[go: up one dir, main page]

Next Issue
Volume 8, December
Previous Issue
Volume 8, October
 
 
remotesensing-logo

Journal Browser

Journal Browser

Remote Sens., Volume 8, Issue 11 (November 2016) – 91 articles

Cover Story (view full-size image): The first operational Sentinel-2 data service platform for obtaining atmospherically-corrected images and generating value-added products, such as maps of leaf area index (LAI). Using the European Space Agency’s (ESA) Sen2Cor algorithm, the platform processes ESA’s Level-1C top-of-atmosphere reflectance to atmospherically-corrected bottom-of-atmosphere (BoA) reflectance (Level-2A). The processing runs on-demand on the Earth Observation Data Centre (EODC), which is a public–private collaborative IT infrastructure in Austria for archiving, processing, and distributing earth observation data. Users can submit processing requests and access the results via a user-friendly web page or use an application programming interface (API). View this paper
  • 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:
4355 KiB  
Article
Climate-Induced Extreme Hydrologic Events in the Arctic
by Toru Sakai, Tsuneo Matsunaga, Shamil Maksyutov, Semen Gotovtsev, Leonid Gagarin, Tetsuya Hiyama and Yasushi Yamaguchi
Remote Sens. 2016, 8(11), 971; https://doi.org/10.3390/rs8110971 - 23 Nov 2016
Cited by 9 | Viewed by 6208
Abstract
The objectives were (i) to evaluate the relationship between recent climate change and extreme hydrological events and (ii) to characterize the behavior of hydrological events along the Alazeya River. The warming rate of air temperature observed at the meteorological station in Chersky was [...] Read more.
The objectives were (i) to evaluate the relationship between recent climate change and extreme hydrological events and (ii) to characterize the behavior of hydrological events along the Alazeya River. The warming rate of air temperature observed at the meteorological station in Chersky was 0.0472 °C·year−1, and an extraordinary increase in air temperatures was observed in 2007. However, data from meteorological stations are somewhat limited in sparsely populated regions. Therefore, this study employed historical remote sensing data for supplementary information. The time-series analysis of the area-averaged Global Precipitation Climatology Project (GPCP) precipitation showed a positive trend because warming leads to an increase in the water vapor content in the atmosphere. In particular, heavy precipitation of 459 ± 113 mm was observed in 2006. On the other hand, the second-highest summer National Oceanic and Atmospheric Administration (NOAA) Advanced Very High Resolution radiometer (AVHRR) brightness temperature (BT) was observed in 2007 when the highest air temperature was observed in Chersky, and the anomaly from normal revealed that the summer AVHRR BTs showed mostly positive values. Conversely, riverbank, lakeshore and seashore areas were much cooler due to the formation, expansion and drainage of lakes and/or the increase in water level by heavy precipitation and melting of frozen ground. The large lake drainage resulted in a flood. Although the flooding was triggered by the thermal erosion along the riverbanks and lakeshores—itself induced by the heat wave in 2007—the increase in soil water content due to the heavy precipitation in 2006 appeared to contribute the magnitude of flood. The flood was characterized by the low streamflow velocity because the Kolyma Lowlands had a very gentle gradient. Therefore, the flood continued for a long time over large areas. Information based on remote sensing data gave basic insights for understanding the mechanism and behavior of climate-induced extreme hydrologic events. Full article
(This article belongs to the Special Issue Remote Sensing of Land Degradation and Drivers of Change)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of study area in the Kolyma lowlands.</p>
Full article ">Figure 2
<p>Time series of mean (<b>a</b>) annual and (<b>b</b>) monthly air temperatures observed at the meteorological station in Chersky for the period 1960–2015. The red line indicates the air temperatures in 2007, and the gray lines show the air temperatures for the other years.</p>
Full article ">Figure 3
<p>Time series of area-averaged (<b>a</b>) brightness temperature (BT) calculated using Advanced Very High Resolution Radiometer (AVHRR) data during the summer period (June–September) and (<b>b</b>) annual precipitation calculated using the Global Precipitation Climatology Project (GPCP) data, which combines meteorological observations and satellite estimates. The average (solid line) and standard deviation (shaded area) of AVHRR summer BT and GPCP annual precipitation data are shown for the region shown in <a href="#remotesensing-08-00971-f001" class="html-fig">Figure 1</a> (67°N–72°N, 147°E–162°E).</p>
Full article ">Figure 4
<p>AVHRR summer BT anomaly in 2007 with respect to the 1982–2015 baseline. Red/blue areas indicate higher/lower than normal summer BT values. Black rectangles show the area of Phased Array type L-band Synthetic Aperture Radar (PALSAR) images shown in <a href="#remotesensing-08-00971-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 5
<p>Inundated areas around three villages on the Alazeya River visualized as a RGB composite from different three-year PALSAR images (R:G:B=2006:2007:2008). (<b>a</b>) Svatai (upper reaches); (<b>b</b>) Argakhtakh (middle reaches); and (<b>c</b>) Andryushkino (lower reaches). Black areas show continuous water bodies (e.g., rivers and lakes) for the three years. Dark-blue, purple and red areas show the inundated areas in 2006, 2007, and 2007 and 2008, respectively. White circles show the location of the villages.</p>
Full article ">Figure 6
<p>Water area detected by PALSAR data during the summer season from 2006 to 2009 in the 20 km × 20 km area shown in <a href="#remotesensing-08-00971-f005" class="html-fig">Figure 5</a>. Blue, green and red dots show the villages of Svatai, Argahtah and Andryushkino, respectively. Closed and open circles show fine and ScanSAR modes of PALSAR, respectively.</p>
Full article ">Figure 7
<p>Lake after drainage.</p>
Full article ">Figure 8
<p>Argakhtakh village surrounded by the flood in 2007.</p>
Full article ">
8015 KiB  
Article
Modeling and Reconstruction of Time Series of Passive Microwave Data by Discrete Fourier Transform Guided Filtering and Harmonic Analysis
by Haolu Shang, Li Jia and Massimo Menenti
Remote Sens. 2016, 8(11), 970; https://doi.org/10.3390/rs8110970 - 23 Nov 2016
Cited by 4 | Viewed by 5627
Abstract
Daily time series of microwave radiometer data obtained in one-orbit direction are full of observation gaps due to satellite configuration and errors from spatial sampling. Such time series carry information about the surface signal including surface emittance and vegetation attenuation, and the atmospheric [...] Read more.
Daily time series of microwave radiometer data obtained in one-orbit direction are full of observation gaps due to satellite configuration and errors from spatial sampling. Such time series carry information about the surface signal including surface emittance and vegetation attenuation, and the atmospheric signal including atmosphere emittance and atmospheric attenuation. To extract the surface signal from this noisy time series, the Time Series Analysis Procedure (TSAP) was developed, based on the properties of the Discrete Fourier Transform (DFT). TSAP includes two stages: (1) identify the spectral features of observation gaps and errors and remove them with a modified boxcar filter; and (2) identify the spectral features of the surface signal and reconstruct it with the Harmonic Analysis of Time Series (HANTS) algorithm. Polarization Difference Brightness Temperature (PDBT) at 37 GHz data were used to illustrate the problems, to explain the implementation of TSAP and to validate this method, due to the PDBT sensitivity to the water content both at the land surface and in the atmosphere. We carried out a case study on a limited heterogeneous crop land and lake area, where the power spectrum of the PDBT time series showed that the harmonic components associated with observation gaps and errors have periods ≤8 days. After applying the modified boxcar filter with a length of 10 days, the RMSD between raw and filtered time series was above 11 K, mainly related to the power reduction in the frequency range associated with observation gaps and errors. Noise reduction is beneficial when applying PDBT observations to monitor wet areas and open water, since the PDBT range between dryland and open water is about 20 K. The spectral features of the atmospheric signal can be revealed by time series analysis of rain-gauge data, since the PDBT at 37 GHz is mainly attenuated by hydrometeors that yield precipitation. Thus, the spectral features of the surface signal were identified in the PDBT time series with the help of the rain-gauge data. HANTS reconstructed the upper envelope of the signal, i.e., correcting for atmospheric influence, while retaining the spectral features of the surface signal. To evaluate the impact of TSAP on retrieval accuracy, the fraction of Water Saturated Surface (WSS) in the region of Poyang Lake was retrieved with 37 GHz observations. The retrievals were evaluated against estimations of the lake area obtained with MODerate-resolution Imaging Spectroradiometer (MODIS) and Advanced Synthetic Aperture Radar (ASAR) data. The Relative RMSE on WSS was 39.5% with unfiltered data and 23% after applying TSAP, i.e., using the estimated surface signal only. Full article
Show Figures

Figure 1

Figure 1
<p>Raw space-borne Polarization Difference Brightness Temperature (PDBT) time series at two sample pixels from 1998 to 2007. All gaps are assigned a value of 0. The dominant land cover in the 1st sample pixel (<b>a</b>) is farm land, centered at 28.603°N and 115.835°E. The dominant land cover in the 2nd sample pixel (<b>b</b>) is open water and wet land, centered at 29.049°N and 116.356°E. Their locations are shown in <a href="#remotesensing-08-00970-f002" class="html-fig">Figure 2</a>. Date format: yyyy/mm/dd.</p>
Full article ">Figure 2
<p>Location of the first (numbered with 1) and the second (numbered with 2) sample pixel in the Poyang Lake floodplain, China.</p>
Full article ">Figure 3
<p>Power spectrum of the space-borne Polarization Difference Brightness Temperature (PDBT) time series for the: 1st (<b>a</b>); and 2nd (<b>b</b>) example pixels from 1998 to 2007. The dominant land cover of the 1st sample pixel is farm land, centered at 28.603°N and 115.835°E. The dominant land cover of the 2nd sample pixel is open water and wet land, centered at 29.049°N and 116.356°E. FFT: Fast Fourier Transform</p>
Full article ">Figure 4
<p>Power spectrum of the square wave in Equation (8) with <span class="html-italic">L</span> = 8 days (<b>a</b>); and power spectrum of the sum of two square waves with <span class="html-italic">L</span> = 7 days and <span class="html-italic">L</span> = 8 days with unit amplitude (<b>b</b>).</p>
Full article ">Figure 4 Cont.
<p>Power spectrum of the square wave in Equation (8) with <span class="html-italic">L</span> = 8 days (<b>a</b>); and power spectrum of the sum of two square waves with <span class="html-italic">L</span> = 7 days and <span class="html-italic">L</span> = 8 days with unit amplitude (<b>b</b>).</p>
Full article ">Figure 5
<p>The Normalized Difference (ND) between amplitudes of harmonic components in Equation (13) and their filtered amplitudes for 10-year time series.</p>
Full article ">Figure 6
<p>The Polarization Difference Brightness Temperature (PDBT) time series after applying the boxcar filter, its HANTS reconstruction and the daily cumulated precipitation from 1998 to 2007 for: the 1st sample pixel (<b>a</b>); and the 2nd sample pixel (<b>b</b>). Date format: yyyy/mm/dd.</p>
Full article ">Figure 6 Cont.
<p>The Polarization Difference Brightness Temperature (PDBT) time series after applying the boxcar filter, its HANTS reconstruction and the daily cumulated precipitation from 1998 to 2007 for: the 1st sample pixel (<b>a</b>); and the 2nd sample pixel (<b>b</b>). Date format: yyyy/mm/dd.</p>
Full article ">Figure 7
<p>The cumulated power spectrum of 10-year precipitation time series for: the 1st sample pixel (<b>a</b>); and the 2nd sample pixel (<b>b</b>).</p>
Full article ">Figure 8
<p>Harmonic components with peak power in the power spectrum of rain-gauge time series in the period range [20 days, 182 days] for both the 1st and the 2nd sample pixels.</p>
Full article ">Figure 9
<p>Harmonic components with main lobe peaks in the power spectrum of the PDBT time series in the period range [20 days, 182 days] for both the 1st and the 2nd sample pixels. The main lobe peaks are defined as the peak values in the upper envelope of the power spectrum.</p>
Full article ">Figure 10
<p>The retrieved Poyang lake area from original PDBT data, filtered PDBT and HANTS reconstructed PDBT, compared with lake area observed with the MODIS and ASAR by [<a href="#B58-remotesensing-08-00970" class="html-bibr">58</a>].</p>
Full article ">Figure 11
<p>The Water Saturated Surface (WSS) area within each 25 km × 25 km pixel on: 7 April (<b>a</b>); 23 April (<b>b</b>); 9 May (<b>c</b>); and 25 May (<b>d</b>) of 2010 in the Dongting Lake and Poyang Lake floodplains; (<b>e</b>) the legend of images from (<b>a</b>) to (<b>d</b>). Dongting Lake is located on the left side and Poyang Lake is located on the right side, with their boundary lines shown in each image.</p>
Full article ">Figure 12
<p>the WSS area within each 25 km × 25 km pixel on: 7 April (<b>a</b>); 23 April (<b>b</b>); 9 May (<b>c</b>); and 25 May (<b>d</b>) of 2011 in the Dongting Lake and Poyang Lake floodplains; (<b>e</b>) the legend of images from (<b>a</b>) to (<b>d</b>). Dongting Lake is located on the left side and Poyang Lake is located on the right side, with their boundary lines shown in each image.</p>
Full article ">
10174 KiB  
Article
Estimation of Building Density with the Integrated Use of GF-1 PMS and Radarsat-2 Data
by Yi Zhou, Chenxi Lin, Shixin Wang, Wenliang Liu and Ye Tian
Remote Sens. 2016, 8(11), 969; https://doi.org/10.3390/rs8110969 - 23 Nov 2016
Cited by 13 | Viewed by 6925
Abstract
Building density, as a component of impervious surface fraction, is a significant indicator of population distribution as essentially all humans live and conduct activities in buildings. Because population spatialization usually occurs over large areas, large-scale building density estimation through a proper, time-efficient, and [...] Read more.
Building density, as a component of impervious surface fraction, is a significant indicator of population distribution as essentially all humans live and conduct activities in buildings. Because population spatialization usually occurs over large areas, large-scale building density estimation through a proper, time-efficient, and relatively precise way is urgently required. Therefore, this study constructed a decision tree by the Classification and Regression Tree (CART) algorithm combining synthetic aperture radar (SAR) with optical images. The input features included four spectral bands (B14) of GF-1 PMS imagery; Normalized Difference Vegetation Index (NDVI), Normalized Difference Water Index (NDWI), and Ratio Built-up Index (RBI) derived from them; and backscatter intensity (BI) of Radarsat-2 SAR data. In addition, a new index called amended backscatter intensity (ABI), which takes the influence created by different spatial patterns into account, was introduced and calculated through fractal dimension and lacunarity. Result showed that before the integration use of multisource data, a model using B14, NDVI, NDWI, and RBI had the highest accuracy, with RMSE of 10.28 and R2 of 0.63 for Jizhou and RMSE of 20.34 and R2 of 0.36 for Beijing. In Comparison, the best model after combining two data sources (i.e., the model employing B14, NDVI, NDWI, RBI and ABI) reduced the RMSE to 8.93 and 16.21 raised the R2 to 0.80 and 0.64, respectively. The result indicated that the synergistic use of optical and SAR data has the potential to improve the building density estimation performance and the addition of ABI has a better capacity for improving the model than other input features. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>left</b>) The geographical location; and (<b>right</b>) the GF-1 PMS images of the studied regions (areas in the red square are used for training and validation).</p>
Full article ">Figure 2
<p>Flowchart of building density modeling and mapping.</p>
Full article ">Figure 3
<p>Principle of pixel aggregate method: (<b>a</b>) input image; (<b>b</b>) output image; and (<b>c</b>) overlap image.</p>
Full article ">Figure 4
<p>Samples with similar density and different distribution pattern: (<b>a</b>) building density: 42.43%, intensity: −15.81; (<b>b</b>) building density: 42.22%, intensity: −13.33; (<b>c</b>) building density: 44.38%, intensity: −15.33; (<b>d</b>) building density: 42.02%, intensity: −9.86; (<b>e</b>) building density: 43.28%, intensity: −16.91; and (<b>f</b>) building density: 44.71%, intensity: −16.24.</p>
Full article ">Figure 5
<p>Triangular prism method: (<b>a</b>) Top view of the corner pixels and the center point used in improved triangular prism method (an example with step size = 4); and (<b>b</b>) 3D view of improved triangular prism method.</p>
Full article ">Figure 6
<p>Illustration of the reiteration procedure to calculate top surface area. (<b>a</b>) <math display="inline"> <semantics> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> <mi>s</mi> <mi>i</mi> <mi>z</mi> <mi>e</mi> <mo>=</mo> <mn>2</mn> <mo>,</mo> <mtext> </mtext> <mi>t</mi> <mi>o</mi> <mi>p</mi> <mtext> </mtext> <mi>s</mi> <mi>u</mi> <mi>r</mi> <mi>f</mi> <mi>a</mi> <mi>c</mi> <mi>e</mi> <mtext> </mtext> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>a</mi> <mo>=</mo> <mstyle displaystyle="true"> <mo>∑</mo> <mrow> <mrow> <mo>(</mo> <mrow> <mn>64</mn> <mo>×</mo> <mi>s</mi> <mi>q</mi> <mi>u</mi> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>s</mi> </mrow> <mo>)</mo> </mrow> </mrow> </mstyle> </mrow> </semantics> </math>; (<b>b</b>) <math display="inline"> <semantics> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> <mi>s</mi> <mi>i</mi> <mi>z</mi> <mi>e</mi> <mo>=</mo> <mn>4</mn> <mo>,</mo> <mtext> </mtext> <mi>t</mi> <mi>o</mi> <mi>p</mi> <mtext> </mtext> <mi>s</mi> <mi>u</mi> <mi>r</mi> <mi>f</mi> <mi>a</mi> <mi>c</mi> <mi>e</mi> <mtext> </mtext> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>a</mi> <mo>=</mo> <mstyle displaystyle="true"> <mo>∑</mo> <mrow> <mrow> <mo>(</mo> <mrow> <mn>16</mn> <mo>×</mo> <mi>s</mi> <mi>q</mi> <mi>u</mi> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>s</mi> </mrow> <mo>)</mo> </mrow> </mrow> </mstyle> </mrow> </semantics> </math>; (<b>c</b>) <math display="inline"> <semantics> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> <mi>s</mi> <mi>i</mi> <mi>z</mi> <mi>e</mi> <mo>=</mo> <mn>8</mn> <mo>,</mo> <mtext> </mtext> <mi>t</mi> <mi>o</mi> <mi>p</mi> <mtext> </mtext> <mi>s</mi> <mi>u</mi> <mi>r</mi> <mi>f</mi> <mi>a</mi> <mi>c</mi> <mi>e</mi> <mtext> </mtext> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>a</mi> <mo>=</mo> <mstyle displaystyle="true"> <mo>∑</mo> <mrow> <mrow> <mo>(</mo> <mrow> <mn>4</mn> <mo>×</mo> <mi>s</mi> <mi>q</mi> <mi>u</mi> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>s</mi> </mrow> <mo>)</mo> </mrow> </mrow> </mstyle> </mrow> </semantics> </math>; and (<b>d</b>) <math display="inline"> <semantics> <mrow> <mi>s</mi> <mi>t</mi> <mi>e</mi> <mi>p</mi> <mi>s</mi> <mi>i</mi> <mi>z</mi> <mi>e</mi> <mo>=</mo> <mn>16</mn> <mo>,</mo> <mtext> </mtext> <mi>t</mi> <mi>o</mi> <mi>p</mi> <mtext> </mtext> <mi>s</mi> <mi>u</mi> <mi>r</mi> <mi>f</mi> <mi>a</mi> <mi>c</mi> <mi>e</mi> <mtext> </mtext> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>a</mi> <mo>=</mo> <mstyle displaystyle="true"> <mo>∑</mo> <mrow> <mrow> <mo>(</mo> <mrow> <mn>1</mn> <mo>×</mo> <mi>s</mi> <mi>q</mi> <mi>u</mi> <mi>a</mi> <mi>r</mi> <mi>e</mi> <mi>s</mi> </mrow> <mo>)</mo> </mrow> </mrow> </mstyle> </mrow> </semantics> </math>.</p>
Full article ">Figure 7
<p>Relative differential box counting method (in the figure, the moving window size is 9 × 9 and the gliding box size is 3).</p>
Full article ">Figure 8
<p>Distribution pattern of different building areas from Google Map images ((<b>a</b>–<b>l</b>) represent different regions in the study area).</p>
Full article ">Figure 9
<p>Distribution of Deviation Degree (DD): (<b>a</b>) Jizhou; and (<b>b</b>) Beijing.</p>
Full article ">Figure 10
<p>Fractal dimensions and lacunarities of samples: (<b>a</b>) fractal dimension; and (<b>b</b>) lacunarity.</p>
Full article ">Figure 11
<p>(<b>a</b>) Backscatter intensity (<span class="html-italic">BI</span>) distribution of Jizhou; (<b>b</b>) amended backscatter intensity (<span class="html-italic">ABI</span>) distribution of Jizhou; (<b>c</b>) backscatter intensity (<span class="html-italic">BI</span>) distribution of Beijing; and (<b>d</b>) amended backscatter intensity (<span class="html-italic">ABI</span>) distribution of Beijing.</p>
Full article ">Figure 12
<p>Overall building density distribution of Jizhou: (<b>a</b>) model <span class="html-italic">a</span> using four GF-1 PMS spectral bands (<span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>); (<b>b</b>) model <span class="html-italic">b</span> using Normalized Difference Vegetation Index (<span class="html-italic">NDVI</span>), Normalized Difference Water Index (<span class="html-italic">NDWI</span>), and Ratio Built-up Index (<span class="html-italic">RBI</span>); (<b>c</b>) model <span class="html-italic">c</span> using <span class="html-italic">BI</span>; (<b>d</b>) model <span class="html-italic">d</span> using <span class="html-italic">ABI</span>; (<b>e</b>) model <span class="html-italic">e</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>, <span class="html-italic">NDVI</span>, <span class="html-italic">NDBI</span>, and <span class="html-italic">RBI</span>; (<b>f</b>) model <span class="html-italic">f</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub> and <span class="html-italic">BI</span>; (<b>g</b>) model <span class="html-italic">g</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub> and <span class="html-italic">ABI</span>; (<b>h</b>) model <span class="html-italic">h</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>, <span class="html-italic">NDVI</span>, <span class="html-italic">NDWI</span>, <span class="html-italic">RBI</span>, and <span class="html-italic">BI</span>; and (<b>i</b>) model <span class="html-italic">i</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>, <span class="html-italic">NDVI</span>, <span class="html-italic">NDWI</span>, <span class="html-italic">RBI</span>, and <span class="html-italic">ABI</span>.</p>
Full article ">Figure 13
<p>Overall building density distribution of Beijing: (<b>a</b>) model <span class="html-italic">a</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>; (<b>b</b>) model <span class="html-italic">b</span> using <span class="html-italic">NDVI</span>, <span class="html-italic">NDBI</span>, and <span class="html-italic">RBI</span>; (<b>c</b>) model <span class="html-italic">c</span> using <span class="html-italic">BI</span>; (<b>d</b>) model <span class="html-italic">d</span> using <span class="html-italic">ABI</span>; (<b>e</b>) model <span class="html-italic">e</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>, <span class="html-italic">NDVI</span>, <span class="html-italic">NDBI</span>, and <span class="html-italic">RBI</span>; (<b>f</b>) model <span class="html-italic">f</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub> and <span class="html-italic">BI</span>; (<b>g</b>) model <span class="html-italic">g</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub> and <span class="html-italic">ABI</span>; (<b>h</b>) model <span class="html-italic">h</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>, <span class="html-italic">NDVI</span>, <span class="html-italic">NDWI</span>, <span class="html-italic">RBI</span>, and <span class="html-italic">BI</span>; and (<b>i</b>) model <span class="html-italic">i</span> using <span class="html-italic">B</span><sub>1<span class="html-italic">–</span>4</sub>, <span class="html-italic">NDVI</span>, <span class="html-italic">NDWI</span>, <span class="html-italic">RBI</span>, and <span class="html-italic">ABI</span>.</p>
Full article ">
3301 KiB  
Article
Biomass Estimation Using 3D Data from Unmanned Aerial Vehicle Imagery in a Tropical Woodland
by Daud Jones Kachamba, Hans Ole Ørka, Terje Gobakken, Tron Eid and Weston Mwase
Remote Sens. 2016, 8(11), 968; https://doi.org/10.3390/rs8110968 - 23 Nov 2016
Cited by 92 | Viewed by 11584
Abstract
Application of 3D data derived from images captured using unmanned aerial vehicles (UAVs) in forest biomass estimation has shown great potential in reducing costs and improving the estimates. However, such data have never been tested in miombo woodlands. UAV-based biomass estimation relies on [...] Read more.
Application of 3D data derived from images captured using unmanned aerial vehicles (UAVs) in forest biomass estimation has shown great potential in reducing costs and improving the estimates. However, such data have never been tested in miombo woodlands. UAV-based biomass estimation relies on the availability of reliable digital terrain models (DTMs). The main objective of this study was to evaluate application of 3D data derived from UAV imagery in biomass estimation and to compare impacts of DTMs generated based on different methods and parameter settings. Biomass was modeled using data acquired from 107 sample plots in a forest reserve in miombo woodlands of Malawi. The results indicated that there are no significant differences (p = 0.985) between tested DTMs except for that based on shuttle radar topography mission (SRTM). A model developed using unsupervised ground filtering based on a grid search approach, had the smallest root mean square error (RMSE) of 46.7% of a mean biomass value of 38.99 Mg·ha−1. Amongst the independent variables, maximum canopy height (Hmax) was the most frequently selected. In addition, all models included spectral variables incorporating the three color bands red, green and blue. The study has demonstrated that UAV acquired image data can be used in biomass estimation in miombo woodlands using automatically generated DTMs. Full article
(This article belongs to the Special Issue Recent Trends in UAV Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Miombo woodlands in Muyobe forest reserve, Malawi (Photos: Hans Ole Ørka).</p>
Full article ">Figure 2
<p>Map of Malawi showing the location of the study site.</p>
Full article ">Figure 3
<p>Schematic diagram for the different DTM generation approaches.</p>
Full article ">Figure 4
<p>Mean height differences (m) between measured GPS (reference values) and predicted heights for the different DTM generation methods with standard errors: 01, DTM using supervised ground filtering based on visual classification; 02, DTM using supervised ground filtering based on logistic regression; 03, DTM using supervised ground filtering based on quantile regression; 04, DTM using unsupervised ground filtering based on shuttle radar topography mission (SRTM); and 05–13, DTMs using unsupervised ground filtering based on a grid search approach for optimal parameter settings in Agisoft Photoscan Professional software (See <a href="#sec2dot4-remotesensing-08-00968" class="html-sec">Section 2.4</a>. for details).</p>
Full article ">Figure 5
<p>Ground reference versus predicted biomass for different DTMs: 01, DTM using supervised ground filtering based on visual classification; 02, DTM using supervised ground filtering based on logistic regression; 03, DTM using supervised ground filtering based on quantile regression; 04, DTM using unsupervised ground filtering based on shuttle radar topography mission (SRTM); and 05–13, DTMs using unsupervised ground filtering based on a grid search approach for optimal parameter settings in Agisoft Photoscan Professional software (see <a href="#sec2dot4-remotesensing-08-00968" class="html-sec">Section 2.4</a>. for details).</p>
Full article ">
7903 KiB  
Article
Interpolation of GPS and Geological Data Using InSAR Deformation Maps: Method and Application to Land Subsidence in the Alto Guadalentín Aquifer (SE Spain)
by Marta Béjar-Pizarro, Carolina Guardiola-Albert, Ramón P. García-Cárdenas, Gerardo Herrera, Anna Barra, Antonio López Molina, Serena Tessitore, Alejandra Staller, José A. Ortega-Becerril and Ramón P. García-García
Remote Sens. 2016, 8(11), 965; https://doi.org/10.3390/rs8110965 - 23 Nov 2016
Cited by 48 | Viewed by 8365
Abstract
Land subsidence resulting from groundwater extractions is a global phenomenon adversely affecting many regions worldwide. Understanding the governing processes and mitigating associated hazards require knowing the spatial distribution of the implicated factors (piezometric levels, lithology, ground deformation), usually only known at discrete locations. [...] Read more.
Land subsidence resulting from groundwater extractions is a global phenomenon adversely affecting many regions worldwide. Understanding the governing processes and mitigating associated hazards require knowing the spatial distribution of the implicated factors (piezometric levels, lithology, ground deformation), usually only known at discrete locations. Here, we propose a methodology based on the Kriging with External Drift (KED) approach to interpolate sparse point measurements of variables influencing land subsidence using high density InSAR measurements. In our study, located in the Alto Guadalentín basin, SE Spain, these variables are GPS vertical velocities and the thickness of compressible soils. First, we estimate InSAR and GPS rates of subsidence covering the periods 2003–2010 and 2004–2013, respectively. Then, we apply the KED method to the discrete variables. The resulting continuous GPS velocity map shows maximum subsidence rates of 13 cm/year in the center of the basin, in agreement with previous studies. The compressible deposits thickness map is significantly improved. We also test the coherence of Sentinel-1 data in the study region and evaluate the applicability of this methodology with the new satellite, which will improve the monitoring of aquifer-related subsidence and the mapping of variables governing this phenomenon. Full article
(This article belongs to the Special Issue Multi-Sensor and Multi-Data Integration in Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location map. The black-dashed rectangles labelled A and B outline the regions used for the interpolation of GPS vertical velocities and the thickness of compressible soils., respectively. The thickness of compressible deposits, formed by clay and silt layers, is shown as colored circles at 18 boreholes (estimated by [<a href="#B28-remotesensing-08-00965" class="html-bibr">28</a>]). The black line delimits the Alto Guadalentín aquifer. The SRTM-90 Digital Elevation Model was used to generate the background topography. Names for the main cities in the study area are indicated. (<b>b</b>) Location of the 39 points that we measured with GPS between 28 January and 1 March 2013. Each site is identified with a number from 1–39 (<a href="#app1-remotesensing-08-00965" class="html-app">Supplementary Table S1</a>). The height of these points had been previously measured by the Spanish National Geographic Institute (IGNE): red diamonds were measured in August–September 2004; yellow diamonds were measured in March 2005; and blue diamonds were measured in March 2009. (<b>c</b>) NW-SE geological cross-section of the Alto Guadalentín basin, from [<a href="#B28-remotesensing-08-00965" class="html-bibr">28</a>]. The location of the two permanent GPS stations, LOR1 and LORC, is indicated by the black stars in (<b>a</b>,<b>c</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) Deformation map from ENVISAT data for the time period 2003–2010 (Box A in <a href="#remotesensing-08-00965-f001" class="html-fig">Figure 1</a>a). Negative deformation (red regions) indicates land subsidence. Black circles show the GPS location, and blue arrows indicate GPS vertical velocities. GPS velocities are obtained by comparing the reference vertical positions provided by the IGNE, which varies in the network between August–September 2004 and March 2009 (see <a href="#remotesensing-08-00965-f001" class="html-fig">Figure 1</a>b), and the new vertical positions measured in 2013 (see <a href="#app1-remotesensing-08-00965" class="html-app">Table S1</a>). The large blue arrow at the top represents the scale of GPS velocities. (<b>b</b>–<b>d</b>) Cross-sections showing vertical deformation rates from InSAR (red dots) and GPS data (blue circles) measured along the three dashed black lines in (<b>a</b>). (<b>e</b>) LOS time series for the point indicated by the black star in (<b>a</b>).</p>
Full article ">Figure 3
<p>Correlation diagrams: (<b>a</b>) comparison between deformation rates estimated from InSAR and GPS data. Note that InSAR data cover the period 2003–2010, while GPS data cover three different periods (2004–2013, 2005–2013, 2009–2013) depending on the position of the GPS point (see <a href="#sec3dot2-remotesensing-08-00965" class="html-sec">Section 3.2</a> and <a href="#remotesensing-08-00965-f001" class="html-fig">Figure 1</a>b). (<b>b</b>) Comparison between the InSAR-derived deformation rate and compressible deposit thickness.</p>
Full article ">Figure 4
<p>(<b>a</b>) Interpolated GPS-velocity map for period 2004–2013. Dashed black lines indicate cross-sections in (<b>b</b>–<b>d</b>). (<b>b</b>–<b>d</b>) Cross-sections showing vertical ground deformation rates from InSAR (red dots), GPS measurements (blue dots) and GPS interpolations (green dots).</p>
Full article ">Figure 5
<p>Cross-validation procedure for the GPS deformation rates’ interpolation. To characterize the ability of the interpolation to re-estimate (Z*) the data values (Z), errors are standardized by the predicted standard deviation ((Z* – Z)/S*). We define the interval [−2.5; 2.5] to focus on the 1% extreme values of a normal distribution. (<b>a</b>) Histogram of the standardized error. The minimum value, maximum value, mean value and variance are indicated. (<b>b</b>) Standardized error versus interpolated GPS rates in the 39 GPS locations.</p>
Full article ">Figure 6
<p>(<b>a</b>) Map of compressible soil thickness from this study (Box B in <a href="#remotesensing-08-00965-f001" class="html-fig">Figure 1</a>a). (<b>b</b>) Compressible soil thickness map from [<a href="#B28-remotesensing-08-00965" class="html-bibr">28</a>]. Circles indicate the location of the 18 boreholes where the thickness of the compressible deposit was measured. Map of residuals between maps (<b>a</b>,<b>b</b>) is shown in <a href="#app1-remotesensing-08-00965" class="html-app">Supplementary Figure S2</a>.</p>
Full article ">Figure 7
<p>Cross-validation procedure for the compressible deposit thickness interpolation. (<b>a</b>) Histogram of the standardized error. The minimum value, maximum value, mean value and variance are indicated. (<b>b</b>) Standardized error versus the interpolated thickness of the compressible deposit in the 18 borehole locations.</p>
Full article ">Figure 8
<p>Cross-sections showing ground deformation rates from ENVISAT data (red dots), ALOS data (gray dots), COSMO-SkyMed (CSK) data (green dots) and GPS measurements (blue circles). The temporal interval of each InSAR dataset is indicated in (<b>a</b>). GPS velocities were obtained by comparing the reference vertical positions provided by the IGNE, which varies in the network between August–September 2004 and March 2009 (see <a href="#remotesensing-08-00965-f001" class="html-fig">Figure 1</a>b), and the new vertical positions measured in 2013. The location of the cross-sections is indicated in <a href="#remotesensing-08-00965-f002" class="html-fig">Figure 2</a>a: (<b>a</b>) shows ground deformation rates across A–B, (<b>b</b>) shows ground deformation rates across C–D and (<b>c</b>) shows ground deformation rates across E–F.</p>
Full article ">
30825 KiB  
Article
Incremental and Enhanced Scanline-Based Segmentation Method for Surface Reconstruction of Sparse LiDAR Data
by Weimin Wang, Ken Sakurada and Nobuo Kawaguchi
Remote Sens. 2016, 8(11), 967; https://doi.org/10.3390/rs8110967 - 22 Nov 2016
Cited by 18 | Viewed by 8319
Abstract
The segmentation of point clouds is an important aspect of automated processing tasks such as semantic extraction. However, the sparsity and non-uniformity of the point clouds gathered by the popular 3D mobile LiDAR devices pose many challenges for existing segmentation methods. To improve [...] Read more.
The segmentation of point clouds is an important aspect of automated processing tasks such as semantic extraction. However, the sparsity and non-uniformity of the point clouds gathered by the popular 3D mobile LiDAR devices pose many challenges for existing segmentation methods. To improve the segmentation results of point clouds from mobile LiDAR devices, we propose an optimized segmentation method based on Scanline Continuity Constraint (SLCC) in this work. Unlike conventional scanline-based segmentation methods, SLCC clusters scanlines using the continuity constraints in terms of the distance as well as the direction of two consecutive points. In addition, scanline clusters are agglomerated not only into primitive geometrical shapes but also irregular shapes. Another downside to existing segmentation methods is that they are not capable of incremental processing. This causes unnecessary memory and time consumption for applications that require frame-wise segmentation or when new point clouds are added. In order to address this, we propose an incremental scheme—the Incremental Recursive Segmentation (IRIS), that can be easily applied to any segmentation method. IRIS is achieved by combining the segments of newly added point clouds and the previously segmented results. Furthermore, as an example application, we construct a processing pipeline consisting of plane fitting and surface reconstruction using the segmentation results. Finally, we evaluate the proposed methods on three datasets acquired from a handheld Velodyne HDL-32E LiDAR device. The experimental results verify the efficiency of IRIS for any segmentation method and the advantages of SLCC for processing mobile LiDAR data. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Automatic segmentation and modeling for a single frame of a point cloud or multiple frames. Left (<b>a</b>,<b>c</b>) show the single frame of spare point cloud input and multiple frames of sparse point cloud input respectively; Right (<b>b</b>,<b>d</b>) are output results; the surface is reconstructed using segmented information. Different segments are visualized using different colors.</p>
Full article ">Figure 2
<p>Notations defined in this paper.</p>
Full article ">Figure 3
<p>Block diagram of the pipeline for a processing point cloud.</p>
Full article ">Figure 4
<p>Two constraints of Scanline continuity. 1. <span class="html-italic">d</span> : distance of two consecutive points (green lines show the distance of two consecutive points from the same object, black lines show the same from different objects); 2. <span class="html-italic">θ</span>: the angle between two vectors of two groups of consecutive points. Two points are considered to be from two objects if either <span class="html-italic">d</span> or <span class="html-italic">θ</span> exceeds the user-defined thresholds.</p>
Full article ">Figure 5
<p>Theoretical interval of two consecutive points for different angular resolution and ranges. We can see that the distance of consecutive points increases greatly as range increases. This is the reason why we introduce an adaptive threshold for clustering scanlines.</p>
Full article ">Figure 6
<p>The effect of the incident angle to the theoretical interval of point <math display="inline"> <semantics> <msub> <mi mathvariant="bold-italic">p</mi> <mi>m</mi> </msub> </semantics> </math>. <math display="inline"> <semantics> <mrow> <mi mathvariant="normal">Δ</mi> <mi>θ</mi> </mrow> </semantics> </math> is the resolution angle in a horizontal direction. <span class="html-italic">r</span> is the range from point <math display="inline"> <semantics> <msub> <mi mathvariant="bold-italic">p</mi> <mi>m</mi> </msub> </semantics> </math> to the LiDAR. <span class="html-italic">α</span> is the incident angle. <math display="inline"> <semantics> <msub> <mi mathvariant="bold-italic">d</mi> <mi mathvariant="bold-italic">i</mi> </msub> </semantics> </math> represents the theoretical interval.</p>
Full article ">Figure 7
<p>Example for combining segments. The middle red, green, blue point clouds are three scanned frames of the left object from different positions and angles, the right one shows the combined segment.</p>
Full article ">Figure 8
<p>Linear classifier to judge whether two segments should be combined. (<b>a</b>) A separating plane exists between the red and blue point cloud, so we consider them as derived from different objects; (<b>b</b>) No plane could completely separate them, so we consider them as being from the same object.</p>
Full article ">Figure 9
<p>A challenge caused by registration error. The red and blue point clouds show scanned data of a plane such as a wall or roof, and they should be intersecting as in <a href="#remotesensing-08-00967-f008" class="html-fig">Figure 8</a>b. However, a separating plane exists in the gap caused by registration errors, which indicates that they are from different objects. For this situation, we map the red and blue points into the same plane and then judge with a linear classifier.</p>
Full article ">Figure 10
<p>Processing for planar segments. (<b>a</b>) An input point cloud segment that is detected as a planar shape; (<b>b</b>) All points are mapped in the detected plane, and boundary points are extracted; (<b>c</b>) Planar surface is reconstructed.</p>
Full article ">Figure 11
<p>Surface reconstruction comparison. Green points represents vertexes of the point cloud. (<b>a</b>) Shows the reconstructed surface by 3D Delaunay triangulation, where a 3D convex hull is constructed; (<b>b</b>) shows the result based on alpha shape. We can see that surfaces are better reconstructed based on alpha shape rather than Delaunay triangulation.</p>
Full article ">Figure 12
<p>Three datasets are used in this work. (<b>a</b>–<b>c</b>) Point clouds of the Corridor, Lobby and Underground shopping mall dataset respectively; (<b>d</b>–<b>f</b>) Panoramic images of the three datasets.</p>
Full article ">Figure 13
<p>Time performance of IRIS . Both Region Growing and IRIS Region Growing are applied to the three datasets from one frame to sixteen frames of point clouds. (<b>a</b>–<b>c</b>) Results of the Corridor, Lobby and Underground shopping mall are shown respectively. The time consumption of IRIS Region Growing is calculated without considering the computation time of previous frames.</p>
Full article ">Figure 14
<p>Scanline clustering comparison. (<b>a</b>,<b>b</b>) show the result of the method in [<a href="#B14-remotesensing-08-00967" class="html-bibr">14</a>], and (<b>c</b>,<b>d</b>) show the result of our proposed method for Scanline clustering. Magnified images are shown in (<b>b</b>,<b>d</b>). Compared with the left part in (<b>d</b>), the left part in (<b>b</b>) will cause under-segmentation after the agglomeration of Scanline clusters. Compared with the right part in (<b>d</b>), the right part in (<b>b</b>) will cause over-segmentation after the agglomeration of scanline clusters (see the red dash line ).</p>
Full article ">Figure 15
<p>Segmentation results for one frame of point cloud. (<b>a</b>,<b>b</b>) show results with Region Growing, and (<b>c</b>,<b>d</b>) show results with our proposed SLCC. From the magnified images in the right column, our proposed SLCC performs better in avoiding miss-segmentation (marked by boxes with black, red and magenta lines) and over-segmentation (marked by the box with blue lines).</p>
Full article ">Figure 16
<p>Examples of irregularly shaped objects segmented by SLCC. (<b>a</b>–<b>d</b>) Pedestrian, pedestrian, two pedestrians side by side, pedestrian segmented from the Underground shopping mall dataset; (<b>e</b>–<b>h</b>) cyclist, cyclist, pedestrian, car segmented from the KITTI dataset [<a href="#B36-remotesensing-08-00967" class="html-bibr">36</a>].</p>
Full article ">Figure 17
<p>Comparison of segmentation results on multiple frames. Results of Region Growing, IRIS Region Growing and SLCC are shown from upper to lower rows respectively. Magnified images are shown in the right column. (<b>a</b>,<b>b</b>) Region Growing is prone to over-segmentation due to the sparsity and non-uniformity; (<b>c</b>,<b>d</b>) Over-segmentation of Region Growing is also suppressed by IRIS; (<b>e</b>,<b>f</b>) A small car is well segmented by IRIS SLCC (marked by the box with magenta lines).</p>
Full article ">Figure 18
<p>Surface reconstruction of a single frame point cloud without regularization. (<b>a</b>,<b>b</b>) show surface reconstruction based on alpha shape without plane regularization for comparison. (<b>c</b>,<b>d</b>) show the result after plane regularization. We can see that creases are flatter.</p>
Full article ">Figure 19
<p>Surface reconstruction results for a multiple point cloud of LiDAR. Different colors of the plane indicate over-segmentation. (<b>a</b>,<b>b</b>) Region Growing; (<b>c</b>,<b>d</b>) IRIS Region Growing; (<b>e</b>,<b>f</b>) IRIS SLCC.</p>
Full article ">Figure 20
<p>(<b>a</b>–<b>l</b>) More surface reconstruction results on three datasets. (<b>a</b>,<b>c</b>,<b>e</b>) Results using Region Growing for one frame on three datasets; (<b>b</b>,<b>d</b>,<b>f</b>) results using SLCC for one frame on three datasets; (<b>g</b>–<b>l</b>) results using Region Growing, IRIS Region Growing and IRIS SLCC for the Corridor and Entrance datasets.</p>
Full article ">
36024 KiB  
Article
Multi-Sensor Geomagnetic Prospection: A Case Study from Neolithic Thessaly, Greece
by Tuna Kalaycı and Apostolos Sarris
Remote Sens. 2016, 8(11), 966; https://doi.org/10.3390/rs8110966 - 22 Nov 2016
Cited by 9 | Viewed by 5463
Abstract
Multi-sensor prospecting is a fast-emerging paradigm in archaeological geophysics. Given suitable ground conditions for navigation, sensor arrays drastically increase efficiency in data collection. In particular, geomagnetic prospecting benefits from this development. Despite these advancements, data processing still lacks a best-practice approach. Conventional processing [...] Read more.
Multi-sensor prospecting is a fast-emerging paradigm in archaeological geophysics. Given suitable ground conditions for navigation, sensor arrays drastically increase efficiency in data collection. In particular, geomagnetic prospecting benefits from this development. Despite these advancements, data processing still lacks a best-practice approach. Conventional processing methods developed for gridded data has been challenged by sensor arrays “roaming” in the landscape. In realization of the issue, the Innovative Geophysical Approaches for the Study of Early Agricultural Villages of Neolithic Thessaly (IGEAN) Project explored various innovative techniques for the betterment of the multi-sensor geomagnetic data processing. As a result, a modular pipeline is produced with minimal user intervention. In addition to standard steps, such as data clipping, various other algorithms have been introduced. This pipeline is tested over 20 Neolithic settlements in Thessaly, Greece, three of which are presented here in detail. The proposed workflow provides drastic improvements over raw data. As a result of these improvements, the IGEAN project revealed astonishing details on architectural elements, settlement enclosures, and paleolandscapes, changing completely the existing perspective of the Neolithic habitation in Thessaly. Full article
(This article belongs to the Special Issue Archaeological Prospecting and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Thessaly was the gateway for the Neolithization of Europe. The Innovative Geophysical Approaches for the Study of Early Agricultural Villages of Neolithic Thessaly (IGEAN) project examined 24 settlements in detail using geospatial technologies. Three sites (Almyriotiki, Perdika 1, and Velestino-Mati) constitute the demonstration case studies of this paper.</p>
Full article ">Figure 2
<p>The SENSYS Sensorik and Systemtechnologie can carry up to 16 sensors, offering variations for sensor numbers and separations. The IGEAN setup included eight sensors with 50 cm separation. A high-precision Global Navigation Satellite System (GNSS) receiver helps in navigation and sensor-data registration.</p>
Full article ">Figure 3
<p>The IGEAN workflow for processing geomagnetic data. The first three steps (ASCII parsing, quality control, and data clipping) are standardized. The proceeding steps and their related parameters are case dependent.</p>
Full article ">Figure 4
<p>The raw data (<b>above</b>) is imported from the SENSYS software in GeoTIFF format and displayed in a Geographic Information System (GIS). The processed data (<b>below</b>) shows drastic improvements over the original dataset. The data is an excerpt from Magoula Almyriotiki (see <a href="#sec5dot1-remotesensing-08-00966" class="html-sec">Section 5.1</a>).</p>
Full article ">Figure 5
<p>The grid collection of geomagnetic data (red triangles) results in far fewer data points than multi-sensor collection. The comparison is the most clear in the figure inlay. For both images, the north is up.</p>
Full article ">Figure 6
<p>The space-phase method can detect data outliers in an efficient manner. Geomagnetic readings falling outside of the 3D ellipsoid (in red) are tagged as data spikes.</p>
Full article ">Figure 7
<p>Distribution of data points is determined by track navigation. (<b>a</b>) Horizontal hatching indicates overlapping areas and thus these areas contain denser data. Diagonal hatching indicates data gaps, which can be avoided by good navigation skills in most cases; (<b>b</b>) The alpha-shape approach can remove most of the data overlaps and even data distribution can be achieved for further processing.</p>
Full article ">Figure 8
<p>Geomagnetic data from Magoula Almyriotiki. Two distinct architectural groups and settlement enclosures are clearly visible.</p>
Full article ">Figure 9
<p>Geomagnetic data from Perdika 1. Palimpsest of two architectural groups are visible in the dataset. The settlement enclosure is the most visible in the north.</p>
Full article ">Figure 10
<p>Geomagnetic data from Magoula Velestino-Mati. The bean shape settlement mound reveals some information about the built environment. There is little information on settlement enclosures but, when available, they are significantly distinct from the other two sites mentioned in this study.</p>
Full article ">Figure 11
<p>Magnetic susceptibility data from the enclosure in Magoula Almyriotiki suggest a single running feature; unlike the geomagnetic data which reveal a double-ditch like feature.</p>
Full article ">Figure 12
<p>Magnetometers are the work-horses of archaeological prospection. However, it is the best practice to integrate other sensors into the work schema. (<b>a</b>) Geomagnetic data from Magoula Almyriotiki reveals a large building to the south of the settlement; (<b>b</b>) GPR survey in the same area further corrects this argument and suggests three buildings; two rectangular buildings to the west and another one, equally partitioned, to the east.</p>
Full article ">
17184 KiB  
Article
Cloud Extraction from Chinese High Resolution Satellite Imagery by Probabilistic Latent Semantic Analysis and Object-Based Machine Learning
by Kai Tan, Yongjun Zhang and Xin Tong
Remote Sens. 2016, 8(11), 963; https://doi.org/10.3390/rs8110963 - 22 Nov 2016
Cited by 42 | Viewed by 11273
Abstract
Automatic cloud extraction from satellite imagery is a vital process for many applications in optical remote sensing since clouds can locally obscure the surface features and alter the reflectance. Clouds can be easily distinguished by the human eyes in satellite imagery via remarkable [...] Read more.
Automatic cloud extraction from satellite imagery is a vital process for many applications in optical remote sensing since clouds can locally obscure the surface features and alter the reflectance. Clouds can be easily distinguished by the human eyes in satellite imagery via remarkable regional characteristics, but finding a way to automatically detect various kinds of clouds by computer programs to speed up the processing efficiency remains a challenge. This paper introduces a new cloud detection method based on probabilistic latent semantic analysis (PLSA) and object-based machine learning. The method begins by segmenting satellite images into superpixels by Simple Linear Iterative Clustering (SLIC) algorithm while also extracting the spectral, texture, frequency and line segment features. Then, the implicit information in each superpixel is extracted from the feature histogram through the PLSA model by which the descriptor of each superpixel can be computed to form a feature vector for classification. Thereafter, the cloud mask is extracted by optimal thresholding and applying the Support Vector Machine (SVM) algorithm at the superpixel level. The GrabCut algorithm is then applied to extract more accurate cloud regions at the pixel level by assuming the cloud mask as the prior knowledge. When compared to different cloud detection methods in the literature, the overall accuracy of the proposed cloud detection method was up to 90 percent for ZY-3 and GF-1 images, which is about a 6.8 percent improvement over the traditional spectral-based methods. The experimental results show that the proposed method can automatically and accurately detect clouds using the multispectral information of the available four bands. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Graphic examples of the three types of clouds visually recognizable from remotely sensed data: thick cloud, thin cloud, Cirrus cloud.</p>
Full article ">Figure 2
<p>Proposed cloud detection framework.</p>
Full article ">Figure 3
<p>Superpixel segmentation by SLIC.</p>
Full article ">Figure 4
<p>Basic features for cloud detection. (<b>a</b>) Original image; (<b>b</b>) Hue channel; (<b>c</b>) Intensity channel; (<b>d</b>) Saturation channel; (<b>e</b>) NIR channel.</p>
Full article ">Figure 5
<p>Complicated features of a graphic example. (<b>a</b>) Original image; (<b>b</b>) SF image; (<b>c</b>) TF image; (<b>d</b>) FF image; (<b>e</b>) LSD.</p>
Full article ">Figure 6
<p>Statistics of the features information for the three cloud types. Each box plot shows the location of the 25, 50, and 75 percentiles using horizontal lines. The two whiskers cover the 99.3% percentile of the data, and the red “+” is the outlier outside the two whiskers.</p>
Full article ">Figure 7
<p>Example showing BS extraction based on four conditions. (<b>a</b>) Superpixels that missed condition 1; (<b>b</b>) Superpixels that missed condition 2; (<b>c</b>) Superpixels that missed condition 3; (<b>d</b>) Superpixels that missed condition 4; (<b>e</b>) BS extraction result.</p>
Full article ">Figure 8
<p>Flow chart of cloud mark extraction.</p>
Full article ">Figure 9
<p>Four categories of superpixels (thick cloud superpixels, thin cloud superpixels, cirrus cloud superpixel, others) and their feature histogram STFH. The histogram describes the SF, TF, and FF distribution of the superpixel, the green line histograms denote the positive superpixels, and the yellow ones are the negative superpixels. The x label of the feature histogram is the feature value, ranging from 0 to 255, and the y label is the proportion of the corresponding feature value.</p>
Full article ">Figure 10
<p>The features distribution of snow and thick clouds. Each pentagram represents a superpixel, plotted by the SF value on the horizontal axis and the NIR value on the vertical. The red pentagrams denote snow superpixels and the green ones are thick cloud superpixels. The circles are the boundary superpixels, and the yellow dotted line denotes the reference boundary fitted by the Least Squares algorithm.</p>
Full article ">Figure 11
<p>Example of cloud mask extraction using the proposed method. (<b>a</b>) Original image; (<b>b</b>) Cloud mask extracted by SVM. (<b>c</b>) Refined cloud mask. The white superpixles in (<b>b</b>) and (<b>c</b>) denote the foreground superpixels, the light gray ones denote possible foreground superpixels, the dark gray ones denote possible background superpixels, and the black ones denote the background superpixels. The red lines with double arrows between (<b>b</b>) and (<b>c</b>) point out the difference between before and after the refined cloud mask.</p>
Full article ">Figure 12
<p>Example of accurate cloud detection by the proposed method. The first row shows the original images, the second row shows the cloud mask extracted by the method described in <a href="#sec3dot2-remotesensing-08-00963" class="html-sec">Section 3.2</a>., and the third row shows the accurate cloud detection results.</p>
Full article ">Figure 13
<p>Cloud detection precision analysis of internal feasibility (Red denotes extracted true cloud regions, yellow denotes missed true cloud regions, and green denotes non-cloud regions misjudged as cloud regions).</p>
Full article ">Figure 14
<p>Effects of parameter S for cloud detection.</p>
Full article ">Figure 15
<p>Effects of parameter K for cloud detection.</p>
Full article ">Figure 16
<p>Results of comparison with automatic cloud detection algorithms and comparison of visually extracted clouds with those obtained by automatic detection algorithms (Red denotes the extracted true cloud regions, yellow denotes the missed true cloud regions, green denote the non-cloud regions misjudged as cloud regions).</p>
Full article ">Figure 17
<p>Results of comparison with automatic image segmentation algorithms (Red denotes the extracted true cloud regions, yellow denotes the missed true cloud regions, green denotes the non-cloud regions misjudged as cloud regions).</p>
Full article ">Figure 18
<p>Results of comparison with interactive image segmentation algorithms (Red regions in the columns of GrabCut, watershed and the proposed method denote the extracted true cloud regions, yellow denotes the missed cloud regions, and green denotes the non-cloud regions that were misjudged as cloud regions. The green scribbles in the columns of the GrabCut mask and the Watershed mask indicate non-cloud regions, and the red scribbles indicate cloud pixels).</p>
Full article ">Figure 19
<p>Results of comparison with other algorithms in the failed cases (Red denotes the extracted true cloud regions, yellow denotes the missed true cloud regions, green denotes the non-cloud regions misjudged as cloud regions).</p>
Full article ">
3877 KiB  
Article
Spatial Distribution of Diffuse Attenuation of Photosynthetic Active Radiation and Its Main Regulating Factors in Inland Waters of Northeast China
by Jianhang Ma, Kaishan Song, Zhidan Wen, Ying Zhao, Yingxin Shang, Chong Fang and Jia Du
Remote Sens. 2016, 8(11), 964; https://doi.org/10.3390/rs8110964 - 21 Nov 2016
Cited by 21 | Viewed by 6060
Abstract
Light availability in lakes or reservoirs is affected by optically active components (OACs) in the water. Light plays a key role in the distribution of phytoplankton and hydrophytes, thus, is a good indicator of the trophic state of an aquatic system. Diffuse attenuation [...] Read more.
Light availability in lakes or reservoirs is affected by optically active components (OACs) in the water. Light plays a key role in the distribution of phytoplankton and hydrophytes, thus, is a good indicator of the trophic state of an aquatic system. Diffuse attenuation of photosynthetic active radiation (PAR) (Kd(PAR)) is commonly used to quantitatively assess the light availability. The PAR and the concentration of OACs were measured at 206 sites, which covered 26 lakes and reservoirs in Northeast China. The spatial distribution of Kd(PAR) was depicted and its association with the OACs was assessed by grey incidences(GIs) and linear regression analysis. Kd(PAR) varied from 0.45 to 15.04 m−1. This investigation revealed that reservoirs in the east part of Northeast China were clear with small Kd(PAR) values, while lakes located in plain areas, where the source of total suspended matter (TSM) varied, displayed high Kd(PAR) values. The GIs and linear regression analysis indicated that the TSM was the dominant factor in determining Kd(PAR) values and best correlated with Kd(PAR) (R2 = 0.906, RMSE = 0.709). Most importantly, we have demonstrated that the TSM concentration is a reliable measurement for the estimation of the Kd(PAR) as 74% of the data produced a relative error (RE) of less than 0.4 in a leave-one-out cross validation (LOO-CV) analysis. Spatial transferability assessment of the model also revealed that TSM performed well as a determining factor of the Kd(PAR) for the majority of the lakes. However, a few exceptions were identified where the optically regulating dominant factors were chlorophyll-a (Chl-a) and/or the chromophroic dissolved organic matter (CDOM). These extreme cases represent lakes with exceptionally clear waters. Full article
(This article belongs to the Special Issue Water Optics and Water Colour Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area map: (<b>a</b>) Digital elevation model (DEM) and sampling lakes with corresponding lake ID; and (<b>b</b>) type of vegetation and spatial distribution of the K<sub>d</sub>(PAR) derived from in situ measurements, combined with isotherm and isohyets in Northeast China.</p>
Full article ">Figure 2
<p>Correlation between K<sub>d</sub>(PAR) and the concentration of the water’s quality parameters: (<b>a</b>) water transparency (SDD); (<b>b</b>) total suspended matter (TSM); (<b>c</b>) chlorophyll-<span class="html-italic">a</span> (Chl-<span class="html-italic">a</span>); and (<b>d</b>) absorption coefficient of chromophroic dissolved organic matter (CDOM) at 355 nm (a<sub>CDOM</sub>(355)).</p>
Full article ">Figure 3
<p>Frequency distribution and cumulative percentage of the relative errors (RE) derived from the leave-out-one cross validation (LOO-CV) of the regression model that had been used to calibrate the K<sub>d</sub>(PAR) from the concentration of TSM.</p>
Full article ">Figure 4
<p>Relative contribution of phytoplankton (a<sub>ph</sub>), NAP (a<sub>NAP</sub>) and CDOM (a<sub>CDOM</sub>) to the total absorption of the sample points with relative errors (RE) greater than 0.6 (based on LOO-CV analysis). Sample points are represented on the x-axis with the abbreviations of the location’s name.</p>
Full article ">Figure 5
<p>The mean relative error (MRE) and root-mean-square error (RMSE) of the K<sub>d</sub>(PAR) regression models for each lake for the evaluation of the spatial transferability of the model. LakeID was corresponding to <a href="#remotesensing-08-00964-t001" class="html-table">Table 1</a> and was represented on the x-axis.</p>
Full article ">Figure 6
<p>The bars represent the determination coefficients (R<sup>2</sup>) of the linear regression analysis between the OACs and the K<sub>d</sub>(PAR). The lines represent the relative contribution of phytoplankton (a<sub>ph</sub>), non-algal particles (a<sub>NAP</sub>) and CDOM (a<sub>CDOM</sub>) to the total absorption. Abbreviations of the lakes correspond to those in <a href="#remotesensing-08-00964-t001" class="html-table">Table 1</a> and are represented on the x-axis.</p>
Full article ">
14337 KiB  
Article
Global Daily High-Resolution Satellite-Based Foundation Sea Surface Temperature Dataset: Development and Validation against Two Definitions of Foundation SST
by Kohtaro Hosoda and Futoki Sakaida
Remote Sens. 2016, 8(11), 962; https://doi.org/10.3390/rs8110962 - 21 Nov 2016
Cited by 8 | Viewed by 6725
Abstract
This paper describes a global, daily sea surface temperature (SST) analysis based on satellite microwave and infrared measurements. The SST analysis includes a diurnal correction method to estimate foundation SST (SST free from diurnal variability) using satellite sea surface wind and solar radiation [...] Read more.
This paper describes a global, daily sea surface temperature (SST) analysis based on satellite microwave and infrared measurements. The SST analysis includes a diurnal correction method to estimate foundation SST (SST free from diurnal variability) using satellite sea surface wind and solar radiation data, frequency splitting to reproduce intra-seasonal variability and a quality control procedure repeated twice to avoid operation errors. An optimal interpolation method designed for foundation SST is applied to blend the microwave and infrared satellite measurements. Although in situ SST measurements are not used for bias correction adjustments in the analysis, the output product, with a spatial grid size of 0.1°, has an accuracy of 0.48 C and 0.46 C compared to the in situ foundation SST measurements derived by drifting buoys and Argo floats, respectively. The same quality against the two types of in situ foundation SST (drifters and Argo) suggests that the two definitions of foundation SST proposed by past studies can provide same-quality information about the sea surface state underlying the diurnal thermocline. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Example of the satellite-based foundation sea surface temperature (SST) (SST<math display="inline"> <semantics> <msub> <mrow/> <mi>fnd</mi> </msub> </semantics> </math>: top) and spatial gradient (<math display="inline"> <semantics> <msqrt> <mrow> <msup> <mrow> <mo>(</mo> <msub> <mi>∂</mi> <mi>x</mi> </msub> <msub> <mi>SST</mi> <mi>fnd</mi> </msub> <mo>)</mo> </mrow> <mn>2</mn> </msup> <mo>+</mo> <msup> <mrow> <mo>(</mo> <msub> <mi>∂</mi> <mi>y</mi> </msub> <msub> <mi>SST</mi> <mi>fnd</mi> </msub> <mo>)</mo> </mrow> <mn>2</mn> </msup> </mrow> </msqrt> </semantics> </math>) on 1 January 2014 of this work. The contour interval (C.I.) in the SST<math display="inline"> <semantics> <msub> <mrow/> <mi>fnd</mi> </msub> </semantics> </math> field is 1 <math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>C.</p>
Full article ">Figure 2
<p>(<b>a</b>) Example of Argo near-surface profiling data and the determination of foundation SST by this study. Black dots represent observations by Argo float, and the solid line is the line interpolated by the Akima method from the observations. The dashed line denotes the foundation sea surface temperature (SST) determined from the observation. The difference between the foundation SST and 1-m depth SST was 1.88 <math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>C. (<b>b</b>) Satellite-based foundation SST field of this study. (<b>c</b>) Same as (<b>b</b>), but for the daily maximum minus minimum SST estimated from the sea surface wind (SSW) and solar radiation data. Cross points in (<b>b</b>) and (<b>c</b>) corresponded to the Argo float.</p>
Full article ">Figure 3
<p>Flowchart of the data integration in this study for the low-frequency part. Dashed arrows denote the usage of the reference data for quality control (QC) of the input data indicated by solid arrows. Data storage indicated by blue color is used in the next step (<a href="#remotesensing-08-00962-f004" class="html-fig">Figure 4</a>).</p>
Full article ">Figure 4
<p>Same as <a href="#remotesensing-08-00962-f003" class="html-fig">Figure 3</a>, but for the high-frequency part and combined low- and high-frequency parts used in this study. Data storage indicated by blue color is derived from the previous step (<a href="#remotesensing-08-00962-f003" class="html-fig">Figure 3</a>).</p>
Full article ">Figure 5
<p>Example of the frequency splitting for foundation SST in this study. (<b>a</b>) Original foundation SST derived from the composite of microwave (AMSR-E and WindSat) and infrared (MODIS) measurements on 21 May 2005; (<b>b</b>) low-frequency component of blended (microwave and infrared) foundation SST; (<b>c</b>,<b>d</b>) high-frequency component (anomaly from the low-frequency component) of infrared/microwave foundation SST, respectively. C.I. in (<b>a</b>) and (<b>b</b>) is 1 <math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>C.</p>
Full article ">Figure 6
<p>(<b>a</b>) Histograms of satellite-based foundation SST and in situ measurements (red: drifting buoy; green: Argo profiling floats). Statistics of these (bias, RMSE and the number of match-ups are summarized in the upper part of the figure. (<b>b</b>,<b>c</b>) Distributions of validation match-ups between drifting buoys (<b>b</b>) and Argo floats (<b>c</b>). The number of match-ups were calculated in a 5° × 10° (lat. × lon.) box. Their comparison periods are 2003–2015 (<b>b</b>) and 2012–2015, respectively.</p>
Full article ">Figure 7
<p>Distributions of standard deviation and bias of the satellite-based blended (infrared + microwave) foundation SST in this study. (<b>a</b>) Standard deviation of the low-frequency part; (<b>b</b>) standard deviation of the combination of low- and high-frequency parts; (<b>c</b>) bias of the low-frequency part; and (<b>d</b>) bias of the combination of low- and high-frequency parts. The statistics were obtained by comparisons with 2003–2015 drifting buoy measurements.</p>
Full article ">Figure 8
<p>Global validation statistics of the satellite foundation SST against in situ foundation SST by drifting buoy (red, semi-annually). and Argo floats (blue, semi-annually). Calculated RMSE and bias are shown by solid and dashed lines, respectively.</p>
Full article ">Figure 9
<p>Accuracy (RMSE) of the blended SST product depending on the infrared (IR) SST coverage (horizontal axis) in the three-dimensional range searching data for optimal interpolation (OI) and the SST gradient (vertical axis) derived from the blended SST data. The coastal areas (oceans from coastline ≤ 200 km) were excluded from the calculation. (<b>a</b>) Period without AMSR2 10-GHz SST data (until July 2012); and (<b>b</b>) period with AMSR2 10-GHz data (after July 2012, except for low-temperature areas (&lt;11 <math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>C)). RMSEs were calculated in each 10% × 1 <math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>C/100 km bin.</p>
Full article ">Figure 10
<p>Seasonal validation statistics of the daily-mean SST (green: New Generation SST for Open Ocean (NGSST-O) Version 1.6) and foundation SST (black: this study) products in the western North Pacific (13–63°N, 113–163°E) during 2003–2011. Calculated RMSE and bias are shown by solid and dashed lines, respectively.</p>
Full article ">
458 KiB  
Correction
Correction: Singh, A., et al. Remote Sensing of Storage Fluctuations of Poorly Gauged Reservoirs and State Space Model (SSM)-Based Estimation. Remote Sens. 2015, 7, 17113–17134
by Alka Singh, Ujjwal Kumar and Florian Seitz
Remote Sens. 2016, 8(11), 960; https://doi.org/10.3390/rs8110960 - 21 Nov 2016
Cited by 2 | Viewed by 3242
Abstract
The authors wish to make the following corrections to their paper [1].[...] Full article
Show Figures

Figure 1

Figure 1
<p>Lake Mead SSM analysis. (<b>Top left</b>) The combined SSM estimate (CSSME) (magenta line) and the forecast (green line) for 2013 and 2014; (<b>Bottom left</b>) difference between CSSME and in situ observations; (<b>Top right</b>) estimated seasonal component; (<b>Bottom right</b>) estimated trend component. The dashed cyan lines indicate the upper and lower 95% confidence limit.</p>
Full article ">Figure 2
<p>Lake Mead SSM analysis. (<b>Top left</b>) The combined SSM estimate (CSSME) (black line) and the forecast (green line) for 2013 and 2014; (<b>Bottom left</b>) difference between CSSME and in situ observations; (<b>Top right</b>) estimated seasonal component; (<b>Bottom right</b>) estimated trend component. The dashed cyan lines indicate the upper and lower 95% confidence limit.</p>
Full article ">
21134 KiB  
Article
Estimating and Up-Scaling Fuel Moisture and Leaf Dry Matter Content of a Temperate Humid Forest Using Multi Resolution Remote Sensing Data
by Hamed Adab, Kasturi Devi Kanniah and Jason Beringer
Remote Sens. 2016, 8(11), 961; https://doi.org/10.3390/rs8110961 - 19 Nov 2016
Cited by 16 | Viewed by 7661
Abstract
Vegetation moisture and dry matter content are important indicators in predicting the behavior of fire and it is widely used in fire spread models. In this study, leaf fuel moisture content such as Live Fuel Moisture Content (LFMC), Leaf Relative Water Content (RWC), [...] Read more.
Vegetation moisture and dry matter content are important indicators in predicting the behavior of fire and it is widely used in fire spread models. In this study, leaf fuel moisture content such as Live Fuel Moisture Content (LFMC), Leaf Relative Water Content (RWC), Dead Fuel Moisture Content (DFMC), and Leaf Dry Matter Content (LDMC) (hereinafter known as moisture content indices (MCI)) were calculated in the field for different forest species at 32 sites in a temperate humid forest (Zaringol forest) located in northeastern Iran. These data and several relevant vegetation-biophysical indices and atmospheric variables calculated using Landsat 7 Enhanced Thematic Mapper Plus (ETM+) data with moderate spatial resolution (30 m) were used to estimate MCI of the Zaringol forest using Artificial Neural Network (ANN) and Multiple Linear Regression (MLR) methods. The prediction of MCI using ANN showed that ETM+ predicted MCI slightly better (Mean Absolute Percentage Error (MAPE) of 6%–12%)) than MLR (MAPE between 8% and 17%). Once satisfactory results in estimating MCI were obtained by using ANN from ETM+ data, these data were then upscaled to estimate MCI using MODIS data for daily monitoring of leaf water and leaf dry matter content at 500 m spatial resolution. For MODIS derived LFMC, LDMC, RWC, and DLMC, the ANN produced a MAPE between 11% and 29% for the indices compared to MLR which produced an MAPE of 14%–33%. In conclusion, we suggest that upscaling is necessary for solving the scale discrepancy problems between the indicators and low spatial resolution MODIS data. The scaling up of MCI could be used for pre-fire alert system and thereby can detect fire prone areas in near real time for fire-fighting operations. Full article
(This article belongs to the Special Issue Remote Sensing of Vegetation Structure and Dynamics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location of the province of Golestan in Iran; (<b>b</b>) boundary of Zaringol forest in Golestan province and location of the 32 sampling sites that cover different forest species in the Zaringol forest area. Locations are shown on ETM+ satellite image dated 20 July 2000.</p>
Full article ">Figure 2
<p>Box plots showing the distribution of MCI calculated using data collected in the field. Whiskers represent highest and lowest value of MCI metrics, solid lines of the box represent the median and the 25th and 75th percentiles.</p>
Full article ">Figure 3
<p>Visual comparison of residual standard errors between estimated LFMC, RWC, LDMC, and DFMC from aggregated Landsat products and MODIS by MLR and ANN. Blue color shows the values estimated by MODIS was more than ETM+ (overestimates) and the brown color displays the values estimated by MODIS was less than ETM+ (underestimates). There were no differences between the value aggregated Landsat products and MODIS in cream colored area.</p>
Full article ">Figure 4
<p>Estimated LFMC, RWC, LDMC, and DFMC (%) from ETM+ (30 m) using the ANN technique for the Zaringol forest at 11:20 a.m. local time, on 6 August 2012.</p>
Full article ">Figure 5
<p>The MCI data (%) upscaled from 30 m Landsat ETM+ to the spatial resolution of the Terra MODIS (500 m) using ANN technique for the Zaringol forest at 11:40 a.m. local time, on 6 August 2012.</p>
Full article ">
16495 KiB  
Article
Long Term Global Surface Soil Moisture Fields Using an SMOS-Trained Neural Network Applied to AMSR-E Data
by Nemesio J. Rodríguez-Fernández, Yann H. Kerr, Robin Van der Schalie, Amen Al-Yaari, Jean-Pierre Wigneron, Richard De Jeu, Philippe Richaume, Emanuel Dutra, Arnaud Mialon and Matthias Drusch
Remote Sens. 2016, 8(11), 959; https://doi.org/10.3390/rs8110959 - 18 Nov 2016
Cited by 38 | Viewed by 6118
Abstract
A method to retrieve soil moisture (SM) from Advanced Scanning Microwave Radiometer—Earth Observing System Sensor (AMSR-E) observations using Soil Moisture and Ocean Salinity (SMOS) Level 3 SM as a reference is discussed. The goal is to obtain longer time series of SM with [...] Read more.
A method to retrieve soil moisture (SM) from Advanced Scanning Microwave Radiometer—Earth Observing System Sensor (AMSR-E) observations using Soil Moisture and Ocean Salinity (SMOS) Level 3 SM as a reference is discussed. The goal is to obtain longer time series of SM with no significant bias and with a similar dynamical range to that of the SMOS SM dataset. This method consists of training a neural network (NN) to obtain a global non-linear relationship linking AMSR-E brightness temperatures ( T b ) to the SMOS L3 SM dataset on the concurrent mission period of 1.5 years. Then, the NN model is used to derive soil moisture from past AMSR-E observations. It is shown that in spite of the different frequencies and sensing depths of AMSR-E and SMOS, it is possible to find such a global relationship. The sensitivity of AMSR-E T b ’s to soil temperature ( T s o i l ) was also evaluated using European Centre for Medium-Range Weather Forecast Interim/Land re-analysis (ERA-Land) and Modern-Era Retrospective analysis for Research and Applications-Land (MERRA-Land) model data. The best combination of AMSR-E T b ’s to retrieve T s o i l is H polarization at 23 and 36 GHz plus V polarization at 36 GHz. Regarding SM, several combinations of input data show a similar performance in retrieving SM. One NN that uses C and X bands and T s o i l information was chosen to obtain SM in the 2003–2011 period. The new dataset shows a low bias (<0.02 m3/m3) and low standard deviation of the difference (<0.04 m3/m3) with respect to SMOS L3 SM over most of the globe’s surface. The new dataset was evaluated together with other AMSR-E SM datasets and the Climate Change Initiative (CCI) SM dataset against the MERRA-Land and ERA-Land models for the 2003–2011 period. All datasets show a significant bias with respect to models for boreal regions and high correlations over regions other than the tropical and boreal forest. All of the global SM datasets including AMSR-E NN were also evaluated against a large number of in situ measurements over four continents. Over Australia, all datasets show a strong level of agreement with in situ measurements. Models perform better over Europe and mountainous regions in North America. Remote sensing datasets (in particular NN and the Land Parameter Retrieval Model (LPRM)) perform as well as models for other North American sites and perform better than models over the Sahel region. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Locations of the in situ sensors.</p>
Full article ">Figure 2
<p>Schema of the methodology used in this study.</p>
Full article ">Figure 3
<p>Scatter plot of the soil temperature retrieved with the neural network using <math display="inline"> <semantics> <msup> <msub> <mi>T</mi> <mi mathvariant="normal">b</mi> </msub> <mi>H</mi> </msup> </semantics> </math> at 23 GHz and both <math display="inline"> <semantics> <msup> <msub> <mi>T</mi> <mi mathvariant="normal">b</mi> </msub> <mi>H</mi> </msup> </semantics> </math> and <math display="inline"> <semantics> <msup> <msub> <mi>T</mi> <mi mathvariant="normal">b</mi> </msub> <mi>V</mi> </msup> </semantics> </math> at 36 GHz (Model 7 in <a href="#remotesensing-08-00959-t002" class="html-table">Table 2</a>) versus ERA-Land soil temperature in the first layer (0–7 cm). The scatter plot is represented as the density of points in a logarithmic scale.Scatter plot of the soil temperature retrieved with the neural network versus ERA-Land soil temperature</p>
Full article ">Figure 4
<p>Scatter plot of the AMSR-E SM retrieved with the neural network (Model 15 in <a href="#remotesensing-08-00959-t003" class="html-table">Table 3</a>) versus SMOS L3 SM. The scatter plot is represented as the density of points in a logarithmic scale.</p>
Full article ">Figure 5
<p>Maps of different statistics metrics for the SMOS L3 SM dataset (left panels) and AMSR-E NN SM (right panels) computed from June 2010–October 2011: minimum values (<b>a</b>,<b>b</b>); mean (<b>c</b>,<b>d</b>); maximum (<b>e</b>,<b>f</b>); and standard deviation (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 6
<p>Comparison of AMSR-E NN SM and SMOS L3 SM from June 2010–October 2011. (<b>a</b>) Local Pearson correlation of the two SM datasets; (<b>b</b>) local Pearson correlation of the anomalies of the two SM datasets; (<b>c</b>) bias (mean AMSR-E NN SM minus mean SMOS L3 SM); (<b>d</b>) standard deviation of the difference of the two SM datasets.</p>
Full article ">Figure 7
<p>Comparison of AMSR-E NN SM and AMSR-E LPRM (applied to the C-band) from January 2003–October 2011. (<b>a</b>) Bias (mean of Land Parameter Retrieval Model (LPRM) SM minus the mean of NN SM); (<b>b</b>) Pearson correlation; (<b>c</b>) standard deviation of the difference of both datasets.</p>
Full article ">Figure 8
<p>Pearson correlation of remote sensing products with respect to Modern-Era Retrospective analysis for Research and Applications-Land (MERRA-Land) (left panels) and ERA-Land (right panels). The remote sensing products evaluated are, from top to bottom: ESA CCI (<b>a</b>,<b>b</b>); AMSR-E LPRM C-Band (<b>c</b>,<b>d</b>); AMSR-E NN (<b>e</b>,<b>f</b>); and AMSR-E Reg (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 9
<p>Bias of remote sensing products with respect to MERRA-Land (left panels) and ERA-Land (right panels). The remote sensing products evaluated are, from top to bottom: ESA CCI (<b>a</b>,<b>b</b>); AMSR-E LPRM C-Band (<b>c</b>,<b>d</b>); AMSR-E NN (<b>e</b>,<b>f</b>); and AMSR-E Reg (<b>g</b>,<b>h</b>). The bias has been computed as the mean of the remote sensing data minus the mean of model data, except for AMSR-E Reg and NN, for which the bias shown is the mean of the model data minus the mean of the remote sensing data.</p>
Full article ">Figure 10
<p>Standard deviation of the difference of remote sensing products with respect to MERRA-Land (left panels) and ERA-Land (right panels). The remote sensing products evaluated are, from top to bottom: ESA CCI (<b>a</b>,<b>b</b>); AMSR-E LPRM C-Band (<b>c</b>,<b>d</b>); AMSR-E NN (<b>e</b>,<b>f</b>); and AMSR-E Reg (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 11
<p>(<b>a</b>–<b>f</b>) Examples of time series: in situ measurements (black), AMSR-E NN (blue), AMSR-E LPRM (red), AMSR-E Reg (green), CCI (magenta), ERA-Land (cyan) and MERRA-Land (yellow).</p>
Full article ">
31180 KiB  
Article
Scanning, Multibeam, Single Photon Lidars for Rapid, Large Scale, High Resolution, Topographic and Bathymetric Mapping
by John J. Degnan
Remote Sens. 2016, 8(11), 958; https://doi.org/10.3390/rs8110958 - 18 Nov 2016
Cited by 79 | Viewed by 10719
Abstract
Several scanning, single photon sensitive, 3D imaging lidars are herein described that operate at aircraft above ground levels (AGLs) between 1 and 11 km, and speeds in excess of 200 knots. With 100 beamlets and laser fire rates up to 60 kHz, we, [...] Read more.
Several scanning, single photon sensitive, 3D imaging lidars are herein described that operate at aircraft above ground levels (AGLs) between 1 and 11 km, and speeds in excess of 200 knots. With 100 beamlets and laser fire rates up to 60 kHz, we, at the Sigma Space Corporation (Lanham, MD, USA), have interrogated up to 6 million ground pixels per second, all of which can record multiple returns from volumetric scatterers such as tree canopies. High range resolution has been achieved through the use of subnanosecond laser pulsewidths, detectors and timing receivers. The systems are presently being deployed on a variety of aircraft to demonstrate their utility in multiple applications including large scale surveying, bathymetry, forestry, etc. Efficient noise filters, suitable for near realtime imaging, have been shown to effectively eliminate the solar background during daytime operations. Geolocation elevation errors measured to date are at the subdecimeter level. Key differences between our Single Photon Lidars, and competing Geiger Mode lidars are also discussed. Full article
(This article belongs to the Special Issue Airborne Laser Scanning)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A comparison of Single Photon Lidars with conventional Discrete Return and Digitized Waveform lidars in interacting with a tree canopy (Courtesy of D. Harding, NASA GSFC).</p>
Full article ">Figure 2
<p>Leafcutter was the first Sigma Single Photon Lidar (SPL) to split the laser beam into 100 beamlets. In early mapping missions, the dual wedge scanner was used to generate either linear raster scans at 45° to the flight line or a conical scan with cone half angles up to 13.5 degrees. At the design AGL of 1 km, pixels on the ground were separated by 15 cm. Contiguous alongtrack and crosstrack mapping on a single pass was achieved by ensuring: (1) that the distance traveled by the aircraft during one scan cycle did not exceed the 1.5 m dimension of the single pulse array; and (2) that ground array patterns from subsequent pulses overlapped along the full circumference of the conical scan and the length of the linear scans.</p>
Full article ">Figure 3
<p>A collage of daytime images created on a single overflight by the Leafcutter SPL. The images in the left half were over low reflectance (10% to 15%) surfaces at above ground levels (AGLs) of 1 km or less while those in the right half were high reflectance cryospheric measurements in Greenland and Antarctica from AGLs up to 2.5 km. The images are color-coded according to the lidar-derived surface elevation (blue = low, red = high). Note the bathymetry results in the bottom two images.</p>
Full article ">Figure 4
<p>NASA Mini-ATM (Airborne Topographic Mapper) and its designated host aircraft, the Viking 300 micro-UAV.</p>
Full article ">Figure 5
<p>Moderate altitude HRQLS-1 and HRQLS-2 lidars and the King Air B200 host aircraft.</p>
Full article ">Figure 6
<p>The NASA MABEL pushbroom lidar, jointly developed by NASA Goddard Space Flight Center and Sigma Space Corporation, has successfully generated 2D surface profiles in Greenland from an AGL of 20 km. The surface returns are highly spatially correlated and stand out against the dense “salt-and-pepper” solar noise background resulting from the high reflectance (typically 80% to 96%) of snow and ice at 532 nm.</p>
Full article ">Figure 7
<p>The automated filtering of HRQLS-1 lidar data taken on a single overflight of a residential community in Oakland, MD. The raw/unfiltered point cloud data is taken with a range gate of 4.6 microseconds corresponding to a total range interval of 690 m. The color scheme is deep blue to red in order of increasing elevation, and it should be mentioned that the solar noise is equally dense below the surface but does not show up as well in the raw unfiltered image because of the poor contrast against the black background. The first stage filter isolates a 90 m interval that contains the surface data as well as roughly 13% of the total noise, and the second stage filter uses narrower range bins (~5 m) to eliminate the vast majority of the remaining noise.</p>
Full article ">Figure 8
<p>This color-coded elevation map of Garrett County, occupying approximately 1700 km<sup>2</sup> in the state of Maryland, was generated by HRQLS-1 from an AGL of 2.3 km. Total flight duration was approximately 12 h at an air speed of 278 km/h which included a 50% overlap between flight line, ferries, and turn maneuvers. The scanner was operating with a cone half angle of 17° resulting in a swath of 1.36 km and a mapping rate of 378 km<sup>2</sup>/h. Highest and lowest elevations are: red = 857 m, blue = 551 m.</p>
Full article ">Figure 9
<p>A Garrett County coal mine in which buildings, conveyor belts, and even black coal piles are clearly visible. Elevation Scales: Top Left red = 803.4 m, blue = 759.8 m; Bottom Left and Bottom Right red = 795.2 m, blue = 767.3 m.</p>
Full article ">Figure 10
<p>HRQLS-1 SPL point cloud profiles showing different growth patterns within a 1 square kilometer of forested area in Garrett County, MD. (<b>a</b>) Short even aged stand with little understory vegetation; (<b>b</b>) Uneven aged stand composed of tall trees and dense midstory vegetation; (<b>c</b>) Even aged stand with some mid and understory growth; (<b>d</b>) Tall open stand with distinct understory vegetation (Courtesy of the University of Maryland [<a href="#B11-remotesensing-08-00958" class="html-bibr">11</a>]).</p>
Full article ">Figure 11
<p>HRQLS-1 lidar image and digital color photograph of the area surrounding the Naval Post Graduate School in Monterey, California.</p>
Full article ">Figure 12
<p>“Fused” HRQLS-1 lidar-photographic 3D image of the Naval Post Graduate School in Monterrey, California.</p>
Full article ">Figure 13
<p><b>Top</b>: Colored HRQLS-1 lidar topo-bathymetric 3D pointcloud of a hilltop monastery and the beach at Pt. Lobos near Monterey, CA; <b>Bottom</b>: The bottom image shows the 2D lidar profile along the blue line in the top figure and extending from the monastery to the beach and into the Pacific Ocean to an optical depth of 17.3 m or a physical depth of 13 m. Vertical grid size = 10 m, Horizontal grid size = 50 m.</p>
Full article ">Figure 14
<p><b>Top</b>: Two passes of HRQLS-1 over a cruise ship docked at Ft. Lauderdale, Florida; <b>Bottom</b>: Multiple HRQLS-1 passes over a power line grid in North Carolina yielding over 40 points per square meter from an AGL of 1.83 km and an aircraft velocity of 296 km/h.</p>
Full article ">Figure 15
<p>(<b>a</b>) Fraction of pixels recording surface returns as a function of surface signal strength, <span class="html-italic">n</span>, and mean number of noise photons detected within a half range gate; (<b>b</b>) ratio of signal to noise counts as a function of the same two parameters.</p>
Full article ">Figure 16
<p>Surface detection probabilities for SPL and Geiger Mode (GM) lidars as a function of the unobscured signal strength for a tree canopy having a one way transmission of 40%. Unlike the GM lidar which has an unobscured signal strength that optimizes the surface detection probability, the SPL lidar can “power” through the canopy by increasing the laser pulse energy.</p>
Full article ">Figure 17
<p>The relative performance of SPL and GM lidars over a wide range of one-way tree canopy transmissions (<span class="html-italic">T<sub>c</sub></span> = 0.1 to 1) A value γ = 1 is assumed. The top left graph demonstrates that, as the tree canopy transmission decreases, the optimum unobscured signal for maximum penetration decreases, further reducing the detectability of the under canopy surface by the GM lidar. The bottom right graph describes the increasing advantages of the SPL technique in detecting the under canopy surface as the one way canopy transmission decreases.</p>
Full article ">
27221 KiB  
Article
Study of Subsidence and Earthquake Swarms in the Western Pakistan
by Jingqiu Huang, Shuhab D. Khan, Abduwasit Ghulam, Wanda Crupa, Ismail A. Abir, Abdul S. Khan, Din M. Kakar, Aimal Kasi and Najeebullah Kakar
Remote Sens. 2016, 8(11), 956; https://doi.org/10.3390/rs8110956 - 18 Nov 2016
Cited by 21 | Viewed by 7776
Abstract
In recent years, the Quetta Valley and surrounding areas have experienced unprecedented levels of subsidence, which has been attributed mainly to groundwater withdrawal. However, this region is also tectonically active and is home to several regional strike-slip faults, including the north–south striking left-lateral [...] Read more.
In recent years, the Quetta Valley and surrounding areas have experienced unprecedented levels of subsidence, which has been attributed mainly to groundwater withdrawal. However, this region is also tectonically active and is home to several regional strike-slip faults, including the north–south striking left-lateral Chaman Fault System. Several large earthquakes have occurred recently in this area, including one deadly Mw 6.4 earthquake that struck on 28 October 2008. This study integrated Interferometric Synthetic Aperture Radar (InSAR) results with GPS, gravity, seismic reflection profiles, and earthquake centroid-moment-tensor (CMT) data to identify the impact of tectonic and anthropogenic processes on subsidence and earthquake patterns in this region. To detect and map the spatial-temporal features of the processes that led to the surface deformation, this study used two Synthetic Aperture Radar (SAR) time series, i.e., 15 Phased Array L-band Synthetic Aperture Radar (PALSAR) images acquired by an Advanced Land Observing Satellite (ALOS) from 2006–2011 and 40 Environmental Satellite (ENVISAT) Advanced Synthetic Aperture Radar (ASAR) images spanning 2003–2010. A Small Baseline Subset (SBAS) technique was used to investigate surface deformation. Five seismic lines totaling ~60 km, acquired in 2003, were used to map the blind thrust faults beneath a Quaternary alluvium layer. The median filtered SBAS-InSAR average velocity profile supports groundwater withdrawal as the dominant source of subsidence, with some contribution from tectonic subsidence in the Quetta Valley. Results of SBAS-InSAR multi-temporal analysis provide a better explanation for the pre-, co-, and post-seismic displacement pattern caused by the 2008 earthquake swarms across two strike-slip faults. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of major structures and earthquake hypocenters for the Quetta Valley and surrounding areas. The focal mechanisms of the October–December 2008 earthquakes are shown. Thrust faults in the Quetta Valley from Jones et al. [<a href="#B17-remotesensing-08-00956" class="html-bibr">17</a>] are also shown. Two blind thrust faults identified using 2D seismic data are shown in indigo. The beach ball symbol indicates the centroid-moment-tensor (CMT) solution [<a href="#B18-remotesensing-08-00956" class="html-bibr">18</a>,<a href="#B19-remotesensing-08-00956" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Water level decline from 1987–2006 in alluvial and carbonate aquifers of Quetta Valley [<a href="#B21-remotesensing-08-00956" class="html-bibr">21</a>]. The hydrographs of these wells clearly indicate a constant decline in their water levels. The water level decline rate was lower at ~1 m/year in alluvial aquifers and higher at ~5 m/year in carbonate aquifers. Locations of water wells in alluvium and carbonate are indicated by blue and red dots, respectively, in <a href="#remotesensing-08-00956-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Normal baselines (y-axis) and the acquisition dates (x-axis). The yellow diamonds refer to the Small Baseline Subset (SBAS) super master image (1 October 2009, for Phased Array L-band Synthetic Aperture Radar (PALSAR) and 8 November 2007, for Advanced Synthetic Aperture Radar (ASAR).</p>
Full article ">Figure 4
<p>Bouguer gravity anomaly ranging from −85.6 to −223.1 mGal shows the crustal density variations in the Quetta Valley and surrounding areas. AA’ is a 2D gravity profile used in <a href="#remotesensing-08-00956-f007" class="html-fig">Figure 7</a>, and BB’ is a 2D gravity profile extracted across the Karahi and Harnai Faults, shown in <a href="#remotesensing-08-00956-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 5
<p>(<b>Top</b>) Seismic profile 4 and 5 show the top of Shrinab Limestone (red horizons) and two thrust faults, shown on profile 4, dipping westward. (<b>Bottom</b>) Seismic profile 1, 2 and 3 show the top of Shrinab Limestone (red horizons) and many blind thrust faults. Only one thrust fault mapped on profile 3 serves as a boundary between the Shrinab and Chiltan Formations.</p>
Full article ">Figure 6
<p>(<b>A</b>) SBAS results generated from forty descending ENVISAT ASAR images from 2003–2010 show surface deformations ranging from −5.8–1.8 mm/year in the Quetta Valley and surrounding areas. A 2D velocity profile, CC’, was extracted from the InSAR results. The low coherence area is marked by a dotted line on the CC’ velocity profile. (<b>B</b>) GPS rates from 2006–2016 in a vertical direction around –7.7 mm/year, which matches the InSAR vertical movement rate in the Quetta Valley. (<b>C</b>) A five-point median filter was applied on the CC’ cross-profile (<a href="#remotesensing-08-00956-f006" class="html-fig">Figure 6</a>A) to distinguish the subsidence caused by tectonic processes (shown as a red line) from subsidence caused by anthropogenic processes (water withdrawal) (shown as a green line).</p>
Full article ">Figure 7
<p>Schematic cross-section AA’ (parallel with seismic profile 3) modified from a previous study [<a href="#B20-remotesensing-08-00956" class="html-bibr">20</a>] to explain subsidence in the Quetta Valley and surrounding areas. The Bouguer anomaly profile of AA’ shows highs and lows forming a trend likely related to mountain ranges and basins. This model shows the activity of blind thrust faults 1 and 2 mapped in subsurface using 2D seismic profiles, which are part of the Chiltan Thrust System. High-risk locations for potential fissures are shown near the Murdar Range based on differential compaction theory in <a href="#sec2dot5-remotesensing-08-00956" class="html-sec">Section 2.5</a>.</p>
Full article ">Figure 8
<p>(<b>A</b>) SBAS results generated from 15 ascending PALSAR scenes from 2006 to 2011 show a surface deformation velocity ranging from −10–40 mm/year. (<b>B</b>) The profile BB’ shows changes in gravity across the Karahi and Harnai Faults. The gravity data indicate the presence of three different fault blocks that cross faults. (<b>C</b>) The SBAS multi-temporal analysis shows a change in pre-, co-, and post-seismic velocity rates for zones 1, 2 and 3 at the epicenters of the 2008 earthquake swarms.</p>
Full article ">
169 KiB  
Editorial
Preface: The Environmental Mapping and Analysis Program (EnMAP) Mission: Preparing for Its Scientific Exploitation
by Saskia Foerster, Véronique Carrère, Michael Rast and Karl Staenz
Remote Sens. 2016, 8(11), 957; https://doi.org/10.3390/rs8110957 - 17 Nov 2016
Cited by 11 | Viewed by 4924
Abstract
The imaging spectroscopy mission EnMAP aims to assess the state and evolution of terrestrial and aquatic ecosystems, examine the multifaceted impacts of human activities, and support a sustainable use of natural resources. Once in operation (scheduled to launch in 2019), EnMAP will provide [...] Read more.
The imaging spectroscopy mission EnMAP aims to assess the state and evolution of terrestrial and aquatic ecosystems, examine the multifaceted impacts of human activities, and support a sustainable use of natural resources. Once in operation (scheduled to launch in 2019), EnMAP will provide high-quality observations in the visible to near-infrared and shortwave-infrared spectral range. The scientific preparation of the mission comprises an extensive science program. This special issue presents a collection of research articles, demonstrating the potential of EnMAP for various applications along with overview articles on the mission and software tools developed within its scientific preparation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">
20037 KiB  
Article
Evaluation of the Quality of NDVI3g Dataset against Collection 6 MODIS NDVI in Central Europe between 2000 and 2013
by Anikó Kern, Hrvoje Marjanović and Zoltán Barcza
Remote Sens. 2016, 8(11), 955; https://doi.org/10.3390/rs8110955 - 17 Nov 2016
Cited by 40 | Viewed by 7368
Abstract
Remote sensing provides invaluable insight into the dynamics of vegetation with global coverage and reasonable temporal resolution. Normalized Difference Vegetation Index (NDVI) is widely used to study vegetation greenness, production, phenology and the responses of ecosystems to climate fluctuations. The extended global NDVI3g [...] Read more.
Remote sensing provides invaluable insight into the dynamics of vegetation with global coverage and reasonable temporal resolution. Normalized Difference Vegetation Index (NDVI) is widely used to study vegetation greenness, production, phenology and the responses of ecosystems to climate fluctuations. The extended global NDVI3g dataset created by Global Inventory Modeling and Mapping Studies (GIMMS) has an exceptional 32 years temporal coverage. Due to the methodology that was used to create NDVI3g inherent noise and uncertainty is present in the dataset. To evaluate the accuracy and uncertainty of application of NDVI3g at regional scale we used Collection-6 data from the MODerate resolution Imaging Spectroradiometer (MODIS) sensor on board satellite Terra as a reference. After noise filtering, statistical harmonization of the NDVI3g dataset was performed for Central Europe based on MOD13 NDVI. Mean seasonal NDVI profiles, start, end and length of the growing season, magnitude and timing of peak NDVI were calculated from NDVI3g (original, noise filtered and harmonized) and MODIS NDVI and compared with each other. NDVI anomalies were also compared and evaluated using simple climate sensitivity metrics. The results showed that (1) the original NDVI3g has limited applicability in Central Europe, which was also implied by the significant disagreement between the NDVI3g and MODIS NDVI datasets; (2) the harmonization of NDVI3g with MODIS NDVI is promising since the newly created dataset showed improved quality for diverse vegetation metrics. For NDVI anomaly detection NDVI3g showed limited applicability, even after harmonization. Climate–NDVI relationships are not represented well by NDVI3g. The presented results can help researchers to assess the expected quality of the NDVI3g-based studies in Central Europe. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mean NDVI cycles for the 14 year time period for different land cover types calculated from NDVI3g<sub>O</sub>, NDVI3g<sub>F</sub> and NDVI<sub>MODIS</sub>. The light gray shaded area represents the area covered by the individual pixels for NDVI3g<sub>O</sub>, while the light green area shows the same for NDVI<sub>MODIS</sub> (area for NDVI3g<sub>F</sub> is not shown for clarity). The number of pixels (n) used to create the NDVI cycles are also shown.</p>
Full article ">Figure 2
<p>Bias map of the original NDVI3g<sub>O</sub> relative to NDVI<sub>MODIS</sub> during the overlapping 14 years, calculated for the time period between the second half of April and the first part of September. Positive bias means higher NDVI3g<sub>O</sub>.</p>
Full article ">Figure 3
<p>Frequency distribution of (<b>a</b>) SOS bias; (<b>b</b>) EOS bias; (<b>c</b>) length of growing season bias during the time period of 2000–2013 for NDVI3g<sub>O</sub> (black), NDVI3g<sub>F</sub> (light blue) and NDVI3g<sub>H</sub> (dark blue) datasets relative to NDVI<sub>MODIS</sub>. Positive bias means earlier SOS, EOS and shorter LOS for MODIS (n = 20191).</p>
Full article ">Figure 4
<p>Frequency distribution of (<b>a</b>) maximum NDVI bias and (<b>b</b>) the bias of the maximum NDVI timing during the 2000–2013 time period for NDVI3g<sub>O</sub> (black), NDVI3g<sub>F</sub> (light blue) and NDVI3g<sub>H</sub> (dark blue) relative to NDVI<sub>MODIS</sub>. Positive NDVI bias means higher maximum NDVI for NDVI3g, while the positive bias of peak NDVI dates means earlier maximum NDVI for MODIS (n = 20191).</p>
Full article ">Figure 5
<p>NDVI anomaly maps for the first half of September 2003 after a severe heatwave and exceptional drought in Europe for (<b>a</b>) NDVI3g<sub>O</sub>; (<b>b</b>) NDVI3g<sub>F</sub>; (<b>c</b>) NDVI3g<sub>H</sub> and (<b>d</b>) (Terra) NDVI<sub>MODIS</sub>, where the anomaly was referring to the period of 2000–2013.</p>
Full article ">Figure 6
<p>Map of R values between the anomalies derived from Terra/NDVI<sub>MODIS</sub> and (<b>a</b>) NDVI3g<sub>O</sub>, (<b>b</b>) NDVI3g<sub>F</sub> and (<b>c</b>) NDVI3g<sub>H</sub> datasets during the overlapping 14 years taking into account only the time period between the second half of April and the first part of September period. Pixels with dots inside indicate statistically significant R (p &lt; 0.01).</p>
Full article ">Figure 7
<p>Box-whisker plots of the correlation coefficients between the temperature anomaly and the NDVI anomaly in August derived from the NDVI3g<sub>O</sub> (dark blue, the first column in every LC group), NDVI3g<sub>F</sub> (purple, the second column in every LC group), NDVI3g<sub>H</sub> (light blue, the third column in every LC group) and NDVI<sub>MODIS</sub> (green, the fourth column in every group) datasets for all vegetated pixels of the study area, and for different land cover types. (Box-whisker plots are indicating maximum, upper quartile, median, lower quartile, and minimum values.)</p>
Full article ">Figure 8
<p>Map of R values between the temperature anomaly in August and the NDVI anomaly in August derived from (<b>a</b>) the NDVI3g<sub>O</sub>; (<b>b</b>) NDVI3g<sub>F</sub>; (<b>c</b>) NDVI3g<sub>H</sub> and (<b>d</b>) NDVI<sub>MODIS</sub> datasets. Pixels with dots inside indicate statistically significant R (p &lt; 0.01).</p>
Full article ">Figure 8 Cont.
<p>Map of R values between the temperature anomaly in August and the NDVI anomaly in August derived from (<b>a</b>) the NDVI3g<sub>O</sub>; (<b>b</b>) NDVI3g<sub>F</sub>; (<b>c</b>) NDVI3g<sub>H</sub> and (<b>d</b>) NDVI<sub>MODIS</sub> datasets. Pixels with dots inside indicate statistically significant R (p &lt; 0.01).</p>
Full article ">Figure 9
<p>Box-whisker plots of the correlation coefficients between the precipitation anomaly in August and the NDVI anomaly in September derived from the NDVI3g<sub>O</sub> (dark blue, the first column in every LC group), NDVI3g<sub>F</sub> (purple, the second column in every LC group), NDVI3g<sub>H</sub> (light blue, the third column in every LC group) and NDVI<sub>MODIS</sub> (green, the fourth column in every group) datasets for all vegetated pixels of the study area, and for different land cover types. (Box-whisker plots are indicating maximum, upper quartile, median, lower quartile, and minimum values.)</p>
Full article ">Figure 10
<p>Map of R values between the precipitation anomaly in August and the NDVI anomaly in September derived from (<b>a</b>) the NDVI3g<sub>O</sub>; (<b>b</b>) NDVI3g<sub>F</sub>; (<b>c</b>) NDVI3g<sub>H</sub> and (<b>d</b>) NDVI<sub>MODIS</sub> datasets. Pixels with dots inside indicate statistically significant R (p &lt; 0.01).</p>
Full article ">
4190 KiB  
Article
Random Forest Classification of Wetland Landcovers from Multi-Sensor Data in the Arid Region of Xinjiang, China
by Shaohong Tian, Xianfeng Zhang, Jie Tian and Quan Sun
Remote Sens. 2016, 8(11), 954; https://doi.org/10.3390/rs8110954 - 16 Nov 2016
Cited by 155 | Viewed by 12106
Abstract
The wetland classification from remotely sensed data is usually difficult due to the extensive seasonal vegetation dynamics and hydrological fluctuation. This study presents a random forest classification approach for the retrieval of the wetland landcover in the arid regions by fusing the Pléiade-1B [...] Read more.
The wetland classification from remotely sensed data is usually difficult due to the extensive seasonal vegetation dynamics and hydrological fluctuation. This study presents a random forest classification approach for the retrieval of the wetland landcover in the arid regions by fusing the Pléiade-1B data with multi-date Landsat-8 data. The segmentation of the Pléiade-1B multispectral image data was performed based on an object-oriented approach, and the geometric and spectral features were extracted for the segmented image objects. The normalized difference vegetation index (NDVI) series data were also calculated from the multi-date Landsat-8 data, reflecting vegetation phenological changes in its growth cycle. The feature set extracted from the two sensors data was optimized and employed to create the random forest model for the classification of the wetland landcovers in the Ertix River in northern Xinjiang, China. Comparison with other classification methods such as support vector machine and artificial neural network classifiers indicates that the random forest classifier can achieve accurate classification with an overall accuracy of 93% and the Kappa coefficient of 0.92. The classification accuracy of the farming lands and water bodies that have distinct boundaries with the surrounding land covers was improved 5%–10% by making use of the property of geometric shapes. To remove the difficulty in the classification that was caused by the similar spectral features of the vegetation covers, the phenological difference and the textural information of co-occurrence gray matrix were incorporated into the classification, and the main wetland vegetation covers in the study area were derived from the two sensors data. The inclusion of phenological information in the classification enables the classification errors being reduced down, and the overall accuracy was improved approximately 10%. The results show that the proposed random forest classification by fusing multi-sensor data can retrieve better wetland landcover information than the other classifiers, which is significant for the monitoring and management of the wetland ecological resources in arid areas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographic location of the study area.</p>
Full article ">Figure 2
<p>The normalized difference vegetation index (NDVI) time series curves of three wetland vegetation after harmonic analysis.</p>
Full article ">Figure 3
<p>The flowchart for the random forest classification of the wetland. OLI: Operational land imager; NDVI: normalized difference vegetation index; FFT: fast Fourier transform.</p>
Full article ">Figure 4
<p>The classification results of the RFC (<b>a</b>); SVM (<b>b</b>); and ANN (<b>c</b>) models.</p>
Full article ">Figure 5
<p>The feature distribution in different land cover types (<b>a</b>) normalized area; (<b>b</b>) normalized rectangle fitting; (<b>c</b>) normalized difference water index (NDWI); (<b>d</b>) normalized difference vegetation index (NDVI); (<b>e</b>) normalized variation texture; (<b>f</b>) 10,000 × reflectance of Pléiades-1B blue band).</p>
Full article ">Figure 6
<p>Relative contributions of the features to the classification.</p>
Full article ">
5264 KiB  
Article
Water Budget Analysis within the Surrounding of Prominent Lakes and Reservoirs from Multi-Sensor Earth Observation Data and Hydrological Models: Case Studies of the Aral Sea and Lake Mead
by Alka Singh, Florian Seitz, Annette Eicker and Andreas Güntner
Remote Sens. 2016, 8(11), 953; https://doi.org/10.3390/rs8110953 - 16 Nov 2016
Cited by 12 | Viewed by 6759
Abstract
The hydrological budget of a region is determined based on the horizontal and vertical water fluxes acting in both inward and outward directions. These integrated water fluxes vary, altering the total water storage and consequently the gravitational force of the region. The time-dependent [...] Read more.
The hydrological budget of a region is determined based on the horizontal and vertical water fluxes acting in both inward and outward directions. These integrated water fluxes vary, altering the total water storage and consequently the gravitational force of the region. The time-dependent gravitational field can be observed through the Gravity Recovery and Climate Experiment (GRACE) gravimetric satellite mission, provided that the mass variation is above the sensitivity of GRACE. This study evaluates mass changes in prominent reservoir regions through three independent approaches viz. fluxes, storages, and gravity, by combining remote sensing products, in-situ data and hydrological model outputs using WaterGAP Global Hydrological Model (WGHM) and Global Land Data Assimilation System (GLDAS). The results show that the dynamics revealed by the GRACE signal can be better explored by a hybrid method, which combines remote sensing-based reservoir volume estimates with hydrological model outputs, than by exclusive model-based storage estimates. For the given arid/semi-arid regions, GLDAS based storage estimations perform better than WGHM. Full article
(This article belongs to the Special Issue Remote Sensing of Climate Change and Water Resources)
Show Figures

Figure 1

Figure 1
<p>Study box for the Lake Mead region and the Aral Sea region. River discharge: 1 = Colorado River inflow, 2 = Virgin River, 3 = Muddy River, 4 = Colorado River outflow, 5 = Syr Darya and 6 = Amu Darya.</p>
Full article ">Figure 2
<p>Runoff: (<b>left</b>) Lake Mead and (<b>right</b>) The Aral Sea.</p>
Full article ">Figure 3
<p>Precipitation: (<b>left</b>) The Lake Mead region and (<b>right</b>) The Aral Sea region.</p>
Full article ">Figure 4
<p>ET: (<b>left</b>) The Lake Mead region and (<b>right</b>) The Aral Sea region.</p>
Full article ">Figure 5
<p>Mean reduced reservoir volume: (<b>left</b>) Lake Mead and (<b>right</b>) The Aral Sea.</p>
Full article ">Figure 6
<p>Mean reduced Snow Water Equivalent: (<b>left</b>) The Lake Mead region and (<b>right</b>) The Aral Sea region.</p>
Full article ">Figure 7
<p>Mean reduced Soil Moisture: (<b>left</b>) The Lake Mead region and (<b>right</b>) The Aral Sea region.</p>
Full article ">Figure 8
<p>Gravity Recovery and Climate Experiment (GRACE)-derived trend of the equivalent water height (meter/year) between 2003 and 2014. The size of the study area was chosen according to the mascon grid. (<b>left</b>) The Lake Mead region is 3° × 3° where Lake Mead is located at the center. (<b>right</b>) The Aral Sea region is 4° × 6°covering the entire lake and two mascon grid cells.</p>
Full article ">Figure 9
<p>GRACE-derived mass variations with the uncertainty range of the measurements as provided by GRACE Tellus: (<b>left</b>) The Lake Mead region, (<b>right</b>) The Aral Sea region</p>
Full article ">Figure 10
<p>Lake Mead: (<b>top</b>) Net surface runoff of the lake: inflow–outflow and (<b>bottom</b>) Reservoir volume variation compared with the hydrological fluxes.</p>
Full article ">Figure 11
<p>Lake Mead region (3° × 3°) mass variations observed by net fluxes, net storages, and GRACE: (<b>top</b>) Monthly mass variations and (<b>bottom</b>) Non-seasonal water storage variability. All time series in the lower panel have been reduced for their mean (i.e., mean value over the study period). The large numbers at the top of the figure are the periods of different mass evolution, discussed in <a href="#sec4dot2-remotesensing-08-00953" class="html-sec">Section 4.2</a>. Here symbol δ indicates derivative and <math display="inline"> <semantics> <mrow> <mo>∫</mo> </mrow> </semantics> </math> indicates integral of the signal.</p>
Full article ">Figure 12
<p>Aral Sea region (4° × 6°) mass variations observed by net flux, net storages, and GRACE: (<b>top</b>) Monthly mass variations and (<b>bottom</b>) Non-seasonal water storage variability. All time series in the lower panel have been reduced for their mean (i.e., mean value over the study period). The large numbers at the top of the figure are the periods of different mass evolution, discussed in <a href="#sec4dot3-remotesensing-08-00953" class="html-sec">Section 4.3</a>. Here symbol δ indicates derivative and <math display="inline"> <semantics> <mrow> <mo>∫</mo> </mrow> </semantics> </math> indicates integral of the signal.</p>
Full article ">Figure 13
<p>Total water storage (TWS) observed by GRACE compared with the best estimates and the reservoir volume: (<b>left</b>) The Lake Mead region and (<b>right</b>) The Aral Sea region.</p>
Full article ">
8288 KiB  
Article
Using Landsat, MODIS, and a Biophysical Model to Evaluate LST in Urban Centers
by Mohammad Z. Al-Hamdan, Dale A. Quattrochi, Lahouari Bounoua, Asia Lachir and Ping Zhang
Remote Sens. 2016, 8(11), 952; https://doi.org/10.3390/rs8110952 - 16 Nov 2016
Cited by 13 | Viewed by 8194
Abstract
In this paper, we assessed and compared land surface temperature (LST) in urban centers using data from Landsat, MODIS, and the Simple Biosphere model (SiB2). We also evaluated the sensitivity of the model’s LST to different land cover types, fractions (percentages), and emissivities [...] Read more.
In this paper, we assessed and compared land surface temperature (LST) in urban centers using data from Landsat, MODIS, and the Simple Biosphere model (SiB2). We also evaluated the sensitivity of the model’s LST to different land cover types, fractions (percentages), and emissivities compared to reference points derived from Landsat thermal data. This was demonstrated in three climatologically- and morphologically-different cities of Atlanta, GA, New York, NY, and Washington, DC. Our results showed that in these cities SiB2 was sensitive to both the emissivity and the land cover type and fraction, but much more sensitive to the latter. The practical implications of these results are rather significant since they imply that the SiB2 model can be used to run different scenarios for evaluating urban heat island (UHI) mitigation strategies. This study also showed that using detailed emissivities per land cover type and fractions from Landsat-derived data caused a convergence of the model results towards the Landsat-derived LST for most of the studied cases. This study also showed that SiB2 LSTs are closer in magnitude to Landsat-derived LSTs than MODIS-derived LSTs. It is important, however, to emphasize that both Landsat and MODIS LSTs are not direct observations and, as such, do not represent a ground truth. More studies will be needed to compare these results to in situ LST data and provide further validation. Full article
(This article belongs to the Special Issue Multi-Sensor and Multi-Data Integration in Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The spatial domains of the study areas (5 km × 5 km cells) and NLCD-2001 LC within those areas.</p>
Full article ">Figure 2
<p>An example of Landsat-derived LST for Atlanta on 1 October 2001 and inputs needed for that derivation.</p>
Full article ">Figure 3
<p>An example of Landsat-derived LST for New York on 14 April 2001 and inputs needed for that derivation.</p>
Full article ">Figure 4
<p>An example of Landsat-derived LST for Washington, DC on 26 August 2001 and inputs needed for that derivation (the extracted block represents the 5 km × 5 km cell where analyses were performed).</p>
Full article ">Figure 5
<p>Differences in daily composite of canopy temperature computed with Landcover type-dependent emissivity and constant emissivity. Results are averaged during winter (December, January, and February) and summer (June, July and August).</p>
Full article ">Figure 6
<p>Time series of LST results from all SiB2 model scenarios, Landsat, and MODIS for Atlanta, Washington, DC, and New York.</p>
Full article ">Figure 7
<p>Landsat LST versus MODIS LST or SiB2 LST for Atlanta, Washington, DC, and New York.</p>
Full article ">Figure 8
<p>Root mean square difference (RMSD) between Landsat thermal data (TM, ETM<sup>+</sup>, and TM and ETM<sup>+</sup> combined), each SiB2 model scenario, and MODIS.</p>
Full article ">Figure 9
<p>Comparison of nighttime SiB2 simulated canopy temperature for each model scenario and MODIS nighttime LST for the 5 km × 5 km cell in Washington, DC.</p>
Full article ">Figure 10
<p>Landsat TM- and ETM<sup>+</sup>-derived land surface temperatures and SiB2 canopy temperature for each land cover type present in the 5 km × 5 km cell in Washington, DC.</p>
Full article ">Figure 11
<p>Comparison of SiB2 canopy temperature (<b>blue</b>) to Landsat TM (<b>a</b>–<b>c</b>) and ETM<sup>+</sup> (<b>d</b>–<b>f</b>) LST (<b>red</b>) for each land cover type in the 5 km × 5 km cell in Washington, DC.</p>
Full article ">Figure 12
<p>SUHI amplitude computed using SiB2 Canopy temperature, Landsat TM- and Landsat ETM<sup>+</sup>-derived LST over the 5 km × 5 km cell in Washington, DC.</p>
Full article ">
26592 KiB  
Article
Investigation on Mining Subsidence Based on Multi-Temporal InSAR and Time-Series Analysis of the Small Baseline Subset—Case Study of Working Faces 22201-1/2 in Bu’ertai Mine, Shendong Coalfield, China
by Chao Ma, Xiaoqian Cheng, Yali Yang, Xiaoke Zhang, Zengzhang Guo and Youfeng Zou
Remote Sens. 2016, 8(11), 951; https://doi.org/10.3390/rs8110951 - 16 Nov 2016
Cited by 61 | Viewed by 7712
Abstract
High-intensity coal mining (large mining height, shallow mining depth, and rapid advancing) frequently causes large-scale ground damage within a short period of time. Understanding mining subsidence under high-intensity mining can provide a basis for mining-induced damage assessment, land remediation in a subsidence area, [...] Read more.
High-intensity coal mining (large mining height, shallow mining depth, and rapid advancing) frequently causes large-scale ground damage within a short period of time. Understanding mining subsidence under high-intensity mining can provide a basis for mining-induced damage assessment, land remediation in a subsidence area, and ecological reconstruction in vulnerable ecological regions in Western China. In this study, the mining subsidence status of Shendong Coalfield was investigated and analyzed using two-pass differential interferometric synthetic aperture radar (DInSAR) technology based on high-resolution synthetic aperture radar data (RADARSAT-2 precise orbit, multilook fine, 5 m) collected from 20 January 2012 to June 2013. Surface damages in Shendong Coalfield over a period of 504 days under open-pit mining and underground mining were observed. Ground deformation of the high-intensity mining working faces 22201-1/2 in Bu’ertai Mine, Shendong Coalfield was monitored using small baseline subset (SBAS) InSAR technology. (1) DInSAR detected and located 85 ground deformation areas (including ground deformations associated with past-mining activity). The extent of subsidence in Shendong Coalfield presented a progressive increase at an average monthly rate of 13.09 km2 from the initial 54.98 km2 to 225.20 km2, approximately, which accounted for 7% of the total area of Shendong Coalfield; (2) SBAS-InSAR reported that the maximum cumulative subsidence area reached 5.58 km2 above the working faces 22201-1/2. The advance speed of ground destruction (7.9 m/day) was nearly equal to that of underground mining (8.1 m/day). Full article
(This article belongs to the Special Issue Earth Observations for Geohazards)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geographical location of Shendong Coalfield. Background data: Satellite imagery acquired by Landsat 7 on 10 April 2013.</p>
Full article ">Figure 2
<p>The photographs of subsidence damage types in Shendong Coalfield (2009–2015).</p>
Full article ">Figure 3
<p>Mining layout plan of 22201-1/2 long wall.</p>
Full article ">Figure 4
<p>Distribution of the spatiotemporal baseline. (<b>a</b>) Spatiotemporal baseline distribution of consecutive DInSAR; (<b>b</b>) Spatiotemporal baseline distribution of cumulative DInSAR.</p>
Full article ">Figure 5
<p>The results of the consecutive DInSAR pattern (January 2012–June 2013).</p>
Full article ">Figure 5 Cont.
<p>The results of the consecutive DInSAR pattern (January 2012–June 2013).</p>
Full article ">Figure 6
<p>The results of the cumulative DInSAR pattern (January 2012–June 2013).</p>
Full article ">Figure 6 Cont.
<p>The results of the cumulative DInSAR pattern (January 2012–June 2013).</p>
Full article ">Figure 7
<p>Spatiotemporal baseline distribution of the connection diagram and unwrapping results: (<b>a</b>) Spatiotemporal baseline distribution of the connection diagram; (<b>b</b>) 3D unwrapping sub-pairs.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>q</b>) Interferograms of SBAS-InSAR.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>–<b>q</b>) Interferograms of SBAS-InSAR.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>–<b>q</b>) Interferograms of SBAS-InSAR.</p>
Full article ">Figure 9
<p>Example I: verification and validation of Interferometric Measurement.</p>
Full article ">Figure 10
<p>Example II: verification and validation of Interferometric Measurement.</p>
Full article ">Figure 11
<p>Example III: verification and validation of Interferometric Measurement.</p>
Full article ">Figure 11 Cont.
<p>Example III: verification and validation of Interferometric Measurement.</p>
Full article ">Figure 12
<p>Interpretation results of the consecutive interferometry (January 2012–June2013).</p>
Full article ">Figure 13
<p>Interpretation results of the cumulative interferometry (January 2012–June2013).</p>
Full article ">Figure 14
<p>(<b>A</b>–<b>H</b>) Subsidence areas obtained from the InSAR time-series analysis (Western region to the Wulanmulun River).</p>
Full article ">Figure 15
<p>(<b>a</b>–<b>l</b>) Time-series analysis based on multi-temporal DInSAR.</p>
Full article ">Figure 15 Cont.
<p>(<b>a</b>–<b>l</b>) Time-series analysis based on multi-temporal DInSAR.</p>
Full article ">Figure 16
<p>Comparative analysis between multi-temporal DInSAR and SBAS-InSAR. (<b>a</b>) Time-series subsidence areas of the consecutive DInSAR; (<b>b</b>) Time-series subsidence curves of DInSAR; (<b>c</b>) Time-series subsidence areas of SBAS-InSAR; (<b>d</b>) Time-series subsidence curves of SBAS-InSAR.</p>
Full article ">
3038 KiB  
Article
An Alternative Quality Control Technique for Mineral Chemistry Analysis of Portland Cement-Grade Limestone Using Shortwave Infrared Spectroscopy
by Nasrullah Zaini, Freek Van der Meer, Frank Van Ruitenbeek, Boudewijn De Smeth, Fadli Amri and Caroline Lievens
Remote Sens. 2016, 8(11), 950; https://doi.org/10.3390/rs8110950 - 15 Nov 2016
Cited by 25 | Viewed by 7040
Abstract
Shortwave infrared (SWIR) spectroscopy can be applied directly to analyze the mineral chemistry of raw or geologic materials. It provides diagnostic spectral characteristics of the chemical composition of minerals, information that is invaluable for the identification and quality control of such materials. The [...] Read more.
Shortwave infrared (SWIR) spectroscopy can be applied directly to analyze the mineral chemistry of raw or geologic materials. It provides diagnostic spectral characteristics of the chemical composition of minerals, information that is invaluable for the identification and quality control of such materials. The present study aims to investigate the potential of SWIR spectroscopy as an alternative quality control technique for the mineral chemistry analysis of Portland cement-grade limestone. We used the spectroscopic (wavelength position and depth of absorption feature) and geochemical characteristics of limestone samples to estimate the abundance and composition of carbonate and clay minerals on rock surfaces. The depth of the carbonate (CO3) and Al-OH absorption features are linearly correlated with the contents of CaO and Al2O3 in the samples, respectively, as determined by portable X-ray fluorescence (PXRF) measurements. Variations in the wavelength position of CO3 and Al-OH absorption features are related to changes in the chemical compositions of the samples. The results showed that the dark gray and light gray limestone samples are better suited for manufacturing Portland cement clinker than the dolomitic limestone samples. This finding is based on the CaO, MgO, Al2O3, and SiO2 concentrations and compositions. The results indicate that SWIR spectroscopy is an appropriate approach for the chemical quality control of cement raw materials. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Carbonate rock samples: (<b>a</b>) <b>dark gray</b>; and (<b>b</b>) <b>light gray</b> limestone samples collected from the Lhoknga limestone quarry of PT. Lafarge Cement Indonesia, Aceh Besar, Indonesia; and (<b>c</b>) dolomitic limestone sample collected from the Bédarieux dolomite mine, Hérault department of the Languedoc-Roussillon region, southern France.</p>
Full article ">Figure 2
<p>Examples of hand specimens of carbonate rock samples collected from the mines: (<b>a</b>) <b>dark gray</b> limestones (samples B28 and B45); (<b>b</b>) <b>light gray</b> limestones (samples B1 and B41); and (<b>c</b>) dolomitic limestones (samples F18 and F20), with white circles pointing out the locations of portable X-ray fluorescence (PXRF) and Analytical Spectral Device (ASD) spot measurements on both fresh surfaces of the rock samples.</p>
Full article ">Figure 3
<p>Continuum removed spectra of <b>dark gray</b> limestone samples: (<b>a</b>) sample B28; and (<b>b</b>) sample B45 derived from five ASD spot measurements on the rock sample surfaces in the SWIR wavelength region. The spectral curves exhibit variations of depths of absorption features at ~2.20 µm and ~2.34 µm corresponding to changes in mineral contents on the rock surfaces.</p>
Full article ">Figure 4
<p>Continuum removed spectra of <b>light gray</b> limestone samples: (<b>a</b>) sample B1; and (<b>b</b>) sample B41 derived from five ASD spot measurements on the rock sample surfaces in the SWIR wavelength region, showing the variability in the depth of carbonate absorption feature at ~2.34 µm.</p>
Full article ">Figure 5
<p>Continuum removed spectra of dolomitic limestone samples: (<b>a</b>) sample F18; and (<b>b</b>) sample F20 derived from five ASD spot measurements on the rock sample surfaces in the SWIR wavelength region, showing the variability in the wavelength position and depth of carbonate absorption feature at ~2.32 µm.</p>
Full article ">Figure 6
<p>Correlation between spectral and geochemical characteristics derived from the same spot measurements of (<b>a</b>) <b>dark gray</b> limestone samples (samples B20, B28, B33, B45 and B49); (<b>b</b>) <b>light gray</b> limestone samples (samples B1, B26, B32, B36 and B41); and (<b>c</b>) dolomitic limestones samples (samples F17, F18, F20, F23 and F26). Physical models relating depths of carbonate features and wt % CaO PXRF results, color coded with the wt % MgO results (<b>left</b>) and the depths of Al-OH features and wt % Al<sub>2</sub>O<sub>3</sub> PXRF results, color coded with the wt % SiO<sub>2</sub> results (<b>right</b>).</p>
Full article ">Figure 7
<p>Geochemical charts of CaO vs. MgO (<b>left</b>) and SiO<sub>2</sub> vs. Al<sub>2</sub>O<sub>3</sub> (<b>right</b>) contents derived from PXRF spot measurements of (<b>a</b>) <b>dark gray</b> limestone samples (samples B20, B28, B33, B45 and B49); (<b>b</b>) <b>light gray</b> limestone samples (samples B1, B26, B32, B36 and B41); and (<b>c</b>) dolomitic limestones samples (samples F17, F18, F20, F23 and F26), color coded with the wavelength position of the carbonate and Al-OH absorption features.</p>
Full article ">
4744 KiB  
Article
Variability of Particle Size Distributions in the Bohai Sea and the Yellow Sea
by Zhongfeng Qiu, Deyong Sun, Chuanmin Hu, Shengqiang Wang, Lufei Zheng, Yu Huan and Tian Peng
Remote Sens. 2016, 8(11), 949; https://doi.org/10.3390/rs8110949 - 15 Nov 2016
Cited by 20 | Viewed by 5385
Abstract
Particle size distribution (PSD) is an important parameter that is relevant to many aspects of marine ecosystems, such as phytoplankton functional types, optical absorption and scattering from particulates, sediment fluxes, and carbon export. However, only a handful of studies have documented the PSD [...] Read more.
Particle size distribution (PSD) is an important parameter that is relevant to many aspects of marine ecosystems, such as phytoplankton functional types, optical absorption and scattering from particulates, sediment fluxes, and carbon export. However, only a handful of studies have documented the PSD variability in different regions. Here, we investigate the PSD properties and variability in two shallow and semi-enclosed seas (the Bohai Sea (BS) and Yellow Sea (YS)), using in situ laser diffraction measurements (LISST-100X Type C) and other measurements at 79 stations in November 2013. The results show large variability in particle concentrations (in both volume and number concentrations), with volume concentrations varying by 57-fold. The median particle diameter (Dv50) from each of the water samples also covers a large range (22.4–307.0 μm) and has an irregular statistical distribution, indicating complexity in the PSD. The PSD slopes (2.7–4.5), estimated from a power-law model, cover nearly the entire range reported previously for natural waters. Small mineral particles (with large PSD slopes) are characteristic of near-shore waters prone to sediment resuspension by winds and tides, while large biological particles (with small PSD slopes) dominate the total suspended particulates for waters away from the coast. For the BS and YS, this study provides the first report on the properties and spatial variability of the PSD, which may influence the optical properties of the ocean surface and remote sensing algorithms that are based on estimations of particle concentrations and sizes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The study area and stations in the Bohai Sea and the Yellow Sea during the winter 2013 cruise. Bathymetric contours are color coded (n = 79, measured only once).</p>
Full article ">Figure 2
<p>(<b>A</b>) The modeling performance of PSD at two typical stations (H5 and I1) using two power-law models (i.e., without/with correction of parameters), where the dotted lines and solid lines represent the correction of the model’s independent (particle size or D/D<sub>0</sub>) and dependent (N′(D)/N′(D<sub>0</sub>)) variables at stations H5 and I1, respectively, and the colors black, red and blue represent the results of observation data, by using the power-law model with constants and without constants, respectively. (<b>B</b>) Errors (MAPE, %) in the modeled PSD for all stations (n = 79) in the whole particle size range.</p>
Full article ">Figure 3
<p>The variability in the bio-optical properties: (<b>A</b>) <span class="html-italic">V</span>(D) in μL/L; (<b>B</b>) <span class="html-italic">N</span>(D) in 1 × 10<sup>10</sup> count/m<sup>3</sup>; (<b>C</b>) <span class="html-italic">D</span><sub>v</sub><sup>50</sup> in μm; (<b>D</b>) <span class="html-italic">c</span><sub>p</sub> in m<sup>−1</sup> at a wavelength of 670 nm; and (<b>E</b>) PSD slope determined from the modified power-low model) in the Bohai Sea and the Yellow Sea, measured during the cruise (n = 79) in November 2013. SD: standard deviation; CV: coefficient of variation.</p>
Full article ">Figure 4
<p>The relationships between the PSD parameters and the bio-optical properties in the Bohai Sea and the Yellow Sea, measured in November 2013 (n = 79): (<b>A</b>) <span class="html-italic">V</span>(D) vs. <span class="html-italic">c</span><sub>p</sub>; (<b>B</b>) <span class="html-italic">N</span>(D) vs. <span class="html-italic">c</span><sub>p</sub>; (<b>C</b>) <span class="html-italic">D</span><sub>v</sub><sup>50</sup> vs. <span class="html-italic">c</span><sub>p</sub>; and (<b>D</b>) <span class="html-italic">V</span>(D) vs. <span class="html-italic">N</span>(D).</p>
Full article ">Figure 5
<p>The spatial distribution of: (<b>A</b>) the PSD slopes; and (<b>B</b>) <span class="html-italic">D</span><sub>v</sub><sup>50</sup> in surface waters of the Bohai Sea and Yellow Sea in November 2013 (n = 79); and the relationship between the two parameters (<b>C</b>).</p>
Full article ">Figure 6
<p>The statistics of the PSD slopes in the: (<b>A</b>) upper layer (0–5 m) (n = 375); (<b>B</b>) middle layer (5–20 m) (n = 1065); and (<b>C</b>) deep layer (20–90 m) (n = 1354) of the water column. These sample numbers were determined by LISST sampling frequency (1 count·s<sup>−1</sup>) and residence time during the observation period in each specific layer, which is the time required for the LISST to travel down through the upper layer (0–5 m), middle layer (5–20 m), and deep layer (20–90 m).</p>
Full article ">Figure 7
<p>The vertical profiles of the mean and the standard deviation (SD) of the PSD slopes at 5 m intervals: (<b>A</b>) all stations (n = 79); (<b>B</b>) Bohai Sea (n = 20); (<b>C</b>) North Yellow Sea (n = 29); and (<b>D</b>) South Yellow Sea (n = 30). Note that the Bohai Sea is relatively shallow, and the profiles were limited to the first 40 m.</p>
Full article ">Figure 8
<p>(<b>A</b>–<b>L</b>) The vertical distributions of the size-specific particle numbers and volume concentrations for each of the size bins (<b>left two columns</b>), total particle number and volume concentrations (<b>the third column</b>), and PSD slopes and c<sub>p</sub>(670) (<b>last column</b>) from three representative stations located in the BS ((<b>top</b>) Stn. M4, Stn means Station), NYS ((<b>middle</b>) Stn. J4), and SYS ((<b>bottom</b>) Stn. A5), respectively.</p>
Full article ">Figure 9
<p>(<b>A</b>–<b>F</b>)The statistical relationships between the different PSD and the bio-optical parameters obtained at different water depths from the three representative stations (M4, J4, and A5 in the BS, NYS, and SYS, respectively).</p>
Full article ">Figure 9 Cont.
<p>(<b>A</b>–<b>F</b>)The statistical relationships between the different PSD and the bio-optical parameters obtained at different water depths from the three representative stations (M4, J4, and A5 in the BS, NYS, and SYS, respectively).</p>
Full article ">Figure 10
<p>A comparison of MAPE (%) derived from the power-law models with/without constants, by using an independent validation dataset (n = 75) collected from the Bohai Sea and the Yellow Sea in November 2014.</p>
Full article ">
4921 KiB  
Article
Environmental and Anthropogenic Degradation of Vegetation in the Sahel from 1982 to 2006
by Khaldoun Rishmawi and Stephen D. Prince
Remote Sens. 2016, 8(11), 948; https://doi.org/10.3390/rs8110948 - 13 Nov 2016
Cited by 17 | Viewed by 6018
Abstract
There is a great deal of debate on the extent, causes, and even the reality of land degradation in the Sahel. Investigations carried out before approximately 2000 using remote sensing data suggest widespread reductions in biological productivity, while studies extending beyond 2000 consistently [...] Read more.
There is a great deal of debate on the extent, causes, and even the reality of land degradation in the Sahel. Investigations carried out before approximately 2000 using remote sensing data suggest widespread reductions in biological productivity, while studies extending beyond 2000 consistently reveal a net increase in vegetation production, strongly related to the recovery of rainfall following the extreme droughts of the 1970s and 1980s, and thus challenging the notion of widespread, long-term, subcontinental-scale degradation. Yet, the spatial variations in the rates of vegetation recovery are not fully explained by rainfall trends. It is hypothesized that, in addition to rainfall, other meteorological variables and human land use have contributed to vegetation dynamics. Throughout most of the Sahel, the interannual variability in growing season ΣNDVIgs (measured from satellites, used as a proxy of vegetation productivity) was strongly related to rainfall, humidity, and temperature (mean r2 = 0.67), but with rainfall alone was weaker (mean r2 = 0.41). The mean and upper 95th quantile (UQ) rates of change in ΣNDVIgs in response to climate were used to predict potential ΣNDVIgs—that is, the ΣNDVIgs expected in response to climate variability alone, excluding any anthropogenic effects. The differences between predicted and observed ΣNDVIgs were regressed against time to detect any long-term (positive or negative) trends in vegetation productivity. Over most of the Sahel, the trends did not significantly depart from what is expected from the trends in meteorological variables. However, substantial and spatially contiguous areas (~8% of the total area of the Sahel) were characterized by negative, and, in some areas, positive trends. To explore whether the negative trends were human-induced, they were compared with the available data of population density, land use, and land biophysical properties that are known to affect the susceptibility of land to degradation. The spatial variations in the trends of the residuals were partly related to soils and tree cover, but also to several anthropogenic pressures. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Properties of the models used to estimate potential ΣNDVI<sub>gs</sub> from the relationship between observed ΣNDVI<sub>gs</sub> and climate variables. (<b>a</b>) The OLS and UQ regression lines of ΣNDVI<sub>gs</sub> on precipitation and their prediction intervals (Pred. Int) at the 95% confidence level for a cropland site (3.725°W, 11.525°N); (<b>b</b>) The difference between the OLS and UQ precipitation coefficient values for all sites throughout the Sahel study region; (<b>c</b>) The ability of precipitation (model A), precipitation and its intra-seasonal distribution (model B) and precipitation, specific humidity and temperature (model C) to account for the variations in ΣNDVI<sub>gs</sub>. SHUM—specific humidity.</p>
Full article ">Figure 2
<p>Potential ΣNDVI<sub>gs</sub> prediction errors: (<b>a</b>) errors of the OLS regression between ΣNDVI<sub>gs</sub> and precipitation (model A); Compared with model A are: (<b>b</b>) percentage reduction in potential ΣNDVI<sub>gs</sub> prediction errors of the OLS regression between ΣNDVI<sub>gs</sub> and precipitation, specific humidity, and temperature (model B); and (<b>c</b>) percentage reduction in potential ΣNDVI<sub>gs</sub> prediction errors of the OLS regression between ΣNDVI<sub>gs</sub> and precipitation, its seasonal distribution variance, and skewness (model C); (<b>d</b>) Frequency distribution of prediction errors for the three models normalized by the range of ΣNDVI<sub>gs</sub> values (prediction error/maximum ΣNDVI<sub>gs</sub>–minimum ΣNDVI<sub>gs</sub>); and (<b>e</b>) frequency distribution of the values in (<b>b</b>,<b>c</b>). The red lines from north to south are the 300 mm, 700 mm and 1100 mm rainfall isohyets.</p>
Full article ">Figure 3
<p>Temporal trends (slopes) of ΣNDVI<sub>gs</sub> residuals (observed–potential) regressed over time at four locations in the Sahel. The trends in sites (<b>a</b>,<b>b</b>) are significantly different from zero (<span class="html-italic">p</span> value of the F test &lt;0.05 and their absolute values are greater than their respective uncertainty), whereas the trends for the sites shown in (<b>c</b>,<b>d</b>) have a p values of the F test &lt;0.05, but the slope values are less than their respective uncertainty.</p>
Full article ">Figure 4
<p>Residual trends and two maps of human land use. (<b>a</b>–<b>f</b>) Trends (slopes) of NDVI<sub>gs</sub> residuals (observed–potential) over time as obtained from the six RESTREND models (A) through (F); see <a href="#remotesensing-08-00948-t002" class="html-table">Table 2</a>); (<b>g</b>) % Human Appropriation of Net Primary Production (%HANPP); and (<b>h</b>) proportion of cropland datasets used to explore the relationship between RESTREND value and land use [<a href="#B5-remotesensing-08-00948" class="html-bibr">5</a>,<a href="#B59-remotesensing-08-00948" class="html-bibr">59</a>]. The red lines from north to south are the 300 mm, 700 mm and 1100 mm rainfall isohyets.</p>
Full article ">Figure 5
<p>Mean RESTREND values (observed–potential) within groupings of: (<b>a</b>) population density (persons/ha); (<b>b</b>) percentage human appropriation of NPP (HANPP%); (<b>c</b>) livestock unit density (units/ha); (<b>d</b>) livestock unit density normalized by site productivity; (<b>e</b>) fraction land area used for crops; and (<b>f</b>) fraction land area used for crops normalized by mean annual precipitation. Filled circles are trends of the residuals where potential NDVI was obtained from OLS multivariate regression between NDVI and precipitation, specific humidity, and temperature. Open circles are trends of the residuals where potential NDVI was obtained from OLS multivariate regression between NDVI and precipitation, its seasonal variance, and skewness. Error bars are ±1 standard deviation around the mean. Dashed lines mark the zero RESTREND values.</p>
Full article ">Figure 6
<p>The relationship between significant RESTREND values and four explanatory variables, soil erodibility factor, livestock unit density normalized by site productivity (LSU/NPP), fraction land used for agriculture (cropping density), and population density. (<b>a</b>) biplot of the first and second principal component loadings of a principal component analysis; (<b>b</b>) variable importance values calculated by the Random Forest (RF) regression tree model; (<b>c</b>) comparison between RESTREND values modeled from the NDVI<sub>gs</sub> time series (x-axis) and RESTREND values predicted by RF analysis (y-axis); and (<b>d</b>) histogram of the differences between the plotted values in (<b>c</b>). Residual trends insignificantly different from zero were excluded.</p>
Full article ">Figure 7
<p>The relationship between significant RESTREND and the three explanatory variables, fraction tree cover (fTree), percentage human appropriation of NPP (%HANPP), and soil bulk density. (<b>a</b>) biplot of the first and second principal component loadings of a principal component analysis; (<b>b</b>) variable importance values calculated by the regression tree model Random Forest (RF); (<b>c</b>) comparison between RESTREND values modeled from the NDVI<sub>gs</sub> data time series (x-axis) and RESTREND values predicted by RF analysis (y-axis); and (<b>d</b>) histogram of the differences between the plotted values in (<b>c</b>). Residual trends insignificantly different from zero were excluded.</p>
Full article ">Figure 8
<p>Pruned regression tree showing the hierarchical relations of RESTREND to land use, population, and soil erodibility. Regression tree r<sup>2</sup> = 0.6 and RMSE = 0.23; LSU/NPP (livestock unit density normalized by site net primary productivity); Cropping density (fraction land area used for crops); <span class="html-italic">S.V.</span>—node split value; <span class="html-italic">S.D.</span>—standard deviation of values at the node; <span class="html-italic">V.</span>—value at terminal node; <span class="html-italic">N</span>—number of observations at terminal node.</p>
Full article ">Figure 9
<p>Pruned regression tree showing the hierarchical relations of RESTREND to %HANPP and to soil and land cover properties. Regression tree r<sup>2</sup> = 0.65 and RMSE = 0.21; <span class="html-italic">f</span> Tree—fraction tree cover; %HANPP—percentage human appropriation of NPP; <span class="html-italic">S.V.</span>—node split value; <span class="html-italic">S.D.</span>—standard deviation of values at the node; <span class="html-italic">V.</span>—value at terminal node; <span class="html-italic">N</span>—number of observations at terminal node.</p>
Full article ">
2359 KiB  
Article
Development of a New BRDF-Resistant Vegetation Index for Improving the Estimation of Leaf Area Index
by Su Zhang, Liangyun Liu, Xinjie Liu and Zefei Liu
Remote Sens. 2016, 8(11), 947; https://doi.org/10.3390/rs8110947 - 12 Nov 2016
Cited by 11 | Viewed by 5368
Abstract
The leaf area index (LAI) is one of the most important Earth surface parameters used in the modeling of ecosystems and their interaction with climate. Numerous vegetation indices have been developed to estimate the LAI. However, because of the effects of the bi-directional [...] Read more.
The leaf area index (LAI) is one of the most important Earth surface parameters used in the modeling of ecosystems and their interaction with climate. Numerous vegetation indices have been developed to estimate the LAI. However, because of the effects of the bi-directional reflectance distribution function (BRDF), most of these vegetation indices are also sensitive to the effect of BRDF. In this study, we aim to present a new BRDF-resistant vegetation index (BRVI), which is sensitive to the LAI but insensitive to the effect of BRDF. Firstly, the BRDF effects of different bands were investigated using both simulated data and in-situ measurements of winter wheat made at different growth stages. We found bi-directional shape similarity in the solar principal plane between the green and the near-infrared (NIR) bands and between the blue and red bands for farmland soil conditions and with medium chlorophyll content level. Secondly, the consistency of the shape of the BRDF across different bands was employed to develop a new BRDF-resistant vegetation index for estimating the LAI. The reflectance ratios of the NIR band to the green band and the blue band to the red band were reasonably assumed to be resistant to the BRDF effects. Nevertheless, the variation amplitude of the bi-directional reflectance in the solar principal plane was different for different bands. The divisors in the two reflectance ratios were improved by combining the reflectances at the red and green bands. The new BRVI was defined as a normalized combination of the two improved reflectance ratios. Finally, the potential of the proposed BRVI for estimation of the LAI was evaluated using both simulated data and in-situ measurements and also compared to other popular vegetation indices. The results showed that the influence of the BRDF on the BRVI was the weakest and that the BRVI retrieved LAI values well, with a coefficient of determination (R2) of 0.84 and an RMSE of 0.83 for the field data and with an R2 of 0.97 and an RMSE of 0.25 for the simulated data. It was concluded, therefore, that the new BRVI is resistant to BRDF effect and is also promising for use in estimating the LAI. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photo of instrumentation using the multi-angle observation system (MAOS).</p>
Full article ">Figure 2
<p>The simulated bi-directional reflectance in the solar principle plane at the blue, green, red and the near-infrared (NIR) bands for different leaf area index (LAI) values. A view zenith angle of 0° indicates a direction vertically downwards from the observation point. Negative values of the view zenith angle (VZA) indicate the backwards direction with a relative azimuth angle of 0°; positive values of the VZA indicate the forwards direction with a relative azimuth angle of 180°. A peak in reflectance, known as the hotspot, is assumed to occur for the negative VZA value equal to each corresponding solar zenith angle (SZA) value.</p>
Full article ">Figure 3
<p>The simulated bi-directional reflectances in the solar principle plane at the blue, green, red and NIR bands for different SZA values.</p>
Full article ">Figure 4
<p>Different vegetation indices calculated using simulated bi-directional reflectances in the solar principle plane for different SZA values.</p>
Full article ">Figure 5
<p>The values of five different vegetation indices for different SZA values calculated from canopy bi-directional reflectances of winter wheat cross the solar principle plane, acquired from 8:08 to 17:00 on 9 April 2016.</p>
Full article ">Figure 6
<p>Different vegetation indices after normalizing each vegetation indices (dividing the maximum value in the different view zenith angle) calculated by canopy bi-directional reflectances of winter wheat in the solar principle plane under different SZA values, acquired from 8:08 to 17:00 on 9 April 2016.</p>
Full article ">Figure 7
<p>Relationship between LAI and vegetation indices. LAI predictive function for VZA varying from −60° to 60° and SZA of 30.</p>
Full article ">Figure 8
<p>Relationship between vegetation indices BRVI, normalized difference vegetation index (NDVI), the soil-adjusted vegetation index (SAVI), simple ratio (SR), enhanced vegetation index (EVI) and the LAI.</p>
Full article ">Figure 9
<p>Reflectance of three types of farmland soil.</p>
Full article ">Figure 10
<p>The simulated bi-directional reflectance in the solar principle plane at the blue, green, red and NIR bands for different LAI values (with wet soil background). A view zenith angle of 0° indicates a direction vertically downwards from the observation point. Negative values of the VZA indicate the backwards direction with a relative azimuth angle of 0°; positive values of the VZA indicate the forwards direction with a relative azimuth angle of 180°. A peak in reflectance, known as the hotspot, is assumed to occur for the negative VZA value equal to each corresponding SZA value.</p>
Full article ">
23653 KiB  
Article
Mapping Urban Impervious Surface by Fusing Optical and SAR Data at the Decision Level
by Zhenfeng Shao, Huyan Fu, Peng Fu and Li Yin
Remote Sens. 2016, 8(11), 945; https://doi.org/10.3390/rs8110945 - 12 Nov 2016
Cited by 91 | Viewed by 8672
Abstract
The proliferation of impervious surfaces results in a series of environmental issues, such as the decrease of vegetated areas and the aggravation of the urban heat island effects. The mapping of impervious surface and its spatial distributions is of significance for the ecological [...] Read more.
The proliferation of impervious surfaces results in a series of environmental issues, such as the decrease of vegetated areas and the aggravation of the urban heat island effects. The mapping of impervious surface and its spatial distributions is of significance for the ecological study of urban environment. Currently, the integration of optical and synthetic aperture radar (SAR) data has shown advantages in accurately characterizing impervious surface. However, the fusion mainly occurs at the pixel and feature levels which are subject to influences of data noises and feature selections, respectively. In this paper, an innovative and effective method was developed to extract urban impervious surface by synergistically utilizing optical and SAR images at the decision level. The objective of this paper was to obtain an accurate urban impervious surface map based on the random forest classifier and the evidence theory and to provide a detailed uncertainty analysis accompanying the fused impervious surface maps. In this study, both the GaoFen (GF-1) and Sentinel-1A imagery were first used as independent data sources for mapping urban impervious surfaces. Then additional spectral features and texture features were extracted and integrated with the original GF-1 and Sentinel-1A images in generating impervious surfaces. Finally, based on the Dempster-Shafer (D-S) theory, impervious surfaces were produced by fusing the previously estimated impervious surfaces from different datasets at the decision level. Results showed that impervious surfaces estimated from the combined use of original images and features yielded a higher accuracy than those from the original optical or SAR data. Further validations suggested that optical data was better than SAR data in separating impervious surfaces from non-impervious surfaces. The fused impervious surfaces at the decision level had a higher overall accuracy than those produced independently by optical or SAR data. It was also highlighted that the fusion of GF-1 and Sentinel-1A images reduced the amount of confusions among the low reflectance of impervious surface and water, as well as for low reflectance of bare land. An overall accuracy of 95.33% was achieved for extracting urban impervious surfaces by fused datasets. The spatial distributions of uncertainties provided by the evidence theory displayed a confidence level of at least 75% for the impervious surfaces derived from the fused datasets. Full article
(This article belongs to the Special Issue Monitoring of Land Changes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographic location of the study area.</p>
Full article ">Figure 2
<p>The overall workflow for extracting urban impervious surfaces by fusing optical and synthetic aperture radar (SAR) data.</p>
Full article ">Figure 3
<p>The flowchart of the random forest (RF) algorithm.</p>
Full article ">Figure 4
<p>Feature importance.</p>
Full article ">Figure 5
<p>The frame of discernment Θ = {IS_H, IS_L, W, VE, BL_H, and BL_L}.</p>
Full article ">Figure 6
<p>The uncertainty interval measured by the belief and plausibility functions.</p>
Full article ">Figure 7
<p>The producer’s accuracy of land cover types extracted from different data sources. (<b>a</b>) The producer’s accuracy of land cover types extracted from GF-1 image; (<b>b</b>) The producer’s accuracy of land cover types extracted from Sentinel-1A image; (<b>c</b>) The producer’s accuracy of land cover types extracted from fusing the GF-1 and Sentinel-1A image.</p>
Full article ">Figure 8
<p>The user’s accuracy of land cover types extracted from different data sources. (<b>a</b>) The user’s accuracy of land cover types extracted from GF-1 image; (<b>b</b>) The user’s accuracy of land cover types extracted from Sentinel-1A image; (<b>c</b>) The user’s accuracy of land cover types extracted from fusing the GF-1 and Sentinel-1A image.</p>
Full article ">Figure 9
<p>The producer’s accuracy of land cover types extracted from different data sources. (<b>a</b>) The producer’s accuracy of land cover types extracted from GF-1 image and features; (<b>b</b>) The producer’s accuracy of land cover types extracted from Sentinel-1A image and features; (<b>c</b>) The producer’s accuracy of land cover types extracted from fusing the GF-1 and Sentinel-1A images with features.</p>
Full article ">Figure 10
<p>The user’s accuracy of land cover types extracted from different data sources. (<b>a</b>) The user’s accuracy of land cover types extracted from GF-1 image and features; (<b>b</b>) The user’s accuracy of land cover types extracted from Sentinel-1A image and features; (<b>c</b>) The user’s accuracy of land cover types extracted from fusing the GF-1 and Sentinel-1A images with features.</p>
Full article ">Figure 11
<p>Impervious surface extracted from different data sources. (<b>a</b>) Impervious surface (IS) from the GF-1 image; (<b>b</b>) IS from the Sentinel-1A image; (<b>c</b>) IS from the combined use of GF-1 image and its spectral features; (<b>d</b>) IS from the combined use the Sentinel-1A and its textural features; (<b>e</b>) IS from the fusion of the original optical and SAR images; (<b>f</b>) IS from the fusion of the original optical and SAR images and their features.</p>
Full article ">Figure 12
<p>The spatial distributions of the uncertainty levels for the fused impervious surfaces. (<b>a</b>) The uncertainty values of impervious surfaces derived by fusing the GF-1 and Sentinel-1A images; (<b>b</b>) The uncertainty values of impervious surfaces derived by fusing the GF-1 and Sentinel-1A images and their features.</p>
Full article ">
14204 KiB  
Article
Grassland and Cropland Net Ecosystem Production of the U.S. Great Plains: Regression Tree Model Development and Comparative Analysis
by Bruce Wylie, Daniel Howard, Devendra Dahal, Tagir Gilmanov, Lei Ji, Li Zhang and Kelcy Smith
Remote Sens. 2016, 8(11), 944; https://doi.org/10.3390/rs8110944 - 11 Nov 2016
Cited by 14 | Viewed by 6751
Abstract
This paper presents the methodology and results of two ecological-based net ecosystem production (NEP) regression tree models capable of up scaling measurements made at various flux tower sites throughout the U.S. Great Plains. Separate grassland and cropland NEP regression tree models were trained [...] Read more.
This paper presents the methodology and results of two ecological-based net ecosystem production (NEP) regression tree models capable of up scaling measurements made at various flux tower sites throughout the U.S. Great Plains. Separate grassland and cropland NEP regression tree models were trained using various remote sensing data and other biogeophysical data, along with 15 flux towers contributing to the grassland model and 15 flux towers for the cropland model. The models yielded weekly mean daily grassland and cropland NEP maps of the U.S. Great Plains at 250 m resolution for 2000–2008. The grassland and cropland NEP maps were spatially summarized and statistically compared. The results of this study indicate that grassland and cropland ecosystems generally performed as weak net carbon (C) sinks, absorbing more C from the atmosphere than they released from 2000 to 2008. Grasslands demonstrated higher carbon sink potential (139 g C·m−2·year−1) than non-irrigated croplands. A closer look into the weekly time series reveals the C fluctuation through time and space for each land cover type. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The map of the Great Plains with the National Land Cover Database (NLCD) 2006 as the backdrop, along with grassland and cropland flux towers used in this study as green and yellow points. Numbered flux tower labels refer to individual flux towers as designated in <a href="#app1-remotesensing-08-00944" class="html-app">Table S1</a>. Source: [<a href="#B21-remotesensing-08-00944" class="html-bibr">21</a>].</p>
Full article ">Figure 2
<p>Rangeland productivity strata (above ground biomass) classes for a normal year.</p>
Full article ">Figure 3
<p>Crop mapping model randomized test and training variations in mean (of nine replications) MAD as a function of the size of the training dataset.</p>
Full article ">Figure 4
<p>Grassland NEP mapping model randomized test and training mean (from nine replications) error terms (MAD) with varying test sizes.</p>
Full article ">Figure 5
<p>Cumulative annual NEP of the U.S. Great Plains for 2000–2008 [<a href="#B67-remotesensing-08-00944" class="html-bibr">67</a>]. Data published at [<a href="#B68-remotesensing-08-00944" class="html-bibr">68</a>].</p>
Full article ">Figure 6
<p>Cumulative seasonal NEP of the U.S. Great Plains for 2000–2008.</p>
Full article ">Figure 7
<p>Mean annual NEP map (2000–2008) with overlaid Level 3 Ecoregions. The spatial mean of cropland NEP for the U.S. Great Plains was calculated to be 30.63 g C·m<sup>−2</sup>·year<sup>−1</sup> and the spatial mean of grassland NEP was calculated to be 45.37 g C·m<sup>−2</sup>·year<sup>−1</sup>.</p>
Full article ">Figure 8
<p>Distribution of leave-one-out site cross validation RMSE and degree of extrapolation.</p>
Full article ">Figure 9
<p>NEP comparison of all crops and grasslands through time. Note that these are median values because of non-normality in some year and class combinations whereas the grass and crop NEP values reported in <a href="#remotesensing-08-00944-f007" class="html-fig">Figure 7</a> are inter-annual means.</p>
Full article ">Figure 10
<p>Major crop type NEP comparisons to grassland NEP.</p>
Full article ">Figure 11
<p>Mean weekly NEP based on land cover type in the U.S. Great Plains (2000–2008).</p>
Full article ">Figure 12
<p>Grass NEP regressed on non-irrigated crop NEP using regional class rangeland biomass median 2000–2008 mean NEP.</p>
Full article ">Figure 13
<p>Agreement of NEP between flux tower synthesis analysis [<a href="#B2-remotesensing-08-00944" class="html-bibr">2</a>,<a href="#B3-remotesensing-08-00944" class="html-bibr">3</a>,<a href="#B4-remotesensing-08-00944" class="html-bibr">4</a>] and regional dominant land cover mapped areas.</p>
Full article ">
3174 KiB  
Article
An Optimal Sample Data Usage Strategy to Minimize Overfitting and Underfitting Effects in Regression Tree Models Based on Remotely-Sensed Data
by Yingxin Gu, Bruce K. Wylie, Stephen P. Boyte, Joshua Picotte, Daniel M. Howard, Kelcy Smith and Kurtis J. Nelson
Remote Sens. 2016, 8(11), 943; https://doi.org/10.3390/rs8110943 - 11 Nov 2016
Cited by 47 | Viewed by 6704
Abstract
Regression tree models have been widely used for remote sensing-based ecosystem mapping. Improper use of the sample data (model training and testing data) may cause overfitting and underfitting effects in the model. The goal of this study is to develop an optimal sampling [...] Read more.
Regression tree models have been widely used for remote sensing-based ecosystem mapping. Improper use of the sample data (model training and testing data) may cause overfitting and underfitting effects in the model. The goal of this study is to develop an optimal sampling data usage strategy for any dataset and identify an appropriate number of rules in the regression tree model that will improve its accuracy and robustness. Landsat 8 data and Moderate-Resolution Imaging Spectroradiometer-scaled Normalized Difference Vegetation Index (NDVI) were used to develop regression tree models. A Python procedure was designed to generate random replications of model parameter options across a range of model development data sizes and rule number constraints. The mean absolute difference (MAD) between the predicted and actual NDVI (scaled NDVI, value from 0–200) and its variability across the different randomized replications were calculated to assess the accuracy and stability of the models. In our case study, a six-rule regression tree model developed from 80% of the sample data had the lowest MAD (MADtraining = 2.5 and MADtesting = 2.4), which was suggested as the optimal model. This study demonstrates how the training data and rule number selections impact model accuracy and provides important guidance for future remote-sensing-based ecosystem modeling. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Land cover types as identified by the 2011 National Land Cover Database [<a href="#B5-remotesensing-08-00943" class="html-bibr">5</a>] and the location of the study area (red box in the USA map). The legend is for the entire USA, some land cover types are not present within the study area.</p>
Full article ">Figure 2
<p>(<b>a</b>) Two-hundred fifty meter pixels with 0% of “0” values (light blue color) used to develop regression tree models and (<b>b</b>) randomly stratified samples (orange dots) overlaid on the 250 m MODIS NDVI (scaled NDVI, value from 0–200) map for the study area.</p>
Full article ">Figure 3
<p>(<b>a</b>) Mean MAD and (<b>b</b>) SD of MAD for the training and testing (validation) results from step 2. Eighty percent of the sample data used for training is highlighted by the red oval.</p>
Full article ">Figure 4
<p>Training and testing MAD based on the different number of rules for the 80% training data size. The red oval indicates the optimal number of rules and the associated training and testing MAD.</p>
Full article ">Figure 5
<p>(<b>a</b>) MAD values for all error (all AD), small error (AD ≤ 10), and large error (AD &gt; 10, the MAD value for this class was scaled to get a better view with the other classes) categories based on the different model accuracy conditions; (<b>b</b>) percent of the total population (map area) for perfect prediction (AD = 0), small error (AD ≤ 10, % of total value was scaled to get a better view with the other classes), and large error (AD &gt; 10) categories based on the different model accuracy conditions.</p>
Full article ">Figure 6
<p>Extreme AD (AD &gt; 10 at 250 m resolution) map from the four models (underfit, overfit, good fit, and all samples) for the study area. Three small representative boxes (boxes 1–3) were selected and zoomed for MADs and 30 m NLCD 2011.</p>
Full article ">Figure 7
<p>Extreme AD maps overlaid on the NLCD 2011 map for the “underfit”, “overfit”, and “good fit” models. Box 1 in the study area was zoomed for illustration.</p>
Full article ">
7939 KiB  
Article
A Novel Approach for Retrieving Tree Leaf Area from Ground-Based LiDAR
by Ting Yun, Feng An, Weizheng Li, Yuan Sun, Lin Cao and Lianfeng Xue
Remote Sens. 2016, 8(11), 942; https://doi.org/10.3390/rs8110942 - 11 Nov 2016
Cited by 62 | Viewed by 7951
Abstract
Leaf area is an important plant canopy structure parameter with important ecological significance. Light detection and ranging technology (LiDAR) with the application of a terrestrial laser scanner (TLS) is an appealing method for accurately estimating leaf area; however, the actual utility of this [...] Read more.
Leaf area is an important plant canopy structure parameter with important ecological significance. Light detection and ranging technology (LiDAR) with the application of a terrestrial laser scanner (TLS) is an appealing method for accurately estimating leaf area; however, the actual utility of this scanner depends largely on the efficacy of point cloud data (PCD) analysis. In this paper, we present a novel method for quantifying total leaf area within each tree canopy from PCD. Firstly, the shape, normal vector distribution and structure tensor of PCD features were combined with the semi-supervised support vector machine (SVM) method to separate various tree organs, i.e., branches and leaves. In addition, the moving least squares (MLS) method was adopted to remove ghost points caused by the shaking of leaves in the wind during the scanning process. Secondly, each target tree was scanned using two patterns, i.e., one scan and three scans around the canopy, to reduce the occlusion effect. Specific layer subdivision strategies according to the acquisition ranges of the scanners were designed to separate the canopy into several layers. Thirdly, 10% of the PCD was randomly chosen as an analytic dataset (ADS). For the ADS, an innovative triangulation algorithm with an assembly threshold was designed to transform these discrete scanning points into leaf surfaces and estimate the fractions of each foliage surface covered by the laser pulses. Then, a novel ratio of the point number to leaf area in each layer was defined and combined with the total number of scanned points to retrieve the total area of the leaves in the canopy. The quantified total leaf area of each tree was validated using laborious measurements with a LAI-2200 Plant Canopy Analyser and an LI-3000C Portable Area Meter. The results showed that the individual tree leaf area was accurately reproduced using our method from three registered scans, with a relative deviation of less than 10%. Nevertheless, estimations from only one scan resulted in a deviation of >25% in the retrieved individual tree leaf area due to the occlusion effect. Indeed, this study provides a novel connection between leaf area estimates and scanning sensor configuration and supplies an interesting method for estimating leaf area based on PCD. Full article
(This article belongs to the Special Issue Digital Forest Resource Monitoring and Uncertainty Analysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation to illustrate the main steps of our method.</p>
Full article ">Figure 2
<p>Experimental setup and scanning data collection: (<b>a</b>) Using terrestrial laser scanner (TLS) for various scenarios and for scanning different tree species; (<b>b</b>) Experimental set up in a triangular or rectangular plot with scanners placed at each side and lateral locations to ensure full coverage of the target trees, which included one michelia tree and two sakura trees; and (<b>c</b>) Only one scan was arranged for <span class="html-italic">Fatsia japonica</span> and rubber tree to analyse the occlusion effect existing in the LiDAR data.</p>
Full article ">Figure 3
<p>Comparison of manual measurement means to verify our method: (<b>a</b>) Measured target leaf area using the LI-3000C. (<b>b</b>) Measured deciduous leaf area using image scanner and image analysis software. (<b>c</b>) Estimated variations of the mean leaf area indexes (LAIs) of the chosen rubber trees using the LAI-2200. Different colours represent the estimated LAIs of the target rubber trees with different ages. The error bars show the standard deviation. (<b>d</b>) The lower part of one rubber tree was selected for this study and is shown in the blue box.</p>
Full article ">Figure 4
<p>Point clouds acquired using two scanning patterns: (<b>a</b>) Top view of an individual <span class="html-italic">Fatsia japonica</span> tree. The canopy was represented by one scan, which was separated into several layers according to the distance to the scanner. Each purple region contains one analytic dataset (ADS) that provides a prediction of point cloud data (PCD) density variations with changing acquisition distance. (<b>b</b>) Top view of three co-registered scans collected from different scanning locations and the canopies of the target trees, including the michelia, sakura and soapberry trees, were separated into two annulus zones (layers) according to the distance from the scanners to the canopy centre. Each purple region contains one ADS and includes both internal and external canopy data.</p>
Full article ">Figure 5
<p>Schematic illustrating the relationship between the minimum angle interval and the corresponding sampling spaces. The laser beams are emitted from a point light source with an interval angle of <math display="inline"> <semantics> <mi>τ</mi> </semantics> </math>: (<b>a</b>) the leaf surface perpendicular to the direction of the incident laser beam with sampling space <math display="inline"> <semantics> <mrow> <msub> <mi>a</mi> <mn>1</mn> </msub> </mrow> </semantics> </math>; and (<b>b</b>) the actual leaf surface with a random inclinational angle in the canopy. <math display="inline"> <semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> </mrow> </semantics> </math> is the intersection angle between the normal vector of the leaf surface and incident laser beam, which is assigned as the mean value around <math display="inline"> <semantics> <mrow> <mn>45</mn> <mo>°</mo> </mrow> </semantics> </math>. Thus, the average sample spacing of the scanned data is adjusted to <math display="inline"> <semantics> <mrow> <msub> <mi>b</mi> <mn>1</mn> </msub> </mrow> </semantics> </math>.</p>
Full article ">Figure 6
<p>Scatter plots of point cloud data using our proposed method to automatically classify into leaves (green) and branches (brown). The obtained foliage points are prepared for leaf area estimates: (<b>a</b>,<b>b</b>) the upper and lower parts of sakura tree 1; (<b>c</b>) <span class="html-italic">Fatsia japonica</span>; and (<b>d</b>) michelia tree.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>g</b>) Schematic diagrams illustrating the scanning data separated into several layers according to acquisition distance and leaf surface and reconstructed by using Delaunay triangulation. (<b>a</b>) Three scan registrations of a sakura tree separated into two layers and zones for the centre and residual edge areas. (<b>b</b>) One scan of a <span class="html-italic">Fatsia japonica</span> separated into three layers according to acquisition distance. (<b>c</b>,<b>e</b>) The partial foliage points (green) chosen from the scanning data of different trees as the ADS and red points represent the resampling leaf points through MLS filtering. (<b>d</b>,<b>f</b>,<b>g</b>) The ADSs of the sakura tree and <span class="html-italic">Fatsia japonica</span>, and triangle meshes were built based on the red points, where the green triangles with larger perimeters represent the gaps within the canopy or the partially occluded regions and the blue triangles with small perimeters represent the local leaf surface contacted by laser pulses.</p>
Full article ">Figure 8
<p>Comparisons of the estimated leaf areas between our algorithm and the manual method for five target trees: (<b>a</b>) scatter plots of the reference values obtained using LI-3000C versus the LiDAR-predicted individual tree leaf area obtained using our method; and (<b>b</b>) the obtained average area of single leaf (AASL) variations of different tree species with changes in threshold and different ADSs. The light blue areas labelled as optimal forecast threshold ranges defined by our algorithm for AASL prediction.</p>
Full article ">Figure 9
<p>Schematic diagrams illustrating the superiority of our algorithm: (<b>a</b>) Wire-frame representation of the 3-D convex hull that contains partial scanned points belonging to several broad leaves. Obviously, one-half of the total surface area of this hull is not equal to the actual foliage area. (<b>b</b>) Based on the scanned points, our method can clearly delineate the foliage surface (blue triangles) covered by the laser pulses. The summation of these triangle areas allows for precise estimates of the actual foliage area in the voxel. (<b>c</b>) Several methods derive LAI by using a line-point intersection algorithm. However, this method is arduous for computing intersections because the distance between the neighbouring laser pulses is always inconsistent with varied point spacing. (<b>d</b>) After transforming the discrete scanned points into leaf surfaces, the line–surface intersection algorithm can be used to precisely determine the number of return pulses trigged inside the voxel by the leaves.</p>
Full article ">Figure 10
<p>Schematic diagrams illustrating the method applied to the forest plot measurements. Each colour and circle represents various tree models resulting from tree crown delineation algorithms; therefore, the distance from the scanner to each tree can be calculated and combined with the minimal angle increment of various scanning patterns, rendering leaf area retrieval possible. (<b>a</b>) Top view of the experimental setup, where the scanner is alongside the forest plot; and (<b>b</b>) side view of the experimental setup, using drone loading with 3D scanner to collect forest PCD.</p>
Full article ">
8981 KiB  
Article
Water Constituents and Water Depth Retrieval from Sentinel-2A—A First Evaluation in an Oligotrophic Lake
by Katja Dörnhöfer, Anna Göritz, Peter Gege, Bringfried Pflug and Natascha Oppelt
Remote Sens. 2016, 8(11), 941; https://doi.org/10.3390/rs8110941 - 11 Nov 2016
Cited by 102 | Viewed by 11988
Abstract
Satellite remote sensing may assist in meeting the needs of lake monitoring. In this study, we aim to evaluate the potential of Sentinel-2 to assess and monitor water constituents and bottom characteristics of lakes at spatio-temporal synoptic scales. In a field campaign at [...] Read more.
Satellite remote sensing may assist in meeting the needs of lake monitoring. In this study, we aim to evaluate the potential of Sentinel-2 to assess and monitor water constituents and bottom characteristics of lakes at spatio-temporal synoptic scales. In a field campaign at Lake Starnberg, Germany, we collected validation data concurrently to a Sentinel-2A (S2-A) overpass. We compared the results of three different atmospheric corrections, i.e., Sen2Cor, ACOLITE and MIP, with in situ reflectance measurements, whereof MIP performed best (r = 0.987, RMSE = 0.002 sr−1). Using the bio-optical modelling tool WASI-2D, we retrieved absorption by coloured dissolved organic matter (aCDOM(440)), backscattering and concentration of suspended particulate matter (SPM) in optically deep water; water depths, bottom substrates and aCDOM(440) were modelled in optically shallow water. In deep water, SPM and aCDOM(440) showed reasonable spatial patterns. Comparisons with in situ data (mean: 0.43 m−1) showed an underestimation of S2-A derived aCDOM(440) (mean: 0.14 m−1); S2-A backscattering of SPM was slightly higher than backscattering from in situ data (mean: 0.027 m−1 vs. 0.019 m−1). Chlorophyll-a concentrations (~1 mg·m−3) of the lake were too low for a retrieval. In shallow water, retrieved water depths exhibited a high correlation with echo sounding data (r = 0.95, residual standard deviation = 0.12 m) up to 2.5 m (Secchi disk depth: 4.2 m), though water depths were slightly underestimated (RMSE = 0.56 m). In deeper water, Sentinel-2A bands were incapable of allowing a WASI-2D based separation of macrophytes and sediment which led to erroneous water depths. Overall, the results encourage further research on lakes with varying optical properties and trophic states with Sentinel-2A. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>S2-A true-colour composite (R-G-B: 665 nm-560 nm-490 nm, 13 August 2015, 10 m) of Lake Starnberg, Germany (<b>a</b>) and location of measurement sites (<b>b</b>).</p>
Full article ">Figure 2
<p>Simplified methodological workflow of the study.</p>
Full article ">Figure 3
<p>NEΔR<sub>rsE</sub> for each band calculated for the MIP atmospherically corrected dataset.</p>
Full article ">Figure 4
<p>S2-A resampled irradiance reflectance spectra of the two considered bottom types. The sediment spectrum is the average of reflectance measurements on 13 August 2015, the macrophyte spectrum originates from the WASI-2D database.</p>
Full article ">Figure 5
<p>Comparisons of resampled in situ and mean atmospherically corrected <math display="inline"> <semantics> <mrow> <msubsup> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> <mrow> <mi>B</mi> <mi>O</mi> <mi>A</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mrow> <msup> <mn>0</mn> <mo>+</mo> </msup> <mo>,</mo> <mi>λ</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics> </math> spectra. Error bars represent the standard deviation within a 3 × 3 ((<b>a</b>–<b>e</b>) shallow water) and 7 × 7 ((<b>f</b>–<b>g</b>) deep water) pixel environment and standard deviation of in situ spectra. Note different scaling of ordinate axis.</p>
Full article ">Figure 6
<p>Results of deep water inversion using the MIP 20 m pixel size dataset. Background is gray scaled S2-A band B05 (705 nm).</p>
Full article ">Figure 7
<p>Results of g<sub>dd</sub> (<b>a</b>–<b>c</b>) and a<span class="html-italic"><sub>CDOM</sub></span>(440) (<b>d</b>–<b>f</b>) of shallow water inversion using the MIP 10 m pixel size dataset. Background is gray scaled S2-A band B05 (705 nm).</p>
Full article ">Figure 8
<p>Results of bottom substrate unmixing using the MIP 10 m dataset (<b>a</b>); Zoomed areas are Roseninsel (<b>b</b>); Karpfenwinkel (<b>c</b>) and Seeshaupt (<b>d</b>). Low shares of sediment are illustrated as high macrophyte coverage. Background is gray scaled S2-A band B05 (705 nm).</p>
Full article ">Figure 9
<p>Results of water depths retrieval during shallow water inversion using the MIP 10 m pixel size dataset (<b>a</b>); Zoomed areas are Karpfenwinkel (<b>b</b>); validation area Seeshaupt (<b>c</b>) and Roseninsel (<b>d</b>). Background is gray scaled S2-A band B05 (705 nm).</p>
Full article ">Figure 10
<p>Scatterplots comparing echo sounding data (acquisition June 2012) and S2-A (WASI-2D) derived water depths (10 m) while unmixing two bottom types; the colour gradient highlights bottom type (<b>a</b>). Scatterplot (<b>b</b>) results from modelling water depths with fixed sediment coverage (fA[Sediment] = 1.0).</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop