[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 10, August
Previous Issue
Volume 10, June
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
remotesensing-logo

Journal Browser

Journal Browser

Remote Sens., Volume 10, Issue 7 (July 2018) – 187 articles

Cover Story (view full-size image): In the last 10 years, developments in robotics, computer vision, and sensor technology have provided new spectral remote sensing tools to capture unprecedented ultra-high spatial and high spectral resolution data with unmanned aerial vehicles (UAVs). This development has led to a revolution in geospatial data collection in which not only a small number of specialists collect and deliver remotely sensed data, but a whole diverse community is potentially able to gather geospatial data that fit their needs. However, the diversification of sensing systems challenges the common application of good practice procedures that ensure the quality of the data. This challenge can only be met by establishing and communicating common procedures. In our review, we evaluate the state-of-the-art methods in UAV spectral remote sensing that have proven successful in scientific experiments and operational demonstrations. [...] Read more.
  • 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:
21 pages, 3818 KiB  
Article
2-D Coherent Integration Processing and Detecting of Aircrafts Using GNSS-Based Passive Radar
by Hong-Cheng Zeng, Jie Chen, Peng-Bo Wang, Wei Yang and Wei Liu
Remote Sens. 2018, 10(7), 1164; https://doi.org/10.3390/rs10071164 - 23 Jul 2018
Cited by 21 | Viewed by 6281
Abstract
Long time coherent integration is a vital method for improving the detection ability of global navigation satellite system (GNSS)-based passive radar, because the GNSS signal is not radar-designed and its power level is very low. For aircraft detection, the large range cell migration [...] Read more.
Long time coherent integration is a vital method for improving the detection ability of global navigation satellite system (GNSS)-based passive radar, because the GNSS signal is not radar-designed and its power level is very low. For aircraft detection, the large range cell migration (RCM) and Doppler frequency migration (DFM) will seriously affect the coherent processing of azimuth signals, and the traditional range match filter will also be mismatched due to the Doppler-intolerant characteristic of GNSS signals. Accordingly, the energy loss of 2-dimensional (2-D) coherent processing is inevitable in traditional methods. In this paper, a novel 2-D coherent integration processing and algorithm for aircraft target detection is proposed. For azimuth processing, a modified Radon Fourier Transform (RFT) with range-walk removal and Doppler rate estimation is performed. In respect to range compression, a modified matched filter with a shifting Doppler is applied. As a result, the signal will be accurately focused in the range-Doppler domain, and a sufficiently high SNR can be obtained for aircraft detection with a moving target detector. Numerical simulations demonstrate that the range-Doppler parameters of an aircraft target can be obtained, and the position and velocity of the aircraft can be estimated accurately by multiple observation geometries due to abundant GNSS resources. The experimental results also illustrate that the blind Doppler sidelobe is suppressed effectively and the proposed algorithm has a good performance even in the presence of Doppler ambiguity. Full article
(This article belongs to the Special Issue Radar Imaging Theory, Techniques, and Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Air target detection geometric configuration with global navigation satellite system (GNSS) illuminators.</p>
Full article ">Figure 2
<p>Variations of <span class="html-italic">f<sub>d</sub></span><sub>,<span class="html-italic">v</span></sub> and <span class="html-italic">f<sub>r</sub></span><sub>,<span class="html-italic">v</span></sub> caused by air target motion with 220 m/s in different geometric configurations (the X-axis is the angle between the target velocity and the Y-axis). (<b>a</b>) Variation of <span class="html-italic">f<sub>d</sub></span><sub>,<span class="html-italic">v</span></sub> (<b>b</b>) Variation of <span class="html-italic">f<sub>r</sub></span><sub>,<span class="html-italic">v</span></sub>.</p>
Full article ">Figure 3
<p>Framework of target detection using GNSS-based passive radar.</p>
Full article ">Figure 4
<p>Processing steps of parallelized estimation of the Doppler rate.</p>
Full article ">Figure 5
<p>Processing steps of the fast implementation for azimuth coherent integration.</p>
Full article ">Figure 6
<p>Coherent integration processing results of target-1 with SVN #7 transmitter: (<b>a</b>) Range-Doppler results using the proposed algorithm; (<b>b</b>) range-Doppler results using traditional Radon Fourier Transform (RFT); (<b>c</b>) range profiles of the results; (<b>d</b>) Doppler frequency profiles of the results.</p>
Full article ">Figure 7
<p>Coherent integration processing results for all targets by the proposed algorithm with different transmitter. (<b>a</b>) Results with SVN #2; (<b>b</b>) Results with SVN #7; (<b>c</b>) Results with SVN #12.</p>
Full article ">Figure 8
<p>Available coherent time and power budget for aircraft detection using GNSS-based passive radar. (<b>a</b>) Available coherent time. (<b>b</b>) Power budget analysis.</p>
Full article ">Figure 9
<p>The impact of searching step of Doppler rate on the amplitude of the final result.</p>
Full article ">Figure 10
<p>The range and Doppler profile when the Doppler ambiguity exists. (<b>a</b>) Range profile (<b>b</b>) Doppler profile.</p>
Full article ">Figure A1
<p>Delay-Doppler ambiguity function of LFM and GPS-L5 signal. (<b>a</b>) Three-dimensional results of LFM (<b>b</b>) Doppler frequency profile of LFM (<b>c</b>) Three-dimensional results of GPS-L5 (<b>d</b>) Doppler frequency profile of GPS-L5.</p>
Full article ">
22 pages, 3882 KiB  
Article
GRACE-Based Terrestrial Water Storage in Northwest China: Changes and Causes
by Yangyang Xie, Shengzhi Huang, Saiyan Liu, Guoyong Leng, Jian Peng, Qiang Huang and Pei Li
Remote Sens. 2018, 10(7), 1163; https://doi.org/10.3390/rs10071163 - 23 Jul 2018
Cited by 38 | Viewed by 7557
Abstract
Monitoring variations in terrestrial water storage (TWS) is of great significance for the management of water resources. However, it remains a challenge to continuously monitor TWS variations using in situ observations and hydrological models because of a limited number of gauge stations and [...] Read more.
Monitoring variations in terrestrial water storage (TWS) is of great significance for the management of water resources. However, it remains a challenge to continuously monitor TWS variations using in situ observations and hydrological models because of a limited number of gauge stations and the complicated spatial distribution characteristics of TWS. In contrast, the Gravity Recovery and Climate Experiment (GRACE) could overcome the aforementioned restrictions, providing a new reliable means of observing TWS variation. Thus, GRACE was employed to investigate TWS variations in Northwest China (NWC) between April 2002 and March 2016. Unlike previous studies, we focused on the interactions of multiple climatic and vegetational factors, and their combined effects on TWS variation. In addition, we also analyzed the relationship between TWS variations and socioeconomic water consumption. The results indicated that (i) TWS had obvious seasonal variations in NWC, and showed significant decreasing trends in most parts of NWC at the 95% confidence level; (ii) decreasing sunshine duration and wind speed resulted in an increase in TWS in Qinghai province, whereas the increasing air temperature, ameliorative vegetational coverage, and excessive groundwater withdrawal jointly led to a decrease in TWS in the other provinces of NWC; (iii) TWS variations in NWC had a good correlation with water storage variations in cascade reservoirs of the upper Yellow River; and (iv) the overall interactions between multiple climatic and vegetational factors were obvious, and the strong effects of some climatic and vegetational factors could mask the weak influences of other factors in TWS variations in NWC. Hence, it is necessary to focus on the interactions of multiple factors and their combined effects on TWS variations when exploring the causes of TWS variations. Full article
(This article belongs to the Special Issue Observations, Modeling, and Impacts of Climate Extremes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of Northwest China (NWC), including meteorological stations and cascade reservoirs in the upstream of the Yellow River.</p>
Full article ">Figure 2
<p>Correlation coefficients between monthly time series of terrestrial water storage anomalies (TWSA) of NWC measured by three data centers from April 2002 to March 2016. CSR—the University of Texas Center for Space Research; GFZ—the German Research Center for Geosciences; JPL—the Jet Propulsion Laboratory.</p>
Full article ">Figure 3
<p>Multi-year averages of monthly TWS in the five provinces in NWC between April 2002 and March 2016.</p>
Full article ">Figure 4
<p>Trends of monthly TWSA time series in NWC from April 2002 to March 2016: (<b>a</b>) significance test results of the overall trends; Z is the statistic of the trend of a time series, and <span class="html-italic">Z</span>_ub and <span class="html-italic">Z</span>_lb separately denote the upper and lower bounds of non-significant trends under the significance level, <span class="html-italic">α</span> = 0.05; (<b>b</b>) linear trend rates of TWS in the five provinces of NWC.</p>
Full article ">Figure 5
<p>Cross-wavelet transformations (CWTs) between time series of the variations in TWS (ΔTWS) and precipitation (P) in Xinjiang from May 2002 to February 2016: (<b>a</b>) Original ΔTWS and P time series; (<b>b</b>) ΔTWS and P time series with periodic components and linear trend components removed.</p>
Full article ">Figure 6
<p>Significance tests of the trends of annual climatic (precipitation (P), sunshine duration (SD), air temperature (AT), and wind speed (WS)) and vegetational (normalized difference vegetation index (NDVI)) factor time series in NWC from 1982 to 2015. (<span class="html-italic">Z</span> is the statistic of the trend of a time series, and <span class="html-italic">Z</span>_ub and <span class="html-italic">Z</span>_lb separately denote the upper and lower bounds of non-significant trends under the significance level, <span class="html-italic">α</span> = 0.05.).</p>
Full article ">Figure 7
<p>Annual water withdrawals of the four provinces from the Yellow River between 2002 and 2015.</p>
Full article ">Figure 8
<p>Annual groundwater withdrawals of the five provinces in NWC.</p>
Full article ">Figure 8 Cont.
<p>Annual groundwater withdrawals of the five provinces in NWC.</p>
Full article ">
26 pages, 84876 KiB  
Article
Resolving Surface Displacements in Shenzhen of China from Time Series InSAR
by Peng Liu, Xiaofei Chen, Zhenhong Li, Zhenguo Zhang, Jiankuan Xu, Wanpeng Feng, Chisheng Wang, Zhongwen Hu, Wei Tu and Hongzhong Li
Remote Sens. 2018, 10(7), 1162; https://doi.org/10.3390/rs10071162 - 23 Jul 2018
Cited by 31 | Viewed by 7293
Abstract
Over the past few decades, the coastal city of Shenzhen has been transformed from a small fishing village to a mega city as China’s first Special Economic Zone. The rapid economic development was matched by a sharp increase in the demand for usable [...] Read more.
Over the past few decades, the coastal city of Shenzhen has been transformed from a small fishing village to a mega city as China’s first Special Economic Zone. The rapid economic development was matched by a sharp increase in the demand for usable land and coastal reclamation has been undertaken to create new land from the sea. However, it has been reported that subsidence occurred in land reclamation area and around subway tunnel area. Subsidence and the additional threat of coastal inundation from sea-level rise highlight the necessity of displacement monitoring in Shenzhen. The time Series InSAR technique is capable of detecting sub-centimeter displacement of the Earth’s surface over large areas. This study uses Envisat, COSMO-SkyMed, and Sentinel-1 datasets to determine the surface movements in Shenzhen from 2004 to 2010 and from 2013 to 2017. Subsidence observed can be attributable to both land reclamation and subway construction. Seasonal displacements are likely to be associated with precipitation. The influence of ocean tidal level changes on seasonal displacement is not strongly evident from the results and requires further investigations. In general, InSAR has proven its ability to provide accurate measurements of ground stability for the city of Shenzhen. Full article
(This article belongs to the Special Issue Radar Interferometry for Geohazards)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Shenzhen and its coastlines extracted from Landsat images in 1979, 1991, 2000, 2005, 2010 and 2015. Places shown on the map are Shafu (SF), Fuyong (FY), Gushu (GS), Bihai (BiH), Baoti (BT), Fanshen (FS), Baohua (BaH), Xin’an (XA), Linhai (LH), Qianhaiwan (QHW), Zhenhai (ZH), Mawan (MW), Qianhai (QH), Houhai (HH), and Huaqiaocheng (HQC).</p>
Full article ">Figure 2
<p>Study area and major faults. Footprints of ASAR T175, ASAR T025, CSK and S1 images are outlined in dashed lines. The NE trending faults in our study area are Jiuweiling fault (JWF), Henggang-Luohu fault (HGF), Liantang fault (LTF), and Yantian fault (YTF). The NW trending faults in our study area are Tongluojing fault (TLF), Zhengkeng fault (ZKF), Niuweiling fault (NWF), Wentang-guanlan fault (WTF), Jigongshan fault (JGF), Yangtaishan fault (YSF), and Taoyuan fault (TYF) [<a href="#B33-remotesensing-10-01162" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) ASAR T175, (<b>b</b>) ASAR T025, (<b>c</b>) CSK and (<b>d</b>) S1 archive with acquisition time and length of perpendicular baseline relative to the reference image.</p>
Full article ">Figure 4
<p>(<b>a</b>) Mean rates derived from InSAR time series of ASAR T175 between 21 November 2004 and 4 April 2010; (<b>b</b>) Mean rates from ASAR T025 between 6 September 2006 and 17 February 2010. Coastlines extracted from Landsat images in 1979, 1991, 2000, 2005, 2010 and 2015 are displayed in grey, light slate grey, slate grey, dim grey, dark slate grey, and black dashed lines respectively. Two points P1 and P2 are marked here for time series analysis. Reference points for ASAR T175 and T025 datasets are shown in black triangles.</p>
Full article ">Figure 5
<p>(<b>a</b>) Mean rates derived from CSK between 7 December 2013 and 14 January 2016; (<b>b</b>) Mean rates from S1 between 15 June 2015 and 28 February 2017. Coastlines are the same as <a href="#remotesensing-10-01162-f004" class="html-fig">Figure 4</a>. Reference points for CSK and S1A datasets are shown in black triangles.</p>
Full article ">Figure 6
<p>Vertical rates estimated from time series of ASAR T175 and T025. Scatter plot shows all pixels common to both tracks. Best-fit line is shown (solid line) along with the 1:1 line (dashed).</p>
Full article ">Figure 7
<p>(<b>a</b>) CSK and (<b>b</b>) S1 mean rates in Qianhai area. The white dashed line is the coastline in 1979. Background is a SPOT-5 image collected on 30 November 2013. Subsidence is outlined by white rectangle.</p>
Full article ">Figure 8
<p>(<b>a</b>) CSK and (<b>b</b>) S1 mean rates in Houhai area. The white dashed line is the coastline in 1979. The white solid rectangle outlines the Shenzhen Bay Sports Center (also known as Spring Cocoon).</p>
Full article ">Figure 9
<p>(<b>a</b>) CSK and (<b>b</b>) S1 mean rates to the south of Shenzhen Bao’an International Airport. The white dashed line is the coastline in 1979. Subsidence is outlined by white rectangle.</p>
Full article ">Figure 10
<p>(<b>a</b>) CSK and (<b>b</b>) S1 mean rates in Shajing area.</p>
Full article ">Figure 11
<p>Mean rates of ASAR T175 (21 November 2004–4 April 2010), ASAR T025 (6 September 2006–17 February 2010), CSK (7 December 2013–14 January 2016), and S1 (15 June 2015–28 February 2017) along the 1979, 1991, 2000, 2005, and 2015 coastlines. “CLYYYY” is short for coastline in year YYYY.</p>
Full article ">Figure 12
<p>(<b>a</b>) Mean rates along subways in Shenzhen from ASAR T175 (21 November 2004–4 April 2010) with buffer distance of 250 m to both sides of subway Line 1, Line 2, Line 5 and Line 11; (<b>b</b>–<b>d</b>) are the same for T025 (6 September 2006–17 February 2010), CSK (7 December 2013–14 January 2016) and S1 (15 June 2015–28 February 2017) respectively.</p>
Full article ">Figure 13
<p>(<b>a</b>) Mean rates along Line 1 from T175 and T025 with buffer distance of 250 m to both sides of the subway; (<b>b</b>–<b>d</b>) Are the same for Line 2, Line 5, and Line 11 respectively; (<b>e</b>) Mean rates along Line 1 from CSK and S1 with buffer distance of 250 m to both sides of the subway; (<b>f</b>–<b>h</b>) Are the same for Line 2, Line 5, and Line 11 respectively.</p>
Full article ">Figure 14
<p>(<b>a</b>) Mean rates along subways in Shenzhen from CSK with buffer distance of 250 m to both sides of Section 2 of Line 5 and Section 2 of Line 9; (<b>b</b>) Is the same for S1; (<b>c</b>) Mean rates along Section 2 of Line 5 from CSK and S1 with buffer distance of 250 m to both sides of the subway; (<b>d</b>) Is the same for Section 2 of Line 9.</p>
Full article ">Figure 15
<p>(<b>a</b>) Detrend of Envisat T175 time series at Point P1 (<a href="#remotesensing-10-01162-f005" class="html-fig">Figure 5</a>b); (<b>b</b>) Continuous wavelet power of the Envisat T175 time series; (<b>c</b>) Daily precipitation recorded by China Meteorological Administration, specifically meteorology station No. 59493 in Shenzhen [<a href="#B51-remotesensing-10-01162" class="html-bibr">51</a>]; (<b>d</b>) Continuous wavelet power of precipitation time series; (<b>e</b>) Cross wavelet transform of Envisat T175 and precipitation time series at P1; (<b>f</b>) Wavelet coherence of Envisat T175 and precipitation time series at P1; (<b>g</b>–<b>l</b>) Are the same for Envisat T025 and precipitation time series at P1; (<b>m</b>–<b>r</b>) Are the same for CSK and precipitation time series at P1; (<b>s</b>–<b>x</b>) Are the same for S1 and precipitation time series at P1. The thick contour is the 95% confidence level. The cone of influence (COI) is shown in light shadow.</p>
Full article ">Figure 16
<p>(<b>a</b>) Envisat T175 and T025 time series at Point P2 (<a href="#remotesensing-10-01162-f005" class="html-fig">Figure 5</a>b); (<b>b</b>) Monthly mean sea levels (MSL) at Tsim Bei Tsui (TBT) and Shek Pik (SP) tidal stations managed by Hong Kong Observatory [<a href="#B57-remotesensing-10-01162" class="html-bibr">57</a>]; (<b>c</b>) Daily tidal level at 00:00 and 12:00 (local time in Hong Kong) at Chek Lap Kok (CLK) tidal station managed by Hong Kong International Airport. The water levels refer to the Hong Kong Principle Datum (HKPD), which is approximately 1.23 m below mean sea level between 1965 and 1983 [<a href="#B58-remotesensing-10-01162" class="html-bibr">58</a>]; (<b>d</b>) Daily precipitation recorded by Shenzhen National Metrological Station [<a href="#B51-remotesensing-10-01162" class="html-bibr">51</a>]; (<b>e</b>–<b>h</b>) Are the same for CSK and S1 time series at P2.</p>
Full article ">Figure A1
<p>The clustering of mean rates using Getis–Ord Gi* statistic for ASAR T175. (<b>a</b>) Gi* values; (<b>b</b>) Hotspot map derived from a kernel density estimation.</p>
Full article ">Figure A2
<p>The clustering of mean rates using Getis–Ord Gi* statistic for ASAR T025. (<b>a</b>) Gi* values; (<b>b</b>) Hotspot map derived from a kernel density estimation.</p>
Full article ">Figure A3
<p>The clustering of mean rates using Getis–Ord Gi* statistic for CSK. (<b>a</b>) Gi* values; (<b>b</b>) Hotspot map derived from a kernel density estimation.</p>
Full article ">Figure A4
<p>The clustering of mean rates using Getis–Ord Gi* statistic for S1. (<b>a</b>) Gi* values; (<b>b</b>) Hotspot map derived from a kernel density estimation.</p>
Full article ">
27 pages, 20990 KiB  
Article
Evaluation of Satellite-Based Soil Moisture Products over Four Different Continental In-Situ Measurements
by Yangxiaoyue Liu, Yaping Yang and Xiafang Yue
Remote Sens. 2018, 10(7), 1161; https://doi.org/10.3390/rs10071161 - 23 Jul 2018
Cited by 27 | Viewed by 5110
Abstract
Global, near-real-time satellite-based soil moisture (SM) datasets have been developed over recent decades. However, there has been a lack of comparison among different passing times, retrieving algorithms, and sensors between SM products over various regions. In this study, we assessed seven types of [...] Read more.
Global, near-real-time satellite-based soil moisture (SM) datasets have been developed over recent decades. However, there has been a lack of comparison among different passing times, retrieving algorithms, and sensors between SM products over various regions. In this study, we assessed seven types of SM products (AMSR_A, AMSR_D, ECV_A, ECV_C, ECV_P, SMOS_A, and SMOS_D) over four different continental in-situ networks in North America, the Tibetan Plateau, Western Europe, and Southeastern Australia. Bias, R, root mean square error (RMSE), unbiased root mean square difference (ubRMSD), anomalies, and anomalies R were calculated to explore the agreement between satellite-based SM and in-situ measurements. Taylor diagrams were drawn for an inter-comparison. The results showed that (1) ECV_C was superior both in characterizing the SM temporal variation tendency and absolute value, while ECV_A produced numerous abnormal values over all validation regions. ECV_P was able to basically express the SM variation tendency, except for a few overestimations and underestimations. (2) The ascending data (AMSR_A, SMOS_A) generally outperformed the corresponding descending data (AMSR_D, SMOS_D). (3) AMSR exceeded SMOS in terms of the coefficient of correlation. (4) The validation result of SMOS_D over the NAN and OZN networks was unsatisfactory, with a rather poor correlation for both original data and anomalies. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of soil moisture (SM) networks and elevation for (<b>a</b>) Oklahoma Mesonet (OKM), (<b>b</b>) REMEDHUS (REM), (<b>c</b>) Naqu Network (NAN), and (<b>d</b>) OZNET (OZN). The grids represent the size of a Soil Moisture and Ocean Salinity soil moisture (SMOS_SM) pixel.</p>
Full article ">Figure 2
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the seven SM products over the OKM area.</p>
Full article ">Figure 3
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of anomalies over the OKM area.</p>
Full article ">Figure 4
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the seven SM products over the REM area.</p>
Full article ">Figure 5
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the anomalies over the REM area.</p>
Full article ">Figure 5 Cont.
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the anomalies over the REM area.</p>
Full article ">Figure 6
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the seven SM products over the NAN area.</p>
Full article ">Figure 7
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the anomalies over the NAN area.</p>
Full article ">Figure 7 Cont.
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the anomalies over the NAN area.</p>
Full article ">Figure 8
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the seven SM products over the OZN area.</p>
Full article ">Figure 9
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the anomalies over the OZN area.</p>
Full article ">Figure 9 Cont.
<p>Bias (<b>a</b>), R (<b>b</b>), RMSE (<b>c</b>), and ubRMSD (<b>d</b>) of the anomalies over the OZN area.</p>
Full article ">Figure 10
<p>Taylor diagrams of the OKM (<b>a</b>), REM (<b>b</b>), NAN (<b>c</b>), and OZN areas (<b>d</b>).</p>
Full article ">Figure 11
<p>Taylor diagrams of the anomalies in the OKM (<b>a</b>), REM (<b>b</b>), NAN (<b>c</b>), and OZN areas (<b>d</b>).</p>
Full article ">Figure 12
<p>Time series variation over the OKM area.</p>
Full article ">Figure 13
<p>Time series variation over the REM area.</p>
Full article ">Figure 14
<p>Time series variation over the NAN area.</p>
Full article ">Figure 15
<p>Time series variation over the OZN area.</p>
Full article ">Figure 16
<p>Time series variation in anomalies over the OKM area.</p>
Full article ">Figure 17
<p>Time series variation in anomalies over the REM area.</p>
Full article ">Figure 17 Cont.
<p>Time series variation in anomalies over the REM area.</p>
Full article ">Figure 18
<p>Time series variation in anomalies over the NAN area.</p>
Full article ">Figure 19
<p>Time series variation in anomalies over the OZN area.</p>
Full article ">Figure 19 Cont.
<p>Time series variation in anomalies over the OZN area.</p>
Full article ">
17 pages, 1637 KiB  
Article
Doppler Frequency Estimation of Point Targets in the Single-Channel SAR Image by Linear Least Squares
by Joong-Sun Won
Remote Sens. 2018, 10(7), 1160; https://doi.org/10.3390/rs10071160 - 23 Jul 2018
Cited by 6 | Viewed by 4531
Abstract
This paper presents a method and results for the estimation of residual Doppler frequency, and consequently the range velocity component of point targets in single-channel synthetic aperture radar (SAR) focused single-look complex (SLC) data. It is still a challenging task to precisely retrieve [...] Read more.
This paper presents a method and results for the estimation of residual Doppler frequency, and consequently the range velocity component of point targets in single-channel synthetic aperture radar (SAR) focused single-look complex (SLC) data. It is still a challenging task to precisely retrieve the radial velocity of small and slow-moving objects, which requires an approach providing precise estimates from only a limited number of samples within a few range bins. The proposed method utilizes linear least squares, along with the estimation of signal parameters via rotational invariance techniques (ESPRIT) algorithm, to provide optimum estimates from sets of azimuth subsamples that have different azimuth temporal distances. The ratio of estimated Doppler frequency to root-mean square error (RMSE) is suggested for determining a critical threshold, optimally selecting a number of azimuth subsample sets to be involved in the estimation. The proposed method was applied to TerraSAR-X and KOMPSAT-5 X-band SAR SLC data for on-land and coastal sea estimation, with speed-controlled, truck-mounted corner reflectors and ships, respectively. The results demonstrate its performance of the method, with percent errors of less than 5%, in retrieved range velocity for both on-land and in the sea. It is also robust, even for weak targets with low peak-to-sidelobe ratios (PSLRs) and signal-to-clutter ratios (RCSs). Since the characteristics of targets and clutter on land and in the sea are different, it is recommended that the method is applied separately with different thresholds. The limitations of the approach are also discussed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Two synthetic aperture radar (SAR) images used in this study: (<b>a</b>) TerraSAR-X SLC image, in which a speed-controlled vehicle A is observed. (<b>b</b>) KOMPSAT-5 SLC image over coastal areas, in which a speed and heading controlled ship A’ for the experiment, as well as numerous high velocity ships, are observed.</p>
Full article ">Figure 2
<p>(<b>a</b>) Residual Doppler frequency estimation of the speed-controlled vehicle denoted by “A” in <a href="#remotesensing-10-01160-f001" class="html-fig">Figure 1</a>a. The SAR-measured range velocity by the proposed method is −51.4 ± 1.20 km/h, while that recorded by GPS on the car is 49.6 km/h. The peak-to-sidelobe ratio (PSLR) of the vehicle is −11.4 dB, with a signal-to-clutter ratio (SCR) of 41.1 dB. (<b>b</b>) Residual Doppler frequency estimation of the detected target “B” in <a href="#remotesensing-10-01160-f001" class="html-fig">Figure 1</a>a. The SAR measured range velocity is approximately 5.1 ± 0.91 km/h, with an azimuth velocity of approximately 9.8 km/h. The PSLR of the target is −7.4 dB, with an SCR of 27.3 dB. While the Doppler spectrum (bottom right in each sub-figure) of targets with high PSLR and SCR well-depicts the Doppler centre frequency as in (<b>a</b>), it is difficult to determine a centre frequency from the Doppler spectrum of weak point targets, as seen in (<b>b</b>).</p>
Full article ">Figure 3
<p>An application result for a weak stationary target, to determine an optimum ratio of Doppler frequency to RMSE for estimation on land. The PSLR of the target is −4.3 dB, with an SCR of 9.6 dB. A ratio of larger than 5–10 would be acceptable for the on-land point targets.</p>
Full article ">Figure 4
<p>Results from the ship for the experiment denoted by “A’” in <a href="#remotesensing-10-01160-f001" class="html-fig">Figure 1</a>b. The ship velocity measured by on-board GPS at the time of SAR data acquisition was 16.4 km/h with an azimuth and range component of 16.1 and 2.91 km/h, respectively. The PSLR and SCR are −6.5 dB and 31.7 dB, respectively. The estimated residual Doppler frequency of 24.8 Hz corresponds to the range of −2.85 ± 0.15 km/h. From this analysis, the optimum ratio of Doppler frequency to RMSE in the sea is determined as 17.</p>
Full article ">Figure 5
<p>The SAR measured Doppler frequency of the high-speed ship B’ in <a href="#remotesensing-10-01160-f001" class="html-fig">Figure 1</a>b. The corresponding range velocity is 18.3 ± 0.72 km/h. The PSLR and SCR of the target are −7.9 dB and 23.9 dB, respectively. The estimated range velocity by this method was confirmed by measuring the azimuth distance of the target, shifted from the apex of wake in the image.</p>
Full article ">
18 pages, 6964 KiB  
Article
Interdependent Dynamics of LAI-Albedo across the Roofing Landscapes: Mongolian and Tibetan Plateaus
by Li Tian, Jiquan Chen and Changliang Shao
Remote Sens. 2018, 10(7), 1159; https://doi.org/10.3390/rs10071159 - 23 Jul 2018
Cited by 24 | Viewed by 4355
Abstract
The Mongolian Plateau (MP) and Tibetan Plateau (TP) have experienced higher-than-global average warming in recent decades, resulting in many significant changes in ecosystem structure and function. Among them are the leaf area index (LAI) and albedo, which play a fundamental role in understanding [...] Read more.
The Mongolian Plateau (MP) and Tibetan Plateau (TP) have experienced higher-than-global average warming in recent decades, resulting in many significant changes in ecosystem structure and function. Among them are the leaf area index (LAI) and albedo, which play a fundamental role in understanding many causes and consequences of land surface processes and climate. Here, we focused on the spatiotemporal changes of LAI, albedo, and their spatiotemporal relationships on the two roofing landscapes in Eurasia. Based on the MODIS products, we investigated the spatiotemporal changes of albedo(VIS, NIR and SHO) and LAI from 2000 through 2016. We found that there existed a general negative logarithmic relationship between LAI and three measures of albedo on both plateaus. No significant relationship was found for LAI-albedoNIR on the TP, due to more complex land surface canopy characteristics affected by the NIR reflection there. During 2000–2016, overall, annual mean LAI increased significantly by 119.40 × 103 km2 on the MP and by 28.35 × 103 km2 on the TP, while the decreased areas for annual mean albedoVIS were 585.59 × 103 km2 and 235.73 × 103 km2 on the MP and TP, respectively. More importantly, the LAI-albedo relationships varied substantially across the space and over time, with mismatches found in some parts of the landscapes. Substantial additional efforts with observational and/or experimental investigations are needed to explore the underlying mechanisms responsible for these relationships, including the influences of vegetation characteristics and disturbances. Full article
(This article belongs to the Special Issue Earth Radiation Budget)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The spatial distributions of land cover types. (<b>b</b>) The average annual LAI during 2000–2016. (<b>c</b>) The average annual albedo<span class="html-italic"><sub>SHO</sub></span> on the two plateaus during 2000–2016. (<b>d</b>–<b>e</b>) The probability density function (PDF) for LAI and albedo<span class="html-italic"><sub>SHO</sub></span>.</p>
Full article ">Figure 2
<p>Spatial distributions of the changing trends for LAI during 2000–2016 in the growing season (Annual, May, June, July, August, and September). The inset maps show the significant (<span class="html-italic">p</span> &lt; 0.05) increases in blue and the significant decreases in red.</p>
Full article ">Figure 3
<p>Spatial distributions of the changing trends for albedo<span class="html-italic"><sub>NIR</sub></span> during 2000–2016 in the growing season (Annual, May, June, July, August, and September). The inset maps show the significant increasing (<span class="html-italic">p</span> &lt; 0.05) (blue) or decreasing (red) trends at a significant level of <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Spatial distributions of the changing trends for albedo<span class="html-italic"><sub>VIS</sub></span> during the growing season (Annual, May, June, July, August, and September); the inset map shows the significant increasing (<span class="html-italic">p</span> &lt; 0.05) (blue) or decreasing (red) trends at a significance level of <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Spatial distributions of the changing trends in albedo<span class="html-italic"><sub>SHO</sub></span> during the growing season (Annual, May, June, July, August, and September); the inset map shows significant increases (blue) and decreases (red) (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>The trends of the normalized LAI by year and month on the grassland vegetated Mongolian and Tibetan plateau.</p>
Full article ">Figure 7
<p>The trends of the normalized albedo<sub>(<span class="html-italic">NIR, VIS, SHO</span>)</sub> by year and month for the grasslands of the Mongolian and the Tibetan plateau during 2000–2016.</p>
Full article ">Figure 8
<p>The logarithm regression of average annual albedo<sub>(<span class="html-italic">NIR,VIS,SHO</span>)</sub> retrieval as a function of estimated LAI classes (0.5 LAI steps), with stratified average data for the two plateaus.</p>
Full article ">
25 pages, 24979 KiB  
Article
Dense Connectivity Based Two-Stream Deep Feature Fusion Framework for Aerial Scene Classification
by Yunlong Yu and Fuxian Liu
Remote Sens. 2018, 10(7), 1158; https://doi.org/10.3390/rs10071158 - 23 Jul 2018
Cited by 67 | Viewed by 7073
Abstract
Aerial scene classification is an active and challenging problem in high-resolution remote sensing imagery understanding. Deep learning models, especially convolutional neural networks (CNNs), have achieved prominent performance in this field. The extraction of deep features from the layers of a CNN model is [...] Read more.
Aerial scene classification is an active and challenging problem in high-resolution remote sensing imagery understanding. Deep learning models, especially convolutional neural networks (CNNs), have achieved prominent performance in this field. The extraction of deep features from the layers of a CNN model is widely used in these CNN-based methods. Although the CNN-based approaches have obtained great success, there is still plenty of room to further increase the classification accuracy. As a matter of fact, the fusion with other features has great potential for leading to the better performance of aerial scene classification. Therefore, we propose two effective architectures based on the idea of feature-level fusion. The first architecture, i.e., texture coded two-stream deep architecture, uses the raw RGB network stream and the mapped local binary patterns (LBP) coded network stream to extract two different sets of features and fuses them using a novel deep feature fusion model. In the second architecture, i.e., saliency coded two-stream deep architecture, we employ the saliency coded network stream as the second stream and fuse it with the raw RGB network stream using the same feature fusion model. For sake of validation and comparison, our proposed architectures are evaluated via comprehensive experiments with three publicly available remote sensing scene datasets. The classification accuracies of saliency coded two-stream architecture with our feature fusion model achieve 97.79%, 98.90%, 94.09%, 95.99%, 85.02%, and 87.01% on the UC-Merced dataset (50% and 80% training samples), the Aerial Image Dataset (AID) (20% and 50% training samples), and the NWPU-RESISC45 dataset (10% and 20% training samples), respectively, overwhelming state-of-the-art methods. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A categorization of the methods for aerial scene classification.</p>
Full article ">Figure 2
<p>The texture coded two-stream deep architecture. The raw RGB network stream takes RGB images as input. The mapped LBP coded network stream takes mapped LBP images as input. Two differnet sets of features are fused using the two-stream deep feature fusion model.</p>
Full article ">Figure 3
<p>The GoogLeNet architecture used in this paper. In the inception module, the convolutions with different sizes can make the network process the input features at different scales.</p>
Full article ">Figure 4
<p>The saliency coded two-stream deep architecture. The raw RGB network stream takes RGB images as input. The saliency coded network stream takes the processed images through saliency detection as input. Two different sets of features are fused using the two-stream deep feature fusion model.</p>
Full article ">Figure 5
<p>The overall architecture of the two-stream deep feature fusion model.</p>
Full article ">Figure 6
<p>The architecture of the basic module in the two-stream deep feature fusion model.</p>
Full article ">Figure 7
<p>The detailed construction of the two-stream deep feature fusion model.</p>
Full article ">Figure 8
<p>Class representatives of the UC-Merced dataset: (1) Agricultural; (2) Airplane; (3) Baseball diamond; (4) Beach; (5) Buildings; (6) Chaparral; (7) Dense residential; (8) Forest; (9) Freeway; (10) Golf course; (11) Harbor; (12) Intersection; (13) Medium residential; (14) Mobile home park; (15) Overpass; (16) Parking lot; (17) River; (18) Runway; (19) Sparse residential; (20) Storage tanks; (21) Tennis court.</p>
Full article ">Figure 9
<p>Class representatives of the Aerial Image Dataset (AID) dataset: (1) Airport; (2) Bare land; (3) Baseball field; (4) Beach; (5) Bridge; (6) Center; (7) Church; (8) Commercial; (9) Dense residential; (10) Desert; (11) Farmland; (12) Forest; (13) Industrial; (14) Meadow; (15) Medium residential; (16) Mountain; (17) Park; (18) Parking; (19) Playground; (20) Pond; (21) Port; (22) Railway station; (23) Resort; (24) River; (25) School; (26) Sparse residential; (27) Square; (28) Stadium; (29) Storage tanks; (30) Viaduct.</p>
Full article ">Figure 10
<p>Class representatives of the NWPU-RESISC45 dataset: (1) Airplane; (2) Airport; (3) Baseball diamond; (4) Basketball court; (5) Beach; (6) Bridge; (7) Chaparral; (8) Church; (9) Circular farmland; (10) Cloud; (11) Commercial area; (12) Dense residential; (13) Desert; (14) Forest; (15) Freeway; (16) Golf course; (17) Ground track field; (18) Harbor; (19) Industrial area; (20) Intersection; (21) Island; (22) Lake; (23) Meadow; (24) Medium residential; (25) Mobile home park; (26) Mountain; (27) Overpass; (28) Palace; (29) Parking lot; (30) Railway; (31) Railway station; (32) Rectangular farmland; (33) River; (34) Roundabout; (35) Runway; (36) Sea ice; (37) Ship; (38) Snowberg; (39) Sparse residential; (40) Stadium; (41) Storage tank; (42) Tennis court; (43) Terrace; (44) Thermal power station; (45) Wetland.</p>
Full article ">Figure 11
<p>Confusion matrixes of the UC-Merced dataset under the training ratio of 80% using the following methods. (<b>a</b>) texture coded two-stream architecture with our proposed fusion model; (<b>b</b>) saliency coded two-stream architecture with our proposed fusion model.</p>
Full article ">Figure 12
<p>Confusion matrixes of the AID dataset under the training ratio of 50% using the following methods. (<b>a</b>) texture coded two-stream architecture with our proposed fusion model; (<b>b</b>) saliency coded two-stream architecture with our proposed fusion model.</p>
Full article ">Figure 13
<p>Confusion matrixes of the NWPU-RESISC45 dataset under the training ratio of 20% using the following methods. (<b>a</b>) texture coded two-stream architecture with our proposed fusion model; (<b>b</b>) saliency coded two-stream architecture with our proposed fusion model.</p>
Full article ">
18 pages, 6604 KiB  
Article
Site-Specific Unmodeled Error Mitigation for GNSS Positioning in Urban Environments Using a Real-Time Adaptive Weighting Model
by Zhetao Zhang, Bofeng Li, Yunzhong Shen, Yang Gao and Miaomiao Wang
Remote Sens. 2018, 10(7), 1157; https://doi.org/10.3390/rs10071157 - 22 Jul 2018
Cited by 41 | Viewed by 4808
Abstract
In Global Navigation Satellite System (GNSS) positioning, observation precisions are frequently impacted by the site-specific unmodeled errors, especially for the code observations that are widely used by smart phones and vehicles in urban environments. The site-specific unmodeled errors mainly refer to the multipath [...] Read more.
In Global Navigation Satellite System (GNSS) positioning, observation precisions are frequently impacted by the site-specific unmodeled errors, especially for the code observations that are widely used by smart phones and vehicles in urban environments. The site-specific unmodeled errors mainly refer to the multipath and other space loss caused by the signal propagation (e.g., non-line-of-sight reception). As usual, the observation precisions are estimated by the weighting function in a stochastic model. Only once the realistic weighting function is applied can we obtain the precise positioning results. Unfortunately, the existing weighting schemes do not fully take these site-specific unmodeled effects into account. Specifically, the traditional weighting models indirectly and partly reflect, or even simply ignore, these unmodeled effects. In this paper, we propose a real-time adaptive weighting model to mitigate the site-specific unmodeled errors of code observations. This unmodeled-error-weighted model takes full advantages of satellite elevation angle and carrier-to-noise power density ratio (C/N0). In detail, elevation is taken as a fundamental part of the proposed model, then C/N0 is applied to estimate the precision of site-specific unmodeled errors. The principle of the second part is that the measured C/N0 will deviate from the nominal values when the signal distortions are severe. Specifically, the template functions of C/N0 and its precision, which can estimate the nominal values, are applied to adaptively adjust the precision of site-specific unmodeled errors. The proposed method is tested in single-point positioning (SPP) and code real-time differenced (RTD) positioning by static and kinematic datasets. Results indicate that the adaptive model is superior to the equal-weight, elevation and C/N0 models. Compared with these traditional approaches, the accuracy of SPP and RTD solutions are improved by 35.1% and 17.6% on average in the dense high-rise building group, as well as 11.4% and 11.9% on average in the urban-forested area. This demonstrates the benefit to code-based positioning brought by a real-time adaptive weighting model as it can mitigate the impacts of site-specific unmodeled errors and improve the positioning accuracy. Full article
(This article belongs to the Special Issue Environmental Research with Global Navigation Satellite System (GNSS))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Twenty-four hour skyplot of visible GPS satellites observed from the permanent reference station.</p>
Full article ">Figure 2
<p>Observation environment of dataset no. 1 near buildings.</p>
Full article ">Figure 3
<p>Observation environment of dataset no. 2 near trees.</p>
Full article ">Figure 4
<p>(<b>a</b>) Trajectory of dataset no. 3 represented as a blue line on the main road at the University of Calgary, where the EEEL and ES buildings are marked with yellow rectangles. The green and red points denote the starting point and finishing point, respectively; and (<b>b</b>) the test environment between the EEEL and ES buildings.</p>
Full article ">Figure 5
<p>Mean C/N0 values of each satellite, where the error bars are the one-time STD ranges for each satellite. (<b>a</b>) C/N0 of C1 (C/N01); and (<b>b</b>) C/N0 of P2 (C/N02).</p>
Full article ">Figure 6
<p>Refined reference C/N0 of C1 (C/N01) and P2 (C/N02) as a function of elevation.</p>
Full article ">Figure 7
<p>(<b>a</b>) Satellite numbers; and (<b>b</b>) PDOP values. The data are calculated from the C1 observations of dataset no. 1.</p>
Full article ">Figure 8
<p>Positioning differences between the SPP solutions and the precise coordinates of dataset no. 1. (<b>a</b>) Results of the equal-weight model (EQUM); (<b>b</b>) results of the elevation model (ELEM); (<b>c</b>) results of the C/N0 model (CN0M); and (<b>d</b>) results of the adaptive model (ADAM).</p>
Full article ">Figure 9
<p>Positioning differences between the RTD solutions and the precise coordinates of dataset no. 1. (<b>a</b>) Results of the equal-weight model (EQUM); (<b>b</b>) results of the elevation model (ELEM); (<b>c</b>) results of the C/N0 model (CN0M); and (<b>d</b>) results of the adaptive model (ADAM).</p>
Full article ">Figure 10
<p>(<b>a</b>) Elevations of used satellites; (<b>b</b>) C/N0 of C1 (CN01) of used satellites; and (<b>c</b>) C/N0 of P2 (CN02) of used satellites. Each color denotes one satellite calculated from dataset no. 1.</p>
Full article ">Figure 11
<p>(<b>a</b>) Satellite numbers; and (<b>b</b>) PDOP values. The data are calculated from the C1 observations of dataset no. 2.</p>
Full article ">Figure 12
<p>Positioning differences between the SPP solutions and the precise coordinates of dataset no. 2. (<b>a</b>) Results of the equal-weight model (EQUM); (<b>b</b>) results of the elevation model (ELEM); (<b>c</b>) results of the C/N0 model (CN0M); and (<b>d</b>) results of the adaptive model (ADAM).</p>
Full article ">Figure 13
<p>Positioning differences between the RTD solutions and the precise coordinates of dataset no. 2. (<b>a</b>) Results of the equal-weight model (EQUM); (<b>b</b>) results of the elevation model (ELEM); (<b>c</b>) results of the C/N0 model (CN0M); and (<b>d</b>) results of the adaptive model (ADAM).</p>
Full article ">Figure 14
<p>(<b>a</b>) Elevations of used satellites; (<b>b</b>) C/N0 of C1 (CN01) of used satellites; and (<b>c</b>) C/N0 of P2 (CN02) of used satellites. Each color denotes one satellite calculated from dataset no. 2.</p>
Full article ">Figure 15
<p>(<b>a</b>) Satellite numbers; and (<b>b</b>) PDOP values. The data are calculated from dataset no. 3.</p>
Full article ">Figure 16
<p>SPP results (blue lines) of dataset no. 3, where the horizontal and vertical coordinates denote the X and Y directions in WGS84 coordinate system respectively, and the detailed results (red points) in meter when the test receiver is located between EEEL and ES buildings. (<b>a</b>) Results of the equal-weight model (EQUM); (<b>b</b>) results of the elevation model (ELEM); (<b>c</b>) results of the C/N0 model (CN0M); and (<b>d</b>) results of the adaptive model (ADAM).</p>
Full article ">Figure 17
<p>RTD results (blue lines) of dataset no. 3, where the horizontal and vertical coordinates denote the X and Y directions in WGS84 coordinate system respectively, and the detailed results (red points) in meter when the test receiver is located between EEEL and ES buildings. (<b>a</b>) Results of the equal-weight model (EQUM); (<b>b</b>) results of the elevation model (ELEM); (<b>c</b>) results of the C/N0 model (CN0M); and (<b>d</b>) results of the adaptive model (ADAM).</p>
Full article ">
19 pages, 1799 KiB  
Article
Spectral-Spatial Classification of Hyperspectral Images: Three Tricks and a New Learning Setting
by Jacopo Acquarelli, Elena Marchiori, Lutgarde M.C. Buydens, Thanh Tran and Twan Van Laarhoven
Remote Sens. 2018, 10(7), 1156; https://doi.org/10.3390/rs10071156 - 21 Jul 2018
Cited by 31 | Viewed by 7354
Abstract
Spectral-spatial classification of hyperspectral images has been the subject of many studies in recent years. When there are only a few labeled pixels for training and a skewed class label distribution, this task becomes very challenging because of the increased risk of overfitting [...] Read more.
Spectral-spatial classification of hyperspectral images has been the subject of many studies in recent years. When there are only a few labeled pixels for training and a skewed class label distribution, this task becomes very challenging because of the increased risk of overfitting when training a classifier. In this paper, we show that in this setting, a convolutional neural network with a single hidden layer can achieve state-of-the-art performance when three tricks are used: a spectral-locality-aware regularization term and smoothing- and label-based data augmentation. The shallow network architecture prevents overfitting in the presence of many features and few training samples. The locality-aware regularization forces neighboring wavelengths to have similar contributions to the features generated during training. The new data augmentation procedure favors the selection of pixels in smaller classes, which is beneficial for skewed class label distributions. The accuracy of the proposed method is assessed on five publicly available hyperspectral images, where it achieves state-of-the-art results. As other spectral-spatial classification methods, we use the entire image (labeled and unlabeled pixels) to infer the class of its unlabeled pixels. To investigate the positive bias induced by the use of the entire image, we propose a new learning setting where unlabeled pixels are not used for building the classifier. Results show the beneficial effect of the proposed tricks also in this setting and substantiate the advantages of using labeled and unlabeled pixels from the image for hyperspectral image classification. Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Unlabeled pixels for each of the considered hyperspectral images.</p>
Full article ">Figure 2
<p>Single hidden convolutional layer CNN architecture. The input of the CNN is the spectral feature vector of a pixel to which 1D convolutions are applied in the convolutional layer. Afterwards, the resulting feature maps are flattened and fed to the last, fully-connected, layer, which outputs the class prediction of the input pixels.</p>
Full article ">Figure 3
<p>Effect of spatial smoothing on the spectra of two neighboring pixels: original image (<b>left</b>); image after spatial-smoothing (<b>right</b>). Spectra look more similar after spatial smoothing.</p>
Full article ">Figure 4
<p>CNN-RSL (Regularization (R), Smoothing-based data augmentation (S) and Label-based data augmentation (L)) data processing flowchart. Data augmentation is applied to the original hyperspectral image. The labeled pixels from each of the three hyperspectral images (original, noisy and smoothed) form the training set <math display="inline"><semantics> <mrow> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>,</mo> <mi mathvariant="bold">y</mi> <mo>)</mo> </mrow> </semantics></math>, which is used to train the CNN with the spectral locality-aware regularization term (CNN-R).</p>
Full article ">Figure 5
<p>Convergence behavior of the CNN-RSL loss function (average over 10 folds of cross-validation on the KSC dataset).</p>
Full article ">Figure 6
<p>Prediction time in milliseconds per pixel (ms/pixel) depending on the Gaussian window size <math display="inline"><semantics> <mi>σ</mi> </semantics></math>.</p>
Full article ">
26 pages, 6504 KiB  
Article
TerraSAR-X Time Series Fill a Gap in Spaceborne Snowmelt Monitoring of Small Arctic Catchments—A Case Study on Qikiqtaruk (Herschel Island), Canada
by Samuel Stettner, Hugues Lantuit, Birgit Heim, Jayson Eppler, Achim Roth, Annett Bartsch and Bernhard Rabus
Remote Sens. 2018, 10(7), 1155; https://doi.org/10.3390/rs10071155 - 21 Jul 2018
Cited by 9 | Viewed by 5539
Abstract
The timing of snowmelt is an important turning point in the seasonal cycle of small Arctic catchments. The TerraSAR-X (TSX) satellite mission is a synthetic aperture radar system (SAR) with high potential to measure the high spatiotemporal variability of snow cover extent (SCE) [...] Read more.
The timing of snowmelt is an important turning point in the seasonal cycle of small Arctic catchments. The TerraSAR-X (TSX) satellite mission is a synthetic aperture radar system (SAR) with high potential to measure the high spatiotemporal variability of snow cover extent (SCE) and fractional snow cover (FSC) on the small catchment scale. We investigate the performance of multi-polarized and multi-pass TSX X-Band SAR data in monitoring SCE and FSC in small Arctic tundra catchments of Qikiqtaruk (Herschel Island) off the Yukon Coast in the Western Canadian Arctic. We applied a threshold based segmentation on ratio images between TSX images with wet snow and a dry snow reference, and tested the performance of two different thresholds. We quantitatively compared TSX- and Landsat 8-derived SCE maps using confusion matrices and analyzed the spatiotemporal dynamics of snowmelt from 2015 to 2017 using TSX, Landsat 8 and in situ time lapse data. Our data showed that the quality of SCE maps from TSX X-Band data is strongly influenced by polarization and to a lesser degree by incidence angle. VH polarized TSX data performed best in deriving SCE when compared to Landsat 8. TSX derived SCE maps from VH polarization detected late lying snow patches that were not detected by Landsat 8. Results of a local assessment of TSX FSC against the in situ data showed that TSX FSC accurately captured the temporal dynamics of different snow melt regimes that were related to topographic characteristics of the studied catchments. Both in situ and TSX FSC showed a longer snowmelt period in a catchment with higher contributions of steep valleys and a shorter snowmelt period in a catchment with higher contributions of upland terrain. Landsat 8 had fundamental data gaps during the snowmelt period in all 3 years due to cloud cover. The results also revealed that by choosing a positive threshold of 1 dB, detection of ice layers due to diurnal temperature variations resulted in a more accurate estimation of snow cover than a negative threshold that detects wet snow alone. We find that TSX X-Band data in VH polarization performs at a comparable quality to Landsat 8 in deriving SCE maps when a positive threshold is used. We conclude that TSX data polarization can be used to accurately monitor snowmelt events at high temporal and spatial resolution, overcoming limitations of Landsat 8, which due to cloud related data gaps generally only indicated the onset and end of snowmelt. Full article
(This article belongs to the Special Issue Ten Years of TerraSAR-X—Scientific Results)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of Qikiqtaruk in the southwestern Beaufort Sea Region and footprints of TerraSAR-X imagery.</p>
Full article ">Figure 2
<p>Data processing scheme for the optical Landsat 8 and TSX data. Layered objects represent input data, diamonds processing steps and rounded rectangles results. FSC = Fractional Snow Cover; NDSI = Normalized Difference Snow Index; SCE = Snow Cover Extent.</p>
Full article ">Figure 3
<p>Examples of Landsat 8 (<b>left</b>), TerraSAR-X in VH (<b>center</b>), and TSX in VV (<b>right</b>) polarization.</p>
Full article ">Figure 4
<p>Comparison of filter methods for SCE generation. White pixels represent snow. The red line represents the Ice Creek catchment limit on the southeastern part of Qikiqtaruk.</p>
Full article ">Figure 5
<p>Backscatter time series of TSX orbits in the Western Ice Creek catchment. The red dot indicates the footprint of the extracted backscatter.</p>
Full article ">Figure 6
<p>Comparison of Landsat 8 SCE (<b>left</b> panels) and corresponding TSX SCE derived using a threshold of −2 dB on the VH (<b>middle</b> panels) and −2.3 dB on the VV (<b>right</b> panels) polarized channels of orbit 61 from three dates in 2015 (first row) and 2016 (third and fourth row). Also shown are the results of the accuracy assessment, U = Users accuracy, P = Producers accuracy, O = Overall accuracy.</p>
Full article ">Figure 7
<p>Comparison of Landsat 8 SCE (<b>left</b> panels) and corresponding TSX SCE derived using a threshold of −2 dB on the VH (<b>middle</b> panels) and −2.3 dB on the VV (<b>right</b> panels) polarized channels of orbit 115 from four dates in 2015 (first row) and 2016 (second, fourth and fifth row). Also shown are the results of the accuracy assessment, U = Users accuracy, P = Producers accuracy, O = Overall accuracy.</p>
Full article ">Figure 8
<p>Comparison of Landsat 8 SCE (left panels) and corresponding TSX SCE derived using a threshold of 1 dB on the VH (middle panels) and VV (right panels) polarized channels of orbit 61 from three dates in 2015 (first row) and 2016 (third and fourth row). Also shown are the results of the accuracy assessment, U = Users accuracy, P = Producers accuracy, O = Overall accuracy.</p>
Full article ">Figure 9
<p>Comparison of Landsat 8 SCE (<b>left</b> panels) and corresponding TSX SCE derived using a threshold of 1 dB on the VH (<b>middle</b> panels) and VV (<b>right</b> panels) polarized channels of orbit 115 from four dates in 2015 (first row) and 2016 (second, fourth and fifth row). Also shown are the results of the accuracy assessment, U = Users accuracy, P = Producers accuracy, O = Overall accuracy.</p>
Full article ">Figure 10
<p>Time series of fractional snow cover (FSC) extent and air temperature of Qikiqtaruk from Landsat 8 and VH polarized TSX data from the orbits 115 and 61 for the years 2015, 2016, and 2017.</p>
Full article ">Figure 11
<p>Top graph: Fractional snow cover from time lapse (in situ), Landsat 8 and TerraSAR-X (TSX) imagery in 2016 in the Ice Creek catchment (white outline in the map inlet on lower right). Dates on the x-axis show month-day. Time-lapse imagery is from Camera TL2 and is located in the lower Ice Creek. It’s location (red dot) and viewing direction (red line) is indicated in the inset map on the lower right. Please note that the camera was unstable and moved between images because of ground thaw. Please note that the acquisition time of orbit 115 is in the morning, potentially affected by refreezing snow layers in early May.</p>
Full article ">Figure 12
<p>Top graph: Fractional snow cover from time lapse (in situ), Landsat 8 (L8) and TerraSAR-X (TSX) imagery in 2017 in a selected small Arctic catchment (white shape in the map inlet on lower right). Dates on the x-axis show month-day. Time-lapse imagery is from the cameras PC2 (first row) and PC3 (second row), PC5 (third row) and PC6 (fourth row), all representing flat upland tundra locations with low vegetation and tussocks. Dates of the images are the 5<sup>th</sup> of May (first column), 15/16<sup>th</sup> May (second column) and 27<sup>th</sup> of May (third column). Camera locations (red dots) and viewing directions (red lines) are shown in the inset map on the lower right.</p>
Full article ">Figure 13
<p>SCE from TerraSAR-X VH from the 29<sup>th</sup> of April 2016 (<b>left</b>) and from the 9<sup>th</sup> of May 2016 (<b>right</b>) for the Ice Creek catchment and surroundings. The red dot shows the position of the time lapse camera.</p>
Full article ">
21 pages, 3173 KiB  
Article
SfM-Based Method to Assess Gorgonian Forests (Paramuricea clavata (Cnidaria, Octocorallia))
by Marco Palma, Monica Rivas Casado, Ubaldo Pantaleo, Gaia Pavoni, Daniela Pica and Carlo Cerrano
Remote Sens. 2018, 10(7), 1154; https://doi.org/10.3390/rs10071154 - 21 Jul 2018
Cited by 27 | Viewed by 7450
Abstract
Animal forests promote marine habitats morphological complexity and functioning. The red gorgonian, Paramuricea clavata, is a key structuring species of the Mediterranean coralligenous habitat and an indicator species of climate effects on habitat functioning. P. clavata metrics such as population structure, morphology [...] Read more.
Animal forests promote marine habitats morphological complexity and functioning. The red gorgonian, Paramuricea clavata, is a key structuring species of the Mediterranean coralligenous habitat and an indicator species of climate effects on habitat functioning. P. clavata metrics such as population structure, morphology and biomass inform on the overall health of coralligenous habitats, but the estimation of these metrics is time and cost consuming, and often requires destructive sampling. As a consequence, the implementation of long-term and wide-area monitoring programmes is limited. This study proposes a novel and transferable Structure from Motion (SfM) based method for the estimation of gorgonian population structure (i.e., maximal height, density, abundance), morphometries (i.e., maximal width, fan surface) and biomass (i.e., coenenchymal Dry Weight, Ash Free Dried Weight). The method includes the estimation of a novel metric (3D canopy surface) describing the gorgonian forest as a mosaic of planes generated by fitting multiple 5 cm × 5 cm facets to a SfM generated point cloud. The performance of the method is assessed for two different cameras (GoPro Hero4 and Sony NEX7). Results showed that for highly dense populations (17 colonies/m2), the SfM-method had lower accuracies in estimating the gorgonians density for both cameras (60% to 89%) than for medium to low density populations (14 and 7 colonies/m2) (71% to 100%). Results for the validation of the method showed that the correlation between ground truth and SfM estimates for maximal height, maximal width and fan surface were between R2 = 0.63 and R2 = 0.9, and R2 = 0.99 for coenenchymal surface estimation. The methodological approach was used to estimate the biomass of the gorgonian population within the study area and across the coralligenous habitat between −25 to −40 m depth in the Portofino Marine Protected Area. For that purpose, the coenenchymal surface of sampled colonies was obtained and used for the calculations. Results showed biomass values of dry weight and ash free dry weight of 220 g and 32 g for the studied area and to 365 kg and 55 Kg for the coralligenous habitat in the Marine Protected Area. This study highlighted the feasibility of the methodology for the quantification of P. clavata metrics as well as the potential of the SfM-method to improve current predictions of the status of the coralligenous habitat in the Mediterranean sea and overall management of threatened ecosystems. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The study site in the Marine Protected Area of Portofino (Punta del Faro, Italy). (<b>a</b>) The border of the Marine Protected Area; (<b>b</b>) a detailed view of the benthic biocenosis mapped in the location of the study site using shapefiles available from [<a href="#B37-remotesensing-10-01154" class="html-bibr">37</a>]; (<b>c</b>) 3D view of the seabed around the study site at Punta del Faro (white polyline) generated via SfM using imagery collected on site.</p>
Full article ">Figure 2
<p>Schematic diagram summarizing the workflow including data collection, imagery pre-processing and alignment, point cloud processing, ground truth data analysis and the analysis of the dried colonies (i.e., biomass).</p>
Full article ">Figure 3
<p>Image footprint obtained with the Gopro Hero4 Black Edition (Woodman Labs, Inc., San Mateo, CA, USA) and the Sony NEX7 alpha (Sony Corporation, Minato, Tokyo, Japan) with both cameras being triggered at the same point over a ground control point. The image depicts the difference in frame size and in extent covered by each frame. The dimension of each of the ground control points was 22 cm × 22 cm × 22 cm. The chessboard extension was 21 cm long by 8 cm wide.</p>
Full article ">Figure 4
<p>(<b>a</b>) the point cloud generated from SfM on one of the dried colonies; (<b>b</b>) the planar fan surface interpolated over the colony; (<b>c</b>) the mosaic of planes representing the filtering gorgonian surface; (<b>d</b>) the 3D reconstruction of the colony’s coenenchymal surface with the detail of one quadrat sample of 5 cm × 5 cm.</p>
Full article ">Figure 5
<p>Three spatial representations of: (<b>a</b>) The unclassified point cloud (blue); (<b>b</b>) the point cloud segmented into matrix (blue background) and gorgonians (other colours); (<b>c</b>) the planar surfaces fitted to the point cloud and representing the colonies (matrix in blue and gorgonians in multiple colours); (<b>d</b>) the 3D canopy surface generated by fitting facets with a maximal dimension of 5 cm × 5 cm to the point cloud. The three spatial representations include (from left to right): the orthoimage, a prospective view from the Western point of the case study area and a close-up view of a gorgonian.</p>
Full article ">Figure 6
<p>Histogram of the morphometric values obtained for the SfM method across the study site of Punta del Faro (Portofino, Italy).</p>
Full article ">Figure 7
<p>Correlation between the SfM method vs. the ground truth data for each of the morphometrics calculated and cameras considered: (<b>a</b>) Gopro Hero4 and (<b>b</b>) Sony NEX7. Height refers to maximal height (m), width refers to maximal width (m) and surface refers to planar fan surface area (m<sup>2</sup>).</p>
Full article ">Figure 8
<p>Correlation between the coenenchymal surface estimated through the SfM method and the laboratory measurements over the nine dried gorgonian colonies.</p>
Full article ">
17 pages, 1135 KiB  
Technical Note
Estimation of Gap Fraction and Foliage Clumping in Forest Canopies
by Andres Kuusk, Jan Pisek, Mait Lang and Silja Märdla
Remote Sens. 2018, 10(7), 1153; https://doi.org/10.3390/rs10071153 - 21 Jul 2018
Cited by 28 | Viewed by 5978
Abstract
The gap fractions of three mature hemi-boreal forest stands in Estonia were estimated using the LAI-2000 plant canopy analyzer ( LI-COR Biosciences, Lincoln, NE, USA), the TRAC instrument (Edgewall, Miami, FL, USA), Cajanus’ tube, hemispherical photos, as well as terrestrial (TLS) and airborne [...] Read more.
The gap fractions of three mature hemi-boreal forest stands in Estonia were estimated using the LAI-2000 plant canopy analyzer ( LI-COR Biosciences, Lincoln, NE, USA), the TRAC instrument (Edgewall, Miami, FL, USA), Cajanus’ tube, hemispherical photos, as well as terrestrial (TLS) and airborne (ALS) laser scanners. ALS measurements with an 8-year interval confirmed that changes in the structure of mature forest stands are slow, and that measurements in the same season of different years should be well comparable. Gap fraction estimates varied considerably depending on the instruments and methods used. None of the methods considered for the estimation of gap fraction of forest canopies proved superior to others. The increasing spatial resolution of new ALS devices allows the canopy structure to be analyzed in more detail than was possible before. The high vertical resolution of point clouds improves the possibility of estimating the stand height, crown length, and clumping of foliage in the canopy. The clumping/regularity of the foliage in a forest canopy is correlated with tree height, crown length, and basal area. The method suggested herein for the estimation of foliage clumping allows the leaf area estimates of forest canopies to be improved. Full article
(This article belongs to the Special Issue Optical Remote Sensing of Boreal Forests)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the test site. Yellow squares are the RAMI stands, red points are the leaf area index (LAI) points in RAMI stands, empty blue polygons mark the area of airborne laser scanner (ALS) data in 2017, and filled polygons mark the area of ALS data in 2009.</p>
Full article ">Figure 2
<p>Distribution function of ALS hits in the RAMI pine stand. 1—first hits, 2—all hits, 3—weighted hits, 4—last hits.</p>
Full article ">Figure 3
<p>Distribution function of ALS hits in the RAMI pine stand in 2009 and 2017.</p>
Full article ">Figure 4
<p>Distribution function of ALS hits in the RAMI birch stand in 2009 and 2017.</p>
Full article ">Figure 5
<p>Distribution function of ALS hits in the RAMI spruce stand in 2009 and 2017.</p>
Full article ">Figure 6
<p>Gap fraction estimates: (<b>a</b>) pine stand; (<b>b</b>) birch stand; (<b>c</b>) spruce stand. TRAC—TRAC instrument, TLS—terrestrial laser scanner, HSPI—hemispherical images, ALS—airborne laser scanner.</p>
Full article ">Figure 7
<p>Dependence of the gap fraction estimate <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>0</mn> </msub> <mrow> <mo>(</mo> <mi>H</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> on the thickness of differential layers <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>z</mi> </mrow> </semantics></math>; 1—the pine stand, 2—the spruce stand, 3—the birch stand.</p>
Full article ">Figure 8
<p>Tree height and clumping factor.</p>
Full article ">Figure 9
<p>Crown length and clumping factor.</p>
Full article ">Figure 10
<p>Basal area and clumping factor.</p>
Full article ">Figure 11
<p>Allometric [<a href="#B34-remotesensing-10-01153" class="html-bibr">34</a>] and ALS estimates of LAI.</p>
Full article ">Figure 12
<p>Allometric [<a href="#B35-remotesensing-10-01153" class="html-bibr">35</a>] and ALS estimates of LAI.</p>
Full article ">
16 pages, 4889 KiB  
Article
Sentinel-1 InSAR Measurements of Elevation Changes over Yedoma Uplands on Sobo-Sise Island, Lena Delta
by Jie Chen, Frank Günther, Guido Grosse, Lin Liu and Hui Lin
Remote Sens. 2018, 10(7), 1152; https://doi.org/10.3390/rs10071152 - 21 Jul 2018
Cited by 34 | Viewed by 6679
Abstract
Yedoma—extremely ice-rich permafrost with massive ice wedges formed during the Late Pleistocene—is vulnerable to thawing and degradation under climate warming. Thawing of ice-rich Yedoma results in lowering of surface elevations. Quantitative knowledge about surface elevation changes helps us to understand the freeze-thaw processes [...] Read more.
Yedoma—extremely ice-rich permafrost with massive ice wedges formed during the Late Pleistocene—is vulnerable to thawing and degradation under climate warming. Thawing of ice-rich Yedoma results in lowering of surface elevations. Quantitative knowledge about surface elevation changes helps us to understand the freeze-thaw processes of the active layer and the potential degradation of Yedoma deposits. In this study, we use C-band Sentinel-1 InSAR measurements to map the elevation changes over ice-rich Yedoma uplands on Sobo-Sise Island, Lena Delta with frequent revisit observations (as short as six or 12 days). We observe significant seasonal thaw subsidence during summer months and heterogeneous inter-annual elevation changes from 2016–17. We also observe interesting patterns of stronger seasonal thaw subsidence on elevated flat Yedoma uplands by comparing to the surrounding Yedoma slopes. Inter-annual analyses from 2016–17 suggest that our observed positive surface elevation changes are likely caused by the delayed progression of the thaw season in 2017, associated with mean annual air temperature fluctuations. Full article
(This article belongs to the Special Issue Remote Sensing of Dynamic Permafrost Regions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Google Earth Image of Sobo-Sise Island. The red polygon outlines the island. The red crosses mark the four positions for which we present the results from InSAR time-series analysis. The three blue polygons are used to calibrate the InSAR measurements. The inset shows the location of Sobo-Sise Island in the southeastern Lena River Delta. (<b>b</b>) Elevation above the mean sea level over the Yedoma uplands and slopes. Thermokarst basins and water bodies are masked out.</p>
Full article ">Figure 2
<p>The connections between 12 Sentinel-1A/B SAR images with the perpendicular and temporal baselines.</p>
Full article ">Figure 3
<p>Surface elevation changes over the Yedoma uplands on Sobo-Sise Island from 23 June to 9 September in 2017. Minus values denote ground subsidence.</p>
Full article ">Figure 4
<p>Box plot (minimum, first quartile, median, third quartile, maximum) for seasonal subsidence at different Yedoma elevations during the thaw season of 2017.</p>
Full article ">Figure 5
<p>Averaged inter-annual elevation changes between the late thaw seasons of 2016 and 2017.</p>
Full article ">Figure 6
<p>Inter-annual elevation changes between image pairs (<b>a</b>) 20160722–20170723 (<b>b</b>) 20160803–20170804; (<b>c</b>) 20160815–20170816; (<b>d</b>) 20160827–20170828; (<b>e</b>) 20160908–20170909 and (<b>f</b>) 20160926–20170921.</p>
Full article ">Figure 7
<p>Time series of elevation changes based on our InSAR measurements at points (<b>a</b>) A; (<b>b</b>) B; (<b>c</b>) C; (<b>d</b>) D relative to the first acquisition on 22 July 2016. The error bars denote the uncertainty of the InSAR time series. The locations of the four points are shown in <a href="#remotesensing-10-01152-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 8
<p>Mean coherence over the Yedoma upland for all the 2017 interferograms. The legend symbols denote the months of the master images used to construct the interferograms.</p>
Full article ">Figure 9
<p>Daily air temperature for the thaw season of 2016 (red) and 2017 (black). The six vertical solid lines denote the dates of the inter-annual SAR acquisition pairs.</p>
Full article ">Figure A1
<p>All the possible coherence maps for the thaw season of 2017.</p>
Full article ">
23 pages, 6896 KiB  
Article
Aboveground Forest Biomass Estimation Combining L- and P-Band SAR Acquisitions
by Michael Schlund and Malcolm W. J. Davidson
Remote Sens. 2018, 10(7), 1151; https://doi.org/10.3390/rs10071151 - 20 Jul 2018
Cited by 44 | Viewed by 8641
Abstract
While considerable research has focused on using either L-band or P-band SAR (Synthetic Aperture Radar) on their own for forest biomass retrieval, the use of the two bands simultaneously to improve forest biomass retrieval remains less explored. In this paper, we make use [...] Read more.
While considerable research has focused on using either L-band or P-band SAR (Synthetic Aperture Radar) on their own for forest biomass retrieval, the use of the two bands simultaneously to improve forest biomass retrieval remains less explored. In this paper, we make use of L- and P-band airborne SAR and in situ data measured in the field together with laser scanning data acquired over one hemi-boreal (Remningstorp) and one boreal (Krycklan) forest study area in Sweden. We fit statistical models to different combinations of topographic-corrected SAR backscatter and forest heights estimated from PolInSAR for the biomass estimation, and evaluate retrieval performance in terms of R2 and using 10-fold cross-validation. The study shows that specific combinations of radar observables from L- and P-band lead to biomass predictions that are more accurate in comparison with single-band retrievals. The correlations and accuracies between the combinations of SAR features and aboveground biomass are consistent across the two study areas, whereas the retrieval performance varied for individual bands. P-band-based retrievals were more accurate than L-band for the hemi-boreal Remningstorp site and less accurate than L-band for the boreal Krycklan site. The aboveground biomass levels as well as the ground topography differ between the two sites. The results suggest that P-band is more sensitive to higher biomass and L-band to lower biomass forests. The forest height from PolInSAR improved the results at L-band in the higher biomass substantially, whereas no improvement was observed at P-band in both study areas. These results are relevant in the context of combining information over boreal forests from future low-frequency SAR missions such as the European Space Agency (ESA) BIOMASS mission, which will operate at P-band, and future L-band missions planned by several space agencies. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the Remningstorp and Krycklan study areas in Sweden.</p>
Full article ">Figure 2
<p>Log relationship between HV backscatter (<math display="inline"><semantics> <msup> <mi>γ</mi> <mn>0</mn> </msup> </semantics></math> [dB]) and aboveground biomass of forest stands for P- (<b>left</b>) and L-band (<b>right</b>) in Remningstorp (dashed line is for the uncertainty of the intercept; 2 May 2007).</p>
Full article ">Figure 3
<p>Comparison of co-polarized phase difference of P- and L-band in the forest stands in Remningstorp ((<b>a</b>) 2 May 2007) and Krycklan ((<b>b</b>) 313<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>).</p>
Full article ">Figure 4
<p>Log relationship between HV backscatter (<math display="inline"><semantics> <msup> <mi>γ</mi> <mn>0</mn> </msup> </semantics></math> [dB]) and aboveground biomass of forest stands for P- (<b>left</b>) and L-band (<b>right</b>) in Krycklan (dashed line is for the uncertainty of the intercept; 313<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math> heading).</p>
Full article ">Figure 5
<p>PolInSAR height and aboveground biomass regression based on P- (<b>left</b>) and L-band (<b>right</b>) in Remningstorp (Data acquired on 2 May 2007).</p>
Full article ">Figure 6
<p>Spatial representation of L-band backscatter coefficient <math display="inline"><semantics> <msup> <mi>γ</mi> <mn>0</mn> </msup> </semantics></math> with location of forest stands, aboveground biomass estimated with backscatter and PolInSAR height at P- and L-band (center) and biomass model combining the two bands (right) in Remningstorp (<b>a</b>) and Krycklan (<b>b</b>). Please note that the aboveground biomass of the stands ranged to 183 t ha<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> in Krycklan and 287 t ha<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> in Remningstorp and thus, estimations above these values may not be reliable.</p>
Full article ">Figure 7
<p>Estimated aboveground biomass using biomass models <math display="inline"><semantics> <msub> <mi>B</mi> <mn>1</mn> </msub> </semantics></math> with P-band (<b>left</b>), <math display="inline"><semantics> <msub> <mi>B</mi> <mn>1</mn> </msub> </semantics></math> with L-band (<b>center</b>) and <math display="inline"><semantics> <msub> <mi>B</mi> <mn>10</mn> </msub> </semantics></math> (<b>right</b>) compared to actual aboveground biomass in Remningstorp (<b>a</b>) and Krycklan (<b>b</b>; line corresponds to a 1:1 relationship).</p>
Full article ">Figure 8
<p>PolInSAR height and aboveground biomass regression based on P- (<b>left</b>) and L-band (<b>right</b>) in Krycklan.</p>
Full article ">
14 pages, 2758 KiB  
Technical Note
How Well Can Global Precipitation Measurement (GPM) Capture Hurricanes? Case Study: Hurricane Harvey
by Ehsan Omranian, Hatim O. Sharif and Ahmad A. Tavakoly
Remote Sens. 2018, 10(7), 1150; https://doi.org/10.3390/rs10071150 - 20 Jul 2018
Cited by 64 | Viewed by 8713
Abstract
Hurricanes and other severe coastal storms have become more frequent and destructive during recent years. Hurricane Harvey, one of the most extreme events in recent history, advanced as a category IV storm and brought devastating rainfall to the Houston, TX, region during 25–29 [...] Read more.
Hurricanes and other severe coastal storms have become more frequent and destructive during recent years. Hurricane Harvey, one of the most extreme events in recent history, advanced as a category IV storm and brought devastating rainfall to the Houston, TX, region during 25–29 August 2017. It inflicted damage of more than $125 billion to the state of Texas infrastructure and caused multiple fatalities in a very short period of time. Rainfall totals from Harvey during the 5-day period were among the highest ever recorded in the United States. Study of this historical devastating event can lead to better preparation and effective reduction of far-reaching consequences of similar events. Precipitation products based on satellites observations can provide valuable information needed to understand the evolution of such devastating storms. In this study, the ability of recent Integrated Multi-satellitE Retrievals for Global Precipitation Mission (GPM-IMERG) final-run product to capture the magnitudes and spatial (0.1° × 0.1°)/temporal (hourly) patterns of rainfall resulting from hurricane Harvey was evaluated. Hourly gridded rainfall estimates by ground radar (4 × 4 km) were used as a reference dataset. Basic and probabilistic statistical indices of the satellite rainfall products were examined. The results indicated that the performance of IMERG product was satisfactory in detecting the spatial variability of the storm. It reconstructed precipitation with nearly 62% accuracy, although it systematically under-represented rainfall in coastal areas and over-represented rainfall over the high-intensity regions. Moreover, while the correlation between IMERG and radar products was generally high, it decreased significantly at and around the storm core. Full article
(This article belongs to the Special Issue Remote Sensing of Precipitation)
Show Figures

Figure 1

Figure 1
<p>Cumulative precipitation measurement (mm) and spatial pattern captured by (<b>a</b>) NCEP stage-IV Radar (<b>b</b>) GPM-IMERG satellite final-run product, during Hurricane Harvey from 25 to 29, August 2018.</p>
Full article ">Figure 2
<p>Spatial distributions of the Correlation Coefficient (CC) factor for hourly IMERG precipitation product at 0.1° × 0.1° resolution over Texas from 25 to 29 August 2017.</p>
Full article ">Figure 3
<p>Scatterplots of IMERG satellite product and stage-IV NCEP radar precipitation product with an hourly temporal resolution between 25th and 29th August 2017.</p>
Full article ">Figure 4
<p>Box plots of probabilistic (POD, FAR, CSI and PSS) and basic (ME, MBF, BIAS and RMSE) statistical indices over grid boxes for IMERG hourly precipitation products.</p>
Full article ">Figure 5
<p>Spatial distribution of basic statistical indices for the final-run IMERG hourly product during Hurricane Harvey (<b>a</b>) ME (<b>b</b>) BIAS (<b>c</b>) MBF (<b>d</b>) RSME.</p>
Full article ">Figure 6
<p>Spatial distribution of probabilistic statistical indices for the final-run IMERG during hurricane Harvey (<b>a</b>) POD (<b>b</b>) CSI (<b>c</b>) FAR (<b>d</b>) PSS.</p>
Full article ">
20 pages, 7564 KiB  
Article
Intercomparison of Three Two-Source Energy Balance Models for Partitioning Evaporation and Transpiration in Semiarid Climates
by Yongmin Yang, Jianxiu Qiu, Renhua Zhang, Shifeng Huang, Sheng Chen, Hui Wang, Jiashun Luo and Yue Fan
Remote Sens. 2018, 10(7), 1149; https://doi.org/10.3390/rs10071149 - 20 Jul 2018
Cited by 28 | Viewed by 5635
Abstract
Evaporation (E) and transpiration (T) information is crucial for precise water resources planning and management in arid and semiarid areas. Two-source energy balance (TSEB) methods based on remotely-sensed land surface temperature provide an important modeling approach for estimating evapotranspiration (ET) and its components [...] Read more.
Evaporation (E) and transpiration (T) information is crucial for precise water resources planning and management in arid and semiarid areas. Two-source energy balance (TSEB) methods based on remotely-sensed land surface temperature provide an important modeling approach for estimating evapotranspiration (ET) and its components of E and T. Approaches for accurate decomposition of the component temperature and E/T partitioning from ET based on TSEB requires careful investigation. In this study, three TSEB models are used: (i) the TSEB model with the Priestley-Taylor equation, i.e., TSEB-PT; (ii) the TSEB model using the Penman-Monteith equation, i.e., TSEB-PM, and (iii) the TSEB using component temperatures derived from vegetation fractional cover and land surface temperature (VFC/LST) space, i.e., TSEB-TC-TS. These models are employed to investigate the impact of component temperature decomposition on E/T partitioning accuracy. Validation was conducted in the large-scale campaign of Heihe Watershed Allied Telemetry Experimental Research-Multi-Scale Observation Experiment on Evapotranspiration (HiWATER-MUSOEXE) in the northwest of China, and results showed that root mean square errors (RMSEs) of latent and sensible heat fluxes were respectively lower than 76 W/m2 and 50 W/m2 for all three approaches. Based on the measurements from the stable oxygen and hydrogen isotopes system at the Daman superstation, it was found that all three models slightly overestimated the ratio of E/ET. In addition, discrepancies in E/T partitioning among the three models were observed in the kernel experimental area of MUSOEXE. Further intercomparison indicated that different temperature decomposition methods were responsible for the observed discrepancies in E/T partitioning. The iterative procedure adopted by TSEB-PT and TSEB-PM produced higher LEC and lower TC when compared to TSEB-TC-TS. Overall, this work provides valuable insights into understanding the performances of TSEB models with different temperature decomposition mechanisms over semiarid regions. Full article
(This article belongs to the Special Issue Remote Sensing of Evapotranspiration (ET))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The distribution of flux towers and the land use classifications in MUSOEXE over the Zhangye oasis. The yellow rectangular in the left shows the kernel experimental area in MUSOEXE, and the subset figure in the lower right shows the location of MUSOEXE (marked in red triangle) in the Heihe River Basin (marked by pink polygon) and in China.</p>
Full article ">Figure 2
<p>Validation of energy balance components of TSEB-PT, TSEB-PM, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub> during the HiWATER experiment at times of ASTER overpass. Energy balance components are (<b>a</b>) <span class="html-italic">R</span><sub>n</sub>, (<b>b</b>) <span class="html-italic">G</span>, (<b>c</b>) <span class="html-italic">LE</span> and (<b>d</b>) <span class="html-italic">H</span>.</p>
Full article ">Figure 3
<p>The spatial distributions of <span class="html-italic">LE</span> (first row) and <span class="html-italic">H</span> (second row) over the Zhangye oasis based on TSEB-PT, TSEB-PM, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub> for the satellite overpass time on 10 July 2012.</p>
Full article ">Figure 4
<p>The spatial distribution of <span class="html-italic">LE</span>c (first row) and <span class="html-italic">LE</span>s (second row) over the Zhangye oasis based on TSEB-PT, TSEB-PM, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub> for the satellite overpass time on 10 July 2012.</p>
Full article ">Figure 5
<p>The spatial distribution of <span class="html-italic">CWSI</span>c (first row) and <span class="html-italic">SWDI</span>s (second row) over the Zhangye oasis based on TSEB-PM, TSEB-PT, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub> for the satellite overpass time on 10 July 2012. The white areas correspond to the sandy and Gobi desert, and these pixels are masked.</p>
Full article ">Figure 6
<p>Comparison of <span class="html-italic">LE</span><sub>C</sub>/<span class="html-italic">LE</span> (%) between the three TSEB models and ground measurements by stable oxygen and hydrogen isotopes technique at Daman superstation.</p>
Full article ">Figure 7
<p>The intercomparison of <span class="html-italic">LE</span><sub>C</sub> (first row) and <span class="html-italic">LE</span><sub>S</sub> (second row) derived from TSEB-PT, TSEB-PM, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub> in the kernel experimental area on 10 July 2012. The intercompared <span class="html-italic">LE</span><sub>C</sub> pairs are (<b>a</b>) TSEB-PM vs. TSEB-PT; (<b>b</b>) TSEB-PM vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>; (<b>c</b>) TSEB-PT vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>. The intercompared <span class="html-italic">LE</span><span class="html-italic">s</span> pairs are (<b>d</b>) TSEB-PM vs. TSEB-PT; (<b>e</b>)TSEB-PM vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>; (<b>f</b>) TSEB-PT vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>.</p>
Full article ">Figure 8
<p>The spatial distribution of <span class="html-italic">T</span><sub>C</sub> (first row) and <span class="html-italic">T</span>s (second row) over HiWATER-MUSOEXE derived from TSEB-PT, TSEB-PM, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub> on 10 July 2012.</p>
Full article ">Figure 9
<p>Intercomparison of <span class="html-italic">T</span>c (first row) and <span class="html-italic">T</span>s (second row) derived from TSEB-PT, TSEB-PM, and TSEB-<span class="html-italic">T</span>c-<span class="html-italic">T</span>s in the kernel experimental area on 10 July 2012. The intercompared <span class="html-italic">T</span>c pairs are (<b>a</b>) TSEB-PM vs. TSEB-PT; (<b>b</b>) TSEB-PM vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>; (<b>c</b>) TSEB-PT vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>. The intercompared <span class="html-italic">T</span>s pairs are (<b>d</b>) TSEB-PM vs. TSEB-PT; (<b>e</b>)TSEB-PM vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>; (<b>f</b>) TSEB-PT vs. TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>.</p>
Full article ">Figure 10
<p>Illustration of the temperature decomposition methods adopted in the three TSEB models. <span class="html-italic">T</span><sub>s1</sub>, <span class="html-italic">T</span><sub>s2</sub>, and <span class="html-italic">T</span><sub>s3</sub> denote the soil surface temperatures derived from TSEB-PM, TSEB-PT, and TSEB-<span class="html-italic">T</span><sub>C</sub>-<span class="html-italic">T</span><sub>S</sub>, respectively, and <span class="html-italic">T</span><sub>v1</sub>, <span class="html-italic">T</span><sub>v2</sub>, and <span class="html-italic">T</span><sub>v3</sub> denote the vegetation canopy temperatures derived from the same three models.</p>
Full article ">
26 pages, 8924 KiB  
Article
Large-Scale Accurate Reconstruction of Buildings Employing Point Clouds Generated from UAV Imagery
by Shirin Malihi, Mohammad Javad Valadan Zoej and Michael Hahn
Remote Sens. 2018, 10(7), 1148; https://doi.org/10.3390/rs10071148 - 20 Jul 2018
Cited by 29 | Viewed by 6257
Abstract
High-density point clouds are valuable and detailed sources of data for different processes related to photogrammetry. We explore the knowledge-based generation of accurate large-scale three-dimensional (3D) models of buildings employing point clouds derived from UAV-based photogrammetry. A new two-level segmentation approach based on [...] Read more.
High-density point clouds are valuable and detailed sources of data for different processes related to photogrammetry. We explore the knowledge-based generation of accurate large-scale three-dimensional (3D) models of buildings employing point clouds derived from UAV-based photogrammetry. A new two-level segmentation approach based on efficient RANdom SAmple Consensus (RANSAC) shape detection is developed to segment potential facades and roofs of the buildings and extract their footprints. In the first level, the cylinder primitive is implemented to trim point clouds and split buildings, and the second level of the segmentation produces planar segments. The efficient RANSAC algorithm is enhanced in sizing up the segments via point-based analyses for both levels of segmentation. Then, planar modelling is carried out employing contextual knowledge through a new constrained least squares method. New evaluation criteria are proposed based on conceptual knowledge. They can examine the abilities of the approach in reconstruction of footprints, 3D models, and planar segments in addition to detection of over/under segmentation. Evaluation of the 3D models proves that the geometrical accuracy of LoD3 is achieved, since the average horizontal and vertical accuracy of the reconstructed vertices of roofs and footprints are better than (0.24, 0.23) m, (0.19, 0.17) m for the first dataset, and (0.35, 0.37) m, (0.28, 0.24) m for the second dataset. Full article
Show Figures

Figure 1

Figure 1
<p>The general diagram of the proposed 3D reconstruction method.</p>
Full article ">Figure 2
<p>Two building blocks and their surrounding areas.</p>
Full article ">Figure 3
<p>The first level of segmentation: (<b>a</b>,<b>c</b>) using cylinder primitive in dividing point clouds of the buildings, to (<b>d</b>,<b>f</b>) several parts; (<b>b</b>) using cylinder primitive to (<b>e</b>) segment point cloud of the footprint. (Units of axes are m).</p>
Full article ">Figure 4
<p>Approximation of µ. (<b>a</b>) The eigenvectors of an ellipse and a circle-shaped neighborhood around the centre point. (<b>b</b>) The histogram of <math display="inline"> <semantics> <mrow> <mn>1</mn> <mo>/</mo> <mrow> <mo>(</mo> <mrow> <msub> <mi>λ</mi> <mrow> <mn>3</mn> <mtext> </mtext> <mi>e</mi> </mrow> </msub> <mtext> </mtext> <mo>.</mo> <mtext> </mtext> <msub> <mi>λ</mi> <mrow> <mn>2</mn> <mtext> </mtext> <mi>e</mi> </mrow> </msub> </mrow> <mo>)</mo> </mrow> </mrow> </semantics> </math>.</p>
Full article ">Figure 5
<p>Approximating number of points of the planar segments for the efficient RANSAC. (<b>a</b>) The histogram of cosine of the inclination angle of <math display="inline"> <semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>λ</mi> <mn>1</mn> </msub> </mrow> <mo stretchy="true">→</mo> </mover> </mrow> </semantics> </math>. (<b>b</b>–<b>e</b>) Histograms of the inclination angle (in radian) of the footprint, roof, step roof and walls. (<b>b’</b>,<b>c’</b>,<b>d’</b>,<b>e’</b>) corresponding marked points in clusters (units of axes are m).</p>
Full article ">Figure 5 Cont.
<p>Approximating number of points of the planar segments for the efficient RANSAC. (<b>a</b>) The histogram of cosine of the inclination angle of <math display="inline"> <semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>λ</mi> <mn>1</mn> </msub> </mrow> <mo stretchy="true">→</mo> </mover> </mrow> </semantics> </math>. (<b>b</b>–<b>e</b>) Histograms of the inclination angle (in radian) of the footprint, roof, step roof and walls. (<b>b’</b>,<b>c’</b>,<b>d’</b>,<b>e’</b>) corresponding marked points in clusters (units of axes are m).</p>
Full article ">Figure 6
<p>Planarity. (<b>a</b>,<b>c</b>) for the point clouds of the roof (5–8, 7–19), the walls (1–4, 1–7), and (<b>b</b>,<b>d</b>) for the point clouds of the grounds. (<b>a</b>,<b>b</b>) point clouds of building 1, and (<b>c</b>,<b>d</b>) point clouds of building 2.</p>
Full article ">Figure 7
<p>Superimposing the constructed models on point clouds of (<b>a</b>) building 1 and (<b>b</b>) building 2. (Units of axes are m).</p>
Full article ">Figure 8
<p>The reconstructed B-rep models. (<b>a</b>,<b>d</b>) superimposing the reconstructed buildings on the point clouds, (<b>b</b>,<b>e</b>) the reconstructed roofs, (<b>c</b>,<b>f</b>) the reconstructed footprints (green) besides the reference data (blue). (Units of axes are m).</p>
Full article ">Figure 8 Cont.
<p>The reconstructed B-rep models. (<b>a</b>,<b>d</b>) superimposing the reconstructed buildings on the point clouds, (<b>b</b>,<b>e</b>) the reconstructed roofs, (<b>c</b>,<b>f</b>) the reconstructed footprints (green) besides the reference data (blue). (Units of axes are m).</p>
Full article ">Figure 9
<p>The corrected footprint, (<b>a</b>) of building 1 and (<b>b</b>) of building 2. (Units of axes are m).</p>
Full article ">Figure 9 Cont.
<p>The corrected footprint, (<b>a</b>) of building 1 and (<b>b</b>) of building 2. (Units of axes are m).</p>
Full article ">Figure 10
<p>Number of iterations for “minimum points per primitive” computing for planes.</p>
Full article ">Figure A1
<p>Approximation of the step roof’s number of points. (<b>a</b>,<b>b</b>) Recognizing the edges of the roof as the borders of half-planes via histogram of <math display="inline"> <semantics> <mrow> <mfrac bevelled="true"> <mrow> <msub> <mi>λ</mi> <mn>2</mn> </msub> </mrow> <mrow> <msub> <mi>λ</mi> <mn>3</mn> </msub> </mrow> </mfrac> </mrow> </semantics> </math>, and (<b>c</b>) generation of the main edges of the roof. (Units of axes in (<b>b</b>,<b>c</b>) are m).</p>
Full article ">Figure A1 Cont.
<p>Approximation of the step roof’s number of points. (<b>a</b>,<b>b</b>) Recognizing the edges of the roof as the borders of half-planes via histogram of <math display="inline"> <semantics> <mrow> <mfrac bevelled="true"> <mrow> <msub> <mi>λ</mi> <mn>2</mn> </msub> </mrow> <mrow> <msub> <mi>λ</mi> <mn>3</mn> </msub> </mrow> </mfrac> </mrow> </semantics> </math>, and (<b>c</b>) generation of the main edges of the roof. (Units of axes in (<b>b</b>,<b>c</b>) are m).</p>
Full article ">Figure A2
<p>Point cloud of the building and roof (m).</p>
Full article ">Figure A3
<p>The histogram of cosine of the inclination angle of <math display="inline"> <semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>λ</mi> <mn>1</mn> </msub> </mrow> <mo stretchy="true">→</mo> </mover> </mrow> </semantics> </math>.</p>
Full article ">Figure A4
<p>The point cloud from (<b>a</b>) <span class="html-italic">x</span>–<span class="html-italic">z</span> view and (<b>b</b>) <span class="html-italic">y</span>–<span class="html-italic">z</span> view (m).</p>
Full article ">Figure A5
<p>Reconstructed planar models (m).</p>
Full article ">Figure A6
<p>Reconstructed B-rep model of the roof is superimposed on the point cloud (m).</p>
Full article ">
28 pages, 11236 KiB  
Article
Radiation Component Calculation and Energy Budget Analysis for the Korean Peninsula Region
by Bu-Yo Kim and Kyu-Tae Lee
Remote Sens. 2018, 10(7), 1147; https://doi.org/10.3390/rs10071147 - 20 Jul 2018
Cited by 16 | Viewed by 7063
Abstract
In this study, a radiation component calculation algorithm was developed using channel data from the Himawari-8 Advanced Himawari Imager (AHI) and meteorological data from the Unified Model (UM) Local Data Assimilation and Prediction System (LDAPS). In addition, the energy budget of the Korean [...] Read more.
In this study, a radiation component calculation algorithm was developed using channel data from the Himawari-8 Advanced Himawari Imager (AHI) and meteorological data from the Unified Model (UM) Local Data Assimilation and Prediction System (LDAPS). In addition, the energy budget of the Korean Peninsula region in 2016 was calculated and its regional differences were analyzed. Radiation components derived using the algorithm were calibrated using the broadband radiation component data from the Clouds and the Earth’s Radiant Energy System (CERES) to improve their accuracy. The calculated radiation components and the CERES data showed an annual mean percent bias of less than 3.5% and a high correlation coefficient of over 0.98. The energy budget of the Korean Peninsula region was −2.4 Wm−2 at the top of the atmosphere (RT), −14.5 Wm−2 at the surface (RS), and 12.1 Wm−2 in the atmosphere (RA), with regional energy budget differences. The Seoul region had a high surface temperature (289.5 K) and a RS of −33.4 Wm−2 (surface emission), whereas the Sokcho region had a low surface temperature (284.7 K) and a RS of 5.0 Wm−2 (surface absorption), for a difference of 38.5 Wm−2. In short, regions with relatively high surface temperatures tended to show energy emission, and regions with relatively low surface temperatures tended to show energy absorption. Such regional energy imbalances can cause weather and climate changes and bring about meteorological disasters, and thus research on detecting energy budget changes must be continued. Full article
(This article belongs to the Special Issue Earth Radiation Budget)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Korean Peninsula region (red box) and altitude distribution for energy budget calculation.</p>
Full article ">Figure 2
<p>Scatter plot of calculated DLR (red: calibrated result in this study, blue: [<a href="#B71-remotesensing-10-01147" class="html-bibr">71</a>] method) and CERES DLR.</p>
Full article ">Figure 3
<p>Scatter plot of infrared atmospheric transmittance according to surface temperature and OLR.</p>
Full article ">Figure 4
<p>Density scatter plot of calculated shortwave radiation components and CERES shortwave radiation before (<b>a</b>,<b>c</b>,<b>e</b>) and after (<b>b</b>,<b>d</b>,<b>f</b>) calibration. Here, N is the total number of samples used for verification. The black dash line is the 1:1 line and the red line is the regression line. The unit of RMSE is Wm<sup>−2</sup>.</p>
Full article ">Figure 4 Cont.
<p>Density scatter plot of calculated shortwave radiation components and CERES shortwave radiation before (<b>a</b>,<b>c</b>,<b>e</b>) and after (<b>b</b>,<b>d</b>,<b>f</b>) calibration. Here, N is the total number of samples used for verification. The black dash line is the 1:1 line and the red line is the regression line. The unit of RMSE is Wm<sup>−2</sup>.</p>
Full article ">Figure 5
<p>Density scatter plot of calculated longwave radiation components and CERES longwave radiation before (<b>a</b>,<b>c</b>,<b>e</b>) and after (<b>b</b>,<b>d</b>,<b>f</b>) calibration. Here, N is the total number of samples used for verification. The black dash line is the 1:1 line and the red line is the regression line. The unit of RMSE is Wm<sup>−2</sup>.</p>
Full article ">Figure 5 Cont.
<p>Density scatter plot of calculated longwave radiation components and CERES longwave radiation before (<b>a</b>,<b>c</b>,<b>e</b>) and after (<b>b</b>,<b>d</b>,<b>f</b>) calibration. Here, N is the total number of samples used for verification. The black dash line is the 1:1 line and the red line is the regression line. The unit of RMSE is Wm<sup>−2</sup>.</p>
Full article ">Figure 6
<p>Daily mean (<b>a</b>) SHF and (<b>b</b>) LHF of MERRA-2 and LDAPS in 2016 before and after calibration.</p>
Full article ">Figure 7
<p>Daily mean time series distribution of CERES and calculated (<b>a</b>) ISR; (<b>b</b>) RSR; (<b>c</b>) DSR; and (<b>d</b>) ASR.</p>
Full article ">Figure 7 Cont.
<p>Daily mean time series distribution of CERES and calculated (<b>a</b>) ISR; (<b>b</b>) RSR; (<b>c</b>) DSR; and (<b>d</b>) ASR.</p>
Full article ">Figure 8
<p>Daily mean time series distribution of CERES and calculated (<b>a</b>) OLR; (<b>b</b>) DLR; and (<b>c</b>) ULR.</p>
Full article ">Figure 8 Cont.
<p>Daily mean time series distribution of CERES and calculated (<b>a</b>) OLR; (<b>b</b>) DLR; and (<b>c</b>) ULR.</p>
Full article ">Figure 9
<p>Daily mean time series distribution of GWNU observatory, CERES, and calculated (<b>a</b>) DSR and (<b>b</b>) DLR. Observation at GWNU observatory are missing in Julian day 108–110 and 295–301.</p>
Full article ">Figure 10
<p>Annual mean radiation components and energy budget of the Korean Peninsula region in 2016. Red: Radiation components and energy budget calculated using CERES and MERRA-2 data (values in parentheses are the results of analyzing only 350 days of data used in this study); Blue: radiation components and energy budget calculated in this study.</p>
Full article ">Figure 11
<p>Annual mean energy budget for 2016 at (<b>a</b>) the top of the atmosphere; (<b>b</b>) the surface; and (<b>c</b>) in the atmosphere; and (<b>d</b>) surface temperature distribution.</p>
Full article ">Figure 12
<p>Distribution of monthly mean energy budget (<math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>T</mi> </msub> </mrow> </semantics></math>) at the top of the atmosphere for the Korean Peninsula region in 2016.</p>
Full article ">Figure 13
<p>Distribution of monthly mean energy budget (<math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>S</mi> </msub> </mrow> </semantics></math>) at the surface for the Korean Peninsula region in 2016.</p>
Full article ">Figure 13 Cont.
<p>Distribution of monthly mean energy budget (<math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>S</mi> </msub> </mrow> </semantics></math>) at the surface for the Korean Peninsula region in 2016.</p>
Full article ">Figure 14
<p>Distribution of monthly mean energy budget (<math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>A</mi> </msub> </mrow> </semantics></math>) in the atmosphere for the Korean Peninsula region in 2016.</p>
Full article ">Figure 14 Cont.
<p>Distribution of monthly mean energy budget (<math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi>A</mi> </msub> </mrow> </semantics></math>) in the atmosphere for the Korean Peninsula region in 2016.</p>
Full article ">Figure 15
<p>Metropolitan, Yeongdong, and island region cities for studying energy budgets by region in the Korean Peninsula.</p>
Full article ">Figure 16
<p>Monthly mean energy budget time series for the Korean Peninsula and cities by region in 2016 at (<b>a</b>) the top of the atmosphere; (<b>b</b>) the surface; and (<b>c</b>) in the atmosphere.</p>
Full article ">
19 pages, 8371 KiB  
Article
Subsidence Evolution of the Firenze–Prato–Pistoia Plain (Central Italy) Combining PSI and GNSS Data
by Matteo Del Soldato, Gregorio Farolfi, Ascanio Rosi, Federico Raspini and Nicola Casagli
Remote Sens. 2018, 10(7), 1146; https://doi.org/10.3390/rs10071146 - 20 Jul 2018
Cited by 59 | Viewed by 7081
Abstract
Subsidence phenomena, as well as landslides and floods, are one of the main geohazards affecting the Tuscany region (central Italy). The monitoring of related ground deformations plays a key role in their management to avoid problems for buildings and infrastructure. In this scenario, [...] Read more.
Subsidence phenomena, as well as landslides and floods, are one of the main geohazards affecting the Tuscany region (central Italy). The monitoring of related ground deformations plays a key role in their management to avoid problems for buildings and infrastructure. In this scenario, Earth observation offers a better solution in terms of costs and benefits than traditional techniques (e.g., GNSS (Global Navigation Satellite System) or levelling networks), especially for wide area applications. In this work, the subsidence-related ground motions in the Firenze–Prato–Pistoia plain were back-investigated to track the evolution of displacement from 2003 to 2017 by means of multi-interferometric analysis of ENVISAT and Sentinel-1 imagery combined with GNSS data. The resulting vertical deformation velocities are aligned to the European Terrestrial Reference System 89 (ETRS89) datum and can be considered real velocity of displacement. The vertical ground deformation maps derived by ENVISAT and Sentinel-1 data, corrected with the GNSS, show how the area affected by subsidence for the period 2003–2010 and the period 2014–2017 evolved. The differences between the two datasets in terms of the extension and velocity values were analysed and then associated with the geological setting of the basin and external factors, e.g., new greenhouses and nurseries. This analysis allowed for reconstructing the evolution of the subsidence for the area of interest showing an increment of ground deformation in the historic centre of Pistoia Town, a decrement of subsidence in the nursery area between Pistoia and Prato cities, and changes in the industrial sector close to Prato. Full article
(This article belongs to the Special Issue Radar Interferometry for Geohazards)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Locations of the Area of Interest (AoI) and the eight GNSS stations used for the geodetic correction: five stations are inside the AoI and three are outside the AoI but close to the FPP plain.</p>
Full article ">Figure 2
<p>Sample of a stratigraphic column of the AoI (modified from [<a href="#B36-remotesensing-10-01146" class="html-bibr">36</a>]).</p>
Full article ">Figure 3
<p>Representation of the cosine directors in both ascending and descending geometries for decomposing the LOS velocity in Vertical and Horizontal components.</p>
Full article ">Figure 4
<p>Filtering of the data. E: East, W: West, Z: Zenith, and N: Nadir (from [<a href="#B35-remotesensing-10-01146" class="html-bibr">35</a>]).</p>
Full article ">Figure 5
<p>Comparison of the GNSS data recorded by the IGMI station and the nearby ENVISAT and Sentinel-1 PS data.</p>
Full article ">Figure 6
<p>Corrected ENVISAT data (vertical velocity) and isolines of the correction values adopted for the vertical velocities.</p>
Full article ">Figure 7
<p>Corrected Sentinel-1 data (vertical velocity) and isolines of the correction values adopted for the vertical velocities.</p>
Full article ">Figure 8
<p>Time series of the two PS sites in the two areas affected by subsidence in Pistoia Province: city centre of Pistoia (<b>a</b>) and southern area of Pistoia (<b>b</b>).</p>
Full article ">Figure 9
<p>Boundary of the subsidence areas recorded by the ENVISAT (2003–2010, (<b>a</b>)) and Sentinel-1 (2014–2017, (<b>b</b>)) constellations. Different subsidence rates between the two monitored periods may be due to changes in groundwater exploitation rates.</p>
Full article ">Figure 10
<p>Subsidence trace (<b>a</b>) and cross section of Pistoia. One N-S profile (A-A’, (<b>b</b>)) and two E-W profiles (B-B’, (<b>c</b>), and C-C’, (<b>d</b>)). ENVISAT data acquired from 2003 to 2010, Sentinel-1 data acquired from 2014 to 2017.</p>
Full article ">Figure 11
<p>Subsidence line (<b>a</b>) cross sections of Prato. One N-S profile (D-D’, (<b>b</b>)) and two E-W profiles (E-E’, (<b>c</b>), and F-F’, (<b>d</b>)). ENVISAT data acquired from 2003 to 2010, Sentinel-1 data acquired from 2014 to 2017.</p>
Full article ">Figure 12
<p>Geological section of Pistoia town (modified from [<a href="#B37-remotesensing-10-01146" class="html-bibr">37</a>]) across to nearly section A-A’ in <a href="#remotesensing-10-01146-f010" class="html-fig">Figure 10</a>.</p>
Full article ">
16 pages, 14504 KiB  
Article
Supervised Classification of Built-Up Areas in Sub-Saharan African Cities Using Landsat Imagery and OpenStreetMap
by Yann Forget, Catherine Linard and Marius Gilbert
Remote Sens. 2018, 10(7), 1145; https://doi.org/10.3390/rs10071145 - 20 Jul 2018
Cited by 33 | Viewed by 7919
Abstract
The Landsat archives have been made freely available in 2008, allowing the production of high resolution built-up maps at the regional or global scale. In this context, most of the classification algorithms rely on supervised learning to tackle the heterogeneity of the urban [...] Read more.
The Landsat archives have been made freely available in 2008, allowing the production of high resolution built-up maps at the regional or global scale. In this context, most of the classification algorithms rely on supervised learning to tackle the heterogeneity of the urban environments. However, at a large scale, the process of collecting training samples becomes a huge project in itself. This leads to a growing interest from the remote sensing community toward Volunteered Geographic Information (VGI) projects such as OpenStreetMap (OSM). Despite the spatial heterogeneity of its contribution patterns, OSM provides an increasing amount of information on the earth’s surface. More interestingly, the community has moved beyond street mapping to collect a wider range of spatial data such as building footprints, land use, or points of interest. In this paper, we propose a classification method that makes use of OSM to automatically collect training samples for supervised learning of built-up areas. To take into account a wide range of potential issues, the approach is assessed in ten Sub-Saharan African urban areas from various demographic profiles and climates. The obtained results are compared with: (1) existing high resolution global urban maps such as the Global Human Settlement Layer (GHSL) or the Human Built-up and Settlements Extent (HBASE); and (2) a supervised classification based on manually digitized training samples. The results suggest that automated supervised classifications based on OSM can provide performances similar to manual approaches, provided that OSM training samples are sufficiently available and correctly pre-processed. Moreover, the proposed method could reach better results in the near future, given the increasing amount and variety of information in the OSM database. Full article
(This article belongs to the Special Issue Citizen Science and Earth Observation II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Examples of digitized samples in Dakar, Senegal: (<b>a</b>) Built-up; (<b>b</b>) Bare soil; (<b>c</b>) Low vegetation; and (<b>d</b>) High vegetation. The grid corresponds to the 30 meters Landsat pixels. Satellite imagery courtesy of Google Earth.</p>
Full article ">Figure 2
<p>Evolution of OSM data availability in our case studies between 2011 and 2018.</p>
Full article ">Figure 3
<p>Availability and median surface of building footprints in each case study.</p>
Full article ">Figure 4
<p>Urban blocks extracted from the OSM road network in Windhoek, Namibia (transparent: surface greater than 10 ha; red: surface greater than 1 ha; green: surface lower than 1 ha). Satellite imagery courtesy of Google.</p>
Full article ">Figure 5
<p>Quality and quantity of built-up training samples extracted from OSM building footprints according to the minimum coverage threshold in the 10 case studies: (<b>a</b>) mean spectral distance to the reference built-up samples; and (<b>b</b>) mean number of samples (in pixels).</p>
Full article ">Figure 6
<p>Quality and quantity of built-up training samples extracted from OSM urban blocks according to maximum surface threshold in the 10 case studies: (<b>a</b>) mean spectral distance to the reference built-up samples; and (<b>b</b>) number of samples (in pixels) in the five case studies with the lowest data availability.</p>
Full article ">Figure 7
<p>Most similar land cover of each OSM non-built-up object according to its tag. Circles are logarithmically proportional to the number of pixels available.</p>
Full article ">Figure 8
<p>Quality and quantity of non-built-up training samples extracted from the OSM-based urban distance raster: (<b>a</b>) mean spectral distance to the reference built-up samples according to the urban distance; and (<b>b</b>) number of samples (in pixels) in the five case studies with the lowest sample availability.</p>
Full article ">Figure 9
<p>Areas with high rates of misclassifications in: (<b>a</b>) Katsina; (<b>b</b>) Johannesburg; (<b>c</b>) Gao; and (<b>d</b>) Dakar. Satellite imagery courtesy of Google Earth.</p>
Full article ">Figure 10
<p>Relationship between the number of training samples and the classification F1-score (the outlier Johannesburg is excluded from the graph).</p>
Full article ">
18 pages, 9363 KiB  
Article
Multitemporal Classification of River Floodplain Vegetation Using Time Series of UAV Images
by Wimala Van Iersel, Menno Straatsma, Hans Middelkoop and Elisabeth Addink
Remote Sens. 2018, 10(7), 1144; https://doi.org/10.3390/rs10071144 - 19 Jul 2018
Cited by 37 | Viewed by 6334
Abstract
The functions of river floodplains often conflict spatially, for example, water conveyance during peak discharge and diverse riparian ecology. Such functions are often associated with floodplain vegetation. Frequent monitoring of floodplain land cover is necessary to capture the dynamics of this vegetation. However, [...] Read more.
The functions of river floodplains often conflict spatially, for example, water conveyance during peak discharge and diverse riparian ecology. Such functions are often associated with floodplain vegetation. Frequent monitoring of floodplain land cover is necessary to capture the dynamics of this vegetation. However, low classification accuracies are found with existing methods, especially for relatively similar vegetation types, such as grassland and herbaceous vegetation. Unmanned aerial vehicle (UAV) imagery has great potential to improve the classification of these vegetation types owing to its high spatial resolution and flexibility in image acquisition timing. This study aimed to evaluate the increase in classification accuracy obtained using multitemporal UAV images versus single time step data on floodplain land cover classification and to assess the effect of varying the number and timing of imagery acquisition moments. We obtained a dataset of multitemporal UAV imagery and field reference observations and applied object-based Random Forest classification (RF) to data of different time step combinations. High overall accuracies (OA) exceeding 90% were found for the RF of floodplain land cover, with six vegetation classes and four non-vegetation classes. Using two or more time steps compared with a single time step increased the OA from 96.9% to 99.3%. The user’s accuracies of the classes with large similarity, such as natural grassland and herbaceous vegetation, also exceeded 90%. The combination of imagery from June and September resulted in the highest OA (98%) for two time steps. Our method is a practical and highly accurate solution for monitoring areas of a few square kilometres. For large-scale monitoring of floodplains, the same method can be used, but with data from airborne platforms covering larger extents. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Breemwaard study area on the Southern bank of the river Waal. (<b>A</b>) Location in the Rhine Delta; (<b>B</b>) Field impressions of typical vegetation in the study area; (<b>C</b>) Orthophoto of June with classified reference polygons. Polygons marked with a dotted circle were obtained from the imagery and the remaining polygons were obtained from the field.</p>
Full article ">Figure 2
<p>Example of a field plot with mixture of grassland and herbaceous vegetation, which was divided into homogeneous polygons with field sketches of September.</p>
Full article ">Figure 3
<p>Workflow of multitemporal segmentation, classification, and validation, resulting in accuracy by decreasing the number of time steps. The subscript <math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>R</mi> <mi>G</mi> <mi>B</mi> </mrow> </msub> </semantics></math> indicates a true colour layer and <math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>C</mi> <mi>I</mi> <mi>R</mi> </mrow> </msub> </semantics></math> indicates a false colour. Random Forest (RF) and varSelRF are classification models in R.</p>
Full article ">Figure 4
<p>Accuracy of the RF classification by decreasing number of time steps. OA<math display="inline"><semantics> <msub> <mrow/> <mrow> <mi>v</mi> <mi>a</mi> <mi>l</mi> </mrow> </msub> </semantics></math> is the classification accuracy based on validation with an independent data set. For only spectral attributes, <tt>maxnodes</tt> was 25.</p>
Full article ">Figure 5
<p>Classified land-cover map of the Breemwaard floodplain with data from the best performing RF, which used data from all time steps and <tt>maxnodes</tt> set to default. (<b>A</b>) Overview of the classified study area. (<b>B</b>) Orthophoto of September of zoomed in area C’ and (<b>C</b>) zoomed in area of classified grassland and herbaceous vegetation.</p>
Full article ">
18 pages, 7129 KiB  
Article
Variations in Remotely-Sensed Phytoplankton Size Structure of a Cyclonic Eddy in the Southwest Indian Ocean
by Tarron Lamont, Raymond G. Barlow and Robert J. W. Brewin
Remote Sens. 2018, 10(7), 1143; https://doi.org/10.3390/rs10071143 - 19 Jul 2018
Cited by 4 | Viewed by 6171
Abstract
Phytoplankton size classes were derived from weekly-averaged MODIS Aqua chlorophyll a data over the southwest Indian Ocean in order to assess changes in surface phytoplankton community structure within a cyclonic eddy as it propagated across the Mozambique Basin in 2013. Satellite altimetry was [...] Read more.
Phytoplankton size classes were derived from weekly-averaged MODIS Aqua chlorophyll a data over the southwest Indian Ocean in order to assess changes in surface phytoplankton community structure within a cyclonic eddy as it propagated across the Mozambique Basin in 2013. Satellite altimetry was used to identify and track the southwesterly movement of the eddy from its origin off Madagascar in mid-June until mid-October, when it eventually merged with the Agulhas Current along the east coast of South Africa. Nano- and picophytoplankton comprised most of the community in the early phase of the eddy development in June, but nanophytoplankton then dominated in austral winter (July and August). Microphytoplankton was entrained into the eddy by horizontal advection from the southern Madagascar shelf, increasing the proportion of microphytoplankton to 23% when the chlorophyll a levels reached a peak of 0.36 mg·m−3 in the third week of July. Chlorophyll a levels declined to <0.2 mg·m−3 in austral spring (September and October) as the eddy propagated further to the southwest. Picophytoplankton dominated the community during the spring period, accounting for >50% of the population. As far as is known, this is the first study to investigate temporal changes in chlorophyll a and community structure in a cyclonic eddy propagating across an ocean basin in the southwest Indian Ocean. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Figure 1

Figure 1
<p>Main oceanographic features in the Mozambique Channel and Mozambique Basin. The southern branch of the East Madagascar Current (SEMC), the Agulhas Current, Mozambique Channel eddies, as well as dipoles stemming from the SEMC are indicated. Anticlockwise (clockwise) circulation features indicate anticyclonic (cyclonic) eddies. Black contours indicate the 1000 m, 2000 m, 3000 m, 4000 m, and 5000 m bathymetric contours.</p>
Full article ">Figure 2
<p>Verification of the use of the three-component model for studying phytoplankton size structure within the MB eddy. (<b>a</b>–<b>c</b>) show the fractions of micro-, nano-, and picophytoplankton, respectively, as a function of chlorophyll <span class="html-italic">a</span> for in situ measurements collected during the passage of the MB eddy with the three-component model overlain. For comparison, the data from the Lamont et al. [<a href="#B19-remotesensing-10-01143" class="html-bibr">19</a>] (L18) study is also shown, and the grey shading represents uncertainty in the fractions based on the validation in the L18 study (see their <a href="#remotesensing-10-01143-f003" class="html-fig">Figure 3</a>). MAD is the median absolute difference between the model and the in situ size fractions from the MB eddy. (<b>d</b>–<b>g</b>) show the in situ chlorophyll <span class="html-italic">a</span> and size fractions overlain onto MODIS-Aqua estimates from 20 and 24 July 2013, merged (averaged) into a single image. The in situ samples are coloured on the same scale as the satellite images.</p>
Full article ">Figure 3
<p>Sea Surface Height (colour contours) and geostrophic velocity (black arrows) over the Mozambique Basin on 17 June 2013. The black box highlights the location of the cyclonic eddy.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) Daily Sea Surface Height (colour contours) and geostrophic velocity (black arrows) on selected days for 21 June to 22 July 2013; and (<b>e</b>–<b>h</b>) 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> composites for 18 June to 27 July 2013, over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 5
<p>Fractional contribution of (<b>a</b>–<b>d</b>) micro-, (<b>e</b>–<b>h</b>) nano-, and (<b>i</b>–<b>l</b>) picophytoplankton to MODIS Aqua chlorophyll <span class="html-italic">a</span> for 18 June to 27 July 2013 over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>d</b>) Daily Sea Surface Height (colour contours) and geostrophic velocity (black arrows) on selected days for 31 July to 24 August 2013; and (<b>e</b>–<b>h</b>) 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> composites for 28 July to 28 August 2013, over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 7
<p>Fractional contribution of (<b>a</b>–<b>d</b>) micro-, (<b>e</b>–<b>h</b>) nano-, and (<b>i</b>–<b>l</b>) picophytoplankton to MODIS Aqua chlorophyll <span class="html-italic">a</span> for 28 July to 28 August 2013 over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>d</b>) Daily Sea Surface Height (colour contours) and geostrophic velocity (black arrows) on selected days for 1 September to 11 October 2013; and (<b>e</b>–<b>h</b>) 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> composites for 29 August to 15 October 2013, over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 9
<p>Fractional contribution of (<b>a</b>–<b>d</b>) micro-, (<b>e</b>–<b>h</b>) nano-, and (<b>i</b>–<b>l</b>) picophytoplankton to MODIS Aqua chlorophyll <span class="html-italic">a</span> for 29 August to 15 October 2013 over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 10
<p>Temporal variation in (<b>a</b>) daily Sea Surface Height (SSH) (black line) and 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> (green line and dots), and (<b>b</b>) the fractional contribution of micro-, nano-, and picophytoplankton at the centre of the cyclonic eddy as it propagated across the Mozambique Basin. Vertical bars indicate the standard deviation of chlorophyll <span class="html-italic">a</span> and the fractional contributions of micro- (green dots and line), nano- (blue dots and line), and picophytoplankton (red dots and line) for the 3 × 3 pixel window at the centre of the eddy.</p>
Full article ">
25 pages, 5009 KiB  
Article
Improving Spatial-Temporal Data Fusion by Choosing Optimal Input Image Pairs
by Donghui Xie, Feng Gao, Liang Sun and Martha Anderson
Remote Sens. 2018, 10(7), 1142; https://doi.org/10.3390/rs10071142 - 19 Jul 2018
Cited by 42 | Viewed by 5902
Abstract
Spatial and temporal data fusion approaches have been developed to fuse reflectance imagery from Landsat and the Moderate Resolution Imaging Spectroradiometer (MODIS), which have complementary spatial and temporal sampling characteristics. The approach relies on using Landsat and MODIS image pairs that are acquired [...] Read more.
Spatial and temporal data fusion approaches have been developed to fuse reflectance imagery from Landsat and the Moderate Resolution Imaging Spectroradiometer (MODIS), which have complementary spatial and temporal sampling characteristics. The approach relies on using Landsat and MODIS image pairs that are acquired on the same day to estimate Landsat-scale reflectance on other MODIS dates. Previous studies have revealed that the accuracy of data fusion results partially depends on the input image pair used. The selection of the optimal image pair to achieve better prediction of surface reflectance has not been fully evaluated. This paper assesses the impacts of Landsat-MODIS image pair selection on the accuracy of the predicted land surface reflectance using the Spatial and Temporal Adaptive Reflectance Fusion Model (STARFM) over different landscapes. MODIS images from the Aqua and Terra platforms were paired with images from the Landsat 7 Enhanced Thematic Mapper Plus (ETM+) and Landsat 8 Operational Land Imager (OLI) to make different pair image combinations. The accuracy of the predicted surface reflectance at 30 m resolution was evaluated using the observed Landsat data in terms of mean absolute difference, root mean square error and correlation coefficient. Results show that the MODIS pair images with smaller view zenith angles produce better predictions. As expected, the image pair closer to the prediction date during a short prediction period produce better prediction results. For prediction dates distant from the pair date, the predictability depends on the temporal and spatial variability of land cover type and phenology. The prediction accuracy for forests is higher than for crops in our study areas. The Normalized Difference Vegetation Index (NDVI) for crops is overestimated during the non-growing season when using an input image pair from the growing season, while NDVI is slightly underestimated during the growing season when using an image pair from the non-growing season. Two automatic pair selection strategies are evaluated. Results show that the strategy of selecting the MODIS pair date image that most highly correlates with the MODIS image on the prediction date produces more accurate predictions than the nearest date strategy. This study demonstrates that data fusion results can be improved if appropriate image pairs are used. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Experimental sites and the boundaries of the Landsat scenes (<b>a</b>) with corresponding land cover maps from the Cropland Data Layer for 2015 (<b>b</b>).</p>
Full article ">Figure 2
<p>Cloud-free Landsat 7 (green dots) and 8 (red dots) images from 2015 were used in the study. Landsat 8 images from p28r31 and p29r31 were used as pair images while Landsat 7 images were used to validate data fusion results in study site 1. Both Landsat 7 and 8 images for p25r36 can be used as pair images since Landsat 7 ETM+ images in study site 2 are not affected by the SLC-off problem. All validations only used Landsat images that were not used as input pair images.</p>
Full article ">Figure 3
<p>Schematic diagram for the nearest date (ND) and higher correlation (HC) pair selection strategies. The black line and red line represent Pearson correlation coefficient (<span class="html-italic">R</span>) values between MODIS images at pair base dates <span class="html-italic">t<sub>1</sub></span> and <span class="html-italic">t<sub>2</sub></span>, respectively, and the prediction dates. <span class="html-italic">R</span> generally decrease with increasing MODIS image time separation. <span class="html-italic">t<sub>m</sub></span> and <span class="html-italic">t<sub>n</sub></span> are the split dates determined by the HC and ND strategies, individually.</p>
Full article ">Figure 4
<p>Temporally-averaged RMSE (&lt;RMSE&gt;) (<b>a</b>) and R (&lt;R&gt;) (<b>b</b>) for different combinations of MODIS and Landsat pair images for site 2. NBAR, Terra, and Aqua in the x axis represent the combinations of the Terra and Aqua combined NBAR, Terra NBAR, and Aqua NBAR with Landsat, respectively. Labels in the plot show the combinations of MODIS (NBAR, Terra, Aqua) and Landsat (L7 and L8) image pairs.</p>
Full article ">Figure 5
<p>Temporally-averaged mean absolute difference (&lt;MAD&gt;) for different combinations of pair images from MODIS (Terra, Aqua, and combined NBAR) and Landsat (Landsat 7 and 8) for site 1 (<b>a</b>) and site 2 (<b>b</b>). In site 1, &lt;MAD&gt; is calculated based on the predicted image from Landsat 8 (p28r31) in order to compare different pair combinations consistently.</p>
Full article ">Figure 6
<p>Comparison of the composite NBAR images from Terra (250 m), Aqua (250 m), and the combined products (500 m) with Landsat image for site 1 (<b>A</b>,<b>B</b>) and site 2 (<b>C</b>,<b>D</b>) from different Landsat dates. The images are composited using NIR, red and red bands as RGB (only red and NIR bands are available at 250 m resolution). The images for each date are linearly stretched using the same stretch range for better visualization. The view zenith angles (VZA) for Terra and Aqua images are the averaged VZAs from original directional reflectance. The figure shows that even after BRDF correction, the image quality and spatial detail in the MODIS data used as fusion inputs can vary significantly.</p>
Full article ">Figure 6 Cont.
<p>Comparison of the composite NBAR images from Terra (250 m), Aqua (250 m), and the combined products (500 m) with Landsat image for site 1 (<b>A</b>,<b>B</b>) and site 2 (<b>C</b>,<b>D</b>) from different Landsat dates. The images are composited using NIR, red and red bands as RGB (only red and NIR bands are available at 250 m resolution). The images for each date are linearly stretched using the same stretch range for better visualization. The view zenith angles (VZA) for Terra and Aqua images are the averaged VZAs from original directional reflectance. The figure shows that even after BRDF correction, the image quality and spatial detail in the MODIS data used as fusion inputs can vary significantly.</p>
Full article ">Figure 7
<p>Comparison of fused images (site 2) on DOY 2015-219 to the direct Landsat observation on that date (<b>a</b>). Three data combinations used in the fusion include (<b>b</b>) Terra + L8, (<b>c</b>) Aqua + L8, and (<b>d</b>) the Terra and Aqua combined NBAR + L8 from DOY 2015-027. The correlation coefficients (R) between the fused images and observed Landsat image are listed.</p>
Full article ">Figure 8
<p>Comparison of MAD between observed and predicted Landsat-scale NDVI for different prediction dates (x-axis) using different input pair dates (signified by different color lines) for crops at site 1 (<b>a</b>) and forest types at site 2 (<b>b</b>). Mean NDVI from vegetation land cover types (corn and soybean in site 1, and deciduous, evergreen and mixed forests in site 2) were extracted and averaged based on observed Landsat and CDL data. Each symbol represents the MAD between the observed Landsat (not used in data fusion) and the fused result. Four pair dates, including days 80 (non-growing season), 167 (around green-up dates), 224 (around peak growing dates), and 279 (around senescence) are highlighted (enlarged dots) for site 1. Two pair dates, including day 43 from the non-growing season and day 195 from the growing season are highlighted (enlarged dots) for site 2.</p>
Full article ">Figure 9
<p>Comparison of &lt;MAD&gt; (<b>a</b>) and &lt;R&gt; (<b>b</b>) as a function of pair date for different land cover types from both study sites (the results of corn and soybean are from site 1 and those of forests are from site 2).</p>
Full article ">Figure 10
<p>Comparison of NDVI time series for corn (<b>a</b>) and soybean (<b>b</b>) from original Landsat observations (black) and the fused images using pair images on day 80 (red), 224 (green), and 283 (blue) for site 1.</p>
Full article ">Figure 11
<p>MAD for different prediction dates based on one image pair option from either day 80 or 263 when using the nearest date (ND) and higher correlation (HC) strategies. The HC strategy shows the smaller MAD (red) and better data fusion results than the ND strategy (black).</p>
Full article ">Figure 12
<p>&lt;MAD&gt; of NDVI decreases as the available number of pair images increases for both the ND and HC pair selection strategy. Each point in the plot shows a combination of pair dates, as indicated by the adjacent label.</p>
Full article ">Figure 13
<p>MSD of Landsat reflectance (red and NIR) and NDVI for site 1 (<b>a</b>) and site 2 (<b>b</b>). A higher MSD means a more heterogeneous landscape.</p>
Full article ">
16 pages, 4794 KiB  
Article
The Benefit of the Geospatial-Related Waveforms Analysis to Extract Weak Laser Pulses
by Tee-Ann Teo and Wan-Yi Yeh
Remote Sens. 2018, 10(7), 1141; https://doi.org/10.3390/rs10071141 - 19 Jul 2018
Cited by 5 | Viewed by 4056
Abstract
Waveform lidar provides both geometric and waveform properties from the entire returned signals. The waveform analysis is an important process to extract the attributes of the reflecting surface from the waveform. The proposed method analyzes the geospatial relationship between the return signals by [...] Read more.
Waveform lidar provides both geometric and waveform properties from the entire returned signals. The waveform analysis is an important process to extract the attributes of the reflecting surface from the waveform. The proposed method analyzes the geospatial relationship between the return signals by combining the sequential waves. The idea of this method is to analyze the waveform parameters from sequential waves. Since the adjacent return signals are geospatially correlated, they have similar waveform properties that can be used to validate the correctness of the extracted waveform parameters. The proposed method includes three major steps: (1) single-waveform processing for the initial echo detection; (2) multi-waveform processing using waveform alignment and stacking; (3) verification of the enhanced weak return. The experimental waveform lidar data were acquired using Leica ALS60, Optech Pegasus, and Riegl Q680i. The experimental result indicates that the proposed method successfully extracts the weak returns while considering the geospatial relationships. The correctness and increasing rate of the extracted ground points are related to the vegetated coverage such as the complexity and density. The correctness is above 76% in this study. Because the nearest waveform has a higher correlation, the increase in distance of adjacent waveforms will reduce the correctness of the enhanced weak return. Full article
(This article belongs to the Special Issue Future Trends and Applications for Airborne Laser Scanning)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The illustration of initial peaks’ detection: (<b>a</b>) Original extracted points; (<b>b</b>) Results after neighbor point removal; (<b>b</b>) Results after lower energy point removal.</p>
Full article ">Figure 2
<p>The illustration of Gaussian decomposition: (<b>a</b>) Input waveform and initial peaks; (<b>b</b>) Results of Gaussian fitting; (<b>b</b>) Results of Gaussian decomposition.</p>
Full article ">Figure 3
<p>The illustration of the sequential waveforms: (<b>a</b>) Top view of lidar point clouds and colored by height in meter (the color bar indicate the height in meters); (<b>b</b>) Profile of sequential waveforms (the color bar indicate the intensity); (<b>c</b>) Sequential-waveform-covered tree.</p>
Full article ">Figure 4
<p>The illustration of multi-waveform processing: (<b>a</b>) Sequence Waveforms; (<b>b</b>) Result of waveforms alignment; (<b>c</b>) Result of waveform stacking (black color waveform is stacked waveform); (<b>d</b>) Result of extracted ground point using stacked waveform.</p>
Full article ">Figure 5
<p>The original lidar points from Leica ALS60: (<b>a</b>) Area 1 with lower tree density; (<b>b</b>) Area 2 with medium tree density; (<b>c</b>) Area 3 with higher tree density.</p>
Full article ">Figure 6
<p>The histogram of waveform strength: (<b>a</b>) The histogram of adjacent waveforms at time <span class="html-italic">t</span> − <span class="html-italic">1</span>; (<b>b</b>) The histogram of adjacent waveforms at time <span class="html-italic">t</span> + 1; (<b>c</b>) The histogram of master waveforms at time <span class="html-italic">t</span>; (<b>d</b>) The background noise of master waveforms at time <span class="html-italic">t</span>.</p>
Full article ">Figure 7
<p>The different spatial distances: (<b>a</b>) Distance 1: 0.4–0.7 m; (<b>b</b>) Distance 2: 0.7–1.4 m; (<b>c</b>) Distance 3: 1.4–2.1 m.</p>
Full article ">Figure 8
<p>The correctness of different spatial distances.</p>
Full article ">Figure 9
<p>The different waveforms in this study: (<b>a</b>) a waveform with 256 samples from case 1; (<b>b</b>) a waveform with 100 samples from case 4.</p>
Full article ">
15 pages, 5954 KiB  
Article
Detecting and Quantifying a Massive Invasion of Floating Aquatic Plants in the Río de la Plata Turbid Waters Using High Spatial Resolution Ocean Color Imagery
by Ana I. Dogliotti, Juan I. Gossn, Quinten Vanhellemont and Kevin G. Ruddick
Remote Sens. 2018, 10(7), 1140; https://doi.org/10.3390/rs10071140 - 19 Jul 2018
Cited by 32 | Viewed by 7065
Abstract
The massive development of floating plants in floodplain lakes and wetlands in the upper Middle Paraná river in the La Plata basin is environmentally and socioeconomically important. Every year aquatic plant detachments drift downstream arriving in small amounts to the Río de la [...] Read more.
The massive development of floating plants in floodplain lakes and wetlands in the upper Middle Paraná river in the La Plata basin is environmentally and socioeconomically important. Every year aquatic plant detachments drift downstream arriving in small amounts to the Río de la Plata, but huge temporary invasions have been observed every 10 or 15 years associated to massive floods. From late December 2015, heavy rains driven by a strong El Niño increased river levels, provoking a large temporary invasion of aquatic plants from January to May 2016. This event caused significant disruption of human activities via clogging of drinking water intakes in the estuary, blocking of ports and marinas and introducing dangerous animals from faraway wetlands into the city. In this study, we developed a scheme to map floating vegetation in turbid waters using high-resolution imagery, like Sentinel-2/SMI (MultiSpectral Imager), Landsat-8/OLI (Operational Land Imager), and Aqua/MODIS (MODerate resolution Imager Spectroradiometer)-250 m. A combination of the Floating Algal Index (that make use of the strong signal in the NIR part of the spectrum), plus conditions set on the RED band (to avoid misclassifying highly turbid waters) and on the CIE La*b* color space coordinates (to confirm the visually “green” pixels as floating vegetation) were used. A time-series of multisensor high resolution imagery was analyzed to study the temporal variability, covered area and distribution of the unusual floating macroalgae invasion that started in January 2016 in the Río de la Plata estuary. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location of the Paraná–Paraguay fluvial corridor and Río de la Plata estuary; (<b>b</b>) schematic representation of low and high water conditions (water pixels are dark-blue in the Red-Green-Blue Landsat images of the Paraná delta) that can be found in the floodplain (modified from [<a href="#B14-remotesensing-10-01140" class="html-bibr">14</a>]); (<b>c</b>) a passenger ferry terminal invaded by water hyacinth (<b>d</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) In situ reflectance spectra of water hyacinth mats (green) collected in January 2016 and RdP turbid waters (black) collected in previous cruises. Thick lines represent the mean of 9 and &gt;50 measurements from <span class="html-italic">Eichhornia crassipes</span> mats and turbid waters of RdP, respectively. Dashed lines correspond to one standard deviation. Photographs of (<b>b</b>) floating water hyacinth <span class="html-italic">E. crassipes</span>, and (<b>c</b>) measurement setup.</p>
Full article ">Figure 3
<p>Rayleigh-corrected spectra of pixels that look green in the RGB combination thus containing a detectable amount of floating vegetation (green), turbid (blue) and extreme turbid (brown) waters extracted from MODIS-Aqua image of the RdP estuary. The FAI index is schematically indicated as well as the threshold applied to the <span class="html-italic">R<sub>rc</sub></span> in the RED band.</p>
Full article ">Figure 4
<p>(<b>a</b>) Subset of S2 imagery (9 February 2016) and (<b>b</b>) zoom over a patch of floating vegetation and passing ship; (<b>c</b>) a* vs. b* diagram of the pixels in <b>b</b>) colored with the RGB values. Pixels that passed the first spectral criteria (FAI &gt; 0 and RED &gt; 0.09) are indicated with black contour, while the dashed line indicates the a* threshold used to further determine the presence of FV.</p>
Full article ">Figure 5
<p>RGB details of floating vegetation (FV) patches for different systems and dates are shown in the (<b>upper row</b>). Red squares indicate the pixels that are flagged as FV by FAI but not by NDVI, also indicated by the red square area in the NDVI vs. FAI scatter plots (<b>lower row</b>). The grey area corresponds to the pixels not identified as FV neither by FAI nor NDVI.</p>
Full article ">Figure 6
<p>Flowchart of the FAIT scheme used to detect floating vegetation in the turbid waters of RdP estuary.</p>
Full article ">Figure 7
<p>Quasi-true-color red-green-blue (RGB) images from L8 and S2A (<b>left</b>) and pixel flagged as floating vegetation are shown for a subset of each image (dashed squares in RGB) and from the same day MODIS-Aqua image. The acquisition date and time (UTC) of each image is also indicated.</p>
Full article ">Figure 8
<p>Quasi-true-color S2A data on 9 February 2016 over a patch of floating vegetation (<b>upper left</b>) and spatially averaged to 30, 300 and 1000 m pixel size. Corresponding spectra of the original S2A green pixel and the arithmetic mean value of 3 × 3, 29 × 29 and 99 × 99 pixel boxes (<b>lower right</b>).</p>
Full article ">Figure 9
<p>Spectra of selected endmembers of floating vegetation (FV), turbid waters (TW), moderate turbid waters (MT), highly turbid waters caused by intense dredging activity (DRG), and extreme turbid waters (XTW) extracted from different S2 images (only selected bands are shown).</p>
Full article ">Figure 10
<p>(<b>a</b>) Pseudo-true-color image of PROBA-V (R = 650 nm, G = 835 nm, B = 470 nm) image (100 m) acquired on 22 April 2016 showing the region of interest (ROI) used for the FV area coverage analysis (grey dashed square), S2A 21HUB area (dotted-black line) and L8 225/84 path/row area (dotted white line). (<b>b</b>) FV area (km<sup>2</sup>) detected by L8 (green), S2 (magenta), MA (blue) for the 2015–2016 time period overlaid over the RdP outflow anomaly (dashed light-blue line). Availability of non cloudy imagery for each sensor is shown on top.</p>
Full article ">Figure 11
<p>Quasi-true-color (RGB) Sentinel-2A image acquired on 16 October 2016. (<b>a</b>) A light-brown plume parallel to Buenos Aires’ coastline produced by sediments after heavy dredging activities is clearly seen; and (<b>b</b>) a zoom (dashed square in [<b>a</b>]) shows the details of the temporarily vegetated island generated by the accumulation of sediments that this activity produced.</p>
Full article ">
23 pages, 2753 KiB  
Article
Analysis of Airborne Optical and Thermal Imagery for Detection of Water Stress Symptoms
by Max Gerhards, Martin Schlerf, Uwe Rascher, Thomas Udelhoven, Radoslaw Juszczak, Giorgio Alberti, Franco Miglietta and Yoshio Inoue
Remote Sens. 2018, 10(7), 1139; https://doi.org/10.3390/rs10071139 - 19 Jul 2018
Cited by 72 | Viewed by 7320
Abstract
High-resolution airborne thermal infrared (TIR) together with sun-induced fluorescence (SIF) and hyperspectral optical images (visible, near- and shortwave infrared; VNIR/SWIR) were jointly acquired over an experimental site. The objective of this study was to evaluate the potential of these state-of-the-art remote sensing techniques [...] Read more.
High-resolution airborne thermal infrared (TIR) together with sun-induced fluorescence (SIF) and hyperspectral optical images (visible, near- and shortwave infrared; VNIR/SWIR) were jointly acquired over an experimental site. The objective of this study was to evaluate the potential of these state-of-the-art remote sensing techniques for detecting symptoms similar to those occurring during water stress (hereinafter referred to as ‘water stress symptoms’) at airborne level. Flights with two camera systems (Telops Hyper-Cam LW, Specim HyPlant) took place during 11th and 12th June 2014 in Latisana, Italy over a commercial grass (Festuca arundinacea and Poa pratense) farm with plots that were treated with an anti-transpirant agent (Vapor Gard®; VG) and a highly reflective powder (kaolin; KA). Both agents affect energy balance of the vegetation by reducing transpiration and thus reducing latent heat dissipation (VG) and by increasing albedo, i.e., decreasing energy absorption (KA). Concurrent in situ meteorological data from an on-site weather station, surface temperature and chamber flux measurements were obtained. Image data were processed to orthorectified maps of TIR indices (surface temperature (Ts), Crop Water Stress Index (CWSI)), SIF indices (F687, F780) and VNIR/SWIR indices (photochemical reflectance index (PRI), normalised difference vegetation index (NDVI), moisture stress index (MSI), etc.). A linear mixed effects model that respects the nested structure of the experimental setup was employed to analyse treatment effects on the remote sensing parameters. Airborne Ts were in good agreement (∆T < 0.35 K) compared to in situ Ts measurements. Maps and boxplots of TIR-based indices show diurnal changes: Ts was lowest in the early morning, increased by 6 K up to late morning as a consequence of increasing net radiation and air temperature (Tair) and remained stable towards noon due to the compensatory cooling effect of increased plant transpiration; this was also confirmed by the chamber measurements. In the early morning, VG treated plots revealed significantly higher Ts compared to control (CR) plots (p = 0.01), while SIF indices showed no significant difference (p = 1.00) at any of the overpasses. A comparative assessment of the spectral domains regarding their capabilities for water stress detection was limited due to: (i) synchronously overpasses of the two airborne sensors were not feasible, and (ii) instead of a real water stress occurrence only water stress symptoms were simulated by the chemical agents. Nevertheless, the results of the study show that the polymer di-1-p-menthene had an anti-transpiring effect on the plant while photosynthetic efficiency of light reactions remained unaffected. VNIR/SWIR indices as well as SIF indices were highly sensitive to KA, because of an overall increase in spectral reflectance and thus a reduced absorbed energy. On the contrary, the TIR domain was highly sensitive to subtle changes in the temperature regime as induced by VG and KA, whereas VNIR/SWIR and SIF domain were less affected by VG treatment. The benefit of a multi-sensor approach is not only to provide useful information about actual plant status but also on the causes of biophysical, physiological and photochemical changes. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Overview of the study site with the experimental setup. The study site was located in northeastern Italy near the town of Latisana (<b>a</b>). (<b>b</b>) shows the locations of the experimental plots, pool and weather station. A scheme of the experimental design is presented in the legend.</p>
Full article ">Figure 2
<p>Environmental conditions (net radiation (Rn), air temperature (<span class="html-italic">T<sub>air</sub></span>), surface temperature (<span class="html-italic">T<sub>s</sub></span>), <span class="html-italic">T<sub>s</sub></span>–<span class="html-italic">T<sub>air</sub></span>, vapour pressure deficit (VPD), as well as soil water content (SWC)) for 11th and 12th of June 2014 over a non-treated grass surface measured by the weather station.</p>
Full article ">Figure 3
<p>Diurnal changes in chamber flux measurements for Control plot (CR) (green) and Vapor Gard<sup>®</sup> (VG) (orange) treatments. Solid lines are showing H<sub>2</sub>O fluxes in mmols H<sub>2</sub>O m<sup>−2</sup> s<sup>−1</sup> and Gross Ecosystem Productivity (GEP) is represented in dashed lines measured in µmols CO<sub>2</sub> m<sup>−2</sup> s<sup>−1</sup> (<b>upper left</b>). PAR (Photosynthetically Active Radiation) as measured in µmols m<sup>−2</sup> s<sup>−1</sup> is represented with dotted lines (<b>upper right</b>). Boxplots for H<sub>2</sub>O fluxes (<b>lower left</b>) and H<sub>2</sub>O fluxes normalised by PAR (<b>lower right</b>).</p>
Full article ">Figure 4
<p>Diurnal <span class="html-italic">T<sub>s</sub></span> [K] and CWSI maps at 11th June (top), F<sub>687</sub>, F<sub>760</sub>, PRI, NDVI and LWI maps of the same day (11th June, 14:52 CEST) (bottom left), and locations of the treatments (bottom right).</p>
Full article ">Figure 5
<p>Boxplots of the different treatments (CR, KA, VG) at the three overpasses for <span class="html-italic">T<sub>s</sub></span>, <span class="html-italic">T<sub>s</sub></span>–<span class="html-italic">T<sub>air</sub></span> and CWSI. Different letters indicate significant differences (<span class="html-italic">* p</span> ≤ 0.05).</p>
Full article ">Figure 6
<p>VNIR/SWIR mean reflectance spectra from <span class="html-italic">HyPlant</span>’s dual-channel module.</p>
Full article ">Figure 7
<p>Boxplots of the different treatments (CR, KA, and VG) at three <span class="html-italic">HyPlant</span> overpasses for VNIR/SWIR based indices and F<sub>687</sub> and F<sub>760</sub>. Different letters indicate significant differences (<span class="html-italic">* p</span> ≤ 0.05).</p>
Full article ">
24 pages, 8308 KiB  
Article
A Comparison of Crop Parameters Estimation Using Images from UAV-Mounted Snapshot Hyperspectral Sensor and High-Definition Digital Camera
by Jibo Yue, Haikuan Feng, Xiuliang Jin, Huanhuan Yuan, Zhenhai Li, Chengquan Zhou, Guijun Yang and Qingjiu Tian
Remote Sens. 2018, 10(7), 1138; https://doi.org/10.3390/rs10071138 - 18 Jul 2018
Cited by 143 | Viewed by 9925
Abstract
Timely and accurate estimates of crop parameters are crucial for agriculture management. Unmanned aerial vehicles (UAVs) carrying sophisticated cameras are very pertinent for this work because they can obtain remote-sensing images with higher temporal, spatial, and ground resolution than satellites. In this study, [...] Read more.
Timely and accurate estimates of crop parameters are crucial for agriculture management. Unmanned aerial vehicles (UAVs) carrying sophisticated cameras are very pertinent for this work because they can obtain remote-sensing images with higher temporal, spatial, and ground resolution than satellites. In this study, we evaluated (i) the performance of crop parameters estimates using a near-surface spectroscopy (350~2500 nm, 3 nm at 700 nm, 8.5 nm at 1400 nm, 6.5 nm at 2100 nm), a UAV-mounted snapshot hyperspectral sensor (450~950 nm, 8 nm at 532 nm) and a high-definition digital camera (Visible, R, G, B); (ii) the crop surface models (CSMs), RGB-based vegetation indices (VIs), hyperspectral-based VIs, and methods combined therefrom to make multi-temporal estimates of crop parameters and to map the parameters. The estimated leaf area index (LAI) and above-ground biomass (AGB) are obtained by using linear and exponential equations, random forest (RF) regression, and partial least squares regression (PLSR) to combine the UAV based spectral VIs and crop heights (from the CSMs). The results show that: (i) spectral VIs correlate strongly with LAI and AGB over single growing stages when crop height correlates positively with AGB over multiple growth stages; (ii) the correlation between the VIs multiplying crop height and AGB is greater than that between a single VI and crop height; (iii) the AGB estimate from the UAV-mounted snapshot hyperspectral sensor and high-definition digital camera is similar to the results from the ground spectrometer when using the combined methods (i.e., using VIs multiplying crop height, RF and PLSR to combine VIs and crop heights); and (iv) the spectral performance of the sensors is crucial in LAI estimates (the wheat LAI cannot be accurately estimated over multiple growing stages when using only crop height). The LAI estimates ranked from best to worst are ground spectrometer, UAV snapshot hyperspectral sensor, and UAV high-definition digital camera. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of study area and experimental design: (<b>a</b>) location of study area in China; (<b>b</b>) map showing Changping District in Beijing City; (<b>c</b>) design of treatments and images of ground-measurement field acquired from unmanned aerial vehicle mounted high-definition digital camera.</p>
Full article ">Figure 2
<p>UAV-UHD hyperspectral images and corresponding crop height on (<b>a</b>,<b>b</b>) 21 April; (<b>c</b>,<b>d</b>) 26 April; and (<b>e</b>,<b>f</b>) 13 May 2015.</p>
Full article ">Figure 3
<p>Digital camera images and corresponding crop height on (<b>a</b>,<b>b</b>) 21 April; (<b>c</b>,<b>d</b>) 26 April; and (<b>e</b>,<b>f</b>) 13 May 2015.</p>
Full article ">Figure 4
<p>Averaged hyperspectral reflectance spectra, averaged DN values, and crop height in the three growing stages: (<b>a</b>) G-hyperspectral; (<b>b</b>) UHD-hyperspectral; (<b>c</b>) calibrated DC-DN values; (<b>d</b>) G-height; (<b>e</b>) UHD-height; (<b>f</b>) DC height. G- indicates data measured by ground-based ASD spectrometer and measuring stick; UHD- indicates data measured using the UHD 185 mounted on the UAV; and DC- indicates data measured by using the digital camera mounted on the UAV.</p>
Full article ">Figure 5
<p>Pearson correlation coefficient between VIs, crop height (H), AGB and LAI: (<b>a</b>) G-Vis; (<b>b</b>) UHD-Vis; (<b>c</b>) DC-VIs.</p>
Full article ">Figure 6
<p>Relationship between best VIs and AGB: (<b>a</b>) G-LCI; (<b>b</b>) UHD-LCI; (<b>c</b>) DC-r; (<b>d</b>) G-height; (<b>e</b>) UHD-height; (<b>f</b>) DC-height; (<b>g</b>) G-height × LCI; (<b>h</b>) UHD-height × LCI; (<b>i</b>) DC-height × r.</p>
Full article ">Figure 7
<p>Relationship between best VIs and LAI: (<b>a</b>) G-LCI; (<b>b</b>) UHD-LCI; (<b>c</b>) DC-B; (<b>d</b>) G-height; (<b>e</b>) UHD-height; (<b>f</b>) DC-height; (<b>g</b>) G-height × LCI; (<b>h</b>) G-height × LCI; (<b>i</b>) DC-height × B.</p>
Full article ">Figure 8
<p>Relationship between the predicted and measured winter wheat AGB (t/ha): (<b>a</b>) G-height, VIs, and PLSR; (<b>b</b>) G-height, VIs, and RF; (<b>c</b>) G-VIs and PLSR; (<b>d</b>) G-VIs and RF; (<b>e</b>) UHD-height, VIs, and PLSR; (<b>f</b>) UHD-height, VIs, and RF; (<b>g</b>) UHD-VIs and PLSR; (<b>h</b>) UHD-VIs and RF; (<b>i</b>) DC-height, VIs, and PLSR; (<b>j</b>) DC-height, VIs, and RF; (<b>k</b>) DC-VIs and PLSR; (<b>l</b>) DC-VIs and RF (validation dataset, mean AGB = 4.58 t/ha).</p>
Full article ">Figure 9
<p>Relationship between predicted and measured winter wheat LAI (m<sup>2</sup>/m<sup>2</sup>): (<b>a</b>) G-height, VIs, and PLSR; (<b>b</b>) G-height, VIs, and RF; (<b>c</b>) G-VIs and PLSR; (<b>d</b>) G-VIs and RF; (<b>e</b>) UHD-height, VIs, and PLSR; (<b>f</b>) UHD-height, VIs, and RF; (<b>g</b>) UHD-VIs and PLSR; (<b>h</b>) UHD-VIs and RF; (<b>i</b>) DC-height, VIs, and PLSR; (<b>j</b>) DC-height, VIs, and RF; (<b>k</b>) DC-VIs and PLSR; (<b>l</b>) DC-VIs and RF (validation dataset, mean LAI = 3.57 m<sup>2</sup>/m<sup>2</sup>).</p>
Full article ">Figure 9 Cont.
<p>Relationship between predicted and measured winter wheat LAI (m<sup>2</sup>/m<sup>2</sup>): (<b>a</b>) G-height, VIs, and PLSR; (<b>b</b>) G-height, VIs, and RF; (<b>c</b>) G-VIs and PLSR; (<b>d</b>) G-VIs and RF; (<b>e</b>) UHD-height, VIs, and PLSR; (<b>f</b>) UHD-height, VIs, and RF; (<b>g</b>) UHD-VIs and PLSR; (<b>h</b>) UHD-VIs and RF; (<b>i</b>) DC-height, VIs, and PLSR; (<b>j</b>) DC-height, VIs, and RF; (<b>k</b>) DC-VIs and PLSR; (<b>l</b>) DC-VIs and RF (validation dataset, mean LAI = 3.57 m<sup>2</sup>/m<sup>2</sup>).</p>
Full article ">Figure 10
<p>Above-ground Biomass (t/ha) maps made using UAV-UHD, UAV-DC, and PLSR. (<b>a</b>) UHD on 21 April; (<b>b</b>) UHD on 26 April; (<b>c</b>) UHD on 13 May; (<b>d</b>) DC on 21 April; (<b>e</b>) DC on 26 April; and (<b>f</b>) DC on 13 May. Note: UHD- indicates data measured using the snapshot hyperspectral sensor mounted on the UAV; and DC- indicates data measured by using the digital camera mounted on the UAV.</p>
Full article ">Figure 10 Cont.
<p>Above-ground Biomass (t/ha) maps made using UAV-UHD, UAV-DC, and PLSR. (<b>a</b>) UHD on 21 April; (<b>b</b>) UHD on 26 April; (<b>c</b>) UHD on 13 May; (<b>d</b>) DC on 21 April; (<b>e</b>) DC on 26 April; and (<b>f</b>) DC on 13 May. Note: UHD- indicates data measured using the snapshot hyperspectral sensor mounted on the UAV; and DC- indicates data measured by using the digital camera mounted on the UAV.</p>
Full article ">Figure 11
<p>Leaf Area Index (m<sup>2</sup>/m<sup>2</sup>) maps based on UAV-UHD and UAV-DC images and with PLSR. (<b>a</b>) UHD on 21 April; (<b>b</b>) UHD on 26 April; (<b>c</b>) UHD on 13 May; (<b>d</b>) DC on 21 April; (<b>e</b>) DC on 26 April; (<b>f</b>) DC on 13 May. Note: UHD- indicates data measured using the snapshot hyperspectral sensor mounted on the UAV; and DC- indicates data measured by using the digital camera mounted on the UAV.</p>
Full article ">
30 pages, 9457 KiB  
Article
Imaging Multi-Age Construction Settlement Behaviour by Advanced SAR Interferometry
by Francesca Bozzano, Carlo Esposito, Paolo Mazzanti, Mauro Patti and Stefano Scancella
Remote Sens. 2018, 10(7), 1137; https://doi.org/10.3390/rs10071137 - 18 Jul 2018
Cited by 42 | Viewed by 6246
Abstract
This paper focuses on the application of Advanced Satellite Synthetic Aperture Radar Interferometry (A-DInSAR) to subsidence-related issues, with particular reference to ground settlements due to external loads. Beyond the stratigraphic setting and the geotechnical properties of the subsoil, other relevant boundary conditions strongly [...] Read more.
This paper focuses on the application of Advanced Satellite Synthetic Aperture Radar Interferometry (A-DInSAR) to subsidence-related issues, with particular reference to ground settlements due to external loads. Beyond the stratigraphic setting and the geotechnical properties of the subsoil, other relevant boundary conditions strongly influence the reliability of remotely sensed data for quantitative analyses and risk mitigation purposes. Because most of the Persistent Scatterer Interferometry (PSI) measurement points (Persistent Scatterers, PSs) lie on structures and infrastructures, the foundation type and the age of a construction are key factors for a proper interpretation of the time series of ground displacements. To exemplify a methodological approach to evaluate these issues, this paper refers to an analysis carried out in the coastal/deltaic plain west of Rome (Rome and Fiumicino municipalities) affected by subsidence and related damages to structures. This region is characterized by a complex geological setting (alternation of recent deposits with low and high compressibilities) and has been subjected to different urbanisation phases starting in the late 1800s, with a strong acceleration in the last few decades. The results of A-DInSAR analyses conducted from 1992 to 2015 have been interpreted in light of high-resolution geological/geotechnical models, the age of the construction, and the types of foundations of the buildings on which the PSs are located. Collection, interpretation, and processing of geo-thematic data were fundamental to obtain high-resolution models; change detection analyses of the land cover allowed us to classify structures/infrastructures in terms of the construction period. Additional information was collected to define the types of foundations, i.e., shallow versus deep foundations. As a result, we found that only by filtering and partitioning the A-DInSAR datasets on the basis of the above-mentioned boundary conditions can the related time series be considered a proxy of the consolidation process governing the subsidence related to external loads as confirmed by a comparison with results from a physically based back analysis based on Terzaghi’s theory. Therefore, if properly managed, the A-DInSAR data represents a powerful tool for capturing the evolutionary stage of the process for a single building and has potential for forecasting the behaviour of the terrain–foundation–structure combination. Full article
(This article belongs to the Special Issue Imaging Geodesy and Infrastructure Monitoring)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location of the study region over the map by Amenduni [<a href="#B25-remotesensing-10-01137" class="html-bibr">25</a>], in which the ancient Maccarese and Ostia ponds are highlighted (white dashed line identifies the Grande Raccordo Anulare, i.e., the ring-shaped motorways around Rome City Center; (<b>b</b>) Contour map of the unconformity at the base of the Tiber depositional sequence; (<b>c</b>) Stratigraphic cross-sections of the Tiber depositional sequence (see <a href="#remotesensing-10-01137-f001" class="html-fig">Figure 1</a>b for the location of the traces) from Milli et al. [<a href="#B39-remotesensing-10-01137" class="html-bibr">39</a>]; (<b>d</b>) Morphological features of the Tiber delta (redrawn from Giraudi [<a href="#B36-remotesensing-10-01137" class="html-bibr">36</a>]). LST, lowstand system tract; TST, transgressive system tract; HST, highstand system tract.</p>
Full article ">Figure 2
<p>Workflow of the proposed approach to assessing ground response to urbanisation by combining geological and geotechnical data with satellite interferometric synthetic aperture radar (InSAR) data. PS, Persistent Scatterer; 2D, two-dimensional; 3D, three-dimensional; GIS, geographical information system.</p>
Full article ">Figure 3
<p>Location of the collected boreholes (blue dots). The magenta polygon outlines the extent of the region analysed via ascending COSMO-SkyMed imagery; the green polygon represents the extent of the region analysed via ERS-ENVISAT data; the white polygons outline the eight selected sub-regions for the high-resolution analysis.</p>
Full article ">Figure 4
<p>Graph used to connect images related to the temporal baseline (X axis) and normal baseline (Y axis). Every dot represents an image, whereas every line represents an interferogram. Colours from blue to red show an increasing value of the spatial coherence.</p>
Full article ">Figure 5
<p>Multi-temporal view of a representative region showing the intense urbanisation in the time interval 1988–2011: (<b>a</b>) 1988, (<b>b</b>) 1998, (c) 2005, (<b>d</b>) 2011. The Da Vinci shopping mall, Commercity, and Rome Fair region are highlighted.</p>
Full article ">Figure 6
<p>A-DInSAR results plotted on the geological map by Amenduni [<a href="#B25-remotesensing-10-01137" class="html-bibr">25</a>]. (<b>a</b>) ERS ascending (1992–2000); (<b>b</b>) ERS descending (1993–2000); (<b>c</b>) ENVISAT ascending (2002–2010); (<b>d</b>) ENVISAT descending (2003–2010); (<b>e</b>) COSMO-SkyMed ascending (2011–2015); (<b>f</b>) Location of the Ostia and Maccarese ponds before reclamation, regions analysed in detail in this work (white polygons), and airstrip 3 of the Leonardo Da Vinci airport.</p>
Full article ">Figure 7
<p>Geotechnical 3D models of (<b>a</b>) sector 1 (Rome Fair region and Commercity, inner delta) and (<b>b</b>) sector 2 (offices of the Leonardo Da Vinci airport, outer delta). The figure also shows the processing chain for the 3D geotechnical model, which started from the collection and interpretation of borehole data (first images on the top), followed by their correlation as fence diagrams, and finally interpolated as continuous surfaces (last images at the bottom).</p>
Full article ">Figure 8
<p>A detail of sector 1. GU2 + GU3 thickness map and geotechnical cross sections (with a strong vertical exaggeration) of the Rome Fair and Commercity region.</p>
Full article ">Figure 9
<p>Velocities of persistent scatterers for the inner (<b>left column</b>; sector 1 in <a href="#remotesensing-10-01137-f006" class="html-fig">Figure 6</a>f) and outer delta (<b>right column</b>, sector 2 in <a href="#remotesensing-10-01137-f006" class="html-fig">Figure 6</a>f) regions.</p>
Full article ">Figure 10
<p>Classification of the building ages in the eight investigated regions (<b>a</b>); and zoom on sector 1 (<b>b</b>); and sector 2 (<b>c</b>) discussed in detail in this paper. For these sectors, the foundation type of the buildings is also reported (<b>d</b>, <b>e</b>).</p>
Full article ">Figure 11
<p>Ground deformation of airstrip 3 of the Leonardo Da Vinci airport analysed by ERS ascending (<b>a</b>); ERS descending (<b>b</b>); ENVISAT ascending (<b>c</b>); ENVISAT descending (<b>d</b>); and COSMO-SkyMed (<b>e</b>). In (<b>f</b>) are reported the cumulative time series of displacement of the three main sectors detected at the airstrip. In (<b>g</b>) a geological cross section of the airstrip is shown (from Manassero and Dominijanni [<a href="#B46-remotesensing-10-01137" class="html-bibr">46</a>], redrawn). The location of this region is reported in <a href="#remotesensing-10-01137-f006" class="html-fig">Figure 6</a>f.</p>
Full article ">Figure 12
<p>(<b>a</b>) Velocity of persistent scatterers of COSMO-SkyMed over a thickness map of compressible soils; the white rectangle highlights the presence of buildings on deep foundations. Most of the red dots within the rectangle refer to PSs located directly on the ground (i.e., paving) and not connected to the buildings; (<b>b</b>) Correlation diagram between PS velocity and soft-soil thickness in the region regardless of the characteristics (age of construction and foundation type) of the structure; (<b>c</b>) Geotechnical cross section (see <b>a</b> for the location) with a schematic positioning of the persistent scatterers localized over the cross section (the colour of the dots refers to the colour scale of <a href="#remotesensing-10-01137-f012" class="html-fig">Figure 12</a>a).</p>
Full article ">Figure 13
<p>View of the Da Vinci shopping centre and the Roma-Fiumicino highway. The foundation types of the most relevant structures and infrastructures and the displacement velocity measured on PSs with different SAR datasets are reported. The white star represents the location of the 22-year cumulative time series reported at the bottom right in which the decrease in settlement rates over time is evident and in accordance with the typical consolidation curve (bottom left).</p>
Full article ">Figure 14
<p>(<b>a</b>) Velocity of persistent scatterers of COSMO-SkyMed over a thickness map of compressible soils; the classified polygons refer to the date of construction; (<b>b</b>, <b>c</b>) Correlation diagrams between the PS velocity and the soft-soil thickness in the region for structures built between 1994–1998 and 1998–2002, respectively.</p>
Full article ">Figure 15
<p>Analysis of the velocity of persistent scatterers on single buildings. (<b>a</b>) Classified plot of persistent scatterers over a thickness map of compressible soils. (<b>b</b>, <b>c</b>) Correlation diagrams between PS velocity and soft-soil thickness along cross sections (see <b>a</b> for the location) referring to two single buildings.</p>
Full article ">Figure 16
<p>(<b>a</b>) Location of the detailed view (reported in (<b>b</b>) on the north-eastern sector of the deltaic plain; (<b>b</b>) Location of the main farmhouses over a 1:25.000 topographic map (I.G.M). The red polygons in (<b>b</b>) indicate the location of the detailed views of the PS velocity map reported in (<b>c</b>, <b>d</b>), in which the farmhouses are clearly visible.</p>
Full article ">Figure 17
<p>Comparison of theoretical settlements values (obtained by applying the one-dimensional (1D) consolidation theory to known shallow-founded buildings and pavements of known age) with the corresponding COSMO-SkyMed dataset for the period 2011–2015. Blue symbols indicate LOS displacements, whereas orange symbols represent the corresponding results converted to the vertical direction. The slopes of the trend line, quite close to the bisector, show the agreement between the physically based estimations and the actual measurements via Persistent Scatterer Interferometry (PSI).</p>
Full article ">Figure 18
<p>Age interval and time series of displacement of representative persistent scatterers belonging to structures used to check the comparability with the theoretical estimation of settlement. Blue dots represent the localisation of the boreholes used in this phase.</p>
Full article ">
22 pages, 6369 KiB  
Article
Spatial and Temporal Dependency of NDVI Satellite Imagery in Predicting Bird Diversity over France
by Sébastien Bonthoux, Solenne Lefèvre, Pierre-Alexis Herrault and David Sheeren
Remote Sens. 2018, 10(7), 1136; https://doi.org/10.3390/rs10071136 - 18 Jul 2018
Cited by 23 | Viewed by 7228
Abstract
Continuous-based predictors of habitat characteristics derived from satellite imagery are increasingly used in species distribution models (SDM). This is especially the case of Normalized Difference Vegetation Index (NDVI) which provides estimates of vegetation productivity and heterogeneity. However, when NDVI predictors are incorporated into [...] Read more.
Continuous-based predictors of habitat characteristics derived from satellite imagery are increasingly used in species distribution models (SDM). This is especially the case of Normalized Difference Vegetation Index (NDVI) which provides estimates of vegetation productivity and heterogeneity. However, when NDVI predictors are incorporated into SDM, synchrony between biological observations and image acquisition must be questionned. Due to seasonal variations of NDVI during the year, landscape patterns of habitats are revealed differently from one date to another leading to variations in models’ performance. In this paper, we investigated the influence of acquisition time period of NDVI to explain and predict bird community patterns over France. We examined if the NDVI acquisition period that best fit the bird data depends on the dominant land cover context. We also compared models based on single time period of NDVI with one model built from the Dynamic Habitat Index (DHI) components which summarize variations in vegetation phenology throughout the year from the fraction of radiation absorbed by the canopy (fPAR). Bird species richness was calculated as response variable for 759 plots of 4 km2 from the French Breeding Bird Survey. Bird specialists and generalists to habitat were considered. NDVI and DHI predictors were both derived from MODIS products. For NDVI, five time periods in 2010 were compared, from late winter to begin of autumn. A climate predictor was also used and Generalized Additive Models were fitted to explain and predict bird species richness. Results showed that NDVI-based proxies of dominant habitat identity and spatial heterogeneity explain more bird community patterns than DHI-based proxies of annual productivity and seasonnality. We also found that models’ performance was both time and context-dependent, varying according to the bird groups. In general, best time period of NDVI did not match with the acquisition period of bird data because in case of synchrony, differences in habitats are less pronounced. These findings suggest that the most powerful approach to estimate bird community patterns is the simplest one. It only requires NDVI predictors from a single appropriate time period, in addition to climate, which makes the approach very operational. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Distribution of the selected bird plot squares over France for the year 2010. Each plot of 2 km side contains 10 point counts. Plots included in the analyses (<math display="inline"> <semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>759</mn> </mrow> </semantics> </math>) were filtered according to the quality of the satellite imagery.</p>
Full article ">Figure 2
<p>Image data used for the study: five NDVI 16-day-composit products at different time periods in 2010 (250-m spatial resolution) and one map of climate regions based on [<a href="#B31-remotesensing-10-01136" class="html-bibr">31</a>].</p>
Full article ">Figure 3
<p>The components of the Dynamic Habitat Index of 2010 over France derived from the MODIS fPAR 8-day composite product at 500-m spatial resolution: (<b>a</b>) cumulative productivity; (<b>b</b>) minimum productivity; (<b>c</b>) intra-annual variation of productivity expressed as the coefficient of variation; (<b>d</b>) the composite DHI. The components were normalized from 0 to 1 for visualization.</p>
Full article ">Figure 4
<p>Relationship between species richness (SR, <span class="html-italic">y</span>-axis) of woodland birds (<b>left</b>), farmland birds (<b>center</b>), and total richness (<b>right</b>) with the <span class="html-italic">average</span> NDVI variable (<span class="html-italic">x</span>-axis) at the best time period. NDVI values (250-m product) were rescaled with a multiplicative factor of 10,000.</p>
Full article ">Figure 5
<p>Predictive performances (i.e., Spearman’s rank correlations <span class="html-italic">Rho</span> between observed and predicted values) of the bird-habitat models based on climate factor only, climate with NDVI variables for each time period, climate with ensemble NDVI (consensus approach), or climate with DHI. The models are based on 759 bird plots using GAM and NDVI at 250-m spatial resolution. The response variables are species richness (SR) for four groups of birds (woodland, farmland, generalist and all the species). The predictive performances were calculated by 3-fold cross-validation with 100 repetitions. The average values of <span class="html-italic">Rho</span> are provided above each boxplot. Different letters indicate that <span class="html-italic">Rho</span> values are significantly different between datasets.</p>
Full article ">Figure 6
<p>Prediction maps of species richness for farmland (<b>a</b>) and woodland (<b>b</b>) birds at French national scale. Maps were produced using the best GAM models (i.e., climate with NDVI 26-Jun for SR farmland and climate with NDVI 14-Sep for SR woodland at 250-m spatial resolution) using all the 759 bird plots. Spatial concordance is observed between predictions and maps related to cultivated (<b>c</b>) and forested (<b>d</b>) areas. These maps (<b>c</b>,<b>d</b>) were derived from the French land cover map of 2010 (OSO product) [<a href="#B47-remotesensing-10-01136" class="html-bibr">47</a>].</p>
Full article ">Figure 7
<p>Comparison of performances (i.e., Spearman’s rank correlations <span class="html-italic">Rho</span> between observed and predicted values) in predicting species richness of woodland and farmland birds according to different dominant land cover in the bird plots. The bird-habitat models were based on climate only, climate with 250-m NDVI variables of each time period, and climate with ensemble 250-m NDVI (consensus approach). The predictive performances were calculated by 3-fold cross-validation with 100 repetitions. The average values of <span class="html-italic">Rho</span> is provided above each boxplot. Different letters indicate that <span class="html-italic">Rho</span> values are significantly different between datasets.</p>
Full article ">Figure 7 Cont.
<p>Comparison of performances (i.e., Spearman’s rank correlations <span class="html-italic">Rho</span> between observed and predicted values) in predicting species richness of woodland and farmland birds according to different dominant land cover in the bird plots. The bird-habitat models were based on climate only, climate with 250-m NDVI variables of each time period, and climate with ensemble 250-m NDVI (consensus approach). The predictive performances were calculated by 3-fold cross-validation with 100 repetitions. The average values of <span class="html-italic">Rho</span> is provided above each boxplot. Different letters indicate that <span class="html-italic">Rho</span> values are significantly different between datasets.</p>
Full article ">Figure 8
<p>Seasonal patterns of MODIS NDVI 16-day-composit products in 2010 (250-m spatial resolution) according to the landscape context of bird plot.</p>
Full article ">Figure A1
<p>Non parametric spline correlograms, with 95% pointwise bootstrap confidence intervals of the Pearson residuals from GAMs, based on the best model of each bird group.</p>
Full article ">Figure A2
<p>NDVI temporal profiles (average value ± standard deviation) for bird plots dominated by forests, grasslands, winter crops and summer crops.</p>
Full article ">
21 pages, 12337 KiB  
Article
Improved Fully Convolutional Network with Conditional Random Fields for Building Extraction
by Sanjeevan Shrestha and Leonardo Vanneschi
Remote Sens. 2018, 10(7), 1135; https://doi.org/10.3390/rs10071135 - 18 Jul 2018
Cited by 104 | Viewed by 9013
Abstract
Building extraction from remotely sensed imagery plays an important role in urban planning, disaster management, navigation, updating geographic databases, and several other geospatial applications. Several published contributions dedicated to the applications of deep convolutional neural networks (DCNN) for building extraction using aerial/satellite imagery [...] Read more.
Building extraction from remotely sensed imagery plays an important role in urban planning, disaster management, navigation, updating geographic databases, and several other geospatial applications. Several published contributions dedicated to the applications of deep convolutional neural networks (DCNN) for building extraction using aerial/satellite imagery exists. However, in all these contributions, high accuracy is always obtained at the price of extremely complex and large network architectures. In this paper, we present an enhanced fully convolutional network (FCN) framework that is designed for building extraction of remotely sensed images by applying conditional random fields (CRFs). The main objective is to propose a methodology selecting a framework that balances high accuracy with low network complexity. A modern activation function, namely, the exponential linear unit (ELU), is applied to improve the performance of the fully convolutional network (FCN), thereby resulting in more accurate building prediction. To further reduce the noise (falsely classified buildings) and to sharpen the boundaries of the buildings, a post-processing conditional random fields (CRFs) is added at the end of the adopted convolutional neural network (CNN) framework. The experiments were conducted on Massachusetts building aerial imagery. The results show that our proposed framework outperformed the fully convolutional network (FCN), which is the existing baseline framework for semantic segmentation, in terms of performance measures such as the F1-score and IoU measure. Additionally, the proposed method outperformed a pre-existing classifier for building extraction using the same dataset in terms of the performance measures and network complexity. Full article
(This article belongs to the Special Issue Recent Advances in Neural Networks for Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>General pipelines of our proposed approach: the training stage and the classification stage.</p>
Full article ">Figure 2
<p>Visualization of the VGG-FCN architecture. The figure depicts the skip connection architecture that was devised in [<a href="#B29-remotesensing-10-01135" class="html-bibr">29</a>]. Only the pooling and prediction layers are shown, omitting the intermediate convolutional layer. The image shows the FCN-32s variant (without skip connections) on top, the FCN-16s variant in the middle and FCN-8s variant at the bottom.</p>
Full article ">Figure 3
<p>Two sample aerial images from the Massachusetts building dataset; each row contains an original image on the left, which acts as a ground truth of the corresponding image on the right: (<b>a</b>) aerial image, (<b>b</b>) building mask.</p>
Full article ">Figure 4
<p>Iteration plot on the Massachusetts satellite data sets of variations of the proposed methods: fully convolutional network (FCN) and ELU-FCN. The x-axis corresponds to the number of iterations, and the y-axis refers to the measure. Each row refers to a different model. (<b>a</b>) plot of model loss (cross-entropy) on the training and validation datasets for FCN ; (<b>b</b>) plot of accuracy on the training and validation datasets for FCN; (<b>c</b>) plot of model loss on the training and validation datasets for ELU-FCN; and (<b>d</b>) plot of accuracy on the training and validation datasets for ELU-FCN.</p>
Full article ">Figure 4 Cont.
<p>Iteration plot on the Massachusetts satellite data sets of variations of the proposed methods: fully convolutional network (FCN) and ELU-FCN. The x-axis corresponds to the number of iterations, and the y-axis refers to the measure. Each row refers to a different model. (<b>a</b>) plot of model loss (cross-entropy) on the training and validation datasets for FCN ; (<b>b</b>) plot of accuracy on the training and validation datasets for FCN; (<b>c</b>) plot of model loss on the training and validation datasets for ELU-FCN; and (<b>d</b>) plot of accuracy on the training and validation datasets for ELU-FCN.</p>
Full article ">Figure 5
<p>Visual comparison of three variations of the proposed techniques using sample aerial test images of the Massachusetts area. (<b>a</b>) original input image; (<b>b</b>) ground-truth map; (<b>c</b>) output of FCN-8s; (<b>d</b>) output of ELU-FCN; and (<b>e</b>) output of ELU-FCN-CRFs.</p>
Full article ">Figure 6
<p>Visual comparison of three variations of the proposed techniques on large buildings using sample aerial test images of the Massachusetts area. (<b>a</b>) original input image; (<b>b</b>) ground-truth map; (<b>c</b>) output of FCN-8s; (<b>d</b>) output of ELU-FCN; and (<b>e</b>) output of ELU-FCN-CRF.</p>
Full article ">Figure 7
<p>Visualization of predictions of three variations of the proposed model on building detection tasks with original extracted test images from the Massachusetts dataset. (<b>a</b>) input images; (<b>b</b>) results of base FCN network; (<b>c</b>) results of ELU-FCN; and (<b>d</b>) results of ELU-FCN-CRFs. Green pixels are TP, red pixels are FN, blue pixels are FP, and background pixels are TN.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop