[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Issue
Volume 10, May
Previous Issue
Volume 10, March
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 4 (April 2018) – 170 articles

Cover Story (view full-size image): Ocean surface currents and winds are closely coupled essential climate variables and should be observed simultaneously to understand air–sea interactions. Under NASA’s Instrument Incubator Program (IIP), we have built a wide-swath Doppler scatterometer, DopplerScatt, intended to serve as an airborne prototype for a future wind and current spaceborne missions. The cover shows data collected at the outflow of the Mississippi River into Barataria Bay, where the river releases significant sediment, as shown in the lower-left Sentinel-3 image. The lower right image shows the DopplerScatt estimated neutral winds, which are noticeably modified by the currents. The upper panels show the east (left) and north (right) surface current components. In addition to the plume recirculation into Barataria Bay, one can observe a strong submesoscale front coinciding with a front in sediment concentration. View this paper.
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
18 pages, 87543 KiB  
Article
A Novel Affine and Contrast Invariant Descriptor for Infrared and Visible Image Registration
by Xiangzeng Liu, Yunfeng Ai, Juli Zhang and Zhuping Wang
Remote Sens. 2018, 10(4), 658; https://doi.org/10.3390/rs10040658 - 23 Apr 2018
Cited by 37 | Viewed by 6384
Abstract
Infrared and visible image registration is a very challenging task due to the large geometric changes and the significant contrast differences caused by the inconsistent capture conditions. To address this problem, this paper proposes a novel affine and contrast invariant descriptor called maximally [...] Read more.
Infrared and visible image registration is a very challenging task due to the large geometric changes and the significant contrast differences caused by the inconsistent capture conditions. To address this problem, this paper proposes a novel affine and contrast invariant descriptor called maximally stable phase congruency (MSPC), which integrates the affine invariant region extraction with the structural features of images organically. First, to achieve the contrast invariance and ensure the significance of features, we detect feature points using moment ranking analysis and extract structural features via merging phase congruency images in multiple orientations. Then, coarse neighborhoods centered on the feature points are obtained based on Log-Gabor filter responses over scales and orientations. Subsequently, the affine invariant regions of feature points are determined by using maximally stable extremal regions. Finally, structural descriptors are constructed from those regions and the registration can be implemented according to the correspondence of the descriptors. The proposed method has been tested on various infrared and visible pairs acquired by different platforms. Experimental results demonstrate that our method outperforms several state-of-the-art methods in terms of robustness and precision with different image data and also show its effectiveness in the application of trajectory tracking. Full article
(This article belongs to the Special Issue Learning to Understand Remote Sensing Images)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Differences of contrast and viewpoints in input images. (<b>a</b>) Infrared image; (<b>b</b>) Corresponding regions and their gradient images; and (<b>c</b>) Visible image.</p>
Full article ">Figure 2
<p>Illustration of registration by using the proposed method.</p>
Full article ">Figure 3
<p>Feature points detection by the method of salient feature points extraction (MSFPE).</p>
Full article ">Figure 4
<p>Structural features extraction using multi-orientation phase congruency.</p>
Full article ">Figure 5
<p>The construction of the maximally stable phase congruency (MSPC) descriptor from input images. (<b>a</b>) Original patches around the feature points; (<b>b</b>) Rectangle regions from structural features image (SFI) according to the scales and orientations of the feature points; (<b>c</b>) Fine ellipse regions detected by maximally stable extremal regions (MSER) based on the rectangle regions; (<b>d</b>) Normalized circle regions relate to the ellipse regions; (<b>e</b>) MSPC descriptors constructed in the circle regions.</p>
Full article ">Figure 6
<p>Flow chart of the proposed registration.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>d</b>) are different infrared and visible image pairs from CVC datasets.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>–<b>d</b>) are different infrared and visible image pairs from CVC datasets.</p>
Full article ">Figure 8
<p>Matching results using the proposed method for <a href="#remotesensing-10-00658-f007" class="html-fig">Figure 7</a>. (<b>a</b>–<b>d</b>) are the matching results of the <a href="#remotesensing-10-00658-f007" class="html-fig">Figure 7</a>a–d respectively.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>f</b>) are the samples of image pairs captured from electro-optical pod (EOP) on UAV.</p>
Full article ">Figure 10
<p>Matching results by the proposed method for <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a>. (<b>a</b>–<b>f</b>) are the matching results of the <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a>a–f respectively.</p>
Full article ">Figure 10 Cont.
<p>Matching results by the proposed method for <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a>. (<b>a</b>–<b>f</b>) are the matching results of the <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a>a–f respectively.</p>
Full article ">Figure 11
<p>Comparison of matching performance by the related methods. (<b>a</b>) is the matching precision for the six image pairs in <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a> by the related methods; (<b>b</b>) is repeatability for the six image pairs in <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a> by the related methods.</p>
Full article ">Figure 12
<p>Registration results by the proposed method for <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a>. (<b>a</b>–<b>f</b>) are the registration results of the proposed method for <a href="#remotesensing-10-00658-f009" class="html-fig">Figure 9</a>a–f respectively.</p>
Full article ">Figure 13
<p>Reference image download from Google.</p>
Full article ">Figure 14
<p>Samples of the sub-images from the real-time images.</p>
Full article ">Figure 15
<p>Several registration results of the samples in <a href="#remotesensing-10-00658-f014" class="html-fig">Figure 14</a> and the sub-regions of the reference image in <a href="#remotesensing-10-00658-f013" class="html-fig">Figure 13</a>.</p>
Full article ">Figure 16
<p>UAV trajectory tracking results of our registration method.</p>
Full article ">
15 pages, 5082 KiB  
Article
Using Satellite Altimetry to Calibrate the Simulation of Typhoon Seth Storm Surge off Southeast China
by Xiaohui Li, Guoqi Han, Jingsong Yang, Dake Chen, Gang Zheng and Nan Chen
Remote Sens. 2018, 10(4), 657; https://doi.org/10.3390/rs10040657 - 23 Apr 2018
Cited by 20 | Viewed by 6493
Abstract
Satellite altimeters can capture storm surges generated by typhoons and tropical storms, if the satellite flies over at the right time. In this study, we show TOPEX/Poseidon altimeter-observed storm surge features off Southeast China on 10 October 1994 during Typhoon Seth. We then [...] Read more.
Satellite altimeters can capture storm surges generated by typhoons and tropical storms, if the satellite flies over at the right time. In this study, we show TOPEX/Poseidon altimeter-observed storm surge features off Southeast China on 10 October 1994 during Typhoon Seth. We then use a three-dimensional, barotropic, finite-volume community ocean model (FVCOM) to simulate storm surges. An innovative aspect is that satellite data are used to calibrate the storm surge model to improve model performance, by adjusting model wind forcing fields (the National Center for Environment Prediction (NCEP) reanalysis product) in reference to the typhoon best-track data. The calibration reduces the along-track root-mean-square (RMS) difference between model and altimetric data from 0.15 to 0.10 m. It also reduces the RMS temporal difference from 0.21 to 0.18 m between the model results and independent tide-gauge data at Xiamen. In particular, the calibrated model produces a peak storm surge of 1.01 m at 6:00 10 October 1994 at Xiamen, agreeing with tide-gauge data; while the peak storm surge with the NCEP forcing is 0.71 m only. We further show that the interaction between storm surges and astronomical tides contributes to the peak storm surge by 34% and that the storm surge propagates southwestward as a coastally-trapped Kelvin wave. Full article
(This article belongs to the Special Issue Satellite Altimetry for Earth Sciences)
Show Figures

Figure 1

Figure 1
<p>Map showing the study area off Southeast China with bathymetric contours in meters. Typhoon Seth’s track and locations at specific times are shown as blue lines and red dots. The red line is the TOPEX/Poseidon (T/P) satellite ground track. Pingtan (PT), Xiamen (XM) and Dongshan (DS) tide stations are marked by circles. The Dachen (DC) weather station is also depicted (red triangle).</p>
Full article ">Figure 2
<p>Mesh with 29,235 nodes in the waters off the eastern and southern coast of China. The fine grid resolution is around 2 km along the Taiwan Strait coast.</p>
Full article ">Figure 3
<p>M2 co-tidal and co-phase charts from FVCOM. The red and black lines show the phase lag (in degrees relative to Beijing local time (UT + 8 h)) and amplitude (in centimeters), respectively.</p>
Full article ">Figure 4
<p>Sea level anomalies comparison between observations and simulations at Xiamen.</p>
Full article ">Figure 5
<p>Altimetric sea surface height anomalies along Track 88 observed by T/P. Blue: Cycle 76 during Typhoon Seth. Red: Cycle 75. Black: Cycle 77. CTOH, Centre for Topographic studies of the Ocean and Hydrosphere. SLA, Sea Level Anomalies.</p>
Full article ">Figure 6
<p>De-tided altimetric and model surface height anomalies for Cycle 76 during Hurricane Seth, relative to those averaged for Cycle 75 and 77.</p>
Full article ">Figure 7
<p>Comparison between the modified and NCEP wind field at 06:00, 10 October.</p>
Full article ">Figure 8
<p>Non-tidal sea surface height anomalies at Xiamen from tide-gauge observations, from the model run forced by the National Center for Atmospheric (NCAR) wind modified by 1.3-times and from the model run forced by both the tide and the NACR wind modified by 1.3-times.</p>
Full article ">Figure 9
<p>Sea surface height anomalies due to the tide-surge nonlinear interactions and tidal height at (<b>a</b>) XM and (<b>b</b>) PT. The model is forced by both the tide and the NCEP wind modified by 1.3-times.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>f</b>) show temporal change of the model non-tidal sea surface height anomalies and the forcing wind fields at 03:00, 04:00, 05:00, 06:00, 07:00 and 08:00 on 10 October. The model run is forced by both the tide and the NCEP wind modified by 1.3-times.</p>
Full article ">Figure 11
<p>The lagged correlation coefficients of the model non-tidal sea surface height anomalies between PT and XM (blue) and between DS and XM (green). Negative values in the time lag mean the former station leading the latter for each pair. The model run is forced by both the tide and the NCEP wind modified by 1.3-times.</p>
Full article ">Figure 12
<p>(<b>a</b>) The dispersion relationship estimated for the first-mode continental shelf wave in the coastal area nearby the T/P ground track; (<b>b</b>) the corresponding wave speed curve.</p>
Full article ">
21 pages, 6399 KiB  
Article
Measurements on the Absolute 2-D and 3-D Localization Accuracy of TerraSAR-X
by Ulrich Balss, Christoph Gisinger and Michael Eineder
Remote Sens. 2018, 10(4), 656; https://doi.org/10.3390/rs10040656 - 23 Apr 2018
Cited by 35 | Viewed by 5740
Abstract
The German TerraSAR-X radar satellites TSX-1 and TDX-1 are well-regarded for their unprecedented geolocation accuracy. However, to access their full potential, Synthetic Aperture Radar (SAR)-based location measurements have to be carefully corrected for effects that are well-known in the area of geodesy but [...] Read more.
The German TerraSAR-X radar satellites TSX-1 and TDX-1 are well-regarded for their unprecedented geolocation accuracy. However, to access their full potential, Synthetic Aperture Radar (SAR)-based location measurements have to be carefully corrected for effects that are well-known in the area of geodesy but were previously often neglected in the area of SAR, such as wave propagation and Earth dynamics. Our measurements indicate that in this way, when SAR is handled as a geodetic measurement instrument, absolute localization accuracy at better than centimeter level with respect to a given geodetic reference frame is obtained in 2-D and, when using stereo SAR techniques, also in 3-D. The TerraSAR-X measurement results presented in this study are based on a network of three globally distributed geodetic observatories. Each is equipped with one or two trihedral corner reflectors with accurately (<5 mm) known reference coordinates, used as a reference for the verification of the SAR measured coordinates. Because these observatories are located in distant parts of the world, they give us evidence on the worldwide reproducibility of the obtained results. In this paper we report the achieved results of measurements performed over 6 1/2 years (from July 2011 to January 2018) and refer to some first new application areas for geodetic SAR. Full article
(This article belongs to the Special Issue Ten Years of TerraSAR-X—Scientific Results)
Show Figures

Figure 1

Figure 1
<p>Schematic view of measurement arrangement and procedures.</p>
Full article ">Figure 2
<p>Corner reflectors at our test sites: (<b>a</b>) Wettzell, Germany; (<b>b</b>) GARS O’Higgins, Antarctic Peninsula; (<b>c</b>) Metsähovi, Finland.</p>
Full article ">Figure 3
<p>Temporal progression of the azimuth (blue) and range (red) offset obtained at the Metsähovi test site.</p>
Full article ">Figure 4
<p>Temporal progression of the azimuth (blue) and range (red) offset obtained at the Wettzell test site: (<b>a</b>) Wettzell Ascending measurement series; (<b>b</b>) Wettzell Descending.</p>
Full article ">Figure 5
<p>Temporal progression of the azimuth (blue) and range (red) offset obtained at the GARS O’Higgins test site: (<b>a</b>) GARS O’Higgins Ascending measurement series; (<b>b</b>) GARS O’Higgins Descending.</p>
Full article ">Figure 6
<p>Gradients of the interpolated linear trends at the different test sites: (<b>a</b>) range; (<b>b</b>) azimuth. The plotted error bars represent a 95% confidence interval (2σ).</p>
Full article ">Figure 7
<p>Scatter plots of the obtained azimuth and range offsets: (<b>a</b>) Wettzell Ascending; (<b>b</b>) Wettzell Descending; (<b>c</b>) GARS O’Higgins Ascending; (<b>d</b>) GARS O’Higgins Descending; (<b>e</b>) Metsähovi Descending.</p>
Full article ">Figure 8
<p>Close-up of the temporal progression of the obtained azimuth (<b>a</b>) and range (<b>b</b>) offset at the Wettzell test site during the ST300 campaign (yellow) interleaving regular HS300 acquisitions (green).</p>
Full article ">Figure 9
<p>Differences of the TerraSAR-X stereo coordinates to the reference coordinates, and the error ellipsoid of the SAR solution scaled to 95% confidence level: Horizontal cross section in local east/height (<b>a</b>); vertical cross section in local north/east (<b>b</b>). WTZ = Wettzell, OHI = GARS O’Higgins and MET = Metsähovi.</p>
Full article ">Figure 10
<p>Comparison of the mean values (<b>a</b>) and standard deviations (<b>b</b>) in azimuth and range for operational Science orbits (values of <a href="#remotesensing-10-00656-t006" class="html-table">Table 6</a>) and new experimental orbits (values of <a href="#remotesensing-10-00656-t009" class="html-table">Table 9</a>). The different colors represent the different measurement series.</p>
Full article ">Figure 10 Cont.
<p>Comparison of the mean values (<b>a</b>) and standard deviations (<b>b</b>) in azimuth and range for operational Science orbits (values of <a href="#remotesensing-10-00656-t006" class="html-table">Table 6</a>) and new experimental orbits (values of <a href="#remotesensing-10-00656-t009" class="html-table">Table 9</a>). The different colors represent the different measurement series.</p>
Full article ">Figure 11
<p>Comparison of the biases (<b>a</b>) and 95% confidence intervals (<b>b</b>) in north, east and height for operational Science orbits (values of <a href="#remotesensing-10-00656-t007" class="html-table">Table 7</a>) and new experimental orbits (values of <a href="#remotesensing-10-00656-t010" class="html-table">Table 10</a>). The different colors represent the different measurement series.</p>
Full article ">
18 pages, 4321 KiB  
Article
Target Reconstruction Based on Attributed Scattering Centers with Application to Robust SAR ATR
by Jihong Fan and Andrew Tomas
Remote Sens. 2018, 10(4), 655; https://doi.org/10.3390/rs10040655 - 23 Apr 2018
Cited by 16 | Viewed by 4666
Abstract
This paper proposes a synthetic aperture radar (SAR) automatic target recognition (ATR) method by target reconstruction based on attributed scattering centers (ASCs). The extracted ASCs can effectively describe the electromagnetic scattering characteristics of the target, while eliminating the background clutters and noises. Therefore, [...] Read more.
This paper proposes a synthetic aperture radar (SAR) automatic target recognition (ATR) method by target reconstruction based on attributed scattering centers (ASCs). The extracted ASCs can effectively describe the electromagnetic scattering characteristics of the target, while eliminating the background clutters and noises. Therefore, the ASCs are discriminative features for SAR ATR. The neighbor matching algorithm was used to build the correspondence between the test ASC set and corresponding template ASC set. Afterwards, the selected template ASCs were used to reconstruct the template image, whereas all the test ASCs were used to reconstruct the test image based on the ASC model. A similarity measure was further designed based on the reconstructed images for target recognition. Compared with traditional ASC matching methods, the complex one-to-one correspondence between two ASC sets was avoided. Moreover, all the attributes of the ASCs were utilized during the target reconstruction. Therefore, the proposed method can better exploit the discriminability of ASCs to improve the ATR performance. To evaluate the effectiveness and robustness of the proposed method, extensive experiments on the moving and stationary target acquisition and recognition (MSTAR) dataset were conducted under both the standard operating condition (SOC) and typical extended operating conditions (EOCs). Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Technical flowchart of the proposed method.</p>
Full article ">Figure 2
<p>Reconstruction using extracted attributed scattering centers (ASCs) (<b>a</b>) original image (<b>b</b>) reconstructed image.</p>
Full article ">Figure 3
<p>The neighbor matching results between the test and template ASC sets from BMP2 at the radius of 0.3 m.</p>
Full article ">Figure 4
<p>The neighbor matching results between the test and template ASC sets from BMP at the radius of (<b>a</b>) 0.4 m; (<b>b</b>) 0.5 m.</p>
Full article ">Figure 5
<p>The neighbor matching results between the BMP2 test ASC set and the template ASC sets from other targets: (<b>a</b>) T72; (<b>b</b>) BTR70.</p>
Full article ">Figure 6
<p>The reconstructed template images based on the matching results from different targets: (<b>a</b>) BMP2; (<b>b</b>) T72; (<b>c</b>) BTR70.</p>
Full article ">Figure 7
<p>Optical images of the 10 military targets. (<b>a</b>) BMP2; (<b>b</b>) BTR70; (<b>c</b>) T72; (<b>d</b>) T62; (<b>e</b>) BRDM2; (<b>f</b>) BTR60; (<b>g</b>) ZSU23/4; (<b>h</b>) D7; (<b>i</b>) ZIL131; (<b>j</b>) 2S1.</p>
Full article ">Figure 8
<p>Optical and synthetic aperture radar (SAR) images of four configurations of T72 tank. (<b>a</b>) Optical images (<b>b</b>) SAR images.</p>
Full article ">Figure 9
<p>SAR images of 2S1 at different depression angles of (<b>a</b>) 17°; (<b>b</b>) 30°; (<b>c</b>) 45°.</p>
Full article ">Figure 10
<p>Noisy images with different levels of additive white Gaussian noises (AWGN) addition. (<b>a</b>) Original image, (<b>b</b>) 10 dB, (<b>c</b>) 5 dB, (<b>d</b>) 0 dB, (<b>e</b>) −5 dB, (<b>f</b>) −10 dB.</p>
Full article ">Figure 11
<p>Performance of different methods under noise corruption.</p>
Full article ">Figure 12
<p>Twenty percent occluded images from different directions. (<b>a</b>) Original image, (<b>b</b>) direction 1, (<b>c</b>) direction 2, (<b>d</b>) direction 3, (<b>e</b>) direction 4, (<b>f</b>) direction 5, (<b>g</b>) direction 6, (<b>h</b>) direction 7, (<b>i</b>) direction 8.</p>
Full article ">Figure 12 Cont.
<p>Twenty percent occluded images from different directions. (<b>a</b>) Original image, (<b>b</b>) direction 1, (<b>c</b>) direction 2, (<b>d</b>) direction 3, (<b>e</b>) direction 4, (<b>f</b>) direction 5, (<b>g</b>) direction 6, (<b>h</b>) direction 7, (<b>i</b>) direction 8.</p>
Full article ">Figure 13
<p>Performance of different methods under partial occlusion.</p>
Full article ">
23 pages, 6898 KiB  
Article
Characterizing the Spatio-Temporal Pattern of Land Surface Temperature through Time Series Clustering: Based on the Latent Pattern and Morphology
by Huimin Liu, Qingming Zhan, Chen Yang and Jiong Wang
Remote Sens. 2018, 10(4), 654; https://doi.org/10.3390/rs10040654 - 23 Apr 2018
Cited by 40 | Viewed by 8775
Abstract
Land Surface Temperature (LST) is a critical component to understand the impact of urbanization on the urban thermal environment. Previous studies were inclined to apply only one snapshot to analyze the pattern and dynamics of LST without considering the non-stationarity in the temporal [...] Read more.
Land Surface Temperature (LST) is a critical component to understand the impact of urbanization on the urban thermal environment. Previous studies were inclined to apply only one snapshot to analyze the pattern and dynamics of LST without considering the non-stationarity in the temporal domain, or focus on the diurnal, seasonal, and annual pattern analysis of LST which has limited support for the understanding of how LST varies with the advancing of urbanization. This paper presents a workflow to extract the spatio-temporal pattern of LST through time series clustering by focusing on the LST of Wuhan, China, from 2002 to 2017 with a 3-year time interval with 8-day MODerate-resolution Imaging Spectroradiometer (MODIS) satellite image products. The Latent pattern of LST (LLST) generated by non-parametric Multi-Task Gaussian Process Modeling (MTGP) and the Multi-Scale Shape Index (MSSI) which characterizes the morphology of LLST are coupled for pattern recognition. Specifically, spatio-temporal patterns are discovered after the extraction of spatial patterns conducted by the incorporation of k -means and the Back-Propagation neural networks (BP-Net). The spatial patterns of the 6 years form a basic understanding about the corresponding temporal variances. For spatio-temporal pattern recognition, LLSTs and MSSIs of the 6 years are regarded as geo-referenced time series. Multiple algorithms including traditional k -means with Euclidean Distance (ED), shape-based k -means with the constrained Dynamic Time Warping ( c DTW) distance measure, and the Dynamic Time Warping Barycenter Averaging (DBA) centroid computation method ( k - c DBA) and k -shape are applied. Ten external indexes are employed to evaluate the performance of the three algorithms and reveal k - c DBA as the optimal time series clustering algorithm for our study. The study area is divided into 17 geographical time series clusters which respectively illustrate heterogeneous temporal dynamics of LST patterns. The homogeneous geographical clusters correspond to the zoning custom of urban planning and design, and thus, may efficiently bridge the urban and environmental systems in terms of research scope and scale. The proposed workflow can be utilized for other cities and potentially used for comparison among different cities. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The study area represented by the Landsat 8 image (RGB) on 6 October 2014.</p>
Full article ">Figure 2
<p>The date range of the air temperature collected from 2011 to 2017. The red dots represent the daily maximum temperatures, while the blue dots represent the daily minimum temperatures.</p>
Full article ">Figure 3
<p>The representation of the methodological framework used in this study.</p>
Full article ">Figure 4
<p>The operating principles of Euclidean Distance (ED), Dynamic Time Warping (DTW), The Constrained Dynamic Time Warping (<math display="inline"><semantics> <mi>c</mi> </semantics></math>DTW), and the Dynamic Time Warping Barycenter Averaging (DBA). (<b>a</b>) Similarity computation under ED; (<b>b</b>) similarity computation under DTW; (<b>c</b>) Sakoe-Chiba band with a warping window of 5 cells (light blue cells in band) and the warping path computed under <math display="inline"><semantics> <mi>c</mi> </semantics></math>DTW (dark blue cells in band) [<a href="#B23-remotesensing-10-00654" class="html-bibr">23</a>]; (<b>d</b>) centroid computation through the Dynamic Time Warping Barycenter Averaging (DBA) [<a href="#B55-remotesensing-10-00654" class="html-bibr">55</a>].</p>
Full article ">Figure 5
<p>The Latent Pattern of LST (LLST) of 28 July 2014 generated by the Multi-Task Gaussian Process Modeling (MTGP). (<b>a</b>) The original image; (<b>b</b>) the LLST.</p>
Full article ">Figure 6
<p>The LLSTs of the 6 years generated by MTGP. (<b>a</b>) 2002; (<b>b</b>) 2005; (<b>c</b>) 2008; (<b>d</b>) 2011; (<b>e</b>) 2014; (<b>f</b>) 2017.</p>
Full article ">Figure 7
<p>The Multi-Scale Shape Indexes (MSSIs) of the 6 years indicating the surface morphology of LLSTs. The MSSI of (<b>a</b>) 2002, (<b>b</b>) 2005, (<b>c</b>) 2007, (<b>d</b>) 2011, (<b>e</b>) 2014, (<b>f</b>) 2017, (<b>g</b>) the classic morphologies taken from the MSSI of 2014.</p>
Full article ">Figure 8
<p>The determination of the optimal cluster number <math display="inline"><semantics> <mi>k</mi> </semantics></math> for spatial clustering.</p>
Full article ">Figure 9
<p>The spatial classification results of the 6 years. (<b>a</b>) 2002; (<b>b</b>) 2005; (<b>c</b>) 2008; (<b>d</b>) 2011; (<b>e</b>) 2014; (<b>f</b>) 2017.</p>
Full article ">Figure 10
<p>The boxplots of the spatial classification results. LLST of (<b>a</b>) 2002, (<b>b</b>) 2005, (<b>c</b>) 2008, (<b>d</b>) 2011, (<b>e</b>) 2014, (<b>f</b>) 2017; MSSI of (<b>g</b>) 2002, (<b>h</b>) 2005, (<b>i</b>) 2008, (<b>j</b>) 2011, (<b>k</b>) 2014, (<b>l</b>) 2017.</p>
Full article ">Figure 11
<p>The determination of the optimal cluster number <math display="inline"><semantics> <mi>k</mi> </semantics></math> for time series clustering. The left ordinate measures the Sum of Squared Error (SSE) of <math display="inline"><semantics> <mi>k</mi> </semantics></math>-cDBA and <math display="inline"><semantics> <mi>k</mi> </semantics></math>-shape; the right ordinate measures the SSE of <math display="inline"><semantics> <mi>k</mi> </semantics></math>-means.</p>
Full article ">Figure 12
<p>The time series clustering result of the algorithms. (<b>a</b>) <math display="inline"><semantics> <mi>k</mi> </semantics></math>-means; (<b>b</b>) <math display="inline"><semantics> <mi>k</mi> </semantics></math>-cDBA; (<b>c</b>) <math display="inline"><semantics> <mi>k</mi> </semantics></math>-shape.</p>
Full article ">Figure 13
<p>The time series centroids of all 17 clusters. <math display="inline"><semantics> <mrow> <msub> <mrow> <mn>2002</mn> </mrow> <mi>L</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mn>2002</mn> </mrow> <mi>M</mi> </msub> </mrow> </semantics></math> represent the LLST and MSSI of the year 2002, and so on.</p>
Full article ">Figure 14
<p>The boxplots of the time series clustering results. The whole boxplot of (<b>a</b>) LLSTs; (<b>b</b>) MSSIs; The temporal variance (TV) boxplot of (<b>c</b>) the mean LLSTs of each time series; (<b>d</b>) the mean MSSIs of each time series; (<b>e</b>) the standard deviations (Std) of LLSTs of each time series; (<b>f</b>) the Std of MSSIs of each time series.</p>
Full article ">Figure 15
<p>The Land Use and Land Cover (LULC) trajectory analysis of Clusters 15 and 13 by taking two areas as examples. (<b>a</b>) The time series clustering result of <math display="inline"><semantics> <mi>k</mi> </semantics></math>-cDBA where the first box represents the example of Cluster 15 while the second represents Cluster 13; the LULC trajectory demonstration of box (<b>b</b>) 1 and (<b>c</b>) 2.</p>
Full article ">
20 pages, 4192 KiB  
Article
Global MODIS Fraction of Green Vegetation Cover for Monitoring Abrupt and Gradual Vegetation Changes
by Federico Filipponi, Emiliana Valentini, Alessandra Nguyen Xuan, Carlos A. Guerra, Florian Wolf, Martin Andrzejak and Andrea Taramelli
Remote Sens. 2018, 10(4), 653; https://doi.org/10.3390/rs10040653 - 23 Apr 2018
Cited by 48 | Viewed by 8662
Abstract
The presence and distribution of green vegetation cover in the biosphere are of paramount importance in investigating cause-effect phenomena at the land/atmosphere interface, estimating primary production rates as part of global carbon and water cycle assessments and evaluating soil protection and land use [...] Read more.
The presence and distribution of green vegetation cover in the biosphere are of paramount importance in investigating cause-effect phenomena at the land/atmosphere interface, estimating primary production rates as part of global carbon and water cycle assessments and evaluating soil protection and land use change over time. The fraction of green vegetation cover (FCover) as estimated from satellite observations has already been demonstrated to be an extraordinarily useful product for understanding vegetation cover changes, for supporting ecosystem service assessments over areas with variable extents and for processes spanning a variable period of time (abrupt events or long-term processes). This study describes a methodology implemented to estimate global FCover (from 2001 to 2015) by applying a linear spectral mixture analysis with global endmembers to an entire temporal series of MODIS satellite observations and gap-filling missing FCover observations in temporal series using the DINEOF algorithm. The resulting global MODV1 FCover product was validated with two global validation datasets and showed an overall good thematic absolute accuracy (RMSE = 0.146) consistent with the validation performance of other FCover global products. Basic statistics performed on the product show changes in average and trend values and allow for the quantification of gross vegetation loss and gain over different temporal scales. To demonstrate the capacity of this global product to monitor specific dynamics, a multitemporal analysis was performed on selected sites and vegetation responses (i.e., cover changes), and specific dynamics resulting from cause-effect phenomena are briefly discussed. The product is intended to be used for monitoring vegetation dynamics, but it also has the potential to be integrated in other modeling frameworks (e.g., the carbon cycle, primary production, and soil erosion) in conjunction with other spatial datasets such as those on climate and soil type. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Processing chain developed to generate the global FCover product.</p>
Full article ">Figure 2
<p>Global composite spectral mixing space and spectral endmembers. Scatterplots displayed with density colors are the orthogonal projections of three primary principal components (PCs) computed from global MOD13A3 data (monthly, at 1-km spatial resolution) acquired during 2001. (<b>a</b>) Global composite spectral mixing space of the first and second principal components. Black dots indicate substrate (S), vegetation (V) and dark surface (D) global spectral endmember positions. (<b>b</b>) Global composite spectral mixing space of the second and third principal components. (<b>c</b>) Global composite spectral mixing space of the first and third principal components. (d) Spectra of global endmembers. (<b>e</b>) Reflectance values of global spectral endmembers. The gap-filling procedure was able to reconstruct a consistent number of missing values to allow the use of complete FCover time series, mostly at lower latitudes. Looking at differences between the percentage of missing values before and after the gap-filling procedure (<a href="#remotesensing-10-00653-f003" class="html-fig">Figure 3</a>b and <a href="#app1-remotesensing-10-00653" class="html-app">Figure S3</a>), it is possible to see that the FCover product has been reconstructed for almost the entire globe except for the higher latitudes in the northern hemisphere.</p>
Full article ">Figure 3
<p>Direct validation results. (<b>a</b>) Comparison of FCover variable estimates with the corresponding FCover values from the CEOS BELMANIP2 and ImagineS datasets (FCover observed). Dotted lines represent the target accuracy range, and dashed lines represent the optimal target accuracy range. (<b>b</b>) Percentage of valid monthly composite pixels after MODV1 gap-filling and location of ground data. The grid represents the new tiling system adopted for WGS84 MODV1 projection.</p>
Full article ">Figure 4
<p>Over the period of 2001–2015 (<b>a</b>) Average of global FCover product and site locations of the break detection analysis (bounding boxes); (<b>b</b>) global FCover temporal trend slope: percentage increase (green) and decrease (red) are shown as annual rates.</p>
Full article ">Figure 5
<p>Temporal profile (black line), temporal trend (dashed red line), and strong change in temporal profile (i.e., the break represents the highest change in pixel temporal profile; dotted vertical black line) of the selected case study sites: (<b>a</b>) Portugal, (<b>b</b>) Indonesia, (<b>c</b>) Brazil, (<b>d</b>) Australia, (<b>e</b>) Mexico.</p>
Full article ">
16 pages, 4420 KiB  
Article
Salient Object Detection via Recursive Sparse Representation
by Yongjun Zhang, Xiang Wang, Xunwei Xie and Yansheng Li
Remote Sens. 2018, 10(4), 652; https://doi.org/10.3390/rs10040652 - 23 Apr 2018
Cited by 13 | Viewed by 5388
Abstract
Object-level saliency detection is an attractive research field which is useful for many content-based computer vision and remote-sensing tasks. This paper introduces an efficient unsupervised approach to salient object detection from the perspective of recursive sparse representation. The reconstruction error determined by foreground [...] Read more.
Object-level saliency detection is an attractive research field which is useful for many content-based computer vision and remote-sensing tasks. This paper introduces an efficient unsupervised approach to salient object detection from the perspective of recursive sparse representation. The reconstruction error determined by foreground and background dictionaries other than common local and global contrasts is used as the saliency indication, by which the shortcomings of the object integrity can be effectively improved. The proposed method consists of the following four steps: (1) regional feature extraction; (2) background and foreground dictionaries extraction according to the initial saliency map and image boundary constraints; (3) sparse representation and saliency measurement; and (4) recursive processing with a current saliency map updating the initial saliency map in step 2 and repeating step 3. This paper also presents the experimental results of the proposed method compared with seven state-of-the-art saliency detection methods using three benchmark datasets, as well as some satellite and unmanned aerial vehicle remote-sensing images, which confirmed that the proposed method was more effective than current methods and could achieve more favorable performance in the detection of multiple objects as well as maintaining the integrity of the object area. Full article
(This article belongs to the Special Issue Pattern Analysis and Recognition in Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Examples of the limitations of previous contrast and boundary prior-based methods. The images in the first row are the examples of boundary prior testing: (<b>a</b>) input; (<b>b</b>) ground truth; (<b>c</b>) saliency map of saliency optimization from robust background detection (RBD) [<a href="#B29-remotesensing-10-00652" class="html-bibr">29</a>] using boundary prior; (<b>d</b>) saliency map of dense and spares reconstruction based method (DSR) [<a href="#B31-remotesensing-10-00652" class="html-bibr">31</a>] related to boundary prior; (<b>e</b>) initial saliency map of the proposed recursive sparse representation (RSR) generated by Itti’s visual attention model (IT) [<a href="#B5-remotesensing-10-00652" class="html-bibr">5</a>]; (<b>f</b>) final saliency map of the proposed RSR. The images in the second row are the examples of contrast prior testing: (<b>a</b>) input; (<b>b</b>) ground truth; (<b>c</b>) saliency map of low level-features of luminance and color based method (AC) [<a href="#B32-remotesensing-10-00652" class="html-bibr">32</a>] related to local contrast; (<b>d</b>) saliency map of histogram-based contrast method (HC) [<a href="#B33-remotesensing-10-00652" class="html-bibr">33</a>] related to global contrast; (<b>e</b>) initial saliency map of the proposed RSR generated by IT [<a href="#B5-remotesensing-10-00652" class="html-bibr">5</a>]; (<b>f</b>) final saliency map of the proposed RSR.</p>
Full article ">Figure 2
<p>The framework of the proposed approach. Only one scale of SLIC segmentation is illustrated in detail.</p>
Full article ">Figure 3
<p>Saliency results based on different dictionaries: (<b>a</b>) input image; (<b>b</b>) ground truth; (<b>c</b>) IT fixation result; (<b>d</b>) saliency result by RSR with background-based sparse representation only; (<b>e</b>) saliency result by RSR with foreground-based sparse representation and background-based sparse representation without saliency map restricting the dictionary extraction; (<b>f</b>) saliency result by complete RSR. Judging from the two groups of the experiments shown in (<b>e</b>,<b>f</b>), the foreground-based combined methods can get an obviously better result when compared to the single representation by background dictionary as shown in (<b>d</b>).</p>
Full article ">Figure 4
<p>Examples of recursive processing.</p>
Full article ">Figure 5
<p>Integration of multiscale results.</p>
Full article ">Figure 6
<p>Visual comparison on MRSA-ASD, ECSSD and SED2 datasets, where GT represents the ground truth.</p>
Full article ">Figure 7
<p>Quantitative comparison results on MSRA-ASD, SED2 and ECSSD datasets. The first row is <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>R</mi> </mrow> </semantics></math> curve; the second row is <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>-</mo> <mi>M</mi> <mi>e</mi> <mi>a</mi> <mi>s</mi> <mi>u</mi> <mi>r</mi> <mi>e</mi> </mrow> </semantics></math> curve; the third row is the <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>r</mi> <mi>e</mi> <mi>c</mi> <mi>i</mi> <mi>s</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mi>c</mi> <mi>a</mi> <mi>l</mi> <mi>l</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>F</mi> <mi>β</mi> </msup> </mrow> </semantics></math> values with adaptive threshold; and the last row is the <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>A</mi> <mi>E</mi> </mrow> </semantics></math> measure value with adaptive threshold.</p>
Full article ">Figure 8
<p>Comparison of the remote sensing dataset. In the right side, the first three graphs are the <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>r</mi> <mi>e</mi> <mi>c</mi> <mi>i</mi> <mi>s</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mi>c</mi> <mi>a</mi> <mi>l</mi> <mi>l</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>F</mi> <mi>β</mi> </msup> </mrow> </semantics></math> values with adaptive threshold; and the last one is the <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>A</mi> <mi>E</mi> </mrow> </semantics></math> measure value with adaptive threshold.</p>
Full article ">Figure 9
<p>Some results with shortcomings of the proposed RSR in benchmark datasets. Columns 1 and 2 are over-detections and columns 3–5 are edge limitations.</p>
Full article ">
17 pages, 5267 KiB  
Article
Impacts of 3D Aerosol, Cloud, and Water Vapor Variations on the Recent Brightening during the South Asian Monsoon Season
by Zengxin Pan, Feiyue Mao, Wei Wang, Bo Zhu, Xin Lu and Wei Gong
Remote Sens. 2018, 10(4), 651; https://doi.org/10.3390/rs10040651 - 23 Apr 2018
Cited by 14 | Viewed by 5163
Abstract
South Asia is experiencing a levelling-off trend in solar radiation and even a transition from dimming to brightening. Any change in incident solar radiation, which is the only significant energy source of the global ecosystem, profoundly affects our habitats. Here, we use multiple [...] Read more.
South Asia is experiencing a levelling-off trend in solar radiation and even a transition from dimming to brightening. Any change in incident solar radiation, which is the only significant energy source of the global ecosystem, profoundly affects our habitats. Here, we use multiple observations of the A-Train constellation to evaluate the impacts of three-dimensional (3D) aerosol, cloud, and water vapor variations on the changes in surface solar radiation during the monsoon season (June–September) in South Asia from 2006 to 2015. Results show that surface shortwave radiation (SSR) has possibly increased by 16.2 W m−2 during this period. However, an increase in aerosol loading is inconsistent with the SSR variations. Instead, clouds are generally reduced and thinned by approximately 8.8% and 280 m, respectively, with a decrease in both cloud water path (by 34.7 g m−2) and particle number concentration under cloudy conditions. Consequently, the shortwave cloud radiative effect decreases by approximately 45.5 W m−2 at the surface. Moreover, precipitable water in clear-sky conditions decreases by 2.8 mm (mainly below 2 km), and related solar brightening increases by 2.5 W m−2. Overall, the decreases in 3D water vapor and clouds distinctly result in increased absorption of SSR and subsequent surface brightening. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Temporal variations in spatial average AOD from (<b>a</b>) CALIPSO and MODIS during the monsoon season and (<b>b</b>) the vertical average aerosol extinction coefficient at 532 nm in South Asia from 2006 to 2015, respectively. The dashed lines indicate the linear fit line of the solid line with the same color; <span class="html-italic">p</span> is the significance level.</p>
Full article ">Figure 2
<p>Temporal variations in spatial average vertical cloud physical parameters from CloudSat during the monsoon season in South Asia from 2006 to 2015: (<b>a</b>) cloud vertical frequency distribution, and liquid and ice (<b>b</b>,<b>e</b>) water content, (<b>c</b>,<b>f</b>) effective radius, and (<b>d</b>,<b>g</b>) number concentration.</p>
Full article ">Figure 3
<p>Temporal variations in spatial average (<b>a</b>) cloud fraction and CWP, as well as (<b>b</b>) uppermost CTH, lowermost CBH and CGD; spatial distributions of the temporal changes in average (<b>c</b>) cloud fraction, (<b>d</b>) CH, (<b>e</b>) CWP, and (<b>f</b>) CGD from CloudSat during the monsoon season in South Asia from 2006 to 2015. The dashed lines indicate the linear fit line of the solid line with the same color; <span class="html-italic">p</span> is the significance level.</p>
Full article ">Figure 4
<p>Temporal variations in spatial average vertical (<b>a</b>) SW, (<b>b</b>) LW, and (<b>c</b>) net heat rating in all-sky conditions from CloudSat during the monsoon season in South Asia from 2006 to 2015.</p>
Full article ">Figure 5
<p>(<b>a</b>) Temporal variations in spatial average CRE at the TOA and surface, and (<b>b</b>) spatial distribution of the temporal changes in average SW CRE from CloudSat at the surface during the monsoon season in South Asia from 2006 to 2015. The dashed lines indicate the linear fit line of the solid line with the same color; <span class="html-italic">p</span> is the significance level.</p>
Full article ">Figure 6
<p>Temporal variations in spatial average vertical RH in (<b>a</b>) all-sky and (<b>c</b>) clear-sky conditions; spatial distribution of temporal changes in the average PW in (<b>b</b>) all-sky and (<b>d</b>) clear-sky conditions from the ECMWF-AUX during the monsoon season in South Asia from 2006 to 2015.</p>
Full article ">Figure 7
<p>Temporal variations in spatial average (<b>a</b>) PW (blue line) and SSR in clear-sky conditions from BUGSrad (red line) and CloudSat (orange line); spatial distribution of temporal changes in average (<b>b</b>) SSR from BUGSrad in clear-sky conditions during the monsoon season in South Asia from 2006 to 2015. The dashed lines indicate the linear fit line of the solid line with the same color; <span class="html-italic">p</span> is the significance level.</p>
Full article ">Figure 8
<p>Temporal variations in spatial average PW in (<b>a</b>) all-sky and (<b>b</b>) clear-sky conditions at different ranges of altitude during the monsoon season in South Asia from 2006 to 2015. The dashed lines indicate the linear fit line of the solid line with the same color; <span class="html-italic">p</span> is the significance level.</p>
Full article ">Figure 9
<p>Temporal variations in (<b>a</b>) spatial average AOD from MODIS and (<b>b</b>) SSR from CERES during the pre-monsoon, monsoon, and dry seasons in South Asia from 2006 to 2015, respectively. The dashed lines indicate the linear fit line of the solid line with the same color; <span class="html-italic">p</span> is the significance level.</p>
Full article ">Figure 10
<p>Schematic of the impacts of 3D aerosol, cloud, and water vapor variations on brightening during the monsoon season in South Asia from 2006 to 2015. Background gradient color represents the changing RH. The relative humidity is high when the color is dark.</p>
Full article ">
1 pages, 2626 KiB  
Article
Enhanced Modeling of Annual Temperature Cycles with Temporally Discrete Remotely Sensed Thermal Observations
by Zhaoxu Zou, Wenfeng Zhan, Zihan Liu, Benjamin Bechtel, Lun Gao, Falu Hong, Fan Huang and Jiameng Lai
Remote Sens. 2018, 10(4), 650; https://doi.org/10.3390/rs10040650 - 23 Apr 2018
Cited by 31 | Viewed by 5741
Abstract
Satellite thermal remote sensing provides land surface temperatures (LST) over extensive areas that are vital in various applications, but this technique suffers from its sampling style and the impenetrability of clouds, which frequently generates data gaps. Annual temperature cycle (ATC) models can fill [...] Read more.
Satellite thermal remote sensing provides land surface temperatures (LST) over extensive areas that are vital in various applications, but this technique suffers from its sampling style and the impenetrability of clouds, which frequently generates data gaps. Annual temperature cycle (ATC) models can fill these gaps and estimate continuous daily LST dynamics from a number of thermal observations. However, the standard ATC model (termed ATCS) remains incapable of quantifying the short-term LST variations caused by synoptic conditions. By incorporating in-situ surface air temperatures (SATs) and satellite-derived normalized difference vegetation indexes (NDVIs), here we proposed an enhanced ATC model (ATCE) to describe the daily LST fluctuations. With Aqua/MODIS LST products as validation data, we implemented and tested the ATCE over the Yangtze River Delta region of China. The results demonstrate that, when compared with the ATCS, the overall root mean square errors of the ATCE decrease by 1.0 and 0.8 K for the day and night, respectively. The accuracy improvements vary with land cover types with greater improvements over the forest, grassland, and built-up areas than over cropland and wetland. The assessments at different time scales further confirm that LST fluctuations can be better described by the ATCE. Though with limitations, we consider this new model and its associated parameters hold great potentials in various applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Information on the study area (i.e., the Yangtze River Delta, YRD). (<b>a</b>–<b>c</b>) denote the general location, digital elevation model, and land cover type map, respectively.</p>
Full article ">Figure 2
<p>A demonstration of the standard ATC model (represented by the ATC<sub>S</sub>) and the observed continuous daily LST dynamics. LST dynamics and the model predictions are presented for (<b>a</b>) an entire year and (<b>b</b>) a shorter period ranging from Jun. to Oct. <span class="html-italic">T</span><sub>0</sub>, <span class="html-italic">A</span>, and <span class="html-italic">θ</span> are the annual mean LST, ATC amplitude, and the corresponding phase shift relative to the spring equinox, respectively; Δ<span class="html-italic">T</span><sub>air</sub> denotes the SAT fluctuation; and <span class="html-italic">γ</span> is the ratio between the LST and SAT fluctuations.</p>
Full article ">Figure 3
<p>Performances of the ATC<sub>S</sub> and ATC<sub>E</sub> for the day and night. (<b>a</b>) is the boxplot on the predicted RMSEs; while (<b>b</b>) denotes the probability distribution functions (PDFs) of the associated RMSEs. Δ<span class="html-italic">T</span><sub>day</sub> and Δ<span class="html-italic">T</span><sub>night</sub> are the accuracy improvements for the ATC<sub>E</sub> when referenced to the ATC<sub>S</sub> for the day and night, respectively. The end of each whisker of the boxplot indicates the highest or lowest value within 1.5 times of the inter-quartile range (IQR).</p>
Full article ">Figure 4
<p>Spatial distributions on the predicted RMSEs for the ATC<sub>S</sub> (left) and ATC<sub>E</sub> (right) during the day (<b>a</b>,<b>b</b>) and the night (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 5
<p>Model performances at different time scales. (<b>a</b>) denotes the daily comparisons between the ATC<sub>E</sub>-predicted and observed LSTs at a selected pixel; (<b>b</b>,<b>c</b>) show the model performances at the monthly scale and a series of time scales ranging from 1 to 32 days, respectively.</p>
Full article ">
16 pages, 2408 KiB  
Article
The Use of Three-Dimensional Convolutional Neural Networks to Interpret LiDAR for Forest Inventory
by Elias Ayrey and Daniel J. Hayes
Remote Sens. 2018, 10(4), 649; https://doi.org/10.3390/rs10040649 - 23 Apr 2018
Cited by 77 | Viewed by 15218
Abstract
As light detection and ranging (LiDAR) technology becomes more available, it has become common to use these datasets to generate remotely sensed forest inventories across landscapes. Traditional methods for generating these inventories employ the use of height and proportion metrics to measure LiDAR [...] Read more.
As light detection and ranging (LiDAR) technology becomes more available, it has become common to use these datasets to generate remotely sensed forest inventories across landscapes. Traditional methods for generating these inventories employ the use of height and proportion metrics to measure LiDAR returns and relate these back to field data using predictive models. Here, we employ a three-dimensional convolutional neural network (CNN), a deep learning technique that scans the LiDAR data and automatically generates useful features for predicting forest attributes. We test the accuracy in estimating forest attributes using the three-dimensional implementations of different CNN models commonly used in the field of image recognition. Using the best performing model architecture, we compared CNN performance to models developed using traditional height metrics. The results of this comparison show that CNNs produced 12% less prediction error when estimating biomass, 6% less in estimating tree count, and 2% less when estimating the percentage of needleleaf trees. We conclude that using CNNs can be a more accurate means of interpreting LiDAR data for forest inventories compared to standard approaches. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A comparison of biomass model performance by architecture. Results are shown in terms of root mean square error (RMSE) and bias. Red lines represent the performance of the random forest model trained on traditional height metrics. Models are listed in order of complexity, which refers to the number of trainable parameters in the model. Note that the Y-axis begins at a RMSE of 45 Mg/ha to better highlight differences in model performance.</p>
Full article ">Figure 2
<p>Predicted vs. observed plots for each of the three model types estimating biomass, tree count, and percent needleleaf. The solid line is a one-to-one line, and the dashed red line is a loess regression fit of the data. <b>Left</b> (<b>A</b>, <b>D</b>, and <b>G</b>): linear mixed models with traditional height metrics (LMM-THM). <b>Center</b> (<b>B</b>, <b>E</b>, and <b>H</b>): random forest models with traditional height metrics (RF-THM). <b>Right</b> (<b>C</b>, <b>F</b>, and <b>I</b>): Inception-V3 convolutional neural networks (Inception-V3 CNN).</p>
Full article ">Figure 3
<p>(<b>A</b>) The initial point cloud prior to being input to the CNN model, colored by height; (<b>B</b>–<b>D</b>) feature maps resulting from the first layer of convolutions using LeNet. Red values represent areas of higher interest to the model, blue and white represent areas of lower interest. Note that each convolution detects different patterns and features in the voxelized point cloud.</p>
Full article ">
14 pages, 26661 KiB  
Article
Evaluation of ISS-RapidScat Wind Vectors Using Buoys and ASCAT Data
by Jungang Yang and Jie Zhang
Remote Sens. 2018, 10(4), 648; https://doi.org/10.3390/rs10040648 - 23 Apr 2018
Cited by 10 | Viewed by 4680
Abstract
The International Space Station scatterometer (named ISS-RapidScat) was launched by NASA on 20 September 2014 as a continuation of the QuikSCAT climate data record to maintain the availability of Ku-band scatterometer data after the QuikSCAT missions ended. In this study, the overall archived [...] Read more.
The International Space Station scatterometer (named ISS-RapidScat) was launched by NASA on 20 September 2014 as a continuation of the QuikSCAT climate data record to maintain the availability of Ku-band scatterometer data after the QuikSCAT missions ended. In this study, the overall archived ISS-RapidScat wind vectors in the wind speed range of 0–24 m/s are evaluated by the global moored buoys’ wind observations, including the U.S. National Data Buoy Center (NDBC), the Tropical Atmosphere Ocean (TAO), and the Pilot Research Moored Array in the Tropical Atlantic (PIRATA), the Research Moored Array for African–Asian–Australian Monsoon Analysis and Prediction (RAMA), and Advanced Scatterometer (ASCAT) wind data in the same period of ISS-RapidScat by calculating the statistical parameters, namely, the root mean square error (RMSE), bias (mean of residuals), and correlation coefficient (R) between the collocated data. The comparisons with the global moored buoys show that the RapidScat wind vectors are consistent with buoys’ wind measurements. The average errors of the RapidScat wind vectors are 1.42 m/s and 19.5°. The analysis of the RapidScat wind vector errors at different buoy wind speeds in bins of 1 m/s indicates that the errors of the RapidScat wind speed reduce firstly, and then increase with the increasing buoy wind speed, and the errors of the RapidScat wind direction decrease with increasing buoy wind speed. The comparisons of the errors of the RapidScat wind speed and direction at different months from April 2015 to August 2016 show that the accuracies of the RapidScat wind vectors have no dependence on the time, and the biases of the RapidScat wind speed indicate that there is an annual periodic signal of wind speed errors which are due to the annual cycle variation of ocean winds. The accuracies of the RapidScat wind vectors at different times in one day are also analyzed and the results show that the accuracy of the RapidScat wind vectors at different times of the day is basically consistent and with no diurnal variation. In order to evaluate the ISS-RapidScat wind vectors of the global oceans, the differences (RapidScat-ASCAT) in the wind speed range of 0–30 m/s are analyzed in the different months from October 2014 to August 2016, and the average RMSEs of differences between ISS-RapidScat and ASCAT wind vectors are less than 1.15 m/s and 15.21°. In general, the evaluation of the all-over archived ISS-RapidScat wind vectors show that the accuracies of the ISS-RapidScat wind vectors satisfy the general scatterometer’s mission requirement and are consistent with ASCAT wind data. Full article
(This article belongs to the Special Issue Radar Remote Sensing of Oceans and Coastal Areas)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Locations of the National Data Buoy Center (NDBC), Tropical Atmosphere Ocean (TAO), Pilot Research Moored Array (PIRATA), and Research Moored Array for African-Asian-Australian Monsoon (RAMA) moored buoys.</p>
Full article ">Figure 2
<p>Percentage of the number of observations by the NDBC, TAO, PIRATA, and RAMA moored buoys.</p>
Full article ">Figure 3
<p>The number distribution of RapidScat and The Advanced Scatterometer (ASCAT) collocated data pairs.</p>
Full article ">Figure 4
<p>Scatterplots for wind speed and direction of the comparisons between RapidScat and moored buoys. (<b>a</b>) NDBC wind speed; (<b>b</b>) NDBC wind direction; (<b>c</b>) PIRATA wind speed; (<b>d</b>) PIRATA wind direction; (<b>e</b>) RAMA wind speed; (<b>f</b>) RAMA wind direction; (<b>g</b>) TAO wind speed; (<b>h</b>) TAO wind direction.</p>
Full article ">Figure 4 Cont.
<p>Scatterplots for wind speed and direction of the comparisons between RapidScat and moored buoys. (<b>a</b>) NDBC wind speed; (<b>b</b>) NDBC wind direction; (<b>c</b>) PIRATA wind speed; (<b>d</b>) PIRATA wind direction; (<b>e</b>) RAMA wind speed; (<b>f</b>) RAMA wind direction; (<b>g</b>) TAO wind speed; (<b>h</b>) TAO wind direction.</p>
Full article ">Figure 5
<p>Dependence of statistical parameters and wind speed and direction residuals (RapidScat-buoy) on the buoy wind speed.</p>
Full article ">Figure 6
<p>Errors of RapidScat wind vectors in different months from April 2015 to August 2016 for NDBC, PIRATA, RAMA, and TAO buoys.</p>
Full article ">Figure 7
<p>Errors of RapidScat wind vectors at different times of the day for NDBC, PIRATA, RAMA, and TAO buoys.</p>
Full article ">Figure 8
<p>PDFs of RapidScat and buoy wind speeds and directions in the bin of 1 m/s and 10°.</p>
Full article ">Figure 9
<p>Scatterplots for wind speed and direction of the comparisons between RapidScat and ASCAT in December 2014 and June 2016. (<b>a</b>) wind speed in December 2014; (<b>b</b>) wind direction in December 2014; (<b>c</b>) wind speed in June 2016; (<b>d</b>) wind direction in June 2016.</p>
Full article ">Figure 10
<p>statistical parameters of difference between RapidScat and ASCAT wind vectors in different months from October 2014 to August 2016.</p>
Full article ">
15 pages, 2622 KiB  
Article
Spatial Downscaling of Gross Primary Productivity Using Topographic and Vegetation Heterogeneity Information: A Case Study in the Gongga Mountain Region of China
by Xinyao Xie, Ainong Li, Huaan Jin, Gaofei Yin and Jinhu Bian
Remote Sens. 2018, 10(4), 647; https://doi.org/10.3390/rs10040647 - 22 Apr 2018
Cited by 14 | Viewed by 4183
Abstract
Due to the spatial heterogeneity of land surfaces, downscaling is an important issue in the development of carbon cycle models when evaluating the role of ecosystems in the global carbon cycle. In this study, a downscaling algorithm was developed to model gross primary [...] Read more.
Due to the spatial heterogeneity of land surfaces, downscaling is an important issue in the development of carbon cycle models when evaluating the role of ecosystems in the global carbon cycle. In this study, a downscaling algorithm was developed to model gross primary productivity (GPP) at 500 m in a time series over rugged terrain, which considered the effects of spatial heterogeneity on carbon flux simulations. This work was carried out for a mountainous area with an altitude ranging from 2606 to 4744 m over the Gongga Mountain (Sichuan Province, China). In addition, the Moderate Resolution Imaging Spectroradiometer (MODIS) GPP product at 1 km served as the primary dataset for the downscaling algorithm, and the 500 m MODIS GPP product was used as the reference dataset to evaluate the downscaled GPP results. Moreover, in order to illustrate the advantages and benefits of the proposed downscaling method, the downscaled results in this work, along with ordinary kriging downscaled results, spline downscaled results and inverse distance weighted (IDW) downscaled results, were compared to the MODIS GPP at 500 m. The results showed that (1) the GPP difference between the 500 m MODIS GPP and the proposed downscaled GPP results was primarily in the range of [−1, 1], showing that both vegetation heterogeneity factors (i.e., LAI) and topographic factors (i.e., altitude, slope and aspect) were useful for GPP downscaling; (2) the proposed downscaled results (R2 = 0.89, RMSE = 1.03) had a stronger consistency with the 500 m MODIS GPP than those of the ordinary kriging downscaled results (R2 = 0.43, RMSE = 1.36), the spline downscaled results (R2 = 0.40, RMSE = 1.50) and the IDW downscaled results (R2 = 0.42, RMSE = 1.10) for all Julian days; and (3) the inconsistency between MODIS GPP at 500 m and 1 km increased with the increase in altitude and slope. The proposed downscaling algorithm could provide a reference when considering the effects of spatial heterogeneity on carbon flux simulations and retrieving other fine resolution ecological-physiology parameters (e.g., net primary productivity and evaporation) over topographically complex terrains. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area.</p>
Full article ">Figure 2
<p>Flow chart of the proposed downscaling algorithm.</p>
Full article ">Figure 3
<p>Spatial distributions of altitude (<b>a</b>); slope (<b>b</b>); aspect (<b>c</b>) and LAI (<b>d</b>) at 500 m and 1 km.</p>
Full article ">Figure 4
<p>Density scatterplot between MODIS GPP and downscaled GPP at 500 m during Julian days 169–209. (<b>a</b>) represents the comparison between MODIS GPP products; and (<b>b</b>–<b>e</b>) represent the comparison between downscaled GPP and MODIS GPP at 500 m. The solid lines are the regression lines, while the dashed lines are the 1:1 lines. The green and red represent the low-density and high-density areas, respectively.</p>
Full article ">Figure 5
<p>Relationships between topographic factors and inconsistency in the MODIS GPP at 500 m and 1 km during Julian days 169–209.</p>
Full article ">Figure 6
<p>The spatial distribution of absolute GPP differences between the 500 m MODIS GPP and the downscaled GPP results for Julian days 169–209. Figures (<b>a</b>–<b>f</b>) indicate Julian days 169–209, respectively.</p>
Full article ">Figure 7
<p>The density distributions concerning the GPP differences between the 500 m MODIS GPP and the downscaled GPP results for Julian days 169–209. Figures (<b>a</b>–<b>f</b>) indicate Julian days 169–209, respectively.</p>
Full article ">
25 pages, 3641 KiB  
Article
Evaluation of Heavy Precipitation Simulated by the WRF Model Using 4D-Var Data Assimilation with TRMM 3B42 and GPM IMERG over the Huaihe River Basin, China
by Lu Yi, Wanchang Zhang and Kai Wang
Remote Sens. 2018, 10(4), 646; https://doi.org/10.3390/rs10040646 - 22 Apr 2018
Cited by 36 | Viewed by 6477
Abstract
To obtain independent, consecutive, and high-resolution precipitation data, the four-dimensional variational (4D-Var) method was applied to directly assimilate satellite precipitation products into the Weather Research and Forecasting (WRF) model. The precipitation products of the Tropical Rainfall Measuring Mission 3B42 (TRMM 3B42) and its [...] Read more.
To obtain independent, consecutive, and high-resolution precipitation data, the four-dimensional variational (4D-Var) method was applied to directly assimilate satellite precipitation products into the Weather Research and Forecasting (WRF) model. The precipitation products of the Tropical Rainfall Measuring Mission 3B42 (TRMM 3B42) and its successor, the Integrated Multi-satellitE Retrievals for Global Precipitation Measurement (GPM IMERG) were assimilated in this study. Two heavy precipitation events that occurred over the Huaihe River basin in eastern China were studied. Before assimilation, the WRF model simulations were first performed with different forcing data to select more suitable forcing data and determine the control experiments for the subsequent assimilation experiments. Then, TRMM 3B42 and GPM IMERG were separately assimilated into the WRF. The simulated precipitation results in the outer domain (D01), with a 27-km resolution, and the inner domain (D02), with a 9-km resolution, were evaluated in detail. The assessments showed that (1) 4D-Var with TRMM 3B42 or GPM IMERG could both significantly improve WRF precipitation predictions at a time interval of approximately 12 h; (2) the WRF simulated precipitation assimilated with GPM IMERG outperformed the one with TRMM 3B42; (3) for the WRF output precipitation assimilated with GPM IMERG over D02, which has spatiotemporal resolutions of 9 km and 50 s, the correlation coefficients of the studied events in August and November were 0.74 and 0.51, respectively, at the point and daily scales, and the mean Heidke skill scores for the two studied events both reached 0.31 at the grid and hourly scales. This study can provide references for the assimilation of TRMM 3B42 or GPM IMERG into the WRF model using 4D-Var, which is especially valuable for hydrological applications of GPM IMERG during the transition period from the TRMM era into the GPM era. Full article
(This article belongs to the Special Issue Assimilation of Remote Sensing Data into Earth System Models)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Outer domain (D01, 27-km resolution) and inner domain (D02, 9-km resolution) defined in the Weather Research and Forecasting (WRF) model; (<b>b</b>) location and meteorological stations of the Huaihe River basin.</p>
Full article ">Figure 2
<p>(<b>a</b>) Contribution (%) of accumulated daily precipitation to the total annual precipitation for 30 CMA stations across the HRB in 2015 (sorted in decreasing order); (<b>b</b>) mean monthly precipitation of the HRB in 2015; (<b>c</b>) daily precipitation at the 30 CMA stations and the mean daily precipitation of the basin (HRB).</p>
Full article ">Figure 3
<p>Scatter plots of daily precipitation (mm/day) observed by the China Meteorological Administration (CMA) meteorological stations and predicted by the WRF control experiments: CTL1 and CTL2 (<b>a</b>) and CTL3 and CTL4 (<b>b</b>).</p>
Full article ">Figure 4
<p>Box plots * of evaluation scores <span class="html-italic">BIAS</span> (<b>a</b>), <span class="html-italic">FAR</span> (<b>b</b>), <span class="html-italic">POD</span> (<b>c</b>), <span class="html-italic">POFD</span> (<b>d</b>), and <span class="html-italic">HSS</span> (<b>e</b>) for hourly precipitation (exceeding 0.1 mm/h) simulated by the WRF CTL experiments and estimated by the merged Climate Prediction Center Morphing technique (CMORPH) data. * The lower and upper edges of the central box represent the first and third quartiles (25% and 75%, respectively), and the band and the circle inside the box represent the 50th percentiles and the mean values, respectively. The ends of the outliers represent the minimum and maximum values of the score distributions. The asterisks represent several possible alternative values.</p>
Full article ">Figure 5
<p>Spatial patterns of 12-hour accumulated precipitation obtained from experiments CTL2, CTL4, DA1-4, and the merged CMORPH estimations over a subset of D01.</p>
Full article ">Figure 6
<p>Scatter plots of daily precipitation (mm/day) observed by CMA meteorological stations and simulated by the CTL and DA experiments for rainfall events in August (<b>a</b>) and November (<b>b</b>) of 2015.</p>
Full article ">Figure 7
<p>Skill scores of BIAS (<b>a</b>), FAR (<b>b</b>), POD (<b>c</b>), POFD (<b>d</b>), and HSS (<b>e</b>) for the hourly precipitation (48 h in all during the study period for each event) obtained from the CLT2, CTL4, and DA experiments.</p>
Full article ">Figure 8
<p>Mean hourly precipitation (mm/h) of the basin via the merged CMORPH estimates and the CTL2, CTL4, and DA experiments for rainfall events in August (<b>a</b>) and November (<b>b</b>).</p>
Full article ">Figure 9
<p>(<b>a</b>) Error indices for daily precipitation between the CTL experiment simulations and the CMA observations; (<b>b</b>) skill score averages of hourly precipitation between the CTL experiment simulations and the merged CMORPH estimations.</p>
Full article ">Figure 10
<p>Skill scores of daily precipitation (mm/day) obtained from the CTL2, CTL4, and DA experiments in comparison with the merged CMORPH estimations at different thresholds ranging from 1 mm/day to 70 mm/day. (<b>a</b>-<b>d</b>), (<b>e</b>-<b>h</b>), (<b>i</b>-<b>l</b>), (<b>m</b>-<b>p</b>) and (<b>q</b>-<b>t</b>) are the skill scores of <span class="html-italic">BIAS</span>, <span class="html-italic">FAR</span>, <span class="html-italic">POD</span>, <span class="html-italic">POFD</span> and <span class="html-italic">HSS</span>, respectively, for the daily precipitation in the first and second days of the event A and the event N.</p>
Full article ">Figure 11
<p>Increments of skill scores of <span class="html-italic">BIAS</span> (<b>a</b>), <span class="html-italic">FAR</span> (<b>b</b>), <span class="html-italic">POD</span> (<b>c</b>), <span class="html-italic">POFD</span> (<b>d</b>), and <span class="html-italic">HSS</span> (<b>e</b>) in the DA experiments compared to their corresponding CTL experiments. The table in each graph lists the evaluation scores of CTL2 and CTL4. The x-axis represents the first, second, third, and fourth 12-hour periods in the overall study period.</p>
Full article ">Figure 12
<p>Normalized Taylor diagrams of the precipitation simulated by the CTL2, CTL4, and DA1-4 experiments on the first day (<b>a</b>), second day (<b>b</b>), and over the whole 48 h (<b>c</b>).</p>
Full article ">
21 pages, 37047 KiB  
Article
Pushbroom Hyperspectral Data Orientation by Combining Feature-Based and Area-Based Co-Registration Techniques
by Kévin Barbieux
Remote Sens. 2018, 10(4), 645; https://doi.org/10.3390/rs10040645 - 22 Apr 2018
Cited by 10 | Viewed by 6475
Abstract
Direct georeferencing of airborne pushbroom scanner data usually suffers from the limited precision of navigation sensors onboard of the aircraft. The bundle adjustment of images and orientation parameters, used to perform geocorrection of frame images during the post-processing phase, cannot be used for [...] Read more.
Direct georeferencing of airborne pushbroom scanner data usually suffers from the limited precision of navigation sensors onboard of the aircraft. The bundle adjustment of images and orientation parameters, used to perform geocorrection of frame images during the post-processing phase, cannot be used for pushbroom cameras without difficulties—it relies on matching corresponding points between scan lines, which is not feasible in the absence of sufficient overlap and texture information. We address this georeferencing problem by equipping our aircraft with both a frame camera and a pushbroom scanner: the frame images and the navigation parameters measured by a couple GPS/Inertial Measurement Unit (IMU) are input to a bundle adjustment algorithm; the output orientation parameters are used to project the scan lines on a Digital Elevation Model (DEM) and on an orthophoto generated during the bundle adjustment step; using the image feature matching algorithm Speeded Up Robust Features (SURF), corresponding points between the image formed by the projected scan lines and the orthophoto are matched, and through a least-squares method, the boresight between the two cameras is estimated and included in the calculation of the projection. Finally, using Particle Image Velocimetry (PIV) on the gradient image, the projection is deformed into a final image that fits the geometry of the orthophoto. We apply this algorithm to five test acquisitions over Lake Geneva region (Switzerland) and Lake Baikal region (Russia). The results are quantified in terms of Root Mean Square Error (RMSE) between matching points of the RGB orthophoto and the pushbroom projection. From a first projection where the Interior Orientation Parameters (IOP) are known with limited precision and the RMSE goes up to 41 pixels, our geocorrection estimates IOP, boresight and Exterior Orientation Parameters (EOP) and produces a new projection with an RMSE, with the reference orthophoto, around two pixels. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Illustrates the setting for our airborne acquisitions. (<b>b</b>) Represents the different frame coordinates for the three sensors: the frame camera, the pushbroom (PB) camera, and the IMU.</p>
Full article ">Figure 2
<p>Flowchart of the proposed geocorrection method.</p>
Full article ">Figure 3
<p>Processing of the frame images using Agisoft Photoscan: (<b>a</b>) alignment of frames and computation of the orthophoto; and (<b>b</b>) computation of the DEM of the area (Selenga Delta Village).</p>
Full article ">Figure 4
<p>Orthoprojection of the scan lines (cyan) on top of the reference orthomosaic (red) above a village near the Selenga Delta of Lake Baikal. (<b>a</b>) is the orthoprojection of the pushbroom pixels (without interpolation) and (<b>b</b>) is the image as seen after bilinear interpolation of the projected values.</p>
Full article ">Figure 5
<p>Illustrations of the various steps of our algorithm for a test flight over a village of the Selenga Delta (Russia, 17 August 2014). All images show the superimposed scan lines (in cyan) on top of the reference orthophoto (in red).</p>
Full article ">Figure 6
<p>Scatter of the difference vectors for the pairs of points matched by the SURF , for a flight over a village of the Selenga Delta. All the matches are represented by red crosses. Tilted blue crosses correspond to the pairs kept after removing outliers.</p>
Full article ">Figure 7
<p>Averages of the cross-correlations of corresponding patterns between a reference orthophoto and a co-registered mosaic, for the patterns and their gradients.</p>
Full article ">Figure 8
<p>Examples of co-registered patterns using PIV on original images (<b>a</b>–<b>c</b>) and on gradient images (<b>d</b>–<b>f</b>).</p>
Full article ">Figure 9
<p>Superposition of the reference orthophotos and the mosaic produced with the scan lines: (<b>a</b>), (<b>c</b>) before geocorrection and (<b>b</b>), (<b>d</b>) after geocorrection. (<b>a</b>,<b>b</b>): Selenga Village 2; (<b>c</b>,<b>d</b>): Selenga Rivers.</p>
Full article ">Figure 10
<p>Superposition of the reference orthophotos and the mosaic produced with the scan lines: (<b>a</b>,<b>c</b>) before geocorrection and (<b>b</b>,<b>d</b>) after geocorrection. (<b>a</b>,<b>b</b>): Gremyachinsk; (<b>c</b>,<b>d</b>): shore of Lake Geneva.</p>
Full article ">
14 pages, 2877 KiB  
Article
SAR Mode Altimetry Observations of Internal Solitary Waves in the Tropical Ocean Part 1: Case Studies
by Adriana M. Santos-Ferreira, José C. B. Da Silva and Jorge M. Magalhaes
Remote Sens. 2018, 10(4), 644; https://doi.org/10.3390/rs10040644 - 22 Apr 2018
Cited by 24 | Viewed by 6478
Abstract
It is well known that internal waves (IWs) of tidal frequency (i.e., internal tides) are successfully detected in sea surface height (SSH) by satellite altimetry. Shorter period internal solitary waves (ISWs), whose periods (and spatial scales) are an order of magnitude smaller than [...] Read more.
It is well known that internal waves (IWs) of tidal frequency (i.e., internal tides) are successfully detected in sea surface height (SSH) by satellite altimetry. Shorter period internal solitary waves (ISWs), whose periods (and spatial scales) are an order of magnitude smaller than tidal internal waves, have been generally assumed too small to be detected with conventional altimeters. This is because conventional (pulse-limited) radar altimeter footprints are somewhat larger than or of similar size, at best, as the typical wavelengths of the ISWs. Here we demonstrate that the synthetic aperture radar altimeter (SRAL) on board the Sentinel-3A can detect short-period ISWs. A variety of signatures owing to the surface manifestations of the ISWs are apparent in the SRAL Level-2 products over the ocean. These signatures are identified in several geophysical parameters, such as radar backscatter (sigma0), sea level anomaly (SLA), and significant wave height (SWH). Radar backscatter is the primary parameter in which ISWs can be identified owing to the measurable sea surface roughness perturbations in the along-track sharpened SRAL footprint. The SRAL footprint is sufficiently small to capture radar power fluctuations over successive wave crests and troughs, which produce rough and slick surface patterns arrayed in parallel bands with scales of a few kilometers. The ISW signatures are unambiguously identified in the SRAL because of the exact synergy with OLCI (Ocean Land Colour Imager) images, which in cloud-free conditions allow clear identification of the ISWs in the sunglint OLCI images. We show that both sigma0 and SLA yield realistic estimates for routine observation of ISWs with the SRAL, which is a significant improvement from previous observations recently reported for conventional pulse-limited altimeters (Jason-2). Several case studies of ISW signatures are interpreted in light of our knowledge of radar backscatter in the internal wave field. An analysis is presented for the tropical Atlantic Ocean off the Amazon shelf to infer the frequency of the phenomena, being consistent with previous satellite observations in the study region. Full article
(This article belongs to the Special Issue Radar Remote Sensing of Oceans and Coastal Areas)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Subset of an Envisat-ASAR (Advanced Synthetic Aperture Radar) image acquired in image precision mode dated 10 May 2005 at 03:29 UTC. The image shows a typical internal solitary wave (ISW) packet propagating northeast in the Andaman Sea (see inset in top-right corner for location). (<b>b</b>) ISW relative intensity retrieved from the black rectangle in the panel (<b>a</b>). (<b>c</b>) Schematic representation of ISW sea surface roughness patterns alongside typical altimetry footprints for Jason-2 and the synthetic aperture radar altimeter (SRAL). See text for more details.</p>
Full article ">Figure 2
<p>(<b>a</b>) Composite map of Sentinel-3A Ocean Land Colour Imager (OLCI) (partially shown in orange tones) with wave crests delineated as orange curves on top of a Terra Moderate Resolution Imaging Spectroradiometer (MODIS) acquisition (blue tones). The Sentinel-3A satellite track projected on the ground is marked with a red line while longitudes and latitudes are marked in white. The Jason-3 ground track is also shown in a black line. (<b>b</b>) Jason-3 along-track radar backscatter record for the Ku-band (20 Hz) in (dB) units (σ<sub>0</sub>). (<b>c</b>) Along-track SRAL record in backscatter (dB) units (σ<sub>0</sub>) for the Ku-band. The labels P1 to P4 indicate different ISW trains and latitudes are displayed on the top of the vertical grid. (<b>d</b>) Enlargement of image section in the latitude interval [8.0, 8.5]°N showing the surface manifestations of ISWs due to sunglint co-located with the SRAL backscatter record. Note that the backscatter record was displaced by 5.9 km to account for the internal wave propagation between the times of MODIS and Sentinel-3A acquisitions (see text for details).</p>
Full article ">Figure 3
<p>(<b>a</b>) Sentinel-3A OLCI image (Level 1b) dated 2017.10.11 acquired at 12:54:29 UTC (start time) presented in quasi-true color. The red line represents the Sentinel-3A track on the ground. (<b>b</b>) Radargram of SRAL composed of a sequence of 20 Hz waveforms. Backscatter power is represented in color with arbitrary units displayed in the color bar. (<b>c</b>) Averaged waveforms for perturbed (red) and unperturbed (blue) sea surface conditions.</p>
Full article ">Figure 4
<p>(<b>a</b>) Along-track backscatter (σ<sub>0</sub>) for Ku-band (black) and C-band (light blue). (<b>b</b>) Along-track SRAL record of SWH (significant wave height) showing a pronounced increase of retrieved SWH at approximately 4.3°N, from the average 2 m to 4 m. (<b>c</b>) Sea level anomaly (SLA) in SAR mode for the same ground segment where a sea surface elevation anomaly of approximately 0.4 m can be seen at the same location of the ISW signature. (<b>d</b>) Schematic representation of an interfacial ISW of depression showing a maximum of the surface water displacement over its trough, in between its rough and smooth surface manifestations. See text for more details.</p>
Full article ">Figure 5
<p>Selection of Sentinel-3A SRAL observations (Ku-band) over the tropical Atlantic Ocean off the Amazon shelf where signatures of surface manifestations of ISWs can be identified (in light blue rectangles, see also <a href="#remotesensing-10-00644-t001" class="html-table">Table 1</a> for details).</p>
Full article ">Figure 6
<p>Graphic representation of a sequence of 15 cycles for relative orbit number 95 of Sentinel-3A (SRAL sigma0) for a stretch over more than 300 km off the Amazon shelf in the tropical Atlantic Ocean. The green segments indicate detection of ISWs by the algorithm developed in this paper.</p>
Full article ">
22 pages, 4059 KiB  
Article
The Combined ASTER MODIS Emissivity over Land (CAMEL) Part 1: Methodology and High Spectral Resolution Application
by E. Eva Borbas, Glynn Hulley, Michelle Feltz, Robert Knuteson and Simon Hook
Remote Sens. 2018, 10(4), 643; https://doi.org/10.3390/rs10040643 - 21 Apr 2018
Cited by 43 | Viewed by 6910
Abstract
As part of a National Aeronautics and Space Administration (NASA) MEaSUREs (Making Earth System Data Records for Use in Research Environments) Land Surface Temperature and Emissivity project, the Space Science and Engineering Center (UW-Madison) and the NASA Jet Propulsion Laboratory (JPL) developed a [...] Read more.
As part of a National Aeronautics and Space Administration (NASA) MEaSUREs (Making Earth System Data Records for Use in Research Environments) Land Surface Temperature and Emissivity project, the Space Science and Engineering Center (UW-Madison) and the NASA Jet Propulsion Laboratory (JPL) developed a global monthly mean emissivity Earth System Data Record (ESDR). This new Combined ASTER (Advanced Spaceborne Thermal Emission and Reflection Radiometer) and MODIS (Moderate Resolution Imaging Spectroradiometer) Emissivity over Land (CAMEL) ESDR was produced by merging two current state-of-the-art emissivity datasets: the UW-Madison MODIS Infrared emissivity dataset (UW BF) and the JPL ASTER Global Emissivity Dataset Version 4 (GEDv4). The dataset includes monthly global records of emissivity and related uncertainties at 13 hinge points between 3.6–14.3 µm, as well as principal component analysis (PCA) coefficients at 5-km resolution for the years 2000 through 2016. A high spectral resolution (HSR) algorithm is provided for HSR applications. This paper describes the 13 hinge-points combination methodology and the high spectral resolutions algorithm, as well as reports the current status of the dataset. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>Left</b>) Advanced Spaceborne Thermal Emission and Reflection Radiometer Global Emissivity Dataset (ASTER GEDv3) mean summertime (July–September) emissivity for band 12 (9.1 µm), (<b>Right</b>) UW-Madison MODIS Infrared emissivity dataset Baseline Fit emissivity (UW BF) mean summertime (July–September) emissivity at 9.1 µm, which is the same as Moderate Resolution Imaging Spectroradiometer (MODIS) Band 29, 8.5 µm.</p>
Full article ">Figure 2
<p>Emissivity comparison on January 2003 over Sahara Desert (Lat = 25.075N, Lon = 26.058E) of UW BF (solid black line), UW high spectral resolution (HSR) (blue dots), the AIRS L2 (V5.0) Standard products (red line), and the UW/AIRS (green dots) emissivity products. (<b>Left</b>) The UW BF and UW HSR emissivity products have been derived from the Col 4, and (<b>Right</b>) from the Col 5 MODIS emissivity products.</p>
Full article ">Figure 3
<p>Time series of Col 4/4.1 (<b>left</b>) and Col 6 (<b>right</b>) band 20, 29, 31, and 32 MxD11C3 monthly mean emissivity for Aqua (red) and Terra (blue) MODIS over a Namib desert location (Lat: 24.25S, Lon:15.25E). Note: In Col 6, band 20 appears to be a copy of band 29 for both Terra and Aqua MODIS.</p>
Full article ">Figure 4
<p>Time series of the Combined ASTER and MODIS Emissivity over Land (CAMEL) database and its input UW BF dataset at 10.8 μm (<b>top panel</b>) and 12.1 μm (<b>bottom panel</b>) is shown over the Atmospheric Radiation Measurement (ARM) Southern Great Plains (SGP) Cart Site. The CAMEL algorithm effectively eliminates the emissivity degradation observed after 2009 for the long-wave MODIS MOD11 product based UW BF hinge-points.</p>
Full article ">Figure 5
<p>The CAMEL high spectral resolution (HSR) emissivity algorithm now includes three sets of laboratory spectra: (<b>a</b>) 55 selected spectra for general use (called version 8), (<b>b</b>) 82 spectra for surface types including carbonates (called version 10; version 8 + carbonates), and (<b>c</b>) four snow/ice selected spectra (version 12). The UW HSR 123 selected laboratory measurements (<b>d</b>) are shown for comparison.</p>
Full article ">Figure 6
<p>Making Earth System Data Records for Use in Research Environments (MEaSUREs) CAMEL Emissivity Earth System Data Record (ESDR) flowchart.</p>
Full article ">Figure 7
<p>The UW BF (<b>a</b>), CAMEL (<b>b</b>) and the ASTER GEDv4 (<b>c</b>) emissivity at 8.6 μm for February 2004. The difference maps of the emissivity between CAMEL and UW BF database (<b>d</b>) and ASTER GED v4 (<b>e</b>) are also shown for February 2004.</p>
Full article ">Figure 8
<p>The advantages of combining the ASTER GEDv4 and UW BF databases are evident here, showing the emissivity spectra over the Namib Desert, Namibia. UW BF emissivity for January 2004 (crosses) and hyperspectral fit (red line), the CAMEL 13 hinge-point emissivity (blue dots) and hyperspectral fit (blue line), and lab spectra (black) of sand samples collected over the Namib Desert. Note the improved spectral shape in CAMEL HSR (blue) in the quartz doublet regions between 8–10 μm and 12–13 μm.</p>
Full article ">Figure 9
<p>(<b>Left</b>) Percentage cumulative variance (PCV) function of the three selected laboratory measurement sets as a function of the number of principal components (PCs). The chosen number of PCs is indicated with blue or red stars. The legend contains the corresponding PCV values. Black stars stand for the number of PCVs, which reached the 0.999 value. (<b>Right</b>) The mean eigenvalues of the laboratory datasets for the first 10 eigenvectors.</p>
Full article ">Figure 10
<p>Emissivity on January 2004 at (<b>a</b>) Yemen, (<b>b</b>) Atmospheric Radiation Measurement (ARM) Southern Great Plains (SGP) Cart site, (<b>c</b>) Namib Desert, and (<b>d</b>) Greenland case sites. High spectral resolution emissivity from CAMEL with a different number of PCs used (different colored lines) and lab or Atmospheric Emitted Radiance Interferometer (AERI) measurements (black) are shown in the top row panels. The selected number of PCs is solid red for each case.</p>
Full article ">Figure 11
<p>The CAMEL dataset is currently available for the years 2003–2015, and makes use of the Aqua MODIS data as input to the 13 hinge-point product. The dataset is now extended to 2000, and uses the Terra MODIS data for the months of 2000 through December 2002. Time series of the CAMEL, ASTER GEDv4, UW BF (UW BF), and MODIS emissivity are shown, and demonstrate consistency between the Aqua and Terra products over the Rocky Mountain case site.</p>
Full article ">Figure 12
<p>IASI observed brightness temperatures are compared to those calculated using the Radiative Transfer for TOVS (RTTOV) UW IR emissivity module based on the UW BF emissivity database (black) and the CAMEL emissivity database (red) for the granule at 17:56 UTC, on 29 September 2008. The debiased variances are included over the 8–9 μm and 10–13 μm spectral regions.</p>
Full article ">Figure 13
<p>Same as <a href="#remotesensing-10-00643-f012" class="html-fig">Figure 12</a> but for the short IR spectral region (between 3.6–7 μm).</p>
Full article ">
17 pages, 2784 KiB  
Article
NWP-Based Adjustment of IMERG Precipitation for Flood-Inducing Complex Terrain Storms: Evaluation over CONUS
by Xinxuan Zhang, Emmanouil N. Anagnostou and Craig S. Schwartz
Remote Sens. 2018, 10(4), 642; https://doi.org/10.3390/rs10040642 - 21 Apr 2018
Cited by 15 | Viewed by 5343
Abstract
This paper evaluates the use of precipitation forecasts from a numerical weather prediction (NWP) model for near-real-time satellite precipitation adjustment based on 81 flood-inducing heavy precipitation events in seven mountainous regions over the conterminous United States. The study is facilitated by the National [...] Read more.
This paper evaluates the use of precipitation forecasts from a numerical weather prediction (NWP) model for near-real-time satellite precipitation adjustment based on 81 flood-inducing heavy precipitation events in seven mountainous regions over the conterminous United States. The study is facilitated by the National Center for Atmospheric Research (NCAR) real-time ensemble forecasts (called model), the Integrated Multi-satellitE Retrievals for GPM (IMERG) near-real-time precipitation product (called raw IMERG) and the Stage IV multi-radar/multi-sensor precipitation product (called Stage IV) used as a reference. We evaluated four precipitation datasets (the model forecasts, raw IMERG, gauge-adjusted IMERG and model-adjusted IMERG) through comparisons against Stage IV at six-hourly and event length scales. The raw IMERG product consistently underestimated heavy precipitation in all study regions, while the domain average rainfall magnitudes exhibited by the model were fairly accurate. The model exhibited error in the locations of intense precipitation over inland regions, however, while the IMERG product generally showed correct spatial precipitation patterns. Overall, the model-adjusted IMERG product performed best over inland regions by taking advantage of the more accurate rainfall magnitude from NWP and the spatial distribution from IMERG. In coastal regions, although model-based adjustment effectively improved the performance of the raw IMERG product, the model forecast performed even better. The IMERG product could benefit from gauge-based adjustment, as well, but the improvement from model-based adjustment was consistently more significant. Full article
(This article belongs to the Special Issue Remote Sensing of Precipitation)
Show Figures

Figure 1

Figure 1
<p>Location of the seven study regions over the conterminous United States (CONUS).</p>
Full article ">Figure 2
<p>Map of terrain elevation for the seven study regions (meters). DEM data are from the USGS Shuttle Radar Topography Mission (SRTM, <a href="https://lta.cr.usgs.gov/SRTM" target="_blank">https://lta.cr.usgs.gov/SRTM</a>). Data are available at <a href="https://dds.cr.usgs.gov/srtm/version2_1/SRTM3/North_America/" target="_blank">https://dds.cr.usgs.gov/srtm/version2_1/SRTM3/North_America/</a>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Quantile-quantile plot of the raw IMERG-L (Late run) product vs. the NCAR model product; (<b>b</b>) scatter plot of the values of parameters <span class="html-italic">a</span> and <span class="html-italic">b</span> for events in all regions.</p>
Full article ">Figure 4
<p>Error statistics of precipitation rate in Western Washington Region. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 5
<p>Error statistics of precipitation rate in Western Oregon Region. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 6
<p>Error statistics of precipitation rate in Northern and Central California Region. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 7
<p>Error statistics of precipitation rate in Southern California Region. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 8
<p>Error statistics of precipitation rate in Northern Idaho and Western Montana Region. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 9
<p>Error statistics of precipitation rate in Central Colorado Region. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 10
<p>Error statistics of precipitation rate in Southern Appalachians Regions. (<b>a</b>) Bias Ratio of frequency (BS), (<b>b</b>) Heidke Skill Score (HSS) and (<b>c</b>) Critical Success Index (CSI).</p>
Full article ">Figure 11
<p>Event total precipitation maps for selected storms. Top panel: event in Northern and Central California (start at 15 October 2016 18:00 UTC, 42-h length). Middle panel: event in Central Colorado (start at 7 May 2015 18:00 UTC, 84-h length). Bottom panel: event in Southern Appalachians (start at 29 September 2016 00:00 UTC, 48-h length).</p>
Full article ">
28 pages, 17324 KiB  
Review
On the Use of Unmanned Aerial Systems for Environmental Monitoring
by Salvatore Manfreda, Matthew F. McCabe, Pauline E. Miller, Richard Lucas, Victor Pajuelo Madrigal, Giorgos Mallinis, Eyal Ben Dor, David Helman, Lyndon Estes, Giuseppe Ciraolo, Jana Müllerová, Flavia Tauro, M. Isabel De Lima, João L. M. P. De Lima, Antonino Maltese, Felix Frances, Kelly Caylor, Marko Kohv, Matthew Perks, Guiomar Ruiz-Pérez, Zhongbo Su, Giulia Vico and Brigitta Tothadd Show full author list remove Hide full author list
Remote Sens. 2018, 10(4), 641; https://doi.org/10.3390/rs10040641 - 20 Apr 2018
Cited by 553 | Viewed by 39644
Abstract
Environmental monitoring plays a central role in diagnosing climate and management impacts on natural and agricultural systems; enhancing the understanding of hydrological processes; optimizing the allocation and distribution of water resources; and assessing, forecasting, and even preventing natural disasters. Nowadays, most monitoring and [...] Read more.
Environmental monitoring plays a central role in diagnosing climate and management impacts on natural and agricultural systems; enhancing the understanding of hydrological processes; optimizing the allocation and distribution of water resources; and assessing, forecasting, and even preventing natural disasters. Nowadays, most monitoring and data collection systems are based upon a combination of ground-based measurements, manned airborne sensors, and satellite observations. These data are utilized in describing both small- and large-scale processes, but have spatiotemporal constraints inherent to each respective collection system. Bridging the unique spatial and temporal divides that limit current monitoring platforms is key to improving our understanding of environmental systems. In this context, Unmanned Aerial Systems (UAS) have considerable potential to radically improve environmental monitoring. UAS-mounted sensors offer an extraordinary opportunity to bridge the existing gap between field observations and traditional air- and space-borne remote sensing, by providing high spatial detail over relatively large areas in a cost-effective way and an entirely new capacity for enhanced temporal retrieval. As well as showcasing recent advances in the field, there is also a need to identify and understand the potential limitations of UAS technology. For these platforms to reach their monitoring potential, a wide spectrum of unresolved issues and application-specific challenges require focused community attention. Indeed, to leverage the full potential of UAS-based approaches, sensing technologies, measurement protocols, postprocessing techniques, retrieval algorithms, and evaluation techniques need to be harmonized. The aim of this paper is to provide an overview of the existing research and applications of UAS in natural and agricultural ecosystem monitoring in order to identify future directions, applications, developments, and challenges. 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>Number of articles extracted from the database ISI-web of knowledge published from 1990 up to 2017 (last access 15 January 2018).</p>
Full article ">Figure 2
<p>A thermal survey over an Aglianico vineyard in the Basilicata region (southern Italy) overlaying an RGB orthophoto obtained by a multicopter mounted with both optical and FLIR Tau 2 cameras. Insets (<b>A</b>) and (<b>B</b>) provide magnified portions of the thermal map, where it is possible to distinguish vineyard rows (<b>B</b>) and surface temperature distribution on bare soil with a spot of colder temperature due to higher soil water content (<b>B</b>).</p>
Full article ">Figure 3
<p>Multi-spectral false colour (near infrared, red, green) imagery collected over the RoBo Alsahba date palm farm near Al Kharj, Saudi Arabia. Imagery (from L-R) shows the resolution differences between: (<b>A</b>) UAV mounted Parrot Sequoia sensor at 50 m height (0.05 m); (<b>B</b>) a WorldView-3 image (1.24 m); and (<b>C</b>) Planet CubeSat data (approx. 3 m), collected on the 13th, 29th and 27th March 2018, respectively.</p>
Full article ">Figure 4
<p>(<b>A</b>) A single RGB image of mangrove forest clearances, Matang Mangrove Forest Reserve, Malaysia, as observed using an RGB digital camera mounted on a DJI Phantom 3; (<b>B</b>) RGB orthomosaic from which individual (upper canopy) tree crowns can be identified as well as different mangrove species; and (<b>C</b>) the Canopy Height Model (CHM) derived from stereo RGB imagery, with darker green colors representing tall mangroves (typically &gt; 15 m) [<a href="#B121-remotesensing-10-00641" class="html-bibr">121</a>].</p>
Full article ">Figure 5
<p>Comparison of the most important aspects of UAS and satellite monitoring.</p>
Full article ">
12 pages, 3405 KiB  
Article
Influence of Tropical Instability Waves on Phytoplankton Biomass near the Marquesas Islands
by Elodie Martinez, Hirohiti Raapoto, Christophe Maes and Keitapu Maamaatuaihutapu
Remote Sens. 2018, 10(4), 640; https://doi.org/10.3390/rs10040640 - 20 Apr 2018
Cited by 8 | Viewed by 5017
Abstract
The Marquesas form an isolated group of small islands in the Central South Pacific where quasi-permanent biological activity is observed. During La Niña events, this biological activity, shown by a net increase of chlorophyll-a concentration (Chl, a proxy of phytoplankton biomass), is particularly [...] Read more.
The Marquesas form an isolated group of small islands in the Central South Pacific where quasi-permanent biological activity is observed. During La Niña events, this biological activity, shown by a net increase of chlorophyll-a concentration (Chl, a proxy of phytoplankton biomass), is particularly strong. It has been hypothesized that this strong activity is due to iron-rich waters advected from the equatorial region to the Marquesas by tropical instability waves (TIWs). Here we investigate this hypothesis over 18 years by combining satellite observations, re-analyses of ocean data, and Lagrangian diagnostics. Four La Niña events ranging from moderate to strong intensity occurred during this period, and our results show that the Chl plume within the archipelago can be indeed influenced by such equatorial advection, but this was observed during the strong 1998 and 2010 La Niña conditions only. Chl spatio-temporal patterns during the occurrence of other TIWs rather suggest the interaction of large-scale forcing events such as an uplift of the thermocline or the enhancement of coastal upwelling induced by the tropical strengthening of the trades with the islands leading to enhancement of phytoplankton biomass within the surface waters. Overall, whatever the conditions, our analyses suggest that the influence of the TIWs is to disperse, stir, and, therefore, modulate the shape of the existing phytoplankton plume. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The 1998–2014 annual average of chlorophyll-a concentrations (Chl, mg/m<sup>3</sup>) from the satellite-derived GlobColour Chl AVE product. The purple line delineates the French Polynesian Exclusive Economic Zone (EEZ).</p>
Full article ">Figure 2
<p>Annual average conditions calculated over the whole-time period for the Marquesas physical and biological environment. (<b>a</b>) Finite-size lyapunov exponents (FSLEs) (d<sup>−1</sup>); (<b>b</b>) Sea surface temperature (SST) (°C); (<b>c</b>) surface sigma (kg/m<sup>3</sup>); (<b>d</b>) Chl (mg/m<sup>3</sup>) and surface current (m/s). The islands are shown in black.</p>
Full article ">Figure 3
<p>Root mean square (RMS) of FLSEs in 1998. The black box delineates the area over which data are averaged to provide time-series of Figure 6. The islands are shown in black.</p>
Full article ">Figure 4
<p>(<b>Left</b>) SST (°C) and FSLEs (d<sup>−1</sup>, isocontours are plotted from 0.07 to 1 every 0.1); (<b>Centre</b>) surface sigma (kg/m<sup>3</sup>); (<b>Right</b>) Chl (mg/m<sup>3</sup>) and surface current (m/s) on 6, 14, and 22 September 1998 (top to bottom, respectively). The islands are shown in black.</p>
Full article ">Figure 5
<p>Annual RMS of FLSE for (<b>a</b>) a neutral year (as in 2003); and during La Niña events as in (<b>b</b>) 1999; (<b>c</b>) 2007; and (<b>d</b>) 2010.</p>
Full article ">Figure 6
<p>Time series of (<b>a</b>) El Niño 3–4 index provided by the Climate Prediction Center (CPC)/National Centers for Environmental Prediction (NCEP) services (the y-axis is inverted). Values lower than the −1 threshold (dash line) highlight moderate to strong La Niña years; (<b>b</b>) FSLE (d<sup>−1</sup>; blue line and left axis) and SST (°C; black line and right axis) monthly anomalies averaged over the Marquesas archipelago (11°S–18°S/142°W–138°W); (<b>c</b>) Number of particles launched from the northeastern equatorial area and reaching the Marquesas after 40, 60, and 90 days of drift (red, blue, and black lines, respectively); (<b>d</b>) Chl monthly anomalies (mg/m<sup>3</sup>) averaged over the same area as in (<b>b</b>).</p>
Full article ">Figure 7
<p>(<b>Left</b>) SST (°C) and FSLEs (d<sup>−1</sup>, isocontours are plotted from 0.07 to 1 every 0.1); (<b>Centre</b>) surface sigma (kg/m<sup>3</sup>); (<b>Right</b>) Chl (mg/m<sup>3</sup>) and surface current (m/s) during the five moderate to strong La Niña events over 1997–2014 (top to bottom). The islands are shown in black.</p>
Full article ">Figure 8
<p>Annual mean of eddy kinetic energy (EKE; cm²/s<sup>2</sup>) issued from (<b>a</b>) the Geostrophic and Ekman Current Observatory (GECKO) climatology with a ¼° spatial resolution [<a href="#B36-remotesensing-10-00640" class="html-bibr">36</a>] and (<b>b</b>) a climatological 1/45° resolution simulation from the Regional Ocean Modeling System (ROMS model) (see [<a href="#B37-remotesensing-10-00640" class="html-bibr">37</a>]); (<b>c</b>) Daily vorticity field (10<sup>−5</sup>/s<sup>1</sup>) at 10 m for 14 June of Year 6 from the ROMS climatological simulation; (<b>d</b>) Chl (mg/m<sup>3</sup>) for 20 July 2006, from the satellite-derived Chl AVE GlobColour product.</p>
Full article ">
25 pages, 14915 KiB  
Article
Application of Ground Penetrating Radar Supported by Mineralogical-Geochemical Methods for Mapping Unroofed Cave Sediments
by Teja Čeru, Matej Dolenec and Andrej Gosar
Remote Sens. 2018, 10(4), 639; https://doi.org/10.3390/rs10040639 - 20 Apr 2018
Cited by 10 | Viewed by 8333
Abstract
Ground penetrating radar (GPR) using a special unshielded 50 MHz Rough Terrain Antenna (RTA) in combination with a shielded 250 MHz antenna was used to study the capability of this geophysical method for detecting cave sediments. Allochthonous cave sediments found in the study [...] Read more.
Ground penetrating radar (GPR) using a special unshielded 50 MHz Rough Terrain Antenna (RTA) in combination with a shielded 250 MHz antenna was used to study the capability of this geophysical method for detecting cave sediments. Allochthonous cave sediments found in the study area of Lanski vrh (W Slovenia) are now exposed on the karst surface in the so-called “unroofed caves” due to a general lowering of the surface (denudation of carbonate rocks) and can provide valuable evidence of the karst development. In the first phase, GPR profiles were measured at three test locations, where cave sediments are clearly evident on the surface and appear with flowstone. It turned out that cave sediments are clearly visible on GPR radargrams as areas of strong signal attenuation. Based on this finding, GPR profiling was used in several other places where direct indicators of unroofed caves or other indicators for speleogenesis are not present due to strong surface reshaping. The influence of various field conditions, especially water content, on GPR measurements was also analysed by comparing radargrams measured in various field conditions. Further mineralogical-geochemical analyses were conducted to better understand the factors that influence the attenuation in the area of cave sediments. Samples of cave sediments and soils on carbonate rocks (rendzina) were taken for X-ray diffraction (XRD) and X-ray fluorescence (XRF) analyses to compare the mineral and geochemical compositions of both sediments. Results show that cave sediments contain higher amounts of clay minerals and iron/aluminium oxides/hydroxides which, in addition to the thickness of cave sediments, can play an important role in the depth of penetration. Differences in the mineral composition also lead to water retention in cave sediments even through dry periods which additionally contribute to increased attenuation with respect to surrounding soils. The GPR method has proven to be reliable for locating areas of cave sediments at the surface and to determine their spatial extent, which is very important in delineating the geometry of unroofed cave systems. GPR thus proved to be a very valuable method in supporting geological and geomorphological mapping for a more comprehensive recognition of unroofed cave systems. These are important for understanding karstification and speleogenetic processes that influenced the formation of former underground caves and can help us reconstruct the direction of former underground water flows. Full article
(This article belongs to the Special Issue Recent Advances in GPR Imaging)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area Lanski vrh and the basic geology [<a href="#B27-remotesensing-10-00639" class="html-bibr">27</a>] with the selected lithostratigraphic units of the narrow area on the LiDAR shaded relief image [<a href="#B28-remotesensing-10-00639" class="html-bibr">28</a>].</p>
Full article ">Figure 2
<p>Cave sediments and flowstone as typical indicators for speleogenesis: (<b>a</b>) Comparison of the colour difference between cave sediments and soils on carbonate rocks (rendzina); (<b>b</b>) Closer view of the cave sediments; (<b>c</b>) Cave sediments are accompanied with occurrences of flowstone in some places; (<b>d</b>) Outcrops of flowstone are often covered with moss, making it easier to recognize them compared to outcrops of limestones.</p>
Full article ">Figure 3
<p>Indirect indicators for the possible presence of cave sediments: (<b>a</b>) Cave sediments are sometimes associated with surface deepening, so the road must be frequently filled to keep it passable; (<b>b</b>) Increased water retention in sediments even during dry periods is evident in certain places. In the area below the increased water retention septarian concretions were found as one of the most reliable indicators of speleogenesis.</p>
Full article ">Figure 4
<p>Schematic model of unroofed cave development: (<b>a</b>) Underground horizontal passage formed during karstification. When the water table started dropping, cave sediments were deposited and speleothems began to form. Dry passage filled with allochthonous and autochthonous sediments; (<b>b</b>) Denudation processes cut off the passage’s upper parts, so cave sediments are revealed on the surface. Weathering and surface processes reshape the karst surface.</p>
Full article ">Figure 5
<p>Locations of ground penetrating radar (GPR) profiles at the investigated area on the LiDAR shaded relief image. Testing locations (blue rectangles) were chosen where cave sediments and flowstone occur on the surface.</p>
Full article ">Figure 6
<p>Processing steps (1–9) applied in profiles. Raw data and some processing steps are presented for Profile 1: (<b>a</b>) Raw data; (<b>b</b>) After dewow and time zero correction; (<b>c</b>) After background removal, manually-defined exponential gain correction, frequency and 2D filtering.</p>
Full article ">Figure 7
<p>(<b>a</b>) Direction of Profile 1 acquired over an area of cave sediments; (<b>b</b>) A processed radargram with highly attenuated area presenting cave sediments with 50 MHz antenna and (<b>c</b>) with 250 MHz antenna (vertical exaggeration 1:4); (<b>d</b>) Locations of V1 and V2 boreholes in cave sediments; (<b>e</b>) Sampling with a manual drilling machine in soils on carbonate rocks (V3).</p>
Full article ">Figure 8
<p>(<b>a</b>) Direction of perpendicular Profiles 2a and 2b recorded with a 50 and 250 MHz antenna (vertical exaggeration of the 250 MHz GPR images is 1:2) in the area of the cave sediments outcrop to determine their spatial extent; (<b>b</b>) Extent of cave sediments was determined in Profile 2a; (<b>c</b>) Boundaries of cave sediments detected by GPR are limited to the area of the outcrop visible on the surface.</p>
Full article ">Figure 9
<p>Surface topography in the vicinity of detected cave sediments: (<b>a</b>) An elongated depression at the location of Profile 1; (<b>b</b>) The surface morphology in the area of Profiles 2a and 2b, where depressions pass from one to another with intervening small ridges.</p>
Full article ">Figure 10
<p>Radargram acquired with a 250 MHz antenna, where the spatial extent of cave sediments was revealed along Profile 3.</p>
Full article ">Figure 11
<p>The GPR images obtained in various field conditions using the 250 MHz antenna and the area of cave sediments (dashed red lines): (<b>a</b>) Profile 1 acquired in July 2016 (very dry field conditions) and in March 2017 (dry field conditions); (<b>b</b>) Profile 2b acquired in March 2017 (dry field conditions) and in January 2018 (moderately wet field conditions).</p>
Full article ">Figure 12
<p>(<b>a</b>) The starting point and direction of Profile 4 with the surface depression which appears in the area of anomaly A detected by GPR. The surface deepening is also visible on the road; (<b>b</b>) The radargram conducted with the 50 MHz antenna with the detected anomaly A and attenuated areas B and C interpreted as cave sediments; (<b>c</b>) The radargram conducted with the 250 MHz antenna (note the vertical exaggeration 1:5) with a comparison of signal amplitudes, where the signal is strongly attenuated by the depth in the area of cave sediments even after the processing steps; (<b>d</b>) Surface morphology in the area of anomaly B, where the depression and the deepening on the road occur—the attenuated areas were interpreted as cave sediments; (<b>e</b>) The extent of cave sediments on the field as detected by GPR and the accompanying occurrences of the road deepening.</p>
Full article ">Figure 13
<p>(<b>a</b>) Roadwork, where slightly deepened areas need to be filled with coarse-grained material to keep the road in function; (<b>b</b>) Large outcrops of flowstone occur below the road; (<b>c</b>) A detail of flowstone structure; (<b>d</b>) In the radargram of Profile 5, the area of the possible presence of cave sediments is not as obvious as in other profiles.</p>
Full article ">Figure 14
<p>The radargram of Profile 6 with four detected attenuated areas. Flowstone only occurs on the surface at one location, so other locations were interpreted as “possible locations of cave sediments” (see legend). The attenuated area at the distance of 155–170 m corresponds to the direction of the cave sediments extent detected in Profile 3.</p>
Full article ">Figure 15
<p>X-ray diffraction (XRD) patterns of air-dried oriented samples with the determined minerals (Vrm-vermiculite, Chl-chlorite group minerals, Ms/Ill-muscovite/illite, Kln-kaolinite, Qz-quartz).</p>
Full article ">Figure 16
<p>PCA (Principal Component Analysis) diagram for the overall XRF and XRD datasets.</p>
Full article ">Figure 17
<p>Diagram shows the used methods with the main results, as needed to delineate unroofed cave systems into a whole.</p>
Full article ">Figure 18
<p>Profiles with interpreted anomalies detected by GPR and the occurrences of the superficial indicators of cave sediments.</p>
Full article ">
22 pages, 12316 KiB  
Article
Lithological Classification Using Sentinel-2A Data in the Shibanjing Ophiolite Complex in Inner Mongolia, China
by Wenyan Ge, Qiuming Cheng, Yunwei Tang, Linhai Jing and Chunsheng Gao
Remote Sens. 2018, 10(4), 638; https://doi.org/10.3390/rs10040638 - 20 Apr 2018
Cited by 78 | Viewed by 11460
Abstract
As a source of data continuity between Landsat and SPOT, Sentinel-2 is an Earth observation mission developed by the European Space Agency (ESA), which acquires 13 bands in the visible and near-infrared (VNIR) to shortwave infrared (SWIR) range. In this study, a Sentinel-2A [...] Read more.
As a source of data continuity between Landsat and SPOT, Sentinel-2 is an Earth observation mission developed by the European Space Agency (ESA), which acquires 13 bands in the visible and near-infrared (VNIR) to shortwave infrared (SWIR) range. In this study, a Sentinel-2A imager was utilized to assess its ability to perform lithological classification in the Shibanjing ophiolite complex in Inner Mongolia, China. Five conventional machine learning methods, including artificial neural network (ANN), k-nearest neighbor (k-NN), maximum likelihood classification (MLC), random forest classifier (RFC), and support vector machine (SVM), were compared in order to find an optimal classifier for lithological mapping. The experiment revealed that the MLC method offered the highest overall accuracy. After that, Sentinel-2A image was compared with common multispectral data ASTER and Landsat-8 OLI (operational land imager) for lithological mapping using the MLC method. The comparison results showed that the Sentinel-2A imagery yielded a classification accuracy of 74.5%, which was 2.5% and 5.08% higher than those of the ASTER and OLI imagery, respectively, indicating that Sentinel-2A imagery is adequate for lithological discrimination, due to its high spectral resolution in the VNIR to SWIR range. Moreover, different data combinations of Sentinel-2A + ASTER + DEM (digital elevation model) and OLI + ASTER + DEM data were tested on lithological mapping using the MLC method. The best mapping result was obtained from Sentinel-2A + ASTER + DEM dataset, demonstrating that OLI can be replaced by Sentinel-2A, which, when combined with ASTER, can achieve sufficient bandpasses for lithological classification. Full article
(This article belongs to the Collection Sentinel-2: Science and Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Simplified tectonic index map showing the position of Beishan orogenic belt, modified after Jolivet [<a href="#B29-remotesensing-10-00638" class="html-bibr">29</a>]. BH, Bayan Har; HK, Hindu Kush; Kh, Kohistan; Ku, Kudi; NQi, North Qiangtang; P, Pamir; Qi, Qilian Shan; SG, Songpan–Garze; SQi, South Qiangtang. (<b>b</b>) simplified geological map of the western Beishan orogenic belt, modified after Davis et al. [<a href="#B30-remotesensing-10-00638" class="html-bibr">30</a>]; and (<b>c</b>) geological map of the Shibanjing ophiolite.</p>
Full article ">Figure 2
<p>Flowchart of the lithological classification process.</p>
Full article ">Figure 3
<p>Lithological classification of the S2A_DEM dataset using machine learning methods. (<b>a</b>) <span class="html-italic">k</span>-nearest neighbor (<span class="html-italic">k</span>-NN); (<b>b</b>) random forest classifier (RFC); (<b>c</b>) artificial neural network (ANN); (<b>d</b>) support vector machine (SVM); (<b>e</b>) maximum likelihood classification (MLC).</p>
Full article ">Figure 3 Cont.
<p>Lithological classification of the S2A_DEM dataset using machine learning methods. (<b>a</b>) <span class="html-italic">k</span>-nearest neighbor (<span class="html-italic">k</span>-NN); (<b>b</b>) random forest classifier (RFC); (<b>c</b>) artificial neural network (ANN); (<b>d</b>) support vector machine (SVM); (<b>e</b>) maximum likelihood classification (MLC).</p>
Full article ">Figure 4
<p>(<b>a</b>) The lithological classification accuracies of each class; (<b>b</b>) overall accuracies; and (<b>c</b>) Kappa coefficient of the S2A_DEM dataset using different machine learning methods.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) The lithological classification accuracies of each class; (<b>b</b>) overall accuracies; and (<b>c</b>) Kappa coefficient of the S2A_DEM dataset using different machine learning methods.</p>
Full article ">Figure 5
<p>Lithological classification of different datasets using the MLC method. (<b>a</b>) S2A, (<b>b</b>) partial magnification of (<b>a</b>), (<b>c</b>) S2A_DEM, and (<b>d</b>) partial magnification of (<b>c</b>).</p>
Full article ">Figure 5 Cont.
<p>Lithological classification of different datasets using the MLC method. (<b>a</b>) S2A, (<b>b</b>) partial magnification of (<b>a</b>), (<b>c</b>) S2A_DEM, and (<b>d</b>) partial magnification of (<b>c</b>).</p>
Full article ">Figure 6
<p>The lithological classification accuracies of the S2A and S2A_DEM datasets using the MLC method.</p>
Full article ">Figure 7
<p>The overall accuracies and Kappa coefficients of datasets using the MLC method.</p>
Full article ">Figure 8
<p>MLC-generated lithological classifications of two different datasets. (<b>a</b>) OLI_DEM; and (<b>b</b>) AST_DEM.</p>
Full article ">Figure 9
<p>The classification accuracies of datasets of OLI + DEM and AST + DEM using the MLC method.</p>
Full article ">Figure 10
<p>Lithological classification of three different datasets using MLC. (<b>a</b>) The classification of OLI_AST_DEM; (<b>b</b>) the classification of S2A_AST_DEM; (<b>c</b>) partial magnification of quartz diorite in (<b>a</b>); (<b>d</b>) partial magnification of quartz diorite in (<b>b</b>); (<b>e</b>) partial magnification of quartz diorite in geological map; (<b>f</b>) partial magnification of basalt in (<b>a</b>); (<b>g</b>) partial magnification of basalt in (<b>b</b>); (<b>h</b>) partial magnification of basalt in geological map.</p>
Full article ">Figure 11
<p>The classification accuracies of the OLI + AST + DEM and S2A + AST + DEM datasets using the MLC method.</p>
Full article ">Figure 12
<p>The topographic map of the Shibanjing ophiolite obtained from DEM.</p>
Full article ">
21 pages, 4183 KiB  
Article
Reducing Uncertainty in Mapping of Mangrove Aboveground Biomass Using Airborne Discrete Return Lidar Data
by Francisca Rocha de Souza Pereira, Milton Kampel, Mário Luiz Gomes Soares, Gustavo Calderucio Duque Estrada, Cristina Bentz and Gregoire Vincent
Remote Sens. 2018, 10(4), 637; https://doi.org/10.3390/rs10040637 - 20 Apr 2018
Cited by 26 | Viewed by 7195
Abstract
Remote sensing techniques offer useful tools for estimating forest biomass to large extent, thereby contributing to the monitoring of land use and landcover dynamics and the effectiveness of environmental policies. The main goal of this study was to investigate the potential use of [...] Read more.
Remote sensing techniques offer useful tools for estimating forest biomass to large extent, thereby contributing to the monitoring of land use and landcover dynamics and the effectiveness of environmental policies. The main goal of this study was to investigate the potential use of discrete return light detection and ranging (lidar) data to produce accurate aboveground biomass (AGB) maps of mangrove forests. AGB was estimated in 34 small plots scatted over a 50 km2 mangrove forest in Rio de Janeiro, Brazil. Plot AGB was computed using either species-specific or non-species-specific allometric models. A total of 26 descriptive lidar metrics were extracted from the normalized height of the lidar point cloud data, and various model forms (random forest and partial least squares regression with backward selection of predictors (Auto-PLS)) were tested to predict the recorded AGB. The models developed using species-specific allometric models were distinctly more accurate (R2(calibration) = 0.89, R2(validation) = 0.80, root-mean-square error (RMSE, calibration) = 11.20 t·ha−1, and RMSE(validation) = 14.80 t·ha−1). The use of non-species-specific allometric models yielded large errors on a landscape scale (+14% or −18% bias depending on the allometry considered), indicating that using poor quality training data not only results in low precision but inaccuracy at all scales. It was concluded that under suitable sampling pattern and provided that accurate field data are used, discrete return lidar can accurately estimate and map the AGB in mangrove forests. Conversely this study underlines the potential bias affecting the estimates of AGB in other forested landscapes where only non-species-specific allometric equations are available. Full article
(This article belongs to the Special Issue Biomass Remote Sensing in Forest Landscapes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area located in (<b>a</b>) Brazil; (<b>b</b>) Rio de Janeiro State (the black box indicates the location of the Guanabara Bay); (<b>c</b>) the Northeast region of the Guanabara Bay showing the Guapimirim Environmental Protection Area (APA Guapimirim) and the Guanabara Ecological Station (ESEC Guanabara) with the locations of the 34 ground plots in yellow (WorldView2 image (band 5 in red, band 6 in green, and band 3 in blue) of 01 October 2012 -EPSG Projection: 32723).</p>
Full article ">Figure 2
<p>Histogram of aboveground biomass (ABG, t·ha<sup>−1</sup>) for: (<b>a</b>) species-specific equations (AGB_ssp) and the pantropical equation of Komiyama et al. [<a href="#B44-remotesensing-10-00637" class="html-bibr">44</a>]; (<b>b</b>) (AGB_K) and Chave et al. [<a href="#B56-remotesensing-10-00637" class="html-bibr">56</a>]; (<b>c</b>) (AGB_C) of 34 field plots scattered across the mangrove area inside the APA Guapimirim and ESEC Guanabara.</p>
Full article ">Figure 3
<p>Correlation plot for field-estimated AGB with species-specific equations (AGB_ssp) versus the pantropical equations of Chave et al. [<a href="#B56-remotesensing-10-00637" class="html-bibr">56</a>] (AGB_C) (<b>a</b>), and Komiyama et al. [<a href="#B44-remotesensing-10-00637" class="html-bibr">44</a>] (AGB_K) (<b>b</b>).</p>
Full article ">Figure 4
<p>Correlation between structural parameters and the lidar metrics for 34 field plots scattered over the mangrove area inside the APA Guapimirim and ESEC Guanabara. The AGB used is computed using species-specific allometric equations.</p>
Full article ">Figure 5
<p>Lidar Cloud point data for (<b>a</b>) Plot 25 (field AGB_ssp: 45.56 t·ha<sup>−1</sup>); (<b>b</b>) Plot 09 (field AGB_ssp: 126.37 t·ha<sup>−1</sup>); and (<b>c</b>) Plot 20 (field AGB_ssp: 186.10 t·ha<sup>−1</sup>).</p>
Full article ">Figure 6
<p>Scatter plots of predicted versus observed AGB (t·ha<sup>−1</sup>) for (<b>a</b>) M2sp.autopls model (Auto-PLS); (<b>b</b>) M2sp.rf model (RF); (<b>c</b>) M3K.autopls and; (<b>d</b>) M4C.autopls. R<sup>2</sup> for RF are comparable to R<sup>2</sup>(CAL) for Auto-PLS. R<sup>2</sup>(CAL) is the coefficient of determination in calibration.</p>
Full article ">Figure 7
<p>Mangrove aboveground biomass map (AGB in t·ha<sup>−1</sup>) of APA Guapimirim and ESEC Guanabara obtained using the best regression model, M2sp.autopls (Auto-PLS). Non-mangrove areas appear in black. The black lines delimit the APA Guapimirim (outside) and the ESEC Guanabara (inside). (EPSG Projection: 32723).</p>
Full article ">Figure 8
<p>Field plot distribution (in red) on a space plane of principal component 1 (PC1) versus principal component 2 (PC2). The grey circles represent the mangrove forest area and the black circles represent the non-mangrove area. The internal ellipse (solid line) corresponds to 95% confidence and the dashed ellipse to 98% confidence.</p>
Full article ">Figure 9
<p>(<b>a</b>) Probability density of average return height for mangrove area on a landscape level and (<b>b</b>) Probability density of average return height for 34 sample plots; height in meters.</p>
Full article ">
15 pages, 5877 KiB  
Article
Soil Moisture and Vegetation Water Content Retrieval Using QuikSCAT Data
by Shadi Oveisgharan, Ziad Haddad, Joe Turk, Ernesto Rodriguez and Li Li
Remote Sens. 2018, 10(4), 636; https://doi.org/10.3390/rs10040636 - 20 Apr 2018
Cited by 18 | Viewed by 4972
Abstract
Climate change and hydrological cycles can critically impact future water resources. Uncertainties in current climate models result in disagreement on the amount of water resources. Soil moisture and vegetation water content are key environmental variables on evaporation and transpiration at the land–atmosphere boundary. [...] Read more.
Climate change and hydrological cycles can critically impact future water resources. Uncertainties in current climate models result in disagreement on the amount of water resources. Soil moisture and vegetation water content are key environmental variables on evaporation and transpiration at the land–atmosphere boundary. Radar remote sensing helps to improve our estimate of water resources spatially and temporally. This work proposes a backscattered power formulation for the Ku-band. Li et al. (2010) retrieved soil moisture and vegetation water content values using Windsat data and simultaneous collocated QuikSCAT backscattered power are used to estimate different parameters of backscatter formulation. These parameters are used to estimate soil moisture and vegetation water content using QuikSCAT power everywhere and every day during the summer season. The 2-folded cross validation method is used to evaluate the performance of soil moisture and vegetation water content retrieval. A relatively large correlation is observed between vegetation water content using WindSat and QuikSCAT data in land classes of Evergreen Needleleaf, Evergreen Broadleaf, Deciduous Broadleaf, and Mixed Forests. Similarly, the retrieved soil moisture using QuikSCAT in areas with bare surface fraction of greater than 60% shows relatively high correlation with WindSat values. QuikSCAT satellite collects data over land globally almost every day. Therefore, QuikSCAT data can be used to generate a global map of soil moisture and vegetation water content daily from 2000 to 2009. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flow chart showing the steps for 2-folded cross-validation in retrieving soil moisture and vegetation water content.</p>
Full article ">Figure 2
<p>Spatial correlation of different parameters in our backscattered power formulation.</p>
Full article ">Figure 3
<p>Global map of different classes of the Earth’s surface. 0: Water, 1: Evergreen Needleleaf Forest, 2: Evergreen Broadleaf Forest, 3: Deciduous Needleleaf Forest, 4: Deciduous Broadleaf Forest, 5: Mixed Forest, 6: Woodland, 7: Wooded Grassland, 8: Closed Shrubland, 9: Open Shrubland, 10: Grassland, 11: Cropland, 12: Bare Ground, 13: Urban and Built classes are shown by different colors changing from dark blue to dark red.</p>
Full article ">Figure 4
<p>(<b>a)</b> land cover classes map of the U.S.; (<b>b</b>) 2001 bare surface fraction map of the U.S.; (<b>c</b>) estimated bare surface fraction map of the U.S.</p>
Full article ">Figure 5
<p>(<b>a</b>) land cover classes map of Australia; (<b>b</b>) 2001 bare surface fraction map of Australia; (<b>c</b>) estimated bare surface fraction map of Australia.</p>
Full article ">Figure 6
<p>2D logarithmic histogram of retrieved vegetation water content (kg/m<sup>2</sup>) using QuikSCAT backscattered power vs. estimated vegetation water content using WindSat data for all the points around the globe over month of summer 2006 for bare surface fraction of (<b>a</b>) 0 &lt; <span class="html-italic">f</span> &lt; 15%; (<b>b</b>) 15 &lt; <span class="html-italic">f</span> &lt; 30%; (<b>c</b>) 30 &lt; <span class="html-italic">f</span> &lt; 45%; (<b>d</b>) 45 &lt; <span class="html-italic">f</span> &lt; 60%; (<b>e</b>) 60 &lt; <span class="html-italic">f</span> &lt; 75%; (<b>f</b>) 75 &lt; <span class="html-italic">f</span> &lt; 85%; (<b>g</b>) 85 &lt; <span class="html-italic">f</span> &lt; 90%; (<b>h</b>) 90 &lt; <span class="html-italic">f</span> &lt; 100%.</p>
Full article ">Figure 7
<p>2D logarithmic histogram of retrieved soil moisture (m<sup>3</sup>/m<sup>3</sup>) using QuikSCAT backscattered power vs. estimated soil moisture using WindSat data for all the points around the globe over summer 2006 for bare surface fraction of (<b>a</b>) 0 &lt; <span class="html-italic">f</span> &lt; 15%; (<b>b</b>) 15 &lt; <span class="html-italic">f</span> &lt; 30%; (<b>c</b>) 30 &lt; <span class="html-italic">f</span> &lt; 45%; (<b>d</b>) 45 &lt; <span class="html-italic">f</span> &lt; 60%; (<b>e</b>) 60 &lt; <span class="html-italic">f</span> &lt; 75%; (<b>f</b>) 75 &lt; <span class="html-italic">f</span> &lt; 85%; (<b>g</b>) 85 &lt; <span class="html-italic">f</span> &lt; 90%; (<b>h</b>) 90 &lt; <span class="html-italic">f</span> &lt; 100%.</p>
Full article ">
13 pages, 17492 KiB  
Article
A Method for Robust Estimation of Vegetation Seasonality from Landsat and Sentinel-2 Time Series Data
by Per Jönsson, Zhanzhang Cai, Eli Melaas, Mark A. Friedl and Lars Eklundh
Remote Sens. 2018, 10(4), 635; https://doi.org/10.3390/rs10040635 - 19 Apr 2018
Cited by 102 | Viewed by 12451
Abstract
Time series from Landsat and Sentinel-2 satellites have great potential for modeling vegetation seasonality. However, irregular time sampling and frequent data loss due to clouds, snow, and short growing seasons, makes this modeling a challenge. We describe a new method for modeling seasonal [...] Read more.
Time series from Landsat and Sentinel-2 satellites have great potential for modeling vegetation seasonality. However, irregular time sampling and frequent data loss due to clouds, snow, and short growing seasons, makes this modeling a challenge. We describe a new method for modeling seasonal vegetation index dynamics from satellite time series data. The method is based on box constrained separable least squares fits to logistic model functions combined with seasonal shape priors. To enable robust estimates, we extract a base level (i.e., the minimum dormant season value) from the frequency distribution of clear-sky vegetation index values. A seasonal shape prior is computed from several years of data, and in the final fits local parameters are box constrained. More specifically, if enough data values exist in a certain time period, the corresponding local parameters determining the shape of the model function over this period are relaxed and allowed to vary freely. If there are no observations in a period, the corresponding local parameters are locked to the parameters of the shape prior. The method is flexible enough to model interannual variations, yet robust enough when data are sparse. We test the method with Landsat, Sentinel-2, and MODIS data over a forested site in Sweden, demonstrating the feasibility and potential of the method for operational modeling of growing seasons. Full article
(This article belongs to the Special Issue Land Surface Phenology )
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area in Central Sweden. The image to the left is a false color composite from Sentinel-2, acquired on 8 July 2016. The white line marks the area of the MODIS data used, and the blue line marks the area of the Sentinel-2 and Landsat data used.</p>
Full article ">Figure 2
<p>Landsat observations for 2006–2014 extracted for a pine forest pixel (60.0863°N, 17.4795°E) denoted as “Clear” or “Other” (i.e., assigned another QA class than clear-sky in the FMASK algorithm). Note the large proportion of non-clear observations and the weak seasonal dynamics for this pixel.</p>
Full article ">Figure 3
<p>Shape prior for time series data from Landsat NDVI 2000–2005. The base level has, after analyzing the histogram for this pixel, been fixed to NDVI = 0.59. As the shape prior does not describe individual years, its main use is for stabilizing the fitting procedure during data-sparse periods.</p>
Full article ">Figure 4
<p>(<b>a</b>) Double logistic fits with free seasonal parameters. Note the unrealistically short second growing season due to lack of clear observations at the end of the season (arrow); (<b>b</b>) fit where the right inflexion point and the parameter determining the fall time are constrained and taken as the corresponding values of the shape prior.</p>
Full article ">Figure 5
<p>Seven regions in which data points must exist, according to <a href="#remotesensing-10-00635-t002" class="html-table">Table 2</a>, to allow free parameters to be used. Circles denote levels 0.01, 0.25, 0.75 and 0.99 of the amplitude to the left and right of the center.</p>
Full article ">Figure 6
<p>Examples of time series over deciduous (<b>top</b>), coniferous (<b>middle</b>), and agricultural (<b>bottom</b>) areas from Landsat (<b>left</b>) and S2 (<b>right</b>).</p>
Full article ">Figure 7
<p>Phenology data from Landsat (<b>top row</b>) and Sentinel 2a (<b>bottom row</b>). Left hand images show estimated start of season (unit: day-of-year, DOY), and the center images show zoom-ins over an agricultural area. The right hand top image shows a false color composite (FCC) from Landsat 8 for comparison, and the bottom right image shows pixels where shape prior was used for estimating parameter <math display="inline"> <semantics> <mrow> <msubsup> <mi>x</mi> <mn>1</mn> <mi>i</mi> </msubsup> </mrow> </semantics> </math>, determining the start of season. Coordinates of the study area are shown in <a href="#remotesensing-10-00635-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 8
<p>Reduced RMSE and bias when estimating start of season (<b>left</b>) and end of season (<b>right</b>) for 2016 from simulated S2 NDVI data by double logistic function fitting without shape prior (SP; blue dots) as compared to fitting with shape prior (red dots). Parameters from simulated S2 data are plotted against reference data of SOS and EOS from daily MODIS NDVI. Equations and statistics of the linear relationships are printed in the graph in blue (no SP) and red (with SP).</p>
Full article ">
22 pages, 3247 KiB  
Article
Model Selection for Parametric Surfaces Approximating 3D Point Clouds for Deformation Analysis
by Xin Zhao, Boris Kargoll, Mohammad Omidalizarandi, Xiangyang Xu and Hamza Alkhatib
Remote Sens. 2018, 10(4), 634; https://doi.org/10.3390/rs10040634 - 19 Apr 2018
Cited by 25 | Viewed by 5775
Abstract
Deformation monitoring of structures is a common application and one of the major tasks of engineering surveying. Terrestrial laser scanning (TLS) has become a popular method for detecting deformations due to high precision and spatial resolution in capturing a number of three-dimensional point [...] Read more.
Deformation monitoring of structures is a common application and one of the major tasks of engineering surveying. Terrestrial laser scanning (TLS) has become a popular method for detecting deformations due to high precision and spatial resolution in capturing a number of three-dimensional point clouds. Surface-based methodology plays a prominent role in rigorous deformation analysis. Consequently, it is of great importance to select an appropriate regression model that reflects the geometrical features of each state or epoch. This paper aims at providing the practitioner some guidance in this regard. Different from standard model selection procedures for surface models based on information criteria, we adopted the hypothesis tests from D.R. Cox and Q.H. Vuong to discriminate statistically between parametric models. The methodology was instantiated in two numerical examples by discriminating between widely used polynomial and B-spline surfaces as models of given TLS point clouds. According to the test decisions, the B-spline surface model showed a slight advantage when both surface types had few parameters in the first example, while it performed significantly better for larger numbers of parameters. Within B-spline surface models, the optimal one for the specific segment was fixed by Vuong’s test whose result was quite consistent with the judgment of widely used Bayesian information criterion. The numerical instabilities of B-spline models due to data gap were clearly reflected by the model selection tests, which rejected inadequate B-spline models in another numerical example. Full article
(This article belongs to the Special Issue 3D Modelling from Point Clouds: Algorithms and Methods)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sketch map of the experimental design concerning the locations of the instruments and relevant targets in side view (<b>upper</b>) and top view (<b>bottom</b>) [<a href="#B16-remotesensing-10-00634" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>Reflectance image generated by reflectivity values of TLS data [<a href="#B16-remotesensing-10-00634" class="html-bibr">16</a>].</p>
Full article ">Figure 3
<p>Extracted Arc-shape object and the target segments in our numerical example (within the red boundary) shown by the software CloudCompare.</p>
Full article ">Figure 4
<p>Histogram of the sampled log-likelihood ratio <math display="inline"><semantics> <msub> <mi>L</mi> <mrow> <mn>0</mn> <mo>,</mo> <mn>1</mn> </mrow> </msub> </semantics></math> under the polynomial (<b>left</b>) and B-spline (<b>right</b>) surface model, approximated by a Gaussian density functions (in red).</p>
Full article ">Figure 5
<p>Statistic values of Vuong’s test in comparison with critics.</p>
Full article ">Figure 6
<p>Side-view (<b>a</b>) and top-view (<b>b</b>) of approximated B-spline surface (<math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>18</mn> <mo>,</mo> <mi>m</mi> <mo>=</mo> <mn>18</mn> </mrow> </semantics></math>) with the original measurements (blue points).</p>
Full article ">Figure 7
<p>Polynomial (<b>a</b>,<b>c</b>) and B-spline surface models (<b>b</b>,<b>d</b>) in terms of differences of the 1st and 13th epochs in Segment I.</p>
Full article ">Figure 8
<p>Deformation of segment I reflected by block means of the point cloud differences based on the 1st and 13th epochs.</p>
Full article ">Figure 9
<p>Deformation of Segment I between 1st and 13th epochs reflected by various approaches.</p>
Full article ">Figure 10
<p>Polynomial (<b>a</b>,<b>c</b>) and B-spline surface models (<b>b</b>,<b>d</b>) in reflecting deformation of segment II based on the 1st and 13th epochs.</p>
Full article ">Figure 11
<p>Deformation of Segment II reflected by block means of the point cloud differences based on the 1st and 13th epochs.</p>
Full article ">Figure 12
<p>AIC (<b>red</b>) and BIC (<b>green</b>) values with an increasing number of parameters.</p>
Full article ">
20 pages, 11622 KiB  
Article
Downscaling of ASTER Thermal Images Based on Geographically Weighted Regression Kriging
by Osvaldo José Ribeiro Pereira, Adolpho José Melfi, Célia Regina Montes and Yves Lucas
Remote Sens. 2018, 10(4), 633; https://doi.org/10.3390/rs10040633 - 19 Apr 2018
Cited by 27 | Viewed by 5128
Abstract
The lower spatial resolution of thermal infrared (TIR) satellite images and derived land surface temperature (LST) is one of the biggest challenges in mapping temperature at a detailed map scale. An extensive range of scientific and environmental applications depend on the availability of [...] Read more.
The lower spatial resolution of thermal infrared (TIR) satellite images and derived land surface temperature (LST) is one of the biggest challenges in mapping temperature at a detailed map scale. An extensive range of scientific and environmental applications depend on the availability of fine spatial resolution temperature data. All satellite-based sensor systems that are equipped with a TIR detector depict a spatial resolution that is coarser than most of the multispectral bands of the same system. Certain studies may therefore be not feasible if applied in areas that depict a high spatial variation in temperature at small spatial scales, such as urban centers and flooded pristine areas. To solve this problem, this study applied an image downscaling method to enhance the spatial resolution of LST data by combining TIR, multispectral images, and derived data, such as Normalized Difference Vegetation Index (NDVI), according to the geographically weighted regression (GWRK) and area-to-point kriging of regressed residuals. The resulting LST images of the natural and anthropogenic urban areas of the Brazilian Pantanal are very highly correlated to the reference LST images. The approach, combining ASTER TIR with ASTER visible/infrared (VNIR) and Sentinel-2 images according to the GWRK method, performed better than all of the remaining state-of-the-art downscaling methods. Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area in the context of the Brazilian portion of the Pantanal biome.</p>
Full article ">Figure 2
<p>Flowchart of the methodology applied to evaluate the quality of the land surface temperature (LST) downscaled image according to the different downscaling methods adopted in this study.</p>
Full article ">Figure 3
<p>Visual results of the merged images at a 90-m spatial resolution in a sector of the Campo Grande study area. The reference (<b>a</b>) image (mean of ASTER bands 13 and 14) as compared to the following methods: (<b>b</b>) ATPRK (Area-To-Point Regression Kriging); (<b>c</b>) DCK (Downscaling Cokriging); (<b>d</b>) GWRK (Geographically Weighted Regression Kriging); (<b>e</b>) ATWT (“À Trous” Wavelet Transform); (<b>f</b>) GS (Gram-Schmidt); (<b>g</b>) MDMR (MultiDirection-MultiResolution); (<b>h</b>) PCA (Principal Component Analysis); and (<b>i</b>) TSHARP.</p>
Full article ">Figure 4
<p>Visual results of the merged images at a 90-m spatial resolution in a sector of the Nhecolândia study area. The reference (<b>a</b>) image (mean of ASTER bands 13 and 14) as compared to the following methods: (<b>b</b>) ATPRK; (<b>c</b>) DCK; (<b>d</b>) GWRK; (<b>e</b>) ATWT; (<b>f</b>) GS; (<b>g</b>) MDMR; (<b>h</b>) PCA; and (<b>i</b>) TSHARP.</p>
Full article ">Figure 5
<p>Box-plots showing the fusion quality values for the Campo Grande region.</p>
Full article ">Figure 6
<p>Box-plots showing the merging quality values for the Nheoclândia region.</p>
Full article ">Figure 7
<p>Results of GWRK pan-sharpening for the Campo Grande study area: (<b>a</b>) Pixel-based comparison between the original 90-m TIR ASTER bands (mean of bands 13 and 14) and the merged GWRK image obtained from the degraded TIR and VNIR ASTER bands; (<b>b</b>) original TIR ASTER band; (<b>c</b>) result of GWRK pan-sharpening of TIR bands without degradation with a 15-m spatial resolution. Lower and higher temperatures range from blue to red, respectively.</p>
Full article ">Figure 8
<p>Results of GWRK pan-sharpening for the Nhecolândia study area: (<b>a</b>) Pixel-based comparison between the original 90-m TIR ASTER band (mean bands 13 and 14) and the merged GWRK image obtained from the degraded TIR and Sentinel-2 bands; (<b>b</b>) original TIR ASTER; (<b>c</b>) result of GWRK pan-sharpening of TIR bands without degradation, with a 20-m spatial resolution. Lower and higher temperatures range from blue to red, respectively.</p>
Full article ">Figure 9
<p>Results of GWR (Geographically Weighted Regression) with one and multiple covariates. Frequency distribution of the coefficient of correlation of the surfaces regressed by GWR (ancillary and TIR) within each local window without kriging in the (<b>a</b>) Campo Grande and (<b>b</b>) Nhecolândia study areas.</p>
Full article ">
19 pages, 6508 KiB  
Article
Woody Cover Estimates in Oklahoma and Texas Using a Multi-Sensor Calibration and Validation Approach
by Kyle A. Hartfield and Willem J. D. Van Leeuwen
Remote Sens. 2018, 10(4), 632; https://doi.org/10.3390/rs10040632 - 19 Apr 2018
Cited by 7 | Viewed by 4400
Abstract
Woody cover encroachment/expansion/conversion is a complex phenomenon that has environmental and economic impacts around the world. This research demonstrates the development of highly accurate models for estimating percent woody cover using high spatial resolution image data in combination with multi-seasonal Landsat reflectance products. [...] Read more.
Woody cover encroachment/expansion/conversion is a complex phenomenon that has environmental and economic impacts around the world. This research demonstrates the development of highly accurate models for estimating percent woody cover using high spatial resolution image data in combination with multi-seasonal Landsat reflectance products. We use a classification and regression tree (CART) approach to classify woody cover using fine resolution multispectral National Agricultural Imaging Program (NAIP) data. A continuous classification and regression tree (Cubist) ingests the aggregated woody cover classification along with the seasonal Landsat data to create a continuous woody cover model. We applied the models, derived by Cubist, across several Landsat scenes to estimate the percentage of woody plant cover, within each Landsat pixel, over a larger regional extent. We measured an average absolute error of 12.1 percent and a correlation coefficient of 0.78 for the models performed. The method of modelling percent woody cover established in this manuscript outperforms currently available woody cover estimates including Landsat Vegetation Continuous Fields (VCF), on average by 26 percent, and Web-Enabled Landsat Data (WELD) products, on average by 16 percent, for the region of interest. Current woody cover products are also limited to certain years and not available pre-2000. This manuscript describes a novel Cubist-based technique to model woody cover for any area of the world, as long as fine (~1–2 m) spatial resolution and Landsat data are available. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Edwards Plateau (blue) and the Rolling Red Plains (green) are the two ecoregions of focus in this study. Demonstration of the proposed methodology uses ten counties (orange) within nine Landsat tiles (white; with Path (P) and Row(R)).</p>
Full article ">Figure 2
<p>Flowchart summarizing the processing steps and training of a continuous fractional woody cover model using a CART-based thematic high resolution woody cover classification. Cubist extrapolates the estimated woody over large areas using multi-temporal Landsat data.</p>
Full article ">Figure 3
<p>Stratified Cluster Sampling Strategy employed to classify 2004 3-band NAIP data for Sutton County, Texas.</p>
Full article ">Figure 4
<p>2004 NAIP binary classification for Sutton County, Texas. Woody cover is green, while all other cover is tan.</p>
Full article ">Figure 5
<p>Estimated woody cover for 2004 within Edwards Plateau modelled using the Schleicher County- and Menard County-based Cubist models.</p>
Full article ">Figure 6
<p>Estimated woody cover for 2004 within an area of the Rolling Red Plains modelled using the Ellis County-, Comanche County-, and Payne County-based Cubist models.</p>
Full article ">
19 pages, 3174 KiB  
Article
Icing Detection over East Asia from Geostationary Satellite Data Using Machine Learning Approaches
by Seongmun Sim, Jungho Im, Sumin Park, Haemi Park, Myoung Hwan Ahn and Pak-wai Chan
Remote Sens. 2018, 10(4), 631; https://doi.org/10.3390/rs10040631 - 19 Apr 2018
Cited by 25 | Viewed by 6705
Abstract
Even though deicing or airframe coating technologies continue to develop, aircraft icing is still one of the critical threats to aviation. While the detection of potential icing clouds has been conducted using geostationary satellite data in the US and Europe, there is not [...] Read more.
Even though deicing or airframe coating technologies continue to develop, aircraft icing is still one of the critical threats to aviation. While the detection of potential icing clouds has been conducted using geostationary satellite data in the US and Europe, there is not yet a robust model that detects potential icing areas in East Asia. In this study, we proposed machine-learning-based icing detection models using data from two geostationary satellites—the Communication, Ocean, and Meteorological Satellite (COMS) Meteorological Imager (MI) and the Himawari-8 Advanced Himawari Imager (AHI)—over Northeast Asia. Two machine learning techniques—random forest (RF) and multinomial log-linear (MLL) models—were evaluated with quality-controlled pilot reports (PIREPs) as the reference data. The machine-learning-based models were compared to the existing models through five-fold cross-validation. The RF model for COMS MI produced the best performance, resulting in a mean probability of detection (POD) of 81.8%, a mean overall accuracy (OA) of 82.1%, and mean true skill statistics (TSS) of 64.0%. One of the existing models, flight icing threat (FIT), produced relatively poor performance, providing a mean POD of 36.4%, a mean OA of 61.0, and a mean TSS of 9.7%. The Himawari-8 based models also produced performance comparable to the COMS models. However, it should be noted that very limited PIREP reference data were available especially for the Himawari-8 models, which requires further evaluation in the future with more reference data. The spatio-temporal patterns of the icing areas detected using the developed models were also visually examined using time-series satellite data. Full article
(This article belongs to the Special Issue Remote Sensing Methods and Applications for Traffic Meteorology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A flow chart of the proposed machine-learning-based icing detection approach.</p>
Full article ">Figure 2
<p>An example of icing and non-icing sample extraction from COMS data. Preprocessed images are masked using the threshold at the infrared-1 (IR1) channel and a spatial buffer. In this example, icing was reported on 5 July 2011 23:50 UTC at N37.42° and E129.2° and non-icing was reported on 5 July 2011 23:48 UTC at N36.58° and E130.25°.</p>
Full article ">Figure 3
<p>Relative variable importance identified by machine learning models. Mean decrease accuracy is calculated using out-of-bag samples when a variable is perturbed by random forest (RF). Absolute normalized coefficients are considered as a variable importance in the multinomial log-linear (MLL) models. The larger the value of a variable, the more significant the variable. The most significant (top 20%) variables are displayed in orange.</p>
Full article ">Figure 4
<p>Accuracy assessment results of each icing detection model. (POD = Probability of detection; POFD = Probability of false detection; OA = Overall accuracy; TSS = True skill statistics; E = Standard error).</p>
Full article ">Figure 5
<p>Maps of potential icing areas produced by the six algorithms for the COMS and Himawari-8 images collected on 24 April 2016 01:00 UTC. The 11 μm channel Tb image ranging from 228 K to 280 K is used as the background image. Areas with no cloud or a Tb of 11 μm &gt; 270 K were masked out and appear black in the figures.</p>
Full article ">Figure 6
<p>Maps of potential icing areas produced by the six algorithms for the COMS and Himawari-8 images collected on 26 April 2016 01:00 UTC. The 11 μm channel Tb image ranging from 228 K to 280 K is used as the background image. Areas with no cloud or a Tb of 11 μm &gt; 270 K were masked out and appear black in the figures.</p>
Full article ">Figure 7
<p>Time-series of potential icing areas when the COMS RF model was applied for two icing cases (i.e., the upper images on 6 December 2015 04:15 UTC with an icing pilot report (PIREP) at E22.19° and N114.60° and the bottom images on 5 January 2016 03:30 UTC with an icing PIREP at E21.58° and N115.41°). The corresponding 11 μm channel images were used as the background.</p>
Full article ">Figure 8
<p>Time-series of potential icing areas when the Himawari-8 RF model was applied for two icing cases (i.e., the upper images on 6 December 2015 04:15 UTC at E22.19° and N114.60° and the bottom images on 5 January 2016 03:30 UTC at E21.58° and N115.41°). The corresponding 11 μm channel images were used as the background.</p>
Full article ">
16 pages, 10301 KiB  
Article
Vegetation Response to the 2012–2014 California Drought from GPS and Optical Measurements
by Eric E. Small, Carolyn J. Roesler and Kristine M. Larson
Remote Sens. 2018, 10(4), 630; https://doi.org/10.3390/rs10040630 - 19 Apr 2018
Cited by 16 | Viewed by 6242
Abstract
We compare microwave GPS and optical-based remote sensing observations of the vegetation response to a recent drought in California, USA. The microwave data are based on reflected GPS signals that were collected by a geodetic network. These data are sensitive to temporal variations [...] Read more.
We compare microwave GPS and optical-based remote sensing observations of the vegetation response to a recent drought in California, USA. The microwave data are based on reflected GPS signals that were collected by a geodetic network. These data are sensitive to temporal variations in vegetation water content and are made available via the Normalized Microwave Reflection Index (NMRI). NMRI data are complementary to information of plant greenness provided by the Normalized Difference Vegetation Index (NDVI). NMRI data from 146 sites in California are compared to collocated NDVI observations, over the interval of 2007–2016. This period includes a severe, three-year drought (2012–2014). We quantify the seasonal variations in vegetation state by calculating a series of phenology metrics at each site, using both NMRI and NDVI. We examine how the phenology metrics vary from year-to-year, as related to the observed fluctuations in accumulated precipitation. The amplitude of seasonal vegetation growth exhibits the greatest sensitivity to prior accumulated precipitation. Above-normal precipitation from 4 to 12 months before peak growth yields a stronger seasonal growth pulse, and vice versa. The amplitude of seasonal growth, as determined from NDVI, varies linearly with precipitation during dry years, but is largely insensitive to precipitation amount in years with above-normal precipitation. In contrast, the amplitude of seasonal growth from NMRI varies approximately linearly with precipitation across the entire range of conditions observed. The length of season is positively correlated with prior accumulated precipitation, more strongly with NDVI than NMRI. The recovery from drought was similar for a one-year (2007) and the more severe three-year drought (2012–2014). In both cases, the amplitude of growth returned to typical values in the first year with near-normal precipitation. Growing season length, only based on NDVI, was greatly reduced in 2014, the driest and final year of the three-year California drought. 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>Location of the 146 GPS stations used in this study. The background represents the MODIS IGBP land cover classification. The white rectangle is the region chosen for the regional-average analysis. The inset shows the number of Normalized Microwave Reflection Index (NMRI) sites by land cover type (GRA grassland, SHR shrubland, WSV woody savanna, SAV savanna, CRP cropland, FOR forest).</p>
Full article ">Figure 2
<p>Time series for GPS site P531: (<b>a</b>) NMRI time series after application of a five-day centered median filter; (<b>b</b>) Normalized Difference Vegetation Index (NDVI) data including linear interpolation between observed values. (<b>a</b>,<b>b</b>) data analysis: Green horizontal lines show the baseline level for amplitude calculation, determined separately between peaks. The lavender (yellow) circles for NMRI (NDVI) show the interval above the 25% cut-off around the main seasonal peak. The red vertical dashed line is the Mean Peak day of year. (<b>c</b>) North American Land Data Assimilation System (NLDAS) cumulative precipitation for each Water Year (red), and the 10-year average (gray). P531 is a grasslands site with an annual precipitation of 428 mm at latitude/longitude of N 35.8° and W 120.54°.</p>
Full article ">Figure 3
<p>Example analysis used to retrieve the amplitude of the seasonal peak, the Start of Season (SOS) and End of Season (EOS) day of year. The seasonal peak is within the Water Year (thick light blue line on <span class="html-italic">x</span>-axis labeled WY). The baseline (green line B) on each side of the peak is the level of pre/post-growth vegetation. The amplitude (black line A) is the difference between the peak and the average baseline pre/post-growth. The red dotted lines are the 25% cut off levels with respect to the pre/post-growth baselines, and they intersect the main seasonal growth cycle at the SOS and EOS day of year (red squares). (<b>a</b>) the daily five-day median NMRI uses the 30th percentile to define the baseline; and, (<b>b</b>) the interpolated bi-weekly NDVI uses the 20th percentile.</p>
Full article ">Figure 4
<p>Histograms of (<b>a</b>) Length of season; (<b>b</b>) Mean peak day of year; (<b>c</b>) Relative amplitude; and, (<b>d</b>) Deviation from the mean peak day of year. Histograms are for all sites over all years.</p>
Full article ">Figure 5
<p>Correlation between the Percent of Normal Precipitation (PNP) computed for the month prior to the NMRI mean seasonal peak day of year (DOY) and NMRI metrics as a function of the PNP scale in months (blue circles). (<b>a</b>) Relative Amplitude; (<b>b</b>) Deviation from the Mean Peak DOY; and, (<b>c</b>) Length of Season. The dashed lines are the 95% confidence interval. The green triangles are the correlation results using NDVI, with PNP computed for the month prior to the NDVI mean seasonal peak.</p>
Full article ">Figure 6
<p>NMRI (<b>a</b>) and NDVI (<b>b</b>) relative amplitude (small dots) versus the corresponding 6-month Percent of Normal Precipitation (PNP). The large symbols are the averaged values over 40% wide PNP bins incremented every 20%. Each bin has a number of points &gt;50.</p>
Full article ">Figure 7
<p>NMRI (<b>left</b>) and NDVI (<b>middle</b>) relative peak amplitude for Plate Boundary Observatory (PBO) H<sub>2</sub>O sites in California, during two dry years (2007 and 2014), a wet year (2011), and a normal year (2015) following a multi-year drought; (<b>Right</b>) six-month Percent of Normal Precipitation (PNP). Both vegetation metrics and PNP are compared to the average for 2007–2016 and reported as a percentage, as shown in the color bar.</p>
Full article ">Figure 8
<p>Central California averaged (inset <a href="#remotesensing-10-00630-f001" class="html-fig">Figure 1</a>, 55 sites) phenology metrics and one standard deviation as a function of year during 2007–2016, using NDVI (green triangles) and NMRI (blue circles); (<b>a</b>) Relative Amplitude; (<b>b</b>) Length of Season; and, (<b>c</b>) Deviation in time of peak growth. The averaged PNP is displayed in all plots (pink squares), with one standard deviation shown only in (<b>a</b>).</p>
Full article ">
19 pages, 15447 KiB  
Article
Topography and Three-Dimensional Structure Can Estimate Tree Diversity along a Tropical Elevational Gradient in Costa Rica
by Chelsea Robinson, Sassan Saatchi, David Clark, Johanna Hurtado Astaiza, Anna F. Hubel and Thomas W. Gillespie
Remote Sens. 2018, 10(4), 629; https://doi.org/10.3390/rs10040629 - 18 Apr 2018
Cited by 11 | Viewed by 5780
Abstract
This research seeks to understand how tree species richness and diversity relates to field data (1-ha plots) on forest structure (stems, basal area) and lidar derived data on topography and three-dimensional forest structure along an elevational gradient in Braulio Carrillo National Park, Costa [...] Read more.
This research seeks to understand how tree species richness and diversity relates to field data (1-ha plots) on forest structure (stems, basal area) and lidar derived data on topography and three-dimensional forest structure along an elevational gradient in Braulio Carrillo National Park, Costa Rica. In 2016 we calculated tree species richness and diversity indices for twenty 1-ha plots located along a gradient ranging from 56 to 2814 m in elevation. Field inventory data were combined with large footprint (20 m) airborne lidar data over plots in 2005, in order to quantify variations in topography and three-dimensional structure across plots and landscapes. A distinct pattern revealing an increase in species’ richness and the Shannon diversity index was observed in correlation with increasing elevation, up to about 600 m; beyond that, at higher elevations, a decrease was observed. Stem density and basal area both peaked at the 2800 m site, with a mini-peak at 600 m, and were both negatively associated with species richness and diversity. Species richness and diversity were negatively correlated with elevation, while the two tallest relative height metrics (rh100, rh75) derived from lidar were both significantly positively correlated with species richness and diversity. The best lidar-derived topographical and three-dimensional forest structural models showed a strong relationship with the Shannon diversity index (r2 = 0.941, p < 0.01), with ten predictors; conversely, the best species richness model was weaker (r2 = 0.599, p < 0.01), with two predictors. We realize that our high r² has to be interpreted with caution due to possible overfitting, since we had so few ground plots in which to develop the relationship with the numerous topographical and structural explanatory variables. However, this is still an interesting analysis, even with the issue of overfitting. To reduce issues with overfitting we used ridge regression, which acted as a regularization method, shrinking coefficients in order to decrease their variability and multicollinearity. This study is unique because it uses paired 1-ha plot and airborne lidar data over a tropical elevation gradient, and suggests potential for mapping species richness and diversity across elevational gradients in tropical montane ecosystems using topography and relative height metrics from spaceborne lidar with greater spatial coverage (e.g., GEDI). Full article
(This article belongs to the Special Issue Remote Sensing of Tropical Forest Biodiversity)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of Braulio Carrillo National Park using Landsat 8 data collected 26 January 2017, with 20 1-ha plot locations/Braulio Carrillo National Park (black outline) and La Selva Biological Station (white outline). Inset shows location of BCNP in the country of Costa Rica. Plots used in analysis are in Red (TEAM) and Green (Carbono), red circles are around TEAM elevation transect plots. On the right is a Digital Elevation Model (DEM) from the Shuttle Radar Topography Mission (SRTM).</p>
Full article ">Figure 2
<p>Species richness and Shannon Diversity Index along an elevation gradient in Costa Rica. The inset shows the variations of diversity across lowland sites where more plots were available for this study. Similar variability may exist in higher elevations but no additional replicates of plots were available to verify.</p>
Full article ">Figure 3
<p>Stem density and Shannon Diversity Index from field data along the elevation gradient. The inset shows the variations of the diversity and stem numbers across lowland sites where more plots were available for this study. Similar variability may exist in higher elevations but no additional replicates of plots were available to verify.</p>
Full article ">Figure 4
<p>Multiple diversity metrics within plots versus lidar-derived relative height metrics (<b>a</b>) Species richness: <span class="html-italic">r</span><sup>2</sup> for rh100, rh75, rh50, and rh25: 0.55 0.56, 0.55, 0.50 respectively; (<b>b</b>) Shannon diversity index: <span class="html-italic">r</span><sup>2</sup> for rh100, rh75, rh50, and rh25: 0.41, 0.40, 0.34, 0.26 respectively.</p>
Full article ">Figure 5
<p>Map of predicted Shannon Diversity.</p>
Full article ">Figure 6
<p>Plot-level and landscape-level mean lidar-derived relative height (rh) metrics across the elevational gradient (<b>a</b>). Red/orange shades are at the plot-level, green shades are the landscape patterns. Standard deviations of lidar-derived relative height (rh) metrics across the elevational gradients (<b>b</b>) Plot-level; (<b>c</b>) and landscape-level mean.</p>
Full article ">Figure 7
<p>Simple beta diversity metric of shared species between sites based on elevation difference between the two sites. Jaccard and Sorensen indices showed same pattern, as did geographic distance instead of elevation difference.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop