is also readily adsorbed to solid organic matter,
oxidase in anaerobic wetland soils allows for the accumula- soil minerals, and even bacterial cells (Young et al. 2004;
tion of polyphenolic substances, which in turn make hydro- Moore 2009). The rate of discharge from wetlands is par-
lase exoenzymes inactive, thus reducing nutrient availabil- ticularly important in the dynamics of DOM export, with
ity and carbon availability. In turn, this reduces microbial high discharge rates causing rapid export of DOM, whereas
growth and activity, which spirals the system into a state of longer retention times result in the DOM being susceptible
very low carbon mineralization and high carbon sequestra- to biodegradation and retention (Pastor et al. 2003; Moore
tion capacity. 2009). There is often a strong relationship between DOM
transport from peatlands and water-table depth because of
the much lower hydraulic conductivity of deeply buried
Dissolved Organic Matter Fluxes
peat (Moore 2009). Factors that increase rates of biodegra-
from Wetlands
dation will tend to increase the ratio of the production of
Dissolved organic matter (surface water DOM concentra- carbon dioxide to DOM, such as higher temperatures, lon-
tions) is an essential component of the carbon dynam- ger hydraulic residence times, and aerobic versus anaerobic
ics of wetlands (Figs. 7.1, 7.2, and 7.10) and can reach very conditions (Moore 2009).
high concentrations in wetland soils—​for example, ranging A long-term increase in surface water concentrations
between 20 to > 60 mg C L –​1 in northern peatlands (Blodau has been observed in many boreal and sub-boreal regions
2002). DOM is thought to be a heterogeneous mixture of and that has been at least partly tied to wetland dynam-
compounds that is typically dominated by high-molecular- ics (Fenner et al. 2009). Possible reasons for the increase
weight compounds (e.g., “humic” and “fulvic” acids) that include greater DOM production due to warming, increased
decompose very slowly (Docherty et al. 2006; Tfaily et al. drought frequency, and elevated atmospheric carbon diox-
2013), although pooling of low-molecular-weight acetate ide concentrations. Decreasing atmospheric sulfate deposi-
(CH3COOH) is sometimes observed in northern bogs for tion may lead to both enhanced DOM production and solu-
reasons that are poorly understood (Bridgham et al. 2013). bility. Melting permafrost may also cause increased release
High-molecular-weight DOM compounds have light-absorb- of DOM (Moore 2009).
ing (chromophoric) properties that give waters with high
concentrations of these compounds a tea-stained color.
Wetlands are a major source of DOM in surface waters, The Big Picture
and the proportion of wetlands in a catchment is often the
Methane Fluxes
best landscape predictor of surface water DOM concentra-
tions (Xenopoulos et al. 2003; Johnston et al. 2008; Moore Methane (CH4) is a major product of anaerobic microbial
2009). DOM in surface waters has myriad important effects, respiration in wetlands (Fig. 7.9), and its greenhouse gas
including degradation of drinking water quality (Fenner effects are about 25 times stronger than carbon dioxide on
et al. 2009), mercury methylation and transport (Moore a mass basis (Forster et al. 2007; see also Chapter 10). It is
2009), fueling of heterotrophic metabolism and control also responsible for about 18% of anthropogenic warming
over foodweb dynamics (Docherty et al. 2006; Frost et al. (Forster et al. 2007). Of the total global flux of 500 to 600
2007), and attenuation of light and harmful ultraviolet Tg CH4 yr–​1, wetlands are responsible for about 164 Tg CH4
radiation (Frost et al. 2006). yr–​1 (range from 80 to 280 depending on study and meth-
Export of DOM ranges from <1 to 50 g C m –​2 yr–​1 in odology, Fig. 7.11). An estimated 73% of these emissions are
northern peatlands (Blodau 2002); potentially, even higher from tropical wetlands, with their large areal extent and
DOM export can occur in bottomland hardwood forests high flux rates (Bridgham et al. 2013). Open freshwater
(Trettin and Jurgensen 2003). However, emergent wet- aquatic ecosystems also emit 93 Tg CH4 yr–​1, with another
lands may be net sinks for DOM (Blodau 2002; Johnston 10 Tg CH4 yr–​1 coming from shallow littoral zones, but there
et al. 2008), and peatlands can be net sinks for DOM at may be substantial overlap between freshwater aquatic eco-
low rates of discharge (Pastor et al. 2003). The net trans- systems and wetlands in these estimates. Rice fields are
port of DOM from wetlands is controlled by the interactive responsible for about 53 Tg CH4 yr–​1, and these are essen-
effects of production, consumption, sorption, and trans- tially agricultural wetlands with many of the same underly-
port mechanisms (Fenner et al. 2009; Moore 2009). DOM is ing biogeochemical controls over methane fluxes.
produced through the decomposition of soil organic matter There is substantial concern that wetlands may pro-
and release of plant exudates. Even at great depth within vide a positive feedback to future climate change for sev-
peatlands, DOM is of relatively recent origin compared eral reasons. First, CH4 production is a temperature-limited
with the solid phase organic matter, demonstrating the process—as are all microbially mediated biogeochemi-
importance of plant sources of DOM (Chanton et al. 2008). cal pathways. Second, a large concentration of wetlands
DOM is consumed through heterotrophic microbial respira- exists at high northern latitudes where the largest tempera-
tion, although, as discussed above, at least under anaerobic ture increases are predicted to occur, and the possibility of

Carbon Dy namics and Ecosystem Processes   197

53488p151-286.indd 197 8/19/14 9:06 AM


(a)

(b)

FIGURE 7.11. Dot density graph of global methane sources. Horizontal FIGURE 7.12. (A) Atmospheric growth rate of methane in dry air mole
lines are the median for each category. Anthropogenic sources fractions in blue and the de-seasonalized trend curve as a red, –​
include rice fields and natural sources include freshwater aquatic dashed line. (B) The instantaneous growth rate of (A). The symbols
ecosystems and wetlands, but they are also presented separately. are the annual increase calculated from January 1 in one year to
Adapted from Bridgham et al. (2013). January 1 in the next year, plotted in the middle of the year. Data
and graphic are from E. Dlugokencky (Dlugokencky et al. 2009).

widespread permafrost melting in the arctic is of special ecosystems, whereas the value for salt marshes in Table 7.2
concern. Third, the stimulatory effects of increased atmo- probably is an underestimate because it is calculated only
spheric carbon dioxide concentrations on temperature may to 1 m depth. Also, the organic content of salt marsh soils
also increase methane fluxes (Bridgham et al. 2013 and ref- varies substantially (Chmura et al. 2003).
erences therein). Wetlands contain 33% of the global soil organic pool to
There is substantial evidence that methane emissions 2 m depth according to the calculations in Table 7.2. Peat-
from wetlands have affected climate and been at least par- lands alone contain 27% of the global soil carbon pool.
tially responsible as a feedback mechanism driving past These estimates are based upon soil carbon densities by Job-
glacial-interglacial cycles (Loulergue et al. 2008). Moreover, bagy and Jackson (2000) in which wetlands were included
although longer-term recent trends in atmospheric concen- within the vegetation categories used to define the vari-
trations of methane appear to be driven by anthropogenic ous global terrestrial ecosystem types. However, since most
activities (Fig. 7.12), numerous studies suggest that much peatlands occur in the boreal zone, and the total soil car-
of the large interannual variations in atmospheric meth- bon estimated for the peatlands is substantially greater
ane concentrations, and potentially the large increases seen than the total for boreal forests and shrub lands—​which
since 2008, are due to climate effects on methane emissions should be the categories that largely include peatlands, but
from wetlands (Bridgham et al. 2013). it appears that peatlands were not adequately represented
in the soil carbon densities determined by Jobbagy and
Jackson (2000). Thus, wetland soil carbon was added to the
Global Soil Carbon Pools and Accumulation
terrestrial carbon pool to calculate a global total in Table
The soil carbon pool stored with in a particular ecosystem 7.2, potentially underestimating the contribution of wet-
depends on its area and the amount of soil carbon per unit lands. For comparison, if one uses the global soil carbon
surface area, termed soil carbon density. Wetlands occupy pool estimate of 2,416 Pg (1 Pg = 1015) by (Batjes 1996), then
only about 6% of the global terrestrial surface, but some wetlands contain about 25% of the soil carbon pool.
wetland types have very high soil carbon densities (Table All wetland types apparently have much higher soil car-
7.2). Soil carbon density is particularly impressive in peat- bon accumulation rates than terrestrial ecosystems have
lands at 1,497 Mg C ha –​1 (I Mg = 106) and in mangroves (Table 7.2), probably reflecting their prolonged water-
at 864 Mg C ha –​1, where the depth of the organic layer is logged conditions and their anaerobic pathways of carbon
often many meters thick. These values are 6.3-fold and 3.6- mineralization. Soil carbon accumulation rates in terres-
fold greater, respectively, than the highest value found in trial ecosystems decline exponentially with age (Zehet-
any terrestrial ecosystem. However, carbon pools and fluxes ner 2010), so that ecosystems older than several thousand
vary greatly among different types of wetlands. For exam- years have very low accumulation rates. Schlesinger (1990)
ple, the soil carbon densities of freshwater mineral-soil wet- estimated that the overall rate of soil carbon accumulation
lands and seagrass beds are in the range of other terrestrial in deglaciated soils over the last 10,000 years is less than

198  CHAP TER Seven

53488p151-286.indd 198 8/19/14 9:06 AM


table 7.2
Area and Soil Organic Carbon Pools and Accumulation in Global Wetlands and Other Ecosystems

Soil Carbon Global Soil Soil Carbon Soil Carbon


Area Density Organic Carbon Accumulation Rates Accumulation
Ecosystem Type (km2 × 1,000) (Mg C ha-1) Pool (Pg) (g C m-2 yr-1) (Tg C yr-1)

Wetlands
Peatlands 3,341a 1,497h 500n 11n 36h
Freshwater
mineral-soil 5,172b 199i 103h 17p 88p
Salt marsh 51c 162j 0.8h 151q 7.7h
Mangrove 138d 864k 12h 163r 23r
Seagrass beds 177e 140l 2.5h 101s 18h
Wetland total 8,879 618 173

Global Terrestrial
Ecosystems
Tropical forests 19,623f 238m 466h 4.0f 79f
Temperate forests 10,400f 196m 204h 5.1f 53f
Boreal forests 13,700f 117m 160h 4.6h 49f
Tropical savanna/
grasslands 15,000g 187m 280h
Temperate
grasslands 9,000g 159m 143h 2.2g 20h
Shrublands 8,500g 122m 104 h
Deserts 18,000g 91m 164h 0.8g 14h
Tundra 8,000g 166m 133h 1.2g 10h
Cropland 14,000g 150m 210h
Extreme desert, rock, ice 24,000g 1g 2.4h

Overall total 149,102 1,867o 400t

Wetland % of total 6.0 33o 43

From assessment of published values.


a o Wetlands are supposedly included in the terrestrial ecosystem classes

Wetland area from Lehner and Döll (2004) excluding coastal wet-
b for both area and soil carbon density, but a comparison of peatlands to
lands minus peatland area in table. boreal forest suggests that wetlands are poorly represented. Hence, wet-
c Pendleton et al. (2012). land soil carbon was added to the terrestrial soil carbon pool, potentially
d Giri et al. (2011). underestimating the wetland contribution. For comparison, Batjes (1996)
e Waycott et al. (2009). estimated the global soil carbon pool to 2 m depth to be 2,416 Pg, in
f McLeod et al. (2011). which case wetlands would represent 25.6% of the global total.
g Schlesinger (1997). p Using sedimentation rate of 2.2 Mg sediment ha-1 yr-1 and 7.7% C

h Calculated from other values in table. as in Bridgham et al. (2006), but the global rate is higher because of the
i Soil carbon density to 2 m depth from Batjes (1996). larger assumed area.
j Soil carbon density to 1 m depth from Pendleton et al. (2012). q Duarte et al. (2005).

k Soil carbon density from Donato et al. (2011). r Breithaupt et al. (2012).

l Soil carbon density from Fourqurean et al. (2012). s Average from Duarte et al. (2005) using sediment burial rates and

m Jobbagy and Jackson (2000) to 2 m depth. Duarte et al. (2010) using net ecosystem exchange.
n Yu (2012). t A maximum value from Schlesinger (1990).

53488p151-286.indd 199 8/19/14 9:06 AM


FIGURE 7.13. (A) The current radiative balance and
(B) historical net radiative forcing for wetlands
in the conterminous United States as a result
of soil carbon sequestration, oxidation of soil
carbon from drained wetlands, methane flux,
and the sum of these fluxes. Units are Tg CO2 -C
equivalents yr-1. A warming effect is represented as
negative number, and a cooling effect as a positive
number. Data from Bridgham et al. (2006).

2.4 g m–​2 yr–​1. However, very young terrestrial soils (e.g., their conservation, and even in selling carbon credits for
after volcanic eruptions) and those in abandoned agricul- wetland restoration activities (Galatowitsch 2009; Irving et
ture fields can have accumulation rates that rival those in al. 2011; Pendleton et al. 2012). However, caution is needed
wetlands (Schlesinger 1997). when using carbon sequestration as a rationale for wetland
The size of the soil carbon pool within a wetland type conservation and restoration activities. Two important
does not necessarily reflect its annual rate of soil carbon caveats must be weighed against the ability of wetlands to
accumulation. Peatlands have a comparatively small rate of sequester large amounts of carbon. First, the drainage of
annual soil carbon sequestration, but that carbon has typ- wetlands leads to the eventual oxidation of the stored car-
ically accumulated over several thousand years (Yu 2011) bon back into the atmosphere as carbon dioxide. Second,
and yields very large stores. The modest rates in freshwa- wetlands emit large amounts of the potent greenhouse gas
ter mineral-soil wetlands may be substantially overesti- methane, as discussed above. An estimate of the current
mated for reasons discussed in Bridgham et al. (2006). It radiative balance of wetlands in the conterminous United
appears that marine-associated wetlands have exception- States is shown in Figure 7.13, where methane fluxes have
ally high soil carbon accumulation rates for reasons that are been multiplied by their global warming potential on a
not clear. Wetlands sequester 173 Tg C yr–​1 (1 Tg = 1012) in 100-year timeframe (i.e., 6.3 on a mass C basis) into CO2 -C
their soils, which is 43% of the global total. Freshwater min- equivalents (Bridgham et al. 2006). According to this anal-
eral-soil wetlands are particularly important in this regard ysis, peatlands, freshwater mineral-soil (FWMS) wetlands,
because of their large global areas. Wetlands can also store and estuarine wetlands have net global warming balances
substantial amounts of carbon in aboveground plant bio- of –​10.2, –​5.1, and 5.2 Tg CO2 -C equivalents yr–​1, respec-
mass, particularly in those with early successional woody tively, where a negative number indicates a net warming
vegetation (Trettin and Jurgensen 2003). effect. Thus, the carbon sequestration capacity of peatlands
There is substantial recent interest in using the high car- and FWMS wetlands is more than offset by the oxidation of
bon sequestration capacity of wetlands as justification for peat in drained peatlands and methane fluxes in both wet-

200  CHAP TER Seven

53488p151-286.indd 200 8/19/14 9:06 AM


land types. However, the low methane fluxes in estuarine Even more caveats need to be considered here. Many
wetlands—​because of their high concentrations of porewa- of these estimates have very low confidence in their mag-
ter sulfate—​cause them to have a net cooling effect. nitudes (Bridgham et al. 2006), although the signs of the
Another important caveat to consider is the differ- net carbon balance and forcing for each wetland type
ence between a radiative balance and a radiative forcing are probably correct. The role of wetlands in the carbon
(Bridgham et al. 2006). The radiative balance reflects the balance and its subsequent effect on radiative forcing in
static effect of a substance, whereas the radiative forcing North America is relatively small compared with that of
refers to an externally imposed perturbation on the Earth’s other ecosystems (Pacala et al. 2007). Maybe most impor-
radiative energy balance (Ramaswamy et al. 2001). Thus, tantly, wetlands have many functions in the landscape
a change in the radiative balance leads to a radiative forc- that far outweigh their effects on carbon cycles and trace
ing, which in turn causes a change in temperature. Wet- gas emissions. However, the analysis presented in Figure
lands have long emitted large amounts of methane to the 7.13 indicates the unintended conundrums that can result
atmosphere, and it is only the change in methane flux that from narrow policy prescriptions based upon the carbon
causes a radiative forcing. Thus, increased methane fluxes dynamics and radiative forcing effects of wetlands. Much
caused by climate change, or potentially even by resto- more work needs to be done on these questions, particu-
ration activities, would have a warming effect, whereas larly in restored wetlands that almost certainly have vari-
destruction of wetlands would have a cooling effect if the able carbon balances dependent upon their hydrology,
decrease in methane fluxes outweighed the loss in their car- vegetation, time since restoration, climate, etc. Because of
bon sequestration rates. In peatlands in the conterminous their low methane fluxes and large soil carbon sequestra-
United States, the oxidation of soil carbon due to drainage tion potential, the restoration of marine-influenced wet-
(a warming effect) has had a much larger effect on radia- lands such as salt marshes, mangroves, and seagrass beds
tive forcing than their loss in sequestration capacity (also is likely to have a large net cooling effect (Bridgham et al.
a warming effect) and the decrease in their methane fluxes 2006; Fourqurean et al. 2012; Pendleton et al. 2012). This
(a cooling effect) (Fig. 7.13). The drainage of peatlands has may well be true of other wetlands such as young succes-
had a net warming effect of 12.7 Tg CO2 -C equivalents yr–​1. sional, forested, FWMS wetlands that experience only brief
In contrast, the loss of FWMS wetlands in the contermi- periods with water tables close to the soil surface. Preserva-
nous United States has dramatically reduced their meth- tion of wetlands with large soil carbon accumulation, such
ane fluxes, which more than offsets their loss in soil carbon as peatlands and mangroves, will reduce the potential for
sequestration (oxidation is unimportant), so the net effect large soil oxidation losses due to drainage. Other types of
is a cooling of 18.3 Tg CO2 -C equivalents yr–​1. The loss of wetlands may have a minimal or warming effect because of
estuarine wetlands is estimated to have a small warming restoration activities, so other functions and values need to
effect of 0.5 TG CO2 -C equivalents yr–​1. be emphasized in these cases.

Carbon Dy namics and Ecosystem Processes   201

53488p151-286.indd 201 8/19/14 9:06 AM


53488p151-286.indd
View publication stats 202 8/19/14 9:06 AM

You might also like