You are on page 1of 17

Remote Sensing of Environment 183 (2016) 65–81

Contents lists available at ScienceDirect

Remote Sensing of Environment

journal homepage: www.elsevier.com/locate/rse

Mapping spatial distribution and biomass of coastal wetland vegetation


in Indonesian Papua by combining active and passive remotely
sensed data
Aslan Aslan a, Abdullah F. Rahman b,⁎, Matthew W. Warren c, Scott M. Robeson d
a
School of Public and Environmental Affairs, Indiana University, Bloomington, IN 47405, USA
b
Coastal Studies Lab, University of Texas Rio Grande Valley, San Padre Island, TX 78597, USA
c
Northern Research Station, USDA Forest Service, Durham, NH 03824, USA
d
Department of Geography, Indiana University, Bloomington, IN 47405, USA

a r t i c l e i n f o a b s t r a c t

Article history: There is ongoing interest to develop remote sensing methods for mapping and monitoring the spatial distribution
Received 30 July 2015 and biomass of mangroves. In this study, we develop a suite of methods to evaluate the combination of Landsat-8,
Received in revised form 13 April 2016 ALOS PALSAR, and SRTM data for mapping spatial distribution of mangrove composition, canopy height, and
Accepted 30 April 2016
aboveground biomass in the wide intertidal zones and coastal plains of Mimika district, Papua, Indonesia.
Available online xxxx
Image segmentation followed by visual interpretation of composite PALSAR images was used to delineate man-
Keywords:
grove areas, whereas a flexible statistical rule based classification of spectral signatures from Landsat-8 images
Indonesia was used to classify mangrove associations. The overall accuracy of land cover classification was 94.38% with a
Mangroves kappa coefficient of 0.94 when validated with field inventory data and Google Earth images. Mangrove height
Classification and aboveground biomass were mapped using the SRTM DEM, which were calibrated with field-measured
Biomass data via quantile regression models. There was a strong correlation between the SRTM DEM and the 0.98 quantile
Landsat-8 of field canopy heights (H.98), which was used to represent the tallest trees in each of 196 10 m radius subplots
Radar (r = 0.84 and R2 = 0.804). Model performance was evaluated through 10,000 bootstrapped simulations, produc-
ing a mean absolute error (MAE) of 3.0 m for canopy height estimation over 30 m pixels of SRTM data. Quantile
regression revealed a relatively strong non-linear relationship between the SRTM derived canopy height model
and aboveground biomass measured in 0.5 ha mangrove inventory plots (n = 33, R2 = 0.46). The model results
produced estimates of mean standing biomass of 237.52 ± 98.2 Mg/ha in short canopy (Avicennia/Sonneratia)
stands to 353.52 ± 98.43 Mg/ha in mature tall canopy (Rhizophora) dominated forest. The model estimates of
mangrove biomass were within 90% confidence intervals of area-weighted biomass derived from field measure-
ments. When validated at the landscape scale, the difference between modeled and measured aboveground
mangrove biomass was 3.48% with MAE of 105.75 Mg/ha. These results indicate that the approaches developed
here are reliable for mapping and monitoring mangrove composition, height, and biomass over large areas of
Indonesia.
© 2016 Elsevier Inc. All rights reserved.

1. Introduction et al., 2010). Mangrove diversity and occurrence are highest in South-
east Asia, where nearly 75% of the World's mangrove species occur.
Mangroves grow exclusively within the intertidal zones of coastal, Indonesia alone contains about 19% of the World's mangrove forest
estuarine, and riverine landforms of the tropics and subtropics (Giri area, a total of 3.2 million hectares, and 45 of about 70 “true” mangrove
et al., 2011). The ecological functions of mangroves are critical to the en- species (species which meet physiological criteria and are adapted to
vironmental health of near-shore marine environments and adjacent occupy saline, hypoxic intertidal habitats; see Bakosurtanal, 2009;
upstream terrestrial systems, as they are transitional ecosystems FAO, 2007; Tomlinson, 1986). Approximately half of Indonesia's man-
where ocean, land, and freshwater converge (Kathiresan & Bingham, groves are located in the Papua and West Papua provinces and remain
2001; Lugo & Snedaker, 1974; Suratman, 2008; Tomlinson, 1986). largely intact, whereas mangroves in other parts of Indonesia have ex-
They provide at least US $1.6 billion each year globally in ecosystem ser- perienced very high deforestation rates in recent decades due to aqua-
vices while supporting the livelihoods of coastal communities (Polidoro culture and infrastructure development (Ilman, Wibisono, &
Suryadiputra, 2011; Rahman, Dragoni, Didan, Barreto-Munoz, &
Hutabarat, 2013). A recent assessment suggests Indonesia has lost up
⁎ Corresponding author. to 40% of its original mangrove cover since 1980. Most of these losses

http://dx.doi.org/10.1016/j.rse.2016.04.026
0034-4257/© 2016 Elsevier Inc. All rights reserved.
66 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

have occurred in Sumatra, Kalimantan, Sulawesi and Java (Ilman et al., Kauffman, & Verchot, 2013). Mangroves in the Indo-Pacific region are
2011; Murdiyarso et al., 2015). reported to contain carbon stocks exceeding 1000 MgC/ha, mostly
The southern coast of Indonesian Papua contains some of the largest stored in buried sediments (Donato et al., 2011; Fujimoto et al., 1999).
contiguous mangrove forests on Earth, including large tracts on Bintuni Per unit area, this capacity to store carbon is several times higher than
Bay and on the wide coastal plains of Mimika and Asmat districts that of tropical forests on dry land (Donato et al., 2011). Therefore, con-
(Fig. 1). An estimated 193,226 ha of mangrove forest cover the coastal servation and improved management of the remaining mangroves pro-
lowlands of Mimika, a southern district of Papua Province (MoF, vides an avenue to avoid large-scale CO2 emissions and loss of
Ministry of Forestry of Indonesia., 2008). These mangroves are mostly sequestered C from mangrove conversion. Such Blue Carbon mitigation
pristine and of high ecological value as they support productive near- strategies could contribute substantially to Indonesia's voluntary pledge
shore fisheries, harbor unique biodiversity, and are critical sources of to reduce greenhouse gas emissions 26% by 2020 (GOI, Government of
food and other natural products to local communities (ARD, 2014). Indonesia, 2011). However, involving mangrove conservation in climate
However, despite the importance of Mimika's mangrove forests and mitigation programs such as Reducing Emissions from Deforestation
coastal wetlands in providing ecosystem services (Barbier, 2012; and Forest Degradation (REDD +) and Blue Carbon (McLeod et al.,
Clough & Abdullah, 1993; Costanza et al., 1997; Mumby et al., 2004; 2011; Pendleton et al., 2012) requires intensive mapping and monitor-
Vo, Kuenzer, Vo, Moder, & Oppelt, 2012), they are vulnerable to the ing to create baseline C stock data and to validate conservation efforts.
rapid large-scale conversion and mismanaged deforestation observed Such mapping is also needed for improved spatial planning and inte-
elsewhere in Indonesia (Margono, Potapov, Turubanova, Stolle, & grated coastal management.
Hansen, 2014; Miettinen, Shi, &Liew, 2011; Rahman et al., 2013). As The goal of this study was to develop a systematic approach for map-
coastal resources of other Indonesian Islands become depleted, it is ping the distribution and biomass of different species and associations
likely that Papuan mangroves will become more attractive to devel- of mangroves at a consistent pixel scale by combined use of multispec-
opers intending to expand aquaculture production and other industries. tral and Synthetic Aperture Radar (SAR) remote sensing data. Specific
In addition to the well-known ecological benefits they provide, man- objectives include:
groves are gaining increasing attention for their capacity to sequester
carbon (they are a key component of “Blue Carbon”) and potential in- 1) To investigate whether improved mapping of mangroves at genus
volvement in climate change mitigation efforts (Alongi, 2012; Chmura, or even species level may be achieved by using images from the
Anisfeld, Cahoon, & Lynch, 2003; Donato et al., 2011; Murdiyarso, recently launched (February 2013) Landsat-8, which has 16-bit

Fig. 1. Map of the study area in the Mimika district of Papua, Indonesia, showing the location of field sampling plots and random validation points bounded between 0 and 150 m ASL
elevation.
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 67

radiometric resolution (65,536 grey levels). To address this objec- Use of radar data in mangrove mapping is challenging because of the
tive, we develop a Flexible Statistical Expert Based (FSEB) system complexity of the backscatter signal received from mangroves (Lucas
and apply an image segmentation technique, rather than using tra- et al., 2007). Different bands of radar backscatter are affected differently
ditional unsupervised or supervised classification methods; and by the interactions between the transmitted signal and biophysical
properties of mangrove plants, such as size, geometry, orientation of
2) To evaluate the effectiveness of using a combination of SAR and
leaves, trunks, branches, different types of roots, and the moisture con-
Landsat-8 data for retrieving a reliable estimation of mangrove can-
tent of both plants and underlying soil (Kuenzer, Bluemel, Gebhardt,
opy heights and its standing biomass.
Quoc, & Dech, 2011; Lucas et al., 2007). For example, the longer wave-
length L-band (wavelengths of 15–30 cm) has a greater capacity of pen-
Following this introduction section the paper is structured as fol-
etrating the foliage and small branches of the upper canopies of forests
lows. Section 2 describes a short literature review of past research.
including woody trunk and the underlying ground surfaces (Lucas,
The materials and methods are presented in Section 3, which provides
Moghaddam, & Cronin, 2004), whereas the shorter wavelength C-
a brief description of study area, field and remote sensing data (SAR
band (wavelengths of 3.75–7.5 cm) not only has a limited penetration
and Landsat-8) used in this study, as well as approaches used to process
through the canopy of forest, but is also heavily influenced by scattering
the field and remote sensing data. Section 4 demonstrates the imple-
caused by the canopy layer (Lang & Kasischke, 2008). SAR data have
mentation of the analysis methods developed in Section 3 and its corre-
been used for mapping mangrove cover, structure, and biomass with
sponding discussion, whereas Section 5 presents the conclusions
some success (Fatoyinbo & Simard, 2013; Held et al., 2003; Lucas
derived from this study.
et al., 2009; Mougin et al., 1999; Proisy, Mougin, Fromard, & Karam,
2000; Proisy, Mougin, Fromard, Trichon, & Karam, 2002; Simard et al.,
2. Literature review 2006, 2008). Several recent studies combined elevation data from Shut-
tle Radar Topographic Mission (SRTM) with spaceborne lidar data from
Remote sensing provides a fast, cost-effective, and efficient method the ICESat/GLAS satellite to estimate canopy height and biomass of
of mapping the spatial distribution and biomass of forests, and monitor- mangroves, producing promising results (Fatoyinbo, Simard,
ing their disturbance dynamics (Held, Ticehurst, Lymburner, & Washington-Allen, & Shugart, 2008; Fatoyinbo & Simard, 2013; Simard
Williams, 2003). It is particularly useful for mangrove monitoring et al., 2006, 2008).
since many mangrove forests are located in remote locations accessible A few studies have also explored the potential of combining SAR and
only by boat, where field measurements are difficult, time consuming, optical remotely sensed data for mapping mangrove ecosystems
and expensive. Producing accurate maps of mangrove distribution and (Aschbacher et al., 1995; Held et al., 2003; Kushwaha, Dwivedi, & Rao,
biomass requires image classification procedures that can discern varia- 2000; Nascimento Jr. et al., 2013; Ramsey III et al., 1998). Lucas et al.
tions in mangrove forest structure, which often depends on age, canopy (2007) have shown the potential of using L-band SAR data for mapping
height, and species associations. mangrove extent and zonation in Australia, New Guinea, and Malaysia
A few studies have demonstrated the potential of hyperspectral and where differences in structure occurred between zones as a function
high spatial resolution optical remote sensing data for detailed mapping of species, growth stage, and biomass distributions. A recent study by
of mangrove forest composition (Green, Mumby, et al., 1998a; Held Flores De Santiago, Kovacs, and Lafrance (2013) has successfully sepa-
et al., 2003; Wang, Sousa, & Gong, 2004; Neukermans, Dahdouh- rated mangrove areas from saltpan and water/shallow zones with an ac-
Guebas, Kairo, & Koedam, 2008). Green, Mumby, et al. (1998a) have curacy of 91.1% for single polarization and 92.3% for dual polarization of
successfully classified the coastal forests of the Turks and Caicos Islands PALSAR images, respectively. Cloud-free multispectral data offer the
into six mangrove and three non-mangrove classes, with overall accu- ability to characterize canopy coverage including different shades of
racy of 78.2%, using hyperspectral data from the Compact Airborne leaf greenness and composition of mangrove species in a given area,
Spectrographic Imager (CASI) sensor. Another study used high spatial whereas SAR data provide images that are sensitive to mangrove forest
resolution (0.7 m) QuickBird pan-sharpened imagery to map man- structure including canopy height.
groves in Gazy Bay, Kenya, into four dominant species with overall accu-
racy of 72% (Neukermans et al., 2008). Yet, one major problem with the 3. Materials and methods
use of hyperspectral or high spatial resolution imagery is that unlike
Landsat or SPOT data, these datasets are not commonly available for 3.1. Study area
the vast majority of mangrove areas, and when available, are cost pro-
hibitive for monitoring and mapping mangroves changes at regional This research was carried out within the Mimika district of
to national scales. Indonesia's Papua province, located between 4° and 5°20′ S and
Several studies have classified mangroves into broad classes using 135°40′ to 137°45′ E, in the south-central coast of Papua. As this study
medium spatial resolution multispectral optical sensors such as SPOT focused on mangroves and associated coastal wetland vegetation, we
(Système Pour l'Observation de la Terre) (Gao, 1998; Jensen et al., limited the study area to 0–150 m above sea level (ASL), extending
1991; Rasolofoharinoro et al., 1998), and Landsat Thematic Mapper from the coastline landward. Based on global mangrove data, our as-
(TM) (Brondízio, Moran, Mausel, & Wu, 1996; Green, Clark, et al., sumption is that mangroves do not occur at elevations greater than
1998b; Long & Skewes, 1996). These studies have indicated that de- the 150 m threshold (Fig.1). The Mimika coastal plains are characterized
tailed mangrove characterization is generally cumbersome at such by broad swampy areas extending from the coast in a northeast direc-
moderate spatial resolution. Other studies have shown that vegetation tion with mountains rising abruptly from the edge of the plains. The
indices from multispectral sensors can be linked to mangrove leaf area swamps are intersected into narrow catchments by a number of
index (LAI) and canopy heights (Ramsey & Jensen, 1996). According north-south parallel rivers flowing from the mountains to the sea.
to Green, Clark, et al. (1998b), two probable reasons for the limited suc- Tide records from Puriri Island (Fig.1) indicate diurnal to semi-diurnal
cess in differentiating among many classes of mangroves using moder- patterns with a tidal range between 3.3 and 3.6 m (Ellison, 2005). The
ate resolution multispectral satellite data may have been related to: wet and dry seasons are not clearly distinguishable, but the southeast
1) sensor's resolution, particularly the relatively low 8-bit (256 grey monsoons typically increase rainfall from June to mid-September
levels) radiometric resolution of those satellite sensors (Al-Wassai & (Pouwer, 1970). Small settlements of the indigenous Kamoro and
Kalyankar, 2013; Masek et al., 2001); and 2) the ability of the image pro- semi-nomadic clans of Asmat ethnic groups inhabit the coastal wet-
cessing method (i.e., classification procedure) to discriminate between lands; however, traditional lifestyles have little direct impact on the
various mangrove types. coastal wetland forests (Marshall, 2007). The rich natural and cultural
68 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

resources of coastal Mimika are currently facing increasing threats due 3.2.2. Multispectral data
to rapid population growth and in-migration, a booming mining indus- Four Landsat-8 Level 1 terrain-corrected (L1T) scenes were used in
try, and rapid forest conversion to oil palm plantations in a recently this study (path/row 102/64 captured on 30 March 2013, path/row
granted 38,000 ha concession to the west of Timika (Fig. 1). 102/63 capture on 30 March 2013, path/row 103/63 capture on 1
April 2014, and path/row 104/63 capture on 23 May 2013). The images
3.2. Data acquisition and analysis were downloaded from the US Geological Survey National Center for
Earth Resources Observation and Science through the EarthExplorer
3.2.1. Field data data portal (http://earthexplorer.usgs.gov/). These images have been
During July–December of 2013, two major field campaigns were geocoded in the Universal Transverse Mercator (UTM) projection sys-
completed, including forest inventory measurements of 41 transect tem zone 53 S and WGS84 datum. The digital number (DN) of each
plots of 0.5 ha each and 25 field verification plots of 0.09 ha each band was converted to reflectance value using the ATCOR tool in
(Fig. 2). Each of the 41 transect plots contained six nested circular sub- ERDAS Imagine®. For image analysis, we used the bands 1 to 7 from
plots (10 m radius), resulting in 246 subplots. For these 41 plots, we the operational land imager (OLI) sensor of Landsat-8. As mentioned
followed the sample plot design and procedure used in the Indo- earlier, Landsat-8 data has 16-bit radiometric resolution, which may
Pacific Forest Carbon Study and current Sustainable Wetlands Adaption have the potential to identify spectrally distinct species/genus of man-
and Mitigation Program (SWAMP; www.cifor.org/swamp) described in groves. In addition, Landsat-8 also has a low geolocation uncertainty,
detail by Kauffman and Donato (2012). At each of the field plots, tree b12 m circular error (Irons, Dwyer, & Barsi, 2012).
height (H) was measured using hypsometer and estimated to the
nearest meter, and stem diameter (cm) was measured at breast height 3.2.3. SAR data
(DBH) or relevant point of measurement. Trees N 50 cm in diameter Two types of SAR data were used in this study: 26 scenes of Phased
were sampled in 125 m by 40 m rectangular plots, whereas trees Array L-band Synthetic Aperture Radar (PALSAR) images from the Ad-
N5 cm were sampled in six 10 m radius subplots arranged every 25 m vanced Land Observing Satellite (ALOS) platform and four scenes of dig-
along the 125 m center line of the plot (Fig. 2). We identified trees to ital elevation model (DEM) images from the C-band SAR onboard the
genus or species level, measured height and diameter, and recorded Shuttle Radar Topography Mission (SRTM) spacecraft. PALSAR data
the coordinates of subplot centers using a handheld global positioning were used for land cover classification to distinguish mangroves from
system (GPS) receiver. non-mangroves. We used the level 1.5 Fine Beam Dual (FBD) Polariza-
We employed Ward's method of hierarchical cluster analysis to clas- tion PALSAR images that have 12.5 m spatial resolution, obtained from
sify dominant mangrove forest types (Ward Jr., 1963). Input data for the Alaska Satellite Facility (ASF) data portal (University of Alaska Fair-
this analysis were species-level ‘Importance Values’ of mangrove and banks, https://ursa.asfdaac.alaska.edu/). The FBD PALSAR images were
non-mangrove species recorded in each 0.5 ha plot (Peet, 1974). Impor- composed of HH (horizontal transmitted and horizontal received) and
tance Value is an integrated measurement of density (number of trees HV (horizontal transmitted and vertical received) polarizations and
per plot), dominance (proportion of total basal area occupied by a their acquisition dates were during June and July of 2007. These images
given species), and frequency (number of subplots the species was re- were processed, calibrated, and geo-rectified to UTM projection (zone
corded within the larger 0.5 ha plot). A stepwise quantile regression 53 S and WGS-84 datum) with ASF MapReady software version 3.2.1
technique was used to obtain the best representation of field canopy in order to convert the DN values to radar backscatter coefficients (σ°,
heights at the scale of the 20 m diameter subplots because the sizes of unit of decibels dB) using the following formula (Shimada, Isoguchi,
Landsat-8 and SRTM pixels (30 m) and PALSAR pixels (12.5 m) are rel- Tadono, & Isono, 2009):
atively more comparable to the subplot size than to the sizes of individ-
ual canopies. Aboveground biomass of mangrove and non-mangrove 
 
trees was calculated at subplot and plot levels using published σ ðdBÞ ¼ 10 x log10 DN2 −83:4: ð1Þ
species-specific and general allometric equations (Chave et al., 2005;
Clough & Scott, 1989; Kauffman & Donato, 2012; Komiyama,
Poungparn, & Kato, 2005). Allometric equations describe the close rela- These backscatter images were then mosaicked to cover the entire
tionship between tree biomass and a combination of diameter, wood area of study. Additional georectification adjustment of the mosaicked
density, and height (Baker et al., 2004; Chambers, Santos, Ribeiro, & PALSAR image was done using the Landsat-8 images as reference,
Higuchi, 2001; Chave et al., 2005; Nascimento & Laurance, 2002; resulting in Root Mean Squared Error (RMSE) of 7.28 m. A 7 by 7 Lee fil-
Phillips et al., 1998). When species-specific values were not available ter was applied to the georectified and stacked PALSAR mosaic to re-
for general allometric equations, average wood density for the genus move speckle noise (Lee, 1980). Using the HH and HV data, we
or most closely related species was used (a complete list of allometric calculated the Radar Forest Degradation Index (RFDI), which is a ratio
equations used for this study can be found in Table S1). of the power difference between HH and HV polarizations and the

Fig. 2. Mangroves in Papua, Indonesia, and plot design. A demonstration of measuring and inventorying mangrove trees (A); design of main transect plots (0.5 ha) and six subsequent
subplots in the left, with additional plots for land cover map verification in the right (B). Coordinates of each location were taken with a GPS receiver.
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 69

total power of HH and HV polarizations (Mitchard, 2012). 3.2.4. Mangrove delineation


We separated mangrove areas from non-mangroves by using image
segmentation on the composite PALSAR image containing HH, HV, and
HH−HV
RFDI ¼ ð2Þ RFDI (Fig. 4). Image segmentation refers to the process in which raster
HH þ HV
images are partitioned into spatially continuous, disjointed and homo-
geneous regions (segments) based on pixel values and locations
The RFDI is potentially a useful input layer for wetlands vegetation (Jensen, 2005). This method is suitable for initial separation of man-
mapping because HH polarized data has been found to be sensitive for grove types based upon statistical similarity of regions (Held et al.,
detecting water beneath canopy (Hess, Melack, Filoso, & Wang, 1995), 2003), and focuses on separating groups of similar pixels (textural and
whereas HV polarized data has been known for its sensitivity to volume spectral), rather than individual pixels (Blaschke, 2010). The resulting
scattering and biomass (Bourgeau-Chavez, Riordan, Powell, & Miller, segmented mangrove delineation image was then exported to a vector
2009; Mitchard, 2012). Finally, the HH, HV and RFDI layers were stacked map to be processed further through refining the mangrove boundary
to make one multi-layered image for further analyses. with visual interpretation using on-screen digitization in ArcGIS (ESRI,
In addition, we also used the SRTM DEM data with 30 m spatial res- 2014, version 10.2.1). In the process of refining the mangrove boundary,
olution to identify potential mangrove areas based on elevation. The we used atmospherically corrected Landsat-8 images as a background
DEM data were provided by the US Forest Service. Although the primary (Green, Clark, et al., 1998b).
mission of SRTM was to deliver high-resolution topographic data, man-
grove trees normally thrive in intertidal zones; hence, a ground eleva- 3.2.5. Land cover classification using flexible statistical expert-based
tion close to the mean sea level can be used to identify potential method
mangrove areas (Treuhaft, Law, & Asner, 2004). The SRTM DEM derived We introduce a novel technique in this study for classifying multispec-
from the Interferometric Synthetic Aperture Radar (InSAR) technique tral images, called Flexible Statistical Expert Based (FSEB) classification.
has been shown to correspond with mangrove canopy height (Simard The FSEB is a “pixel-based” sequential classification technique developed
et al., 2006). Although there is a 13 + year difference between the based on flexible determination of standard deviation thresholds of spec-
SRTM data collected in February of 2000 and field data collected in tral profiles generated for “surface objects” that have been derived from
late 2013, this discrepancy does not affect our results because these the training area of 16-bit multispectral Landsat-8 data.
are intact ecosystems and presumably in a steady state. Any error at With the FSEB method, the land cover classification process starts by
the pixel level because of treefall gaps or natural forest dynamics should extracting the most distinct surface features on the input Landsat-8 im-
be inherent in the random variability. There may be some dynamics on agery (water and bare soil) by applying water and soil masks built from
the edge of the coastlines but that is such a small proportion of the total normalized difference water index (NDWI) (Gao, 1996) and soil ad-
area it should not have any impacts on the results. We georectified the justed vegetation index (SAVI) (Huete, 1988) respectively. Both of
SRTM images with Landsat-8 images resulting in an RMSE of 4.82 m. these classes are excluded from the rest of classification procedure.
Overall, the image processing procedure of this study is schematically Then the Landsat-8 images were masked into two major areas: ‘man-
shown in Fig. 3. groves’ and ‘non-mangroves,’ based on input from the mangrove

Fig. 3. A flow diagram showing the steps involved in the image analysis for producing land cover map and estimating mangrove canopy height and standing biomass.
70 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

Fig. 4. An RGB composite image (R: HH polarization, G: HV polarization, and B: RFDI) of mosaicked ALOS PALSAR images showing delineation of mangrove forests in Mimika district (green
line). Mangrove extent was then verified and refined using atmospherically corrected Landsat-8 images. (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

delineation file, as described earlier. Then we identified several training riparian forest. Finally, the mosaic tool in ERDAS Imagine® was used
sites of targeted classes (e.g., Rhizophora dominated mangroves) from to combine all extracted classes and produce a coastal land cover map
the high-resolution QuickBird images available on Google Earth. Google of the entire area of study. The sequential nature of FSEB is an important
Earth was utilized as a visualization tool in locating samples of the train- aspect of this study, which we found to be an improved classifier of
ing area of each targeted class, and the spectral signature used for clas- mangroves.
sification were derived from the same areas of the Landsat-8 images.
Other studies have demonstrated the potential of using Google Earth
as visualization tool or reference data to collect training area samples 3.2.6. Estimation of mangrove canopy height and biomass
for assisting land use/cover classification (Cha & Park, 2007; Hu et al., The SRTM based elevation data were used to estimate canopy height
2013; Yu & Gong, 2012). Based on the spectral information (Landsat-8 and aboveground biomass of the mangroves. Canopy height was
bands 1–7 reflectance) from these training sites, we identified all pixels modeled by regressing the SRTM data onto the field canopy height mea-
in the ‘mangrove’ areas of the Landsat-8 image using the following clas- surements at the 246 subplots within the 41 forest plots. The 25 m tran-
sification algorithm: sect interval of linearly arranged subplots inside each main plot was
comparable to the size of a 30 m SRTM pixel. Ideally, when overlaid
if μ i −σ ij ≤bandi ≤μ i þ σ ij then bandi ¼ 1; otherwise bandi ¼ 0 ð3Þ with SRTM imagery, the coordinate of each of these 246 subplots should
closely match a single pixel. However, in some cases multiple subplot
where μ and σ respectively represent mean and standard deviation of coordinates were located within a single SRTM pixel because of the var-
grouped pixel samples from the training sites, i represents band num- iable accuracy of handheld GPS devices used in the field. In cases where
bers of the OLI sensor of Landsat-8, and j represents the positive real this occurred, duplicates were averaged to achieve one value of canopy
number of selected standard deviation. The output is a binary image height per pixel. This procedure resulted in 196 out of 246 subplots
consisting of the ‘targeted’ class and ‘all other mangroves.’ being used to develop the canopy height model for the 30 m SRTM
After one mangrove class was successfully extracted, we masked out DEM pixels. The best model for mangrove canopy height was deter-
that extracted class from the mangrove areas of the Landsat-8 scene. On mined from the highest coefficient of determination (R2). In addition,
the rest of the pixels we then repeated the procedure for other targeted a nonlinear quantile regression (NLQR) technique (Koenker & Bassett
classes of mangrove types until all pixels within the mangrove areas Jr., 1978) was used to model standing biomass of mangrove via
were properly assigned a class. Once the mangrove areas were classi- regressing the mean of estimated SRTM canopy height resulting from
fied, we repeated the same procedure for several classes of coastal veg- the previous process and field measured biomass of 33 0.5 ha mangrove
etation in non-mangrove areas including marshlands, sago palm, and forest inventory plots (Fig. 5). Analysis of variance (ANOVA) was used to
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 71

4. Results and discussion

4.1. Spatial and structural heterogeneity of Mimika's coastal wetlands


Forest

A total of 8353 stems from 20 mangrove species and 38 non-


mangrove species were measured from the 41 0.5 ha sample plots.
The 20 mangrove species documented in this study represent approxi-
mately half of the total 43 mangrove species reported in Papua and
New Guinea (Duke, Ball, & Ellison, 1998). However, our study was not
designed to be an exhaustive survey of biodiversity and many trees
were identified only to the genus level. Therefore, our data do not rep-
resent total mangrove species diversity of the region.
Cluster analysis based on forest community composition led to eight
major forest associations in the inventory plots (Appendix Fig. S1);
however, many species remained unidentified especially in the highly
diverse transitional forests which included numerous non-mangrove
lowland swamp species. A summary of structural characteristics and
biomass of the field plots are presented in Table 1 and their percent con-
tribution are shown in Fig. 6. The forest types are dominated by typical
mangrove species in the Rhizophoraceae family (Rhizophora, Bruguiera,
and Ceriops), with co-dominant Camptostemon and Lumnitzera being re-
markably similar in forest structure and aboveground biomass
Fig. 5. Different intervals of nonlinear quantile regression (NLQR) were examined using (Table 1).
ANOVA to determine the best predictor of biomass model. The nonlinear least squares Substantial variation remains within forest types (Fig. 6), attribut-
(NLS) relationship is shown for reference. able to plot-level differences in the abundance of large trees (N50 cm di-
ameter) among plots. Based on field data, Bruguiera dominated forests,
composed of mixed Bruguiera, Rhizophora, and Lumnitzera species, were
more dominant than other forest types in Mimika, illustrated by the
evaluate and select the best model for estimating aboveground biomass largest share of basal area (28.78%), stem density (18.94%), and above-
of mangroves in the study area from the series of NLQR. ground biomass (37.43%). Their dominance was also observed by the
largest average biomass (416.79 ± 175.41 Mg/ha) and the number of
occupied plots assigned to this forest type (10 plots). In fact, two plots
3.2.7. Accuracy assessment and model performance evaluation from this forest type had biomass in excess of 600 Mg/ha. The tallest
Due to the limited number of ground validation points (66 points, tree within Bruguiera dominated forests varied between 16 m and
the total number of field plots), we collected additional random valida- 30 m. These findings coincide with those of Ellison (2005), who sug-
tion points from the available high resolution DigitalGlobe images on gested that species of Bruguiera were dominant in the Ajkwa and
Google Earth to examine the accuracy of the land cover map that was Tipoeka estuaries of Mimika, Papua.
produced using the FSEB method (Cha & Park, 2007; Hu et al., 2013; Spatial analysis of remote sensing data, however, shows that
Yu & Gong, 2012). Due to the limited high resolution image coverage Bruguiera-dominated forests contain less overall area than Rhizophora-
of the study site, we were able to collect 94 random validation points. dominated forests with a total occupancy of 23,623 ha compared to
Thus, our total validation points were 160 (Fig. 1). Kappa statistic was 77,496 ha, respectively, mostly covering the central part of coastal
used to evaluate classification accuracy of produced land cover map area extending to the northwest (Fig. 7). As noted in other publications,
and presented in the form of an error matrix. The error matrix is a sim- the Bruguiera dominated forests tend to occupy landward mangrove
ple cross-tabulation of the mapped class label against the observed in zones (Ellison, 2005), but thrive in areas with sufficient drainage
the ground or reference data (Congalton & Green, 2008). Model perfor- (Tomlinson, 1986), and occasionally mix with transitional forests. Tran-
mance of canopy height and aboveground biomass estimation derived sitional forests refer to areas where a diverse mix of freshwater swamp
from SRTM DEM were evaluated from 10,000 bootstrap analysis runs species dominate and mangrove associates and true mangroves are
(Efron, 1979; Efron & Tibshirani, 1994) of field measurements of canopy present but not abundant. During our field campaigns we encountered
height and standing biomass with replacement to yield mean absolute six Bruguiera species but only five of them were positively identified.
error (MAE) and root mean squared error (RMSE) (Willmott, 1982). The five species include Bruguiera exaristata, Bruguiera gymnorrhiza,

Table 1
Eight major mangrove associations were identified from the 41 0.5 ha forest inventory plots in Mimika district, Papua Indonesia. For landscape scale biomass inventory, mangroves dom-
inated by Rhizophoraceae (Rhizophora, Bruguiera, and Ceriops) were combined in subsequent analyses as these forests were similar in structure and aboveground biomass. Values are
mean ± standard deviation.

Forest type Basal area (m2/ha) Stem density (#/ha) Aboveground biomass (Mg/ha) Range of max height (m) # of occupied plot

Avicennia, Sonneratia 15.6 ± 3.5 1926 ± 247 106.3 ± 38.6 8–10 4


Rhizophora 27.1 ± 4.8 1045 ± 438 333.3 ± 122.9 15–32 7
Camptostemon, Rhizophora, Bruguiera 34.8 ± 1.5 733 ± 197 300.9 ± 96.7 25–31 2
Bruguiera, Rhizophora, Lumnitzera 33.3 ± 7.4 786 ± 215 416.8 ± 175.4 16–30 10
Ceriops 34.4 ± 6.4 827 ± 491 387.1 ± 18.4 28–29 2
Transitional (coastal) 35.3 ± 9.8 1132 ± 250 286.7 ± 68.2 30–35 3
Transitional (landward) 26.4 ± 11.6 1108 ± 313 182.0 ± 97.9 25–27 6
Sago 26.4 ± 5.1 960 ± 289 137.7 ± 24.6 25–31 7
72 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

Fig. 6. Percentage of basal area, stem density, and biomass across eight major groups of forest types within the areas of Mimika derived from cluster analysis. AVI = Avicennia, BRU =
Bruguiera, CAM = Camptostemon, CER = Ceriops, RHI = Rhizophora, SON = Sonneratia, TRANS = Transitional. Percentages were calculated from 41 0.5 ha forest inventory plots.

Bruguiera hainesii, Bruguiera parviflora, and Bruguiera sexangula. Based lowest mean aboveground biomass, (106.26 ± 38.55 Mg/ha), which is
on Ellison's (2005) study, the sixth unidentified species is likely approximately 300 Mg/ha lower than the biomass of Bruguiera-
Bruguiera cylindrica. Bruguiera hainesii is listed by the International dominated forest.
Union for Conservation of Nature (IUCN) as a critically endangered Based on land cover classification and field inventory data from non-
mangrove species (Duke et al., 2010), and was recorded in the transi- mangrove plots, Sago palms tend to be more predominant in transi-
tional coastal zones of the study area where freshwater swamp species tional landward and coastal forests within the study area. Sago palms
mix with mangroves on elevated beach ridges. represented 15% of basal area, 16.18% of stem density, and 8.66% of bio-
The second dominant group of mangroves was composed of mass in the study area. Transitional mixed (non-mangrove) landward
Rhizophora species that according to field data, occupied 16.37% basal forests occupied 12.45% basal area, 13.02% stem density, 9.05% biomass,
area, 17.62% stem density, and 20.96% of the aboveground biomass of whereas transitional mixed coastal forests contained 9.14% basal area,
the delineated mangrove area (Fig. 6). Rhizophora inhabit seaward mar- 81.18% stem density, and 7.72% biomass of the study area respectively.
gins with lower elevations (Ellison, 2005), and were comprised of The mean biomass of Sago was 137.72 ± 24.62 Mg/ha with maximum
Rhizophora mucronata, Rhizophora apiculata, and Rhizophora stylosa spe- height varying from 25 to 31 m among plots. However, the biomass es-
cies, with R. apiculata the most abundant. Their mean biomass was rel- timation of Sago was calculated from the best information available, as
atively high, (333.31 ± 122.95 Mg/ha) and the tallest trees varied from Sago specific allometric equations have not been developed. Equations
14 to 32 m. Rhizophora-dominated forests spread from the central part used for this study were derived from data collected for a different pur-
of the mangrove zone to the southeast, although Rhizophora was also pose (Table. S1; Gusmayanti, unpublished data). Since palm fibers are
abundant in the western part of the study area (Fig. 7). In addition, much less dense than wood, the biomass of Sago palms in adult life
the co-dominant Ceriops- and Camptostemon-dominated forests were stages is much lower than trees of similar diameter. This was reflected
similar in their percentage of basal area occupied (5.93% and 6%, respec- in the low biomass estimates for Sago plots, where the majority of
tively) and stem density (3.98% and 3.53%, respectively), with slight var- standing biomass occurs in the interspersed hardwoods. Sago palm is
iations in standing biomass (6.95% and 5.41%, respectively) (Fig. 6). distributed throughout the eastern part of study area, intermixed with
They also possessed relatively high mean biomass, (387.14 ± swamp forest and/or swamp shrublands, pandanus, and grasslands cov-
18.40 Mg/ha, Ceriops-dominated forests; and 300.89 ± 96.75 Mg/ha, ering a total area of 107,129 ha (Fig. 6). Transitional forests represent an
Camptostemon-dominated forests), with maximum canopy height vary- ecotone between mangroves and freshwater swamp forests. They can
ing between 28 and 29 m for Ceriops and from 25 to 31 m for be further subdivided into “coastal” and “landward” types in which
Camptostemon. Bruguiera, Ceriops and Camptostemon species inhabit both forests have unique characteristics and occupy distinct landforms
the landward portion of the mangrove zone (Ellison, 2005) with higher (Marshall, 2007). These forests are where the endangered Bruguiera
drainage (Tomlinson, 1986). Avicennia/Sonneratia forests colonize hainesii were documented. The mean stem density of the coastal tran-
newly exposed mudflats as pioneering stands. The land cover map indi- sitional forests was almost similar to that of landward transitional
cates that the Avicennia/Sonneratia forests are confined to coastal zones forests; however, their average biomass was about 100 Mg/ha higher
covering 5472 ha of the study area. They are characterized by a lower than landward forests. The difference could be attributed to signifi-
canopy height (between 8 and 10 m), the lowest basal area (5.38%) cant variation in canopy height between both types of transitional
and aboveground biomass (3.82%), but the highest density of smaller forest (Table 1). For example, the tallest trees in the transitional
stems (18.55%) compared to other mangrove types. Their stem density coastal forests plots were between 30 and 35 m; conversely, the
surpassed Rhizophora-dominated forest and was similar to Bruguiera- maximum canopy heights recorded for landward forest plots were
dominated forest (Fig. 6). Avicennia/Sonneratia stands contained the between 25 and 27 m.
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 73

4.2. Evaluation of land cover classification methods Rhizophora forest, Avicennia/Sonneratia forest, Bruguiera dominated for-
est, other (mixed) mangrove forest, transitional landward and coastal
The area and statistical accuracy of the 17 classes included in the forests, and Nypa stands.
land cover map derived from a combination of image segmentation The rule-based FSEB classification substantially improved mangrove
and FSEB classification methods are shown in Table 2. The total area of classification accuracy compared to supervised classification methods.
mangroves in the study was 186,291 ha with Rhizophora forest account- These results are in agreement with those of Gao, Chen, Zhang, and
ing for the largest coverage (5.78%) and Avicennia/Sonneratia forest with Zha (2004), who reported overall accuracy of 46.7%and 68.3% for super-
the smallest coverage (0.41%). vised parametric classification used to map stunted and lush mangroves
Moreover, the total classified forest area, including swamp forest, in the western portion of the Waitemata Harbor, Auckland, New
dryland/riparian forest, and all types of mangrove forest, was Zealand, whereas after the spatial knowledge (spatial situation and
941,891 ha. It is about 3% of the 35.2 million hectares of primary forest spectral properties) was sequentially incorporated, these accuracies im-
in Papua Island (Margono et al., 2014). Our results indicate that al- proved substantially, to 83.3% and 96.7%, respectively. Another study
though the settlements, roads and palm oil concessions accounted for that mapped Alpine vegetation cover in southwest Yukon Territory,
only 0.67% of the total area of the study site, rapid land clearing for Canada demonstrated improved classification accuracy from 74%
palm oil plantation to the west of the city of Timika is contributing to when using satellite data alone to 85% accuracy when incorporating
an increase in deforestation of Mimika's coastal wetlands. The classifica- geomorphometric variables extracted from DEM data (Wilson &
tion methods presented in this study produced an overall classification Franklin, 1992).
accuracy of 94.38% with kappa statistics of 0.94, surpassing many typical Three factors were important for our classification technique's over-
recommended targets of 85% of overall accuracy (Foody, 2002). The all accuracy: proper band selection, proper determination of standard
combination of image segmentation and FSEB methods were effective deviation value (σ), and sequential classification procedure. Proper
for mapping tropical wetland vegetation indicated by high percentages band selection was done by analyzing the spectral profiles of seven
of producer's and user's accuracy (N81.82% and N 85.71%), respectively. multi-spectral bands of Landsat-8 and finding the appropriate bands
Our methods have been successful in classifying 10 groups of wetland for characterizing certain genus/species of mangrove/non-mangrove
vegetation including several genera of mangroves (Fig. 7). Within man- forest (Fig. 8A). Proper determination of σ was introduced to include di-
grove genera category, for example, all groups of mangrove have versity in a class. For instance, 1σ may or may not be enough to classify
producer's accuracy N 85% whereas for user's accuracy, only mixed man- particular classes of land cover, and a factor of 2σ or larger may be
grove class had b 85% user's accuracy. Wetland cover classes included: needed to include the variance when a relatively high dispersion of
swamp grassland, swamp shrubland, swamp forest, Sago forest, spectral reflectance occurs within a species in an area (Figs. 8B and C).

Fig. 7. Coastal land cover map of study area shown 17 distinct land cover classes derived from Landsat-8 imagery using FSEB method.
74 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

Table 2
Accuracy matrix of land cover map showed producer's and user's accuracy including area for each class cover types.

Class name Area (ha) Area (%) Reference Classified # Correct Producer's (%) User's (%) Kappa

Dryland/riparian forest 278,136.45 20.75 17 16 15 88.24 93.75 0.93


Swamp grassland 13,254.12 0.99 6 6 6 100.00 100.00 1.00
Swamp scrubland 54,036.00 4.03 3 3 3 100.00 100.00 1.00
Sago palm dominated forest 107,129.07 7.99 10 9 9 90.00 100.00 1.00
Mixed swamp forest 301,918.23 22.52 15 16 15 100.00 93.75 0.93
Grassland/scrubland 18,416.07 1.37 7 5 5 71.43 100.00 1.00
Settlements/roads 3663.81 0.27 4 6 4 100.00 66.67 0.66
Palm oil concession 5328.00 0.40 7 7 7 100.00 100.00 1.00
Open water 277,186.41 20.68 15 15 15 100.00 100.00 1.00
Cloud 2059.20 0.15 7 7 7 100.00 100.00 1.00
Barren land – soil 24,583.86 1.83 8 8 8 100.00 100.00 1.00
Rhizopora dominated forest 77,496.39 5.78 22 20 20 90.91 100.00 1.00
Avicennia/Sonneratia forest 5471.19 0.41 7 6 6 85.71 100.00 1.00
Bruguiera dominated forest 23,622.66 1.76 7 8 7 100.00 87.50 0.87
Mixed mangrove forest 68,696.73 5.13 10 11 9 90.00 81.82 0.81
Transitional landward and coastal forest 68,416.47 5.10 9 10 9 100.00 90.00 0.89
Nypa palm forest 11,004.12 0.82 6 7 6 100.00 85.71 0.85
Totals 1,340,418.78 100.00 160 160 151

Fig. 8. A demonstration of the spectral response profiles used for the FSEB classification method. The difference in mean spectral profile response from Rhizophora and Metroxylon Sago on
multispectral band of Landsat-8 (A); the spectral profile of Rhizophora associations species with ±2 σ used for FSEB classification (B); and the spectral profile of Sago palm (Metroxylon
Sago) with ±2.575 σ used for FSEB classification (C).
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 75

Fig. 9. Optimal quantile for the mangrove canopy height model (between SRTM DEM and field canopy heights) was determined using a stepwise quantile regression technique at subplot
scale (A). The scatterplot (B) shows the agreement between SRTM DEM and the 0.98 quantile of field canopy height, H.98.

We found that the incremental use of σ is important for implemen- image, and the remaining image to be used for classifying each subse-
tation of the FSEB classification method, instead of depending solely on quent land cover class until all surface features within Landsat-8 images
a fixed σ as a default. The sequential classification procedure allowed an are properly classified. This procedure helped to avoid overlapping re-
easily identifiable land cover to be classified and removed from the gions between classes.

Fig. 10. Map of mangrove canopy height derived from SRTM elevation data. The map showed cluster patterns in canopy heights corresponding to different forest types.
76 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

This study shows for the first time that the Landsat-8 data of higher
radiometric resolution can be used to map mangroves up to genus levels
at broad spatial scales. Nonetheless, although the FSEB method was suc-
cessful in classifying mangrove forests of Papua, it also showed some
limitations where it failed to separate some genera of mangroves such
as stands of Lumnitzera, Ceriops, and Camptostemon. A possible cause
of this limitation is that we could not select any “pure” spectra of
these mangrove types for training because they are intermixed with
other mangrove species/genera, including Rhizophora and Bruguiera.
The ability of the image segmentation method applied to PALSAR
composite images (i.e., HH, HV, and RFDI) to separate mangroves from
non-mangroves indicates that the methodology can be applied to refine
current boundaries of Landsat-based global mangrove distribution
maps produced by Giri et al. (2011). Several studies have emphasized
the effectiveness of using L-band SAR data for mapping mangrove
(Flores De Santiago et al., 2013; Kovacs et al., 2013; Lucas et al., 2007;
Lucas et al., 2014; Rocha de Souza Pereira et al., 2012). Flores De
Santiago et al. (2013) have also shown that image segmentation of
ALOS PALSAR data can be used for mapping mangrove areas in tropical
areas with persistent cloud cover. Our study demonstrates that in addi-
tion to separating mangroves from non-mangroves, the combined use
of ALOS PALSAR and Landsat-8 imagery can be used to map mangrove
composition to the genus level at regional scales. The methods pre-
sented here to produce detailed coastal wetland maps has the potential
Fig. 11. The 0.5 NLQR was the best predictor of relationship between SRTM-derived
to provide decision makers with a valuable tool for assessing mangrove
canopy height model (based on Eq. 3) and measured biomass from 33 0.5 ha mangroves
inventory plots (R2 = 0.46). Two outliers from the atypical sandy soil plots are shown resources, as well as their use in climate adaptation and mitigation
inside a rectangle. efforts.

Fig. 12. A map of mangrove standing biomass (Mg/ha) derived from the model based on SRTM elevation data. The inbox shows clustered patterns of biomass that correspond to different
forest types.
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 77

Table 3
A comparison of biomass obtained from the modeled, measured, and extrapolated to an entire class from the field measurements, for each land cover types. Values are mean ± standard
deviation.

Forest types Area (Ha) Modeled Measured Modeled total (Tg) Measured (extrapolated) % Difference
estimation (Tg)

Mean ± STD (Mg/ha) Mean ± STD (Mg/ha) Lower 90% CI Total Upper 90% CI

Avicennia/Sonneratia forest 5472 237.52 ± 98.20 106.26 ± 38.55 1.24 0.32 0.58 0.85 72.53
Bruguiera dominated forest 23,623 295.09 ± 86.17 416.79 ± 175.41 6.95 4.64 9.85 15.05 34.52
Other mangroves forest 79,701 284.93 ± 103.89 344.02 ± 57.56 22.29 21.59 27.42 33.25 20.64
Rhizophora dominated forest 77,496 353.30 ± 98.43 333.31 ± 122.95 27.32 13.85 25.83 37.81 5.61
All mangroves 186,291 313.58 ± 105.28 299.56 ± 98.62 57.79 32.71 55.81 78.9 3.48
Transitional forest 68,417 308.20 ± 96.51 234.34 ± 83.05 21.02 8.89 16.03 23.17 26.94

4.3. Evaluation of canopy height model temperate forest and tropical rainforests (Clark, Clark, & Roberts,
2004; Gaveau & Hill, 2003; Rönnholm et al., 2004; Wang & Glenn, 2008).
We found a strong linear correlation between the SRTM DEM and Using Eq. (4), we produced a canopy height map at 30 m spatial res-
the 0.98 quantile of field canopy height (H.98; the tallest trees in each olution of mangrove areas in Mimika district (Fig. 10). The modeled
10 m radius subplot) indicated by high correlation coefficient (r = mangrove canopy heights ranged from 4.3 to 36.2 m with an estimated
0.84) and predicted coefficient of determination (Pred-R2 = 0.804) mean of 20.8 m. Mangrove trees in Papua and New Guinea can reach up
(Fig. 9). to 40 m in height (Alongi, 2007); however, in general, their canopy
The high correlation was expected because the SRTM DEM was de- heights are between 25 and 30 m, typically due to the mangrove species
rived from single-pass interferometry of the C-band radar signal from in the Rhizophoraceae family (Ellison, 2005).
which heights are measured according to scattering phase center, or For comparison, Simard et al. (2008) reported that 15 m was the
the weighted mean of all contributing backscatter throughout the maximum height of mangroves grown in Ciénaga Grande de Santa
depth of canopy (Madsen et al.,1993). Since mangrove forest canopies Marta, Colombia, while Lucas et al. (2002) demonstrated that tall man-
are typically dense and SRTM C-band contains short wavelengths, the groves of northern Australia have maximum heights varying from 16 to
InSAR scattering phase center would be located close to the top of can- 20 m. A study by Pool, Snedaker, and Lugo (1977) reported that the
opy, corresponding closely with field-based canopy-height estimation tallest mangrove trees varied from 9 m in Florida to 15 m in Puerto
(Kellndorfer et al., 2004; Madsen, Zebker, & Martin, 1993). The resulting Rico, 16 m in Costa Rica, and 17 m in Mexico. Based on our findings,
optimal model fit between SRTM DEM and H.98 in our study is mangrove trees in our study area are consistently among the tallest
expressed by the following equation: reported.
Superimposing the land-cover map (Fig. 7) on the canopy height
map (Fig. 10) showed a clustered pattern (zonation) among mangrove
CHMSRTM ðmÞ ¼ 2:7191 ðSRTM DEMÞ0:676 ð4Þ
forest types in Mimika. For example, the majority of tall mangroves
(from the Rhizophoraceae family) occupy the central part of the study
where CHMSRTM represents a canopy-height model derived from SRTM area downstream of larger rivers. In contrast, the shortest mangrove
DEM data. A bootstrapping error analysis (10,000 resamples with re- trees, mainly Nypa, were concentrated in the southeastern portions of
placement) produced average errors of 3.0 m for MAE and 3.7 m for our study area which receive less freshwater inflow from rivers. Short
RMSE for the 30 m SRTM pixels. The RMSE and predicted R2- canopy Avicennia- and Sonneratia-dominated mangroves were also
coefficient of the canopy heights of our model indicated less uncertainty present in the coastal fringe and along aggrading coastlines. Such clus-
than those previously reported by Kellndorfer et al. (2004). Another tered patterns are attributable to periodic changes in the intertidal envi-
study by Simard, Rivera-Monroy, Mancera-Pineda, Castañeda-Moya, ronment and hydrogeological settings. Zonation of mangrove species is
and Twilley (2008) calibrated SRTM DEM with field measurement of influenced by the frequency of inundation, drainage, and/or elevation,
mangrove canopy height, resulting in an RMSE of 2.7 m. However, as estuarine shorelines are largely determined by the interaction be-
their error estimation is more limited as it was estimated from 30 tween sediment supply and variation in sea level (Ellison, 2000, 2005).
plots with only 166 trees in total, whereas our study is derived from
41 field plots with N 8000 trees measured. Similar research by 4.4. Evaluation of biomass model
Fatoyinbo and Simard (2013) estimated mangrove canopy height across
Africa and obtained the overall RMSE of 3.6 m, whereas another study Based on an ANOVA of the series of non-linear quantile regressions,
by Simard et al. (2006) in Everglades National Park revealed the RMSE the 0.5 quantile was the best model to calculate the spatial distribution
of 2.0 m. Although the uncertainty estimation of our canopy height of aboveground mangrove biomass at the landscape scale in the study
model is comparable to Fatoyinbo and Simard (2013) and Simard area (Fig. 11). The two outliers highlighted in Fig. 11 are plots domi-
et al. (2006); their error estimations were not based on ground-truth nated by stunted Lumnitzera littoralis and Rhizophora apiculata found
measurements, rather inferred from the fusion of SRTM DEM data and in coarse sands rather than fine silt/clay substrates that are typical to
Light Detection and Ranging (LiDAR) data. Several studies have shown this region. The architecture, and therefore allometry, of trees growing
a consistent underestimation of canopy height using LiDAR in in the coarse sand plots is atypical, dominated by secondary growth

Table 4
Quantitative error measures and paired mean t-test of mangrove standing biomass model performance. Values are mean ± standard deviation.

Forest types N Predicted mean ± STD (Mg/ha) Observed mean ± STD (Mg/ha) MAE (Mg/ha) RMSE (Mg/ha) r

All mangroves 145 292.72 ± 121.89 314.37 ± 172.62 105.75 147.98 0.55
Avicennia/Sonneratia, Bruguiera, Rhizophora 104 303.96 ± 128.02 313.43 ± 176.20 91.30 129.51 0.67
Other mangroves 41 264.22 ± 100.59 316.97 ± 165.29 142.39 186.84 0.14
Transitional forest 45 346.10 ± 105.29 254.20 ± 134.79 123.57 143.27 0.60
Mangroves and non-mangroves 190 305.36 ± 120.07 300.12 ± 166.11 109.97 146.88 0.50
78 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

Fig. 13. Comparison between the observed and predicted standing biomass of mangroves: all mangroves combined (A), Avicennia + Sonneratia + Bruguiera + Rhizopora combined (B),
mixed mangroves (C), transitional forests (D) and mangroves and non-mangroves combined (E).
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 79

that produce short, stunted trees that grow large in diameter but not (Fig. 13A). The difference between means is nearly 7% of the observed
height. Forest biomass in these areas is underestimated, since the archi- mean biomass with MAE = 105.75 Mg/ha and RMSE = 147.98 Mg/ha.
tecture of this mangrove type is not represented in species-specific allo- Mangroves that are shorter statured, such as those dominated by
metric equations. Ceriops, Lumnitzera, and stunted Rhizophora (included in the “other
Unlike the previous studies that used linear regression to model mangroves” cover class) generally have lower predicted biomass than
mangrove biomass from canopy heights (Fatoyinbo & Simard, 2013; observed in the field, which is an artifact of the allometric equations
Fatoyinbo et al., 2008; Simard et al., 2006, 2008), we applied a non- used rather than any error in the model. When the “other mangroves”
linear regression (exponential function) to model aboveground biomass cover class is excluded from analysis, model fit improves substantially
of mangroves because theoretically is a more justifiable approach and and the difference between predicted and observed means is 3% of the
can be explained as follows. Normally, at an early stage (seedlings and observed mean (Fig. 13B). Mangroves dominated by Avicennia/
saplings) different mangrove species will grow and have heights that Sonneratia, Bruguiera and Rhizophora stands account for 57% of the
linearly correspond with their biomass. However, in nature, each tree total mangrove area in Mimika. Nevertheless, when applied to other
species will compete and respond differently to the sunlight. Therefore, mangroves, the canopy height-biomass model does not fit observed bio-
some trees from the same species or from different species may have mass data very well even though predicted biomass values are within
higher rate of growth and biomass than others. As a result, these partic- the range of observed values (Fig. 13C). The difference between
ular trees will be taller than others and wider in diameter and biomass means is about 18% of the observed values. Again, the lack of fit could
as well, since they received more sunlight. Naturally, at one point the be partly due to the underestimation of biomass in the field using allo-
height and diameter of trunks of these taller trees will slow their growth metric equations that are not representative of shorter statured, large
rate due to their maturity, but their biomass will still increase as a result diameter Rhizophora and Lumnitzera mangroves growing on beach
of secondary growth from branches. Several other studies have used ridges. SRTM data may also not capture the few larger trees present in
non-linear regression approach, particularly an exponential function, low canopy mangroves. These larger diameter trees contribute most
to model aboveground biomass in tropical rainforests based on field to the estimates of plot level (0.5 ha) aboveground biomass. Although
data collection (Chave et al., 2005; Kauffman & Donato, 2012). In this the Pearson's correlation coefficient between predicted and observed
study, the relationship between CHMSRTM (m) and measured above- is relatively high (r = 0.60) for transitional forests growing on elevated
ground biomass (Mg/ha) at the 0.5 quantile was found to be: beach ridges, their standing biomass were overestimated by the canopy
height-biomass model (Fig. 13D). Biomass observed in the field may be
BIOMASSSRTM ðMg=haÞ ¼ expð3:9042 þ 0:0858 ðCHMSRTM Þ ðmÞÞ ð5Þ lower than predicted because of different height-biomass relationships
of non-mangrove species, lower wood density, and smaller diameter
The map of spatial distribution of mangrove biomass derived from trees dominating the canopy. Nevertheless, when all forest inventory
Eq. 5 is provided in Fig. 12. To compare biomass from the model and data are combined, the differences between predicted and observed
that from simple extrapolation of field data (Table 3), total mangrove mean biomass values are within 2% of the observed value (r = 0.5;
biomass was scaled to the landscape by applying the canopy height- Fig. 13E).
biomass model to all forest types within the study area (modeled),
and multiplying the mean observed biomass value of each cover class 5. Conclusions
by its coverage (extrapolated) (Fig.7). There was considerable variation
in aboveground biomass among forest types within mangroves both In the development of new approaches for applying remote sensing
from field calculation and modeled biomass (Table 3). data to derive ecosystem properties, accuracy and simplicity are critical
For instance, average aboveground biomass varies from 237.52 ± factors. Furthermore, the methods must be efficient in terms of effort
98.2 Mg/ha, in a low canopy young Avicennia/Sonneratia stands to (time and cost) to implement and should be capable of broad applica-
353.52 ± 98.43 Mg/ha in a mature tall canopy Rhizophora-dominated tion with minimum supervision. In this study, a remote sensing ap-
stands. Modeled biomass estimation of Avicennia/Sonneratia forests proach was developed by combining multispectral and SAR data for
was higher than the biomass estimates resulting from simple extrapola- mapping the spatial distribution and standing biomass of coastal wet-
tion from field plots by 131.26 Mg/ha. The difference could be partly at- land forests of southern Papua, Indonesia. A new classification method,
tributed to taller statured Avicennia/Sonneratia forests not well FSEB, was introduced for mapping spatial distribution of coastal wet-
represented in the field dataset. Nonetheless, at the landscape scale, land forests, and nonlinear quantile regression was employed to
the canopy height-biomass model indicated only a 3.48% difference be- model aboveground biomass of mangrove. The results showed that
tween the modeled and the measured estimates of total biomass of all the incremental use of variable standard deviations is a realistic ap-
mangroves combined, meaning that the height and biomass models de- proach for implementing the FSEB classification method, instead of de-
veloped in this study are reliable and can be applied broadly for map- pending solely on a fixed standard deviation classification approach.
ping and monitoring mangrove biomass over large areas of Indonesia. This study shows for the first time that Landsat-8 data can be used to
Furthermore, modeled standing biomass estimations were within the map mangroves at the genus level rather than classifying mangroves as
90% confidence intervals for Rhizophora-dominated forest, Bruguiera- a single general class. In addition, contrary to the previously published
dominated forests, and all combined mangroves forests when com- mangrove biomass studies, the use of a nonlinear regression (exponen-
pared with measured area weighted biomass. The average biomass tial function) to model standing biomass of mangroves was used be-
value of all mangrove forest types except for the field-measured cause it follows the growth pattern of trees regardless of their species.
Avicennia/Sonneratia forest greatly exceeded the Intergovernmental The reliability of our approach is demonstrated by high overall accuracy
Panel on Climate Change (IPCC) default value for aboveground biomass and kappa coefficient (94.38% and 0.94, respectively) for the land cover
of wet tropical mangroves (192 Mg/ha) (Hiraishi et al., 2014). The mean classification (surpassing many typical recommended targets of overall
values of field-measured biomass for Bruguiera-dominated forest were accuracy; e.g., Foody, 2002). A point of caution is that given the rela-
double the IPCC default value (416.79 versus 192 Mg/ha, respectively), tively limited number of validation data points (160) that we had avail-
indicating that mangrove biomass in southern Papua is much higher able for this vast area of study (N 1.3 million ha), the high accuracy of
than the global average, and among the highest on Earth. classification achieved by our method may require further validation
A summary of error analysis between predicted and observed man- in other areas. Additionally, our method also produced a high percent-
grove standing biomass at the subplot level is given in Table 4. age of the predicted R2 with low margins of error (MAE and RMSE) for
At the subplot level, the predicted values of aboveground biomass the canopy height model, and predicted estimates within 90% confi-
were well within the range of observed values for all mangroves dence intervals between modeled and measured values for the
80 A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81

landscape-scale mangrove standing biomass. Our results show that the Donato, D. C., Kauffman, J. B., Murdiyarso, D., Kurnianto, S., Stidham, M., & Kanninen, M.
(2011). Mangroves among the most carbon-rich forests in the tropics. Nature
methods proposed in this study can be used can potentially provide de- Geoscience, 4, 293–297.
cision makers with a valuable tool for assessing mangrove resources, Duke, N., Ball, M., & Ellison, J. (1998). Factors influencing biodiversity and distributional
and can be used for climate change adaptation and mitigation efforts. gradients in mangroves. Global Ecology and Biogeography Letters, 7, 27–47.
Duke, N., Kathiresan, K., Salmo, S. G., III, Fernando, E. S., Peras, J. R., Sukardjo, S., ... Ngoc
Nam, V. (2010). Bruguiera hainesii. The IUCN red list of threatened species Version
2014.3. bwww.iucnredlist.orgN. Downloaded on 05 April 2015.
Acknowledgements Efron, B. (1979). Bootstrap methods: Another look at the jackknife. The Annals of Statistics,
1–26.
This work was supported by the USAID Indonesia Forest and Climate Efron, B., & Tibshirani, R. J. (1994). An introduction to the bootstrap. CRC press.
Ellison, J. C. (2005). Holocene palynology and sea-level change in two estuaries in
Support (IFACS) (contract # AG-3187-C-13-0010) through Kamoro Southern Irian Jaya. Palaeogeography, Palaeoclimatology, Palaeoecology, 220,
Collaborative Wetlands Management Program. The authors would like 291–309.
to thank to the USAID Indonesia Forest and Climate Support (IFACS) Ellison, J. C. (2000). How South Pacific mangroves may respond to predicted climate
change and sea-level rise. Climate change in the South Pacific: Impacts and responses
Program and their Timika field office for support in collecting field in Australia, New Zealand, and small island states (pp. 289–300). Springer.
data. Field data collection was accomplished by the tireless efforts of ESRI (2014). ArcGIS desktop release 10.2.1. Redlands CA: Environmental Systems Research
Taryono Darusman, Rudi, Hendri Saleh, Irvansyah, and Leman Noerman Institute.
FAO (2007). The world's mangroves 1980–2005. FAO forestry paper 153. Italy: Rome.
of Puter Foundation, Indonesia. The ALOS PALSAR imagery used in this Foody, G. M. (2002). Status of land cover classification accuracy assessment. Remote
study was provided by the Alaska Satellite Facility (ASF) under ALOS Sensing of Environment, 80, 185–201.
PALSAR and Student Proposal Program. Fatoyinbo, T. E., & Simard, M. (2013). Height and biomass of mangroves in Africa from
ICESat/GLAS and SRTM. International Journal of Remote Sensing, 34, 668–681.
Fatoyinbo, T. E., Simard, M., Washington-Allen, R. A., & Shugart, H. H. (2008). Landscape-
scale extent, height, biomass, and carbon estimation of Mozambique's mangrove for-
Appendix A. Supplementary data ests with Landsat ETM+ and Shuttle Radar Topography Mission elevation data.
Journal of Geophysical Research, 113 G02S06.
Supplementary data to this article can be found online at http://dx. Flores De Santiago, F., Kovacs, J. M., & Lafrance, P. (2013). An object-oriented classification
method for mapping mangroves in Guinea, West Africa, using multipolarized ALOS
doi.org/10.1016/j.rse.2016.04.026. PALSAR L-band data. International Journal of Remote Sensing, 34, 563–586.
Fujimoto, K., Imaya, A., Tabuchi, R., Kuramoto, S., Utsugi, H., & Murofushi, T. (1999). Be-
lowground carbon storage of Micronesian mangrove forests. Ecological Research, 14,
References 409–413.
Gao, B. C. (1996). NDWI—A normalized difference water index for remote sensing of veg-
Alongi, D. M. (2007). 5.4. Mangrove forests of Papua. Ecology of Indonesian Papua part two etation liquid water from space. Remote Sensing of Environment, 58, 257–266.
(pp. 824). Gao, J. (1998). A hybrid method toward accurate mapping of mangroves in a marginal
Alongi, D. M. (2012). Carbon sequestration in mangrove forests. Carbon Management, 3, habitat from SPOT multispectral data. International Journal of Remote Sensing, 19,
313–322. 1887–1899.
Al-Wassai, F. A., & Kalyankar, N. (2013). Major limitations of satellite images. arXiv preprint Gao, J., Chen, H., Zhang, Y., & Zha, Y. (2004). Knowledge-based approaches to accurate
arXiv:1307.2434. mapping of mangroves from satellite data. Photogrammetric Engineering & Remote
ARD, T. T. (2014). USAID IFACS quarterly report year 4. Second quarter of year 4 workplan: Sensing, 70, 1241–1248.
Indonesia forest and climate support. Gaveau, D. L., & Hill, R. A. (2003). Quantifying canopy height underestimation by laser
Aschbacher, J., Tiangco, P., Giri, C., Ofren, R., Paudyal, D., & Ang, Y. (1995). Comparison of pulse penetration in small-footprint airborne laser scanning data. Canadian Journal
different sensors and analysis techniques for tropical mangrove forest mapping. Geo- of Remote Sensing, 29, 650–657.
science and remote sensing symposium IGARSS'95.'. Quantitative remote sensing for Giri, C., Ochieng, E., Tieszen, L., Zhu, Z., Singh, A., Loveland, T., ... Duke, N. (2011). Status
science and applications', international. (pp. 2109–2111) IEEE. and distribution of mangrove forests of the world using earth observation satellite
Baker, T. R., Phillips, O. L., Malhi, Y., Almeida, S., Arroyo, L., Di Fiore, A., ... Laurance, W. F. data. Global Ecology and Biogeography, 20, 154–159.
(2004). Variation in wood density determines spatial patterns in Amazonian forest GOI (Government of Indonesia) (2011). Presidential regulation no 61/2011 on national ac-
biomass. Global Change Biology, 10, 545–562. tion plan to reduce green house gasses (GHGs) emission. Jakarta, Indonesia: Office of the
Bakosurtanal (2009). Peta mangroves Indonesia. Cibinong: Pusat Survey Sumberdaya Alam President.
Laut. Badan Koordinasi Survey dan Pemetaan Nasional. Green, E., Mumby, P., Edwards, A., Clark, C., & Ellis, A. (1998a). The assessment of man-
Barbier, E. B. (2012). Progress and challenges in valuing coastal and marine ecosystem grove areas using high resolution multispectral airborne imagery. Journal of Coastal
services. Review of Environmental Economics and Policy, 6, 1–19. Research, 433–443.
Blaschke, T. (2010). Object based image analysis for remote sensing. ISPRS Journal of Green, E. P., Clark, C. D., Mumby, P. J., Edwards, A. J., & Ellis, A. (1998b). Remote sensing
Photogrammetry and Remote Sensing, 65, 2–16. techniques for mangrove mapping. International Journal of Remote Sensing, 19,
Nowels, M. (2009). Improving wetland characterization with multi-sensor, multi- 935–956.
temporal SAR and optical/infrared data fusion. In L. L. Bourgeau-Chavez, K. Riordan, Held, A., Ticehurst, C., Lymburner, L., & Williams, N. (2003). High resolution mapping of
R. B. Powell, N. Miller, & G. Jedlovec (Eds.), Advances in geoscience and remote sensing tropical mangrove ecosystems using hyperspectral and radar remote sensing.
(pp. 679–708). India: InTech Publishers. International Journal of Remote Sensing, 24, 2739–2759.
Brondízio, E., Moran, E., Mausel, P., & Wu, Y. (1996). Land cover in the Amazon estuary: Hess, L. L., Melack, J. M., Filoso, S., & Wang, Y. (1995). Delineation of inundated area and
Linking of the Thematic Mapper with botanical and historical data. Photogrammetric vegetation along the Amazon floodplain with the SIR-C synthetic aperture radar.
Engineering and Remote Sensing, 62, 921–930. IEEE Transactions on Geoscience and Remote Sensing, 33, 896–904.
Chambers, J. Q., Santos, J. d., Ribeiro, R. J., & Higuchi, N. (2001). Tree damage, allometric Hiraishi, T., Krug, T., Tanabe, K., Srivastava, N., Jamsranjav, B., Fukuda, M., & Troxler, T.
relationships, and above-ground net primary production in central Amazon forest. (2014). 2013 supplement to the 2006 IPCC guidelines for national greenhouse gas inven-
Forest Ecology and Management, 152, 73–84. tories: Wetlands.
Cha, S., & Park, C. (2007). The utilization of Google Earth images as reference data for the Hu, Q., Wu, W., Xia, T., Yu, Q., Yang, P., Li, Z., & Song, Q. (2013). Exploring the use of Google
multitemporal land cover classification with MODIS data of North Korea. Korean Earth imagery and object-based methods in land use/cover mapping. Remote Sensing,
Journal of Remote Sensing, 23, 483–491. 5, 6026–6042.
Chave, J., Andalo, C., Brown, S., Cairns, M., Chambers, J., Eamus, D., ... Kira, T. (2005). Tree Huete, A. R. (1988). A soil-adjusted vegetation index (SAVI). Remote Sensing of
allometry and improved estimation of carbon stocks and balance in tropical forests. Environment, 25, 295–309.
Oecologia, 145, 87–99. Ilman, M., Wibisono, I. T. C., & Suryadiputra, I. N. N. (2011). State of the art information on
Chmura, G. L., Anisfeld, S. C., Cahoon, D. R., & Lynch, J. C. (2003). Global carbon sequestra- mangrove ecosystems in Indonesia. Bogor: Wetlands International–Indonesia
tion in tidal, saline wetland soils. Global Biogeochemical Cycles, 17. Programme.
Clark, M. L., Clark, D. B., & Roberts, D. A. (2004). Small-footprint lidar estimation of sub- Irons, J. R., Dwyer, J. L., & Barsi, J. A. (2012). The next Landsat satellite: The Landsat data
canopy elevation and tree height in a tropical rain forest landscape. Remote Sensing continuity mission. Remote Sensing of Environment, 122, 11–21.
of Environment, 91, 68–89. Jensen, J. R. (2005). Introductory digital image processing 3rd edition. Upper saddle river:
Clough, B., & Scott, K. (1989). Allometric relationships for estimating above-ground bio- Prentice hall.
mass in six mangrove species. Forest Ecology and Management, 27, 117–127. Jensen, J. R., Lin, H., Yang, X., Ramsey, E., III, Davis, B. A., & Thoemke, C. W. (1991). The
Clough, B., & Abdullah, A. (1993). The economic and environmental values of mangrove for- measurement of mangrove characteristics in southwest Florida using SPOT multi-
ests and their present state of conservation in the south-East Asia/Pacific region: Techni- spectral data. Geochemistry International, 6, 13–21.
cal report of the project. (International Society for Mangrove Ecosystems). Kathiresan, K., & Bingham, B. L. (2001). Biology of mangroves and mangrove ecosystems.
Congalton, R. G., & Green, K. (2008). Assessing the accuracy of remotely sensed data: Princi- Advances in Marine Biology, 40, 81–251.
ples and practices. CRC press. Kauffman, J. B., & Donato, D. (2012). Protocols for the measurement, monitoring and
Costanza, R., d'Arge, R., De Groot, R., Farber, S., Grasso, M., Hannon, B., ... Paruelo, J. (1997). reporting of structure, biomass and carbon stocks in mangrove forests. Center for In-
The value of the world's ecosystem services and natural capital. Nature, 387, 253–260. ternational Forestry Research Center (CIFOR) working paper (pp. 86).
A. Aslan et al. / Remote Sensing of Environment 183 (2016) 65–81 81

Kellndorfer, J., Walker, W., Pierce, L., Dobson, C., Fites, J. A., Hunsaker, C., ... Clutter, M. Neukermans, G., Dahdouh-Guebas, F., Kairo, J. G., & Koedam, N. (2008). Mangrove species
(2004). Vegetation height estimation from shuttle radar topography mission and na- and stand mapping in Gazi Bay (Kenya) using Quickbird satellite imagery. Journal of
tional elevation datasets. Remote Sensing of Environment, 93, 339–358. Spatial Science, 53, 75–86.
Koenker, R., & Bassett, G., Jr. (1978). Regression quantiles. Econometrica: journal of the Peet, R. K. (1974). The measurement of species diversity. Annual Review of Ecology and
Econometric Society, 33–50. Systematics, 285–307.
Komiyama, A., Poungparn, S., & Kato, S. (2005). Common allometric equations for esti- Pendleton, L., Donato, D. C., Murray, B. C., Crooks, S., Jenkins, W. A., Sifleet, S., ... Marbà, N.
mating the tree weight of mangroves. Journal of Tropical Ecology, 21, 471–477. (2012). Estimating global “blue carbon” emissions from conversion and degradation
Kovacs, J., Lu, X., Flores-Verdugo, F., Zhang, C., Flores de Santiago, F., & Jiao, X. (2013). Ap- of vegetated coastal ecosystems. PloS One, 7, e43542.
plications of ALOS PALSAR for monitoring biophysical parameters of a degraded black Phillips, O. L., Malhi, Y., Higuchi, N., Laurance, W. F., Núnez, P. V., Vásquez, R. M., ... Brown,
mangrove (Avicennia germinans) forest. ISPRS Journal of Photogrammetry and Remote S. (1998). Changes in the carbon balance of tropical forests: Evidence from long-term
Sensing, 82, 102–111. plots. Science, 282, 439–442.
Kuenzer, C., Bluemel, A., Gebhardt, S., Quoc, T. V., & Dech, S. (2011). Remote sensing of Polidoro, B. A., Carpenter, K. E., Collins, L., Duke, N. C., Ellison, A. M., Ellison, J. C., ... Koedam,
mangrove ecosystems: A review. Remote Sensing, 3, 878–928. N. E. (2010). The loss of species: Mangrove extinction risk and geographic areas of
Kushwaha, S., Dwivedi, R., & Rao, B. (2000). Evaluation of various digital image processing global concern. PloS One, 5, e10095.
techniques for detection of coastal wetlands using ERS-1 SAR data. International Pool, D. J., Snedaker, S. C., & Lugo, A. E. (1977). Structure of mangrove forests in Florida,
Journal of Remote Sensing, 21, 565–579. Puerto Rico, Mexico, and Costa Rica. Biotropica, 195–212.
Lang, M. W., & Kasischke, E. S. (2008). Using C-band synthetic aperture radar data to mon- Pouwer, J. (1970). Mimika land tenure. New Guinea research bulletin. 3. (pp. 24–33). Can-
itor forested wetland hydrology in Maryland's coastal plain, USA. IEEE Transactions on berra: Australian National University.
Geoscience and Remote Sensing, 46, 535–546. Proisy, C., Mougin, E., Fromard, F., & Karam, M. (2000). Interpretation of polarimetric
Lee, J. S. (1980). Digital image enhancement and noise filtering by use of local statistics. radar signatures of mangrove forests. Remote Sensing of Environment, 71, 56–66.
IEEE Transactions on Pattern Analysis and Machine Intelligence, 165–168. Proisy, C., Mougin, E., Fromard, F., Trichon, V., & Karam, M. (2002). On the influence of
Long, B. G., & Skewes, T. D. (1996). A technique for mapping mangroves with Landsat TM canopy structure on the radar backscattering of mangrove forests. International
satellite data and geographic information system. Estuarine, Coastal and Shelf Science, Journal of Remote Sensing, 23, 4197–4210.
43, 373–381. Rahman, A. F., Dragoni, D., Didan, K., Barreto-Munoz, A., & Hutabarat, J. A. (2013). Detect-
Lucas, R., Rebelo, L., Fatoyinbo, L., Rosenqvist, A., Itoh, T., Shimada, M., ... Trettin, C. (2014). ing large scale conversion of mangroves to aquaculture with change point and
Contribution of L-band SAR to systematic global mangrove monitoring. Marine and mixed-pixel analyses of high-fidelity MODIS data. Remote Sensing of Environment,
Freshwater Research. 130, 96–107.
Lucas, R. M., Bunting, P., Clewley, D., Proisy, C., Souza Filho, P. W. M., Viergever, K., ... Ramsey, E. W., & Jensen, J. R. (1996). Remote sensing of mangrove wetlands: Relating can-
Rosenqvist, A. (2009). Characterisation and monitoring of mangroves using ALOS opy spectra to site-specific data. Photogrammetric Engineering and Remote Sensing, 62,
PALSAR data. 939–948.
Lucas, R. M., Mitchell, A. L., Rosenqvist, A., Proisy, C., Melius, A., & Ticehurst, C. (2007). The Ramsey, E. W., III, Nelson, G. A., & Sapkota, S. K. (1998). Classifying coastal resources by
potential of L-band SAR for quantifying mangrove characteristics and change: Case integrating optical and radar imagery and color infrared photography. Mangroves
studies from the tropics. Aquatic Conservation: Marine and Freshwater Ecosystems, and Salt Marshes, 2, 109–119.
17, 245–264. Rasolofoharinoro, M., Blasco, F., Bellan, M., Aizpuru, M., Gauquelin, T., & Denis, J. (1998). A
Lucas, R. M., Ellison, J., Mitchell, A., Donnelly, B., Finlayson, M., & Milne, A. (2002). Use of remote sensing based methodology for mangrove studies in Madagascar.
stereo aerial photography for quantifying changes in the extent and height of man- International Journal of Remote Sensing, 19, 1873–1886.
groves in tropical Australia. Wetlands Ecology and Management, 10, 159–173. de Souza Pereira F., R., Kampel, M., & Cunha-Lignon, M. (2012). Mapping of mangrove for-
Lucas, R. M., Moghaddam, M., & Cronin, N. (2004). Microwave scattering from mixed- ests on the southern coast of São Paulo, Brazil, using synthetic aperture radar data
species forests, Queensland, Australia. IEEE Transactions on Geoscience and Remote from ALOS/PALSAR. IEEE Geoscience and Remote Sensing Letters, 3, 567–576.
Sensing, 42, 2142–2159. Rönnholm, P., Hyyppä, J., Hyyppä, H., Haggrén, H., Yu, X., & Kaartinen, H. (2004). Calibra-
Lugo, A. E., & Snedaker, S. C. (1974). The ecology of mangroves. Annual Review of Ecology tion of laser-derived tree height estimates by means of photogrammetric techniques.
and Systematics, 39–64. Scandinavian Journal of Forest Research, 19, 524–528.
Madsen, S. N., Zebker, H. A., & Martin, J. (1993). Topographic mapping using radar inter- Shimada, M., Isoguchi, O., Tadono, T., & Isono, K. (2009). PALSAR radiometric and geomet-
ferometry: Processing techniques. IEEE Transactions on Geoscience and Remote ric calibration. IEEE Transactions on Geoscience and Remote Sensing, 47, 3915–3932.
Sensing, 31, 246–256. Simard, M., Rivera-Monroy, V. H., Mancera-Pineda, J. E., Castañeda-Moya, E., & Twilley, R.
Masek, J. G., Honzak, M., Goward, S. N., Liu, P., & Pak, E. (2001). Landsat-7 ETM+ as an ob- R. (2008). A systematic method for 3D mapping of mangrove forests based on Shuttle
servatory for land cover: Initial radiometric and geometric comparisons with Radar Topography Mission elevation data, ICEsat/GLAS waveforms and field data: Ap-
Landsat-5 Thematic Mapper. Remote Sensing of Environment, 78, 118–130. plication to Ciénaga Grande de Santa Marta, Colombia. Remote Sensing of
Margono, B. A., Potapov, P. V., Turubanova, S., Stolle, F., & Hansen, M. C. (2014). Primary Environment, 112, 2131–2144.
forest cover loss in Indonesia over 2000–2012. Nature Climate Change, 4, 730–735. Simard, M., Zhang, K., Rivera-Monroy, V. H., Ross, M. S., Ruiz, P. L., Castañeda-Moya, E., ...
Marshall, A. J. (2007). Ecology of Indonesian Papua part two. Tuttle Publishing. Rodriguez, E. (2006). Mapping height and biomass of mangrove forests in Everglades
McLeod, E., Chmura, G. L., Bouillon, S., Salm, R., Björk, M., Duarte, C. M., ... Silliman, B. R. National Park with SRTM elevation data. Photogrammetric Engineering and Remote
(2011). A blueprint for blue carbon: Toward an improved understanding of the role Sensing, 72, 299–311.
of vegetated coastal habitats in sequestering CO2. Frontiers in Ecology and the Suratman, M. N. (2008). Carbon sequestration potential of mangroves in southeast Asia.
Environment, 9, 552–560. Managing forest ecosystems: The challenge of climate change (pp. 297–315). Springer.
Miettinen, J., Shi, C., & Liew, S. C. (2011). Deforestation rates in insular Southeast Asia be- Tomlinson, P. (1986). The botany of mangroves. Cambridge tropical biology series. In (p.
tween 2000 and 2010. Global Change Biology, 17, 2261–2270. 419): Cambridge University Press, Cambridge.
MoF (Ministry of Forestry of Indonesia). (2008). Rekalkulasi Penutupan Lahan (Forest Re- Treuhaft, R. N., Law, B. E., & Asner, G. P. (2004). Forest attributes from radar interferomet-
source Recalculation) Indonesia Tahun 2008. (Badan Planology Kehutanan ric structure and its fusion with optical remote sensing. Bioscience, 54, 561–571.
Departemen Kehutanan Indonesia). Vo, Q. T., Kuenzer, C., Vo, Q. M., Moder, F., & Oppelt, N. (2012). Review of valuation
Mitchard, E. T. A. (2012). Using satellite remote sensing to quantify woody cover and biomass methods for mangrove ecosystem services. Ecological Indicators, 23, 431–446.
across Africa. University of Edinburgh. Wang, L., Sousa, W., & Gong, P. (2004). Integration of object-based and pixel-based clas-
Mougin, E., Proisy, C., Marty, G., Fromard, F., Puig, H., Betoulle, J., & Rudant, J. P. (1999). sification for mapping mangroves with IKONOS imagery. International Journal of
Multifrequency and multipolarization radar backscattering from mangrove forests. Remote Sensing, 25, 5655–5668.
IEEE Transactions on Geoscience and Remote Sensing, 37, 94–102. Wang, C., & Glenn, N. F. (2008). A linear regression method for tree canopy height estima-
Mumby, P. J., Edwards, A. J., Arias-González, J. E., Lindeman, K. C., Blackwell, P. G., Gall, A., ... tion using airborne lidar data. Canadian Journal of Remote Sensing, 34, S217–S227.
Renken, H. (2004). Mangroves enhance the biomass of coral reef fish communities in Ward, J. H., Jr. (1963). Hierarchical grouping to optimize an objective function. Journal of
the Caribbean. Nature, 427, 533–536. the American Statistical Association, 58, 236–244.
Murdiyarso, D., Purbopuspito, J., Kauffman, J. B., Warren, M. W., Sasmito, S. D., Donato, D. Willmott, C. J. (1982). Some comments on the evaluation of model performance. Bulletin
C., ... Kurnianto, S. (2015). The potential of Indonesian mangrove forests for global cli- of the American Meteorological Society, 63, 1309–1313.
mate change mitigation. Nature Climate Change, 5, 1089–1092. Wilson, B. A., & Franklin, S. E. (1992). Characterization of alpine vegetation cover using
Murdiyarso, D., Kauffman, J. B., & Verchot, L. V. (2013). Climate change mitigation strate- satellite remote sensing in the Front Ranges, St. Elias Mountains, Yukon Territory.
gies should include tropical wetlands. Carbon Management, 4, 491–499. Global Ecology and Biogeography Letters, 90–95.
Nascimento, H. E., & Laurance, W. F. (2002). Total aboveground biomass in central Ama- Yu, L., & Gong, P. (2012). Google Earth as a virtual globe tool for Earth science applications
zonian rainforests: A landscape-scale study. Forest Ecology and Management, 168, at the global scale: Progress and perspectives. International Journal of Remote Sensing,
311–321. 33, 3966–3986.
Nascimento, W. R., Jr., Souza-Filho, P. W. M., Proisy, C., Lucas, R. M., & Rosenqvist, A.
(2013). Mapping changes in the largest continuous Amazonian mangrove belt using
object-based classification of multisensor satellite imagery. Coastal and Shelf Science:
Estuarine.

You might also like