<<

WORKING PAPER

Carbon storage in and peatland A preliminary account from plots in

Daniel Murdiyarso Daniel Donato J. Boone Kauffman Sofyan Kurnianto Melanie Stidham Markku Kanninen

WORKING PAPER 48

Carbon storage in mangrove and peatland ecosystems A preliminary account from plots in Indonesia

Daniel Murdiyarso Daniel Donato J. Boone Kauffman Sofyan Kurnianto Melanie Stidham Markku Kanninen Working Paper 48

© 2009, Center for International Forestry Research. All rights reserved.

Printed in Indonesia Photos by Daniel Murdiyarso

CIFOR Jl. CIFOR, Situ Gede Bogor Barat 16115 Indonesia

T +62 (251) 8622-622 F +62 (251) 8622-100 E [email protected] www.cifor.cgiar.org

Any views expressed in this publication are those of the authors. They do not necessarily represent the views of CIFOR, the authors’ institutions or the financial sponsors of this publication.

Center for International Forestry Research CIFOR advances human wellbeing, environmental conservation and equity by conducting research to inform policies and practices that affect in developing countries. CIFOR is one of 15 centres within the Consultative Group on International Agricultural Research (CGIAR). CIFOR’s headquarters are in Bogor, Indonesia. It also has offices in Asia, Africa and South America. Table of contents

Acknowledgements 5

Summary 6

1. Introduction 7 Background 7 Objectives 8

2. Study sites 9 Bunaken National Park 9 National Park 11 Segara Anakan 12

3. Methods 14 Field sampling 14 Sample analysis and allometric computations 16

4. Above and belowground biomass and C-stocks 18 Mangrove ecosystems 18 Riverine : Tanjung Puting National Park 19

5. Diversity, structure and composition of riverine peat forests in Tanjung Puting National Park 25

6. Implications for mitigation and adaptation 28 Mangrove ecosystems 28 Peatland ecosystems 29

7. Conclusions 31

The way forward 32

References 33

Appendix 1. Summary of data collected by species 35 List of figures

1 Study sites in three provinces 9 2 Bunaken National Park, North Sulawesi 10 3 Tanjung Puting National Park, Central 12 4 Segara Anakan ecosystem, on the southern coast of Central 13 5 Field sampling design in a transect to estimate above and belowground carbon biomass 14 6 The distribution of aboveground carbon of living (mean ± SE) by 5 cm diameter classes in the three study areas 21 7 C-pools and total C-stock in sampled , by distance from the ocean 21 8 Depth (mean ± SE) of organic layers in mangroves by distance from the ocean 22 9 Above and belowground carbon pools in riverine peat swamps sampled in Tanjung Puting National Park 22 11 Ecosystem C-stocks (mean, SE) in riverine peat swamp forests of Tanjung Puting National Park, by distance from the river 23 10 Variation in ecosystem carbon storage in riverine peat swamps of Tanjung Puting National Park, according to peat depth 23 12 Distribution of riverine peat swamp trees by 5 cm diameter classes 24 14 Importance Value Index (IVI) of surveyed trees in riverine 25 13 Basal areas of trees surveyed in riverine peat swamp forest 25 15 Relationship between Importance Value Index and aboveground carbon content of trees in riverine peat swamp forest 26 Acknowledgements

We thank the many institutions and numerous individuals who made this work possible. The participation of various national stakeholders made this study possible. Contributions by staff and students of the Tropical Forests and Climate Change Adaptation project were highly valuable. The permit provided by the Ministry of Forestry, Republic of Indonesia (c/o Directorate General for Forest Protection and Nature Conservation) allows us to enter and work in Indonesia’s national parks and conservation areas. Practical support from staff of the Foundation International (OFI) and World Education (WE), represented by Mr Fajar Dewanto and Mr Genpur Sutoyo, respectively, is gratefully acknowledged. We also extend our sincere thanks to Dr Joko Pubopuspito of Sam Ratulangi University, Manado, who provided excellent assistance in terms of both logistical arrangements and technical advice. The authors are grateful to Andrew Wardell for helpful comments provided on drafts and to Imogen Badgery-Parker for her careful editing. Summary

Coastal mangrove forests provide a broad array of Our measurements show that total carbon storage in ecosystem services including fisheries production, mangrove ecosystems is exceptionally high compared sediment regulation, wood production and protection with most forest types, with a mean of 968 Mg C from storms and tsunamis. Similarly, peat swamp ha–1 and range of 863–1073 Mg C ha–1. These carbon forests harbour a diverse range of flora and fauna, stocks result from a combination of large-stature regulate water regimes and store large amounts of forest (trees up to ~2 m in diameter) and organic-rich carbon deposited in organic materials below the peat to a depth of 5 m or more. Aboveground ground. In Southeast Asia, the conversion rates of C-stocks vary widely depending on stand composition mangrove to other land uses, such as shrimp farms and history, but belowground pools comprise a and settlements, are among the highest for any forest large portion of ecosystem C storage in most sites. type. Furthermore, the conversion of peat swamp Although mangrove composition is often stratified forests to oil palm and pulp wood plantations and with distance from the ocean edge, C storage does not the associated fires have been the main sources of vary consistently along this gradient. greenhouse gas (GHG) emissions in the region during the past decade. Ecosystem C-pools at Tanjung Puting exceed 1000 Mg ha–1 in many of the sampled locations. All sampled With accounting for around 17% of stands had a depth to mineral soil of less than 1 m, global anthropogenic GHG emissions, the upcoming with a mean peat depth of 45.5 ± 6.8 cm. Mean total global mechanism known as Reducing Emissions C-stock was 894.3 Mg C ha–1, with a range of 558– from Deforestation and forest Degradation (REDD+) 1213 Mg C ha–1. It should be noted, when considering provides an important climate change mitigation these estimates of ecosystem pools, that peat depths of option. This scheme offers economic incentives for swamp forests may be as much as 20 m conserving forests and associated carbon (C) stores (with an average depth of 3–5 m). in developing countries. Mangrove and peat swamp ecosystems are well suited to such strategies. However, Projected rates of sea-level rise (~1 cm yr–1 over the although their high rates of C assimilation and export next century) are ~5–10 times higher than typical (fluxes) are known, their total C storage—the amount mangrove sediment accrual rates (1–2 mm yr–1), that may be emitted upon conversion—has not been suggesting high susceptibility and a potential positive well quantified. feedback via loss of C-stocks. Thus, the combination of very high C-stocks, susceptibility to land use We measured total ecosystem C storage (above and activities and numerous ecosystem services makes below ground) in mangrove ecosystems in North tropical mangroves ideal candidates for REDD+, Sulawesi, Central Kalimantan and Central Java, particularly if climate change mechanisms can be Indonesia. We assessed variations in mangrove applied to promoting synergies between adaptation to C-pools along transects running inland from the climate change (e.g. local migration) and mitigation. ocean edge, as well as their vulnerability to sea-level rise and land use. In addition, in Tanjung Puting However, additional studies to better quantify National Park, Central Kalimantan, we sampled both ecosystem C-pools and the potential impact of land the total aboveground biomass and the belowground cover change and fire are greatly needed in order peat horizons to ascertain total ecosystem C-pools. to make sound policy decisions related to carbon financing through the REDD+ mechanism. 1. Introduction

1.1 Background magnitude of energy absorption strongly depends Tropical mangroves and peat swamp forests provide on tree density, stem and root diameter, shore slope, numerous ecosystem services, including nutrient bathymetry, spectral characteristics of incident waves cycling, sediment trapping, protection from cyclones and tidal stage upon entering the forest. Alongi also and tsunamis, habitat for numerous organisms (many explores the role of mangroves in responding to economically important) and wood for lumber and various climate-related hazards such as increased fuel (Ellison 2008). Among the most important of precipitation, increased temperature and rise of these functions—but poorly quantified—is ecosystem . carbon (C) storage. The estimated carbon stored in these ecosystems is so large that it makes mangrove Southeast Asia’s tropical forests are experiencing the and peatland important for climate change mitigation. highest rate of deforestation worldwide (Langner and However, these econsystem are especially vulnerable Siegert 2009; Gibbs et al. 2010). Between 1980 and to climate and land use change. 2005, ~25% of the region’s mangroves were logged and converted to other land uses such as commercial The region comprising Southeast Asia and the aquaculture operations (FAO 2007). Lowland western Pacific Islands (the Indo-Pacific) is the global peat forests are also being logged and converted, epicentre of mangroves and tropical peat swamp experiencing still sharper declines (e.g. >56% loss forests (FAO 2007; Hirano et al. 2007). Approximately in protected areas from 1985 to 2001; Curran et al. 40% of the world’s mangroves, or 6 million ha, 2004), with some projections that these forests could occur in this region alone. Standing biomass per be entirely consumed within a decade (Jepson et al. area reaches higher values than in any other locale 2001). Because large quantities of carbon are released (Komiyama et al. 2008). In addition, Southeast Asia to the atmosphere when these forests are cleared and has 25 million ha of peatlands, which is 60% of all converted (Murdiyarso and Adiningsih 2007), this tropical peatlands. Because mangroves and peat deforestation has significant implications for global swamp forests have high root–shoot ratios and occur climate change. For example, owing largely to forest on organic-rich soils several metres deep (Fujimoto conversion, Indonesia has become the world’s third et al. 1999; Page et al. 2002; Komiyama et al. 2008), largest emitter of CO2 behind industrialised USA and total C storage in these ecosystems may be among China (IPCC 2007). Although some regulations are the largest forest C-pools on Earth. However, data in place to limit deforestation in Southeast Asia— quantifying these globally significant carbon stores succeeding in slowing the rate of decline slightly in are lacking. recent years—local socio-economic factors continue to lead to rapid forest losses (Kanninen et al. 2007). Mangrove forests also provide important coastal protection. Following the unprecedented Indian With deforestation accounting for almost 17% of Ocean tsunami on 26 December 2004, mangrove global carbon emissions, conserving existing forest forests were considered to have been the most has been suggested as one of the least expensive important coastal tree vegetation along the coast of options for mitigating human-induced climate Tamil Nadu, India; they had helped lessen damage change. Thus, Reduced Emissions from Deforestation caused by waves (Danielsen et al. 2005). In a more and forest Degradation (REDD+) in developing systematic study of mangroves, Alongi (2008) countries has emerged as a likely component of found that in some circumstances mangroves can the next international policy effort addressing offer limited protection from tsunamis, as the climate change, to be implemented when the Kyoto 8 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

Protocol expires in 2012 (Kanninen et al. 2007). This 1.2. Objectives programme would offer economic incentives for The primary objective of this study was to fill a conserving forests and associated C-stocks, intended critical information gap by working with local to offset the short-term economic factors that forestry professionals to quantify/monitor C-pools in promote deforestation. mangroves and peat swamp forests in Indonesia. This objective may be broadened to attract the attention of However, the viability of such a programme depends the global community while promoting locally wise heavily on having sound information on carbon use of these ecosystems by: storage in various forests, and how much C may be released when these forests are converted. According •• quantifying ecosystem C-pools of tropical to Kanninen et al. (2007; see also Angelsen 2008; mangrove and peat swamp forests in Indonesia Angelsen et al. 2009), the design and implementation that are highly vulnerable to climate change and of REDD+ strategies must be informed by high- land cover change; quality, independent research if they are to succeed. Research is vital to ensure that the inclusion of forests •• providing information on carbon dynamics in a future climate protection regime is efficient and to international processes such as the effective and reflects the interests of forest-dependent Intergovernmental Panel on Climate Change people in developing countries. (IPCC) and the UN Framework Convention on Climate Change (UNFCCC) on these relatively Because of their large carbon stores, Indonesian understudied forest types; mangroves and peat swamp forests would make •• assessing options for managing C-pools of ideal candidates for REDD+ strategies. However, tropical mangrove and peat swamp forests in information on whole-ecosystem carbon storage Indonesia; and is lacking for these forests. At the same time, a •• providing information necessary in the competing land use has already reached the policy development of REDD+ strategies for mitigating sphere: early in 2009, the Government of Indonesia climate change and adaptation strategies for through the Ministry of Agriculture enacted a future climate regimes. regulation allowing the development of oil palm plantations on peatlands that are less than 3 m deep. To achieve these objectives, a measurement campaign was carried out. This study was intended to establish This kind of policy-driven land use decision raises a framework for more extensive research sites to two fundamental questions: capture a broader (regional) context of the issues and challenges. •• How should scientific information on carbon dynamics in peatlands be used to inform policy? • How can sound environmental motives be balanced with sustainable development objectives, assuming that economic reasons for a decision are based on properly assessed opportunity costs? 2. Study sites

The studies were carried out in three contrasting Climate, geology and soils mangrove ecosystems, representing marine or oceanic The area experiences monsoonal wet and dry seasons mangroves in North Sulawesi (Bunaken National (November–April and May–October, respectively). Park), mangroves in Central Kalimantan The mean annual rainfall ranges between 2500 mm (Tanjung Puting National Park) and lagoon-associated and 3500 mm. The average air temperature is 27 °C mangroves in Central Java (Segara Anakan). In with minimum and maximum temperatures of 19 °C addition, a riverine peat swamp ecosystem in Tanjung and 34 °C, respectively. Puting National Park was studied (see Figure 1). The geological formations of the southern section 2.1 Bunaken National Park1 of the park, which is located on mainland Sulawesi, are dominated by Tondano tuff, characterised by Bunaken National Park is located in the province very porous volcanic rock. The northern section is of North Sulawesi, covering 79 056 ha of land and predominantly riverine sediments and alluvial soils, marine areas. The park consists of two sections: (1) especially on Mantehage Island and the eastern part of the northern section, which encompasses five islands Bunaken Island. The western part of Bunaken Island, (Bunaken, Silaken, Manado Tua, Mantehage and Siladen Island and parts of Nain Island are formed Nain) and the coastal area called the Molas-Wori from coral limestone. Manado Tua was formed at the Coast; and (2) the southern section, which spans the same time as the major volcanoes on the mainland. Arakan-Wawontulap Coast (see Figure 2).

Figure 1. Study sites in three provinces

1 Most of the description of Bunaken National Park is adapted from Mehta (1999). 10 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

Figure 2. Bunaken National Park, North Sulawesi. The park is divided into northern and southern sections.

Representative soil types for mangroves sampled in the coastal area of the park to drop directly down the the northern section were taken from Mantehage continental slope. The sea depth between the islands Island, which has a beach land system composed of of the park ranges between 200 m and 1840 m. mainly young alluvial soils. The island is distinguished by extensive mangrove forest flats, which are partially separated by salt water channels. The coastal area in Flora sampled portions of the southern section of the park The trees most commonly found on the shoreline is dominated by an intertidal land system. There is of the park are mangroves, locally known as bakau, also an extensive riverine sediment land system found which can tolerate saltwater or salty soil. Other in this area, which forms extensive mud-sand plains. common trees are bitung (Barringtonia asiatica), a This type of soil is the supporting factor for mangrove waxy-leafed tree with large, sweet-smelling white development in this section of the park. flowers; pandan (Pandanus sp.); and ketapang (Terminalia catappa), which local people grow as a The park has a unique bathymetry, which is one shade tree. There are also several species of beach attraction for divers. The absence of a continental grass and low creeping vines. shelf in the northern part of North Sulawesi allows Carbon storage in mangrove and peatland ecosystems 11

Tree sampling was carried out in the extensive Typical peat swamp habitat (40%–50%) includes mangrove forests found on Mantehage Island and species such as ramin ( bancanus), in the Molas-Wori and Arakan-Wawontulap coastal ulin (Eusideroxylon zwageri) and jelutong (Dyera areas. Mangrove species in most of the park tend to costulata). In disturbed areas, gelam (Melaleuca be short-stature oceanic mangroves. Although some cajuputi), Palaquium sp. and Alstonia scholaris have species can grow as tall as 40 m, most reach a height been found. of only ~10 m. Species found in the heath forests (up to 10% of Twenty-nine species of mangrove tree are found in the park) includes Dacrydium sp., Lithocarpus sp., Bunaken National Park. The dominant species are Eugenia sp., Castanopsis sp., Schima sp., Hoya sp., Rhizophora sp., Avicennia sp. and Sonneratia sp. Local sp. and Vatica sp. people use mangrove trees for their daily needs, such as for construction materials, structural support for The mangrove forests in the river deltas are seaweed farming, firewood, fish traps and traditional most commonly composed of Rhizophora sp., medicines. Much of the mangrove forest degradation Bruguiera sp., Sonneratia sp. and Xylocarpus in Mantehage is due to people cutting mangrove trees granatum. Further inland along the brackish river to sell as firewood in Manado. water, nipah (Nypa fruticans), pandan (Pandanus sp.) and bakung (Lilium spp.) are commonly found Transects were established in six locations and well developed. On the sandy seashores of the representative of the island and mainland mangrove peninsular (less than 5%), patches of Casuarina sp., formations in the northern and southern sections Barringtonia sp., Podocarpus sp. and Scaevola sp. of the park (see Figure 2). Most of these transects are found. sampled short-stature oceanic mangroves in a variety of different soil types and parent materials. The park is a world icon for orangutan (Pongo pygmaeus) rehabilitation and conservation efforts and is home to the world’s single largest population 2.2 Tanjung Puting National Park of orangutan (around 6000). It is also home to 220 Tanjung Puting National Park is located on the species of birds, 17 species of reptiles and 29 species southern tip of Central Kalimantan, bordered by of mammals, including long-nosed monkey (Nasalis Java Sea to the west and south, Sekonyer River to the larvatus), long-tailed macaque (Macaca fascicularis), north and Seruyan River to the east. It is between the agile gibbon (Hylobates agilis), grey gibbon (Hylobates districts of Kota Waringin Barat (whose capital city muelleri), red leaf monkey (Presbytis rubicunda) and is Pangkalan Bun) and Seruyan (whose capital city is Malayan sun bear (Helarctos malayanus). Kuala Pembuang). The two main rivers, Kumai and Seruyan, are very important for transportation and Oil palm expansion, initiated by the private sector the area’s economy (Figure 3). and local governments in the name of economic development, is challenging local forest conservation The park, which covers an area of around 400 000 efforts. It is understood that there has been no ha, lies on alluvial and mostly peat swamp plains, as open dialogue between these parties and the local indicated by the black -rich waters in most communities whose livelihoods depend on the forests river tributaries. The park is divided into dryland and the services they provide, including non-timber forest, peat swamp forests, heath forests, mangrove forest products, watershed protection and other goods forests and sandy beaches. and services.

In the dryland areas, which cover around 30% of Local NGOs led by Orangutan Foundation the park, dipterocarp species such as Shorea spp. International (OFI) have proposed to the Indonesian are commonly found. Also present are Myristica sp., Ministry of Forestry that the area between the eastern Lithocarpus sp., Castanopsis sp., Koompassia sp. and border of the park and Seruyan River be designated Scorodocarpus sp. as a buffer zone, and that plans to expand oil palm plantations in that area be cancelled. Should the 12 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

Figure 3. Tanjung Puting National Park, Central Kalimantan. Five transects each were established in mangrove and peatland ecosystems. central government approve the buffer zone, the park’s another five transects representative of peat swamp legal status would be strengthened, which would ecosystems (see Figure 3). All field measurements eventually generate economic benefits for the local were conducted with support from local NGOs community through ecotourism and sustain wildlife (Orangutan Foundation Indonesia and World sanctuaries. Part of the proposed area has been Education) and National Park Office field staff. developed as a REDD+ pilot and recently successfully completed third-party dual validation under the Voluntary Carbon Standard Guidelines (Rainforest 2.3 Segara Anakan ecosystem Alliance 2009; Bureau Veritas Certification 2010). The Segara Anakan ecosystem is the only lagoon- This is the world’s first new Avoided Planned Peatland mangrove complex in Java. Historically, the ecosystem Conversion Methodology, which is a significant was highly productive, supporting economically step towards protection of peatland ecosystems important coastal fisheries. However, due to high rates in Indonesia. of erosion upstream and heavy sedimentation, the lagoon has rapidly filled, with mud flats and swampy The field measurements were carried out in five lands now dominating the lagoon and causing a transects representative of mangrove forests and decline in productivity (Figure 4). Carbon storage in mangrove and peatland ecosystems 13

Mangrove destruction also is threatening the Segara and diversity are high in the eastern part of the Anakan fishery. Fisheries dependent on mangroves lagoon, located near the city of Cilacap. There, the could disappear within 5 to 10 years if destruction is dominant tree species are Aegiceras corniculatum, sustained at the high rate observed in 2000 (Dudley Nypa fruticans and Rhizophora apiculata, the latter 2000). Mangrove management is critical for the two of which are characteristic of mature forests. success and survival of the Segara Anakan fishery. By contrast, understorey species and the pioneer species (Avicennia alba, Aegiceras corniculatum and In the Segara Anakan lagoon, 21 tree species and 5 Sonneratia caseolaris) dominate the central lagoon understorey genera have been identified (Hinrichs where several rivers discharge. Tree communities in et al. 2009). Average tree density is 0.80 ± 0.99 the eastern lagoon are more stable and represent less individuals/m2 with 48.71% seedlings and an average disturbed mangrove forest. basal area of 9.86 ± 10.54 cm2/m2. Tree density

Figure 4. Segara Anakan ecosystem on the southern coast of Central Java. Sedimentation and mangrove degradation have caused a significant decline in lagoon fisheries. Two transects were established in one of the river deltas. 3. Methods

3.1 Field sampling Saplings In each sampled forest ecosystem, a transect arranged Woody stems with dbh <5 cm, known as saplings, perpendicular from the river or coast shoreline was were recorded in the same way as for trees, but with a established with no a priori knowledge of forest sampling radius of 2 m in each plot. composition or structure. Six plots were established along the transect at 25-m intervals (see Figure 5). Woody debris Measurements and collection of trees, saplings, woody debris, understorey/litter and soil were conducted as •• Woody debris is defined as follows. described below. –– Any dead woody material (twigs, branches or stems of trees or shrubs) that has fallen and Trees lies within 2 m of the ground. Leaning snags that form an angle of > 45° from true vertical The diameter at breast height (dbh = stem diameter at are also included. The piece must be in or 137 cm above the ground) of each tree was measured, above the litter layer to be included; it is not in 6 circular plots with a radius of 7 m in mangroves included if its central axis is buried in soil at and 10 m in peat swamps. Trees include all living the point of intersection. woody stems with dbh ≥5 cm, and any dead woody stem (snag or stump) with dbh ≥5 cm provided its –– Uprooted stumps and roots not encased angle from true vertical is less than 45°. Data on in soil. species, dbh, live/dead and height if dead or broken –– Dead branches and stems still attached to were recorded for all individuals. standing trees or shrubs are excluded.

Trees >5 cm dbh Wood debris transects Trees >5 cm dbh measured in 7 m radius (4 per plot, all plots) measured in 2 m radius (all plots) (all plots)

D A R = 2 m

7 m C B 12 m 3 soil depth Understory/litter sample measurements and 1 (2 per plot, all plots) nutrient core (all plots)

1 2 3 4 5 6 25 m 25 m Marine ecotone

Figure 5. Field sampling design in a transect to estimate above and belowground carbon biomass Carbon storage in mangrove and peatland ecosystems 15

Idul from Orangutan Foundation International measures diameter of breast height (DBH) as one parameter for estimating carbon stock.

•• The planar intersect technique (see Harmon and To determine mean down wood densities, a one-time Sexton 1996) involves counting intersections of collection was made of ~20–25 pieces of each woody woody pieces with a vertical sampling plane. debris class. Each piece was measured for volume using the water mass displacement method and oven- •• A survey tape is run 12 m from the plot centre in dried to obtain dry mass; these values were then used each of 4 directions, oriented at 45° angles from to compute particle density. the main transect line. Woody debris intersecting the transect plane is recorded to a height of 2 m above the forest floor. Understorey and litter •• CWD (coarse woody debris; pieces ≥ 7.6 cm in •• Understorey is defined as all standing vegetation diameter) is recorded for its actual piece diameter matter that does not reach breast height (130 cm). and decay status (sound or rotten) along the entire This includes all aboveground biomass of shrubs, 12-meter transect. FWD (fine woody debris; tree seedlings, herbs and non-vascular . It pieces < 7.6 cm in diameter) is sampled along does not include branches attached to saplings or subsections of each transect starting from the trees. Litter on the forest floor is defined as the distal end and working towards the plot centre. surface detritus and recognisable organic matter Hundred-hour fuels are sampled along 10 m of that lies on top of the mineral soil, excluding transect, 10-hour fuels along 5 m of transect and fragments of wood. 1-hour fuels along 2 m of transect. FWD is tallied •• Litter sampling consists of a 0.46 × 0.46 m square by size class as follows: (using a folding ruler), placed 2 m beyond each 1. 0–0.6 cm (1-hour fuels) plot centre along the site-transect line. Qualifying 2. 0.6–2.5 cm (10-hour fuels) material within this section was collected and placed in a collection bag for drying and weighing 3. 2.5–7.6 cm (100-hour fuels) in the laboratory. 16 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

•• As most mangroves have little to no litter, 3.2 Sample analysis and allometric contributing negligible biomass (Snedaker and computations Lahmann 1988), litter collection was restricted to Our general approach was to use well-established, peat swamp forests. relevant computational techniques from the literature to obtain the most accurate C-stock estimates possible Soil (e.g. Pearson et al. 2005, 2007). Where multiple, •• Mangroves and peat swamps often have an equally valid techniques were available, the more organic-rich soil layer for the first ~0.5–5 m of conservative option was selected in order to support soil depth, with a rather abrupt boundary to a generally conservative estimate of C-stock. Tree and mineral soils or sands underneath (especially in sapling data were entered into allometric equations to oceanic stand types). Three measurements were estimate biomass per individual. obtained to determine total soil organic carbon: organic soil depth (to obtain soil volume per For mangroves, we used allometric equations from area), bulk density (to obtain soil mass per area) Kauffman and Cole (in press), which are specific and % organic carbon (OC) (to convert mass per to the species/genera and size range (>100 cm area to C per area). dbh) encountered in this study. Other mangrove allometries generally derive from trees <50 cm dbh •• Soil depth was measured at three separate and thus lead to inflated biomass estimates for large locations around each plot centre. Depth was trees. Wood densities used in allometric computations measured using a soil auger and an aluminium came from species- and region-specific tables where probe. The probe was capable of measurements to possible (e.g. Oey 1951 cited in Soewarsono 1990; 3 m; it was inserted vertically until the sand layer Hidayat and Simpson 1994; Simpson 1996); genus was felt, and the depth then recorded. If the probe averages were applied otherwise. We used Clough did not encounter sand, the depth was recorded and Scott (1989) for mangrove leaf computations. as >300 cm and calculations were based on the All computations included the Baskerville (1972) top 300 cm of soil only (yielding a conservative correction for log-transformed equations. estimate of soil C storage).

•• A core sample was taken at each plot with a steel Tree coarse root mass was estimated using equations open-faced peat auger with a 1-m extension summarised in Komiyama et al. (2008). Fine roots handle. Five 4-cm subsamples were collected, <2 mm diameter, which were included in soil C representing depth intervals 0–15, 15–30, 30–50, estimates, were subtracted from the root mass 50–100 and 100–300 cm. These cut-offs are based numbers using fine/coarse root ratios summarised approximately on one of the few descriptions of in Cairns et al. (1997) and Chen et al. (2004). Stilt soil horizons in Indo-Pacific mangroves (USDA roots of Rhizophora spp. were included as part of 1983). Depth intervals in peat swamps were aboveground biomass rather than belowground, per more variable depending on the arrangement of Kauffman and Cole (in press). each plot’s peat and mineral horizons. Typically the subsample was taken from just above the For peat swamp forest trees and saplings, we used mid-point of each depth interval, but this rule the general aboveground biomass equations for wet was applied flexibly in order to obtain a good, tropical forests in Chave et al. (2005), belowground undisturbed sample. allometries in Cairns et al. (1997) and species-, genus- •• Coarse roots were avoided with the auger and by or site-specific wood densities. 2-mm sieving in the laboratory. Fine roots (<2 mm) passing through the sieve were considered Dead trees were assumed to have no leaf biomass part of the sample. and an average of 30% density loss above and below ground due to decay. Broken-stemmed dead •• Peat swamp soils were sampled using similar trees were conservatively modelled for volume as a methodology, but due to shallow , sampling modified cylinder of 0.8 πr2. and inference were restricted to the top metre below the soil surface, including both peat and mineral horizons. Carbon storage in mangrove and peatland ecosystems 17

Our field team collects organic material and soil samples using a specially designed peat auger.

Understorey vegetation mass per area was obtained Soil samples were dried to constant mass at 60 °C, by oven-drying the samples to constant mass and the final mass used to compute bulk density by and dividing by the sample area. Where samples dividing by sample volume. (We determined that bulk were large, subsamples were used to determine density estimates derived from drying mangrove soils moisture content. at 60 °C were within 1% of those derived at 105 °C.) Samples were then analysed at the Agricultural Woody debris mass per area was obtained by entering University of Bogor (IPB) analytical laboratory diameters or tallies into the planar intercept volume for OC content using the combustion method equation (see Harmon and Sexton 1996), and then (700 °C), after pretreatment to remove carbonates multiplying by wood density to convert into mass. All (see Schumacher 2002). Soil % OC content was then of the biomass estimates were converted to carbon multiplied by soil bulk density and soil depth to mass using a conservative conversion factor of 0.464 g obtain total soil C storage per unit area. C per 1 g of biomass. 4. Above and belowground biomass and C-stocks

4.1 Mangrove ecosystems trees. Lagoon-associated mangroves of Segara Anakan, Ecosystem C-stocks of sampled mangrove forests Central Java, contained the lowest total C-stocks (586.0 –1 ranged from 437 Mg C ha–1 to 2186 Mg C ha–1 Mg C ha ) due largely to small-stature aboveground (Table 1). The overall mean ecosystem C-stock across vegetation and lower concentrations the three study sites was 993.3 Mg C ha–1. This C (5.6% by mass compared with 9%–15% in other storage is exceptionally high compared with upland sites). Oceanic mangroves of Bunaken National Park, –1 tropical forests —which typically store between 150 North Sulawesi, were intermediate at 939.3 Mg C ha ; and 500 Mg C ha–1 (e.g. Murdiyarso et al. 2002)— and however, the southern section of this site contained the is perhaps second only to the renowned C-stocks of most carbon-dense stand of the study (Sondaken stand –1 peat swamp forests (e.g. Page et al. 2002). = 2186 Mg C ha ).

The river delta mangroves of Tanjung Puting National Dividing trees into diameter classes revealed that stands Park, Central Kalimantan, contained the largest mean in Segara Anakan lagoon had the lowest aboveground C-stocks of all sampled sites at 1220.2 Mg C ha–1, carbon (AGC) because of the absence of trees with having both the deepest soils and the largest-stature diameters >5 cm. The AGC of stems >35 cm in Tanjung

Sonneratia sp. grow along the northern coast of North Sulawesi. The height of these species indicate that most of this ecosystem’s carbon is stored below the ground. Carbon storage in mangrove and peatland ecosystems 19

Table 1. Ecosystem C-pools (Mg C ha–1) in mangroves, by study site (five transects in Tanjung Puting, two transects in Segara Anakan, and six transects in Bunaken)

Tanjung Puting National Park, Segara Anakan, Bunaken National Park,

Central Kalimantan Central Java North Sulawesi Mean (SE) % of total Mean (SE) % of total Mean (SE) % of total Aboveground pools Trees 124.3 (20.8) 10.2 7.5 (5.2) 1.3 61.4 (10.2) 6.5 (wood, foliage, prop roots) Down wood 18.6 (4.9) 1.5 4.3 (0.1) 0.1 42.2 (9.6) 4.5 Total above ground 142.9 (20.0) 11.7 11.8 (5.3) 2.0 103.6 (10.4) 11.0 Belowground pools Roots 18.1 (4.4) 1.5 2.6 (2.2) 0.4 13.7 (4.1) 1.5 Soil 1059.2 (84.5) 86.8 571.6 (141.6) 97.5 822.1 (256.5) 87.5 Total below ground 1077.3 (88.3) 88.3 574.2 (143.8) 98.0 835.7 (256.5) 89.0 Total ecosystem C-stock 1220.2 (92.8) 100.0 586.0 (149.1) 100.0 939.3 (256.7) 100.0

Puting National Park was larger than that in Bunaken less than 5% of the total biomass). This suggests that, National Park. Tanjung Puting National Park had the for accounting purposes, they may be negligible. highest AGC of the three sites (Figure 6).

Trees generally dominated aboveground C storage, 4.2 Riverine peat swamps: Tanjung but down wood was significant in some stands, most Puting National Park notably in Bunaken (Table 1). Although aboveground Ecosystem C storage in riverine peat swamp forests pools were up to 200 Mg C ha–1 at the stand scale, of Tanjung Puting National Park was high despite belowground pools comprised the majority of relatively shallow peat. All sampled stands had a ecosystem C-pools, accounting for 72%–99% of depth to mineral soil of less than 1 m, with a mean the total. peat depth of 45.5 ± 6.8 cm. Nevertheless, mean total C-stock was 894.3 Mg C ha–1, with a range of 558.1 to Soil depths, a major determinant of overall C storage, 1213.3 Mg C ha–1 (Figure 9). ranged from 35 cm to more than 3 m, which was the limit of our sampling tools and therefore our Aboveground pools were fairly consistent at ~200 Mg inference. Some 54% of all sample plots had depths C ha–1 and primarily composed of tree biomass, but >3 m. The deepest soils were in the deltaic deposits belowground pools were again dominant at 63%– of Tanjung Puting (all measured stands had >3 m 82% of total ecosystem C-stock (Figure 9). Because depth), and the shallowest were in Bunaken (mean aboveground C-pools were relatively constant, = 1.22 m); Segara Anakan was intermediate with a most variation in ecosystem C storage was due to mean depth of 2.11 m. variation in peat soil characteristics, especially peat depth (Figure 10). Ecosystem C-pools and soil depths did not vary consistently with distance from the ocean edge One reason that ecosystem C-stock was high despite (Figures 7 and 8), although there was a tendency for shallow peats is that peat bulk densities were higher larger total C-stocks in stands >50 m inland (1051 ± than those reported for most tropical peats, with a 73 Mg C ha–1) than in stands <50 m inland (878 ± 88 mean of 0.46 g cm–3. This may be due to mixing with Mg C ha–1). mineral soils associated with seasonal flooding in these riparian stands. Indeed, OC concentrations in The results clearly indicate that C-pools from dead the peat averaged 25.4%, which is substantially lower biomass and roots tend to be relatively small (typically than for most Southeast Asian peats (~40%–60% C), 20 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

One corner of peat swamp forest in Tanjung Puting National Park, Central Kalimantan, Indonesia

supporting the hypothesis of mixing with mineral aboveground nor belowground pools showed matter. The soil C concentrations of the horizons systematic changes along this gradient. It is uncertain below the peat layers were also relatively high. whether this trend may manifest over further distances inland; this likely depends on the presence There was no strong sign of variation in C-pools and nature of any peat dome. with distance from the river (Figure 11). Neither Carbon storage in mangrove and peatland ecosystems 21

30

25

20

15

10

5 boveground carbon (Mg C/ha) A

0 5 10 15 20 25 30 35 40 45 50 55 60 >60

-5 Diameter classes (cm)

Segara Anakan Bunaken Tanjung Puting

Figure 6. The distribution of aboveground carbon of living trees (mean ± SE) by 5 cm diameter classes in the three study areas. The I bars indicate standard error.

400 Total C stock ( x Mg ha-1): Aboveground pools ound 200 863.3 891.9 1044.4 1038.8 1073.4 1047.8 Trees eg r Down wood bo v ) A -1 0 a h Belowground pools M g ( 200 Roots

ck Soil st o

400 n o b r round a g 600 C lo w e B 800

1000

1200 0 20 40 60 80 100 120 140 Distance from ocean in meters

Figure 7. C-pools and total C-stock in sampled mangroves, by distance from the ocean 22 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

300

250

200

150

Organic soil depth (cm) Organic 100

50

0 0 35 60 85 110 135

Distance from ocean in meters

Figure 8. Depth (mean ± SE) of organic soil layers in mangroves by distance from the ocean

400 Aboveground pools Trees 200 Understory Down wood 0 )

-1 Belowground pools 200 Roots Peat 400 Mineral spoil (to 1 m depth)

600 Carbon stock (Mg ha stock Carbon 800

Total C stock 1000 558.1 743.8 950.3 1005.8 1213.3 1200 Risam Seluang Peramuan Jerumbun Simpang Kancil Site

Figure 9. Above and belowground carbon pools in riverine peat swamps sampled in Tanjung Puting National Park Carbon storage in mangrove and peatland ecosystems 23

1500 Freshwater swamp, 1250 ) -1 ha

g 1000 M (

750 storage C tem s 500 os y

c 2

E R = 0.94

250 P = 0.006

0 0 20 40 60 80 100

Peat depth (cm) Figure 10. Variation in ecosystem carbon storage in riverine peat swamps of Tanjung Puting National Park, according to peat depth

Aboveground pools 400 Total carbon stock ( x Mg ha-1): Trees

un d 1033.7 891.4 823.5 881.5 907.1 828.4 o r g 200 Understory Down wood bov e A ) -1

a 0 h Belowground pools (M g -200 Roots k c

o Peat s t -400 Mineral soil on (to 1 m depth) un d Ca r b o r

g -600 w o l e B -800

-1000 25 75 125 175 225 275

Distance from river (m)

Figure 11. Ecosystem C-stocks (mean, SE) in riverine peat swamp forests of Tanjung Puting National Park, by distance from the river 5. Diversity, structure and composition of riverine peat swamp forests in Tanjung Puting National Park

Across all 30 plots in Tanjung Puting National Park, we measured a total of 2390 trees, representing 55 species in 29 families. In term of species diversity, the Shannon–Wiener diversity index and Thompson’s diversity index were 3.29 and 0.94, respectively; as both these values are high compared with other swamp forest ecosystems, the diversity of the distribution of individuals across species can be described as relatively high in this area. The distribution of tree diameters revealed a typical reverse J-shaped pattern. Approximately 60% of trees have a dbh of less than 10 cm (Figure 12).

Ganua motleyana (Sapotaceae family) had the highest basal area (BA), 18.96 m2/ha, of all trees (Figure 13, Appendix 1). The total BA for all species except those with the 10 highest BAs was 10 m2/ha. Some species had a very low BA, close to zero, because they were represented by only a single individual with small dbh; among these were Diospyros buxifolia Hiern, Macaranga pruinosa and Elaeocarpus angustifolius Blume. Wood density of the trees surveyed ranged Rengas (Gluta wallichii), has a high importance value from 0.26 (Ficus drupacea of Moraceae family) to 1.05 index (IVI) and is frequently found in peat swamp g/cm3 (Parastemon urophyllus of Chrysobalanaceae ecosystems.

900

800

700

600

500

r of trees 400

umb e 300 N 200

100

0 5.0 10.0 15.0 20.0 25.0 30.0 35.0 40.0 45.0 50.0 55.0 60.0 More

Diameter classes (cm)

Figure 12. Distribution of riverine peat swamp forest trees by 5 cm diameter classes Carbon storage in mangrove and peatland ecosystems 25

family); the average across all trees was 0.65 g/cm3 represented by only a single individual, accounted (Appendix 1). for 0.2% of the IVI. Detailed IVI values sorted in descending order, followed by number of trees, basal The most important species in the surveyed area area, wood density and aboveground carbon of trees, was Ganua motleyana with an importance value are given in Appendix 1. index (IVI) of 57.8%, followed by Gluta wallichii, Pternandra azurea, Diospyros maingayi and Ixora The most important species, Ganua motleyana, had tenelliflora with IVIs of 31.6%, 19.3%, 12.2% and the highest AGC of 62.2 Mg/ha. Although Ixora 12%, respectively (Figure 14). Forty-five species tenelliflora has a lower IVI than Pternandra azurea, combined account for a total IVI of 123%. Diospyros its AGC is twice as high because of its higher BA and buxifolia, Macaranga pruinosa, Santiria laevigata, wood density. However, in general, increases in IVI and Elaeocarpus angustifolius, which were each correspond strongly with increases in AGC (Figure 15).

Other species

Symplocos adenophylla

Diospyros frutescens

Ixonanthes petiolaris

Parastemon urophyllus

Shorea leprosula

Diospyros maingayi

Pternandra azurea

Ixora tenelli ora

Gluta wallichii

Ganua mottleyana

0.00 2.00 4.00 6.00 8.00 10.00 12.00 14.00 16.00 18.00 20.00

Basal area (m2/ha) Figure 13. Basal areas of trees surveyed in riverine peat swamp forest. Individual values are given for the 10 species with the highest basal area; the remaining species are grouped in the top-most bar on the graph.

Other species Dialium indum L. Tetractomia tetranda Symplocos adenophylla Baccaurea macrophylla Diospyros frutescens Ixora tenelli ora Diospyros maingayi Ptemandra azurea Gluta wallichii Ganua mottleyana

0.0 20.0 40.0 60.0 80.0 100.0 120.0 140.0 Importance Value Index (%) Figure 14. Importance Value Index (IVI) of surveyed trees in riverine peat swamp forest. Individual values are given for the 10 species with the highest IVI; the remaining species are grouped in the top-most bar on the graph. 26 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

70,0 R = 0,912

60,0

50,0

40,0

30,0

20,0

Aboveground carbon (Mg C/ha) 10,0

0,0 0.0 10.0 20.0 30.0 40.0 50.0 60.0 70.0 -10,0 Importance Value Index (%)

Figure 15. Relationship between Importance Value Index and aboveground carbon content of trees in riverine peat swamp forest 6. Implications for climate change mitigation and adaptation

Rhizophora sp. grow to a low height in oceanic mangrove ecosystems in North Sulawesi. They protect the coastal zone from marine erosion.

Keeping mangroves and peat swamp forests intact mainstreamed into coastal development agendas as a has been a difficult challenge in Indonesia. Because strategy for adapting to climate change. developing and rehabilitating such ecosystems are not easy tasks, an incentive system to conserve these In terms of marketable ecosystem services, mangroves carbon-rich systems is urgently needed. There is a also filter chemical and organic pollution from the great need to raise awareness of the issue and at the water, which keeps the waters on reefs and seagrass same time to build partnerships among stakeholders. beds cleaner. There has been scant monetisation of such regulating services, but these could potentially finance conservation activities. 6.1 Mangrove ecosystems Mangroves provide a natural physical barrier against Other co-benefits of maintaining mangroves are , tsunamis and storm surges. The threat their function as a nursery for juvenile fish and of sea-level rise is quite prominent for small islands shrimp as well as habitat for crabs, oysters, clams, and low-lying coastal zones (IPCC 2007). Therefore, estuarine crocodiles and snakes. Seabirds and fruit conserving and restoring mangrove areas should be bats use mangroves as resting and breeding grounds. 28 Daniel Murdiyarso, Daniel Donato, J. Boone Kauffman, Sofyan Kurnianto, Melanie Stidham and Markku Kanninen

Converted peatlands near Tanjung Puting National Park for oil palm plantations.

Sustaining and restoring standing mangroves 6.2 Peatland ecosystems would continue the flow of these important Most forest conversions planned in the early 1980s ecosystem services. were located in peatlands (Murdiyarso and Lebel 2007). Deforestation of peat forests greatly increased Local communities also benefit from mangroves, greenhouse gas (GHG) emissions from Indonesia in which they use for building materials, food, fuel and the past decade. Contributing to the large land-based medicine. Mangrove ecosystems have an even bigger emissions were large-scale drainage of peatlands, role in aiding communities to adapt to multiple fires and vegetation clearing (Page et al. 2002; stressors associated with global environmental Hooijer et al. 2006). change. Denuded mangroves could be restored by involving communities whose livelihoods depend However, there is significant uncertainty associated on them. with emissions and other factors required to estimate system-wide emissions from peatlands. Such factors Mangroves appear to be one of the most carbon-dense will eventually be needed for carbon accounting types of tropical forest on Earth, possibly second only when benefits from conserving carbon are negotiated to peat swamp. Indonesia has more mangroves than in fund- and market-based mechanisms. Reducing any other country, and their land area adds another these uncertainties to develop reference levels and 3 Mha to the existing peat area (around 22 MHa) in to project future emissions/sequestrations will make Indonesia. Mangroves are an important deep-organic- carbon trading in peatlands more attractive. soil forest type that could benefit from REDD+ and should be considered among strategies for climate REDD+ policies have great potential to aid change mitigation. conservation of peatland ecosystems, which may be Carbon storage in mangrove and peatland ecosystems 29

especially effective candidates given their extremely used, which leads to inaccuracy. This makes carbon high carbon stores. Incentives should be created trading unattractive for both buyers and sellers. to encourage stakeholders to conserve carbon in the same landscapes in which they are entitled to Although this exercise was not intended to provide exploit timber. comprehensive information about peatlands and related emissions, it is does create better To date, the accounting methodology is lacking understanding of the need to prioritise the use credible numerical estimates of land conversions of stratified sampling based on peat depth and (activity factors) and GHG emissions and removals peat formation. (emission factors). Consequently, default values are 7. Conclusions

An adult orangutan with her child hangs on trees in Camp Leaky, Tanjung Puting National Park, Central Kalimantan, Indonesia.

Measurement campaigns of C-stocks in mangrove that even shallow-soiled peat swamps contain very and peatland ecosystems provide a basis for further large C-pools because of the combination of large developing field assessments and monitoring. The trees, denser peats and significant C storage in the data presented in this report show that mangroves upper mineral soil layers. Peatlands are extremely rich hold C-pools that are among the largest in the tropics. in belowground C-stocks, and it is not sufficient to We found no consistent or conclusive differences in keep the trees standing without maintaining natural terms of belowground C-stocks across sites and by inundation processes that prevent the oxidation of distance from the ocean. However, the ocean does organic soils. affect mangrove forest stature and hence aboveground C-stocks. Small-stature stands are more often found Integrating peatlands and mangrove conservation in oceanic (fringe) settings, while large-stature into REDD+ mechanisms would be a strategic stands are found in physically protected, riverine or approach to climate change mitigation in Indonesia, estuarine ecosystems. especially if ecosystem adaptation to global change can be mainstreamed into development agendas. The limited samples from peatlands prevent us from Many aspects of both mangroves and peatlands drawing any broader conclusions. Clearly, more make them unique ecosystems. Improving their studies across Indonesian peat forest and mangrove management, including wise use of resources, would ecosystems would be of great value. Nevertheless, enhance collateral benefits for both global and although our peatland sampling was limited to local communities. relatively shallow riverine forests, these data show The way forward

Wetlands are an important and often under- regionally and nationally appropriate default values appreciated resource. These ecosystems include lakes for the IPCC equation preclude so-called Tier 2 and rivers, swamps and , wet grasslands and accounting. This problem is particularly acute in peatlands, oases, , deltas and tidal flats, near- . shore marine areas, mangroves and coral reefs, as well as artificial sites such as fishponds, rice paddies, CIFOR and US Forest Services are developing a larger reservoirs and salt-pans. Carbon accounting has been project for submission to USAID with the overall goal particularly difficult in wetlands because of limited of supporting the development of the international information on C-stocks and on emissions and REDD+ mechanism in wetlands. The project will removals of other GHGs, particularly methane (CH4) significantly reduce the above problem throughout and nitrous oxide (N2O). the tropics through better characterisation of C-stocks in intact tropical wetlands and in the ecosystems that In the IPCC (2003) Good Practice Guidance for Land are replacing them. Use, Land Use Change and Forestry, wetlands were treated in an annex because of limited information The proposed project will produce maps for the available at that time. The 2006 revised National tropics over four years. It will adopt a regional Greenhouse Gas Inventory Guidelines laid out for approach, refining the methods at each stage and the first time methods for GHG accounting in these updating them with new developments in remote ecosystems. However, these remain incomplete. sensing technology. The project will also develop As more countries take on emission reduction innovative statistical approaches to large-scale commitments and as the international REDD+ assessments of carbon stocks. mechanism becomes operational, it is important that we improve our knowledge of the carbon stores of The resources provided by USAID will be used to these ecosystems, notably in terms of biomass and soil fund the work, initially in Southeast Asia. During organic matter pools. Year 1, we will undertake the detailed conceptual work that will set the stage for the entire project. Using the methodologies and findings described This will involve a major international workshop in in this paper, CIFOR and US Forest Services early 2011. We will establish local partnerships, and plans to undertake measurement activities with a all field measurements in each target country will be broader scope and coverage in order to provide new carried out through local partners with supervision knowledge for international and national policy by CIFOR and the US Forest Service (USFS). This development, early demonstration of REDD+ will contribute to building local capacity to undertake activities and the monitoring, reporting and carbon assessments in ecosystems. As verification of emission reductions. These have been additional funding becomes available, we will move the major technical barriers to the implementation progressively to other regions. of the upcoming global REDD+ scheme. Lack of References

Alongi, D.M. 2008 Mangrove forests: Resilience, Danielsen, F., Sørensen, M.K., Olwig, M.F., Selvam, protection from tsunamis, and responses to V., Parish, F., Burgess, N.D., Hiraishi, T., global climate change. Estuarine, Coastal and Karunagaran, V.M., Rasmussen, M.S., Hansen, Shelf Science 76: 1–13 L.B., et al. 2005 The Asian tsunami: a protective Angelsen, A. (ed.) 2008 Moving ahead with REDD: role for coastal vegetation. Science 310: 643. issues, options and implications. CIFOR, Bogor, Dudley, R. 2000 Segara Anakan fisheries management Indonesia. 156p. plan. Segara Anakan Conservation and Angelsen, A. with Brockhaus, M., Kanninen, M., Sills, Development Project. Asian Development E., Sunderlin, W.D. and Wertz-Kanounnikoff, Bank. 33p. S. (eds.) 2009 Realising REDD+: national Ellison, A.M. 2008 Mangrove ecology–applications in strategy and policy options. CIFOR, Bogor, forestry and coastal zone management. Aquatic Indonesia. 362p. Botany 89: 77. Baskerville, G. 1972 Use of logarithmic regression FAO (Food and Agriculture Organization of the in the estimation of biomass. Canadian United Nations). 2007. The world’s mangroves Journal of Forest Research 2: 49–53. 1980–2005. FAO Forestry Report 153, Rome. Bureau Veritas Certification. Final assessment Fujimoto, K., Imaya, A., Tabuchi, R., Kuramoto, S., report, version 5.2, March 2010 (http://www. Utsugi, H. and Murosushi, T. 1999 Belowground us.bureauveritas.com/wps/wcm/connect/bv_ carbon storage of Micronesian mangrove forests. usnew/Local/Footer/Home/ (10 October 2010). Ecological Research 14: 409–413. Cairns, M.A., Brown, S., Helmer, E.H. and Gibbs, H.K., Ruesch, A.S., Clayton, M.K., Holmgren, Baumgardner, G.H. 1997 Root biomass allocation P., Ramankutty, N. and Foley, J.A. 2010 in the world’s upland forests. Oecologia Tropical forests were the primary sources 111: 1–11. of new agricultural land in the 1980s and Chave, J., Andalo, C., Brown, S., Cairns, M.A., 1990s. PNAS 107: 16732–16737. doi: 10.1073/ Chambers, J.Q., Eamus, D., Fölster, H., Fromard, pnas.0910275107. F., Higuchi, N., Kira, T., et al. 2005 Tree allometry Harmon, M.E. and Sexton, J. 1996 Guidelines for and improved estimation of carbon stocks and measurements of woody detritus in forest balance in tropical forests. Oecologia 145: 87–99. ecosystems. Publication no. 20. United States Chen, W., Zhang, Q., Cihlar, J., Bauhus, J. and Price, Long Term Ecological Research Network Office, D.T. 2004 Estimating fine-root biomass and University of Washington, Seattle. production of boreal and cool temperate forests Hidayat, S. and Simpson, W.T. 1994 Use of green using aboveground measurements: a new moisture content and basic specific gravity to approach. Plant and Soil 265: 31–46. group tropical woods for kiln drying. Research Clough, B.F. and Scott, K. 1989 Allometric Note FPL-RN-0263, USDA Forest Service, Forest relationships for estimating aboveground biomass Products Laboratory. in six mangrove species. Forest Ecology and Hinrichs, S., Nordhaus, I. and Geist, S.J. 2009 Status, Management 27: 117–127. diversity and distribution patterns of mangrove Curran, L.M., Trigg, S.N., McDonald, A.K., Astiani, vegetation in the Segara Anakan lagoon, Java, D., Hardiono, Y.M., Siregar, P., Caniago, I. Indonesia. Regional Environmental Change 9: and Kasischke, E. 2004 Lowland forest loss in 275–289. doi: 10.1007/s10113-008-0074-4 protected areas of Indonesian Borneo. Science Hirano, T., Segah, H., Harada, T., Limin, S., June, T., 303: 1000–1003. Hirata, R. and Osaki, M. 2007 Dahdouh-Guebas, F. and Koedam, N. 2006 Coastal balance of a tropical peat swamp forest in vegetation and the Asian tsunami. Science Kalimantan, Indonesia. Global Change Biology 311: 37–38. 13: 412–425. Carbon storage in mangrove and peatland ecosystems 33

Hooijer, A., Silvius, M., Wösten, H. and Page, S. 2006 Murdiyarso, D. and Lebel, L. 2007 Southeast Asian PEAT-CO2, Assessment of CO2 emissions from fire regimes and land development policy. In: drained peatlands in SE Asia. Delft Hydraulics Canadell, J., Pataki, D. and Pitelka, L. (eds.). Report Q3943. Terrestrial ecosystems in a changing world, IPCC (Intergovernmental Panel on Climate Change). 261–271. Springer, New York. 2003. IPCC Good Practice Guidance for Land Murdiyarso, D., van Noordwijk, M., Wasrin, U.R., Use, Land-Use Change and Forestry Prepared Tomich, T.P. and Gillison, A. 2002 Environmental by the National Greenhouse Gas Inventories benefits and sustainable land-use in Jambi Programme, eds Penman J Gytarsky M Hiraishi transect, , Indonesia. Journal of T, et al. IGES, Hayama. Vegetation Science 13: 429–438. IPCC (Intergovernmental Panel on Climate Change). Oey, D.S. 1951. In: Soewarsono, P.H. 1990 Specific 2006. IPCC Guidelines for National Greenhouse gravity of Indonesian woods and its significance Gas Inventories, Prepared by the National for practical use. FRPDC, Forestry Dept, Bogor, Greenhouse Gas Inventories Programme, eds Indonesia. 42p. Eggleston HS, Buendia L, Miwa K et al. IGES, Page, S.E., Siegert, F., Rieley, J.O., Boehm, H.-D.V., Hayama. Jaya, A. and Limin, S. 2002 The amount of carbon IPCC (Intergovernmental Panel on Climate Change). released from peat and forest fires in Indonesia 2007 Fourth IPCC assessment report: climate during 1997. Nature 420: 61–65. change 2007. Cambridge University Press, Pearson, T., Walker, S. and Brown, S. 2005 Cambridge. Sourcebook for land use, land-use change Jepson, P., Jarvie, J.K., MacKinnon, K. and Monk, K.A. and forestry projects. Report from BioCF and 2001 The end for Indonesia’s lowland forests? Winrock International. Available at: http://www. Science 292: 859–861. winrock.org/ecosystems/tools.asp?BU=9086. Kanninen, M., Murdiyarso, D., Seymour, F., Pearson, T.R.H., Brown, S.L. and Birdsey, R.A. 2007 Angelsen, A., Wunder, S. and German, L. 2007 Measurement guidelines for the sequestration of Do trees grow on money? The implications of forest carbon. General Technical Report-NRS-18, deforestation research for policies to promote USDA Forest Service, Northern Research Station. REDD. CIFOR, Bogor, Indonesia. Rainforest Alliance 2009 Baseline and monitoring Kauffman, J.B. and Cole, T.G. 2010 Effects of a methodology for conservation projects that avoid severe typhoon on Micronesian mangrove forest planned land use conversion in peat swamp structure. Wetlands. In Press. forests. First assessment report, 7 December Komiyama, A., Ong, J.E. and Poungparn, S. 2008 http://www.rainforest-alliance.org/climate. Allometry, biomass, and productivity of cfm?id=peat_swamp (5 October 2010). mangrove forests: A review. Aquatic Botany 89: Schumacher, B. 2002 Methods for the determination 128–137. of total organic carbon (TOC) in soils and Langner, A. and Siegert, F. 2009 Spatiotemporal sediments. NCEA-C-1282, EMASC-001, US fire occurrence in Borneo over a period of 10 Environmental Protection Agency, Las Vegas, years. Global Change Biology 15: 48–62. doi: NV, USA. 10.1111/j.1365-2486.2008.01828.x Simpson, W.T. 1996 Method to estimate dry-kiln Mehta, A. 1999 Bunaken National Park Natural schedules and species groupings: tropical and History. Yayasan Kelola. USAID/NRM/EPIQ temperate hardwoods. RPL-RP-548, USDA Forest Program. Service, Forest Products Laboratory. Murdiyarso, D. and Adiningsih, E.S. 2007 Climate Snedaker, S.C. and Lahmann, E.J. 1988 Mangrove anomalies, Indonesian vegetation fires and understorey absence: a consequence of evolution? terrestrial carbon emissions. Mitigation and Journal of Tropical Ecology 4: 311–314. Adaptation Strategies for Global Change 12: USDA (United States Department of Agriculture). 101–112. 1983 Soil survey of island of Kosrae, Federated States of Micronesia. Soil Conservation Service, Hawaii, USA. Appendix 1. Summary of data collected by tree species

Number of trees, relative density (RD), relative frequency (RF), relative dominance (RDom), importance value index (IVI), basal area (BA), wood density (WD) and aboveground carbon (AGC) of the species in surveyed plots. Means No Family Species Local Name N RD (%) RF (%) RDom (%) IVI (%) BA WD AGC (m2/ha) (g/cm3) (Mg C/ha) 1 Sapotaceae Ganua motleyana Pierre ex Dubard Ketiau 429 17.3 4.4 36.2 57.8 18.96 0.55 62.2 2 Anacardiaceae Gluta wallichii (Hook.f.) Ding Hou Rengas 227 9.1 4.4 18.0 31.6 9.45 0.67 36.6 3 Melastomataceae Pternandra azurea (DC.) Burkill Pesulan 242 9.7 4.4 5.2 19.3 3.07 0.71 9.3 4 Diospyros maingayi Laminaduk 105 4.2 3.9 4.1 12.2 2.16 0.70 7.7 5 Rubiaceae Ixora tenelliflora Merr. Bentuka 60 2.4 3.2 6.3 12.0 3.31 0.97 20.3 6 Ebenaceae Diospyros frutescens Blume Bekunyit 110 4.4 3.8 2.3 10.5 1.27 0.70 4.3 7 Euphorbiaceae Baccaurea macrophylla Muell. Arg. Puak 83 3.3 4.1 1.8 9.2 0.98 0.65 2.6 8 Symplocaceae Symplocos adenophylla Wall. Habu - Habu Rawa 72 2.9 3.6 2.3 8.9 1.20 0.78 4.2 9 Rutaceae Tetractomia spp. (Roxb.) Merr. Jampang 83 3.3 2.8 1.9 8.0 0.99 0.76 3.3 10 Fabaceae Dialium indum L. Keranji 44 1.8 3.8 1.9 7.4 0.99 0.90 4.4 11 Chrysobalanaceae Parastemon urophyllus A.C. DC. Bentan 36 1.4 2.2 3.1 6.7 1.62 1.05 10.8 12 Fabaceae Crudia bantamensis (Hassk.) Benth. Mampai 63 2.5 3.1 1.1 6.7 0.60 0.93 2.4 13 Shorea leprosula Miq. Lanan 27 1.1 2.3 3.3 6.7 1.72 0.52 5.7 14 Ixonanthaceae Ixonanthes petiolaris Blume Lebang 38 1.5 2.6 2.5 6.7 1.30 0.82 6.2 15 Icacinaceae Stemonurus malaccensis (Mast.) Sleumer Bedaru rawa 56 2.3 3.5 0.7 6.5 0.43 0.65 1.0 16 Anacardiaceae Campnosperma auriculata Hook. f. Tentalang 51 2.1 3.1 0.7 5.8 0.50 0.38 0.7 17 Clusiaceae Garcinia beccarii Pierre Kemanjing 43 1.7 3.1 0.4 5.2 0.27 0.75 0.7 18 Stenochlaena palustris Lembiding 95 3.8 1.2 0.0 5.0 0.03 0.65 0.0 19 Euphorbiaceae Baccaurea javanica Muell. Arg. Pintau 53 2.1 2.3 0.4 4.9 0.37 0.64 0.7 20 Melastomataceae Memecylon myrsinoides Blume Tembaras 41 1.7 2.6 0.3 4.6 0.18 0.88 0.5 21 Rubiaceae Ixora fluminalis Ridl. Sariwaka 21 0.8 2.8 0.3 4.0 0.19 0.97 0.7 22 Clusiaceae Calophyllum soulattri Burm.f. ex F. Muell Penaga 34 1.4 1.8 0.8 3.9 0.42 0.54 1.2 23 Lauraceae Endiandra rubescens Blume ex Miq. Kumpang Pasir 35 1.4 2.2 0.3 3.9 0.16 0.75 0.4 24 Annonaceae Polyalthia lateriflora King Banitan kuning 27 1.1 1.9 0.7 3.6 0.44 0.60 1.1 25 Ctenolophonaceae Ctenolophon parvifolius Oliv Templaan 25 1.0 1.6 0.9 3.5 0.45 0.89 1.9 26 Pandanaceae Pandanus motleyanus Solms. Rasan pipit 45 1.8 1.5 0.0 3.3 0.16 0.65 0.2 Carbon storage in mangrove and peatland ecosystems 35 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.1 0.0 0.0 0.9 0.0 0.0 0.0 0.0 0.0 0.2 0.0 2.4 0.5 0.6 0.2 0.6 1.4 0.7 0.2 0.1 1.1 AGC 199.4 (Mg C/ha) ) 3 0.50 0.60 0.39 0.70 0.92 0.72 0.73 0.26 0.52 0.63 0.53 0.65 0.65 0.53 0.64 0.72 0.72 0.35 0.47 0.65 0.45 0.65 0.50 WD 0.52 0.65 0.45 0.70 0.50 0.65 0.65 Means (g/cm /ha) 2 0.00 0.03 0.00 0.00 0.01 0.01 0.00 0.01 0.5 0.00 0.01 0.4 0.0 0.01 0.01 0.01 0.04 0.10 0.02 2.2 0.26 0.21 0.09 0.30 0.55 0.18 0.14 0.06 0.31 BA 56.70 (m 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.2 0.4 0.4 0.4 0.6 0.6 0.6 0.6 0.6 0.6 0.7 1.7 2.0 2.1 2.3 2.4 2.6 2.7 2.9 3.1 3.2 300 IVI (%) 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 0.0 0.1 0.2 0.0 0.0 0.4 0.4 0.2 0.2 1.1 0.4 0.2 0.0 0.6 100 RDom (%) 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.1 0.3 0.3 0.3 0.4 0.4 0.4 0.4 0.3 0.3 0.4 0.6 1.0 1.2 1.3 1.5 1.0 1.8 1.5 1.8 1.9 100 RF (%) 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.1 0.1 0.1 0.1 0.1 0.1 0.2 0.3 0.2 0.2 1.1 0.6 0.5 0.8 0.7 0.5 0.5 1.2 1.3 0.8 100.0 RD (%) 1 1 1 1 2 1 1 1 1 2 3 3 3 3 3 4 7 4 5 27 14 15 20 18 12 13 30 34 19 N 2,390 Pinang tikus Pinang Poga Mahang Ribu Ramania Ubar merah Ular ular Keraya Kumpang arang Kumpang Mata keli Piler warig Piler Rasak Renda Lupis Gembor Kemanjar Ubar Penemplaan Penyarang bukit Penyarang Pinang kara Pinang Jungkung Jangkang Besira Kumpang putih Kumpang Papung Bekapas Medang Putat Ramin Local Name Local Total Mean Elaeocarpus angustifolius Blume Elaeocarpus angustifolius Santiria laevigata Blume Santiria laevigata Macaranga pruinosa Muell. Arg. Arg. Macaranga pruinosa Muell. Diospyros buxifolia Hiern buxifolia Diospyros Bouea oppositifolia (Roxb.) Meisn. (Roxb.) Bouea oppositifolia Syzygium pycnanthum Merr. & L.M. Perry Merr. Syzygium pycnanthum Jack racemosa Fagraea Ficus Ficus Thunb. drupacea Gynotroches axillaris Blume Gynotroches Chisocheton patens Blume patens Chisocheton

Timonius flavescens Baker flavescens Timonius Alseodaphne canescens Boerl. Boerl. canescens Alseodaphne Aporosa lucida (Miq.) AiryAporosa Shaw Syzygium sp. Syzygium sp. Kokoona ochracea Goniothalamus tapis Miq.

Platea excelsa Blume excelsa Platea Xylopia malayana Hook.f. & Thomson Thomson & Hook.f. malayana Xylopia Ilex malaccensis Loes. Loes. Ilex malaccensis Knema cinerea (POiL) Warb Warb Knema (POiL) cinerea Allophyllus cobbe (L.) Raeusch. cobbe Allophyllus Aporosa penangensis (Ridl.) AiryAporosa Shaw Actinodaphne procera Barringtonia lanceolata (Ridl.) Payens (Ridl.) Payens Barringtonia lanceolata Stemonurus grandifolius Becc. Becc. grandifolius Stemonurus Species Elaeocarpaceae Elaeocarpaceae Burseraceae Burseraceae Euphorbiaceae Euphorbiaceae Ebenaceae Ebenaceae Anacardiaceae Anacardiaceae Myrtaceae Loganiaceae Moraceae Moraceae

Rhizophoraceae Rhizophoraceae Meliaceae

Rubiaceae Rubiaceae Lauraceae Lauraceae Euphorbiaceae Euphorbiaceae Myrtaceae

Annonaceae Annonaceae

Icacinaceae Icacinaceae Annonaceae Annonaceae Aquifoliaceae Aquifoliaceae Myristicaceae Sapindaceae Sapindaceae Euphorbiaceae Euphorbiaceae

Lecythidaceae Lecythidaceae Icacinaceae Icacinaceae Family 55 54 53 52 51 49 50 48 47 46 45 44 43 42 41 40 39 38 37 36 35 34 33 32 31 30 29 28 27 No www.cifor.cgiar.org www.ForestsClimateChange.org

Center for International Forestry Research CIFOR advances human wellbeing, environmental conservation and equity by conducting research to inform policies and practices that affect forests in developing countries. CIFOR is one of 15 centres within the Consultative Group on International Agricultural Research (CGIAR). CIFOR’s headquarters are in Bogor, Indonesia. It also has offices in Asia, Africa and South America.