ArticlePDF Available

Carbon sequestration in mangrove forests

Taylor & Francis
Carbon Management
Authors:
  • Tropical Coastal & Mangrove Consultants

Abstract and Figures

Mangrove forests are highly productive, with carbon production rates equivalent to tropical humid forests. Mangroves allocate proportionally more carbon belowground, and have higher below- to above-ground carbon mass ratios than terrestrial trees. Most mangrove carbon is stored as large pools in soil and dead roots. Mangroves are among the most carbon-rich biomes, containing an average of 937 tC ha-1, facilitating the accumulation of fine particles, and fostering rapid rates of sediment accretion (∼5 mm year -1) and carbon burial (174 gC m-2 year -1). Mangroves account for only approximately 1% (13.5 Gt year -1) of carbon sequestration by the world’s forests, but as coastal habitats they account for 14% of carbon sequestration by the global ocean. If mangrove carbon stocks are disturbed, resultant gas emissions may be very high. Irrespective of uncertainties and the unique nature of implementing REDD+ and Blue Carbon projects, mangroves are prime ecosystems for reforestation and restoration.
Content may be subject to copyright.
future science group
313
ISSN 1758-3004
10.4155/CMT.12.20 © 2012 Future Science Ltd
Mangrove forests are the only woody halophytes that live
in salt water along the world’s subtropical and tropical
coastlines. Coincidentally, poverty and dense human
populations flourish along these low-latitude coasts,
partly explaining the high (1–3%) annual de forestation
rates of these tidal forests. Mangroves are true ecotones,
having some components of both marine and terrestrial
biomes, but have also developed a number of unique
structural and functional adaptations, such as viviparous
embryos, physiological mechanisms to tolerate salt and
aerial roots that enable the plants to respire in anoxic,
waterlogged soils [1]. Mangroves are architecturally sim-
ple compared with terrestrial forests, usually harboring
few tree species and lacking an understory of ferns and
scrubs. However, the standing biomass of some mangrove
forests in equatorial regions can be immense, rivaling the
height and weight of many tropical rainforests [1].
Mangroves are ultimately limited by temperature but,
at local and regional scales, variations in pre cipitation,
tides, waves and river flow greatly determine their expanse
and biomass. Attempts have often been made to classify
the sequential changes in forest structure and species dis-
tribution parallel to shore but, in reality, most mangrove
forests represent a continuum of types in relation to
gradients in their physical settings. Variations can be
expressed within a single estuary, where there are usually
upstream–downstream changes in geo morphology, salin-
ity, waves, tides and river flow, with these factors affecting
water circulation by generating mixing and trapping of
coastal water [2]. The development of mangrove forests
occurs where near-horizontal topography coincides with
sea level; a relatively stable period of sea level is, thus, a
prerequisite for the development of old-growth forests
[3]. The response of mangroves to environmental change
is, therefore, often indicative of past changes in coastal
conditions, especially in sea level. Comparing present
patterns in forest species with paleoecological informa-
tion provides considerable insight, not only into how
mangroves responded to past sea level changes, but how
they may respond to climate change in the future.
Human disturbance obscures natural change and
our ability to distinguish one from the another is lim-
ited, as most forests have a history of both natural and
human disturbances, and are often intertwined and
indistinguish able. Mangroves are naturally disturbed by
tsunamis, floods, cyclones, lightning, pests and disease,
Carbon Management (2012) 3(3), 313–322
Carbon sequestration in mangrove forests
Daniel M Alongi*
Mangrove forests are highly productive, with carbon production rates equivalent to tropical humid forests.
Mangroves allocate proportionally more carbon belowground, and have higher below- to above-ground
carbon mass ratios than terrestrial trees. Most mangrove carbon is stored as large pools in soil and dead
roots. Mangroves are among the most carbon-rich biomes, containing an average of 937tCha
-1
, facilitating
the accumulation of ne particles, and fostering rapid rates of sediment accretion (~5 mm year
-1
) and
carbon burial (174gC m
-2
year
-1
). Mangroves account for only approximately 1% (13.5Gtyear
-1
) of carbon
sequestration by the world’s forests, but as coastal habitats they account for 14% of carbon sequestration
by the global ocean. If mangrove carbon stocks are disturbed, resultant gas emissions may be very high.
Irrespective of uncertainties and the unique nature of implementing REDD+ and Blue Carbon projects,
mangroves are prime ecosystems for reforestation and restoration.
Review
*Australian Institute of Marine Science, PMB 3, Townsville MC, Queensland 4810, Australia
E-mail: d.alongi@aims.gov.au
Carbon Management (2012) 3(3)
future science group
314
Review Alongi
and become more susceptible when
human stressors such as pollutants
are introduced. However, mangroves
often exhibit considerable resilience
to disturbance, undergoing perpet-
ual change in ecosystem develop-
ment commensurate with the evo-
lution of the environmental settings
they inhabit, and are, thus, mosaics
of successional stages arrested or
interrupted over time and space by
natural ecological responses in rela-
tion to disturbances both large and
small [4].
Mangrove forests are a valuable
ecological and economic resource,
providing essential services such as
food and fuel resources; nursery
grounds for fish, mammals and
other semi-terrestrial and aquatic fauna; depocenters
for sediment, carbon and other elements; and, in some
instances, offering some protection from coastal erosion
due to tsunamis and intense tropical storms [1]. Despite
their uses to humans, approximately 50% of the worlds
mangrove forests have disappeared over the past 50 years
[5], ironically reflecting their importance as a valuable
economic resource. Major causes for this destruction
have been urban development, aquaculture, mining, and
overexploitation of timber, sh, crustaceans and shell-
fish. The average monetary value of mangroves has been
estimated as second only to the value of estuaries and
seagrass meadows, and greater than the economic value
of coral reefs, continental shelves and the open sea [6].
Of greater eventual value is the role of mangroves in
storing carbon to help ameliorate the impact of climate
change. There is a growing consensus that it will be
impossible to achieve significant cuts in GHG emis-
sions without passive and active means to capture and
store CO
2
[7]. The role of carbon storage in mangroves
has often been overlooked and either underestimated
or overestimated [1], and it is the purpose of this review
to critically assess the role of mangroves in carbon
sequestration
and its global significance.
Carbon production
Mangroves are usually highly productive forests and, as
a significant fraction of their soil carbon is plant-derived
[8], it is crucial to assess rates of net primary productivity
of mangroves and associated plants, especially benthic
microalgae. Measurement of primary production in
mangrove forests is limited by methodological short-
comings, but the best estimates suggest that mangrove
carbon production is more rapid than other estuarine
and marine primary producers [9]. Rates of mangrove net
primary production (NPP) based on different methods
range from 0.5 to 112.1 t dry weight (DW) ha
-1
year
-1
but most methods either significantly overestimate (the
light attenuation method) or underestimate (litterfall)
the true rates of production.
The most reasonable means at present to assess NPP of
forests is to measure aboveground biomass accumulation
plus litterfall, and there are quite a number of such mea-
surements for both mangroves and tropical terrestrial
forests. For mangroves, the mean rate of above ground
NPP is 11.1 t DW ha
-1
year
-1
with a median value of
8.1 t DW ha
-1
year
-1
; for tropical terrestrial forests, the
mean rate of aboveground NPP is 11.9 t DW ha
-1
year
-1
with a median value of 11.4 t DW ha
-1
year
-1
; for both
mangroves and terrestrial forests, NPP declines with
increasing latitude [1]. Considering the differences within
and between both forest groups in biomass, height, age
and species, the rates are very close and clearly imply
that rates of NPP are equivalent between mangroves and
other forests.
Like other forests, mangroves vary in size and age and,
therefore, vary in rates of production and in the bal-
ance between carbon production and respiration. The
few studies that have measured mangrove tree growth
over time or in stands of known age have observed stand
dynamics similar to other forests, identifying stages of
early rapid growth during colonization and early estab-
lishment, followed by a slow decline in growth rate into
maturity and senescence [1,10,11]. The stable-state matu-
rity phase can be prolonged in some mangrove stands
and may represent an alternate succession state in which
the clock for the climax stage is reset by successive dis-
turbances [10]. The relationship between mangrove forest
age and photosynthetic production [11] suggests prolonga-
tion or arrested progression when forests are disturbed;
Rhizophora apiculata forests in southeast Asia show log-
phase photosynthetic rates until approximately 20 years,
after which photosynthesis levels off but does not signifi-
cantly decline for nearly a century [1]. These data imply
that mangroves might indeed constitute a carbon sink
for up to a century if left relatively undisturbed.
Other primary producers inhabit mangrove forests
and their rates of NPP can be significant, especially in
comparatively open canopies and on tidal banks where
sufficient light penetrates to the forest floor [1]. Various
autotrophic and mixotrophic microbes and microalgae,
as well as macroalgae, live on the soil surface and as
epiphytes on tree parts, especially aerial roots and decom-
posing wood. The quantitative contribution of these
smaller autotrophs is dwarfed by tree production, but
belies their importance as food and refugia for consum-
ers. However, some evidence suggests that they can play
an important role in soil carbon and nitrogen cycling,
especially when found as intact mats [12].
Key terms
Mangroves: Trees and associated
plants, microbes and animals that live at
the interface between land and sea.
These tidal ecosystems have both
semi-terrestrial and marine
components.
Coastal: Land, water and aquatic
habitats that reside where the
continents meet the ocean. These
habitats are usually only a few
kilometers in width but are highly
dynamic and interactive with respect to
energy and material flow between land
and sea.
Carbon sequestration: Term used to
describe the acquisition and storage of
carbon. Refers most often in relation to
the ability of ecosystems to reduce the
impact of increasing CO
2
concentrations
in the atmosphere.
Carbon sequestration in mangrove forests Review
future science group
www.future-science.com
315
Carbon allocation & ecosystem storage
Critical to our ability to estimate the role of mangroves
in coastal and global carbon cycling is an accurate under-
standing of where carbonxed by the trees is allocated.
Like other woody plants, mangroves construct new foli-
age, reproductive organs, stem, branches and root tissues
and maintain existing tissue, as well as creating storage
reserves and providing chemical defense. Approximately
half of all CO
2
assimilated by mangroves is returned to
the atmosphere via above- and below-ground respiration
[1,11]. This is only a crude estimate owing to the lack
of empirical data and the difficulty of measuring root
processes and respiration of woody parts. The propor-
tional allocation of xed carbon within trees varies with
many factors, such as light intensity, species composi-
tion, nutrient and water availability, salinity, tides, waves,
temperature and climate [11].
The greatest unknown with regard to carbon alloca-
tion is root production, which is difficult to measure,
especially in waterlogged soils. The few studies that have
measured root growth in situ estimated rates ranging from
18 to 1145 g DW m
-2
year
-1
with most estimates between
300 and 380 g DW m
-2
year
-1
[1]. These estimates are at
the lower end of the range of values measured in tropical
terrestrial forests [13]. However, most measurements were
made in mangrove fringe stands, so it is likely that the
growth and production of mangrove roots is similar to
their terrestrial counterparts. A recent ana lysis of carbon
allocation suggests that mangroves allocate proportion-
ally more carbon belowground than terrestrial trees [14].
Carbon inventories from a number of mangrove eco-
systems show that both above- and below-ground biomass
increases, and that the ratio of below- to above-ground
biomass decreases with increasing stand age (Table1).
These data show that belowground carbon biomass is,
on average (mean = 1.3), equivalent to carbon allocated
aboveground; other studies have indicated that more car-
bon biomass is allocated belowground [15–18] supporting
the notion that mangroves store a disproportionate frac-
tion of fixed carbon underground. Further, the amount
of soil carbon increases with forest age (see Figure 5.1 in
Alongi [1]).
Complete inventories of ecosystem components show
that carbon xed within the forest, as well as carbon
imported from adjacent terrestrial and marine waters,
are stored as large pools of soil carbon [19,20]. Analysis of
carbon in Rhizophora stylosa and Avicennia marina in arid
coastal areas of Western Australia [19] and in R. apiculata
forests in southern Thailand [20] showed that although
most carbon was unassociated with roots, the majority
(75–95%) of tree carbon belowground was vested in
dead, rather than live, roots. The Thai study also showed
that the soil and dead root carbon pools increased in size
with increasing stand age [20].
A recent assessment of carbon stored in various forest
domains found that in comparison with boreal, temper-
ate and tropical upland forests, mangroves throughout
the Indo-Pacific are among the most carbon -rich forests
in the tropics containing, on average, 1023 tC ha
-1
, most
of which is stored in soils >30 cm deep [21]. Adding pub-
lished and unpublished data by authors from southern
China, Vietnam, Indonesia, arid Western Australia,
Queensland, Thailand and Malaysia (Table1) to the data
set of Donato et al. [21] to diversify the geographical,
subtropical and arid-zone forest domains, we obtain a
revised mean whole-ecosystem carbon storage estimate
of 937 tC ha
-1
(Figure1), which still indicates that man-
groves are among the worlds most carbon-rich forests.
It is possible, of course, this statement may not hold true
globally, especially when data is obtained from Central
and South America and Africa, and from more forests in
the arid tropics and subtropics where fringing mangroves
and mangroves growing on hard and/or substrates of
limited depth are common. Nevertheless, throughout the
equatorial regions (e.g., the wet tropics of southeast Asia)
it is true that mature mangrove stands attain highest car-
bon mass compared with other carbon-rich ecosystems,
such as tropical rainforests.
What does inarguably appear to be a global pattern
among mangrove forests is that their belowground pools
of root and soil carbon are large, having a higher below-
to above-ground carbon mass ratio than any other woody
vegetation [22].
With the bulk of belowground carbon stored in dead
roots and soil rather than in live roots, mangroves have
a tendency to accumulate carbon relatively quickly.
Belowground roots may only represent approximately
10–15% of total tree biomass, but the allocation of fixed
carbon to replace sloughed root hairs and fine roots is
considerably greater [23,24]. Moreover, carbon concen-
trations in dead roots are greater than in live roots,
suggesting that dead roots store proportionally more
carbon [19,20].
Vertical profiles of live versus dead root matter in a
number of mangroves show that most living roots are
shallow, within the upper 040 cm of soil [1]. Most fine
roots are dead, probably the net result of rapid root turn-
over coupled with slow rates of root decomposition [23].
Rates of belowground decomposition of fine and coarse
mangrove roots are indeed slow, with most rates ranging
from 0.07 to 0.17% root mass lost per day; only roots of
A. marina decompose more quickly at rates varying from
0.09 to 0.34% root mass lost per day [1]. Roots decom-
pose at equivalent rates regardless of intertidal elevation,
but coarse roots decompose less quickly than fine roots.
These slow decay rates explain the formation of peat in
many mangrove forests as inputs must exceed decay rates
in order for peat to accumulate [23–25].
Carbon Management (2012) 3(3)
future science group
316
Review Alongi
Why do mangrove forests have such large amounts of
carbon vested belowground compared with terrestrial
forests? The presence of a large pool of dead roots can
serve as a nutrient conserving mechanism, and even large
dead roots may serve this purpose. For instance, old root
channels have been found in mangroves in central Belize
with a proliferation of living roots among the decaying
roots, taking paths of least resistance and recovering
nutrients released from decomposing roots [25]. A large
pool of belowground live and dead root biomass mixed
with rich soils may reflect their numerous physiological
and morphological adaptations to life in a harsh, saline
waterlogged environment. Salt negatively affects water
use and under such conditions it may be advantageous
for mangrove trees to invest more fixed carbon in grow-
ing very expensive root systems that turnover rapidly in
order to maximize water gain. Large reservoirs beneath
the forest floor may also help to stabilize the trees and
the entire ecosystem from the continual push and pull
of the tides, wave action, coastal winds and tropical
storms. It makes evolutionary sense for mangroves to
invest in a large belowground pool of carbon biomass
as an effective counterbalance to litter and carbon dis-
solved in interstitial water that is
lost via the tides. Whereas tropical
humid forests recycle nutrients by
rapid soil decomposition of litter in
a relatively thin humus layer, man-
groves reclaim elements by way of
very tight cycling between roots and microbes several
meters deep into the soil, possibly to curtail losses and
to minimize energetic costs.
Mechanisms facilitating sediment accumulation
Lying at the interface between land and sea, it is hardly
surprising that mangroves accumulate sediment and
associated particulate elements, such as inorganic and
organic carbon. What is surprising is that their presence
actively facilitates the accumulation of materials [26].
Carbon is accumulated in mangroves by direct inputs
of mangrove carbon to the soil pool and by increasing
rates of mass sediment accumulation. Carbon produced
by mangroves does have other flow pathways, such as
consumption by living organisms, especially microbes.
Carbon consumed is remineralized and either emitted
back to the atmosphere as CO
2
or exported by dissolved
inorganic carbon. Dissolved and particulate organic
carbon is also exported by tides where it can be either
deposited or eaten or mineralized offshore.
The amount of carbon stored in mangrove soils varies
widely, from <0.1% by soil dry weight to >40% with a
grand median of 2.2% [8]. A highly variable proportion
of this carbon is mangrove-derived as organic matter is
brought in by the tides from adjacent seagrass mead-
ows, coral reefs, macroalgae, rivers and from land-based
sources, and other marine environments [8]. The frac-
tion of mangrove-derived carbon in forest soils depends
on a number of factors, including location of the forest
Key term
Flocculation: Physical, chemical and
microbial processes by which particles
are cemented together; the term ‘floc
refers to the cemented tuft-like mass.
Table1. Whole-ecosystem inventories of above- and below-ground carbon biomass and soil carbon for natural and replanted
mangrove forests.
Location Dominant species Age
(years)
Total
(tCha
-1
)
AGB
(tCha
-1
)
BGB and
soil (tCha
-1
)
Roots/AGB
(tCha
-1
)
Roots
(tCha
-1
)
Soil
(tCha
-1
)
Soil depth
(cm)
Peninsular Malaysia Rhizophora apiculata 80 2205 312 1893 NA NA NA 3800
R. apiculata 18 1117 193 924 NA NA NA 4000
R. apiculata 5 479 87 392 NA NA NA 2800
Southern Vietnam R. apiculata 6 1179 54 1125 NA NA NA 3400
R. apiculata 20 979 72 907 NA NA NA 2750
R. apiculata 35 1904 153 1752 NA NA NA 3600
Southern
China
Kandelia candel NA 619 64 555 2.0 130 425 1850
K. candel NA 391 43 348 2.2 94 254 1900
K. candel NA 332 7 325 1.1 8 317 1175
Indonesia Avicennia marina NA 437 24 413 NA NA NA 80
Rhizophora stylosa NA 703 19 684 NA NA NA 62
Sonneratia caseolaris NA 654 28 626 NA NA NA 1450
Southern Thailand R. apiculata 25 808 138 670 1.0 142 528 1900
R. apiculata 5 579 20 559 2.9 57 502 800
Ceriops decandra 3 600 29 571 4.4 127 444 1000
Western Australia R. stylosa NA 863 115 621 1.1 127 621 1500
A. marina NA 662 55 515 1.7 92 515 775
Queensland, Australia R. stylosa NA 2139 297 1842 1.1 312 1530 3500
AGB: Aboveground biomass; BGB: Belowground biomass; NA: Not available.
Data from [48,50–54,101].
Carbon sequestration in mangrove forests Review
future science group
www.future-science.com
317
in relation to the open coast, dis-
tance to adjacent aquatic habitats,
tidal amplitude, forest position in
the tidal seascape and productivity
of primary producers [27].
Unconsolidated sediments accu-
mulate in relation to the movement
of the turbidity maximum zone,
where incoming bottom flow meets
outward river flow. Tidal mixing
and pumping within the moving
zone facilitate particle occulation
and settlement. Flocculation of
particles begins at salinities <1, and
small flocs and free particles move
downstream where they aggregate
with local particles [28]. As flocs get
larger, they move toward the river
bed where they are entrained back
upstream by baroclinic circulation
and even further upstream at flood
tide due to tidal pumping [28]. As these flocs move into
the forest on flood tides, turbulence generated by flow
around the trees helps to maintain flocs in suspension
[2]. Settling occurs quickly, facilitated by the sticking of
microbial mucus in the soil surface and by pelletization
by invertebrate excreta. Large quantities of nonfloccu-
lated particles are re-exported from the forest on ebb
tide, but most stick to mucus at the water surface.
Mangroves thus actively capture silt, clay and organic
particles, and are not just passive importers of fine par-
ticles [2,28]; mangrove vegetation has a profound impact
on sedimentation. Large trees with complex root sys-
tems, such as Rhizophora species, facilitate the deposi-
tion of particles to a much larger extent than trees that
are smaller and of much simpler architecture, such as
Ceriops species. Until slack water, turbulent wakes cre-
ated by tree trunks, prop roots and pneumatophores
maintain particles in suspension, but most flocs settle
within 30 min just before slack high tide [28]. Despite the
pull of the ebb tide, most flocs are retained within the
forest as water motion and turbulence necessary for their
resuspension is inhibited by the high vegetation density.
Rates of soil accretion & carbon sequestration
Mangroves accumulate carbon in tree biomass, but
much of this carbon is eventually lost in the short- and
medium-term by way of clear-cutting and human use,
decomposition and export to adjacent ecosystems. Over
the long term, carbon is stored primarily belowground as
soil carbon and, eventually, under the right conditions,
as peat. There are a number of methods to measure
soil accretion [29], but some are either highly inaccurate
(a mass balance approach where carbon inputs minus
carbon losses equal carbon either buried or unaccounted
for) or reflect mostly modern rates of accumulation
(measurement of short-term sediment accumulation
using sediment traps or changes in depth of the soil
profile). Analysis of radioactive elements produced by
fallout (excess
210
Pb and
137
Cs) from atomic bomb test-
ing in the atmosphere coupled with estimates of soil
carbon concentrations provide longer term estimates
of accumulation and a chronology of sedimentation of
up to a century. Such methods also have their pitfalls,
including reliance on expensive analytical equipment,
difficulty in interpreting radiotracer profiles in biotur-
bated and disturbed soils and in soils where there are
vertical changes in grain size, and problems with error
induced by compaction of sediments as a result of the
coring process [29].
The rate of soil accretion in mangrove forests averages
5 mm year
-1
, with 94 measurements out of a total of
139 ranging from 0.1 to 10.0 mm year
-1
(Figure2). The
median value is 2.7 mm year
-1
with a few measurements
showing net erosion (minimum value = -11.0 mm year
-1
)
or massive accretion (46.3 mm year
-1
) in highly-impacted
estuaries, such as those in southern China [30].
Frequency of tidal inundation is the primary factor
controlling the rate of accretion [31–33] . Less frequent
inundation by tides means less input of sediment par-
ticles; forests located in the high intertidal area experi-
ence less soil accretion than forests closer to mean sea
level, such as fringing stands at the sea–forest interface.
In fact, mangrove carbon often accumulates on adjacent
margins and intertidal mudflats [34]. Often overlooked,
because empirical data are rare, are contributions to ver-
tical accretion from the growth of belowground roots
Boreal Temperate Tropical Mangroves
0
200
400
600
800
1000
1200
Ecosystem carbon storage (tonnes ha
-1
)
Aboveground carbon storage
Belowground carbon storage
Figure1. Dierences in whole-ecosystem carbon stocks among boreal, temperate and
tropical terrestrial forests, and subtropical and tropical mangrove forests.
Reprinted with permission from Macmillan Publishers Ltd: Nature Geoscience
[21] © (2011).
Mangrove data taken from supplementary data
[102] in[21], and [48,50–54].
Carbon Management (2012) 3(3)
future science group
318
Review Alongi
and surface growth of microbial mats and turf algae
and accumulation of litter. In some cases, such as in the
Caribbean, contributions from these biological sources
can be greater than accretion of mineral particles [33].
On other islands such those in the Federates States of
Micronesia, natural subsidence plays a key role in overall
rates of net elevation [35], although not the actual rates
of soil accretion; nevertheless, such changes are impor-
tant in determining the susceptibility of mangroves
to changes in sea level [35]. Rates of soil accretion can
vary over long timescales. In lagoon mangroves on the
Yucatan Peninsula in Mexico, for instance, natural vari-
ations in accumulation rates and sources of soil carbon
were detected over the past 160 years [36]. These changes
corresponded to fluctuations in climatic variability in
the region.
Mangrove sedimentation in rela-
tion to sea level rise was assessed
by Alongi, who found that most
mangrove forests were currently
keeping pace with local rises in sea
level [37]. However, there are a num-
ber of regions where sedimentation
rates are lower than the rates of
regional relative sea level rise, such
as on some Pacic Islands [36] and
at a number of mangrove stands
in the Caribbean [33,38], although
accretion rates at a number of these
endangered forests is higher than
the eustatic sea level rise.
Available data on burial rates of
carbon in mangrove ecosystems
were rst compiled by Twilley et al.
[39], later updated by Jennerjahn
and Ittekot [40] and Duarte et al.
[9], based on data in Chmura et al.
[41]. Despite the different databases
and methods used, all derived a
similar estimate of a global car-
bon burial rate of approximately
23 TgC year
-1
, which is equiva-
lent to a rate of 167 gC m
-2
year
-1
assuming a total mangrove area of
137,760 km
2
[5]. Bouillon et al. [26,27]
and Alongi [1] derived carbon
burial rates of 18.4 TgC year
-1
(= 134 g C m
-2
year
-1
) and 29 TgC year
-1
(= 211 g C m
-2
year
-1
), respectively.
Adding more recent data derived
from radiochemical methods, we
can revise the mean global burial
rate for soil carbon to 24 TgC year
-1
,
equivalent to 174 g C m
-2
year
-1
with
values ranging from 10 to 920 g C m
-2
year
-1
; the median
burial rate was 16 TgC year
-1
(= 115 g C m
-2
year
-1
).
Like the sediment accretion data, the standard devia-
tion exceeds the mean reflecting the high level of
variability (and uncertainty) in carbon burial rates
among forests worldwide. Nevertheless, most individ-
ual estimates (47 of a total of 66 measurements) are
<200 g C m
-2
year
-1
, with a minority of forests accumu-
lating soil carbon faster (Figure3), mostly in catchments
heavily impacted by human activities, such as those in
southern China [30] and in southeast Asia [42].
Signicance of mangroves to terrestrial &
marine carbon sequestration
How do these new estimates of carbon sequestration
compare with other forested and coastal ecosystems?
Globally, are mangroves a signicant sink for carbon?
Does their loss represent a signicant return of CO
2
to the atmosphere?
The data presented here conrm the notion that
mangroves are among the most carbon-rich eco systems
in the tropics. But at a global level, mangroves occupy
only approximately 137, 760 km
2
, and a simple scal-
ing up of the mean carbon burial rate equates to a
global carbon sequestration rate of 13.53 Gt year
-1
. The
same exercise for boreal, temperate and tropical terres-
trial forests extrapolates to global sequestration rates
of 451.1, 327.6, and 422.4 Gt year
-1
, respectively [43].
Sediment accretion rates (mm year
-1
)
Number of measurements
-20 -10 0 10 20 30 40 50
0
20
40
60
80
100
Mean = 4.996 mm year
-1
SD of mean = 7.402 mm year
-1
Confidence interval of mean = 1.241 mm year
-1
Median = 2.7 mm year
-1
Minimum = -11.0 mm year
-1
Maximum = 46.3 mm year
-1
Figure2. Sediment accretion rates measured in various mangrove forests
worldwide(n=139).
Data from
[1,24,28–31,33,35,38,48,51,55].
Key terms
REDD+: Acronym for Reducing
Emissions from Deforestation and
Forest Degradation. The + refers to the
additional steps of conservation and the
sustainable management of forests and
enhancement of forest carbon stocks.
Blue carbon: Term coined to refer to
steps designed to enhance the
acquisition and storage of carbon in
aquatic ecosystems, especially in coastal
habitats such as seagrass beds and
mangrove forests.
Carbon sequestration in mangrove forests Review
future science group
www.future-science.com
319
Thus, mangroves account for
approximately 3% of carbon seques-
tered by the worlds tropical forests,
although they account for <1% of
total area of tropical forests.
These data do, however, suggest
the potential for significant GHG
emissions if the high per-hectare
carbon stocks of mangroves are
disturbed. Losses of mangroves by
clearing, conversion to industrial
estates/aquaculture and changes in
drainage patterns lead to dramatic
changes in soil chemistry and usu-
ally result in rapid emission rates of
GHGs, especially CO
2
. For exam-
ple, deforesting mangroves that
grow on peat soils results in CO
2
emissions comparable to rates esti-
mated from collapse of terrestrial
peat soils [44 ]. Lovelock et al. mea-
sured CO
2
emissions from cleared
mangrove peat soils in Belize on
the order of 2900 tC km
-2
year
-1
[4 4]; this value compares well
with CO
2
emissions measured
from hurricane-damaged and
aquaculture-impacted mangroves
(1500–1750 tC km
-2
year
-1
), rain-
forests drained for agriculture (3200 tC km
-2
year
-1
)
and thawed Arctic tundra (150430 tC km
-2
year
-1
).
Donato et al. [21] calculated a plausible range of CO
2
emissions of 112392 tC released per hectare of man-
grove forest and soils cleared, which gives a global emis-
sions range of 0.020.12 PgC year
-1
, assuming current
deforestation rates (1–2% per year) and global area.
This range is equivalent to at least 2–10% of global
deforestation emissions (~1.2 PgC year
-1
[45]) and up to
50% of emissions from the worlds tropical peatlands
(0.24 PgC year
-1
[46]). These values are only indicative,
as large uncertainties remain, including the accuracy
of forest areas, temporal and spatial variations in fluxes
and standing stocks, local and regional differences in
the modes of disturbance, and variations in the depth
to which soil is dredged.
If the contribution of mangroves to global forest car-
bon sequestration is very small, their contribution to
carbon burial in the global coastal ocean is considerably
greater. Compared with other coastal ecosystems, man-
groves contribute an average of 14% to carbon seques-
tration in the worlds oceans, although accounting for
only 0.5% of total coastal ocean area (Table2).
Even considering the large uncertainties in these esti-
mates, the average burial rate of carbon in mangroves is
much greater than that from all other habitats, except
for salt marshes. Therefore, considering the data in
Figure1 and in Table2, mangrove forests have the high-
est area rates of carbon sequestration compared with
any other ecosystem, terrestrial or marine, contributing
disproportionately as a carbon sink.
Future perspective
Mangroves are currently being advanced as an essential
component of climate change strategies such as REDD+
and blue carbon. McLeod et al. [47] and Alongi [48] have
recently identified specific actions and issues that need
to be addressed in blue carbon projects:
Careful site selection, preferably at the seaward edge,
based on drivers thought to affect carbon sequestra-
tion rates, such as frequency of tidal inundation, pri-
mary productivity and rates of exchange with adjacent
ecosystems, as not all mangroves accumulate carbon;
Measure and map the spatial and temporal variations
in carbon stocks and burial rates, relating these fac-
tors to environmental and ecological drivers, possibly
determining a set of indicators that can be used to
quickly estimate changes in carbon stocks and
fluxes;
Number of measurements
Annual rates of carbon burial (g m
-2
year
-1
)
0 200 400 600 800 1000
0
5
10
15
20
25
30
Mean = 174.1 mm year
-1
SD of mean = 184.1 mm year
-1
Confidence interval of mean = 45.3 mm year
-1
Median = 114.5 mm year
-1
Minimum = 10.0 mm year
-1
Maximum = 920.0 mm year
-1
Figure3. Annual rates of carbon burial estimated in various mangrove forests
worldwide(n=66).
Data from
[1,9,15–18,21,25,28,30–32, 34,38 ,41,42,48,51,55–59].
Carbon Management (2012) 3(3)
future science group
320
Review Alongi
Remote sensing and aerial photography may be useful
to facilitate changes in restoration/rehabilitation
strategies, and in identifying changes in land use;
Standardization of methods used to measure biomass
and soil carbon stocks and rates of carbon burial;
The execution of any scheme must consider modeled
predictions of future climate changes, such as regional
predicted rises in sea level;
Planting of mixed species to maximize biodiversity,
food web connectivity and net ecosystem production;
Priority must be given to REDD+ schemes that give
priority to old-growth forests as mangrove carbon
stocks increase with stand age;
Studies should be conducted concurrently to assess
the conditions that determine whether or not climate
change impacts such as changes in sea.
Future climate scenarios for the ocean are subject to
large uncertainties, but regional changes in ocean cir-
culation, temperature, salinity and pH patterns, and in
sea level, must be considered as likely to have a strong
impact on the ability of mangroves to sequester carbon
[49]. Large uncertainties exist in our knowledge of carbon
sequestration in mangroves, and such limitations must be
factored into the blueprints of any payment for ecosystem
services, blue carbon or REDD+ schemes. Only then
will management of mangrove ecosystems be sustainable.
Financial & competing interests disclosure
The author has no relevant affiliations or financial involvement with
any organization or entity with a financial interest in or financial
conflict with the subject matter or materials discussed in the manu-
script. This includes employment, consultancies, honoraria, stock
ownership or options, expert t estimony, grants or patents received or
pending, or royalties. No writing assistance was utilized in the
production of this manuscript.
Executive summary
Carbon production
Mangrove net primary production averages 11.1t dry weightha
-1
year
-1
, roughly equivalent to tropical terrestrial forests.
Mangroves may constitute a carbon sink for up to a century.
Carbon allocation & storage
Belowground biomass is equivalent to aboveground biomass in mangroves.
Most carbon in mangroves is stored as large pools of soil carbon and belowground roots.
Storage of carbon in mangroves averages 937tCha
-1
.
Mechanisms facilitating sediment accretion
Mangroves actively facilitate accumulation of carbon and other elements associated to ne particles.
Rates of soil accretion & carbon sequestration
Rates of soil accretion in mangroves average 5mmyear
-1
.
Frequency if tidal inundation is the main factor controlling accretion.
Global carbon burial rates for mangroves approximate 24TgCyear
-1
.
Signicance of mangroves to terrestrial & marine carbon sequestration
Mangroves account for 3% of carbon sequestered by the world’s tropical forests, but 14% of carbon sequestered in the world’s ocean.
If disturbed, mangroves may emit 0.02–0.12 PgCyear
-1
, equal to 2–10% of global deforestation emissions.
Future perspective
Mangroves are prime candidates for REDD+ and blue carbon projects, but a number of issues and specic actions must be carefully
addressed prior to commencement of such projects.
Table2. Global contribution of mangroves and other coastal habitats to carbon sequestration in the global
coastal ocean.
Habitat Area (10
12
m
2
) Sequestration rate
(gCm
-2
year
-1
)
Global carbon sequestration
(Tgyear
-1
)
Mangroves 0.14 (0.5%) 174 24 (14%)
Salt marshes 0.22 (0.8%) 150 33 (20%)
Seagrasses 0.3 (1.1%) 54 16 (10%)
Estuaries 1.1 (4.0%) 45 50 (30%)
Shelves 26 (93.6%) 17 44 (26%)
Total 167
Assumes that depositional areas cover 10% of total shelf area [9].
Data from [41,6062].
Carbon sequestration in mangrove forests Review
future science group
www.future-science.com
321
References
1 Alongi DM. The Energetics of Mangrove
Forests. Springer, Amsterdam, The
Netherlands (2009).
2 Mazda Y, Wolanski E. Hydrodynamics and
modeling of water flow in mangrove areas. In:
Coastal Wetlands: An Integrated Ecosystem
Approach. Perillo GME, Wolanski E, Cahoon
DR, Brinson MM (Eds). Elsevier,
Amsterdam, The Netherlands, 231–262
(2009).
3 Ellison JC. Geomorphology and
sedimentology of mangroves. In: Coastal
Wetlands: An Integrated Ecosystem Approach.
Perillo GME, Wolanski E, Cahoon DR,
Brinson MM (Eds). Elsevier, Amsterdam,
The Netherlands, 565–591 (2009).
4 Berger U, Adams M, Grimm V et al.
Modelling secondary succession of subtropical
mangroves: causes and consequences of
growth reduction in pioneer species. Persp.
Plant Ecol. Evol. Syst. 7, 243–252 (2006).
5 Giri C, Ochieng E, Tiszen LL et al. Status
and distribution of mangrove forests of the
world using earth observation satellite data.
Global Ecol. Biogeogr. 20, 154–159 (2010).
6 Costanza RR, dArge R, deGroot R et al. The
value of the world’s ecosystem services and
natural capital. Ecol. Econ. 25, 3–15 (1998).
7 Mills RM. Capturing Carbon: The New
Weapons in the War against Climate Change.
Columbia University Press, NY, USA, 465
(2011).
8 Kristensen E, Bouillon S, Dittmar T et al.
Organic carbon dynamics in mangrove
ecosystems: a review. Aq. Bot. 89, 201–219
(2008).
9 Duarte CM, Middelburg JJ, Caraco N. Major
role of marine vegetation on the oceanic
carbon cycle. Biogeoscience 2, 1–8 (2005).
10 Fromard F, Puig C, Mougin E et al.
Structure, above-ground biomass and
dynamics of mangrove ecosystems: new data
from French Guiana. Oecologia 115, 39–53
(1998).
11 Clough BF, Ong JE, Gong WK. Estimating
leaf area index and photosynthetic production
in canopies of the mangrove Rhizophora
apiculata. Mar. Ecol. Prog. Ser. 159, 285–292
(1997).
12 Joye SB, Lee RY. Benthic microbial mats:
important sources of fixed nitrogen and
carbon to the Twin Cays, Belize ecosystem.
Atoll Res. Bull. 53, 1–24 (2004).
13 Perry DA, Oren R, Hart SC. Forest Ecosystems
(2nd Edition). Johns Hopkins University
Press, Baltimore, MD, USA, 596 (2008).
14 Lovelock CE. Soil respiration and
belowground carbon allocation in mangrove
forests. Ecosystems 11, 342–354 (2008).
15 Fujimoto K, Imaya A, Tabuchi R et al.
Belowground carbon storage of Micronesian
mangrove forests. Ecol. Res. 14, 409413
(1999).
16 Ren H, Chen H, Li Z et al. Biomass
accumulation and carbon storage of four
different aged Sonneratia apetala plantations
in southern China. Plant Soil 327, 279–291
(2010).
17 Nguyen HT, Yoneda R, Ninomiya I et al. The
effects of stand age and inundation on carbon
accumulation in mangrove plantation soil in
Namdinh, Northern Vietnam. Tropics 14,
21–37 (2004).
18 Ray R, Ganguly D, Chowdhury C et al.
Carbon sequestration and annual increase of
carbon stock in a mangrove forest. Atmos.
Environ. 45, 5016–5024 (2011).
19 Alongi DM, Clough BF, Dixon P et al.
Nutrient partitioning and storage in arid-zone
forests of the mangroves Rhizophora stylosa
and Avicennia marina. Trees 17, 5160 (2003).
20 Alongi DM, Wattayakorn G, Tirendi F et al.
Nutrient capital in different aged forests of the
mangrove Rhizophora apiculata. Bot. Mar. 47,
116–124 (2004).
21 Donato DC, Kauffman JB, Murdiyarso D
et al. Mangroves among the most carbon-rich
forests in the tropics. Nat. Geosci. 4, 293–297
(2011).
22 Saenger P. Mangrove Ecology, Silviculture and
Conservation. Kluwer, Dordrecht, The
Netherlands, 342 (2002).
23 McKee KL, Faulkner PL. Restoration of
biogeochemical function in mangrove forests.
Restor. Ecol. 8, 247–259 (2000).
24 Cahoon DR, Hensel P, Rybczyz J et al. Mass
tree mortality leads to mangrove peat collapse
at Bay Islands, Honduras after Hurricane
Mitch. J. Ecol. 91, 1093–1105 (2003).
25 McKee KL. Root proliferation in decaying
roots and old root channels: a nutrient
conservation mechanism in oligotrophic
mangrove forests? J. Ecol. 89, 876887
(2001).
26 Bouillon S, Rivera-Monroy VH, Twilley RR
et al. Mangroves. In: The Management of
Natural Coastal Carbon Sinks. Laffoley D,
Grimsditch G (Eds). International Union for
Conservation of Nature, Gland, Switzerland,
13–20 (2009).
27 Bouillon, S, Borges AV, Castaneda-Moya E
et al. Mangrove production and carbon sinks:
a revision of global budget estimates. Global
Biogeochem. Cycles 22, GB2013 (2008).
28 Furukawa K, Wolanski E. Sedimentation in
mangrove forests. Mangroves Salt Marshes 1,
3–10 (1996).
29 Bird MI, Fifield LK, Chua S et al. Calculating
sediment compaction for radiocarbon dating
of intertidal sediments. Radiocarbon 46,
421–435 (2004).
30 Alongi DM, Pfitzner J, Trott LA et al. Rapid
sediment accumulation and microbial
mineralization in forests of the mangrove
Kandelia candel in the Jiulongjiang Estuary,
China. Estuar. Coast. Shelf Sci. 63, 605618
(2005).
31 Mahmood H, Misri K, Sidik BJ et al.
Sediment accretion in a protected mangrove
forest of Kuala Selangor, Malaysia. Pakistan
J. Biol. Sci. 8, 149–151 (2005).
32 Ceron-Breton JG, Ceron-Breton RM,
Rangel-Marron M et al. Determination of
carbon sequestration rate in soil of a
mangrove forest in Campeche, Mexico.
WSEAS Trans. Environ. Develop. 7, 5564
(2011).
33 McKee KL. Biophysical controls on accretion
and elevation change in Caribbean mangrove
ecosystems. Estuar. Coast. Shelf Sci. 91,
475483 (2011).
34 Sanders CJ, Smoak JM, Naidu AS et al.
Organic carbon burial in a mangrove forest,
margin and intertidal mud flat. Estuar. Coast.
Shelf Sci. 90, 168–172 (2010).
35 Krauss KW, Cahoon DR, Allen JA et al.
Surface elevation change and susceptibility of
different mangrove zones to sea-level rise on
Pacific high islands of Micronesia. Ecosystems
13, 129–143 (2010).
36 Gonneea ME, Paytan A, Herrera-Silva JA.
Tracing organic matter sources and carbon
burial in mangrove sediments over the past
160 years. Estuar. Coast. Shelf Sci. 61,
211–227 (2004).
37 Alongi DM. Mangrove forests: resilience,
protection from tsunamis, and responses to
global climate change. Estuar. Coast. Shelf Sci.
76, 1–13 (2008).
38 Sanders CJ, Smoak JM, Naidu AS et al.
Mangrove forest sedimentation and its
reference to sea level rise, Cananeia, Brazil.
Environ. Earth Sci. 60, 1291–1301 (2010).
39 Twilley RR, Chen RH, Hargis T. Carbon
sinks in mangroves and their implications to
carbon budget of tropical coastal ecosystems.
Water Air Soil Pollut. 64, 265–288 (1992).
40 Jennerjahn TC, Ittekot V. Relevance of
mangroves for the production and deposition
of organic matter along tropical continental
margins. Naturwissenschaften 89, 23–30
(2002).
Carbon Management (2012) 3(3)
future science group
322
Review Alongi
41 Chmura GL, Anisfield SC, Cahoon DC et al.
Global carbon sequestration in tidal, saline
wetland soils. Global Biogeochem. Cycles 17,
1111 (2003).
42 Fujimoto K. Below-ground carbon
sequestration of mangrove forests in the
Asia–Pacific region. In: Mangrove
Management and Conservation: Present and
Future. Vannucci M (Ed.). United Nations
University Press, Tokyo, Japan, 138–146
(2004).
43 IPCC. Good Practice Guidance for Land Use,
Land-Use Change, and Forestry. Penman J,
Gytarsky M, Hiraishi T et al. (Eds). IPCC,
Kamiyamaguchi, Japan, 632 (2003).
44 Lovelock CE, Ruess RW, Feller IC. CO
2
efflux from cleared mangrove peat. PLoS ONE
6, e21279 (2011).
45 van der Werf GR, Morton DC, DeFries RS
et al. CO
2
emissions from forest loss. Nat.
Geosci. 2, 737–738 (2009).
46 Page SE, Rieley JO, Banks CJ. Global and
regional importance of the tropical peatland
carbon pool. Global Change Biol. 17, 798818
(2011).
47 McLeod E, Chmura GL, Bouillon S et al.
A blueprint for blue carbon: toward an
improved understanding of the role of
vegetated coastal habitats in sequestering
CO
2
. Front. Ecol. Environ. 9, 552–560 (2011).
48 Alongi DM. Carbon payments for mangrove
conservation: ecosystem constraints and
uncertainties of sequestration potential.
Environ. Sci. Policy 14, 462470 (2011).
49 Sen Gupta A, McNeil B. Variability and
change in the ocean. In: The Future of the
Worlds Climate. Henderson-Sellers A,
McGuffie K (Eds). Elsevier, MA, USA,
141–165 (2012).
50 Alongi DM, Trott LA, Rachmansyah et al.
Growth and development of mangrove forests
overlying smothered coral reefs, Sulawesi and
Sumatra, Indonesia. Mar. Ecol. Prog. Ser. 370,
97–109 (2008).
51 Alongi DM, Dixon P. Mangrove primary
production and above- and below-ground
biomass in Sawi Bay, southern Thailand.
Phuket Mar. Biol. Cent. Sp. Publ. 22, 31–38
(2000).
52 Alongi DM, Sasekumar A, Chong VC et al.
Sediment accumulation and organic material
flux in a managed mangrove ecosystem:
estimates of landoceanatmosphere
exchange in peninsular Malaysia. Mar. Geol.
208, 383402 (2004).
53 Alongi DM, Tirendi F, Trott LA et al.
Benthic decomposition rates and pathways in
plantations of the mangrove Rhizophora
apiculata in the Mekong delta, Vietnam. Mar.
Ecol. Prog. Ser. 194, 87–101 (2000).
54 Clough BF. Mangrove forest productivity and
biomass accumulation in Hinchinbrook
Channel, Australia. Mangroves Salt Marshes 2,
191–198 (1998).
55 Duarte CM, Kennedy H, Marba N et al.
Assessing the capacity of seagrass meadows for
carbon burial: current limitations and future
strategies. Ocean Coast. Manag. doi:10.1016/
j.ocecoaman.09.001 (2011) (In press).
56 Tateda Y, Nhan DD, Wattayakorn G et al.
Preliminary evaluation of organic carbon
sedimentation rates in Asian mangrove coastal
ecosystems estimated by
210
Pb chronology.
Radioprotection 40, S527–S532 (2005).
57 Xiaonan D, Xiaoke W, Lu F, Zhiyun O.
Primary evaluation of carbon sequestration
potential of wetlands in China. Acta Ecol.
Sinica 28, 463469 (2008).
58 Kauffman JB, Heider C, Cole TG et al.
Ecosystem carbon stocks of Micronesian
mangrove forests. Wetlands 31, 343–352
(2011).
59 Mitra A, Sengupta K, Banerjee K. Standing
biomass and carbon storage of above-ground
structures in dominant mangrove trees in the
Sunderbans. Forest Ecol. Manag. 261,
1325–1335 (2011).
60 Matsui N. Estimated stocks of organic carbon
in mangrove roots and sediments in
Hinchinbrook Channel, Australia. Mangroves
Salt Marshes 2, 199–204 (1998).
61 Kennedy H, Beggins J, Duarte CM et al.
Seagrass sediments as a global carbon sink:
isotopic constraints. Global Biogeochem. Cycles
24, GB4026, (2010).
62 Cai W-J. Estuarine and coastal ocean carbon
paradox: CO
2
sinks or sites of terrestrial
carbon incineration? Annu. Rev. Mar. Sci. 3,
123–145 (2011).
Websites
101 Data archives of the Australian Institute of
Marine Science, 1979–1999.
http://data.aims.gov.au/metadataviewer/
faces/search.xhtml
102 Nature Geoscience. Supplementary data.
www.nature.com/naturegeoscience
... According to evidences adduced from several studies, the high productivity combined with slow rates of decomposition in the soil significantly improves mangroves' capacity to capture and eventually store organic carbon, especially in the soils (Bouillon et al., 2008;Alongi, 2012;Suello et al., 2022). Estimates by Atwood et al. (2017) indicate that organic carbon stowed in mangrove sediments up to a depth of 1 m, globally equates to 2.6 billion Mg of C. Furthermore, above-ground net primary productivity reported for mangroves (8.1 t DW ha−1 yr−1) match the records from highly productive tropical forests on land (11.1 t DW ha−1 yr−1) (Alongi, 2012;Cooray et al., 2021). ...
... According to evidences adduced from several studies, the high productivity combined with slow rates of decomposition in the soil significantly improves mangroves' capacity to capture and eventually store organic carbon, especially in the soils (Bouillon et al., 2008;Alongi, 2012;Suello et al., 2022). Estimates by Atwood et al. (2017) indicate that organic carbon stowed in mangrove sediments up to a depth of 1 m, globally equates to 2.6 billion Mg of C. Furthermore, above-ground net primary productivity reported for mangroves (8.1 t DW ha−1 yr−1) match the records from highly productive tropical forests on land (11.1 t DW ha−1 yr−1) (Alongi, 2012;Cooray et al., 2021). Research on carbon stocks in the Kenya mangroves report an estimated range of 500-1000 t C ha-1 which is ten times higher than the average carbon content of terrestrial forests in the country (Huxham et al., 2015). ...
Article
Full-text available
Kenya is committed to the global efforts on climate change mitigation and adaptation as seen through investments in various sustainable green and blue economy projects. In this review paper, we present the current status of what has been done, particularly on the blue carbon offset initiatives undertaken in the mangrove and seaweed ecosystems as well as the decarbonization activities at the port of Mombasa and which should form reference information for local, regional, bilateral/multilateral partners, scientists and other climate change stakeholders. The blue carbon offset projects involve mangrove conservation, reforestation and carbon credit sale as well as seaweed farming. The initiatives have several unique features amongst which are the community-led income generation systems that simultaneously act as an inducement for ecosystem preservation, co-management and benefits sharing which are recipes for economic, socio-cultural, and environmental sustainability. A notable project impact is the conferment of economic power to the locals, particularly the women and the youth The model used embraces a collaborative approach involving multisectoral engagements of both the government, multilateral organizations, NGOs, and local communities. This integrated top-down (government) and bottom-up (local community) method deliberately targets the strengthening of economic development while ensuring sustainability.
... Courtesy: Alongi, (2012) ...
Chapter
Full-text available
The concentration of atmospheric CO2 is currently at around 415 ppm and is increasing by approximately 2-3 ppm annually. This rise is mainly caused by human activities, particularly the burning of fossil fuels for energy production, industrial processes, and changes in land use. Although annual emission rates have fluctuated, they generally remain high, with recent estimates indicating around 40-45 GtCO2 being emitted each year. This chapter examines the crucial involvement of terrestrial, aquatic, and marine ecosystems in carbon sequestration, a key process in contending climate change. By investigating various mechanisms such as terrestrial vegetation absorption of carbon dioxide through photosynthesis and the deposition of organic material in marine sediments, the chapter elucidates the intricate connections between these ecosystems and their impact on the global carbon cycle and regulating climate and stresses the vital requirement for cooperative efforts to ensure their ability to sequester carbon for the benefit of future generations
... Some current studies identify macroalgae, particularly kelp forest, as emerging blue carbon which may create a high mitigation impact and research attention on the roles of macroalgae as a blue carbon is growing amongst the scientific community (Kuwae et al., 2022;Howard et al., 2023;Pessarrodona et al., 2023). It is estimated that mangroves store an average of 937 tC.ha -1 with a carbon burial of about 174 gC m -2 year -1 (Alongi, 2012). Seagrass beds could store 19.9 Pg of organic carbon (Fourqurean et al., 2012). ...
Article
Full-text available
Coastal and marine ecosystems play a major role in the global carbon cycle. Connected marine and coastal ecosystems are commonly observed in the Western Gulf of Thailand. Little is known about the blue carbon potential of these interconnected ecosystems and seascapes. This study aims to quantify blue carbon stocks in the interconnected seagrass-coral reef-sandy coastal ecosystems at Samui Island, the Western Gulf of Thailand. At each study site, the samples of seagrasses, algae, and sediments, were collected from the different zones along a transect of interconnected sandy beach-seagrass bed-coral reef habitats, and the organic carbon contents were quantified using elemental analysis and loss on ignition (LOI). Our findings indicate that the habitats may provide a potential blue carbon opportunity. With a total area of 178.04 hectares (ha), consisting of sand (47.70 ha), seagrass beds (122.44 ha), macroalgal beds (2.40 ha), and live corals (5.50 ha), the estimated carbon storage was as much as 9,222.75 MgC; 74.03% of which was stored in sediment, while the rest was as biomass (25.97%). About 96 percent of the total carbon storage was found in seagrass beds (122.44 ha) with a total amount of carbon storage of 8,876.99 MgC, consisting of 8,781.01 MgC and 95.98 MgC of shallow- and deep-seagrass beds, respectively. The carbon stocks in seagrass, algal biomass, and sediment ranged from 1.58 - 19.10 MgC.ha⁻¹, 2.51 -10.45 MgC.ha⁻¹, and 0.93 - 58.46 MgC.ha⁻¹, respectively. Comparing the carbon storage at each study site, Ko Tan showed the highest value of carbon storage, accounting for 4,232.21 MgC, followed by Ao Phangka (2,901.83 MgC), Ao Thong Tanod (1,459.57 MgC) and Ko Mudsum (629.14 MgC). The quantities of carbon stocks varied considerably among microhabitats and the connectivity of these coastal and marine ecosystems may support the carbon stocks potential of the interconnected ecosystems. Ultimately, the findings from this study provide baseline data that supports Thailand’s nationally determined contribution and highlight the importance of interconnected coastal ecosystems in carbon sequestration and storage that should not be overlooked.
Technical Report
Full-text available
This study evaluates the capacity of forests, grasslands, and mangrove forests to achieve carbon neutrality in Kenya's coastal region, focusing on six counties: Kilifi, Kwale, Mombasa, Lamu, Taita Taveta, and Tana River. By analyzing case studies from these regions, the research examines the current state of carbon sequestration, identifies challenges, and proposes strategies to enhance carbon storage capabilities. The study employs a systematic approach, integrating global perspectives from China, India, and Brazil, regional insights from Comoros, Mozambique, Somalia, and Zanzibar, and East African examples from Tanzania and Uganda. The findings highlight the critical role of coastal ecosystems in mitigating climate change and underscore the need for integrated conservation and management policies. Background Information
Article
Full-text available
This study evaluates the capacity of forests, grasslands, and mangrove forests to achieve carbon neutrality in Kenya's coastal region, focusing on six counties: Kilifi, Kwale, Mombasa, Lamu, Taita Taveta, and Tana River. By analyzing case studies from these regions, the research examines the current state of carbon sequestration, identifies challenges, and proposes strategies to enhance carbon storage capabilities. The study employs a systematic approach, integrating global perspectives from China, India, and Brazil, regional insights from Comoros, Mozambique, Somalia, and Zanzibar, and East African examples from Tanzania and Uganda. The findings highlight the critical role of coastal ecosystems in mitigating climate change and underscore the need for integrated conservation and management policies. Background Information
Book
Full-text available
Program kesehatan hutan diarahkan untuk menurunkan laju populasi patogen sehingga dalam jangka panjang mengurangi ledakan populasi karena produktivitas hutan mangrove merupakan tuntutan yang harus diwujudkan sehingga kerusakan hutan harus mendapatkan prioritas dan perhatian utama. Oleh karenanya langkah antisipatif melalui upaya diagnosa dini perlu dilakukan sehingga data dan informasi yang diperoleh dapat dijadikan dasar untuk pengambilan kebijakan. Pulau-pulau kecil memiliki keindahan alam yang masih asli dan alami, serta memiliki berbagai potensi sumber daya alam, budaya, dan jasa lingkungan dapat dimanfaatkan untuk pengembangan ekowisata. Potensi alam di pulau-pulau kecil seperti sungai, hutan mangrove, keanekaragaman hayati baik flora dan fauna endemik, sampai dengan pemandangan sunset dan sunrise merupakan potensi ekowisata yang dapat dimanfaatkan untuk ditawarkan kepada wisatawan/pengunjung. Berbagai aktivitas ekowisata dapat dilakukan dengan memanfaatkan potensi tersebut seperti kegiatan susur sungai, bird watching, menikmati pemandangan, trekking, dan berkano. Lanskap dengan pemandangan yang didominasi oleh fitur alami, memiliki nilai scenic beauty estimation (SBE) yang tinggi karena memiliki karakteristik visual berupa lanskap yang alami, seperti fitur danau, sungai, pantai, hutan, pegunungan, perbukitan, perkebunan dan keragaman vegetasi yang tinggi. Setiap lokasi pulau kecil memiliki perbedaan dalam kondisi fisik wilayah, potensi sumber daya alam, dan permasalahan yang ada. Oleh karena itu, sebelum mengembangkan konsep pengembangan dan sistem pengelolaan ekowisata, perlu dilakukan tahap identifikasi kondisi fisik wilayah, potensi sumber daya alam dan jasa lingkungan, serta permasalahan yang ada. Pengembangan ekowisata di pulau-pulau kecil juga dapat mendorong pelestarian lingkungan dan pengembangan berkelanjutan. Konsep ekowisata tidak hanya memperkenalkan keindahan alam, tetapi juga bertujuan untuk melestarikan lingkungan alam dan mempromosikan praktik berkelanjutan dalam pengelolaan sumber daya.
Chapter
The concentration of atmospheric CO2 is currently at around 415 ppm and is increasing by approximately 2-3 ppm annually. This rise is mainly caused by human activities, particularly the burning of fossil fuels for energy production, industrial processes, and changes in land use. Although annual emission rates have fluctuated, they generally remain high, with recent estimates indicating around 40-45 GtCO2 being emitted each year. This chapter examines the crucial involvement of terrestrial, aquatic, and marine ecosystems in carbon sequestration, a key process in contending climate change. By investigating various mechanisms such as terrestrial vegetation absorption of carbon dioxide through photosynthesis and the deposition of organic material in marine sediments, the chapter elucidates the intricate connections between these ecosystems and their impact on the global carbon cycle and regulating climate and stresses the vital requirement for cooperative efforts to ensure their ability to sequester carbon for the benefit of future generations.
Article
Despite their importance in sustaining livelihoods for many people living along some of the world's most populous coastlines, tropical mangrove forests are disappearing at an alarming rate. Occupying a crucial place between land and sea, these tidal ecosystems provide a valuable ecological and economic resource as important nursery grounds and breeding sites for many organisms, and as a renewable source of wood and traditional foods and medicines. Perhaps most importantly, they are accumulation sites for sediment, contaminants, carbon and nutrients, and offer significant protection against coastal erosion. This book presents a functional overview of mangrove forest ecosystems; how they live and grow at the edge of tropical seas, how they play a critical role along most of the world's tropical coasts, and how their future might look in a world affected by climate change. Such a process-oriented approach is necessary in order to further understand the role of these dynamic forests in ecosystem function, and as a first step towards developing adequate strategies for their conservation and sustainable use and management. The book will provide a valuable resource for researchers in mangrove ecology as well as reference for resource managers.
Article
It was determined carbon sequestration rate, total nitrogen content and several important physicochemical parameters (electrical conductivity, gravimetric moisture, salinity, pH and soil texture) inmangrove forest soil located within the natural protected area named " Términos Lagoon" in Carmen Island, Campeche-México. Six sampling zones were considered within a mangrove forest located in the Botanical Garden of the Autonomous University of Carmen City. Samplings were carried out from February to August during 2009 considering three climatic periods (" Norths" season, dry season and rainy season). The electrical conductivity was within the range of 1.38 to 26.2 dS m-1 and the highest values were found during the rainy season when the study sites were flooded most of the time. The seasonal influence on carbon storage was evident (from 1.2 to 22.2 kg C / m-2), with the highest rate of carbon sequestration in the dry season, in flooded soils with greater predominance of red mangrove, and is lower in those soils rarely flooded with a higher prevalence of buttonwood mangrove individuals. The organic matter content and organic carbon was greater at30 cm depth for flooded areas, where long periods of flood tides and low rates of decomposition maintain anoxic conditions. Due to soils are sandy in the study areas and have high pH values with red mangrove associations, we can suggests that they have a high potential for carbon sequestration and it could be increased in future years.
Article
Accurate inventory of tropical peatland is important in order to (a) determine the magnitude of the carbon pool; (b) estimate the scale of transfers of peat-derived greenhouse gases to the atmosphere resulting from land use change; and (c) support carbon emissions reduction policies. We review available information on tropical peatland area and thickness and calculate peat volume and carbon content in order to determine their best estimates and ranges of variation. Our best estimate of tropical peatland area is 441 025 km 2 ($11% of global peatland area) of which 247 778 km 2 (56%) is in Southeast Asia. We estimate the volume of tropical peat to be 1758 Gm 3 ($ 18–25% of global peat volume) with 1359 Gm 3 in Southeast Asia (77% of all tropical peat). This new assessment reveals a larger tropical peatland carbon pool than previous estimates, with a best estimate of 88.6 Gt (range 81.7–91.9 Gt) equal to 15–19% of the global peat carbon pool. Of this, 68.5 Gt (77%) is in Southeast Asia, equal to 11–14% of global peat carbon. A single country, Indonesia, has the largest share of tropical peat carbon (57.4 Gt, 65%), followed by Malaysia (9.1 Gt, 10%). These data are used to provide revised estimates for Indonesian and Malaysian forest soil carbon pools of 77 and 15 Gt, respectively, and total forest carbon pools (biomass plus soil) of 97 and 19 Gt. Peat carbon contributes 60% to the total forest soil carbon pool in Malaysia and 74% in Indonesia. These results emphasize the prominent global and regional roles played by the tropical peat carbon pool and the importance of including this pool in national and regional assessments of terrestrial carbon stocks and the prediction of peat-derived greenhouse gas emissions.
Chapter
The ocean is host to a vast diversity of life; it is of immense social and economic value to us and it plays a crucial role in the climate system over synoptic to multi millennium timescales. Consequently, it is of great importance to understand changes that oceans have gone through in the past and how they might change in the future. In the context of anthropogenic climate change, the ocean acts as an important buffer. The majority of extra heat entering the climate system through enhanced greenhouse forcing, and almost half of the anthropogenic carbon dioxide (CO2) since preindustrial times have been sequestered and stored in the ocean. While mitigating against more rapid temperature rise in the atmosphere, the oceanic absorption of heat and CO2 has impacts on the ocean environment. The absorption of heat has caused thermal expansion of the ocean, which is a major contributor to global sea-level rise.