[go: up one dir, main page]

Next Issue
Volume 6, March
Previous Issue
Volume 6, January
 
 
remotesensing-logo

Journal Browser

Journal Browser

Remote Sens., Volume 6, Issue 2 (February 2014) – 40 articles , Pages 907-1761

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
148 KiB  
Correction
Correction: Van Beek, J. et al. Stem Water Potential Monitoring in Pear Orchards through WorldView-2 Multispectral Imagery. Remote Sens. 2013, 5, 6647–6666
by Jonathan Van Beek, Laurent Tits, Ben Somers, Pieter Janssens, Wendy Odeurs, Hilde Vandendriessche, Tom Deckers and Pol Coppin
Remote Sens. 2014, 6(2), 1760-1761; https://doi.org/10.3390/rs6021760 - 24 Feb 2014
Viewed by 5968
Abstract
The suitability of high resolution satellite imagery to provide the water status in orchard crops, i.e. stem water potential (Ψstem) was evaluated in [1]. However, the contribution of a number of collaborators was not properly acknowledged. Pieter Janssens, Wendy Odeurs, Hilde [...] Read more.
The suitability of high resolution satellite imagery to provide the water status in orchard crops, i.e. stem water potential (Ψstem) was evaluated in [1]. However, the contribution of a number of collaborators was not properly acknowledged. Pieter Janssens, Wendy Odeurs, Hilde Vandendriessche and Tom Deckers all provided a substantial contribution to the conception and the design of the work. They furthermore had a leading role in the acquisition, processing, analysis, and interpretation of the reference evapotranspiration (ETo) and Ψstem data. The article [1] would not have been possible without their valuable input, and the authors would like to correct the authors list as follows. [...] Full article
4239 KiB  
Article
GIS-Based Roughness Derivation for Flood Simulations: A Comparison of Orthophotos, LiDAR and Crowdsourced Geodata
by Helen Dorn, Michael Vetter and Bernhard Höfle
Remote Sens. 2014, 6(2), 1739-1759; https://doi.org/10.3390/rs6021739 - 24 Feb 2014
Cited by 43 | Viewed by 12518
Abstract
Natural disasters like floods are a worldwide phenomenon and a serious threat to mankind. Flood simulations are applications of disaster control, which are used for the development of appropriate flood protection. Adequate simulations require not only the geometry but also the roughness of [...] Read more.
Natural disasters like floods are a worldwide phenomenon and a serious threat to mankind. Flood simulations are applications of disaster control, which are used for the development of appropriate flood protection. Adequate simulations require not only the geometry but also the roughness of the Earth’s surface, as well as the roughness of the objects hereon. Usually, the floodplain roughness is based on land use/land cover maps derived from orthophotos. This study analyses the applicability of roughness map derivation approaches for flood simulations based on different datasets: orthophotos, LiDAR data, official land use data, OpenStreetMap data and CORINE Land Cover data. Object-based image analysis is applied to orthophotos and LiDAR raster data in order to generate land cover maps, which enable a roughness parameterization. The vertical vegetation structure within the LiDAR point cloud is used to derive an additional floodplain roughness map. Further roughness maps are derived from official land use data, OpenStreetMap and CORINE Land Cover datasets. Six different flood simulations are applied based on one elevation data but with the different roughness maps. The results of the hydrodynamic–numerical models include information on flow velocity and water depth from which the additional attribute flood intensity is calculated of. The results based on roughness maps derived from LiDAR data and OpenStreetMap data are comparable, whereas the results of the other datasets differ significantly. Full article
Show Figures


<p>Location of the study site in the Austrian Alps at the Lake Constance.</p>
Full article ">
<p>Workflow of GIS-based roughness derivation and comparison for flood simulations.</p>
Full article ">
<p>LULC classifications based on (<b>a</b>) orthophotos; (<b>b</b>) nDSM, IFP, ILP; (<b>c</b>) voxel; (<b>d</b>) official land use data; (<b>e</b>) CLC data; (<b>f</b>) OSM data.</p>
Full article ">
<p>Development of the inundation area with time steps (s).</p>
Full article ">
<p>Results of the flood simulation at time step 11,400 s with water depth in meter based on the roughness derivation with (<b>a</b>) orthophotos; (<b>b</b>) nDSM, IFP, ILP; (<b>c</b>) voxel; (<b>d</b>) official land use data; (<b>e</b>) CLC data; (<b>f</b>) OSM data.</p>
Full article ">
<p>Flood intensity with water depth (m) as impact parameter.</p>
Full article ">
215 KiB  
Editorial
Acknowledgement to Reviewers of Remote Sensing in 2013
by Remote Sensing Editorial Office
Remote Sens. 2014, 6(2), 1725-1738; https://doi.org/10.3390/rs6021725 - 24 Feb 2014
Viewed by 6728
Abstract
The publisher and editors of the Remote Sensing would like to express their sincere gratitude to the following reviewers for assessing manuscripts in 2013 for Remote Sensing. [...] Full article
986 KiB  
Article
Evaluating the Ability of NPP-VIIRS Nighttime Light Data to Estimate the Gross Domestic Product and the Electric Power Consumption of China at Multiple Scales: A Comparison with DMSP-OLS Data
by Kaifang Shi, Bailang Yu, Yixiu Huang, Yingjie Hu, Bing Yin, Zuoqi Chen, Liujia Chen and Jianping Wu
Remote Sens. 2014, 6(2), 1705-1724; https://doi.org/10.3390/rs6021705 - 20 Feb 2014
Cited by 513 | Viewed by 34951
Abstract
The nighttime light data records artificial light on the Earth’s surface and can be used to estimate the spatial distribution of the gross domestic product (GDP) and the electric power consumption (EPC). In early 2013, the first global NPP-VIIRS nighttime light data were [...] Read more.
The nighttime light data records artificial light on the Earth’s surface and can be used to estimate the spatial distribution of the gross domestic product (GDP) and the electric power consumption (EPC). In early 2013, the first global NPP-VIIRS nighttime light data were released by the Earth Observation Group of National Oceanic and Atmospheric Administration’s National Geophysical Data Center (NOAA/NGDC). As new-generation data, NPP-VIIRS data have a higher spatial resolution and a wider radiometric detection range than the traditional DMSP-OLS nighttime light data. This study aims to investigate the potential of NPP-VIIRS data in modeling GDP and EPC at multiple scales through a case study of China. A series of preprocessing procedures are proposed to reduce the background noise of original data and to generate corrected NPP-VIIRS nighttime light images. Subsequently, linear regression is used to fit the correlation between the total nighttime light (TNL) (which is extracted from corrected NPP-VIIRS data and DMSP-OLS data) and the GDP and EPC (which is from the country’s statistical data) at provincial- and prefectural-level divisions of mainland China. The result of the linear regression shows that R2 values of TNL from NPP-VIIRS with GDP and EPC at multiple scales are all higher than those from DMSP-OLS data. This study reveals that the NPP-VIIRS data can be a powerful tool for modeling socioeconomic indicators; such as GDP and EPC. Full article
(This article belongs to the Special Issue Remote Sensing with Nighttime Lights)
Show Figures


<p>The NPP-VIIRS nighttime light data of China in 2012.</p>
Full article ">
<p>The DMSP-OLS nighttime light data of China in 2012.</p>
Full article ">
<p>The original nighttime light data of Beijing, Guangzhou, and Shanghai in 2012.</p>
Full article ">
<p>(<b>a</b>) The corrected NPP-VIIRS nighttime light data of China in 2012. Two regions bounded by red rectangles are Aksu, Xinjiang (sampling area No. 1) and Daqing, Heilongjiang (sampling area NO.2). The (<b>b</b>) original and (<b>c</b>) corrected image of sampling area NO.1; and the (<b>d</b>) original and (<b>e</b>) corrected image of sampling area No. 2. Two typical removed lit areas are pointed by blue arrows.</p>
Full article ">
<p>The scatter diagram of linear regression analysis in provincial units: (<b>a</b>) the total nighttime light (TNL) of corrected NPP-VIIRS data and the gross domestic product (GDP); (<b>b</b>) the TNL of corrected NPP-VIIRS data and the electric power consumption (EPC); (<b>c</b>) the TNL of DMSP-OLS data and the GDP; (<b>d</b>) the TNL of DMSP-OLS data and the EPC; (<b>e</b>) the TNL of original NPP-VIIRS data and the GDP; (<b>f</b>) the TNL of original NPP-VIIRS data and the EPC.</p>
Full article ">
<p>The scatter diagram of linear regression analysis in prefectural units: (<b>a</b>) the total nighttime light (TNL) of corrected NPP-VIIRS data and the gross domestic product (GDP); (<b>b</b>) the TNL of corrected NPP-VIIRS data and the EPC; (<b>c</b>) the TNL of DMSP-OLS data and the GDP; (<b>d</b>) the TNL of DMSP-OLS data and the EPC; (<b>e</b>) the TNL of original NPP-VIIRS data and the GDP; (<b>f</b>) the TNL of original NPP-VIIRS data and the EPC.</p>
Full article ">
<p>Comparison among the DMSP-OLS data, corrected NPP-VIIRS data, and Landsat 8 OLI-TIRS images for three cities in China.</p>
Full article ">
1135 KiB  
Article
A Comparative Analysis of EO-1 Hyperion, Quickbird and Landsat TM Imagery for Fuel Type Mapping of a Typical Mediterranean Landscape
by Giorgos Mallinis, Georgia Galidaki and Ioannis Gitas
Remote Sens. 2014, 6(2), 1684-1704; https://doi.org/10.3390/rs6021684 - 20 Feb 2014
Cited by 37 | Viewed by 10049
Abstract
Forest fires constitute a natural disturbance factor and an agent of environmental change with local to global impacts on Earth’s processes and functions. Accurate knowledge of forest fuel extent and properties can be an effective component for assessing the impacts of possible future [...] Read more.
Forest fires constitute a natural disturbance factor and an agent of environmental change with local to global impacts on Earth’s processes and functions. Accurate knowledge of forest fuel extent and properties can be an effective component for assessing the impacts of possible future wildfires on ecosystem services. Our study aims to evaluate and compare the spectral and spatial information inherent in the EO-1 Hyperion, Quickbird and Landsat TM imagery. The analysis was based on a support vector machine classification approach in order to discriminate and map Mediterranean fuel types. The fuel classification scheme followed a site-specific fuel model within the study area, which is suitable for fire behavior prediction and spatial simulation. The overall accuracy of the Quickbird-based fuel type mapping was higher than 74% with a quantity disagreement of 9% and an allocation disagreement of 17%. Both classifications from the Hyperion and Landsat TM fuel type maps presented approximately 70% overall accuracy and 16% allocation disagreement. The McNemar’s test indicated that the overall accuracy differences between the three produced fuel type maps were not significant (p < 0.05). Based on both overall and individual higher accuracies obtained with the use of the Quickbird image, this study suggests that the high spatial resolution might be more decisive than the high spectral resolution in Mediterranean fuel type mapping. Full article
(This article belongs to the Special Issue Quantifying the Environmental Impact of Forest Fires)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>Overall process diagram of this research study.</p>
Full article ">
<p>Location of the study area, spatial extent of the EO-1 Hyperion, Quickbird, and Landsat TM imagery used in the research.</p>
Full article ">
<p>The EO-1 Hyperion (<b>a</b>); Quickbird (<b>b</b>); and Landsat TM (<b>c</b>) satellite images used in the study. In the lower row, respective subsets of the 3 images (<b>a1</b>–<b>c1</b>) on a larger scale illustrating spatial resolution differences.</p>
Full article ">
<p>Fuel type classification obtained from the EO-1 Hyperion (<b>a</b>); Quickbird (<b>b</b>) and Landsat TM (<b>c</b>) data using the SVMs classification algorithm.</p>
Full article ">
<p>Quantity and allocation disagreement observed for the three fuel type maps, derived upon the EO-1 Hyperion, Quickbird and Landsat TM images.</p>
Full article ">
448 KiB  
Review
A Review of Swidden Agriculture in Southeast Asia
by Peng Li, Zhiming Feng, Luguang Jiang, Chenhua Liao and Jinghua Zhang
Remote Sens. 2014, 6(2), 1654-1683; https://doi.org/10.3390/rs6021654 - 20 Feb 2014
Cited by 99 | Viewed by 21352
Abstract
Swidden agriculture is by far the dominant land use system in the mountainous regions of Southeast Asia (SEA). It provides various valuable subsistence products to local farmers, mostly the poor ethnic minority groups. Controversially, it is also closely connected with a number of [...] Read more.
Swidden agriculture is by far the dominant land use system in the mountainous regions of Southeast Asia (SEA). It provides various valuable subsistence products to local farmers, mostly the poor ethnic minority groups. Controversially, it is also closely connected with a number of environmental issues. With the strengthening regional economic cooperation in SEA, swidden agriculture has experienced drastic transformations into other diverse market-oriented land use types since the 1990s. However, there is very limited information on the basic geographical and demographic data of swidden agriculture and the socio-economic and biophysical effects of the transformations. International programs, such as the Reducing Emissions from Deforestation and forest Degradation (REDD), underscore the importance of monitoring and evaluating swidden agriculture and its transition to reduce carbon emission due to deforestation and forest degradation. In this context, along with the accessibility of Landsat historical imagery, remote sensing based techniques will offer an effective way to detect and monitor the locations and extent of swidden agriculture. Many approaches for investigating fire occurrence and burned area can be introduced for swidden agriculture mapping due to the common feature of fire relatedness. In this review paper, four broad approaches involving spectral signatures, phenological characteristics, statistical theory and landscape ecology were summarized for swidden agriculture delineation. Five research priorities about swidden agriculture involving remote sensing techniques, spatial pattern, change, drivers and impacts were proposed accordingly. To our knowledge, a synthesis review on the remote sensing and outlook on swidden agriculture has not been reported yet. This review paper aims to give a comprehensive overview of swidden agriculture studies in the domains of debated definition, trends, remote sensing methods and outlook research in SEA undertaken in the past two decades. Full article
Show Figures


<p>Spatial distribution of swidden practice (including shifting cultivation and slash-and-burn agriculture) in pan-tropical developing countries. Literature source: [<a href="#b9-remotesensing-06-01654" class="html-bibr">9</a>,<a href="#b11-remotesensing-06-01654" class="html-bibr">11</a>–<a href="#b25-remotesensing-06-01654" class="html-bibr">25</a>].</p>
Full article ">
1104 KiB  
Article
Performance Analysis of MODIS 500-m Spatial Resolution Products for Estimating Chlorophyll-a Concentrations in Oligo- to Meso-Trophic Waters Case Study: Itumbiara Reservoir, Brazil
by Igor Ogashawara, Enner H. Alcântara, Marcelo P. Curtarelli, Marcos Adami, Renata F. F. Nascimento, Arley F. Souza, José L. Stech and Milton Kampel
Remote Sens. 2014, 6(2), 1634-1653; https://doi.org/10.3390/rs6021634 - 20 Feb 2014
Cited by 20 | Viewed by 7993
Abstract
Monitoring chlorophyll-a (chl-a) concentrations is important for the management of water quality, because it is a good indicator of the eutrophication level in an aquatic system. Thus, our main purpose was to develop an alternative technique to monitor chl-a [...] Read more.
Monitoring chlorophyll-a (chl-a) concentrations is important for the management of water quality, because it is a good indicator of the eutrophication level in an aquatic system. Thus, our main purpose was to develop an alternative technique to monitor chl-a in time and space through remote sensing techniques. However, one of the limitations of remote sensing is the resolution. To achieve a high temporal resolution and medium space resolution, we used the Moderate Resolution Imaging Spectroradiometer (MODIS) 500-m reflectance product, MOD09GA, and limnological parameters from the Itumbiara Reservoir. With these data, an empirical (O14a) and semi-empirical (O14b) algorithm were developed. Algorithms were cross-calibrated and validated using three datasets: one for each campaign and a third consisting of a combination of the two individual campaigns. Algorithm O14a produced the best validation with a root mean square error (RMSE) of 30.4%, whereas O14b produced an RMSE of 32.41% using the mixed dataset calibration. O14a was applied to MOD09GA to build a time series for the reservoir for the year of 2009. The time-series analysis revealed that there were occurrences of algal blooms in the summer that were likely related to the additional input of nutrients caused by rainfall runoff. During the winter, however, the few observed algal blooms events were related to periods of atmospheric meteorological variations that represented an enhanced external influence on the processes of mixing and stratification of the water column. Finally, the use of remote sensing techniques can be an important tool for policy makers, environmental managers and the scientific community with which to monitor water quality. Full article
Show Figures


<p>Location and sampling sites in the Itumbiara Reservoir.</p>
Full article ">
<p>R<sub>rs</sub> with the limits of Moderate Resolution Imaging Spectroradiometer (MODIS) Bands 1, 3 and 4.</p>
Full article ">
<p>Filtered O14a time-series for two regions of the reservoir.</p>
Full article ">
<p>Meteorological variables for the period from 1–14 May.</p>
Full article ">
<p>Water column temperature and estimated chl-<span class="html-italic">a</span> time-series.</p>
Full article ">
<p>Estimated chl-<span class="html-italic">a</span> distribution in the Itumbiara Reservoir from O14a.</p>
Full article ">
2585 KiB  
Article
Monitoring Wetlands Ecosystems Using ALOS PALSAR (L-Band, HV) Supplemented by Optical Data: A Case Study of Biebrza Wetlands in Northeast Poland
by Katarzyna Dabrowska-Zielinska, Maria Budzynska, Monika Tomaszewska, Maciej Bartold, Martyna Gatkowska, Iwona Malek, Konrad Turlej and Milena Napiorkowska
Remote Sens. 2014, 6(2), 1605-1633; https://doi.org/10.3390/rs6021605 - 20 Feb 2014
Cited by 40 | Viewed by 9622
Abstract
The aim of the study was to elaborate the remote sensing methods for monitoring wetlands ecosystems. The investigation was carried out during the years 2002–2010 in the Biebrza Wetlands. The meteorological conditions at the test site varied from extremely dry to very wet. [...] Read more.
The aim of the study was to elaborate the remote sensing methods for monitoring wetlands ecosystems. The investigation was carried out during the years 2002–2010 in the Biebrza Wetlands. The meteorological conditions at the test site varied from extremely dry to very wet. The authors propose applying satellite remote sensing data acquired in the optical and microwave spectrums to classify wetlands vegetation habitats for the assessment of vegetation changes and estimation of wetlands’ biophysical properties to improve monitoring of these unique, very often physically impenetrable, areas. The backscattering coefficients (σ°) calculated from ALOS PALSAR FBD (Advanced Land Observing Satellite, Phased Array type L-band Synthetic Aperture Radar, Fine Beam Dual Mode) images registered at cross polarization HV on 12 May 2008 were used to classify the main wetland communities using ground truth observations and the visual interpretation method. As a result, the σ° values were distributed among the six wetlands’ vegetation classes: scrubs, sedges-scrubs, sedges, reeds, sedges-reeds, rushes, and the areas of each community and changes were assessed. Also, the change in the biophysical variable as Leaf Area Index (LAI) is described using the information from PALSAR data. Strong linear relationships have been found between LAI and σ° derived for particular wetland classes, which then were applied to elaborate the maps of LAI distribution. The other variables used to characterize the changing environmental conditions are: surface temperature (Ts) calculated from NOAA AVHRR (National Oceanic and Atmospheric Administration Advanced Very High Resolution Radiometer) and Normalized Difference Vegetation Index (NDVI) from ENVISAT MERIS (ENVIronmental SATellite MEdium Resolution Imaging Spectrometer). Differences of almost double Ts between “dry” and “wet” years were noticed that reflect observed weather conditions. The highest values of NDVI occurred in years with a sufficient amount of precipitation with the lowest in “dry” years. NDVI values variances within the same wetlands class resulted mainly from the differences in soil moisture. The results of this study show that the satellite data from microwave and optical spectrum gave the repetitive spatial information about vegetation growth conditions and could be used for monitoring wetland ecosystems. Full article
(This article belongs to the Special Issue Remote Sensing of Peatlands I)
Show Figures


<p>Test site on MERIS RGB (7, 5, 2) composition with transects A–B and C–D marked by the red line for which chosen AVHRR data have been transformed into T<sub>s</sub> and MERIS data into NDVI values.</p>
Full article ">
<p>Accumulated values of air temperature (Acc T).</p>
Full article ">
<p>Accumulated values of precipitation (Acc P).</p>
Full article ">
<p>Red, Green, Blue (RGB) (4, 2, 1) composition of AVHRR image acquired on 24 April 2008.</p>
Full article ">
<p>RGB (10, 5, 3) composition of MERIS image acquired on 25 April 2010.</p>
Full article ">
<p>PALSAR FBD HV image acquired on 12 May 2008.</p>
Full article ">
<p>Map of wetland communities classified based on PALSAR HV image registered on 12 May 2008, placed on CLC classes.</p>
Full article ">
<p>Map of wetland communities classified based on PALSAR HV image registered on 1 May 2010, placed on CLC classes.</p>
Full article ">
<p>Plot of the LAI measured <span class="html-italic">versus</span> estimated from PALSAR HV images.</p>
Full article ">
1145 KiB  
Article
Aerosol Indices Derived from MODIS Data for Indicating Aerosol-Induced Air Pollution
by Junliang He, Yong Zha, Jiahua Zhang and Jay Gao
Remote Sens. 2014, 6(2), 1587-1604; https://doi.org/10.3390/rs6021587 - 20 Feb 2014
Cited by 21 | Viewed by 7823
Abstract
Aerosol optical depth (AOD) is a critical variable in estimating aerosol concentration in the atmosphere, evaluating severity of atmospheric pollution, and studying their impact on climate. With the assistance of the 6S radiative transfer model, we simulated apparent reflectancein relation to AOD in [...] Read more.
Aerosol optical depth (AOD) is a critical variable in estimating aerosol concentration in the atmosphere, evaluating severity of atmospheric pollution, and studying their impact on climate. With the assistance of the 6S radiative transfer model, we simulated apparent reflectancein relation to AOD in each Moderate Resolution Imaging Spectroradiometer (MODIS) waveband in this study. The closeness of the relationship was used to identify the most and least sensitive MODIS wavebands. These two bands were then used to construct three aerosol indices (difference, ratio, and normalized difference) for estimating AOD quickly and effectively. The three indices were correlated, respectively, with in situ measured AOD at the Aerosol Robotic NETwork (AERONET) Lake Taihu, Beijing, and Xianghe stations. It is found that apparent reflectance of the blue waveband (band 3) is the most sensitive to AOD while the mid-infrared wavelength (band 7) is the least sensitive. The difference aerosol index is the most accurate in indicating aerosol-induced atmospheric pollution with a correlation coefficient of 0.585, 0.860, 0.685, and 0.333 at the Lake Taihu station, 0.721, 0.839, 0.795, and 0.629 at the Beijing station, and 0.778, 0.782, 0.837, and 0.643 at the Xianghe station in spring, summer, autumn and winter, respectively. It is concluded that the newly proposed difference aerosol index can be used effectively to study the level of aerosol-induced air pollution from MODIS satellite imagery with relative ease. Full article
Show Figures


<p>Frequency distribution of AOD recorded over September 2011–Augst 2012 at the Lake Taihu station.</p>
Full article ">
<p>Variation of surface reflectance as recorded in the seven MODIS wavebands in four seasons.</p>
Full article ">
<p>Variation of apparent reflectance with seven MODIS wavebands in four seasons.</p>
Full article ">
<p>Relationship between apparent reflectance simulated using the 6S model with AOD by MODIS wavebands. (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">
<p>Regression results of the three proposed aerosol indices (DAI, RAI, NDAI) against AOD (550 nm) as observed at the Lake Taihu station. (<b>a</b>–<b>c</b>) spring; (<b>d</b>–<b>f</b>) summer; (<b>g</b>–<b>i</b>) autumn; (<b>j</b>–<b>l</b>) winter.</p>
Full article ">
<p>Regression results of the three proposed aerosol indices (DAI, RAI, NDAI) against AOD (550 nm) as observed at the Lake Taihu station. (<b>a</b>–<b>c</b>) spring; (<b>d</b>–<b>f</b>) summer; (<b>g</b>–<b>i</b>) autumn; (<b>j</b>–<b>l</b>) winter.</p>
Full article ">
<p>Regression results of the three proposed aerosol indices (DAI, RAI, NDAI) against AOD (550 nm) as observed at the Beijing station. (<b>a</b>–<b>c</b>) spring; (<b>d</b>–<b>f</b>) summer; (<b>g</b>–<b>i</b>) autumn; (<b>j</b>–<b>l</b>) winter.</p>
Full article ">
<p>Regression results of the three proposed aerosol indices (DAI, RAI, NDAI) against AOD (550 nm) as observed at the Beijing station. (<b>a</b>–<b>c</b>) spring; (<b>d</b>–<b>f</b>) summer; (<b>g</b>–<b>i</b>) autumn; (<b>j</b>–<b>l</b>) winter.</p>
Full article ">
<p>Regression results of the three proposed aerosol indices (DAI, RAI, NDAI) against AOD (550 nm) as observed at the Xianghe station. (<b>a</b>–<b>c</b>) spring; (<b>d</b>–<b>f</b>) summer; (<b>g</b>–<b>i</b>) autumn; (<b>j</b>–<b>l</b>) winter.</p>
Full article ">
<p>Regression results of the three proposed aerosol indices (DAI, RAI, NDAI) against AOD (550 nm) as observed at the Xianghe station. (<b>a</b>–<b>c</b>) spring; (<b>d</b>–<b>f</b>) summer; (<b>g</b>–<b>i</b>) autumn; (<b>j</b>–<b>l</b>) winter.</p>
Full article ">
2189 KiB  
Article
Slope Superficial Displacement Monitoring by Small Baseline SAR Interferometry Using Data from L-band ALOS PALSAR and X-band TerraSAR: A Case Study of Hong Kong, China
by Fulong Chen, Hui Lin and Xianzhi Hu
Remote Sens. 2014, 6(2), 1564-1586; https://doi.org/10.3390/rs6021564 - 20 Feb 2014
Cited by 22 | Viewed by 8612
Abstract
Owing to the development of spaceborne synthetic aperture radar (SAR) platforms, and in particular the increase in the availability of multi-source (multi-band and multi-resolution) data, it is now feasible to design a surface displacement monitoring application using multi-temporal SAR interferometry (MT-InSAR). Landslides have [...] Read more.
Owing to the development of spaceborne synthetic aperture radar (SAR) platforms, and in particular the increase in the availability of multi-source (multi-band and multi-resolution) data, it is now feasible to design a surface displacement monitoring application using multi-temporal SAR interferometry (MT-InSAR). Landslides have high socio-economic impacts in many countries because of potential geo-hazards and heavy casualties. In this study, taking into account the merits of ALOS PALSAR (L-band, good coherence preservation) and TerraSAR (X-band, high resolution and short revisit times) data, we applied an improved small baseline InSAR (SB-InSAR) with 3-D phase unwrapping approach, to monitor slope superficial displacement in Hong Kong, China, a mountainous subtropical zone city influenced by over-urbanization and heavy monsoonal rains. Results revealed that the synergistic use of PALSAR and TerraSAR data produces different outcomes in relation to data reliability and spatial-temporal resolution, and hence could be of significant value for a comprehensive understanding and monitoring of unstable slopes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>Coverage of ALOS PALSAR (marked by the black box) and TerraSAR-X data (marked by the green box) in Hong Kong, indicating the former has larger swath than the latter. Background is the SRTM DEM-shaded image, and Tai Mo Shan is marked by the light green mountain symbol.</p>
Full article ">
<p>Small baseline formation, (<b>a</b>) ALOS PALSAR; (<b>b</b>) TerraSAR. Yellow points indicate the reference imagery for temporal image co-registration.</p>
Full article ">
<p>Estimated Delauney network of interferograms in spatial and temporal domain, (<b>a</b>) ALOS PALSAR; (<b>b</b>) TerraSAR. The yellow point indicates the reference imagery for temporal image co-registration; and red points in (b) indicate two TerraSAR images discarded due to the spatial baseline threshold (150 m).</p>
Full article ">
<p>Geometric relationship between SAR incidence angle (ascending path) and slope physical motion trends; (<b>a</b>) slope facing the sensor demonstrates as mild LOS subsidence (“A”) or LOS uplift (“B”); (<b>b</b>) slope facing away from the sensor demonstrates as moderate (“AA”) to mild (“BB”) LOS subsidence.</p>
Full article ">
<p>(<b>a</b>) ALOS PALSAR-derived surface annual motion rates over Hong Kong using the SB-InSAR method; the black arrows indicate reclamation lands; the red and pink arrows indicate the satellite flight path and looking direction of line of sight, respectively; black rectangles “1-Disneyland, Lantau Island” and “2-Tai Lam, Yuen Long, New Territories” indicate the TerraSAR monitoring sub-regions. Black polygons indicate the original Lam Chau and Check Lap Kok Island; (<b>b</b>) SPOT 5 imagery indicates the land cover of Hong Kong.</p>
Full article ">
<p>SB-InSAR derived LOS surface annual motion rates over Disneyland region superimposed on the DEM-shade imagery. The black cross indicates the location of a coherent pixel (CP) for time series comparison shown in Section 5.3, the red and pink arrows indicate the satellite flight path and looking direction of line of sight, respectively; (<b>a</b>) PALSAR for the observation period from June 2007 to January 2011; (<b>b</b>) TerraSAR for the observation period from October 2008 to December 2010.</p>
Full article ">
<p>SB-InSAR derived LOS surface annual motion rates over Disneyland region superimposed on the DEM-shade imagery. The black cross indicates the location of a coherent pixel (CP) for time series comparison shown in Section 5.3, the red and pink arrows indicate the satellite flight path and looking direction of line of sight, respectively; (<b>a</b>) PALSAR for the observation period from June 2007 to January 2011; (<b>b</b>) TerraSAR for the observation period from October 2008 to December 2010.</p>
Full article ">
<p>SB-InSAR derived LOS surface annual motion rates over Tai Lam region, the red and pink arrows indicate the satellite flight path and looking direction of line of sight, respectively. (<b>a</b>) PALSAR for the observation period from June 2007 to January 2011; (<b>b</b>) TerraSAR-X for the observation period from October 2008 to December 2010.</p>
Full article ">
<p>SB-InSAR derived LOS surface annual motion rates over Tai Lam region, the red and pink arrows indicate the satellite flight path and looking direction of line of sight, respectively. (<b>a</b>) PALSAR for the observation period from June 2007 to January 2011; (<b>b</b>) TerraSAR-X for the observation period from October 2008 to December 2010.</p>
Full article ">
<p>Quantitative validation by means of the cross comparison between ALOS PALSAR and TerraSAR; (<b>a</b>) time series at the point marked by the black cross in <a href="#f6-remotesensing-06-01564" class="html-fig">Figure 6</a>; (<b>b</b>) the scatter plot of annual displacement rates from coincident CPs in <a href="#f6-remotesensing-06-01564" class="html-fig">Figure 6</a> (the point in (a) indicated by the red ellipse).</p>
Full article ">
<p>Quantitative validation by means of the cross comparison between ALOS PALSAR and TerraSAR; (<b>a</b>) time series at the point marked by the black cross in <a href="#f6-remotesensing-06-01564" class="html-fig">Figure 6</a>; (<b>b</b>) the scatter plot of annual displacement rates from coincident CPs in <a href="#f6-remotesensing-06-01564" class="html-fig">Figure 6</a> (the point in (a) indicated by the red ellipse).</p>
Full article ">
<p>Performance evaluation of ALOS PALSAR and TerraSAR for the unstable slope monitoring, black arrows indicate the occurrence of CPs on layovers in this study; (<b>a</b>) PALSAR-derived slope-projected data; and (<b>b</b>) TerraSAR-derived slope-projected data overlapped on the slope map of Disneyland region (suspected unstable slopes marked by black polygons); (<b>c</b>) and (<b>d</b>) are visibility maps of PALSAR, TerraSAR indicating layovers and shadows; (<b>e</b>) high resolution optical imagery from Google Earth (GE) shows the vegetation coverage and exposed surface.</p>
Full article ">
<p>Performance evaluation of ALOS PALSAR and TerraSAR for the unstable slope monitoring, black arrows indicate the occurrence of CPs on layovers in this study; (<b>a</b>) PALSAR-derived slope-projected data; and (<b>b</b>) TerraSAR-derived slope-projected data overlapped on the slope map of Disneyland region (suspected unstable slopes marked by black polygons); (<b>c</b>) and (<b>d</b>) are visibility maps of PALSAR, TerraSAR indicating layovers and shadows; (<b>e</b>) high resolution optical imagery from Google Earth (GE) shows the vegetation coverage and exposed surface.</p>
Full article ">
783 KiB  
Article
Early Detection of Crop Injury from Glyphosate on Soybean and Cotton Using Plant Leaf Hyperspectral Data
by Feng Zhao, Yanbo Huang, Yiqing Guo, Krishna N. Reddy, Matthew A. Lee, Reginald S. Fletcher and Steven J. Thomson
Remote Sens. 2014, 6(2), 1538-1563; https://doi.org/10.3390/rs6021538 - 20 Feb 2014
Cited by 21 | Viewed by 8256
Abstract
In this paper, we aim to detect crop injury from glyphosate, a herbicide, by both traditionally used spectral indices and newly extracted features with leaf hyperspectral reflectance data for non-Glyphosate-Resistant (non-GR) soybean and non-GR cotton. The new features were extracted by canonical analysis [...] Read more.
In this paper, we aim to detect crop injury from glyphosate, a herbicide, by both traditionally used spectral indices and newly extracted features with leaf hyperspectral reflectance data for non-Glyphosate-Resistant (non-GR) soybean and non-GR cotton. The new features were extracted by canonical analysis technique, which could provide the largest separability to distinguish the injured leaves from the healthy ones. Spectral bands used for constructing these new features were selected based on the sensitivity analysis results of a physically-based leaf radiation transfer model (leaf optical PROperty SPECTra model, PROSPECT), which could help extend the effectiveness of these features to a wide range of leaf structures and growing conditions. This approach has been validated with greenhouse measured data acquired in glyphosate treatment experiments. Results indicated that glyphosate injury could be detected by NDVI (Normalized Difference Vegetation Index), RVI (Ratio Vegetation Index), SAVI (Soil Adjusted Vegetation Index), and DVI (Difference Vegetation Index) in 48 h After the Treatment (HAT) for soybean and in 72 HAT for cotton, but the other spectral indices either showed little use for separation, or did not show consistent separation for healthy and injured soybean and cotton. Compared with the traditional spectral indices, the new features were more feasible for the early detection of glyphosate injury, with leaves sprayed with a higher rate of glyphosate solution having larger feature values. This trend became more and more pronounced with time. Leaves sprayed with different glyphosate rates showed some separability 24 HAT using the new features and could be totally distinguished at and beyond 48 HAT for both soybean and cotton. These findings demonstrated the feasibility of applying leaf hyperspectral reflectance measurements for the early detection of glyphosate injury using these newly proposed features. Full article
Show Figures


<p>Sketch of the ASD integrating sphere apparatus.</p>
Full article ">
<p>Sensitivity indices of PROSPECT input parameters simulated by EFAST (Extended Fourier Amplitude Sensitivity Test) method. <b>(a)</b> 400–1,000 nm, <b>(b)</b> 1,000–2,500 nm. FOSI: First Order Sensitivity Index, TSI: Total Sensitivity Index. The spectral positions of these selected bands (479, 508, 654, 673, 750 nm) are marked with vertical dotted lines.</p>
Full article ">
<p>(<b>a</b>) <span class="html-italic">FCA<sub>s</sub></span> variation of soybean leaves of the three groups at 6, 24, 48, 72 HAT. (<b>b</b>) <span class="html-italic">FCA<sub>c</sub></span> variation of cotton leaves of the three groups at 6, 24, 48, 72 HAT. Each point is a mean value of six leaves for the same treatment. Error bar presents the standard deviation of each point.</p>
Full article ">
<p>Relationships between <span class="html-italic">FCA</span> and leaf chlorophyll content (Chl): (<b>a</b>) soybean; (<b>b</b>) cotton.</p>
Full article ">
1245 KiB  
Article
Modeling Accumulated Volume of Landslides Using Remote Sensing and DTM Data
by Zhengchao Chen, Bing Zhang, Yongshun Han, Zhengli Zuo and Xiaoyong Zhang
Remote Sens. 2014, 6(2), 1514-1537; https://doi.org/10.3390/rs6021514 - 19 Feb 2014
Cited by 33 | Viewed by 10592
Abstract
Landslides, like other natural hazards, such as avalanches, floods, and debris flows, may result in a lot of property damage and human casualties. The volume of landslide deposits is a key parameter for landslide studies and disaster relief. Using remote sensing and digital [...] Read more.
Landslides, like other natural hazards, such as avalanches, floods, and debris flows, may result in a lot of property damage and human casualties. The volume of landslide deposits is a key parameter for landslide studies and disaster relief. Using remote sensing and digital terrain model (DTM) data, this paper analyzes errors that can occur in calculating landslide volumes using conventional models. To improve existing models, the mechanisms and laws governing the material deposited by landslides are studied and then the mass balance principle and mass balance line are defined. Based on these ideas, a novel and improved model (Mass Balance Model, MBM) is proposed. By using a parameter called the “height adaptor”, MBM translates the volume calculation into an automatic search for the mass balance line within the scope of the landslide. Due to the use of mass balance constraints and the height adaptor, MBM is much more effective and reliable. A test of MBM was carried out for the case of a typical landslide, triggered by the Wenchuan Earthquake of 12 May 2008. Full article
Show Figures


<p>(<b>a</b>) Side projection of a typical landslide; (<b>b</b>) Photo of the La Conchita Landslide, California, USA, 1996. Photo by R.L. Schuster, US Geological Survey.</p>
Full article ">
<p>(<b>a</b>) Schematic diagram of the automatic search process for the balance line in a slice through a landslide. The red line represents the pre-landslide surface profile; the blue solid line represents the real whereas the dotted lines represent the post-landslide slope profile generated by the DTMs; (<b>b</b>) Schematic diagram of the program termination condition for MBM.</p>
Full article ">
<p>Workflow of the proposed model (MBM).</p>
Full article ">
<p>3D image of the case-study site. The tested landslide is enclosed by the red line. This image provides clear figures of the main scarp ①; the head ②; the toe ③; the minor scarp ④; and the main body ⑤ (Enclosed by the yellow line).</p>
Full article ">
<p>Detailed image of the tested landslide. The landslide lies within the yellow line. The red flags and markers are the points where the field samples were taken. The black rectangular boxes are the height adjustment sub-areas used by AHDM. The red dotted line is the contour line where the elevation difference is zero according to AHDM. The red line is the contour line marking zero elevation difference according to MBM. The pink dotted line is the location of the profile which is shown in Section 3.3.2. The blue thick line is the mass balance line derived using MBM.</p>
Full article ">
<p><b>(a)</b> Shaded relief of the pre-landslide DTM; <b>(b)</b> post-landslide DTM with contours. The landslide lies inside the red line.</p>
Full article ">
<p>Changes in the removed volume (black line) and accumulated volume (pink line) with increasing <span class="html-italic">H<sup>b</sup></span>. This process is also the automatic search process for the mass balance line. The final two pairs of <span class="html-italic">V<sup>+</sup></span> and <span class="html-italic">V</span><sup>−</sup>, used to determine the final result, are also shown.</p>
Full article ">
<p>Histograms of the elevation difference data. Red for after co-registration and green for before co-registration.</p>
Full article ">
<p>DTM profile lines corresponding to the pink dotted line in <a href="#f5-remotesensing-06-01514" class="html-fig">Figure 5</a>. The black line is the pre-landslide DTM profile. The blue line is the post-landslide DTM profile without height adaptor adjustment (HDM is applied). The green line and the red line are, respectively, the post-landslide DTM profile with the height adjusted using AHDM (meanID_all) and MBM.</p>
Full article ">
802 KiB  
Article
Remote Sensing-Based Biomass Estimation and Its Spatio-Temporal Variations in Temperate Grassland, Northern China
by Yunxiang Jin, Xiuchun Yang, Jianjun Qiu, Jinya Li, Tian Gao, Qiong Wu, Fen Zhao, Hailong Ma, Haida Yu and Bin Xu
Remote Sens. 2014, 6(2), 1496-1513; https://doi.org/10.3390/rs6021496 - 19 Feb 2014
Cited by 142 | Viewed by 11723
Abstract
Grassland biomass is essential for maintaining grassland ecosystems. Moreover, biomass is an important characteristic of grassland. In this study, we combined field sampling with remote sensing data and calculated five vegetation indices (VIs). Using this combined information, we quantified a remote sensing estimation [...] Read more.
Grassland biomass is essential for maintaining grassland ecosystems. Moreover, biomass is an important characteristic of grassland. In this study, we combined field sampling with remote sensing data and calculated five vegetation indices (VIs). Using this combined information, we quantified a remote sensing estimation model and estimated biomass in a temperate grassland of northern China. We also explored the dynamic spatio-temporal variation of biomass from 2006 to 2012. Our results indicated that all VIs investigated in the study were strongly correlated with biomass (α < 0.01). The precision of the model for estimating biomass based on ground data and remote sensing was greater than 73%. Additionally, the results of our analysis indicated that the annual average biomass was 11.86 million tons and that the average yield was 604.5 kg/ha. The distribution of biomass exhibited substantial spatial heterogeneity, and the biomass decreased from the eastern portion of the study area to the western portion. The interannual biomass exhibited strong fluctuations during 2006–2012, with a coefficient of variation of 26.95%. The coefficient of variation of biomass differed among the grassland types. The highest coefficient of variation was found for the desert steppe, followed by the typical steppe and the meadow steppe. Full article
Show Figures


<p>Spatial distribution of grassland types and sampling sites for three regions of Xilingol in Inner Mongolia, China. Region I is the meadow steppe region, Region II is the typical steppe region and Region III is the desert steppe region. The numbers of model points are 921 samples, and the numbers of verification points are 213 samples.</p>
Full article ">
<p>Relationship between VI and biomass: Region I, Region II and Region III. Region I and Region II selected the power function model based on NDVI; Region III selected the linear function model based on SAVI.</p>
Full article ">
<p>Relationship between the estimated and actual biomass values: Region I, Region II and Region III. The model precision for the relationship between the estimated biomass and the actual biomass was above 73%.</p>
Full article ">
<p>The 7-year-average grassland biomass between 2006 and 2012 in the Xilingol grassland.</p>
Full article ">
<p>Biomass variation in Xilingol grassland from 2006 to 2012. The green line is the Harvest Year threshold, and the red line is the Lean Year threshold. The black line is the interannual variation of biomass estimation.</p>
Full article ">
<p>The interannual variation of biomass in different grassland types.</p>
Full article ">
<p>Biomass-CV plot for different grassland types. Biomass was calculated as the 7-year-average grassland biomass in different grassland types. CV was calculated from the annual biomass of grassland types from 2006 to 2012.</p>
Full article ">
2104 KiB  
Article
Evaluation of InSAR and TomoSAR for Monitoring Deformations Caused by Mining in a Mountainous Area with High Resolution Satellite-Based SAR
by Donglie Liu, Yunfeng Shao, Zhenguo Liu, Björn Riedel, Andrew Sowter, Wolfgang Niemeier and Zhengfu Bian
Remote Sens. 2014, 6(2), 1476-1495; https://doi.org/10.3390/rs6021476 - 19 Feb 2014
Cited by 43 | Viewed by 12106
Abstract
Interferometric Synthetic Aperture Radar (InSAR) and Differential Interferometric Synthetic Aperture Radar (DInSAR) have shown numerous applications for subsidence monitoring. In the past 10 years, the Persistent Scatterer InSAR (PSI) and Small BAseline Subset (SBAS) approaches were developed to overcome the problem of decorrelation [...] Read more.
Interferometric Synthetic Aperture Radar (InSAR) and Differential Interferometric Synthetic Aperture Radar (DInSAR) have shown numerous applications for subsidence monitoring. In the past 10 years, the Persistent Scatterer InSAR (PSI) and Small BAseline Subset (SBAS) approaches were developed to overcome the problem of decorrelation and atmospheric effects, which are common in interferograms. However, DInSAR or PSI applications in rural areas, especially in mountainous regions, can be extremely challenging. In this study we have employed a combined technique, i.e., SBAS-DInSAR, to a mountainous area that is severely affected by mining activities. In addition, L-band (ALOS) and C-band (ENVISAT) data sets, 21 TerraSAR-X images provided by German Aerospace Center (DLR) with a high resolution have been used. In order to evaluate the ability of TerraSAR-X for mining monitoring, we present a case study of TerraSAR-X SAR images for Subsidence Hazard Boundary (SHB) extraction. The resulting data analysis gives an initial evaluation of InSAR applications within a mountainous region where fast movements and big phase gradients are common. Moreover, the experiment of four-dimension (4-D) Tomography SAR (TomoSAR) for structure monitoring inside the mining area indicates a potential near all-wave monitoring, which is an extension of conventional InSAR. Full article
(This article belongs to the Special Issue Analysis of Remote Sensing Image Data)
Show Figures


<p>Five active coal mines are shown below in red polygon on Google Earth (UTM). Blue and yellow frame indicate coverage of ENVISAT and ALOS images, respectively. The SAR amplitude image indicates the coverage of TerraSAR-X acquisitions.</p>
Full article ">
<p>Geography of Xishan shown by 90-m SRTM DEM. The picture at the upper right corner gives a general view of Xishan’s geography. Tunlan mine and Malan town are the AOIs (Area Of Interest) of SBAS-DInSAR processing and TomoSAR processing that will be dealt with later, respectively. Gujiao city is also one object of TomoSAR test.</p>
Full article ">
<p>Baselines of TerraSAR-X data set.</p>
Full article ">
<p>Displacement maps with clear deformation signals generated from ENVISAT and ALOS data. (<b>a</b>) ALOS. 2009.12.25–2010.02.09 B<sub>⊥</sub> = 467 m; (<b>b</b>) ENVISAT. 2008.01.13–2008.12.28 B<sub>⊥</sub> = −12 m; (<b>c</b>) ENVISAT. 2008.12.28–2009.05.17 B<sub>⊥</sub> = 46 m; (<b>d</b>) ENVISAT. 2009.10.04–2010.02.21 B<sub>⊥</sub> = −42 m. Red color indicates area with subsidence. Image centre coordinates: 37°51′27.90″N, 112°7′35.30″E.</p>
Full article ">
<p>Full-scene PSI results of TerraSAR-X data. The abscissa is longitude when the ordinate is latitude.</p>
Full article ">
<p>Time series of a control point displacement during 220 days derived by GPS and InSAR, respectively.</p>
Full article ">
<p>Surface subsidence trough (Adapted, [<a href="#b20-remotesensing-06-01476" class="html-bibr">20</a>]). ω is the advanced influence angle.</p>
Full article ">
<p>Working face 18207 and other vicinity working faces. Green points are corner reflectors. Blue lines indicate the working face while green lines are the progress line. Several abandoned mines (goaf) in other layers were shown as well. Image center coordinates: 37°53′32.51″N, 112°7′16.85″E.</p>
Full article ">
<p>Time ordered subsidence maps around 18207 working face. Green points are corner reflectors and Green lines are progress lines. The unit of subsidence is mm. Image center coordinates: 37°53′32.51″N, 112°7′16.85″E.</p>
Full article ">
5226 KiB  
Article
A Nine-Year Climatology of Arctic Sea Ice Lead Orientation and Frequency from AMSR-E
by David Bröhan and Lars Kaleschke
Remote Sens. 2014, 6(2), 1451-1475; https://doi.org/10.3390/rs6021451 - 18 Feb 2014
Cited by 34 | Viewed by 9226
Abstract
We infer the fractional coverage of sea ice leads (as concentration) in the Arctic from Advanced Microwave Scanning Radiometer for Earth Observing System (EOS) (AMSR-E) brightness temperatures. The lead concentration resolves leads of at least 3 km in width. We introduce a new [...] Read more.
We infer the fractional coverage of sea ice leads (as concentration) in the Arctic from Advanced Microwave Scanning Radiometer for Earth Observing System (EOS) (AMSR-E) brightness temperatures. The lead concentration resolves leads of at least 3 km in width. We introduce a new algorithm based on the progressive probabilistic Hough transform to automatically infer lead positions and orientations from daily AMSR-E satellite observations. Because the progressive probabilistic Hough transform often detects an identical lead several times the algorithm clusters neighboring leads that belong to one lead position. A first comparison of automatically detected lead positions and orientations with manually detected lead positions and orientations reveals that 57% of the reference leads are correctly determined. Around 11% of automatically detected leads are located where no reference lead occurs. The automatically detected lead orientations are distributed slightly differently from the reference lead orientations. A second comparison of automatically detected leads in the Fram Strait to leads in a wide swath mode Advanced Synthetic Aperture Radar scene shows a good agreement. We provide an Arctic-wide time series of lead orientations for winters from 2002 to 2011. For example, while a lead orientation of 110° with respect to the Greenwich meridian prevails in the Fram Strait, lead orientations in the Beaufort Sea are more isotropically distributed. We find significant preferred lead orientations almost everywhere in the Arctic Ocean when averaged over the entire AMSR-E time series. Full article
(This article belongs to the Special Issue Cryospheric Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>Principle of the Hough transform: An arbitrary point (<span class="html-italic">x′,y′</span>) in a coordinate system is intersected by lines with arbitrary slopes <span class="html-italic">m</span> and y-offsets <span class="html-italic">b</span> (<b>a</b>). All lines are described by one parameter line (<b>b</b>). The parameter lines (<b>c</b>) belong to a chain of corresponding points (<b>d</b>). The points are connected via a line with the slope <span class="html-italic">m′</span> and the y-intersect <span class="html-italic">b′</span> indicated by a red arrow.</p>
Full article ">
<p>Comparison between unclustered leads (red lines) and clustered leads (yellow lines) in the Fram Strait and a regional zoom north of Greenland on 14 March 2011. The lead concentration is depicted in blueish colors. The cluster distance <span class="html-italic">d</span> equals 4 pixels or 25 km.</p>
Full article ">
<p>Principle of the cluster algorithm: Leads (black dotted lines) are reduced to lead centers (black cross, abbreviated to L) in a test data set (<b>a</b>). The numbers in the proximity matrix (<b>b</b>) represent the distance between the lead centers. All distances greater than the maximal distance d = 2 are whitened. Intersecting rows and columns with an entry identify lead clusters belonging to one cluster. The clustered lead centers are depicted as gray squares and the clustered leads as gray lines (a).</p>
Full article ">
<p>Sketch of the algorithm outline.</p>
Full article ">
<p>Comparison between reference leads (red) and clustered leads (yellow) detected by the Hough transform in the Fram Strait on 14 March 2011.</p>
Full article ">
<p>Comparison between reference leads (red) and clustered leads (yellow) detected by the Hough transform in the Beaufort Sea on 14 November 2004.</p>
Full article ">
<p>Comparison between lead orientations detected by the Hough transform with reference lead orientations in a histogram (<b>a</b>) and as probability density functions (<b>b</b>) in the Fram Strait on 14 March 2011. The Gauss curve in (b) is shown as reference.</p>
Full article ">
<p>Comparison between lead orientations detected by the Hough transform with reference lead orientations in a histogram (<b>a</b>) and as probability density functions (<b>b</b>) in the Beaufort Sea on 14 November 2004. The Gauss curve in (b) is shown as reference.</p>
Full article ">
<p>Validation of clustered leads detected by the Hough transform with an ASAR scene shows a high accordance for larger leads. Leads (red lines) inferred from AMSR-E lead concentration (blueish pixels) are compared to ASAR observations north of Greenland (box in <a href="#f2-remotesensing-06-01451" class="html-fig">Figure 2</a>) on 14 March 2011.</p>
Full article ">
1090 KiB  
Article
Vicarious Calibration of Beijing-1 Multispectral Imagers
by Zhengchao Chen, Bing Zhang, Hao Zhang and Wenjuan Zhang
Remote Sens. 2014, 6(2), 1432-1450; https://doi.org/10.3390/rs6021432 - 18 Feb 2014
Cited by 18 | Viewed by 8392
Abstract
For on-orbit calibration of the Beijing-1 multispectral imagers (Beijing-1/MS), a field calibration campaign was performed at the Dunhuang calibration site during September and October of 2008. Based on the in situ data and images from Beijing-1 and Terra/Moderate Resolution Imaging Spectroradiometer (MODIS), three [...] Read more.
For on-orbit calibration of the Beijing-1 multispectral imagers (Beijing-1/MS), a field calibration campaign was performed at the Dunhuang calibration site during September and October of 2008. Based on the in situ data and images from Beijing-1 and Terra/Moderate Resolution Imaging Spectroradiometer (MODIS), three vicarious calibration methods (i.e., reflectance-based, irradiance-based, and cross-calibration) were used to calculate the top-of-atmosphere (TOA) radiance of Beijing-1. An analysis was then performed to determine or identify systematic and accidental errors, and the overall uncertainty was assessed for each individual method. The findings show that the reflectance-based method has an uncertainty of more than 10% if the aerosol optical depth (AOD) exceeds 0.2. The cross-calibration method is able to reach an error level within 7% if the images are selected carefully. The final calibration coefficients were derived from the irradiance-based data for 6 September 2008, with an uncertainty estimated to be less than 5%. Full article
Show Figures


<p>Channel layout of the Beijing-1/MS (<b>Left</b>) and the relationship between the cameras and images of the Beijing-1/MS (<b>Right</b>).</p>
Full article ">
<p>Image of Dunhuang site acquired by Beijing-1 on 13 September 2008. The image is a mosaic of images from bank0 (<b>Right</b>) and bank1 (<b>Left</b>). The area within the blue lines is the overlap. The track of the surface reflectance measurements in Area A is shown in the bottom right of the image.</p>
Full article ">
<p>(<b>a</b>) Average reflectance of Area A acquired on 3, 6, and 13 September, and average reflectance of Area C acquired on 18 October. (<b>b</b>) Ratios of band-weight reflectance of Beijing-1/MS at 0 zenith angle compared to that of other zenith angles at the same azimuth angle. Numbers in the legend refer to the azimuth angle (in degrees).</p>
Full article ">
<p>Aerosol optical depth (AOD) at 550 nm (<b>Left</b>) and diffuse-to-global irradiance ratios (<b>Right</b>) during Dunhuang experiments.</p>
Full article ">
<p>Workflow for vicarious calibration of Beijing-1/MS.</p>
Full article ">
<p>Plots of the DNs and corresponding top-of-atmosphere (TOA) radiance (<span class="html-italic">L<sub>TOA</sub></span>) calibrated via the three vicarious methods.</p>
Full article ">
<p>Results derived by reflectance-based method and irradiance-based method under different aerosol types.</p>
Full article ">
25409 KiB  
Article
Temperature and Snow-Mediated Moisture Controls of Summer Photosynthetic Activity in Northern Terrestrial Ecosystems between 1982 and 2011
by Jonathan Barichivich, Keith R. Briffa, Ranga Myneni, Gerard Van der Schrier, Wouter Dorigo, Compton J. Tucker, Timothy J. Osborn and Thomas M. Melvin
Remote Sens. 2014, 6(2), 1390-1431; https://doi.org/10.3390/rs6021390 - 14 Feb 2014
Cited by 78 | Viewed by 13967
Abstract
Recent warming has stimulated the productivity of boreal and Arctic vegetation by reducing temperature limitations. However, several studies have hypothesized that warming may have also increased moisture limitations because of intensified summer drought severity. Establishing the connections between warming and drought stress has [...] Read more.
Recent warming has stimulated the productivity of boreal and Arctic vegetation by reducing temperature limitations. However, several studies have hypothesized that warming may have also increased moisture limitations because of intensified summer drought severity. Establishing the connections between warming and drought stress has been difficult because soil moisture observations are scarce. Here we use recently developed gridded datasets of moisture variability to investigate the links between warming and changes in available soil moisture and summer vegetation photosynthetic activity at northern latitudes (>45°N) based on the Normalized Difference Vegetation Index (NDVI) since 1982. Moisture and temperature exert a significant influence on the interannual variability of summer NDVI over about 29% (mean r2 = 0.29 ± 0.16) and 43% (mean r2 = 0.25 ± 0.12) of the northern vegetated land, respectively. Rapid summer warming since the late 1980s (~0.7 °C) has increased evapotranspiration demand and consequently summer drought severity, but contrary to earlier suggestions it has not changed the dominant climate controls of NDVI over time. Furthermore, changes in snow dynamics (accumulation and melting) appear to be more important than increased evaporative demand in controlling changes in summer soil moisture availability and NDVI in moisture-sensitive regions of the boreal forest. In boreal North America, forest NDVI declines are more consistent with reduced snowpack rather than with temperature-induced increases in evaporative demand as suggested in earlier studies. Moreover, summer NDVI variability over about 28% of the northern vegetated land is not significantly associated with moisture or temperature variability, yet most of this land shows increasing NDVI trends. These results suggest that changes in snow accumulation and melt, together with other possibly non-climatic factors are likely to play a significant role in modulating regional ecosystem responses to the projected warming and increase in evapotranspiration demand during the coming decades. Full article
(This article belongs to the Special Issue Monitoring Global Vegetation with AVHRR NDVI3g Data (1981-2011))
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>Temporal variability in percentages of summer dry and wet areas in the northern latitudes (&gt;45°N) between 1950 and 2009 based on the scPDSI computed with (red) and without (blue) interannual changes in potential evapotranspiration (PET). (<b>a</b>) dry area (summer scPDSI ≤ <span class="html-italic">−</span>2); (<b>b</b>) wet area (summer scPDSI ≥ 2). Also shown are the mean summer temperature anomalies for the region (gray dotted line) and the long-term mean of the percentage area series based on actual PET (red dashed line).</p>
Full article ">
<p>Comparison by region and continent between summer scPDSI averages with (red) and without (blue) interannual changes in potential evapotranspiration (PET). Mean summer temperature anomalies relative to the period 1961–1990 (dotted line) are shown for each biome. The red crosses in each panel indicate values of the average scPDSI series with actual PET within the bottom 20th percentile, which can be thought as regional summer droughts. Note that the severity of most of these drought events occurring during the warming period over the last two decades has been intensified by increasing evapotranspiration demand. The vertical dotted lines denote the years 1988 or 1997.</p>
Full article ">
<p>Correlation maps between summer NDVI and summer temperature, water supply and soil moisture variability since 1982. (<b>a</b>) Summer air temperature; (<b>b</b>) Summer precipitation; (<b>c</b>) Summer scPDSI; (<b>d</b>) Summer satellite microwave surface soil moisture (MW-SMO). All correlations are based on linearly detrended data and the stippling indicates statistically significant (<span class="html-italic">p</span> &lt; 0.1) values.</p>
Full article ">
<p>Correlation maps between summer NDVI and spring (March–May) temperature, snow water supply and soil moisture variability since 1982. (<b>a</b>) Spring air temperature; (<b>b</b>) Maximum Snow Water Equivalent (SWE); (<b>c</b>) Spring scPDSI. All correlations are based on linearly detrended data and the stippling indicates statistically significant (<span class="html-italic">p</span> &lt; 0.1) values.</p>
Full article ">
<p>Geography of moisture and temperature controls on summer NDVI at northern latitudes during the last 30 years and relationship with major land cover types. (<b>a</b>) Bivariate correlation map between detrended summer (June–August) NDVI and detrended variations in spring-summer (March–August) scPDSI and summer temperature during the period 1982–2009. Light greens indicate a strong moisture limitation (<span class="html-italic">i.e.</span>, strong positive correlation with precipitation and negative correlation with temperature), whilst purple shades indicate a dominant temperature limitation (<span class="html-italic">i.e.</span>, strong positive correlation with temperature and weak correlation with precipitation). The stippling indicates grid boxes where either correlations with temperature or scPDSI are statistically significant (<span class="html-italic">p</span> &lt; 0.1); (<b>b</b>) IGBP land cover classification for the study domain. The black polygons in the maps denote the extent of the boreal forests as defined in this study. Also shown is the present position of the latitudinal treeline (purple line).</p>
Full article ">
<p>Fraction of interannual summer NDVI variance explained by spring (March–May) and summer water supply, soil moisture and temperature during the period 1982–2009. The maps show the <span class="html-italic">R</span><sup>2</sup> for a stepwise multiple linear regression model predicting summer NDVI at each grid box based on: (<b>a</b>) water supply (<span class="html-italic">x</span><sub>1</sub> = peak SWE, <span class="html-italic">x</span><sub>2</sub> = summer precipitation); (<b>b</b>) soil moisture (<span class="html-italic">x</span><sub>1</sub> = spring scPDSI, <span class="html-italic">x</span><sub>2</sub> = summer scPDSI); (<b>c</b>) temperature (<span class="html-italic">x</span><sub>1</sub> = spring temperature, <span class="html-italic">x</span><sub>2</sub> = summer temperature); and (<b>d</b>) soil moisture and temperature (<span class="html-italic">x</span><sub>1</sub> = spring scPDSI, <span class="html-italic">x</span><sub>2</sub> = summer scPDSI, <span class="html-italic">x</span><sub>3</sub> = spring temperature, <span class="html-italic">x</span><sub>4</sub> = summer temperature). All the variables were linearly detrended prior to analysis and only predictors significant at the 90% confidence level were retained in the regression models. Only positive associations between summer NDVI and variables representing water supply and soil moisture were considered. The stippling in c indicates grid boxes where NDVI is inversely associated with temperature. Gray shading denotes non-vegetated areas or areas where climate data were not available.</p>
Full article ">
<p>Maps of additional variance (<span class="html-italic">R</span><sup>2</sup>) explained by the univariate dynamic Kalman filter regression model between NDVI and potential climate drivers compared with a standard least squares linear regression model. <span class="html-italic">R</span><sup>2</sup> gain for summer NDVI regressed onto: (<b>a</b>) spring and summer temperature; (<b>b</b>) spring and summer water supply; and (<b>c</b>) spring and summer soil moisture (scPDSI). A 10-year high-pass filter was applied to the data prior to analysis. <span class="html-italic">R</span><sup>2</sup> values are shown for grid points where the dynamic regression model was selected over the standard fixed model based on the minimum Akaike Information Criteria (AIC). The rectangles show two regions for which time-dependent associations are illustrated in <a href="#f8-remotesensing-06-01390" class="html-fig">Figure 8</a>.</p>
Full article ">
<p>Illustration of the time-dependent association between interannual variability in summer NDVI and summer temperature in the regions indicated by the rectangles in <a href="#f7-remotesensing-06-01390" class="html-fig">Figure 7a</a>. (<b>a</b>) Comparison of spatially averaged time series for the region in northern Canada (top) and the corresponding Kalman filter regression coefficients and pointwise confidence intervals over the period 1982–2011 (bottom). Where any of the confidence limits includes zero, the regression weights are not considered statistically significant at that point in time. The monthly number of station temperature records in the region included in the CRU TS 3.20 dataset is also shown (bottom). The Aqua-MODIS NDVI average for the same region is shown as a dotted line for comparison. The overall correlation between NDVI3g and temperature is displayed along with its significance (*: <span class="html-italic">p</span> &lt; 0.05); (<b>b</b>) Same as (a) but for a region in Siberia.</p>
Full article ">
<p>Linear trends in summer NDVI and dominant climate drivers since 1982. (<b>a</b>) Trends in spring and summer temperature between 1982 and 2011; (<b>b</b>) Trends in maximum SWE and summer scPDSI over the periods 1982–2011 and 1982–2009, respectively; (<b>c</b>) Statistically significant (<span class="html-italic">p</span> &lt; 0.1) trends in summer NDVI between 1982 and 2011. The colored stippling indicates regions where spring-summer moisture (purple) and temperature (blue) variability significantly influence interannual summer NDVI anomalies as shown in <a href="#f6-remotesensing-06-01390" class="html-fig">Figure 6b,c</a>. The black thick line in North America denotes the southern edge of the continuous permafrost region [<a href="#b61-remotesensing-06-01390" class="html-bibr">61</a>].</p>
Full article ">
1637 KiB  
Article
A Comparative Study on Satellite- and Model-Based Crop Phenology in West Africa
by Elodie Vintrou, Agnès Bégué, Christian Baron, Alexandre Saad, Danny Lo Seen and Seydou B. Traoré
Remote Sens. 2014, 6(2), 1367-1389; https://doi.org/10.3390/rs6021367 - 13 Feb 2014
Cited by 30 | Viewed by 8694
Abstract
Crop phenology is essential for evaluating crop production in the food insecure regions of West Africa. The aim of the paper is to study whether satellite observation of plant phenology are consistent with ground knowledge of crop cycles as expressed in agro-simulations. We [...] Read more.
Crop phenology is essential for evaluating crop production in the food insecure regions of West Africa. The aim of the paper is to study whether satellite observation of plant phenology are consistent with ground knowledge of crop cycles as expressed in agro-simulations. We used phenological variables from a MODIS Land Cover Dynamics (MCD12Q2) product and examined whether they reproduced the spatio-temporal variability of crop phenological stages in Southern Mali. Furthermore, a validated cereal crop growth model for this region, SARRA-H (System for Regional Analysis of Agro-Climatic Risks), provided precise agronomic information. Remotely-sensed green-up, maturity, senescence and dormancy MODIS dates were extracted for areas previously identified as crops and were compared with simulated leaf area indices (LAI) temporal profiles generated using the SARRA-H crop model, which considered the main cropping practices. We studied both spatial (eight sites throughout South Mali during 2007) and temporal (two sites from 2002 to 2008) differences between simulated crop cycles and determined how the differences were indicated in satellite-derived phenometrics. The spatial comparison of the phenological indicator observations and simulations showed mainly that (i) the satellite-derived start-of-season (SOS) was detected approximately 30 days before the model-derived SOS; and (ii) the satellite-derived end-of-season (EOS) was typically detected 40 days after the model-derived EOS. Studying the inter-annual difference, we verified that the mean bias was globally consistent for different climatic conditions. Therefore, the land cover dynamics derived from the MODIS time series can reproduce the spatial and temporal variability of different start-of-season and end-of-season crop species. In particular, we recommend simultaneously using start-of-season phenometrics with crop models for yield forecasting to complement commonly used climate data and provide a better estimate of vegetation phenological changes that integrate rainfall variability, land cover diversity, and the main farmer practices. Full article
Show Figures


<p>The synoptic station locations and a map of the crop production systems in South Mali [<a href="#b27-remotesensing-06-01367" class="html-bibr">27</a>].</p>
Full article ">
<p>Four transition dates based on the EVI curvature-change rate from the MODIS data time series. The solid line is an ideal time series for the vegetation index data, and the dashed line is the rate of change in the VI data curvature. The circles indicate transition dates: 1: start-of-season (SOS); 2: start-of-maximum (SMAX); 3: end-of-maximum (EMAX); and 4: end-of-season (EOS), adapted from <a href="#f2-remotesensing-06-01367" class="html-fig">Figure 2</a> in [<a href="#b23-remotesensing-06-01367" class="html-bibr">23</a>].</p>
Full article ">
<p>An example of start-of-season extraction on a national scale (averaged on a 20 × 20 km grid) and Bougouni- and Segou-station scale (defined by a 10 × 10 km polygon) and masked using a 2007 crop map [<a href="#b35-remotesensing-06-01367" class="html-bibr">35</a>].</p>
Full article ">
<p>Examples of model-derived phenometrics (dotted lines) calculated with Zhang’s non-linear functions for two sets of LAI simulations using the SARRA-H crop model for the Kita synoptic station in 2007. The green curve represents the maize LAI simulation (fertilized in the dark, non-fertilized in light); the orange curve represents the Guinea Sorghum LAI simulation (fertilized in the dark, non-fertilized in light). For each fertilized curve, the dotted lines correspond to the following from left to right: (i) start-of-season (SOS); (ii) start-of-maximum (SMAX) of season; (iii) end-of-maximum (EMAX) of season; and (iv) end-of-season (EOS).</p>
Full article ">
<p>Satellite- (pink) and model-derived (violet) phenometrics boxplots were calculated for eight synoptic stations, ranged from north (<b>Top</b>) to south (<b>Bottom</b>), in 2007. The mean signed difference (MSD) is the difference between the model- and satellite-derived phenometric values in days.</p>
Full article ">
<p>Linear regression for the median satellite- and model-derived phenometrics values for the eight stations during 2007. The diagonal dotted lines represent the 1:1 line.</p>
Full article ">
<p>Satellite- (pink) and model-derived phenometrics boxplots (violet) calculated for Segou and Sikasso from 2002 to 2008. The mean signed difference (MSD) corresponds to the difference in days between the model- and satellite-derived phenometrics.</p>
Full article ">
<p>Linear regression of the satellite- and model-derived phenometric median values for Segou and Sikasso from 2002 to 2008. The diagonal dotted lines represent the 1:1 lines. The SOS points represented by the orange circles were inconsistent for the rainfall distribution, which is discussed below (Section 4.4).</p>
Full article ">
<p>Rainfall (mm), LAI simulation (blue curve) and model- (dotted blue lines) and satellite-derived (dotted red lines) phenometric barplots for the two years, which indicate inconsistencies (Segou in 2002 and Sikasso in 2006), including for the LAI millet choho simulation in Segou and sorghum guinea in Sikasso (blue curve). From left to right, the dotted lines correspond to the start-of-season, start-of-maximum of season, end-of-maximum of season, and end-of-season, respectively. The red circle indicates a failed sowing date, and the green circle indicates a successful sowing date.</p>
Full article ">
841 KiB  
Article
Ontology-Based Classification of Building Types Detected from Airborne Laser Scanning Data
by Mariana Belgiu, Ivan Tomljenovic, Thomas J. Lampoltshammer, Thomas Blaschke and Bernhard Höfle
Remote Sens. 2014, 6(2), 1347-1366; https://doi.org/10.3390/rs6021347 - 12 Feb 2014
Cited by 97 | Viewed by 13621
Abstract
Accurate information on urban building types plays a crucial role for urban development, planning, and management. In this paper, we apply Object-Based Image Analysis (OBIA) methods to extract buildings from Airborne Laser Scanner (ALS) data and investigate the possibility of classifying detected buildings [...] Read more.
Accurate information on urban building types plays a crucial role for urban development, planning, and management. In this paper, we apply Object-Based Image Analysis (OBIA) methods to extract buildings from Airborne Laser Scanner (ALS) data and investigate the possibility of classifying detected buildings into “Residential/Small Buildings”, “Apartment Buildings”, and “Industrial and Factory Building” classes by means of domain ontology and machine learning techniques. The buildings objects are classified using exclusively the information computed from the ALS data. To select the relevant features for predicting the classes of interest, the Random Forest classifier has been applied. The ontology-based classification yielded convincing results for the “Residential/Small Buildings” class (F-Measure 97.7%), whereas the “Apartment Buildings” and “Industrial and Factory Buildings” classes achieved less accurate results (F-Measure 60% and 51%, respectively). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
<p>Overview of the methodology followed in this study.</p>
Full article ">
<p>Excerpt of the buildings types hierarchy. The evaluated building classes are defined as subclasses of Urban-Features; The Properties of the buildings are related using the <b>AND</b> and <b>OR</b> operator (intersection and union of the selected properties).</p>
Full article ">
<p>The Variable Importance (VI) by Mean Decrease Accuracy (MDA) and Mean Decrease Gini (MDG) measures.</p>
Full article ">
<p>The results of the classification performed using ontology and Random Forest (RF) classifier. The buildings were extracted from ALS data.</p>
Full article ">
1049 KiB  
Article
Absolute Calibration of Optical Satellite Sensors Using Libya 4 Pseudo Invariant Calibration Site
by Nischal Mishra, Dennis Helder, Amit Angal, Jason Choi and Xiaoxiong Xiong
Remote Sens. 2014, 6(2), 1327-1346; https://doi.org/10.3390/rs6021327 - 12 Feb 2014
Cited by 79 | Viewed by 11044
Abstract
The objective of this paper is to report the improvements in an empirical absolute calibration model developed at South Dakota State University using Libya 4 (+28.55°, +23.39°) pseudo invariant calibration site (PICS). The approach was based on use of the Terra MODIS as [...] Read more.
The objective of this paper is to report the improvements in an empirical absolute calibration model developed at South Dakota State University using Libya 4 (+28.55°, +23.39°) pseudo invariant calibration site (PICS). The approach was based on use of the Terra MODIS as the radiometer to develop an absolute calibration model for the spectral channels covered by this instrument from visible to shortwave infrared. Earth Observing One (EO-1) Hyperion, with a spectral resolution of 10 nm, was used to extend the model to cover visible and near-infrared regions. A simple Bidirectional Reflectance Distribution function (BRDF) model was generated using Terra Moderate Resolution Imaging Spectroradiometer (MODIS) observations over Libya 4 and the resulting model was validated with nadir data acquired from satellite sensors such as Aqua MODIS and Landsat 7 (L7) Enhanced Thematic Mapper (ETM+). The improvements in the absolute calibration model to account for the BRDF due to off-nadir measurements and annual variations in the atmosphere are summarized. BRDF models due to off-nadir viewing angles have been derived using the measurements from EO-1 Hyperion. In addition to L7 ETM+, measurements from other sensors such as Aqua MODIS, UK-2 Disaster Monitoring Constellation (DMC), ENVISAT Medium Resolution Imaging Spectrometer (MERIS) and Operational Land Imager (OLI) onboard Landsat 8 (L8), which was launched in February 2013, were employed to validate the model. These satellite sensors differ in terms of the width of their spectral bandpasses, overpass time, off-nadir-viewing capabilities, spatial resolution and temporal revisit time, etc. The results demonstrate that the proposed empirical calibration model has accuracy of the order of 3% with an uncertainty of about 2% for the sensors used in the study. Full article
Show Figures


<p>Temporal uncertainties of various Saharan PICS.</p>
Full article ">
<p>Terra MODIS and ETM+ image over Libya 4. Upper two images are the full-sized MODIS and ETM+ images. The red rectangle in the bottom images marks the chosen region of interest (ROI) with latitude (min and max): 28.45, 28.64, longitude (min and max): 23.29, 23.4 and the size is about 19.75 km by 22.25 km. The spatial resolution of Terra MODIS (<b>Left</b>) is 250 m and that of Landsat (<b>Right</b>) is 30 m.</p>
Full article ">
<p>Terra MODIS and ETM+ image over Libya 4. Upper two images are the full-sized MODIS and ETM+ images. The red rectangle in the bottom images marks the chosen region of interest (ROI) with latitude (min and max): 28.45, 28.64, longitude (min and max): 23.29, 23.4 and the size is about 19.75 km by 22.25 km. The spatial resolution of Terra MODIS (<b>Left</b>) is 250 m and that of Landsat (<b>Right</b>) is 30 m.</p>
Full article ">
<p>Simple linear BRDF correction model for Libya 4 based on solar zenith angle.</p>
Full article ">
<p>Exponential model to express BRDF due to solar zenith angle as a function of wavelength for Libya 4.</p>
Full article ">
<p>Quadratic model to express BRDF due to view zenith angle as a function of wavelength for Libya 4.</p>
Full article ">
<p>Exponential model to express BRDF due to view zenith angle as a function of wavelength for Libya 4.</p>
Full article ">
<p>Magnitude and phase of cosine function used to model the atmospheric effects as a periodic sinusoid for Libya 4.</p>
Full article ">
<p>EO-1 Hyperion TOA reflectance profile.</p>
Full article ">
<p>Standard deviation of 108 EO-1 Hyperion TOA reflectance profile.</p>
Full article ">
3973 KiB  
Article
Segmentation-Based Filtering of Airborne LiDAR Point Clouds by Progressive Densification of Terrain Segments
by Xiangguo Lin and Jixian Zhang
Remote Sens. 2014, 6(2), 1294-1326; https://doi.org/10.3390/rs6021294 - 7 Feb 2014
Cited by 110 | Viewed by 11152
Abstract
Filtering is one of the core post-processing steps for Airborne Laser Scanning (ALS) point clouds. A segmentation-based filtering (SBF) method is proposed herein. This method is composed of three key steps: point cloud segmentation, multiple echoes analysis, and iterative judgment. Moreover, the third [...] Read more.
Filtering is one of the core post-processing steps for Airborne Laser Scanning (ALS) point clouds. A segmentation-based filtering (SBF) method is proposed herein. This method is composed of three key steps: point cloud segmentation, multiple echoes analysis, and iterative judgment. Moreover, the third step is our main contribution. Particularly, the iterative judgment is based on the framework of the classic progressive TIN (triangular irregular network) densification (PTD) method, but with basic processing unit being a segment rather than a single point. Seven benchmark datasets provided by ISPRS Working Group III/3 are utilized to test the SBF algorithm and the classic PTD method. Experimental results suggest that, compared with the PTD method, the SBF approach is capable of preserving discontinuities of landscapes and removing the lower parts of large objects attached on the ground surface. As a result, the SBF approach is able to reduce omission errors and total errors by 18.26% and 11.47% respectively, which would significantly decrease the cost of manual operation required in post-processing. Full article
(This article belongs to the Special Issue Advances in Geographic Object-Based Image Analysis (GEOBIA))
Show Figures


<p>Typical errors of Axelsson’s progressive TIN (triangular irregular network) densification (PTD) filter. (<b>a</b>) A point cloud around a step edge; (<b>b</b>) Filtering result of the point cloud in (a); (<b>c</b>) A point cloud in urban area with dense vehicles; (<b>d</b>) Filtering result of the point cloud in (c).</p>
Full article ">
<p>Flow chart of the segmentation-based filtering (SBF) method.</p>
Full article ">
<p>Parameters of the point cloud smooth segmentation (PCSS) segmentation. For a seed point with normal vector and a fitted plane across the seed point, select its neighboring points whose 3D distance to the seed point within <span class="html-italic">d'</span>. Among the neighboring points, add the points, whose angle difference to seed points is less than <span class="html-italic">α</span> and whose distance to the fitted plane is less than <span class="html-italic">r</span>, to the segment.</p>
Full article ">
<p>The progress of SBF method and some needed parameters. (<b>a</b>) A point cloud; (<b>b</b>) Result of point cloud segmentation; (<b>c</b>) Selection of ground seed segments; (<b>d</b>) Construction of TIN by the points in (c); (<b>e</b>) Measurement of angle and distance; (<b>f</b>) Mirroring process.</p>
Full article ">
<p>Filtering and results of testing data about CSite1: (<b>a</b>) The remaining point cloud after outlier removal; (<b>b</b>) TIN of the data in (a); (<b>c</b>) Segmentation result of PCSS; (<b>d</b>) Detected vegetation measurements by multiple echoes analysis of segments; (<b>e</b>) Detected ground seed segments colored by the labeling number; (<b>f</b>) Detected ground measurements by SBF method; (<b>g</b>) Detected ground measurements by the classic PTD method; (<b>h</b>) Differences between (f) and (g).</p>
Full article ">
<p>Filtering and results of testing data about CSite1: (<b>a</b>) The remaining point cloud after outlier removal; (<b>b</b>) TIN of the data in (a); (<b>c</b>) Segmentation result of PCSS; (<b>d</b>) Detected vegetation measurements by multiple echoes analysis of segments; (<b>e</b>) Detected ground seed segments colored by the labeling number; (<b>f</b>) Detected ground measurements by SBF method; (<b>g</b>) Detected ground measurements by the classic PTD method; (<b>h</b>) Differences between (f) and (g).</p>
Full article ">
<p>Filtering and results of testing data about CSite1: (<b>a</b>) The remaining point cloud after outlier removal; (<b>b</b>) TIN of the data in (a); (<b>c</b>) Segmentation result of PCSS; (<b>d</b>) Detected vegetation measurements by multiple echoes analysis of segments; (<b>e</b>) Detected ground seed segments colored by the labeling number; (<b>f</b>) Detected ground measurements by SBF method; (<b>g</b>) Detected ground measurements by the classic PTD method; (<b>h</b>) Differences between (f) and (g).</p>
Full article ">
<p>Filtering and results of testing data about CSite2: (<b>a</b>) The remaining point cloud after outlier removal; (<b>b</b>) TIN of the data in (a); (<b>c</b>) Segmentation result of PCSS; (<b>d</b>) Detected vegetation measurements by multiple echoes analysis of segments; (<b>e</b>) Detected ground seed segments colored by the labeling number; (<b>f</b>) Detected ground measurements by SBF method; (<b>g</b>) Detected ground measurements by the classic PTD method; (<b>h</b>) Differences between (f) and (g).</p>
Full article ">
<p>Filtering and results of testing data about CSite2: (<b>a</b>) The remaining point cloud after outlier removal; (<b>b</b>) TIN of the data in (a); (<b>c</b>) Segmentation result of PCSS; (<b>d</b>) Detected vegetation measurements by multiple echoes analysis of segments; (<b>e</b>) Detected ground seed segments colored by the labeling number; (<b>f</b>) Detected ground measurements by SBF method; (<b>g</b>) Detected ground measurements by the classic PTD method; (<b>h</b>) Differences between (f) and (g).</p>
Full article ">
<p>Filtering and results of reference data of Sample 11: (<b>a</b>) The digital surface model (DSM); (<b>b</b>) The reference DEM; (<b>c</b>) The DEM generated from the result of the PTD method; (<b>d</b>) The type I errors, type II errors of the PTD method; (<b>e</b>) The DEM generated from the result of the SBF method; (<b>f</b>) The type I errors, type II errors of the SBF method. Note that both of the DSM and the DEM are TIN-based.</p>
Full article ">
<p>Filtering and results of reference data of Sample 24: (<b>a</b>) The DSM; (<b>b</b>) The reference DEM; (<b>c</b>) The DEM generated from the result of the PTD method; (<b>d</b>) The type I errors, type II errors of the PTD method; (<b>e</b>) The DEM generated from the result of the SBF method; (<b>f</b>) The type I errors, type II errors of the SBF method. Note that both of the DSM and the DEM are TIN-based.</p>
Full article ">
<p>Filtering and results of reference data of Sample 51: (<b>a</b>) The DSM; (<b>b</b>) The reference DEM; (<b>c</b>) The DEM generated from the result of the PTD method; (<b>d</b>) The type I errors, type II errors of the PTD method; (<b>e</b>) The DEM generated from the result of the SBF method; (<b>f</b>) The type I errors, type II errors of the SBF method. Note that both of the DSM and the DEM are TIN-based.</p>
Full article ">
1092 KiB  
Article
Validation of the Two Standard MODIS Satellite Burned-Area Products and an Empirically-Derived Merged Product in South Africa
by Philemon Tsela, Konrad Wessels, Joel Botai, Sally Archibald, Derick Swanepoel, Karen Steenkamp and Philip Frost
Remote Sens. 2014, 6(2), 1275-1293; https://doi.org/10.3390/rs6021275 - 4 Feb 2014
Cited by 54 | Viewed by 10068
Abstract
The 500-m Moderate Resolution Imaging Spectroradiometer (MODIS) burned area products, MCD45A1, MCD64A1, and a merged product were validated across six study sites in South Africa using independently-derived Landsat burned-area reference data during the fire season of 2007. The objectives of this study were [...] Read more.
The 500-m Moderate Resolution Imaging Spectroradiometer (MODIS) burned area products, MCD45A1, MCD64A1, and a merged product were validated across six study sites in South Africa using independently-derived Landsat burned-area reference data during the fire season of 2007. The objectives of this study were to: (i) investigate the likelihood of the improved detection of small burns through an empirically-derived merged product; (ii) quantify the probability of detection by each product using sub-pixel burned area measures; and, (iii) compare the mean percent concurrence of burned pixels between the standard products over a ten-year time series in each site. Results show that MCD45A1 presented higher detection probabilities (i.e., 3.0%–37.9%) for small fractions ≤50%, whereas MCD64A1 appeared more reliable (i.e., 12.0%–89.2%) in detecting large fractions >50% of a burned MODIS pixel, respectively. Overall, the merged product demonstrated improved detection of the burned area in all fractions. This paper also demonstrates that, on average, >50% of MODIS burned pixels temporally concur between the MCD45A1 and MCD64A1 products in each site. These findings have significant implications for fire monitoring in southern Africa and contribute toward the understanding of the range and of the sources of errors present in the MODIS burned area products. Full article
(This article belongs to the Special Issue Analysis of Remote Sensing Image Data)
Show Figures


<p>The biome map showing the location of the validation sites and multi-temporal Landsat 5 Thematic Mapper (TM) acquisitions distributed across the four fire-prone biomes in South Africa.</p>
Full article ">
<p>Spatial analysis of the burned areas interpreted and produced from multi-temporal Landsat TM data over the six validation sites. The burned area patterns show a diversity of burned shapes, from narrow and/or fragmented (0.36, 12.5, 50 ha) to large and compact (&gt;1,000 ha).</p>
Full article ">
<p>Comparison of Landsat (interpreted) burned-area reference data and the MCD45A1 product to derive the <span class="html-italic">Ce</span> and <span class="html-italic">Oe</span> by use of a confusion matrix.</p>
Full article ">
<p>The 500-m MODIS reference grids used to compute the fractional sub-pixel burned areas corresponding to the Landsat TM interpreted burns.</p>
Full article ">
<p>Histograms of MODIS sub-pixel burned area analysis used to depict the relationship between burned area proportions according to Landsat TM and the probability of detection by the three MODIS products over the six validation sites.</p>
Full article ">
<p>Mean percent concurrence of burned pixels between the MCD45A1 and MCD64A1 products based on a ten-year time series, January 2002 to December 2011, across the six study sites.</p>
Full article ">
<p>Histograms of the brightness levels for Landsat reference pixels interpreted as burned from the TM temporal-difference between vegetation index (VI)<sub>date2</sub> and VI<sub>date1</sub> across the selected sites.</p>
Full article ">
2902 KiB  
Article
The Stalled Recovery of the Iraqi Marshes
by Richard H. Becker
Remote Sens. 2014, 6(2), 1260-1274; https://doi.org/10.3390/rs6021260 - 30 Jan 2014
Cited by 11 | Viewed by 8153
Abstract
The Iraqi (Mesopotamian) Marshes, an extensive wetlands system in Iraq, has been heavily impacted by both human and climate forces over the past decades. In the period leading up to the Second Gulf War in 2002, the marshlands were shrinking due to both [...] Read more.
The Iraqi (Mesopotamian) Marshes, an extensive wetlands system in Iraq, has been heavily impacted by both human and climate forces over the past decades. In the period leading up to the Second Gulf War in 2002, the marshlands were shrinking due to both a policy of draining and water diversion in Iraq and construction of dams upstream on the Tigris and Euphrates rivers. Following the war through 2006, this trend was reversed as the diversions were removed and active draining stopped. A combination of MODIS and GRACE datasets were used to determine the change in surface water area (SWA) in the marshes, marshland extent and change in mass both upriver in the Tigris and Euphrates watersheds and in the marshlands. Results suggest that the post war dam removal and decreased pumping in 2003 provided only temporary respite for the marshlands (2003–2006 SWA: 1,477 km2 increase (600%), water equivalent depth (WED): +2.0 cm/yr.; 2006–2009: −860 km2 (−41%) WED: −3.9 cm/yr.). Unlike in the period 2003–2006, from 2006 forward the mass variations in the marshes are highly correlated with those in the upper and middle watershed (R = 0.86 and 0.92 respectively), suggesting that any recovery due to that removal is complete, and that all future changes are tied more strongly to any climate changes that will affect recharge in the upper Tigris-Euphrates system. Precipitation changes in the watershed show a reduction of an average of 15% below the 15 yr mean in 2007–2011 This corresponds with published ensemble predictions for the 2071–2099 time period, that suggested similar marshland shrinkage should be expected in that time period. Full article
(This article belongs to the Special Issue Hydrological Remote Sensing)
Show Figures


<p>Extent of the Tigris-Euphrates watershed overlain on MODIS NDVI image showing areas of high vegetation (green). Red Boxes show area of upper watershed (5 × 5 degree box), middle watershed (3 × 3 degree box) and marshes (1 × 1 and 3 × 3 degree box), shown from NW to SE, used in GRACE analysis. Inset shows Tropical Rainfall Measurement Mission (TRMM) precipitation over the watershed is concentrated in the North, through Eastern Turkey, Northern Iraq and Western Iran.</p>
Full article ">
<p>Combined MODIS NDVI-MNDWI images for 2001–2012, over Iraqi Marshes. Open water is shown in blue (MNDWI &gt; 0), high vegetation is shown in green, medium in yellow, and low in brown.</p>
Full article ">
<p>MODIS derived area for open water (blue), high NDVI (green, surrogate for healthy marshland vegetation) in the area shown in <a href="#f2-remotesensing-06-01260" class="html-fig">Figure 2</a>. Water equivalent depth changes derived from GRACE data with annual average signal removed for 1 × 1 degree area centered on the Marshlands is overlain in red showing agreement between GRACE solution and extent of combined open water and high NDVI regions. MODIS bars are centered over the GRACE acquisition for the same month.</p>
Full article ">
<p>Water equivalent depth changes derived from GRACE data for 1 × 1 degree area centered on the Marshlands, a 3 × 3 area centered on the marshlands, a 3 × 3 area in the middle of the T-E watershed, and a 5 × 5 degree area in the upper TE watershed. Standard errors for each measurement are shown in gray. Overlain on these are statistically significant (α = 0.05) best fit model of annual and trend component combined (blue) and linear trend alone (black). Dashed line indicates no trend significantly different from zero slope. Lines show trends in water storage for the time period January 2003–March 2013. Slope and error for each segment are listed in <a href="#t1-remotesensing-06-01260" class="html-table">Table 1</a>.</p>
Full article ">
<p>Average annual rainfall for the Tigris Euphrates watershed from TRMM 3B43 data product. Rainfall is calculated by local water year (August through July of the following year). Mean precipitation line shows average for TRMM (January 1998–October 2013) for the entire basin.</p>
Full article ">
12032 KiB  
Article
Detection of Coal Fire Dynamics and Propagation Direction from Multi-Temporal Nighttime Landsat SWIR and TIR Data: A Case Study on the Rujigou Coalfield, Northwest (NW) China
by Hongyuan Huo, Xiaoguang Jiang, Xianfeng Song, Zhao-Liang Li, Zhuoya Ni and Caixia Gao
Remote Sens. 2014, 6(2), 1234-1259; https://doi.org/10.3390/rs6021234 - 29 Jan 2014
Cited by 49 | Viewed by 11218
Abstract
Coal fires are common and serious phenomena in most coal-producing countries in the world. Coal fires not only burn valuable non-renewable coal reserves but also severely affect the local and global environment. The Rujigou coalfield in Shizuishan City, Ningxia, NW China, is well [...] Read more.
Coal fires are common and serious phenomena in most coal-producing countries in the world. Coal fires not only burn valuable non-renewable coal reserves but also severely affect the local and global environment. The Rujigou coalfield in Shizuishan City, Ningxia, NW China, is well known for being a storehouse of anthracite coal. This coalfield is also known for having more coal fires than most other coalfields in China. In this study, an attempt was made to study the dynamics of coal fires in the Rujigou coalfield, from 2001 to 2007, using multi-temporal nighttime Landsat data. The multi-temporal nighttime short wave infrared (SWIR) data sets based on a fixed thresholding technique were used to detect and monitor the surface coal fires and the nighttime enhanced thematic mapper (ETM+) thermal infrared (TIR) data sets, based on a dynamic thresholding technique, were used to identify the thermal anomalies related to subsurface coal fires. By validating the coal fires identified in the nighttime satellite data and the coal fires extracted from daytime satellite data with the coal fire map (CFM) manufactured by field survey, we found that the results from the daytime satellite data had higher omission and commission errors than the results from the nighttime satellite data. Then, two aspects of coal fire dynamics were analyzed: first, a quantitative analysis of the spatial changes in the extent of coal fires was conducted and the results showed that, from 2001 to 2007, the spatial extent of coal fires increased greatly to an annual average area of 0.167 km2; second, the spreading direction and propagation of coal fires was analyzed and predicted from 2001 to 2007, and these results showed that the coal fires generally spread towards the north or northeast, but also spread in some places toward the east. Full article
Show Figures


<p>Chinese fired coal and Chinese coal production since 1902.</p>
Full article ">
<p>Study area of Rujigou Coal Field: (<b>a</b>) shows the location and direction of study area in Northwest China, (<b>b</b>) shows the Rujigou Coal Field located in the Rujigou district, in Shizuishan city and (<b>c</b>) is a 3-D FCC (False Color Composite) image (generated by coding ETM+7/4/2 in R/G/B) based on Landsat ETM+ data acquired on 12 August 1999, overlaid by coal fire map from local mineral bureau, obtained from a field survey of 2002–2003, Projection: UTM, zone 48 North, WGS 84.</p>
Full article ">
<p>(<b>a</b>) SLC-OFF effects on the thermal band 6 data. (<b>b</b>) Gap-filling with the thermal band 6 data. (<b>c</b>) Image enhancing with Gaussian algorithm. (<b>d</b>) SLC-OFF effects on the SWIR band 6 data. (<b>e</b>) Gap-filling with the SWIR band 6 data. (<b>f</b>) Image enhancing with Gaussian algorithm.</p>
Full article ">
<p>Flowchart of coal fire detection and monitoring in the study.</p>
Full article ">
<p>Subsurface coal fires extracted from the ETM+ thermal band 6 data of 2001, 2002, and 2007. (<b>a</b>) thermal anomalies related subsurface coal fires of 2001; (<b>b</b>) thermal anomalies related subsurface coal fires of 2002; (<b>c</b>) thermal anomalies related subsurface coal fires of 2007; (<b>d</b>) multi-layer thermal anomalies related subsurface coal fires of 2001, 2002, and 2007.</p>
Full article ">
<p>Subsurface coal fires extracted from the ETM+ thermal band 6 data of 2001, 2002, and 2007. (<b>a</b>) thermal anomalies related subsurface coal fires of 2001; (<b>b</b>) thermal anomalies related subsurface coal fires of 2002; (<b>c</b>) thermal anomalies related subsurface coal fires of 2007; (<b>d</b>) multi-layer thermal anomalies related subsurface coal fires of 2001, 2002, and 2007.</p>
Full article ">
<p>Surface coal fires extracted from the ETM+ SWIR band 7 data of 2001, 2002, and 2007. (<b>a</b>) surface coal fires of 2001; (<b>b</b>) surface coal fires of 2002; (<b>c</b>) surface coal fires of 2007; (<b>d</b>) multi-layer surface coal fires of 2001, 2002, and 2007.</p>
Full article ">
<p>Surface coal fires extracted from the ETM+ SWIR band 7 data of 2001, 2002, and 2007. (<b>a</b>) surface coal fires of 2001; (<b>b</b>) surface coal fires of 2002; (<b>c</b>) surface coal fires of 2007; (<b>d</b>) multi-layer surface coal fires of 2001, 2002, and 2007.</p>
Full article ">
<p>Validation of the results of coal fires extracted from the ETM+ data of 2002 by the field survey coal fire map which was developed during the time range of 2002 to 2003. (<b>a</b>) validation of coal fires from 2002 nighttime ETM+ data with CFM by field survey; (<b>b</b>) validation of coal fires from 2002 nighttime ETM+ data with CFM by field survey.</p>
Full article ">
<p>Comparative analysis between the results of subsurface coal fires from thermal band 6 data and the respective results of surface coal fires from SWIR band 7 data. (<b>a</b>) comparative analysis of results of surface coal fires and subsurface coal fires 2001; (<b>b</b>) comparative analysis of results of surface coal fires and subsurface coal fires 2002; (<b>c</b>) comparative analysis of results of surface coal fires and subsurface coal fires 2007; (<b>d</b>) quantify the comparative analysis of results 1 and results 2, from 2001 to 2007.</p>
Full article ">
<p>Delineation and prediction of spreading direction of coal fires based on the results of surface coal fires and the subsurface coal fires respectively extracted from the Landsat ETM+ SWIR band 7 data and the thermal band 6 data.</p>
Full article ">
2912 KiB  
Article
The Generalized Difference Vegetation Index (GDVI) for Dryland Characterization
by Weicheng Wu
Remote Sens. 2014, 6(2), 1211-1233; https://doi.org/10.3390/rs6021211 - 29 Jan 2014
Cited by 155 | Viewed by 25392
Abstract
A large number of vegetation indices have been developed and widely applied in terrestrial ecosystem research in the recent decades. However, a certain limitation was observed while applying these indices in research in dry areas due to their low sensitivity to low vegetation [...] Read more.
A large number of vegetation indices have been developed and widely applied in terrestrial ecosystem research in the recent decades. However, a certain limitation was observed while applying these indices in research in dry areas due to their low sensitivity to low vegetation cover. In this context, the objectives of this study are to develop a new vegetation index, namely, the Generalized Difference Vegetation Index (GDVI), and to examine its applicability to the assessment of dryland environment. Based on the field investigation and crop Leaf Area Index (LAI) measurement, five spring and summer Landsat TM and ETM+ images in the frame with Path/Row number of 174/35, and MODIS (Moderate Resolution Imaging Spectroradiometer) LAI and vegetation indices (VIs) data (MOD15A2 and MOD13Q1), of the same acquisition dates as the Landsat images, were acquired and employed in this study. The results reveal that, despite the same level of correlation with the fractional vegetation cover (FVC) as other VIs, GDVI shows a better correlation with LAI and has higher sensitivity and dynamic range in the low vegetal land cover than other vegetation indices, e.g., the range of GDVI is higher than Normalized Difference Vegetation Index (NDVI),Soil-Adjusted Vegetation Index (SAVI), Enhanced Vegetation Index (EVI), Wide Dynamic Range Vegetation Index (WDRVI), and Soil-Adjusted and Atmospherically Resistant Vegetation Index (SARVI), by 164%–326% in woodland, 185%–720% in olive plantation, and 190%–867% in rangeland. It is, hence, concluded that GDVI is relevant for, and has great potential in, land characterization, as well as land degradation/desertification assessment in dryland environment. Full article
Show Figures


<p>Characteristics of GDVI <span class="html-italic">vs</span>. NDVI ((<b>a</b>) GDVI^2 <span class="html-italic">vs.</span> NDVI, (<b>b</b>) GDVI^3 <span class="html-italic">vs.</span> NDVI and (<b>c</b>) GDVI^4 <span class="html-italic">vs</span>. NDVI).</p>
Full article ">
<p>Location of the test area.</p>
Full article ">
<p>Distribution of the sampling areas taking the Landsat-MODIS image pair of 1 May 2007 as an example.</p>
Full article ">
<p>Logarithmic relationships between vegetation indices and LAI (image pair dated 12 July 2010).</p>
Full article ">
<p>Logarithmic relationships between vegetation indices and LAI (image pair dated 1 May 2007. (<b>a</b>) GDVI <span class="html-italic">vs.</span> LAI, (<b>b</b>) VIs with positive range <span class="html-italic">vs</span>. LAI, and (<b>c</b>) VIs with negative range <span class="html-italic">vs</span>. LAI).</p>
Full article ">
<p>Logarithmic relationships between vegetation indices and LAI (image pair dated 1 May 2007. (<b>a</b>) GDVI <span class="html-italic">vs.</span> LAI, (<b>b</b>) VIs with positive range <span class="html-italic">vs</span>. LAI, and (<b>c</b>) VIs with negative range <span class="html-italic">vs</span>. LAI).</p>
Full article ">
<p>Sensitivity of GDVI <span class="html-italic">vs</span>. other vegetation indices for different biomes (image dated 1 May 2007. (<b>a</b>) GDVI^2 <span class="html-italic">vs.</span> VIs, (<b>b</b>) GDVI^3 <span class="html-italic">vs.</span> VIs, and (<b>c</b>) GDVI^4 <span class="html-italic">vs</span>. VIs).</p>
Full article ">
<p>Sensitivity of GDVI <span class="html-italic">vs</span>. other vegetation indices for different biomes (image dated 1 May 2007. (<b>a</b>) GDVI^2 <span class="html-italic">vs.</span> VIs, (<b>b</b>) GDVI^3 <span class="html-italic">vs.</span> VIs, and (<b>c</b>) GDVI^4 <span class="html-italic">vs</span>. VIs).</p>
Full article ">
<p>Relationships between MODIS Normalized Difference Vegetation Index (NDVI), Enhanced Vegetation Index (EVI) and GDVI and MODIS LAI (1 May 2007).</p>
Full article ">
12276 KiB  
Article
Remotely Sensed Monitoring of Small Reservoir Dynamics: A Bayesian Approach
by Dirk Eilander, Frank O. Annor, Lorenzo Iannini and Nick Van de Giesen
Remote Sens. 2014, 6(2), 1191-1210; https://doi.org/10.3390/rs6021191 - 29 Jan 2014
Cited by 37 | Viewed by 9338
Abstract
Multipurpose small reservoirs are important for livelihoods in rural semi-arid regions. To manage and plan these reservoirs and to assess their hydrological impact at a river basin scale, it is important to monitor their water storage dynamics. This paper introduces a Bayesian approach [...] Read more.
Multipurpose small reservoirs are important for livelihoods in rural semi-arid regions. To manage and plan these reservoirs and to assess their hydrological impact at a river basin scale, it is important to monitor their water storage dynamics. This paper introduces a Bayesian approach for monitoring small reservoirs with radar satellite images. The newly developed growing Bayesian classifier has a high degree of automation, can readily be extended with auxiliary information and reduces the confusion error to the land-water boundary pixels. A case study has been performed in the Upper East Region of Ghana, based on Radarsat-2 data from November 2012 until April 2013. Results show that the growing Bayesian classifier can deal with the spatial and temporal variability in synthetic aperture radar (SAR) backscatter intensities from small reservoirs. Due to its ability to incorporate auxiliary information, the algorithm is able to delineate open water from SAR imagery with a low land-water contrast in the case of wind-induced Bragg scattering or limited vegetation on the land surrounding a small reservoir. Full article
(This article belongs to the Special Issue Earth Observation for Water Resource Management in Africa)
Show Figures


<p>The study area in the Upper East Region of Ghana, overlaid with a base map of small reservoirs in the region.</p>
Full article ">
<p>Flow diagram for the growing Bayesian classifier: first, the seeds are initialized (<b>top right</b>) for which a SAR reservoir image is required (<b>top left</b>); then, the iterative Bayesian classification is performed (<b>right middle</b>); finally, a growing filter is applied (<b>right bottom</b>); the algorithm can readily be extended with additional information (<b>left middle</b>).</p>
Full article ">
<p>Backscatter intensity distributions and scatter plots for the land and water classes from the samples of four distinct small reservoir backscatter scenarios, <span class="html-italic">i. e.,</span> smooth open water, water with vegetation, Bragg scattering and backscatter during a rain event.</p>
Full article ">
<p>Jeffries-Matusita (JM) distances for the samples of three distinct backscatter scenarios from small reservoirs, where the error bars show the mean, minimum and maximum JM distances from the different samples.</p>
Full article ">
<p>Ground truth (yellow line) and delineation (red line) based on the ‘HH, HV’ polarization combination overlaid on Pauli RGB-images, with red colors for double bounce, green for volume scatter and blue for single bounce; note that the different color scales are used for the different Pauli components to enhance the image contrast.</p>
Full article ">
<p>Comparison between classified and ground truth areas of all 29 small reservoir from November 2012, based on the Differential Area Index (DAI) and the Jeffries Matusita (JM) distance.</p>
Full article ">
<p>Rainfall time senes (<b>top</b>), the quality of an acquisition based on the Jeffries Matusita (JM) distance (<b>middle</b>) and the mean backscatter intensity of the minimum delineated small reservoir area (<b>bottom</b>), where the boxplots show the median, first and second quartile boundaries and the red crosses are outliers; the error bars show the mean and one standard deviation boundaries.</p>
Full article ">
<p>Time series of small reservoir delineation based on the ‘HH, HV’ polarization combination and the basic gBC (red line), the gBC updated with temporal prior <span class="html-italic">τ<sup>t−</sup></span><sup>1</sup> (blue line) and the gBC updated with both priors <span class="html-italic">τ<sup>t−</sup></span><sup>1</sup> and <span class="html-italic">τ<sup>t+</sup></span><sup>1</sup> (green line) overlaid on HH backscatter intensity images; the bottom graphs show the areal variation in time for the same reservoirs, where the crosses show the filtered time series without the rain-affected March 28 acquisition.</p>
Full article ">
<p>Time series of the cumulative classified area for 26 small reservoirs, based on the growing Bayesian classifier (gBC) with and without temporal priors (<b>top</b>) and the gBC without temporal priors for different polarization combinations (<b>bottom</b>).</p>
Full article ">
3065 KiB  
Article
An Extended Fourier Approach to Improve the Retrieved Leaf Area Index (LAI) in a Time Series from an Alpine Wetland
by Xingwen Quan, Binbin He, Yong Wang, Zhi Tang and Xing Li
Remote Sens. 2014, 6(2), 1171-1190; https://doi.org/10.3390/rs6021171 - 29 Jan 2014
Cited by 15 | Viewed by 7734
Abstract
An extended Fourier approach was presented to improve the retrieved leaf area index (LAIr) of herbaceous vegetation in a time series from an alpine wetland. The retrieval was performed from the Aqua MODIS 8-day composite surface reflectance product (MYD09Q1) from day [...] Read more.
An extended Fourier approach was presented to improve the retrieved leaf area index (LAIr) of herbaceous vegetation in a time series from an alpine wetland. The retrieval was performed from the Aqua MODIS 8-day composite surface reflectance product (MYD09Q1) from day of year (DOY) 97 to 297 using a look-up table (LUT) based inversion of a two-layer canopy reflectance model (ACRM). To reduce the uncertainty (the ACRM inversion is ill-posed), we used NDVI and NIR images to reduce the influence of the soil background and the priori information to constrain the range of sensitive ACRM parameters determined using the Sobol’s method. Even so the uncertainty caused the LAIr versus time curve to oscillate. To further reduce the uncertainty, a Fourier model was fitted using the periodically LAIr results, obtaining LAIF. We note that the level of precision of the LAIF potentially may increase through removing singular points or decrease if the LAIr data were too noisy. To further improve the precision level of the LAIr, the Fourier model was extended by considering the LAIr uncertainty. The LAIr, the LAI simulated using the Fourier model, and the LAI simulated using the extended Fourier approach (LAIeF) were validated through comparisons with the field measured LAI. The R2 values were 0.68, 0.67 and 0.72, the residual sums of squares (RSS) were 3.47, 3.42 and 3.15, and the root-mean-square errors (RMSE) were 0.31, 0.30 and 0.29, respectively, on DOY 177 (early July 2011). In late August (DOY 233), the R2 values were 0.73, 0.77 and 0.79, the RSS values were 38.96, 29.25 and 27.48, and the RMSE values were 0.94, 0.81 and 0.78, respectively. The results demonstrate that the extended Fourier approach has the potential to increase the level of precision of estimates of the time varying LAI. Full article
Show Figures


<p>Sensitivity of normalized difference vegetation index (NDVI) to leaf area index (LAI) (<b>a</b>) and of NIR reflectance to LAI (<b>b</b>) by running two-layer canopy reflectance model (ACRM) forward at different values of <span class="html-italic">s1</span>.</p>
Full article ">
<p>The study area (Wutumeiren prairie). The color composite Landsat5 image is TM4 (red), TM3 (green), and TM2 (blue). Green points represent the sampling plots from 6 to 9 July 2011, and the yellow points represent sampling plots between 26 and 29 August 2011.</p>
Full article ">
<p>Sensitivity analysis of key input parameters of ACRM at near-infrared (NIR) and RED wavebands using Sobol’s method. TSI stands for total sensitivity index.</p>
Full article ">
<p>Distribution of LAI<sub>r</sub>. Because of the ill-posed inversion problem, LAI<sub>r</sub> is not a single value but a wide range of different frequencies.</p>
Full article ">
<p>LAI<sub>r</sub> compared to the field-measured LAI on DOY 177 (July) (<b>a</b>) and DOY 233 (August) (<b>b</b>) in 2011.</p>
Full article ">
<p>LAI<sub>r</sub> <span class="html-italic">versus</span> day of year (DOY) for three pixels. The solid line is the LAI<sub>r</sub>, and the error bars on each DOY represent <span class="html-italic">LAI<sub>uncertainty</sub></span>. (<b>a</b>) and (<b>b</b>) are the LAI<sub>r</sub> of vegetation near Wutumeiren river; (<b>c</b>) is the LAI<sub>r</sub> of vegetation in drought area. The peak LAI<sub>r</sub> values in the study period decrease from (<b>a</b>) to (<b>b</b>) to (<b>c</b>).</p>
Full article ">
<p>Comparison of the LAI<sub>r</sub> and LAI<sub>F</sub> for three pixels from dense vegetation to sparse vegetation in the time series. (<b>a</b>) and (<b>b</b>) are the LAI<sub>r</sub> of vegetation near Wutumeiren river; (<b>c</b>) is the LAI<sub>r</sub> of vegetation in drought area.</p>
Full article ">
<p>Comparison of the LAI<sub>r</sub> and LAI<sub>F</sub> for three pixels from dense vegetation to sparse vegetation in the time series. (<b>a</b>) and (<b>b</b>) are the LAI<sub>r</sub> of vegetation near Wutumeiren river; (<b>c</b>) is the LAI<sub>r</sub> of vegetation in drought area.</p>
Full article ">
<p>Results of LAI<sub>F</sub> compared to LAI<sub>r</sub> in early July (<b>a</b>) and late August (<b>b</b>) 2011.</p>
Full article ">
<p>Comparison between the LAI<sub>r</sub>, LAI<sub>F</sub>, and LAI<sub>eF</sub> values of three pixels from dense vegetation to sparse vegetation in the time series. (<b>a</b>) and (<b>b</b>) are the LAI<sub>r</sub> of vegetation near Wutumeiren river; (<b>c</b>) is the LAI<sub>r</sub> of vegetation in drought area.</p>
Full article ">
698 KiB  
Article
In-Field Absolute Calibration of Ground and Airborne VIS-NIR-SWIR Hyperspectral Point Spectrometers
by Offer Rozenstein, Adam Devir and Arnon Karnieli
Remote Sens. 2014, 6(2), 1158-1170; https://doi.org/10.3390/rs6021158 - 29 Jan 2014
Cited by 2 | Viewed by 7434
Abstract
Spectrometer calibration and measurements of spectral radiance are often required when performing ground, aerial, and space measurements. While calibrating a spectrometer in the field using an integrating sphere is practically unachievable, calibration against a quartz halogen (QH) lamp is a quite easy and [...] Read more.
Spectrometer calibration and measurements of spectral radiance are often required when performing ground, aerial, and space measurements. While calibrating a spectrometer in the field using an integrating sphere is practically unachievable, calibration against a quartz halogen (QH) lamp is a quite easy and feasible option. We describe a calibration protocol whereby a professional QH lamp, operating with a stabilized current source, is first calibrated in the laboratory against a US National Institute of Standards and Technology (NIST) traceable integrating sphere and, then, used for the field calibration of a spectrometer before a ground or airborne campaign. Another advantage of the lamp over the integrating sphere is its ability to create a continuous calibration curve at the spectrometer resolution, while the integrating sphere is calibrated only for a few discrete wavelengths. A calibrated lamp could also be used for a secondary continuous calibration of an un-calibrated integrating sphere. Full article
Show Figures


<p>The angular dependence of an ASD spectrometer bare optical fiber transmittance. It is apparent that the bare optical fiber has a relatively flat response in the center of the field of view (FOV).</p>
Full article ">
<p>Schematic drawing of the measurement equipment: (<b>A</b>) Blowup of the field of view (FOV) limiter. The bare fiber is attached to the left end of the cylinder. Measurements are in mm; (<b>B</b>) The spectrometer attached to the FOV limiter is aligned in front of an integrating sphere; (<b>C</b>) The spectrometer attached to the FOV limiter is aligned in front of the QH lamp operated by a stabilized direct current (DC) source; (<b>D</b>) The spectrometer attached to the field of view (FOV) FOV limiter is aligned in front of Spectralon panel. An incandescent lamp operated by a stabilized current source is placed near the FOV limiter aperture in front of the Spectralon panel.</p>
Full article ">
<p>Calibration according to integrating sphere radiance.</p>
Full article ">
<p>Calibration according to integrating sphere irradiance.</p>
Full article ">
<p>The signal measured in digital numbers (<span class="html-italic">DN</span>) of the lamp (solid line) and the integrating sphere (dashed line). Radiance of the lamp (dotted line) after employing the calibration coefficients derived by interpolating the discrete calibration points of the integrating sphere (shown in <a href="#f6-remotesensing-06-01158" class="html-fig">Figure 6</a>).</p>
Full article ">
<p>Comparison between the integrating sphere calibration (dots: Represent the discrete wavelengths for which the integrating sphere was calibrated) and the measured lamp radiant intensity (solid line: Derived by interpolating the dots to create a calibration curve). Note that this causes great inaccuracies in areas of water absorption by the barium sulfate inner coating of the integrating sphere since there are not enough calibration points in those areas to capture the absorption features.</p>
Full article ">
<p>Measured <span class="html-italic">vs.</span> calculated lamp radiant intensity. The mismatch between the measured and calculated lamp radiance is corrected when accounting for the spectral changes of the lamp emissivity. The radiance was calculated assuming a temperature of 3,000 °K, a filament area of 3 × 5 mm and an emissivity defined by ε<sub>λ</sub> = 0.745 − 175λ.</p>
Full article ">
<p>Results of integrating sphere calibration by lamp reflectance of a Spectralon panel. The dashed and dotted lines represent calibration curves from two different distances from the lamp to the Spectralon panel (D1 = 23.5 cm, D2 = 49.7 cm). Note that both curves are almost identical, demonstrating that the measuring distance is insignificant. The solid line is the integrating sphere radiance derived by the calibration curve from the last step.</p>
Full article ">
<p>Results of integrating sphere direct calibration from a quartz halogen (QH) lamp. The solid line represents the calibration curve obtained by direct calibration from a QH lamp. The dashed curve is the computed lamp radiance.</p>
Full article ">
3188 KiB  
Article
Mapping and Modelling Spatial Variation in Soil Salinity in the Al Hassa Oasis Based on Remote Sensing Indicators and Regression Techniques
by Amal Allbed, Lalit Kumar and Priyakant Sinha
Remote Sens. 2014, 6(2), 1137-1157; https://doi.org/10.3390/rs6021137 - 29 Jan 2014
Cited by 124 | Viewed by 11717
Abstract
Soil salinity is one of the most damaging environmental problems worldwide, especially in arid and semi-arid regions. An integrated approach using remote sensing in addition to various statistical methods has shown success for developing soil salinity prediction models. The aim of this study [...] Read more.
Soil salinity is one of the most damaging environmental problems worldwide, especially in arid and semi-arid regions. An integrated approach using remote sensing in addition to various statistical methods has shown success for developing soil salinity prediction models. The aim of this study was to develop statistical regression models based on remotely sensed indicators to predict and map spatial variation in soil salinity in the Al Hassa oasis. Different spectral indices were calculated from original bands of IKONOS images. Statistical correlation between field measurements of Electrical Conductivity (EC), spectral indices and IKONOS original bands showed that the Salinity Index (SI) and red band (band 3) had the highest correlation with EC. Combining these two remotely sensed variables into one model yielded the best fit with R2 = 0.65. The results revealed that the high performance of this combined model is attributed to: (i) the spatial resolution of the images; (ii) the great potential of the enhanced images, derived from SI, by enhancing and delineating the spatial variation of soil salinity; and (iii) the superiority of band 3 in retrieving soil salinity features and patterns, which was explained by the high reflectance of the smooth and bright surface crust and the low reflectance of the coarse dark puffy crust. Soil salinity maps generated using the selected model showed that strongly saline soils (>16 dS/m) with variable spatial distribution were the dominant class over the study area. The spatial variability of this class over the investigated areas was attributed to a variety factors, including soil factors, management related factors and climate factors. The results demonstrate that modelling and mapping spatial variation in soil salinity based on regression analysis and remote sensing data is a promising approach, as it facilitates timely detection with a low-cost procedure and allows decision makers to decide what necessary action should be taken in the early stages to prevent soil salinity from becoming prevalent, sustaining agricultural lands and natural ecosystems. Full article
Show Figures


<p>Study area with the location of the study sites.</p>
Full article ">
<p>Scatter plots of predicted <span class="html-italic">vs.</span> measured EC using the developed regression models.</p>
Full article ">
<p>Scatter plots of predicted <span class="html-italic">vs.</span> measured EC using the developed regression models.</p>
Full article ">
<p>Validation of the developed regression models: (<b>a</b>) scatterplots of the predicted <span class="html-italic">vs.</span> measured EC; (<b>b</b>) the histogram of the residuals; (<b>c</b>) the normal plot of the residuals.</p>
Full article ">
<p>Soil salinity maps for different sites in the study region: (<b>a</b>) site one; (<b>b</b>) site two; and (<b>c</b>) site three.</p>
Full article ">
<p>Spectral reflectance of saline soils differs due to surface roughness, crusting and colour.</p>
Full article ">
816 KiB  
Article
Impact of Tree Species on Magnitude of PALSAR Interferometric Coherence over Siberian Forest at Frozen and Unfrozen Conditions
by Christian Thiel and Christiane Schmullius
Remote Sens. 2014, 6(2), 1124-1136; https://doi.org/10.3390/rs6021124 - 28 Jan 2014
Cited by 10 | Viewed by 5774
Abstract
Numerous studies demonstrated the potential of the magnitude of interferometric coherence |γ| for forest growing stock volume (GSV) estimation in boreal forests. Coherence derived from images acquired under frozen conditions proved to be of specific interest. This also applies [...] Read more.
Numerous studies demonstrated the potential of the magnitude of interferometric coherence |γ| for forest growing stock volume (GSV) estimation in boreal forests. Coherence derived from images acquired under frozen conditions proved to be of specific interest. This also applies to PALSAR coherence, although affected by a comparatively large temporal baseline of at least 46 days. However, when working with spaceborne L-band data, acquired under unfrozen conditions, a large spread of |γ| was observed at all GSV levels. This scatter negatively affects the correlation of GSV and |γ|. So far, the impact of tree species on |γ| has rarely been studied in this context, although the different tree geometries are likely to have an impact on volumetric decorrelation. This paper presents the results of a study investigating the impact of tree species on PALSAR coherence employing 36 interferograms. The observations show only a small impact of the tree species on |γ| during frozen conditions. At unfrozen conditions, the impact is about three times larger. Deciduous species (aspen, birch, larch) exhibit the lowest |γ|, while coniferous species (fir, pine) feature the highest |γ|. For example, at unfrozen conditions, the |γ| of fir is 0.15 greater than the |γ| of larch, while the mean |γ| of dense forest is 0.38. Accordingly, the impact of tree species on |γ| under unfrozen conditions causes a portion of the observed spread of the GSV-|γ| relationship. Consequently, when aiming at |γ| based GSV assessment using L-band SAR data acquired during unfrozen conditions, the impact of the species on |γ| needs to be considered. For studies aiming at |γ| based GSV estimation across species, PALSAR data acquired at frozen conditions is preferable. Full article
Show Figures


<p>(<b>a</b>) Map of the study area. Each of the test sites comprises more than 300 stands. Excluding clear-cuts the average growing stock volume (<span class="html-italic">GSV</span>) is 180 m<sup>3</sup>·ha<sup>−1</sup> which corresponds to a tree height of 18 m; (<b>b</b>) Stand-wise canopy proportion of dominating species <span class="html-italic">vs</span>. cumulative proportion of all available stands (in total: 12,243). A stand-wise percentage of the dominating species of 100 means pure stands, a percentage of 80 means that the dominating species covers 80% of the forest stand area. At approximately 90% of the stands the canopy proportion of the dominating tree species is smaller than 80%.</p>
Full article ">
<p>Mean (<b>a</b>) and standard deviation (<b>b</b>) of |<span class="html-italic">γ</span>| for dense forest (200–300 m<sup>3</sup>·ha<sup>−1</sup>)—all sites. Symbols: average (•), standard error (⊥ ⊤), minimum/maximum (–). Mean and standard deviation are higher for non-frozen conditions. The displayed values are based on all 36 interferograms. Each of the 36 |<span class="html-italic">γ</span>| images is one sample in this population (unfrozen: 20 images, frozen: 16 images).</p>
Full article ">
<p>|<span class="html-italic">γ</span>| <span class="html-italic">vs. GSV</span> by dominant tree species for Primorsky South using two single interferograms. (<b>a</b>) and (<b>b</b>): Frozen conditions (18 January 2007–5 March 2007); (<b>c</b>) and (<b>d</b>): Non-frozen conditions (21 July 2007–5 September 2007).</p>
Full article ">
<p>|<span class="html-italic">γ</span>| <span class="html-italic">vs. GSV</span> by dominant tree species for Primorsky South using two single interferograms. (<b>a</b>) and (<b>b</b>): Frozen conditions (18 January 2007–5 March 2007); (<b>c</b>) and (<b>d</b>): Non-frozen conditions (21 July 2007–5 September 2007).</p>
Full article ">
<p>Mean (<b>a</b>) and standard deviation (<b>b</b>) of |<span class="html-italic">γ</span>| over dense forest (200–300 m<sup>3</sup>·ha<sup>−1</sup>) by dominant tree species and environmental conditions—all sites. Symbols: average (•), standard deviation (⊥ ⊤), minimum/maximum (–). The figures at the bottom lines of the diagrams represent the numbers of samples.</p>
Full article ">
<p>Deviation Δ of tree species specific |<span class="html-italic">γ</span>| from average |<span class="html-italic">γ</span>| of dense forest for frozen and non-frozen conditions—all sites. The average |<span class="html-italic">γ</span>| of dense forest is shown in <a href="#f2-remotesensing-06-01124" class="html-fig">Figure 2a</a>.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop