ArticlePDF Available

Abstract and Figures

Tropical peat carbon compound composition (CCC) is a highly understudied subject. Advanced understanding of peat CCC and carbon dynamics in differing conditions is desperately needed due to large-scale utilization of these peatlands. We studied the CCC—i.e. the hemicellulosic carbohydrate and uronic acid composition and concentrations of extractives, cellulose, acid-soluble lignin and acid-insoluble lignin—in association with peat profile depth and physical structure of peat, under representative, common land uses. Samples were gathered from an undrained forest and three sites altered 20–30 years prior to the study, which in aggregate form a continuum of increasing land-use intensity (drainage-affected forest; drained and deforested degraded open site; drained and deforested site under cultivation) in Central Kalimantan, Indonesia. Peat samples were taken from depths between 10 and 115 cm that covered mostly oxic, frequently waterlogged and permanently waterlogged, anoxic conditions. Our results demonstrated greater modification of peat properties when both vegetation and hydrological conditions were altered. The differences between sites were mainly present in the topmost peat and decreased with depth. Peat located at the surface contained more labile compounds (hemicelluloses, extractives, uronic acids, cellulose) on forest sites than at the most intensively altered open sites, where peat was enriched with recalcitrant acid insoluble lignin. The effect of drainage was evident in the drained forest site, where at the approximate median water table depth peat more closely resembled open sites in terms of the peat properties. The increased recalcitrance of peat in reclaimed areas has been a result of enhanced decomposition, reduced litter input rates and, at open sites also by repeated fires.
This content is subject to copyright. Terms and conditions apply.
ORIGINAL PAPER
Land use increases the recalcitrance of tropical peat
M. Ko
¨no
¨nen .J. Jauhiainen .R. Laiho .P. Spetz .
K. Kusin .S. Limin .H. Vasander
Received: 20 November 2015 / Accepted: 23 June 2016
ÓSpringer Science+Business Media Dordrecht 2016
Abstract Tropical peat carbon compound composi-
tion (CCC) is a highly understudied subject. Advanced
understanding of peat CCC and carbon dynamics in
differing conditions is desperately needed due to large-
scale utilization of these peatlands. We studied the
CCC—i.e. the hemicellulosic carbohydrate and uronic
acid composition and concentrations of extractives,
cellulose, acid-soluble lignin and acid-insoluble lig-
nin—in association with peat profile depth and phys-
ical structure of peat, under representative, common
land uses. Samples were gathered from an undrained
forest and three sites altered 20–30 years prior to the
study, which in aggregate form a continuum of
increasing land-use intensity (drainage-affected forest;
drained and deforested degraded open site; drained and
deforested site under cultivation) in Central Kaliman-
tan, Indonesia. Peat samples were taken from depths
between 10 and 115 cm that covered mostly oxic,
frequently waterlogged and permanently waterlogged,
anoxic conditions. Our results demonstrated greater
modification of peat properties when both vegetation
and hydrological conditions were altered. The differ-
ences between sites were mainly present in the topmost
peat and decreased with depth. Peat located at the
surface contained more labile compounds (hemicellu-
loses, extractives, uronic acids, cellulose) on forest
sites than at the most intensively altered open sites,
where peat was enriched with recalcitrant acid insol-
uble lignin. The effect of drainage was evident in the
drained forest site, where at the approximate median
water table depth peat more closely resembled open
sites in terms of the peat properties. The increased
recalcitrance of peat in reclaimed areas has been a
result of enhanced decomposition, reduced litter input
rates and, at open sites also by repeated fires.
Keywords Tropical peat Land-use
Decomposability Carbohydrates Polymers
Introduction
The carbon compound composition (CCC) of organic
matter determines the amount and quality of energy
Electronic supplementary material The online version of
this article (doi:10.1007/s11273-016-9498-7) contains supple-
mentary material, which is available to authorized users.
M. Ko
¨no
¨nen (&)J. Jauhiainen H. Vasander
Department of Forest Sciences, University of Helsinki,
Latokartanonkaari 7, P.O. Box27, 00014 Helsinki, Finland
e-mail: mari.kononen@helsinki.fi
R. Laiho
Natural Resources Institute Finland, Kaironiementie 15,
39700 Parkano, Finland
P. Spetz
Natural Resources Institute Finland, Jokiniemenkuja 1,
01370 Vantaa, Finland
K. Kusin S. Limin
Center for International Cooperation in Sustainable
Management of Tropical Peatland (CIMTROP),
University of Palangkaraya, Palangkaraya, Indonesia
123
Wetlands Ecol Manage
DOI 10.1007/s11273-016-9498-7
available for decomposers (i.e. decomposability), and
is thus an important factor regulating decomposition
(Berg 2000; Bragazza et al. 2007; Zhang et al. 2008;
Sanaullah et al. 2010; Strakova
´et al. 2011). In
particular, the amount of relatively labile C com-
pounds, such as hemicelluloses and cellulose, and
their ratio to more decomposition-resistant (recalci-
trant) polymers, such as lignin, together with nitrogen
(N) and phosphorus (P) concentrations influence the
rate at which decomposition progresses (Berg 2000;
Rejmankova 2001; Strakova
´et al. 2011; Talbot and
Treseder 2012). Organic matter CCC generally firstly
depends on the characteristics of the parent material,
e.g., litter type (woody, roots, leaf, branches, etc.), and
later on the decomposition stage (Berg 2000). Peat is
organic matter formed mainly of partly decomposed
plant litter. Woody, carbon (C)-rich peat is formed in
tropical ombrotrophic forested swamps in Southeast
Asia from incompletely decomposed above- and
below-ground vegetation mainly originated from trees
(Page et al. 1999;Wu
¨st et al. 2008; Hoyos-Santillan
et al. 2015). These thick peat deposits are significant C
stores containing 11–14 % of all C stored in peat soils
globally (Page et al. 2011). However, large-scale land
reclamation including drainage and deforestation has
taken place during the past few decades (Miettinen
et al. 2016). Land reclamation has drastically influ-
enced peat decomposition processes and the quantity
and quality of organic matter inputs leading to
enhanced peat loss through microbial decomposition
and wildfires (Hoscilo et al. 2011; Hooijer et al. 2012;
IPCC 2014), and thus land reclamation is the largest
modifier of peat CCC. An understanding of the impact
of various anthropogenic activities (e.g. drainage,
deforestation, fire, agriculture) on CCC is needed to
understand and predict soil organic matter stability in
tropical peat soils.
Peat is generally formed when the rate of decom-
position does not exceed the litter input rate. Decom-
position in peat soils is regulated by abiotic
environmental factors such as oxygen, moisture,
nutrient availability and pH (Rieley and Page 2005;
Laiho 2006). In an undisturbed peat swamp forest,
there is a typically high soil water-table level (WL).
This creates anoxic conditions in the peat below the
water level where decomposition is slower than the
aerobic decomposition, which takes place in oxic
conditions above the water table. Thus peat accumu-
lates over time (Clymo 1984). These abiotic factors
determine to what extent the decomposability of the
litter inputs or the accumulated peat is realized as
actual decomposition through microbial enzymatic
activity (Freeman et al. 2001; Sjo
¨gersten et al. 2011),
and they may be greatly altered by land use.
Land use reclamation in tropical peatlands usually
includes drainage and deforestation. Drainage lowers
the WL, allowing an increasing potential for aerobic
decomposition deeper in the peat profile. The physical
structure of peat changes due to advancing decompo-
sition, which leads to increased peat dry bulk density
(BD) and decreased porosity (Hooijer et al. 2012;
Ko
¨no
¨nen et al. 2015). Deforestation lowers litter input
rates and alters litter quality, since reclaimed land
uses, such as plantations, cultivated and fallow lands,
typically have lower vegetation cover and biomass
production and less variation in species (crops and
monocultures) than the original peat swamp forests
(Sulistiyanto 2004; Hoscilo et al. 2011; Mehta et al.
2013; Blackham et al. 2014; Yule et al. 2016).
Reclaimed, unmanaged peatlands are vulnerable to
wildfires, and up to several decimetres of peat can be
lost in recurring wildfires (Hoscilo et al. 2011;
Konecny et al. 2016), which results in exposure of
aged peat from the deeper layers and keeps the
vegetation cover low. All these changes in turn likely
contribute to changes in peat CCC.
Despite the increasing utilization of tropical peat-
lands (Miettinen and Liew 2016; Murdiyarso et al.
2010), there is almost no information available
concerning tropical peat CCC from any land-use type.
As far as we know, tropical peat CCC has previously
been reported only by Andriesse (1988). Yet, the
connection between the land-use type and peat CCC
should be understood to minimize both the loss of
substrate and the release of greenhouse gases (GHG).
Here we conduct a detailed analysis on tropical peat
CCC in different land-use types in association with
depth in peat profile and peat physical structure. We
established study sites at both (i) an undrained peat
swamp forest, and at three sites altered by human
activities. The altered sites included in order of
increasing land-use intensity: (ii) a drained forest,
and two sites at clear-cut open peatlands, i.e. (iii) a
drained site burned several times and (iv) a cultivated
site under controlled drainage. To determine the peat
CCC we analysed the concentrations of C, N, neutral
and acidic sugars in non-cellulosic polysaccharides
(hemicellulose), cellulose, extractives, acid-soluble
Wetlands Ecol Manage
123
lignin (ASL) and acid-insoluble lignin (AIL). AIL
forms the most recalcitrant portion of organic matter
and is often referred to as Klason’s lignin. The CCC
data were combined with the peat physical structure
data, defined both by peat BD and by the proportion of
different sized particles forming the peat, as an
indication of the physical decomposition stage of the
peat.
We hypothesized that (i) the impact of land use on
CCC is greatest when both WL and vegetation are
altered, (ii) the differences between land-use types are
greatest in the topmost peat, (iii) the differences
between land-use types can be seen as an increase in
the amount of resistant biopolymers in the peat and
(iv) the peat physical properties implying decompo-
sition stage will correlate with the peat CCC/recalci-
trance level of the peat.
Materials and methods
Study area and sites
The study area was an ombrotrophic lowland peatland
in Central Kalimantan, Indonesia (2°200S, 113°550E).
The reported mean annual temperature in the area was
26.2 ±0.3 °C and the mean annual rainfall was
2540 ±596 mm year
-1
between 2002 and 2010
(Sundari et al. 2012). Climatically, there are two rainy
seasons per year divided by one dry season typically
occurring from June through to the end of August.
We established four study sites located within a
10-km radius in the same watershed of the Sebangau
River. All sites can be assumed to originate from
similar ombrotrophic peat swamp forests, where the
peat depth exceeded 3 m (Fig. 1). The original
Fig. 1 Study area and
sampling sites in Central
Kalimantan, Indonesia
(2°200S, 113°550E)
Wetlands Ecol Manage
123
hydrology at three of the sites was altered due to
human activities, while the vegetation was also altered
at two sites. Two of the sites had original, close to
steady state forest cover (comprising species of a
number of plant families, including Dipterocarpaceae
sp.), although selective logging had occurred in the
past in both areas (Page et al. 1999; Jauhiainen et al.
2005,2008). These forest sites are called undrained
(UF) and drained forest (DF). The hydrology of the UF
was close to natural, but a large drainage canal has
impacted the WL of the drained forest since 1997. Soil
surface microtopography of both forests was formed
of a mosaic of low peat surfaces (forming pools when
high WL) and hummocks around tree bases (Lampela
et al. 2014), although the microtopography was less
apparent in the DF. The other two sites were open due
to deforestation, relatively flat, and both were drained.
The first open site, the degraded site (DO), was
deforested and had been drained by the same large
canal as the drained forest in 1997. The surface at the
degraded site had burned several times (1997, 1999,
2002, and 2006) prior to sampling. The recurring fires
had removed more than 0.5 m of the topmost peat
(Page et al. 1999; Hoscilo et al. 2011), and the
topography was therefore relatively flat excluding
some deeper fire scars. The dominant vegetation at the
degraded site was formed of ferns, and scattered
shrubs and small trees. The second open site, called the
agricultural site (AO), had been under shallow,
controlled drainage and cultivated by smallholder
farmers since the 1980s, but it had been unutilised
fallow land prior to sampling. The agricultural site had
been fertilized during cultivation. Due to repeated soil
surface tillage and management fires the soil
topography was flat. Details of the average annual
long-term WL at the sites are shown in Table 1.
Soil sampling and physical fractioning
Peat samples were collected when annual WL was
deepest at the end of the dry season in September
2009. Samples in the forested sites were taken from
the low peat surfaces between trees and in the open
sites from the flat vegetation-free surfaces. Sampling
depths were 10–15, 40–45, 80–85 and 110–115 cm
below the soil surface. The topmost depth (10–15 cm)
represented surface peat that is mostly above WL and
thus oxic, while deeper peat (40–45 cm) was mostly
waterlogged and thus anoxic at all sites. The long-term
median WL of the UF was relatively high (Table 1)
and thus the conditions under permanent waterlogging
and consequent anoxia were reached at the depth of
80–85, which formed the deepest sampling depth at
the site. The WL of the uncontrollably drained sites
has been deeper more frequently, and the 80–85 cm
depth provided mostly waterlogged and 110–115 cm
depth provided permanently waterlogged conditions.
Where present, recently deposited litter on the soil
surface (e.g. loose cover of leaves or branches) was
removed before sampling. The exposed soil surface
was marked as the 0 cm depth. Samples at all sites
were taken from one wall of a dug pit hole immedi-
ately after reaching each sampling depth. At the end of
sampling the pit holes were approximately
100 9100 cm wide and 150 cm deep. We cut six
samples from each depth using a sharp knife and a
volume-exact frame (10 910 cm wide and 5 cm
high). All samples (total n =90) were sealed in plastic
Table 1 Water table level characteristics at the study sites
UF DF DO AO
Period 1997–2006 2005–2007 2004–2008 2001–2003
n 2733 1231 1421 355
Average -18.3 -50.0 -26.6 -25.7
Median -10.3 -38.6 -16.8 -22.0
Maximum 33.4 1.7 20.2 -5.2
Minimum -60.8 -177.6 -160.0 -71.9
First quartile -30.7 -58.7 -32.3 -30.2
Third quartile 3.1 -27.2 -8.3 -16.6
Negative values indicate depth in centimetres below the soil surface and positive values above the soil surface. Period indicates the
observation period in years, and n indicates the number of observations based on diurnal means
Wetlands Ecol Manage
123
bags and stored in a refrigerator (4 °C) for maximum
of 1 week prior to analyses.
Living roots were removed from all samples. The
samples from each site and depth were divided into
two subsets (n =3 per subset). One sample subset was
dried at 70 °C to determine the dry bulk density
(g cm
-3
, BD). This sample type is called ‘bulk soil’.
The other sample subset was fractioned gently with
running water and two sieves (mesh [1.5 mm and
0.15 mm) to describe the physical decomposition
stage. The physically least decomposed matter cap-
tured by the larger mesh is called a ‘woody’ sample,
and the more decomposed matter, captured by the
smaller mesh, is called a ‘fibric’ sample. Both sieved
fractions were dried at 70 °C and their proportion of
the mass of the corresponding bulk soil samples was
calculated. A summary of the physical structure
characteristics is presented in Table 2, while the
detailed data is reported in Ko
¨no
¨nen et al. (2015).
The proportions of wood and fibre were calculated
from the average BD of the samples from the same site
and depth. The finer the material forming the peat and
the higher the BD, the higher the physical decompo-
sition stage was considered to be.
Chemical analyses
First all samples of the same type (bulk soil, fibric,
woody) sampled from the same depth and site were
pooled and milled to fine powder. Loss in mass on
ignition (LOI, 550 °C for 4 h) was measured. Con-
centrations of total C and N were determined (LECO
CHN-1000). All these analyses were performed using
three laboratory replicates.
We measured several organic matter quality
parameters to examine the CCC of all the fractions.
The parameters were hemicellulose (including pectin),
cellulose, extractives, ASL and AIL. All analyses were
performed using two laboratory replicates.
The carbohydrate composition of hemicellulose and
cellulose concentrations were analysed from
10 ±2 mg of dry sample. The total hemicellulose
concentration was calculated by adding together the
non-cellulosic polysaccharide concentrations formed of
neutral sugars (arabinose, rhamnose, xylose, mannose,
galactose, glucose) and uronic acids (glucuronic,
galacturonic and 4-O-Me-glucuronic acid), which were
determined by acid methanolysis followed by gas
chromatography (GC). The cellulosic glucose concen-
tration was determined by subtracting the concentration
of non-cellulosic glucose determined by acid methanol-
ysis from the total glucose concentration. The total
glucose concentration was determined by acid hydrol-
ysis and silylation followed by GC. The methods for
methanolysis and hydrolysis are described in Sundberg
et al. (1996,2003), respectively.
Concentrations of extractives and lignin-derived
compounds were determined using gravimetric
Table 2 Summary of peat physical structure based on data published in Ko
¨no
¨nen et al. (2015)
Site Depth (cm) BD (g cm
-3
) Wood (% of the sample) Fibre (% of the sample)
UF 10–15 0.13 ±0.01 22.8 ±4.0 48.3 ±8.4
40–45 0.14 ±0.01 5.8 ±1.2 46.1 ±2.8
80–85 0.15 ±0.03 5.8 ±1.7 15.8 ±2.5
DF 10–15 0.17 ±0.03 9.2 ±1.5 65.1 ±9.2
40–45 0.22 ±0.03 2.0 ±1.0 49.0 ±9.1
80–85 0.15 ±0.02 7.2 ±5.1 26.7 ±1.8
110–115 0.12 ±0.01 19.1 ±7.9 36.5 ±5.9
AO 10–15 0.18 ±0.01 4.3 ±2.2 37.8 ±1.1
40–45 0.17 ±0.01 9.9 ±1.0 37.2 ±4.3
80–85 0.13 ±0.01 6.8 ±2.4 32.0 ±50
110–115 0.13 ±0.00 17.3 ±12.0 32.1 ±9.7
DO 10–15 0.20 ±003 2.7 ±1.8 14.1 ±0.5
40–45 0.12 ±0.03 12.1 ±1.3 34.1 ±2.1
80–85 0.14 ±0.02 18.2 ±14.4 30.7 ±6.3
110–115 0.15 ±0.02 15.0 ±6.4 30.2 ±3.6
Wetlands Ecol Manage
123
methods. The concentration of extractives was deter-
mined by mixing 1.2 g of the dry sample with 20 ml of
an acetone:water solution (v:v, 9:1). The mixture was
then sonicated (45 min), filtered (through a no. 4
Pyrex) and the residual was dried (24 h, 105 °C). The
mass loss resulted from dissolving indicated the
amount of extractives in the sample. AIL concentra-
tion was determined from 0.3 g of the extractive-free
sample with a standard procedure (TAPPI Test
Method T 222-om-88, 2000). ASL concentration
was determined from the solution produced as a by-
product when determining AIL using UV spec-
troscopy (Shimadzu UV-2401 PC UV–VIS Recording
Spectrophotometer) at an absorbance of 203 nm
(Brunow et al. 1999). These methods are originally
developed for wood fibres, but have also been
successfully applied to organic soils (Merila
¨et al.
2010; Strakova
´et al. 2010). The chemical character-
ization using these methods is discussed further in
Merila
¨et al. (2010). All concentrations were
expressed as mg g
-1
sample dry mass.
Data analyses
Unconstrained principal component analysis (PCA)
using the determined C compounds as response
variables, and the environmental variables (site,
sampling depth, sample type, and physical structure)
as passive explanatory variables was first conducted to
summarize the variation in peat CCC. Next, con-
strained redundancy analysis (RDA) with variation
partitioning was conducted to test the effect of site,
depth and physical structure and to analyse how much
of the total variation (inertia) was explained by each
environmental variable. This RDA was conducted
with a reduced data set, which included only the
topmost three depths and bulk soil as the only sample
type, to reduce the influence of the lack of the fourth
depth from the undrained site, and to examine the
unique effects of the environmental variables on bulk
soil. Next, to test the sample type effect on peat CCC,
and to further test hypothesis iv, a partial RDA
analysis was conducted with the full data set using
sample type as the explanatory variable, and depth and
site as the covariates. To further test the effect of
sample type a ttest between all samples of the same
type was conducted separately for each measured
variable. Finally, RDA was conducted with interac-
tively chosen explanatory variables, to pinpoint the
best explaining single environmental factors. All
RDAs were followed by Monte Carlo permutation
tests with 999 permutations. All ordination tests were
conducted with Canoco5 for Windows and the t-tests
were conducted with RStudio version 0.98.1102. The
p-value significance level used was 0.05. Otherwise,
due to limitations arising from a relatively small data
pool for any advanced statistical analyses, the main
focus was on comparisons of the calculated means of
the measured variables from different sites and depths.
Results
General patterns in peat carbon compound
composition
In PCA, the strongest gradient in the CCC data, PC1,
separated samples with high AIL (i.e. acid insoluble
lignin) from samples with high hemicellulosic carbo-
hydrates and uronic acids, and explained 42.8 % of the
total variation in CCC. PC2 separated samples rich in
extractives and ash from samples rich in cellulose, and
explained 16.5 % of the total variation in CCC
(Fig. 2). Concentrations of hemicelluloses, uronic
acids and ASL increased towards the UF and two
topmost depths of DF, which were separated from the
deeper depths of DF and the open sites by PC1. Ash
and extractives concentrations increased towards the
80–85 cm depth of DF and two topmost depths of the
degraded site, which were separated from their deepest
depths and all depths of the agricultural site and AIL
concentration by PC2. Of the physical characteristics,
the proportion of wood fraction correlated positively
with cellulose, and BD correlated positively with
extractives and ash. The proportion of fibre fraction
correlated with hemicellulosic carbohydrates and
uronic acids. Of the sample types, the bulk soil and
fibric sample type were separated from the woody
sample type by PC1. Extractives, ASL and AIL
pointed towards, i.e., were higher in, the bulk soil and
fibric sample, and cellulose, hemicelluloses and uronic
acids pointed towards the woody sample type. RDA
with variation partitioning showed that the site
accounted for 27.3 % (p =0.056), depth 9.7 %
(p =0.124) and physical structure (BD and wood
and fibre proportion) 23.6 % (p =0.066) of all the
variation in peat CCC. The best single environmental
factors explaining variation in peat CCC were the
Wetlands Ecol Manage
123
woody sample, UF, fibre proportion and BD, which
explained 23.1, 19.1, 14.1 and 4.4 %, respectively.
The carbon compound composition of bulk soil
in relation to site, depth and physical structure
The average concentrations of C, N, extractives, total
hemicellulose, and ASL were generally higher (2.2,
50.7, 37.8, 646.6 and 37.8 %, respectively) in the bulk
soil of forest sites while concentrations of, ash,
cellulose, and AIL were lower (25.6, 3.6 and
11.3 %, respectively) than at the open sites (Figs. 3,
4; Table 3). The CCC of both undrained and drained
forests was similar in the surface peat (10–15 cm
depth), but at the depth of 40–45 cm and deeper the
peat properties in drained forest and open sites were
increasingly similar, reflecting a higher physical
decomposition stage than in the UF (Tables 2,3;
Fig. 3).
All sites showed a decreasing trend in N concentration
and increasing trends in C concentration and CN-ratio
with increasing peat depth. The hemicellulosic carbohy-
drate concentration of bulk soil was higher in the forests
than in the open sites, but always contained compounds
in the following order: glucose [xylane [man-
nose [galactan [galacturonic acid [arabinose [
rhamnose [glucuronic acid [4-O-Me-glucuronic acid
(Fig. 3;Table4). The total hemicellulose concentration
in the bulk soil of the forest sites usually showed a clear
decreasing trend with depth. At the open sites the total
hemicellulose concentrations in bulk soil was on average
87 % smaller than at the forest sites, and the differences
in concentrations between sites decreased with depth.
The concentration of extractives in bulk soil had a
decreasing trend with depth at the degraded site, and an
increasing trend in the forests and agricultural site
(Fig. 3;Table3). Furthermore, the concentration of
extractives was highest in the drained forest at the depth
of 40–45 cm. At all sites, the cellulose concentration and
ASL decreased and AIL increased with depth.
Generally, fibres comprised 14.1–65.1 % and
woody matter 2.0–22.8 % of the dry mass of bulk soil
(Table 2). Of the physical structure, BD correlated
negatively with the concentrations of cellulose, hemi-
celluloses and ASL, while proportions of wood and
fibres correlated positively with them (Fig. 5). BD
correlated positively with the concentrations of AIL,
ash and extractives.
The carbon compound composition in relation
to sample type
The CCC of fibric samples resembled bulk soil and
responded to depth and site in the same way, but in all
measured variables the woody sample was signifi-
cantly(p \0.05) different (Figs. 3,6). This was in line
with the physical decomposition stage, which was
generally higher in the bulk soils and the fibric sample
type than in the woody sample type at all sites and
depths (Table 2). On average the bulk soil and fibric
samples contained 33.8 % more extractives, 12.7 %
more AIL and 68.4 % more ASL than the woody
samples (Table 5). The woody samples contained on
average over 4.5 times more cellulose and nearly 3
times more total hemicellulose than the other sample
types. The concentrations of hemicellulosic
Fig. 2 Results of the unconstrained PCA. The first principal
component, PC1 (x-axis) explained 42.8 %, and PC2 (y-axis)
explained 28.4 % of the total CCC variation. The response
variables indicated by black arrows were: ASL acid-soluble
lignin, hemicelluloses, uronic acids, extractives, cellulose and
AIL. Site, sampling depth, sample type, and physical structure
(BD and proportion of wood and fibres) were used as passive
explanatory variables. Site: UF undrained forest, DF drained
forest, DO degraded open site, AO agricultural open site,
number following dash indicates sampling depth (e.g.
UF*10 =undrained forest at a depth of 10 cm from the soil
surface). Sample type: bulk soil, fibric sample (between 0.15 and
1.5 mm; fibric) and woody sample ([1.5 mm; woody). Physical
structure: dry bulk density (g cm
-3
)BD,wood % volumetric
proportion of wood, fibre % volumetric proportion of fibres
Wetlands Ecol Manage
123
Wetlands Ecol Manage
123
carbohydrates, especially the concentrations of man-
nose, glucose, xylane, arabinose and galacturonic acid
were significantly higher in the woody samples
throughout the whole sampling profile than in the
fibric samples or bulk soil (Fig. 6). On average, C
concentration was 57 % and N less than 1 % in all
sample types. N concentration was the lowest in the
woody samples, resulting in an average 30 % higher
CN-ratio than other sample types. Generally the CN-
ratio increased with depth at all sample types, but the
increase was remarkable in woody samples, in which
the CN-ratio was as high as 111–167 at the depth of
110 cm (Fig. 4).
Discussion and conclusions
This study provides strong evidence that long-term
land use involving drainage (with or without defor-
estation) leads to the enrichment of recalcitrant C
compounds in peat. This observation is in line with
our hypothesis (iii). Although peat in the UF was
also rich in recalcitrant AIL, it is notable that the
concentrations of labile hemicelluloses and uronic
acids declined especially when both vegetation and
WL had been altered. Our results thus support the
hypothesis (i) that land-use intensity greatly impacts
peat CCC. The differences between land-use types
were greatest at the surface peat down to a depth of
40–45 cm, which supports our hypothesis (ii). The
concentrations of total hemicellulose (including
pectin) and cellulose were higher in the physically
less decomposed bulk soil of surface peat
(10–15 cm) of the forest sites than in the open sites
and in the least physically decomposed woody
sample type at all sites. AIL concentrations were
highest in the physically more decomposed bulk soil
of the surface peat in intensively altered open sites
and in the fibric samples compared to the woody
samples. This supports our fourth hypothesis (iv).
Both vegetation, providing litter supply with a
higher concentration of labile carbohydrates and
decomposition, of the litter and of peat, contribute to
peat CCC structure. Originally the sites were woody
peat-accumulating forest-covered swamps, but peat
formation and decomposition capacity at the
reclaimed sites has been altered in the past. The AIL
content was generally very high in the bulk soil (on
average 77 % of the dry mass). AIL in this study is
comparable to concentrations of lignin measured by
Andriesse (1988), who reported relatively high lignin
concentrations (64–76 %) for tropical peat in com-
parison to boreal Sphagnum peat (18 %) and woody
bFig. 3 CCC of the different sample types in tropical peat. The
right-hand paragraph cumulatively presents the concentrations
of AIL, ASL, extractives, hemicelluloses and uronic acids and
cellulose. Cellulose concentration of wood sample from a depth
of 110 cm is lacking from DF. The left-hand paragraph
cumulatively presents the hemicellulosic carbohydrates (Ara
arabinose, Glc glucose, Gal galactan, GalA galaturonic acid,
GlcA glucuronic acid, Man =mannose, Rha rhamnose, Xyl
xylane, 4-O-me =4-O-methyl glucuronic acid). Ash concen-
tration varied between 0.22 and 1.1 mg g
-1
. Laboratory
analyses were performed with two replicates (n =2). Site: UF
undrained forest, DF drained forest, DO degraded open site, AO
agricultural open site
Fig. 4 The C and N concentrations and CN-ratio for all sample
types, depths and sites. At each depth the observations of woody
samples are presented on the top, the non-fractioned samples in
the middle and the fibric sample at the bottom. Site: UF
undrained forest, DF drained forest, DO degraded open site, AO
agricultural open site
Wetlands Ecol Manage
123
sedge peat (38 %). The relatively high AIL content in
tropical peat is mostly due to the woody origin of the
peat matter (Yule and Gomez 2009; Mehta et al.
2013). In this study, it is shown that the peat CCC
correlates also with the physical decomposition stage,
expressed as BD and proportions of different sized
particles in the bulk soil. Yet, the CCC of bulk soil
cannot be estimated directly from the proportions of
different sized fractions, because the CCC of the
fractions differed between land-uses. However, in all
land-use types the recalcitrance of bulk soil and fibric
samples was higher than in the woody sample type.
This is because during the decomposition process the
physical structure of the peat becomes finer, and
microbes deplete the most labile energy sources
relatively faster from the peat substrate leading to
the enrichment of more resistant compounds such as
AIL (Berg 2000; Hooijer et al. 2012;Ko
¨no
¨nen et al.
2015). Subsequently, peat CCC determines the
decomposability of the residual organic matter, and
thus the amount of GHG emissions released from peat.
Reduced peat decomposability may have contributed
to the lower CO
2
emissions released in oxidation at
degraded peatlands as compared to forests (Jauhiainen
et al. 2008; Hirano et al. 2014; IPCC 2014).
The observed differences in peat CCC (higher AIL
and lower hemicellulose concentrations in the
reclaimed sites) were greatest between forests and
open sites, but drainage also independently affected
peat CCC. In both forest-covered peatlands the litter
decomposition rate depends on both the quality and
position of the litter in the peat profile (Hoyos-
Santillan et al. 2015). Root litter is slower to decom-
pose in comparison to leaves and stems, and their
location deeper in peat profile (often in waterlogged
conditions) makes them a major contributor to peat
accumulation (Page et al. 1999;Wu
¨st et al. 2008;
Hoyos-Santillan et al. 2015). Forest vegetation is also
an important component in the nutrient cycle, because
nutrients in ombrotrophic peatlands mainly bind to
plant biomass and are released during the early phase
of decomposition (Aerts et al. 1999; Rieley and Page
2005). In the topmost peat of the forests the conditions
for decomposition are thus more favourable due to the
higher availability of nutrients and easily decompos-
able substrates, implied in our study as high total
hemicellulose and N concentrations, a CN-ratio close
to 30 and mostly oxic conditions. The easily decom-
posable organic C compounds therefore become
depleted mostly in the topmost peat leading to the
Table 3 The ash content and C compound composition of non-fractioned samples from every site and depth
Site Depth
(cm)
Ash
a
(mg g
-1
)
Extractive
(mg g
-1
)
Total hemicelluloses
(mg g
-1
)
Cellulose
(mg g
-1
)
ASL
(mg g
-1
)
AIL
(mg g
-1
)
UF 10–15 0.60 ±0.04 131.3 ±4.0 56.7 ±5.8 39.3 ±11.8 17.7 ±0.9 683.6 ±14.7
40–45 0.36 ±0.03 135.0 ±3.6 37.2 ±3.9 16.8 ±1.1 19.1 ±4.0 708.6 ±10.8
80–85 0.63 ±0.03 290.9 ±55.9 5.4 ±1.0 14.8 ±0.9 4.8 ±5.2 717.5 ±35.1
DF 10–15 0.61 ±0.08 111.6 ±2.3 59.5 ±7.3 25.7 ±1.8 20.0 ±0.0 674.1 ±20.3
40–45 0.39 ±0.02 380.4 ±11.8 9.3 ±0.3 5.7 ±1.4 4.6 ±1.8 671.0 ±18.2
80–85 0.43 ±0.04 220.9 ±9.0 1.0 ±0.0 12.0 ±0.2 5.1 ±3.1 751.0 ±29.4
110–115 0.32 ±0.06 130.8 ±2.1 2.7 ±0.6 28.8 ±11.3 1.1 ±0.5 811.0 ±14.8
DO 10–15 1.10 ±0.05 207.8 ±17.0 1.0 ±0.6 6.7 ±0.5 12.9 ±6.0 764.5 ±29.2
40–45 0.75 ±0.04 167.2 ±17.5 1.9 ±0.8 17.4 ±0.9 14.3 ±0.5 774.1 ±3.9
80–85 0.62 ±0.36 132.1 ±3.0 1.0 ±0.1 28.7 ±12.7 6.2 ±3.3 808.9 ±1.2
110–115 0.60 ±0.02 129.9 ±12.8 1.2 ±0.7 22.9 ±2.2 5.0 ±1.7 816.2 ±4.5
AO 10–15 0.88 ±0.01 118.5 ±3.5 2.6 ±0.0 19.7 ±0.3 2.9 ±0.0 863.1 ±10.2
40–45 0.73 ±0.03 103.5 ±5.0 8.4 ±9.2 18.7 ±0.9 1.8 ±0.2 820.4 ±5.6
80–85 0.22 ±0.01 173.7 ±8.8 5.0 ±0.5 24.3 ±0.7 3.8 ±0.1 815.8 ±1.0
110–115 0.23 ±0.01 129.6 ±7.6 5.2 ±1.6 31.2 ±1.4 6.1 ±0.8 797.6 ±6.0
Values are mean ±SD of two laboratory replicates
a
Proportion of ash published previously in Ko
¨no
¨nen et al. (2015)
Wetlands Ecol Manage
123
Table 4 The hemicellulosic carbohydrate composition of non-fractioned samples from every site and depth
Site Depth
(cm)
Mannose
(mg g
-1
)
Glucose
(mg g
-1
)
Galactan
(mg g
-1
)
Xylane
(mg g
-1
)
Arabinose
(mg g
-1
)
Rhamnose
(mg g
-1
)
Glucuronic acid
(mg g
-1
)
Galacturonic acid
(mg g
-1
)
4-O-methyl glucuronic
acid (mg g
-1
)
UF 10–15 6.0 ±0.4 21.4 ±2.5 6.5 ±0.4 9.7 ±0.6 2.6 ±0.2 3.2 ±0.2 1.2 ±0.6 5.3 ±0.6 0.7 ±0.2
40–45 5.3 ±0.5 19.8 ±2.6 3.7 ±0.4 3.4 ±0.3 1.7 ±0.4 1.3 ±0.1 1.0 ±0.4 1.1 ±0.0 0.0 ±0.0
80–85 0.9 ±0.1 3.0 ±0.5 0.1 ±0.2 1.3 ±0.3 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
DF 10–15 8.2 ±0.4 3.4 ±2.5 6.9 ±0.9 6.1 ±1.2 1.7 ±1.3 1.8 ±0.1 0.0 ±0.0 1.5 ±0.9 0.0 ±0.0
40–45 2.1 ±0.0 6.2 ±0.1 0.7 ±0.0 0.3 ±0.5 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
80–85 0.0 ±0.0 1.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
110–115 0.0 ±0.0 2.7 ±0.6 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
DO 10–15 0.3 ±0.5 0.7 ±0.2 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
40–45 0.3 ±0.5 1.5 ±0.3 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
80–85 0.0 ±0.0 1.0 ±0.1 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
110–115 0.0 ±0.0 1.2 ±0.7 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
AO 10–15 0.0 ±0.0 1.6 ±0.0 0.0 ±0.0 0.0 ±0.0 1.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
40–45 0.5 ±0.8 1.6 ±0.6 0.0 ±0.0 5.8 ±8.2 0.0 ±0.0 0.1 ±0.1 0.0 ±0.0 0.5 ±0.7 0.0 ±0.0
80–85 0.9 ±0.1 4.1 ±0.4 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0±0.0 0.0 ±0.0 0.0 ±0.0
110–115 0.6 ±0.0 3.3 ±0.3 0.0 ±0.0 0.0 ±0.0 1.3 ±1.9 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0 0.0 ±0.0
Values are mean ±SD of two laboratory replicates
Wetlands Ecol Manage
123
enrichment of more recalcitrant compounds deeper in
the peat over time, where decomposition is more
restricted also due to more frequent waterlogging and
anaerobic conditions. In drained forest the surface peat
CCC is maintained by continuous litter deposition in a
similar manner to the UF, but in the drained forest the
effects of enhanced drainage were seen especially at
the depth of 40–45 cm, where the peat CCC began to
resemble corresponding depths at the open sites. At the
40–45 cm depth, with declining influence of litter
input, the peat properties in drained forest are greatly
impacted by improved conditions for aerobic decom-
position and compaction of peat after WL drawdown
(Laiho 2006; Hooijer et al. 2012; IPCC 2014). It has
also been suggested that vertical and horizontal water
level movements can relocate finer particles (Lampela
et al. 2014), which could have contributed to the
enrichment of finer particles and the high concentra-
tions of extractives detected near the WL median (at
approximately -40 cm depth) in the drained forest.
At open sites, deforestation, drainage and fires have
modified peat CCC properties, and resulted in a low N
content and a high CN-ratio and AIL content seen
especially in the topmost peat. Deforestation had
permanently reduced the vegetation biomass and litter
deposition rate (Sulistiyanto 2004; Mehta et al. 2013;
Blackham et al. 2014), and due to this, recent litter
deposition now barely contributes to the formation of
new peat, and decomposition mainly happens in the
peat formed prior to deforestation and drainage. The
peat at the open sites has also been repeatedly fire-
affected, and especially at the degraded site peat
combustion has exposed older peat from a depth of
several decimetres (Hoscilo et al. 2011; Konecny et al.
2016). The combustion process in smouldering fires
also modifies organic matter into a more recalcitrant
form i.e. ‘black carbon’ (Knicker 2007 and references
within). The consequent decomposition together with
the formation of black carbon has likely contributed to
the AIL concentrations in the surface peat in the open
sites, where it was found to be 35 % higher than in
non-burnt peat at the forest sites. The slightly higher
concentrations of recalcitrant compounds at the agri-
cultural site than at the degraded site may be due to
increased microbial decomposition, which is an out-
come of both longer lasted drainage and active soil
management including fertilizer applications and
tilling, which aerates the topmost peat.
Fig. 5 Results of partial redundancy analysis (RDA) testing the
effect of peat physical structure on CCC. The concentration of
extractives, ash and AIL increased with dry BD, while the
concentrations of cellulose and hemicelluloses and uronic acids
were higher in bulk soils with high proportions of wood and
fibres. Ara arabinose, Extractives acetone: water extractives,
Cellulos cellulose, Glc glucose (hemicellulosic), Gal galactan,
GalA galaturonic acid, GlcA glucuronic acid, AIL acid-insoluble
lignin, Man mannose, Rha rhamnose, ASL acid soluble lignin,
Xyl xylane, 4-O-me 4-O-methyl glucuronic acid Fig. 6 Results of the partial RDA testing of the effect of sample
type to CCC. Variation of the measured chemical properties is
summarized by classifying it according to sample type. The
woody sample type differs from the bulk soil and fibric sample
type. Site: UF undrained forest, DF drained forest, DO degraded
open site, AO agricultural open site
Wetlands Ecol Manage
123
Peat CCC determines the decomposability of peat
and thus will influence to the rate of release of gaseous
C emissions. Several studies have highlighted the
importance of substrate quality in controlling decom-
position processes (Berg 2000; Bragazza et al. 2007;
Merila
¨et al. 2010; Strakova
´et al. 2011). Thus, the
results of our study partly explain why the manage-
ment intensity and time from land-use change have
been observed to reflect the amount of GHG emissions
released from the decomposition. Lower gaseous C
emissions have been reported in the most intensively
altered land-use types, which have been both drained,
deforested and burned, and where litter feed from the
vegetation is low, than from land-use types with
substantial vegetation (i.e. oil palm, acacia) (Couwen-
berg et al. 2010; Carlson et al. 2013; Hirano et al.
2014; IPCC 2014). However, regardless of land-use
intensity, the average C losses are high a few years
after land reclamation and then decrease with time
(Hooijer et al. 2012 and references within). In
reclaimed lands this reduction in C loss with time is
likely due to the progressive enrichment of recalcitrant
compounds in the surface peat under conditions where
litter supply is limited, which together with the altered
environmental conditions (i.e. drought, nutrient lim-
itations, high temperature), have led to reduced rates
of decomposition and gaseous C emissions from the
most intensively managed sites. In undisturbed peat
swamp forest, where litter input provides large
continuous supply of labile carbohydrates and nor-
mally no droughts occur, CO
2
emissions can be higher
than at intensively altered sites (Sundari et al. 2012;
Hirano et al. 2014). However, despite the higher level
of CO
2
emissions, the undisturbed forest supports litter
production and thus preserves existing peat deposits
and indicates the potential for the accumulation of new
peat.
Tropical soils are naturally poor in N and P, and our
study showed that with intensive management they
tend to become rich in recalcitrant compounds, which
limit decomposition and plant growth. In drained and
deforested peatlands, peat enriched with recalcitrant
substrates will continue to decompose especially if the
conditions for microbial activities are improved (e.g.
quality and amount of substrate, and nutrient moisture
and oxic conditions) (Dungait et al. 2012). Productive
cultivation in these nutrient limited soils requires
frequent fertilization (Andriesse 1988; Rieley and
Page 2005; Murdiyarso et al. 2010), which increases
labile N and P availability in peat. This may increase
the microbial activity and thus decomposition of
recalcitrant substrates (Dungait et al. 2012; Jauhiainen
et al. 2014; Comeau et al. 2016).
Drainage and deforestation largely alter peat decom-
posability and, despite the increased recalcitrance, in
intensively managed drained and deforested sites lead to
continuous net C loss from peat deposits. For the
existing large degraded areas the best option to reach
original C dynamics and in the end,hopefully, also peat
accumulation could be rewetting and reforestation.
Some drained areas support the economy by remaining
in production, subject to fertilizing and tilling, which
maintains continued C losses, but this is a trade-off
between productivity and carbon storage. Therefore, the
best sustainable and carbon–neutral management con-
dition for these fascinating ecosystems is to keep them
as undisturbed peat swamp forests.
Acknowledgments We thank Dr. Hidenori Takahashi and Dr.
Takashi Inoue for providing the data used in calculating average
groundwater depths of the study sites prior to 2004. We
acknowledge the Centre for International Cooperation in
Sustainable Management of Tropical Peatland (CIMTROP)
for providing facilities and assistance during field sampling. We
thank Satu Repo from the Natural Resources Institute Finland
Table 5 The average C compound composition of different sample types
Sample
type
Extractives
(mg g
-1
dry
mass)
Total
hemicelluloses and
Uronic acids (mg
g
-1
dry mass)
Cellulose
(mg g
-1
dry
mass)
Acid-soluble
lignin (mg
g
-1
dry mass)
Acid-
insoluble
lignin (mg
g
-1
dry mass)
C (%) N (%) CN (%)
Bulk soil 171.0 ±12.9
a
13.2 ±3.0
a
20.8 ±4.5
a
8.3 ±1.8
a
765.2 ±10.9
a
57.6 ±3.2
a
0.97 ±0.3
a
64.5 ±17.0
a
Fibres 162.0 ±25.9
a
14.7 ±5.4
a
24.5 ±2.3
a
11.0 ±2.7
a
758.7 ±7.6
a
58.0 ±2.3
a
0.95 ±0.4
a
69.5 ±21.6
a
Woody 124.0 ±8.0
b
53.1 ±3.2
b
128.6 ±5.1
b
5.7 ±1.0
b
676.3 ±10.2
b
55.9 ±1.4
b
0.70 ±0.3
b
95.9 ±39.0
b
Woody sample: particle size [1.5 mm; 0.15 mm\fibric sample \1.5 mm. Upper indexes denote statistical differences (p\0.05)
between the sample types. Values are mean ±SD of 30 observations for each sample type
Wetlands Ecol Manage
123
(Luke) and Marjut Wallner from the University of Helsinki for
their patience and expertise in the laboratory analyses. This
research was supported by the Academy of Finland -funded
‘Restoration Impact on Tropical Peat Carbon and Nitrogen
Dynamics’ -project (RETROPEAT), University of Helsinki
prize money for Peatlanders and the Jenny and Antti Wihuri
foundation.
References
Aerts R, Verhoeven J, Whigham D (1999) Plant-mediated
controls on nutrient cycling in temperate fens and bogs.
Ecology 80(7):2170–2181
Andriesse JP (1988) Nature and management of tropical peat
soils. FAO SOIL BULL 59. Food and agriculture organi-
zation of the United Nations, Rome. pp 179
Berg B (2000) Litter decomposition and organic matter turnover
in northern forest soils. Ecol Manag 133:13–22
Blackham G, Webb E, Corlett R (2014) Natural regeneration in
a degraded tropical peatland, Central Kalimantan,
Indonesia: implications for forest restoration. Ecol Manag
324:8–15
Bragazza L, Siffi C, Iacumin P, Gerdol R (2007) Mass loss and
nutrient release during litter decay in peatland: the role of
microbial adaptability to litter chemistry. Soil Biol Bio-
chem 39:257–267
Brunow G, Lundquist K, Gellerstedt G (1999) Lignin. In:
Sjo
¨stro
¨m E, Ale
´n R (eds) Analytical methods in wood
chemistry, pulping, and papermaking. Springer, Berlin,
pp 77–125
Carlson K, Curran L, Asner G, Pittman A, Trigg S, Marion
Adeney J (2013) Carbon emissions from forest conversion
by Kalimantan oil palm plantations. Nat Clim Chang
3:283–287
Clymo R (1984) The limits to peat bog growth. Philos Trans R
Soc Lond B 303:605–654
Comeau L-P, Hergoualc’h K, Hartill J, Smith J, Verchot L, Peak
D, Salim A (2016) How do the heterotrophic and the total
soil respiration of an oil palm plantation on peat respond to
nitrogen fertilizer application? Geoderma 268:41–51
Couwenberg J, Dommain R, Joosten H (2010) Greenhouse gas
fluxes from tropical peatlands in south-east Asia. Glob
Chang Biol 16:1715–1732
Dungait J, Hopkins D, Gregory A, Whitmore A (2012) Soil
organic matter turnover is governed by accessibility not
recalcitrance. Glob Change Biol 18:1781–1796
Freeman C, Ostle N, Kang H (2001) An enzymic ‘latch’ on a
global carbon store. Nature 409:149
Hirano T, Kusin K, Limin S, Osaki M (2014) Carbon dioxide
emissions through oxidative peat decomposition on a burnt
tropical peatland. Glob Chang Biol 20:555–565
Hooijer A, Page S, Jauhiainen J, Lee W, Lu X, Idris A, Anshari
G (2012) Subsidence and carbon loss in drained tropical
peatlands. Biogeosciences 9:1053–1071
Hoscilo A, Page S, Tansey K, Rieley J (2011) Effect of repeated
fires on land-cover change on peatland in southern Central
Kalimantan, Indonesia, from 1973 to 2005. Int J Wildland
Fire 20:578–588
Hoyos-Santillan J, Lomax B, Large D, Turner B, Boom A,
Lopez O (2015) Getting to the root of the problem: litter
decomposition and peat formation in lowland Neotropical
peatlands. Biogeochemistry 126:115–129
IPCC (2014) 2013 Supplement to the 2006 IPCC guidelines for
national greenhouse gas inventories: wetlands. In: Hiraishi
T, Krug T, Tanabe K, Srivastava N, Baasansuren J, Fukuda
M, Troxler TG (eds). IPCC, Geneva, p 353
Jauhiainen J, Takahashi H, Heikkinen J, Martikainen P,
Vasander H (2005) Carbon fluxes from a tropical peat
swamp forest floor. Glob Chang Biol 11:1788–1797
Jauhiainen J, Limin S, Silvennoinen H, Vasander H (2008)
Carbon dioxide and methane fluxes in drained tropical peat
before and after hydrological restoration. Ecology
89(12):3503–3514
Jauhiainen J, Kerojoki O, Silvennoinen H, Limin S, Vasander H
(2014) Heterotrophic respiration in drained tropical peat is
greatly affected by temperature—a passive ecosystem
cooling experiment. Environ Res Lett 9:18
Knicker H (2007) How does fire affect the nature and stability of
soil organic nitrogen and carbon? A review. Biogeo-
chemistry 85:91–118
Konecny K, Ballhorn U, Navratil P, Jubanski J, Page SE, Tansey
KJ, Hooijer A, Vernimmen R, Siegert F (2016) Variable
carbon losses for recurrent fires in drained tropical peat-
lands. Glob Change Biol 22(4):1469–1480
Ko
¨no
¨nen M, Jauhiainen J, Laiho R, Kusin K, Vasander H (2015)
Physical and chemical properties of tropical peat under
stabilised land uses. Mires Peat 16(8):1–13
Laiho R (2006) Decomposition in peatlands: reconciling
seemingly contrasting results on the impacts of lowered
water levels. Soil Biol Biochem 38:2011–2024
Lampela M, Jauhiainen J, Vasander H (2014) Surface peat
structure and chemistry in a tropical peat swamp forest.
Plant Soil 382:329–347
Mehta N, Dinakaran J, Patel S, Laskar A, Yadava M, Ramesh Y,
Krishnayya N (2013) Changes in litter decomposition and
soil organic carbon in a reforested tropical deciduous cover
(India). Ecol Res 28:239–248
Merila
¨P, Malmivaara-La
¨msa
¨M, Spetz P, Stark S, Vierikko K
(2010) Soil organic matter quality as a link between
microbial community structure and vegetation composi-
tion along a successional gradient in a boreal forest. Appl
Soil Ecol 46:259–267
Miettinen J, Shi C, Liew S (2016) Land cover distribution in the
peatlands of Peninsular Malaysia, Sumatra and Borneo in
2015 with changes since 1990. Global Ecol Conserv
6:67–78
Murdiyarso D, Hergoualc’h K, Verchot V (2010) Opportunities
for reducing greenhouse gas emissions in tropical peat-
lands. Proc Natl Acad Sci USA 107(46):19655–19660
Page S, Rieley JO, Shotyk Ø, Weiss D (1999) Interdependence
of peat and vegetation in a tropical peat swamp forest.
Philos Trans R Soc B 354(1391):1885–1897
Page SE, Rieley JO, Banks C (2011) Global and regional
importance of the tropical peatland carbon pool. Glob
Chang Biol 17:798–818
Rejmankova E (2001) Effect of experimental phosphorus
enrichment on oligotrophic tropical marshes in Belize,
Central America. Plant Soil 236:33–53
Wetlands Ecol Manage
123
Rieley J, Page S (eds) 2005. Wise use of tropical peatlands:
focus on South East Asia. ALTERRA - Wageningen
University and Research Centre and the EU INCO -
STRAPEAT and RESTORPEAT. 237 p. http://www.
restorpeat.alterra.wur.nl/download/WUG.pdf. Accessed
13 June 2015
Sanaullah M, Chabbi A, Lemaire G, Charrier X, Rumpel C
(2010) How does plant leaf senescence of grassland species
influence decomposition kinetics and litter compounds
dynamics? Nutr Cycl Agroecosys 88:159–171
Sjo
¨gersten S, Cheesman A, Lopez O, Turner B (2011) Bio-
geochemical processes along a nutrient gradient in a trop-
ical ombrotrophic peatland. Biogeochemistry
104:147–163
Strakova
´P, Anttila J, Spetz P, Kitunen V, Tapanila T, Laiho R
(2010) Litter quality and its response to water level draw-
down in boreal peatlands at plant species and community
level. Plant Soil 335:501–520
Strakova
´P, Niemi R, Freeman C, Peltoniemi K, Toberman H,
Heiskanen I, Fritze H, Laiho R (2011) Litter type affects
the activity of aerobic decomposers in a boreal peatland
more than site nutrient and water table regimes. Biogeo-
sciences 8:2741–2755
Sulistiyanto Y (2004) Nutrient dynamics in different sub-types
of peat swamp forest in central Kalimantan, Indonesia.
PhD thesis, University of Nottingham, UK. p 388
Sundari S, Hirano T, Yamada H, Kusin K, Limin S (2012) Effect
of groundwater level on soil respiration in tropical peat
swamp forests. J Agr Meteor 68:121–134
Sundberg A, Sundberg K, Lillandt C, Holmbom B (1996)
Determination of hemicelluloses and pectins in wood and
pulp fibres by acid methanolysis and gas chromatography.
Nord Pulp Pap Res J 11:216–219
Sundberg A, Pranovich AV, Holmbom B (2003) Chemical
characterization of various types of mechanical pulp fines.
J Pulp Pap Sci 29:173–178
Talbot J, Treseder K (2012) Interactions among lignin, cellu-
lose, and nitrogen drive litter chemistry–decay relation-
ships. Ecology 93(2):345–354
Wu
¨st R, Jacobsen G, van der Gaast H, Smith A (2008) Com-
parison of radiocarbon ages from different organic frac-
tions in tropical peat cores: insights from Kalimantan,
Indonesia. Radiocarbon 50(3):359–372
Yule C, Gomez L (2009) Leaf litter decomposition in a tropical
peat swamp forest in Peninsular Malaysia. Wetl Ecol
manag 17:231–241
Yule C, Lim YY, Lim TY (2016) Degradation of Malaysian
peatlands decreases levels of phenolics in soil and in leaves
of Macaranga pruinosa. Front Earth Sci 4(45):1–9
Zhang D, Hui D, Luo Y, Zhou G (2008) Rates of litter decom-
position in terrestrial ecosystems: global patterns and
controlling factors. J Plant Ecol 1(2):85–93
Wetlands Ecol Manage
123
... Tropical peat is generally formed from the slowly accumulated woody organic materials rich in lignin and cellulose under shallow groundwater table (Sabiham and Furukawa, 1986;Yule and Gomez, 2009;Könönen et al., 2016). In several reports, Sumatra and Kalimantan peatlands had developed between 24,000 and 45,000 years BC, respectively, and are dominated by 50-300 cm depth (Page et al., 2004;Hapsari et al., 2017;Ruwaimana et al., 2020;Anda et al., 2021). ...
... Developing drainage canals in OPP lowers the groundwater level/GWL, deepening the influence of the oxic peat layer. Nevertheless, an accumulation of recalcitrant organic material also potentially occurs in OPP's peat surface, thereby restricting microbial developments (Könönen et al. 2016) and lowering peat decomposition. From a spatial perspective, microbial communities' respiration should also be significantly affected by changes in the peat's physicochemical properties due to OPP management zonation. ...
... From a spatial perspective, microbial communities' respiration should also be significantly affected by changes in the peat's physicochemical properties due to OPP management zonation. Differences in microclimate, nutrient input, and litter heterogeneity over a long-term period might result in different carbon composition and microhabitats, especially around the peat surface (Könönen et al., 2016;Manning et al., 2019;Pulunggono et al., 2022a;2022b). This brought some significant deviations regarding microbial responses to the changes in soil properties on drained tropical peat compared to their general patterns (e.g., Espenberg et al., 2018), linked to the shift of microbial structures (Dhandapani et al., 2020). ...
Article
Full-text available
Article history: Even though their role in mediating tropical peat decomposition and GHG emissions had been widely recognized, information concerning lignocellulolytic microbes, their degrading enzyme ability, and interconnection with soil physicochemical properties and peat heterotrophic respiration on mature oil palm plantation/OPP block level were rudimentary. This study evaluated the effect of sampling depth (0-30, 30-60, and 60-90 cm), OPP management zone (fertilization circle/FTC, frond stack/FRS, and harvesting path/HVP), and peat physicochemical properties on the lignocellulolytic bacteria and fungi, their degrading enzymes activities and peat heterotrophic respiration/Rh using principal component analysis/PCA, multiple linear regression/MLR, and generalized linear mixed effect models/GLMM. This study found that the soil microbiological and physicochemical properties varied widely. Dominant lignocellulolytic bacterial population and their cellulase enzyme activity were higher than fungi, regardless of sampling depth and management zone. PCA and GLMM analyses showed the significant importance of sampling depth and management zone in governing lignocellulolytic microbial population, their enzyme activities, and Rh. Microbial population and cellulase activity were also remarkably affected by the interaction of all studied factors. Peat chemical properties (pH and total Mn) controlled the natural variance of lignocellulolytic microbes and their enzymes, whereas total K regulate Rh. This study suggested that the research on microbiological-related GHG mitigation in OPP should be focused on managing the fungal population and cellulase enzyme activity at the peat surface (0-30 cm) and fertilization circle.
... The model predicted that the initial large decomposition rates were not sustained, and this decrease was highly positively correlated to reduced fraction of the litter pool, the most labile pool contributing to SOM ( Figure S4b). This result is in agreement with observations from ex situ experiments indicating that land-use change renders tropical peat more recalcitrant to decomposition (Jauhiainen et al., 2016;Könönen et al., 2016), implying that soil heterotrophic respiration from a drained tropical peatland will decline over time (Swails et al., 2018). Mechanistically, the reduction in C decomposition rate can be explained by preferential consumption by microbes of labile C compounds, leading to increased ratio of recalcitrant to labile C compounds in degraded peat soils (Könönen et al., 2016;Swails et al., 2018;Wright et al., 2011). ...
... The model predicted that the initial large decomposition rates were not sustained, and this decrease was highly positively correlated to reduced fraction of the litter pool, the most labile pool contributing to SOM ( Figure S4b). This result is in agreement with observations from ex situ experiments indicating that land-use change renders tropical peat more recalcitrant to decomposition (Jauhiainen et al., 2016;Könönen et al., 2016), implying that soil heterotrophic respiration from a drained tropical peatland will decline over time (Swails et al., 2018). Mechanistically, the reduction in C decomposition rate can be explained by preferential consumption by microbes of labile C compounds, leading to increased ratio of recalcitrant to labile C compounds in degraded peat soils (Könönen et al., 2016;Swails et al., 2018;Wright et al., 2011). ...
... This result is in agreement with observations from ex situ experiments indicating that land-use change renders tropical peat more recalcitrant to decomposition (Jauhiainen et al., 2016;Könönen et al., 2016), implying that soil heterotrophic respiration from a drained tropical peatland will decline over time (Swails et al., 2018). Mechanistically, the reduction in C decomposition rate can be explained by preferential consumption by microbes of labile C compounds, leading to increased ratio of recalcitrant to labile C compounds in degraded peat soils (Könönen et al., 2016;Swails et al., 2018;Wright et al., 2011). While chrono-sequential in situ experiments on soil heterotrophic respiration remain limited to young oil palm plantations , Cooper et al. ...
Article
Full-text available
Oil palm plantations on peat and associated drainage generate sizeable GHG emissions. Current IPCC default emission factors (EF) for oil palm on organic soil are based on a very limited number of observations from young plantations, thereby resulting in large uncertainties in emissions estimates. To explore the potential of process-based modeling to refine oil palm peat CO2 and N2O EFs, we simulated peat GHG emissions and biogeophysical variables over 30 years in plantations of Central Kalimantan, Indonesia. The DNDC model simulated well the magnitude of C inputs (litterfall and root mortality) and dynamics of annual heterotrophic respiration and peat decomposition N2O fluxes. The modeled peat onsite CO2-C EF was lower than the IPCC default (11 Mg C ha⁻¹ yr⁻¹) and decreased from 7.7 ± 0.4 Mg C ha⁻¹ yr⁻¹ in the first decade to 3.0 ± 0.2 and 1.8 ± 0.3 Mg C ha⁻¹ yr⁻¹ in the second and third decades of the rotation. The modeled N2O-N EF from peat decomposition was higher than the IPCC default (1.2 kg N ha⁻¹ yr⁻¹) and increased from 3.5 ± 0.3 kg N ha⁻¹ yr⁻¹ in the first decade to 4.7–4.6 ± 0.5 kg N ha⁻¹ yr⁻¹ in the following ones. Modeled fertilizer-induced N2O emissions were minimal and much less than 1.6% of N inputs recommended by the IPCC in wet climates regardless of soil type. Temporal variations in EFs were strongly linked to soil C:N ratio and soil mineral N content for CO2 and fertilizer-induced N2O emissions, and to precipitation, water table level and soil NH4⁺ content for peat decomposition N2O emissions. These results suggest that current IPCC EFs for oil palm on organic soil could over-estimate peat onsite CO2 emissions and underestimate peat decomposition N2O emissions and that temporal variation in emissions should be considered for further improvement of EFs.
... Of course, this ecosystem restoration process is carried out by considering the position of the area itself, whether it is included in conservation, protection, production, cultivation, or area for other uses. The condition of the ecosystem as a result of restoration can be used as an indicator of the success of ecosystem restoration because restoration aims to oversee the ecosystem recovery process so that it can provide its original functions [27], [37], [35], [17] Restoration of the peat ecosystem post-fires is full of various challenges. The peatland ecosystem itself has unique characteristics, not to mention that repair also involves all aspects so that the function of the peat swamp forest ecosystem is restored to its original state [38], [39], [40], [41]. ...
... Enrichment also faces many obstacles due to the hightemperature post-fire land conditions and the significant difference between summer and rainy. In contrast, woody species require wetland conditions, high biomass, relatively cold and warm temperatures [1], [35], [37] Through Presidential Decree No. 1 of 2016, the government decided that peat restoration activities should be carried out through the same coordination door, namely the Peat Restoration Agency (BRG), which carries out peat restoration activities in 3 ways rewetting, revegetation, and peat revitalization. On average, the trend of destruction of tropical rain forests, including peat swamp forests for agricultural extensification, can be restored at least on a local scale [44] Restoration is carried out utilizing vegetation, hydrological, and revitalization. ...
Article
Full-text available
Tropical peat swamp forest is one of the wetland ecosystems on tropical peatlands with many ecological, economic, and socio-cultural functions. In Indonesia, the peat swamp forest ecosystems have been experiencing deforestation and degradation due to land clearing for plantations and agriculture and forest fires. In Central Kalimantan, especially in the ex-area of the 1 million hectares mega rice project (MRP)n in the 1990s, hydrological restoration is done by blocking the canals. We compared the three methods of canal blocking and the areas without canal blocking and the community’s preference on what form of canal blocking is more beneficial for them. Large canal blocking, medium canal blocking, and small canal blocking had positively affected the groundwater level in the driest month above the fire-prone critical point. In contrast, the locations without blocking exceed the necessary fire-prone water level. Small, large, and medium blocking are equally capable of optimizing the peat soil water table. However, the local communities preferred small blocking over other methods because it was simple, labour-intensive, and improved their livelihood when involved in its construction. The local communities choose the big canal blockings less because they block transportation access in and out of the peat swamp forest.
... At lower latitudes and elevations, peat swamp forests are the dominant peatland types (Gumbricht et al., 2017;Page et al., 2011). These are dominated by angiosperm trees and palms, and hence the dominant inputs to these peatlands are woody or palm litter with a relatively high lignin content (Dargie et al., 2017;Gandois et al., 2013;Könönen et al., 2016;Lahteenoja and Page, 2011;Lahteenoja et al., 2009). However, our measurements of Shorea albida, a tree dominant in tropical Southeast Asia, were roughly 40% carbohydrate and 30% aromatic as determined with FTIR (Hodgkins et al., 2018). ...
... The combination of the inhibitory effects of aromatic compounds from plants such as Sphagnum and shrubs (Wang et al., 2015), and the higher recalcitrance of lignin-like compounds, make aromatics one of the most influential factors in peat stability. Aromatic compounds from Sphagnum and vegetation with similar chemical composition found at higher latitudes are significant for peat stability in northern peatlands because the plant tissue decomposition rates are much slower than that of other source vegetation (Könönen et al., 2016;Moore et al., 2007;Scanlon & Moore, 2000), but the higher initial aromatic content of source vegetation at lower latitudes shown in Hodgkins et al. (2018) could increase the recalcitrance of the soil organic matter even without inhibitory compounds from Sphagnum (Könönen et al., 2015(Könönen et al., , 2018. Our FTIR-derived aromatics have been calibrated against "Klason Lignins" (Hodgkins et al., 2018), a quite diverse operationally defined pool that includes not just structural lignin, but also other aromatics such as tannins, and other non-lignin-derived polyphenols (De la Cruz et al., 2016). ...
Article
Full-text available
Peatlands contain a significant fraction of global soil carbon, but how these reservoirs will respond to the changing climate is still relatively unknown. A global picture of the variations in peat organic matter chemistry will aid our ability to gauge peatland soil response to climate. The goal of this research is to test the hypotheses that (a) peat carbohydrate content, an indicator of soil organic matter reactivity, will increase with latitude and decrease with mean annual temperatures, (b) while peat aromatic content, an indicator of recalcitrance, will vary inversely, and (c) elevation will have a similar effect to latitude. We used Fourier Transform Infrared Spectroscopy to examine variations in the organic matter functional groups of 1034 peat samples collected from 10 to 20, 30–40, and 60–70 cm depths at 165 individual sites across a latitudinal gradient of 79°N–65°S and from elevations of 0–4,773 m. Carbohydrate contents of high latitude peat were significantly greater than peat originating near the equator, while aromatic content showed the opposite trend. For peat from similar latitudes but different elevations, the carbohydrate content was greater and aromatic content was lower at higher elevations. Higher carbohydrate content at higher latitudes indicates a greater potential for mineralization, whereas the chemical composition of low latitude peat is consistent with their apparent relative stability in the face of warmer temperatures. The combination of low carbohydrates and high aromatics at warmer locations near the equator suggests the mineralization of high latitude peat until reaching recalcitrance under a new temperature regime.
... Indonesia is a country blessed with the largest area of tropical peatlands in the world, where almost 70% of the total global tropical peatlands are in Indonesia, spread over the islands of Sumatra, Kalimantan, and Papua S. E. Page et al., 2007). Indonesia's tropical peatlands primarily consist of forested wetland areas and are called tropical peat swamp forests (Könönen et al., 2016). These fragile and unique ecosystems have vital roles for human life with various ecological, economic, and socio-cultural benefits. ...
... Forest fires in 2006 showed that fires were the most destructive cause of degradation and deforestation in peat swamp forests, especially from the perspective of vegetation biodiversity. Vegetation does not always return naturally and reduces the chance of natural regeneration (Könönen et al., 2016). Forest fires also remove seeds that fall and are stored on the forest floor. ...
Article
Full-text available
The massive forest fire disasters have left an enormous area of degraded peatland. This study aims to analyze the performance of two species, namely C. arborescens and C. rotundatus, as the natural regeneration post forest fires. This research was conducted in 5 different locations that experienced severe fires in 2006. We made a total of 25 plots for each location to measure biodiversity at four growth levels. We analyzed the data with vegetation analysis formulas from Magurran. The results show that at the tree growth level, C. rotundatus can withstand the fires in 2006 and is currently still growing in more significant numbers than C. arborescens. At the pole, sapling, and seedling growth levels, these species perform well as natural regeneration species with many individuals, but C. arborescens is a bit more dominant. Both species are suitable for natural regeneration after fires in degraded peat swamp forests based on survived and existing individuals. On the other hand, both species could not improve the vegetation diversity in the whole ecosystem. These two species can be the option for natural regeneration if there a limited budget and the degraded areas are in a very remote location.
... Not all microorganisms are able to produce enzymes for lignin decomposition (Asina et al. 2016), so lignin is considered a resistant compound for biodegradation. Peatlands with high lignin content have a low rate of CO2 emission (Könönen et al. 2016). Saidy et al. (2018) developed a model of greenhouse gas emissions of tropical peatlands and used peat lignin content as a parameter to predict the amounts of CO2 produced in tropical peatlands, in which increasing peat lignin content suppresses the rate of CO2 emissions. ...
Article
Full-text available
Saidy AR, Priatmadi BJ, Septiana M. 2022. Diversity in characteristics of tropical peatlands varying in land uses leads to differences in methane and carbon dioxide emissions. Biodiversitas 23: 6293-6301. The use of tropical peatlands for commercial agriculture causes a change in their original function as carbon storage to become sources of methane (CH4) and carbon dioxide (CO2) emissions. Therefore, this research aims to quantify the emission of CH4 and CO2, and peat characteristics in five tropical peatlands with different land uses, namely shrubs-, burned-, Albizia-, spring onion-, and lettuce-peatlands, to determine factors controlling carbon emissions. The results showed that CH4 emission ranged from 0.21 to 0.58 mg C m-2 h-1, with the lowest and the highest obtained in burned- and cultivated-peatlands (spring onion- and lettuce peatlands), respectively. The emission of CO2 ranged from the lowest in burned peatland (34.10-47.06 mg C m-2 h-1) to the highest in shrubs-peatland (136.79-180.87 mg C m-2 h-1). This showed that the diversity in CH4 emissions with different land uses is attributed to variations in the water table, water-filled pore space, ammonium, and nitrate contents. The rates of CO2 emission were controlled by carbohydrate-, fiber-, organic C-, and lignin-contents. This indicated that land and water managements need to be applied to reduce the emissions of CH4 and CO2 in tropical peatlands with different land uses.
... In the same study, a relative enrichment of pyrogenic C with ongoing degradation was observed. A higher contribution of aromatics and other recalcitrant moieties was also measured for drained tropical peatlands (Könönen et al., 2016;Gandois et al., 2013;Cooper et al., 2019). An increase in alkyl-C and a corresponding decline in O-alkyl-C is generally related to microbial decomposition of organic material, which is characterized by the accumulation of recalcitrant aliphatic plant compounds and/or microbial material (Baldock et al., 1997). ...
Chapter
Full-text available
Organic soils of intact peatlands store 1/4 of the global soil organic carbon (SOC). Despite being an important source of methane (CH4), they are climate coolers because they continuously accumulate new organic carbon. However, when these organic soils are drained for agriculture, the resulting aerobic conditions lead to fast decomposition of the peat and the release of carbon dioxide (CO2) and nitrous oxide (N2O), turning them into net greenhouse gas (GHG) sources. Reducing the environmental footprint of managing these soils requires a good understanding of the processes during drainage of formerly anoxic soil horizons and eventual subsequent rewetting. We describe changes in soil properties and carbon dynamics following drainage of peatlands and discuss management strategies to reduce carbon loss from drained peatlands by raising the water table to either restore the peatland ecosystem, or to cultivate water-tolerant crops. In addition to rewetting, engineering approaches with continuous management at deeper water tables are evaluated in terms of SOC loss.
... The magnitude of and balance between different peatland C effluxes are primarily linked to abiotic hydrological constraints on peat microbial communities and their respiration pathways, but also to biotic factors. The woody, lignin-rich nature of much of the plant litter supplied to the peat surface confers resistance to microbial decay under anaerobic conditions [70][71][72][73][74] . However, if the peat is oxygenated, easily degraded fractions, such as cellu lose, and more recalcitrant fractions, such as lignin, can both decompose rapidly 71 . ...
Article
Tropical peatlands store around one-sixth of the global peatland carbon pool (105 gigatonnes), equivalent to 30% of the carbon held in rainforest vegetation. Deforestation, drainage, fire and conversion to agricultural land threaten these ecosystems and their role in carbon sequestration. In this Review, we discuss the biogeochemistry of tropical peatlands and the impacts of ongoing anthropogenic modifications. Extensive peatlands are found in Southeast Asia, the Congo Basin and Amazonia, but their total global area remains unknown owing to inadequate data. Anthropogenic transformations result in high carbon loss and reduced carbon storage, increased greenhouse gas emissions, loss of hydrological integrity and peat subsidence accompanied by an enhanced risk of flooding. Moreover, the resulting nutrient storage and cycling changes necessitate fertilizer inputs to sustain crop production, further disturbing the ecosystem and increasing greenhouse gas emissions. Under a warming climate, these impacts are likely to intensify, with both disturbed and intact peat swamps at risk of losing 20% of current carbon stocks by 2100. Improved measurement and observation of carbon pools and fluxes, along with process-based biogeochemical knowledge, is needed to support management strategies, protect tropical peatland carbon stocks and mitigate greenhouse gas emissions. Tropical peatlands hold around 105 gigatonnes of carbon but are increasingly affected by anthropogenic activities. This Review describes the biogeochemistry of these systems and how deforestation, fire, drainage and agriculture are disturbing them. Tropical peatlands are important in terms of the global carbon cycle and in efforts to combat climate change, with a growing recognition of their potential role in natural climate solutions.Tropical peatlands occupy approximately 440,000 km2 across Southeast Asia, Central Africa and South and Central America, and are mostly forested. They are among the world’s most carbon-dense ecosystems with a belowground carbon stock of about 105 gigatonnes (Gt).Although tropical peatlands in Africa and in South and Central America remain largely intact, those in Southeast Asia have undergone widespread transformations owing to deforestation, drainage and agricultural conversion.Land-use changes result in rapid peat carbon loss, high greenhouse gas emissions, land subsidence, changes in hydrology and nutrient cycling, and an increased risk of fire.Management priorities include protection of the carbon sink function of intact forested peatlands; restoration of degraded, forested peatlands; and improved management of agricultural peatlands by raising water levels to mitigate carbon losses and greenhouse gas emissions.The response of tropical peatlands and their carbon stocks to anthropogenic warming and associated changes in hydroclimate remain an area of uncertainty. Tropical peatlands are important in terms of the global carbon cycle and in efforts to combat climate change, with a growing recognition of their potential role in natural climate solutions. Tropical peatlands occupy approximately 440,000 km2 across Southeast Asia, Central Africa and South and Central America, and are mostly forested. They are among the world’s most carbon-dense ecosystems with a belowground carbon stock of about 105 gigatonnes (Gt). Although tropical peatlands in Africa and in South and Central America remain largely intact, those in Southeast Asia have undergone widespread transformations owing to deforestation, drainage and agricultural conversion. Land-use changes result in rapid peat carbon loss, high greenhouse gas emissions, land subsidence, changes in hydrology and nutrient cycling, and an increased risk of fire. Management priorities include protection of the carbon sink function of intact forested peatlands; restoration of degraded, forested peatlands; and improved management of agricultural peatlands by raising water levels to mitigate carbon losses and greenhouse gas emissions. The response of tropical peatlands and their carbon stocks to anthropogenic warming and associated changes in hydroclimate remain an area of uncertainty.
Article
Full-text available
There are limited data for greenhouse gas (GHG) emissions from smallholder agricultural systems in tropical peatlands, with data for non-CO2 emissions from human-influenced tropical peatlands particularly scarce. The aim of this study was to quantify soil CH4 and N2 O fluxes from smallholder agricultural systems on tropical peatlands in Southeast Asia and assess their environmental controls. The study was carried out in four regions in Malaysia and Indonesia. CH4 and N2 O fluxes and environmental parameters were measured in cropland, oil palm plantation, tree plantation, and forest. Annual CH4 emissions (in kg CH4 ha-1 year-1 ) were: 70.7 ± 29.5, 2.1 ± 1.2, 2.1 ± 0.6 and 6.2 ± 1.9 at the forest, tree plantation, oil palm and cropland land-use classes, respectively. Annual N2 O emissions (in kg N2 O ha-1 year-1 ) were: 6.5 ±2.8, 3.2 ± 1.2, 21.9 ± 11.4 and 33.6 ± 7.3 in the same order as above, respectively. Annual CH4 emissions were strongly determined by water table depth (WTD) and increased exponentially when annual WTD was above -25 cm. In contrast, annual N2 O emissions were strongly correlated with mean total dissolved nitrogen (TDN) in soil water, following a sigmoidal relationship, up to an apparent threshold of 10 mg N L-1 beyond which TDN seemingly ceased to be limiting for N2 O production. The new emissions data for CH4 and N2 O presented here should help to develop more robust country level 'emission factors' for the quantification of national GHG inventory reporting. The impact of TDN on N2 O emissions suggests that soil nutrient status strongly impact emissions, and therefore, policies which reduce N-fertilisation inputs might contribute to emissions mitigation from agricultural peat landscapes. However, the most important policy intervention for reducing emissions is one that reduces the conversion of peat swamp forest to agriculture in peatlands in the first place.
Article
Full-text available
Tropical peatlands are among the most carbon-dense ecosystems on Earth, and their water storage dynamics strongly control these carbon stocks. The hydrological functioning of tropical peatlands differs from that of northern peatlands, which has not yet been accounted for in global land surface models (LSMs). Here, we integrated tropical peat-specific hydrology modules into a global LSM for the first time, by utilizing the peatland-specific model structure adaptation (PEATCLSM) of the NASA Catchment Land Surface Model (CLSM). We developed literature-based parameter sets for natural (PEATCLSM Trop,Nat) and drained (PEATCLSM Trop,Drain) tropical peatlands. Simulations with PEATCLSM Trop,Nat were compared against those with the default CLSM version and the northern version of PEATCLSM (PEATCLSM North,Nat) with tropical vegetation input. All simulations were forced with global meteorological reanalysis input data for the major tropical peatland regions in Central and South America, the Congo Basin, and Southeast Asia. The evaluation against a unique and extensive data set of in situ water level and eddy covariance-derived evapotranspiration showed an overall improvement in bias and correlation compared to the default CLSM version. Over Southeast Asia, an additional simulation with PEATCLSM Trop,Drain was run to address the large fraction of drained tropical peatlands in this region. PEATCLSM Trop,Drain outperformed CLSM, PEATCLSM North,Nat and PEATCLSM Trop,Nat over drained sites. Despite the overall improvements of PEATCLSM Trop,Nat over CLSM, there are strong differences in performance between the three study regions. We attribute these performance differences to regional differences in accuracy of meteorological forcing data, and differences in peatland hydrologic response that are not yet captured by our model.
Article
Full-text available
Peatlands are carbon (C) storage ecosystems sustained by a high water level (WL). High WL creates anoxic conditions that suppress the activity of aerobic decomposers and provide conditions for peat accumulation. Peatland function can be dramatically affected by WL drawdown caused by land-use and/or climate change. Aerobic decomposers are directly affected by WL drawdown through environmental factors such as increased oxygenation and nutrient availability. Additionally, they are indirectly affected via changes in plant community composition and litter quality. We studied the relative importance of direct and indirect effects of WL drawdown on aerobic decomposer activity in plant litter. We did this by profiling 11 extracellular enzymes involved in the mineralization of organic C, nitrogen, phosphorus and sulphur. Our study sites represented a three-stage chronosequence from pristine (undrained) to short-term (years) and long-term (decades) WL drawdown conditions under two nutrient regimes. The litter types included reflected the prevalent vegetation, i.e., Sphagnum mosses, graminoids, shrubs and trees. WL drawdown had a direct and positive effect on microbial activity. Enzyme allocation shifted towards C acquisition, which caused an increase in the rate of decomposition. However, litter type overruled the direct effects of WL drawdown and was the main factor shaping microbial activity patterns. Our results imply that changes in plant community composition in response to persistent WL drawdown will strongly affect the C dynamics of peatlands.
Article
Full-text available
Indo-Malaysian tropical peat swamp forests (PSF) sequester enormous stores of carbon in the form of phenolic compounds, particularly lignin as well as tannins. These phenolic compounds are crucial for ecosystem functioning in PSF through their inter-related roles in peat formation and plant defenses. Disturbance of PSF causes destruction of the peat substrate, but the specific impact of disturbance on phenolic compounds in peat and its associated vegetation has not previously been examined. A scale was developed to score peatland degradation based on the three major human impacts that affect tropical PSF—logging, drainage, and fire. The objectives of this study were to compare the amount of phenolic compounds in Macaranga pruinosa, a common PSF tree, and in the peat substrate along a gradient of peatland degradation from pristine peat swamp forest to cleared, drained, and burnt peatlands. We examined phenolic compounds in M. pruinosa and in peat and found that levels of total phenolic compounds and total tannins decrease in the leaves of M.pruinosa and also in the surface peat layers with an increase in peatland degradation. We conclude that waterlogged conditions preserve the concentration of phenolic compounds in peat, and that even PSF that has been previously logged but which has recovered a full canopy cover will have high levels of total phenolic content (TPC) in peat. High levels of TPC in peat and in the flora are vital for the inhibition of decomposition of organic matter and this is crucial for the accretion of peat and the sequestration of carbon. Thus regional PSF flourish despite the phenolic rich, toxic, waterlogged, nutrient poor, conditions, and reversal of such conditions is a sign of degradation.
Article
Full-text available
Insular Southeast Asian peatlands have experienced rapid land cover changes over the past decades inducing a variety of environmental effects ranging from regional consequences on peatland ecology, biodiversity and hydrology to globally significant carbon emissions. In this paper we present the land cover and industrial plantation distribution in the peatlands of Peninsular Malaysia, Sumatra and Borneo in 2015 and analyse their changes since 1990. We create the 2015 maps by visual interpretation of 30 m resolution Landsat data and combine them with fully comparable and completed land cover maps of 1990 and 2007 (Miettinen and Liew, 2010). Our results reveal continued peatland deforestation and conversion into managed land cover types. In 2015, 29% (4.6 Mha) of the peatlands in the study area remain covered by peat swamp forest (vs. 41% or 6.4 Mha in 2007 and 76% or 11.9 Mha in 1990). Managed land cover types (industrial plantations and small-holder dominated areas) cover 50% (7.8 Mha) of all peatlands (vs. 33% 5.2 Mha in 2007 and 11% 1.7 Mha in 1990). Industrial plantations have nearly doubled their extent since 2007 (2.3 Mha; 15%) and cover 4.3 Mha (27%) of peatlands in 2015. The majority of these are oil palm plantations (73%; 3.1 Mha) while nearly all of the rest (26%; 1.1 Mha) are pulp wood plantations. We hope that the maps presented in this paper will enable improved evaluation of the magnitude of various regional to global level environmental effects of peatland conversion and that they will help decision makers to define sustainable peatland management policies for insular Southeast Asian peatlands.
Article
Full-text available
Land-use change has transformed large areas of tropical peatland into globally significant carbon sources. Associated changes in the properties of peat are important for soil processes including decomposition and nutrient cycling. To characterise the changes induced by stabilised land uses, we studied the physical and chemical properties of peat from four land management conditions (undrained and drained forest, degraded land, and managed agricultural land). Peat was sampled from depths of 10–15 cm, 40–45 cm, 80–85 cm and 110–115 cm then partitioned into woody (Ø >1.5 mm), fibric (Ø 0.15–1.5 mm) and amorphic (Ø < 0.15 mm) fractions. Bulk density and total concentrations of ash, C, N, P, K, Ca, Mg, Mn, Zn, Na, Al, Fe, S and Si were determined. There were clear differences between land uses in the characteristics of surface peat down to the 40–45 cm layer, the primary differences being between forested and open sites. Due to smaller particle sizes, the bulk density of peat was higher at the open sites, where Ca and Mg concentrations were also higher but N and P concentrations were lower. Changes in drainage and vegetation cover had resulted in differing outcomes from decomposition processes, and the properties of fire-impacted peats on the open sites had undergone extreme changes. © 2015 International Mire Conservation Group and International Peat Society. Open access at http://mires-and-peat.net/pages/volumes/map16/map1608.php
Article
Increasing oil palm (OP) plantation establishment on tropical peatlands over the last few decades has major implications for the global carbon (C) budget. This study quantified total and heterotrophic soil carbon dioxide (CO2) emissions in an industrial OP plantation (7 year old, 149 trees ha− 1) on peat located in the eastern coast of the Sumatra Island (Jambi district), Indonesia, after two doses of nitrogen (N) fertilizer application at rates typical of local practice. The first dose applied in March 2012 (first Fertilization event FE) consisted of 0.5 kg urea per palm (equivalent to 371 kg N ha− 1 at the base of the palm which when extrapolated across the plantation was 35 kg N ha− 1) and the second dose applied in February 2013 (second FE) amounted to 1 kg urea per palm. Soil CO2 fluxes were measured using an infrared gas analyzer (IRGA) in dark closed chambers. The measurements were made daily from 1 day before to 7 days after fertilizer application. Soil heterotrophic respiration (Rh) and total soil respiration (Rs) were measured in trenched plots (where root respiration was excluded) and non-trenched plots, respectively. Concomitant with CO2 flux measurements, air and soil temperatures, rainfall and the water table level were measured. To estimate the fertilizer effect during the different times of the day, CO2 fluxes were monitored every 3 h during a 24 h period on days 2 and 3 after fertilizer application during the second FE. Shortly after fertilizer application, substantial pulses of CO2 were detected in the IRGA chambers where the fertilizer was applied. Even though the fertilized area represents 9.4% of the plantation area only, the impact of fertilizer application at the plantation scale on CO2 fluxes was noteworthy when compared to non-fertilized control treatments. The Rs was 36.9 kg CO2–C ha− 1 (7 days)− 1 greater in the fertilized than in the non-fertilized plots after the first FE but no enhancement was observed after the second FE (− 72.2 kg CO2–C ha− 1 7 days− 1). The Rh was 340.5 and 98.9 kg CO2–C ha− 1 (7 days)− 1 greater in the fertilized than in the non-fertilized plots after the first and second FE, respectively. The larger CO2 flux enhancement in Rh as compared to Rs may be the result of fertilizer uptake by the palm roots in the un-trenched plots, while in the trenched ones where roots were absent, microorganisms used the fertilizer to accelerate soil organic matter mineralization. Although the response of Rh to N addition and the priming effect were high as compared to results in the literature, the impacts were short-term only and may not have implications on the annual C budget of the plantation.
Chapter
The term lignin is commonly used for unchanged lignin as well as for lignin-derived materials in pulps and pulping liquors. This meaning of the term lignin is retained in this chapter. However, analytical data for unchanged lignin are usually not valid for lignin-derived materials in pulps and pulping liquors because of chemical modifications. Furthermore, analytical techniques used for the examination of unchanged lignins are not always applicable for the examination of chemically modified lignins. Therefore we use the term protolignin(s) (native lignin(s)) when we specifically refer to unchanged or essentially unchanged lignin(s).
Article
Extracted thermomechanical pulp from spruce was fractionated into microfines, fibrils, flakes, ray cells and fibres using centrifugation, sedimentation and filtration steps. The chemical composition and the amount of anionic groups were determined. The various types of fines differed greatly in chemical composition. The content of cellulose was higher in the fibrils and the microfines than in the flakes and the ray cells, whereas the content of lignin was lower. The flakes and the ray cells contained much arabinogalactans, xylans and pectins, but only little galactoglucomannans. The fibrils contained much arabinogalactans and pectins. The increase in the amount of anionic groups during alkaline treatment was to a great extent due to the demethylation of pectins, corroborated by the amount of methanol released. It can be suggested that the fibrils originate mainly from the primary wall and the outer secondary wall. The flakes originate primarily from the middle lamella and the primary wall.
Article
Tropical peatland fires play a significant role in the context of global warming through emissions of substantial amounts of greenhouse gases. However, the state of knowledge on carbon loss from these fires is still poorly developed with few studies reporting the associated mass of peat consumed. Furthermore, spatial and temporal variations in burn depth have not been previously quantified. This study presents the first spatially explicit investigation of fire-driven tropical peat loss and its variability. An extensive airborne LiDAR (Light Detection and Ranging) dataset was used to develop a pre-fire peat surface modeling methodology, enabling the spatially differentiated quantification of burned area depth over the entire burned area. We observe a strong interdependence between burned area depth, fire frequency and distance to drainage canals. For the first time, we show that relative burned area depth decreases over the first four fire events and is constant thereafter. Based on our results, we revise existing peat and carbon loss estimates for recurrent fires in drained tropical peatlands. We suggest values for the dry mass of peat fuel consumed that are 206 t ha(-1) for initial fires, reducing to 115 t ha(-1) for second, 69 t ha(-1) for third and 23 t ha(-1) for successive fires, which are 58% to 7% of the current IPCC Tier 1 default value for all fires. In our study area, this results in carbon losses of 114, 64, 38 and 13 t C ha(-1) for first to fourth fires, respectively. Furthermore, we show that with increasing proximity to drainage canals both burned area depth and the probability of recurrent fires increase and present equations explaining burned area depth as a function of distance to drainage canal. This improved knowledge enables a more accurate approach to emissions accounting and will support IPCC Tier 2 reporting of fire emissions. This article is protected by copyright. All rights reserved.