Nothing Special   »   [go: up one dir, main page]

Next Issue
Volume 10, July
Previous Issue
Volume 10, May
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 6 (June 2018) – 171 articles

Cover Story (view full-size image): Land use change and reservoir construction alter sediment transport within rivers. These changes can impact river morphology and aquatic ecosystems. The integrity of the Lower Mekong Basin is crucial to surrounding countries for transportation, energy production, sustainable water supply, and food production. In response to this need, countries have developed regional scale water quality programs, but they are limited by point-based measurements. To augment in situ surface sediment concentrations (SSSC) data from the current monitoring program, an empirical model to estimate SSSC across the Lower Mekong Basin using decades of Landsat observations was developed. This operational model was implemented in Google Earth Engine and Google App Engine, allowing users, without any prior knowledge of remote sensing, to freely access and interpret sediment data across the region. 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, 9846 KiB  
Article
Parameterization of Spectral Particulate and Phytoplankton Absorption Coefficients in Sognefjord and Trondheimsfjord, Two Contrasting Norwegian Fjord Ecosystems
by Veloisa J. Mascarenhas and Oliver Zielinski
Remote Sens. 2018, 10(6), 977; https://doi.org/10.3390/rs10060977 - 20 Jun 2018
Cited by 4 | Viewed by 4860
Abstract
We present here parameterizations of particulate and phytoplankton absorption coefficients as functions of pigment concentrations (Tchla) in Sognefjord and Trondheimsfjord along the northwestern coast of Norway. The total particulate and non-algal optical densities were measured via quantitative filter technique (QFT) in a spectrophotometer [...] Read more.
We present here parameterizations of particulate and phytoplankton absorption coefficients as functions of pigment concentrations (Tchla) in Sognefjord and Trondheimsfjord along the northwestern coast of Norway. The total particulate and non-algal optical densities were measured via quantitative filter technique (QFT) in a spectrophotometer with integrating sphere. The spectral parameter coefficients A(λ) and E(λ) of the power law describing variations of particulate and phytoplankton absorption coefficients as a function of Tchla, were not only different from those provided for open ocean case 1 waters, but also exhibited differences in the two fjords under investigation. Considering the influence of glacial meltwater leading to increased inorganic sediment load in Sognefjord we investigate differences in two different parameterizations, developed by excluding and including inner Sognefjord stations. Tchla are modelled to test the parameterizations and validated against data from the same cruise and that from a repeated campaign. Being less influenced by non-algal particles parameterizations performed well in Trondheimsfjord and yielded high coefficients of determination (R2) of modelled vs. measured Tchla. In Sognefjord, the modelled vs. measured Tchla resulted in better R2 with parameter coefficients developed excluding the inner-fjord stations influenced by glacial meltwater influx. 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>Sognefjord and Trondheimsfjord, along the northwestern coast of Norway. Blue dots indicate the stations sampled in summer 2015, cruise ID HE448.</p>
Full article ">Figure 2
<p>(<b>A</b>) Near surface optically active constituent (OAC) concentrations and absorption coefficients in Sognefjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter (CDOM), (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Sognefjord (<span class="html-italic">N</span> = 16). (<span class="html-italic">N</span>: number of sampled stations). (<b>B</b>) Near surface OAC concentrations and absorption coefficients in Trondheimsfjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter CDOM, (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Trondheimsfjord (<span class="html-italic">N</span> = 9). (<span class="html-italic">N</span>: number of sampled stations).</p>
Full article ">Figure 2 Cont.
<p>(<b>A</b>) Near surface optically active constituent (OAC) concentrations and absorption coefficients in Sognefjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter (CDOM), (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Sognefjord (<span class="html-italic">N</span> = 16). (<span class="html-italic">N</span>: number of sampled stations). (<b>B</b>) Near surface OAC concentrations and absorption coefficients in Trondheimsfjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter CDOM, (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Trondheimsfjord (<span class="html-italic">N</span> = 9). (<span class="html-italic">N</span>: number of sampled stations).</p>
Full article ">Figure 3
<p>Average absorption spectra (solid lines) with plus/minus one standard deviation (dotted lines) for total particulate (<span class="html-italic">a<sub>p</sub></span>), phytoplankton (<span class="html-italic">a<sub>phy</sub></span>), non-algal (<span class="html-italic">a<sub>nap</sub></span>), pigment-specific phytoplankton absorption (<span class="html-italic">a*<sub>phy</sub></span>) and colored dissolved organic matter (<span class="html-italic">a<sub>cdom</sub></span>) absorption in Sognefjord (<b>a</b>–<b>e</b>, <span class="html-italic">N</span> = 16) and Trondheimsfjord (<b>f</b>–<b>j</b>, <span class="html-italic">N</span> = 9).</p>
Full article ">Figure 4
<p>Trilinear graphs illustrating the relative contributions of OACs: phytoplankton (<span class="html-italic">a<sub>phy</sub></span>), non-algal particles (<span class="html-italic">a<sub>nap</sub></span>) and CDOM (<span class="html-italic">a<sub>cdom</sub></span>) to the total absorption coefficient at 440 nm in Sognefjord (<b>a</b>, <span class="html-italic">N</span> = 16) and Trondheimsfjord (<b>b</b>, <span class="html-italic">N</span> = 9). CDOM dominated absorption at 440 nm in both fjords under investigation. (<span class="html-italic">N</span>: number of data points).</p>
Full article ">Figure 5
<p>Variations in (<b>a</b>,<b>c</b>,<b>e</b>) total particulate (<span class="html-italic">a<sub>p</sub></span>) and (<b>b</b>,<b>d</b>,<b>f</b>) phytoplankton (<span class="html-italic">a<sub>phy</sub></span>) absorption coefficients as a function of pigment concentration (Tchla) at the blue (440 nm, blue filled circles) and red (675 nm, red filled circles) Chla absorption peaks in Sognefjord (<b>a</b>–<b>d</b>) and Trondheimsfjord (<b>e</b>,<b>f</b>). (<b>a</b>,<b>b</b>) represent trends in Sognefjord corresponding to analysis excluding the inner-fjord stations and (<b>c</b>,<b>d</b>) the trends including the inner-fjord stations. (<span class="html-italic">N</span>: number of data points; R<sup>2</sup>: coefficient of determination).</p>
Full article ">Figure 6
<p>Spectral parameter coefficients <span class="html-italic">A<sub>p</sub></span>, <span class="html-italic">E<sub>p</sub></span> and <span class="html-italic">A<sub>phy</sub></span>, <span class="html-italic">E<sub>phy</sub></span> in the power functions (Equations (7) and (8)) representing variations of the absorption coefficients <span class="html-italic">a<sub>p</sub></span> (<span class="html-italic">λ</span>) and <span class="html-italic">a<sub>phy</sub></span> (<span class="html-italic">λ</span>) respectively as functions of pigment concentration (Tchla) in Sognefjord and Trondheimsfjord. Spectra represent coefficients developed excluding (<b>a</b>,<b>b</b>,<b>g</b>,<b>h</b>) and including (<b>c</b>,<b>d</b>,<b>i</b>,<b>j</b>) the inner Sognefjord stations and Trondheimsfjord (<b>e</b>,<b>f</b>,<b>k</b>,<b>l</b>). (<b>A</b>) (<b>a</b>,<b>c</b>,<b>e</b>) Spectral coefficients of <span class="html-italic">A<sub>p</sub></span> (blue), <span class="html-italic">A<sub>phy</sub></span> (red) in comparison to <span class="html-italic">A<sub>phy</sub></span>. B95 (black) reported for low and mid latitude waters covering Tchla range of 0.02–25.0 mg m<sup>−3</sup> [<a href="#B5-remotesensing-10-00977" class="html-bibr">5</a>]. (<b>b</b>,<b>d</b>,<b>f</b>) Spectral coefficients of <span class="html-italic">E<sub>p</sub></span> (blue), <span class="html-italic">E<sub>phy</sub></span> (red) in comparison to <span class="html-italic">E<sub>phy</sub></span>. B95 (black) reported for low- and mid-latitude waters covering Tchla range of 0.02–25.0 mg m<sup>−3</sup> [<a href="#B5-remotesensing-10-00977" class="html-bibr">5</a>]. (<b>B</b>) Same as in (<b>A</b>) with 95% confidence intervals (ci). Dotted blue and red spectra correspond to ci of the solid blue and red spectra respectively.</p>
Full article ">Figure 7
<p>Variations in pigment-specific phytoplankton absorption coefficients, <span class="html-italic">a*<sub>phy</sub></span> (<span class="html-italic">λ</span>) as functions of Tchla (over a wavelength range 0.96–2.58 mg m<sup>−3</sup>), at selected wavelengths (<b>a</b>–<b>f</b>; 412, 440, 490, 510, 555, 675 nm) in Sognefjord (<span class="html-italic">N</span> = 15). Decrease in <span class="html-italic">a*<sub>phy</sub></span> with increase in Tchla at the blue-green wavelength bands indicating possible effects of pigment packaging. (<span class="html-italic">N</span>: number of data points).</p>
Full article ">
22 pages, 2561 KiB  
Article
Physical Retrieval of Land Surface Emissivity Spectra from Hyper-Spectral Infrared Observations and Validation with In Situ Measurements
by Guido Masiello, Carmine Serio, Sara Venafra, Giuliano Liuzzi, Laurent Poutier and Frank-M. Göttsche
Remote Sens. 2018, 10(6), 976; https://doi.org/10.3390/rs10060976 - 20 Jun 2018
Cited by 32 | Viewed by 6781
Abstract
A fully physical retrieval scheme for land surface emissivity spectra is presented, which applies to high spectral resolution infrared observations from satellite sensors. The surface emissivity spectrum is represented with a suitably truncated Principal Component Analysis (PCA) transform and PCA scores are simultaneously [...] Read more.
A fully physical retrieval scheme for land surface emissivity spectra is presented, which applies to high spectral resolution infrared observations from satellite sensors. The surface emissivity spectrum is represented with a suitably truncated Principal Component Analysis (PCA) transform and PCA scores are simultaneously retrieved with surface temperature and atmospheric parameters. The retrieval methodology has been developed within the general framework of Optimal Estimation and, in this context, is the first physical scheme based on a PCA representation of the emissivity spectrum. The scheme has been applied to IASI (Infrared Atmospheric Sounder Interferometer) and the retrieved emissivities have been validated with in situ observations acquired during a field experiment carried out in 2017 at Gobabeb (Namib desert) validation station. It has been found that the retrieved emissivity spectra are independent of background information and in good agreement with in situ observations. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>KIT’s permanent validation station GBB Wind on the 16 June 2017.</p>
Full article ">Figure 2
<p>Location of in situ measurements and collocated IASI soundings. The tags from A to D show the location of in situ measurements on 17, 22, 23 and 24 June 2017, respectively. Red circles correspond to IASI footprints over desert, those green over the gravel plain; in blue the IASI soundings best collocated with in situ observtaions.</p>
Full article ">Figure 3
<p>Instrumentation setting used for in situ measurements of emissivity soundings.</p>
Full article ">Figure 4
<p>(<b>a</b>) Mean emissivity of the ten sets of in situ measurements. (<b>b</b>) Standard deviation.</p>
Full article ">Figure 5
<p>(<b>a</b>) Exemplifying the emissivity retrieval with Fourier series and PCA expansion. The results have been averaged over 70 IASI soundings recorded over a target area in the Algerian, Sahara desert in June 2007. Note how 20 PC scores are as much effective as 450 Fourier coefficients. (<b>b</b>) Difference between PCA and Fourier.</p>
Full article ">Figure 6
<p>Ensemble of laboratory emissivity spectra used to construct the PCA basis.</p>
Full article ">Figure 7
<p>Mean and standard deviation of the ensemble of emissivity shown in <a href="#remotesensing-10-00976-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 8
<p>Singular values of the PCA decomposition of the ensemble of spectra shown in <a href="#remotesensing-10-00976-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 9
<p>Example of PCA reconstruction of a typical desert sand emissivity with a number of PC scores equal to 5, 10 and 20, respectively.</p>
Full article ">Figure 10
<p>Example of Averaging Kernels for the twenty PC scores in the retrieved state vector. The example corresponds to one of the IASI soundings in <a href="#remotesensing-10-00976-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 11
<p>Example of emissivity retrieval corresponding to one of the IASI soundings in <a href="#remotesensing-10-00976-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 12
<p>Example of emissivity retrieval for one of the IASI soundings in <a href="#remotesensing-10-00976-f002" class="html-fig">Figure 2</a>. Its accuracy was calculated as the square root of the diagonal elements of the a posteriori matrix Equation (<a href="#FD27-remotesensing-10-00976" class="html-disp-formula">27</a>).</p>
Full article ">Figure 13
<p>Retrieval of emissivity from IASI soundings over desert sand.</p>
Full article ">Figure 14
<p>Retrieval of emissivity from IASI soundings over the gravel plain.</p>
Full article ">Figure 15
<p>Comparison of retrieved emissivity for desert sand and the gravel plain, which have been obtained by averaging the spectra retrieved for the soundings shown in <a href="#remotesensing-10-00976-f013" class="html-fig">Figure 13</a> and <a href="#remotesensing-10-00976-f014" class="html-fig">Figure 14</a>. The black straight line connecting the reststrahlen band minima has a positive slope as expected for fine-grained quartz sand.</p>
Full article ">Figure 16
<p>IASI vs. in situ emissivity for the best collocated IASI soundings. The ±3<span class="html-italic">σ</span> interval is the variability (standard deviation) of in situ measurements.</p>
Full article ">Figure 17
<p>Exemplifying the angular dependence of emissivity on the angle of view. The figure shows the mean retrieval for a set of IASI soundings with FOVs smaller than 15° and larger than 25°. The set of soundings is that shown in <a href="#remotesensing-10-00976-f003" class="html-fig">Figure 3</a>.</p>
Full article ">
34 pages, 16499 KiB  
Article
A Hybrid Analytic Network Process and Artificial Neural Network (ANP-ANN) Model for Urban Earthquake Vulnerability Assessment
by Mohsen Alizadeh, Ibrahim Ngah, Mazlan Hashim, Biswajeet Pradhan and Amin Beiranvand Pour
Remote Sens. 2018, 10(6), 975; https://doi.org/10.3390/rs10060975 - 19 Jun 2018
Cited by 96 | Viewed by 10618
Abstract
Vulnerability assessment is one of the prerequisites for risk analysis in disaster management. Vulnerability to earthquakes, especially in urban areas, has increased over the years due to the presence of complex urban structures and rapid development. Urban vulnerability is a result of human [...] Read more.
Vulnerability assessment is one of the prerequisites for risk analysis in disaster management. Vulnerability to earthquakes, especially in urban areas, has increased over the years due to the presence of complex urban structures and rapid development. Urban vulnerability is a result of human behavior which describes the extent of susceptibility or resilience of social, economic, and physical assets to natural disasters. The main aim of this paper is to develop a new hybrid framework using Analytic Network Process (ANP) and Artificial Neural Network (ANN) models for constructing a composite social, economic, environmental, and physical vulnerability index. This index was then applied to Tabriz City, which is a seismic-prone province in the northwestern part of Iran with recurring devastating earthquakes and consequent heavy casualties and damages. A Geographical Information Systems (GIS) analysis was used to identify and evaluate quantitative vulnerability indicators for generating an earthquake vulnerability map. The classified and standardized indicators were subsequently weighed and ranked using an ANP model to construct the training database. Then, standardized maps coupled with the training site maps were presented as input to a Multilayer Perceptron (MLP) neural network for producing an Earthquake Vulnerability Map (EVM). Finally, an EVM was produced for Tabriz City and the level of vulnerability in various zones was obtained. South and southeast regions of Tabriz City indicate low to moderate vulnerability, while some zones of the northeastern tract are under critical vulnerability conditions. Furthermore, the impact of the vulnerability of Tabriz City on population during an earthquake was included in this analysis for risk estimation. A comparison of the result produced by EVM and the Population Vulnerability (PV) of Tabriz City corroborated the validity of the results obtained by ANP-ANN. The findings of this paper are useful for decision-makers and government authorities to obtain a better knowledge of a city’s vulnerability dimensions, and to adopt preparedness strategies in the future for Tabriz City. The developed hybrid framework of ANP and ANN Models can easily be replicated and applied to other urban regions around the world for sustainability and environmental management. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographic location of Tabriz City in the NW of Iran.</p>
Full article ">Figure 2
<p>Three-dimensional perspectives of Tabriz area (black lines) and the position of the North Tabriz Fault (NTF) (red lines). Image generated using SPOT 5 satellite images and digital elevation model (DEM).</p>
Full article ">Figure 3
<p>Geology Map of Tabriz City. <b>Legend: Mmg2 =</b> Interlayer of greenish grey marl associated with interlayer of gypsum- bring sandy marl; <b>Msc5 =</b> Interbedded red conglomerate with sandstone and red marl; <b>Msm4 =</b> Sandstone and red marl; <b>Pldt =</b> Diatomic and fish interbedded with fine particles sediment; <b>Plqc =</b> Interlayer of semi-hard conglomerate associated with sandstone and pumice; <b>Plqc =</b> Interlayer of semi-hard conglomerate associated with sandstone and pumice. <b>Qal =</b> Quaternary alluvium; <b>Qt2 =</b> Young terrace and alluvium deposits [<a href="#B62-remotesensing-10-00975" class="html-bibr">62</a>].</p>
Full article ">Figure 4
<p>Standardized input layers derived from GIS procedure for the indicators used in this study; (<b>a</b>) Age of buildings density; (<b>b</b>) Danger centers; (<b>c</b>) Employed density; (<b>d</b>) Distance to fault (<b>e</b>) Buildings floors; (<b>f</b>) Features of geology (<b>g</b>) Housing density; (<b>h</b>) Household density; (<b>i</b>) Literate density; (<b>j</b>) Size of building; (<b>k</b>) Building materials; (<b>l</b>) Distance to open spaces; (<b>m</b>) Population density; (<b>n</b>) Quality of buildings; (<b>o</b>) Distance to rescue centers; (<b>p</b>) Buildings density; (<b>q</b>) Percentage of slope; (<b>r</b>) Distance to roads network; (<b>s</b>) Unemployed people; (<b>t</b>) Commercial density.</p>
Full article ">Figure 4 Cont.
<p>Standardized input layers derived from GIS procedure for the indicators used in this study; (<b>a</b>) Age of buildings density; (<b>b</b>) Danger centers; (<b>c</b>) Employed density; (<b>d</b>) Distance to fault (<b>e</b>) Buildings floors; (<b>f</b>) Features of geology (<b>g</b>) Housing density; (<b>h</b>) Household density; (<b>i</b>) Literate density; (<b>j</b>) Size of building; (<b>k</b>) Building materials; (<b>l</b>) Distance to open spaces; (<b>m</b>) Population density; (<b>n</b>) Quality of buildings; (<b>o</b>) Distance to rescue centers; (<b>p</b>) Buildings density; (<b>q</b>) Percentage of slope; (<b>r</b>) Distance to roads network; (<b>s</b>) Unemployed people; (<b>t</b>) Commercial density.</p>
Full article ">Figure 5
<p>The framework of the hybrid ANP-ANN model.</p>
Full article ">Figure 6
<p>The super matrix consisting of N clusters.</p>
Full article ">Figure 7
<p>The architecture of ANN as a multi-layered Perceptron (MLP) used in this study.</p>
Full article ">Figure 8
<p>The ANP model for constructing vulnerability index.</p>
Full article ">Figure 9
<p>Training site map extracted from ANP model.</p>
Full article ">Figure 10
<p>Final earthquake vulnerability map extracted from ANP-ANN model.</p>
Full article ">Figure 11
<p>The procedure of determining population vulnerability.</p>
Full article ">Figure 12
<p>The scatter-plot between the PV and EVM according to most vulnerable zones.</p>
Full article ">Figure 13
<p>The scatter-plot between the PV and EVM according to least vulnerable zones.</p>
Full article ">
29 pages, 4817 KiB  
Article
Combining a Two Source Energy Balance Model Driven by MODIS and MSG-SEVIRI Products with an Aggregation Approach to Estimate Turbulent Fluxes over Sparse and Heterogeneous Vegetation in Sahel Region (Niger)
by Bouchra Ait Hssaine, Jamal Ezzahar, Lionel Jarlan, Olivier Merlin, Said Khabba, Aurore Brut, Salah Er-Raki, Jamal Elfarkh, Bernard Cappelaere and Ghani Chehbouni
Remote Sens. 2018, 10(6), 974; https://doi.org/10.3390/rs10060974 - 19 Jun 2018
Cited by 8 | Viewed by 5507
Abstract
Estimates of turbulent fluxes (i.e., sensible and latent heat fluxes H and LE) over heterogeneous surfaces is not an easy task. The heterogeneity caused by the contrast in vegetation, hydric and soil conditions can generate a large spatial variability in terms of surface–atmosphere [...] Read more.
Estimates of turbulent fluxes (i.e., sensible and latent heat fluxes H and LE) over heterogeneous surfaces is not an easy task. The heterogeneity caused by the contrast in vegetation, hydric and soil conditions can generate a large spatial variability in terms of surface–atmosphere interactions. This study considered the issue of using a thermal-based two-source energy model (TSEB) driven by MODIS (Moderate resolution Imaging Spectroradiometer) and MSG (Meteosat Second Generation) observations in conjunction with an aggregation scheme to derive area-averaged H and LE over a heterogeneous watershed in Niamey, Niger (Wankama catchment). Data collected in the context of the African Monsoon Multidisciplinary Analysis (AMMA) program, including a scintillometry campaign, were used to test the proposed approach. The model predictions of area-averaged turbulent fluxes were compared to data acquired by a Large Aperture Scintillometer (LAS) set up over a transect about 3.2 km-long and spanning three vegetation types (millet, fallow and degraded shrubs). First, H and LE fluxes were estimated at the MSG-SEVIRI grid scale by neglecting explicitly the subpixel heterogeneity. Moreover, the impact of upscaling the model’s inputs was investigated using in-situ input data and three aggregation schemes of increasing complexity based on MODIS products: a simple averaging of inputs at the MODIS resolution scale, another simple averaging scheme that considers scintillometer footprint extent, and the weighted average of inputs based on the footprint weighting function. The H and LE simulated using the footprint weighted method were more accurate than for the two other aggregation rules despite the heterogeneity of the landscape. The statistical values are: correlation coefficient (R) = 0.71, root mean square error (RMSE) = 63 W/m2 and mean bias error (MBE) = −23 W/m2 for H and an R = 0.82, RMSE = 88 W/m2 and MBE = 45 W/m2 for LE. This study opens perspectives for the monitoring of convective and evaporative fluxes over heterogeneous landscape based on medium resolution satellite products. Full article
(This article belongs to the Special Issue Remote Sensing of Land-Atmosphere Interactions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Map of the Wankama basin; (<b>b</b>) location of the three study sites (millet, fallow and degraded shrubs); and (<b>c</b>) basin toposequence [<a href="#B26-remotesensing-10-00974" class="html-bibr">26</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) An overview of the Wankama basin and the experimental setup, locations of LAS (T and R stand for transmitter and receiver, respectively) and three EC systems are shown (millet, fallow and degraded shrubs sites). Photos (EC system (<b>b</b>) and scintillometer (<b>c</b>)) were taken by J. Ezzahar as part of the ACN project.</p>
Full article ">Figure 3
<p>Footprint of the LAS, calculated using the footprint model of Horst and Weil, superposed with a Ts/MODIS image on DOY 273.</p>
Full article ">Figure 4
<p>Time series of LAS latent and sensible heat fluxes (daily average between 9:00 a.m. and 5:00 p.m.).</p>
Full article ">Figure 5
<p>Scatterplot of: MODIS Ts (<b>a</b>); and <math display="inline"> <semantics> <mi mathvariant="sans-serif">α</mi> </semantics> </math> (<b>b</b>) versus in-situ data (only millet and fallow sites are considered).</p>
Full article ">Figure 6
<p>Scatterplot of MODIS versus MSG SEVIRI Ts.</p>
Full article ">Figure 7
<p>Sensible (<b>a</b>); and latent (<b>b</b>) heat fluxes scatterplots between LAS measurements and TSEB prediction fed by aggregated in-situ data (at 30 min time).</p>
Full article ">Figure 8
<p>Sensible (<b>a</b>) and latent (<b>b</b>) heat fluxes scatterplots of the comparison between the LAS grid scale measurements and TSEB predictions based on MSG SEVIRI products (at MSG-SEVIRI overpass time).</p>
Full article ">Figure 9
<p>Sensible and latent heat fluxes scatterplots of the comparison between the LAS grid scale measurements and TSEB predictions based on MODIS products for: simple averaging (<b>a</b>,<b>b</b>); area weighted method (<b>c</b>,<b>d</b>); and footprint weighted method (<b>e</b>,<b>f</b>) (at MODIS overpass time).</p>
Full article ">Figure 10
<p>Time series of the comparison between the LAS grid scale measurements (<math display="inline"> <semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>LAS</mi> </mrow> </msub> </mrow> </semantics> </math> and <math display="inline"> <semantics> <mrow> <msub> <mrow> <mi>LE</mi> </mrow> <mrow> <mi>LAS</mi> </mrow> </msub> </mrow> </semantics> </math>) and TSEB predictions based on MODIS products (Hsim and LEsim) for footprint weighted method (at MODIS overpass time).</p>
Full article ">
22 pages, 14068 KiB  
Article
Land Cover Segmentation of Airborne LiDAR Data Using Stochastic Atrous Network
by Hasan Asy’ari Arief, Geir-Harald Strand, Håvard Tveite and Ulf Geir Indahl
Remote Sens. 2018, 10(6), 973; https://doi.org/10.3390/rs10060973 - 19 Jun 2018
Cited by 30 | Viewed by 7284
Abstract
Inspired by the success of deep learning techniques in dense-label prediction and the increasing availability of high precision airborne light detection and ranging (LiDAR) data, we present a research process that compares a collection of well-proven semantic segmentation architectures based on the deep [...] Read more.
Inspired by the success of deep learning techniques in dense-label prediction and the increasing availability of high precision airborne light detection and ranging (LiDAR) data, we present a research process that compares a collection of well-proven semantic segmentation architectures based on the deep learning approach. Our investigation concludes with the proposition of some novel deep learning architectures for generating detailed land resource maps by employing a semantic segmentation approach. The contribution of our work is threefold. (1) First, we implement the multiclass version of the intersection-over-union (IoU) loss function that contributes to handling highly imbalanced datasets and preventing overfitting. (2) Thereafter, we propose a novel deep learning architecture integrating the deep atrous network architecture with the stochastic depth approach for speeding up the learning process, and impose a regularization effect. (3) Finally, we introduce an early fusion deep layer that combines image-based and LiDAR-derived features. In a benchmark study carried out using the Follo 2014 LiDAR data and the NIBIO AR5 land resources dataset, we compare our proposals to other deep learning architectures. A quantitative comparison shows that our best proposal provides more than 5% relative improvement in terms of mean intersection-over-union over the atrous network, providing a basis for a more frequent and improved use of LiDAR data for automatic land cover segmentation. Full article
(This article belongs to the Special Issue Deep Learning for Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of (<b>a</b>) pooling, (<b>b</b>) unpooling, (<b>c</b>) convolution, and (<b>d</b>) transposed convolution operation. Each operation uses a 3 × 3 kernel with a stride/step of two, and for any overlapping cell, a summation of the overlapping values of the cells are performed.</p>
Full article ">Figure 2
<p>Bottleneck building block in deep residual network (ResNet) architecture.</p>
Full article ">Figure 3
<p>Atrous kernel with rates one, two, and four. The atrous kernel with rate one is identical to a normal convolution kernel.</p>
Full article ">Figure 4
<p>The study area (Follo) with 1877 tiles covering an area of 819 km<sup>2</sup>.</p>
Full article ">Figure 5
<p>The preprocessing pipeline that was used for our research. It should be noted that the data augmentation was only applied to the training dataset.</p>
Full article ">Figure 6
<p>Qualitative segmentation results from FCN-8s and atrous network. The mean pixel accuracy (MPA) value is included at the bottom of each prediction map. Classes and colors: <span class="html-italic">other land types</span> (red), settlement (pink), road/transportation (chocolate), cultivation/grass (orange), forest (green), swamp (dark blue), lake–river (eggshell blue), and ocean (light cyan). (<b>a</b>) and (<b>b</b>) show the ability of the atrous network to predict the most common classes better than the FCN-8s. (<b>c</b>) and (<b>d</b>) show the shortcoming of the upsampling technique in the atrous network compared to the one in the FCN-8s.</p>
Full article ">Figure 7
<p>Comparison of the confusion matrix between (<b>a</b>) FCN-8s and (<b>b</b>) atrous network. The class names are (1) <span class="html-italic">other land types</span>, (2) settlement, (3) road/transportation, (4) cultivation/grass, (5) forest, (6) swamp, (7) lake–river, and (8) ocean. The colored bar represents the value of each cell: the higher the value, the darker the cell color.</p>
Full article ">Figure 8
<p>Illustration of the atrous–FCN architecture.</p>
Full article ">Figure 9
<p>Data fusion technique based on the <span class="html-italic">Stochastic Atrous Network</span> (SA-Net) architecture. (<b>a</b>) Our proposed EarlyFusion architecture, which merges red–green–blue (RGB), intensity and height above ground (HAG) in the early convolution layers. (<b>b</b>) The FuseNet style architecture, which encodes RGB values and depth (HAG) using two branches of encoders, as inspired by C. Hazirbas et al.’s work [<a href="#B56-remotesensing-10-00973" class="html-bibr">56</a>].</p>
Full article ">Figure 10
<p>Final prediction results from our EarlyFusion SA-Net with RGB, HAG, and Intensity. The color legend is presented in the caption of <a href="#remotesensing-10-00973-f006" class="html-fig">Figure 6</a>. The figures cover and visualize different types of areas, such as (<b>a</b>) forest and ocean, (<b>b</b>) settlement and road, and (<b>c</b>) cultivation and open land.</p>
Full article ">Figure 11
<p>Training errors and validation accuracies from the utilization of the cross-entropy and IoU loss functions on the deep residual network (ResNet)-FCN architecture. The training errors were calculated as the mean loss for all of the iterations in every epoch of a training process, while the validation accuracies were calculated at the beginning of each epoch using the validation data.</p>
Full article ">Figure 12
<p>Training and validation accuracies for the Atrous-FCN and SA-Net.</p>
Full article ">
21 pages, 6824 KiB  
Article
Assessment of Irrigation Performance in Large River Basins under Data Scarce Environment—A Case of Kabul River Basin, Afghanistan
by Fazlullah Akhtar, Usman Khalid Awan, Bernhard Tischbein and Umar Waqas Liaqat
Remote Sens. 2018, 10(6), 972; https://doi.org/10.3390/rs10060972 - 18 Jun 2018
Cited by 30 | Viewed by 7183
Abstract
The Kabul River basin (KRB) of Afghanistan, a lifeline of around 10 million people, has multiplicity of governance, management, and development-related challenges leading to inequity, inadequacy, and unreliability of irrigation water distribution. Prior to any uplifting intervention, there is a need to evaluate [...] Read more.
The Kabul River basin (KRB) of Afghanistan, a lifeline of around 10 million people, has multiplicity of governance, management, and development-related challenges leading to inequity, inadequacy, and unreliability of irrigation water distribution. Prior to any uplifting intervention, there is a need to evaluate the performance of irrigation system on a long term basis to identify the existing bottlenecks. Although there are several indicators available for the performance evaluation of the irrigation schemes, we used the coefficient of variation (CV) of actual evapotranspiration (ETa) in space (basin, sub-basin, and provincial level), relative evapotranspiration (RET), and temporal CV of RET, to assess the equity, adequacy, and reliability of water distribution, respectively, from 2003 to 2013. The ETa was estimated through a surface energy balance system (SEBS) algorithm and the ETa estimates were validated using the advection aridity (AA) method with a R2 value of 0.81 and 0.77 at Nawabad and Sultanpur stations, respectively. The global land data assimilation system (GLDAS) and moderate-resolution imaging spectroradiometer (MODIS) products were used as main inputs to the SEBS. Results show that the mean seasonal sub-based RET values during summer (May–September) (0.37 ± 0.06) and winter (October–April) (0.40 ± 0.08) are below the target values (RET ≥ 0.75) during 2003–2013. The CV of the mean ETa, within sub-basins and provinces for the entire study period, has an equitable distribution of water from October–January (0.09 ± 0.04), whereas the highest inequity (0.24 ± 0.08) in water distribution is during early summer. The range of the CV of the mean ETa (0.04–0.06) on a monthly and seasonal basis shows the unreliability of water supplies in several provinces or sub-basins. The analysis of the temporal CV of mean RET highlights the unreliable water supplies across the entire basin. The maximum ETa during the study period was estimated for the Shamal sub-basin (552 ± 43 mm) while among the provinces, Kunar experienced the highest ETa (544 ± 39 mm). This study highlights the dire need for interventions to improve the irrigation performance in time and space. The proposed methodology can be used as a framework for monitoring and implementing water distribution plans in future. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the Kabul River Basin with its sub-basins and provincial.</p>
Full article ">Figure 2
<p>Climatograph of the downstream (<b>left</b>) and the central upstream (<b>right</b>) of the Kabul River Basin.</p>
Full article ">Figure 3
<p>Methodological framework to estimate actual evapotranspiration in different spatial units of Kabul River Basin with strategic time steps.</p>
Full article ">Figure 4
<p>Comparison of the daily actual evapotranspiration estimated through SEBS algorithm and AA model (1 January 2013–31 December 2013) at (<b>a</b>) Nawabad (Kunar) and (<b>b</b>) Sultanpur (Nangarhar) stations of the Kabul River Basin.</p>
Full article ">Figure 5
<p>(<b>a</b>) Box and whiskers plot of annual evapotranspiration (ET<sub>a</sub>), showing the temporal (2003–2013) and spatial variation of the ET<sub>a</sub>. The horizontal line inside each box represents the median, the lower and upper whiskers show the ET<sub>a</sub> range during the study period. The outliers are encircled with a dotted line. (<b>b</b>) Temporal variation (2003–2013) of actual evapotranspiration of main land covers across the Kabul River Basin.</p>
Full article ">Figure 6
<p>Mean seasonal (2003–2013) variation of actual evapotranspiration in winter and summer seasons at the (<b>a</b>) and (<b>b</b>) sub-basins and (<b>c</b>) and (<b>d</b>) provinces of the Kabul River Basin (Note: The statistical details have been explained in the text).</p>
Full article ">Figure 6 Cont.
<p>Mean seasonal (2003–2013) variation of actual evapotranspiration in winter and summer seasons at the (<b>a</b>) and (<b>b</b>) sub-basins and (<b>c</b>) and (<b>d</b>) provinces of the Kabul River Basin (Note: The statistical details have been explained in the text).</p>
Full article ">Figure 7
<p>Seasonal distribution of actual evapotranspiration (mm) during winter (October–April, 2012/2013) and summer (May–September 2013) in the Kabul River Basin.</p>
Full article ">Figure 8
<p>Physical distribution of cropped and non-cropped area of the KRB.</p>
Full article ">Figure 9
<p>Coefficient of variation of mean actual evapotranspiration at the spatial administrative units of (<b>a</b>) and (<b>b</b>) inter and intra sub-basin and (<b>c</b>) and (<b>d</b>) provincial level.</p>
Full article ">Figure 10
<p>Seasonal relative evapotranspiration in (<b>a</b>) Winter and (<b>b</b>) Summer across the Kabul River Basin.</p>
Full article ">Figure 11
<p>Inter-seasonal distribution of relative evapotranspiration across (<b>a</b>) sub-basins and (<b>b</b>) provinces of the Kabul River Basin in winter and summer seasons.</p>
Full article ">Figure 12
<p>Temporal coefficient of variation (2003–2013) of relative evapotranspiration during (<b>a</b>) winter and (<b>b</b>) summer.</p>
Full article ">
21 pages, 4727 KiB  
Article
Improving Geometric Performance for Imagery Captured by Non-Cartographic Optical Satellite: A Case Study of GF-1 WFV Imagery
by Kai Xu, Guo Zhang, Mingjun Deng, Qingjun Zhang and Deren Li
Remote Sens. 2018, 10(6), 971; https://doi.org/10.3390/rs10060971 - 18 Jun 2018
Cited by 1 | Viewed by 3719
Abstract
Numerous countries have established their own Earth observing systems (EOSs) for global change research. Data acquisition efforts are generally only concerned with the completion of the mission regardless of the potential to expand into other areas, which reduces the application effectiveness of Earth [...] Read more.
Numerous countries have established their own Earth observing systems (EOSs) for global change research. Data acquisition efforts are generally only concerned with the completion of the mission regardless of the potential to expand into other areas, which reduces the application effectiveness of Earth observation data. This paper explores the cartographic possibility of images being not initially intended for surveying and mapping, and a novel method is proposed to improve the geometric performance. First, the rigorous sensor model (RSM) is recovered from the rational function model (RFM), and then the system errors of the non-cartographic satellite’s imagery are compensated by using the conventional geometric calibration method based on RSM; finally, a new and improved RFM is generated. The advantage of the method over traditional ones is that it divides the errors into static errors and non-static errors for each image during the improvement process. Experiments using images collected with the Gaofen-1 (GF-1) wide-field view (WFV) camera demonstrate that the orientation accuracy of the proposed method is within 1 pixel for both calibration and validation images, and the obvious high-order system errors are eliminated. Moreover, a block adjustment test shows that the vertical accuracy is improved from 21 m to 11 m with four ground control points (GCPs) after compensation, which can fulfill requirements for 1:100,000 stereo mapping in mountainous areas. Generally, the proposed method can effectively improve the geometric potential for images captured by non-cartographic satellite. Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of the principle of recovering the rigorous sensor model (RSM) from the rational function model (RFM).</p>
Full article ">Figure 2
<p>Schematic diagram of ground control points (GCPs) acquisition strategy for wide swath image. (<b>a</b>) Geometric calibration field (GCF) image 1 covers the right half of calibration image 1 and the left half of calibration image 3; (<b>b</b>) GCF image 2 covering the middle part of calibration image 2.</p>
Full article ">Figure 3
<p>Schematic diagram of the system error compensation model with constrained conditions.</p>
Full article ">Figure 4
<p>Schematic diagram of wide-field view (WFV) camera onboard GF-1 satellite.</p>
Full article ">Figure 5
<p>(<b>a</b>) Spatial coverage of the images of GCF; (<b>b</b>) image showing a corresponding 1:5000 GCP obtained via photographs taken in aerial photography field work.</p>
Full article ">Figure 6
<p>Residual error before and after calibration. The horizontal axis denotes the image row across the track and the vertical axis denotes residual errors after orientation. (<b>a</b>) WFV-1 residual error before calibration; (<b>b</b>) WFV-1 residual error after calibration; (<b>c</b>) WFV-4 residual error before calibration; (<b>d</b>) WFV-4 residual error after calibration.</p>
Full article ">Figure 7
<p>Residual error before and after compensation. The horizontal axis denotes the image row across the track and the vertical axis denotes residual errors after orientation. (<b>a</b>) WFV-1 residual error before compensation; (<b>b</b>) WFV-1 residual error after compensation; (<b>c</b>) WFV-4 residual error before compensation; (<b>d</b>) WFV-4 residual error after compensation.</p>
Full article ">Figure 8
<p>Orientation errors before and after applying compensation parameters. (<b>a</b>) Image 2143625 before compensation; (<b>b</b>) image 2143625 after compensation; (<b>c</b>) image 2986583 before compensation; (<b>d</b>) image 2986583 after compensation.</p>
Full article ">
23 pages, 9192 KiB  
Article
Remote Sensing of Leaf Area Index from LiDAR Height Percentile Metrics and Comparison with MODIS Product in a Selectively Logged Tropical Forest Area in Eastern Amazonia
by Yonghua Qu, Ahmed Shaker, Carlos Alberto Silva, Carine Klauberg and Ekena Rangel Pinagé
Remote Sens. 2018, 10(6), 970; https://doi.org/10.3390/rs10060970 - 17 Jun 2018
Cited by 27 | Viewed by 6208
Abstract
Leaf area index (LAI) is an important parameter to describe the capacity of forests to intercept light and thus affects the microclimate and photosynthetic capacity of canopies. In general, tropical forests have a higher leaf area index and it is a challenge to [...] Read more.
Leaf area index (LAI) is an important parameter to describe the capacity of forests to intercept light and thus affects the microclimate and photosynthetic capacity of canopies. In general, tropical forests have a higher leaf area index and it is a challenge to estimate LAI in a forest with a very dense canopy. In this study, it is assumed that the traditional Light Detection and Ranging (LiDAR)-derived fractional vegetation cover (fCover) has weak relationship with leaf area index in a dense forest. We propose a partial least squares (PLS) regression model using the height percentile metrics derived from airborne LiDAR data to estimate the LAI of a dense forest. Ground inventory and airborne LiDAR data collected in a selectively logged tropical forest area in Eastern Amazonia are used to map LAI from the plot level to the landscape scale. The results indicate that the fCover, derived from the first return or the last return, has no significant correlations with the ground-based LAI. The PLS model evaluated by the leave-one-out validation shows that the estimated LAI is significantly correlated with the ground-based LAI with an R2 of 0.58 and a root mean square error (RMSE) of 1.13. A data comparison indicates that the Moderate Resolution Imaging Spectrometer (MODIS) LAI underestimates the landscape-level LAI by about 22%. The MODIS quality control data show that in the selected tile, the cloud state is not the primary factor affecting the MODIS LAI performance; rather, the LAI from the main radiative transfer (RT) algorithm contributes much to the underestimation of the LAI in the tropical forest. In addition, the results show that the LiDAR-based LAI has a better response to the logging activities than the MODIS-based LAI, and that the leaf area reduction caused by logging is about 13%. In contrast, the MODIS-based LAI exhibits no apparent spatial correlation with the LiDAR-based LAI. It is suggested that the main algorithm of MODIS should be improved with regard to tropical forests. The significance of this study is the proposal of a framework to produce ground-based LAI using forest inventory data and determine the plot-level LAI at the airborne and satellite scale using LiDAR data. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The location of the experiment in Brazil (<b>a</b>–<b>c</b>) and the plot distribution along the transect overlaid on the Light Detection and Ranging (LiDAR)-derived canopy height model (<b>d</b>). The enlarged images of the plots in the unlogged and selectively logged areas are shown in subplots (<b>d1</b>,<b>d2</b>) respectively.</p>
Full article ">Figure 2
<p>Workflow for the generation of the plot-based leaf area index (LAI) from ground measured diameter at breast height (DBH). The superscripts a, b, c denote the references [<a href="#B42-remotesensing-10-00970" class="html-bibr">42</a>,<a href="#B43-remotesensing-10-00970" class="html-bibr">43</a>,<a href="#B44-remotesensing-10-00970" class="html-bibr">44</a>] respectively.</p>
Full article ">Figure 3
<p>Illustration of the LiDAR leaf area index (LAI) (50 × 50 m) grids overlaid on the Moderate Resolution Imaging Spectrometer (MODIS) LAI map (463 × 463 m).</p>
Full article ">Figure 4
<p>Frequency distribution of ground plot variables, (<b>a</b>) aboveground biomass (AGB, Mg/ha) and (<b>b</b>) leaf area index (LAI). The frequency distribution is fitted using a Gaussian model (black curve). The mean values and standard deviations are plotted as the vertical dashed lines and horizontal lines, respectively.</p>
Full article ">Figure 5
<p>Scatter plots of ground LAI and LiDAR-derived fractional cover. The subplots (<b>a</b>,<b>b</b>) show the fractional cover calculated from the first and last return points respectively. The points in the red circles are used as examples in <a href="#remotesensing-10-00970-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 6
<p>LiDAR-derived height cumulative percentile of the two example points of LAI = 2.05 and LAI = 5.96, as indicated in <a href="#remotesensing-10-00970-f005" class="html-fig">Figure 5</a>a</p>
Full article ">Figure 7
<p>Correlation coefficients between plot-based LAI with LiDAR height metrics. The labels of the <span class="html-italic">x</span>-axis ticks, P01, P05, …, P99 indicate the percentile values of the canopy height at the 1st, 5th, …, 99th percentiles.</p>
Full article ">Figure 8
<p>Color plot of Pearson’s correlation coefficient between the height percentile metrics (<span class="html-italic">x</span> and <span class="html-italic">y</span> labels).</p>
Full article ">Figure 9
<p>Percent of the explained variance (<b>a</b>) and the mean squared error (MSE) vs. the number of partial least squares (PLS) components (<b>b</b>).</p>
Full article ">Figure 10
<p>The weight (contribution) of the LiDAR-derived height percentile metrics to the first PLS component. The <span class="html-italic">x</span>-axis labels represent the height percentile metrics as indicated in the <span class="html-italic">x</span>-axis of <a href="#remotesensing-10-00970-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 11
<p>Leave-one-out cross-validation result of the PLS model. The scatter plot (<b>a</b>) shows the paired comparison and the Q-Q plot (<b>b</b>) shows the similarity in the distribution of the two LAI datasets.</p>
Full article ">Figure 12
<p>Boxplots of LiDAR landscape-level LAI and the MODIS MCD15A3H product. The central line in the box indicates the median and the bottom and top edges indicate the 25th and 75th percentiles. Outliers are plotted using the “+” symbol.</p>
Full article ">Figure 13
<p>Landscape-level LiDAR LAI values of the 43 pixels and the corresponding MODIS LAI values. The vertical bars indicate one standard deviation of the LiDAR LAI values.</p>
Full article ">Figure 14
<p>LiDAR-based LAI and MODIS-based LAI values in resolution of 463 m (<b>a</b>,<b>b</b>) and aggregated into work-unit scale (<b>c</b>,<b>d</b>). The left column shows the LiDAR LAI (<b>a</b>,<b>c</b>) and the right column the MODIS LAI (<b>b</b>,<b>d</b>). The pixel indices are indicated in subplot (<b>a</b>,<b>b</b>). The years of the reduced-impact logging (RIL) or the unlogged status are indicated in the labels of the subplots in (<b>c</b>,<b>d</b>) in red text.</p>
Full article ">Figure 15
<p>Comparison of MODIS and LiDAR LAI values in logged and unlogged areas.</p>
Full article ">Figure 16
<p>Three images obtained in different seasons of an Amazon tropical forest near the study area.</p>
Full article ">
24 pages, 4716 KiB  
Article
Relation between Convective Rainfall Properties and Antecedent Soil Moisture Heterogeneity Conditions in North Africa
by Irina Y. Petrova, Diego G. Miralles, Chiel C. Van Heerwaarden and Hendrik Wouters
Remote Sens. 2018, 10(6), 969; https://doi.org/10.3390/rs10060969 - 17 Jun 2018
Cited by 9 | Viewed by 5915
Abstract
Recent observational studies have demonstrated the relevance of soil moisture heterogeneity and the associated thermally-induced circulation on deep convection and rainfall triggering. However, whether this dynamical mechanism further influences rainfall properties—such as rain volume or timing—has yet to be confirmed by observational data. [...] Read more.
Recent observational studies have demonstrated the relevance of soil moisture heterogeneity and the associated thermally-induced circulation on deep convection and rainfall triggering. However, whether this dynamical mechanism further influences rainfall properties—such as rain volume or timing—has yet to be confirmed by observational data. Here, we analyze 10 years of satellite-based sub-daily soil moisture and precipitation records and explore the potential of strong spatial gradients in morning soil moisture to influence the properties of afternoon rainfall in the North African region, at the 100-km scale. We find that the convective rain systems that form over locally drier soils and anomalously strong soil moisture gradients have a tendency to initiate earlier in the afternoon; they also yield lower volumes of rain, weaker intensity and lower spatial variability. The strongest sensitivity to antecedent soil conditions is identified for the timing of the rain onset; it is found to be correlated with the magnitude of the soil moisture gradient. Further analysis shows that the early initiation of rainfall over dry soils and strong surface gradients yet requires the presence of a very moist boundary layer on that day. Our findings agree well with the expected effects of thermally-induced circulation on rainfall properties suggested by theoretical studies and point to the potential of locally drier and heterogeneous soils to influence convective rainfall development. The systematic nature of the identified effect of soil moisture state on the onset time of rainstorms in the region is of particular relevance and may help foster research on rainfall predictability. Full article
(This article belongs to the Special Issue Remote Sensing of Land-Atmosphere Interactions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area and summary of the methodology. (<b>a</b>) Study domain in North Africa; the color-scale indicates elevation (m); black lines mark major rivers; and the orography mask is shown in red shading. The main orographic features in the region are the Air Mountains (AM), Darfur Mountains (DM), Ethiopian Highlands (EH), Cameroon Mountains (CM), Jos Plateau (JP) and Guinea Highlands (GH). The five smaller domains marked with black rectangles indicate the areas selected for the analyses of atmospheric conditions over the relatively flat West African region. (<b>b</b>) Schematic representation of afternoon rainfall event geometry. (<b>c</b>) Generalized example illustrating the calculation of the deviation, <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>S</mi> </mrow> </semantics></math>, of the observed spatial soil moisture gradient, <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>e</mi> </mrow> </semantics></math>, from the expected value of climatological soil moisture gradients, <math display="inline"><semantics> <mover> <msub> <mi>S</mi> <mrow> <mi>c</mi> <mi>n</mi> <mi>t</mi> <mi>r</mi> </mrow> </msub> <mo>¯</mo> </mover> </semantics></math>, for an individual rainfall event.</p>
Full article ">Figure 2
<p>Sensitivity of afternoon rain properties to preceding soil moisture heterogeneity conditions. D-statistic of the Anderson–Darling (A–D) test, i.e., weighted average squared difference between the two selected Empirical Distribution Functions (EDF), for every <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range (<span class="html-italic">x</span>-axis) and six rain properties: (<b>a</b>) rain area, (<b>b</b>) volume, (<b>c</b>) spatial heterogeneity, (<b>d</b>) maximum intensity, (<b>e</b>) time of maximum intensity and (<b>f</b>) onset time. The first EDF results from the sample of the specific AR property in the particular <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range, while the second EDF results from the total sample of that AR property independent of the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range. Differences significant at the 99% level (<span class="html-italic">p</span>-value &lt; 0.01) are shaded in gray. In the case of the two discrete AR time parameters (<b>e</b>,<b>f</b>), the calculation of the A–D statistic for every <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range was repeated 100 times for every 30 randomly-selected event values. The two inset plots in (<b>e</b>) and (<b>f</b>) show the spread of the estimated <span class="html-italic">p</span>-values for every sample and every <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range. The number of events in every <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range is the same, i.e., ∼1500, because they are binned per decile. The same qualitative result was obtained for the two discrete distributions using the Chi-squared non-parametric test for a complete 1500 sample size (not shown).</p>
Full article ">Figure 3
<p>Preference of each rain property over antecedent soil moisture heterogeneity conditions. As in <a href="#remotesensing-10-00969-f002" class="html-fig">Figure 2</a>, the histogram difference between two probability density functions is shown: one histogram corresponding to the sample of that particular AR property in the particular <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range; the second one resulting from the total sample of that AR property independent of the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> range. Unlike in <a href="#remotesensing-10-00969-f002" class="html-fig">Figure 2</a>, the values of the rain properties are separated into bins and shown in actual units (from left to right): rain area (%), volume (mm/9 h), maximum intensity (mm/h), spatial heterogeneity (mm/9 h), time of maximum intensity (LST) and onset time (LST).</p>
Full article ">Figure 4
<p>Spatial distribution of the correlations between the rain properties and soil moisture heterogeneity. Spearman’s rank correlation coefficient between the soil moisture gradient, <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>e</mi> </mrow> </semantics></math>, and corresponding AR property value is estimated for each 5 × 5 degree grid box. Only the extreme negative <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> values (Range #1) are considered. Boxes with significant correlation (<span class="html-italic">p</span>-value &lt; 0.05) are marked with a cross. Grid boxes where the number of events is below 20 are masked out.</p>
Full article ">Figure 5
<p>Co-variability of antecedent soil moisture and boundary layer conditions on rain event days. Scatter plots between selected boundary layer parameters at 12 LST and soil moisture at 13:30 LST measured in the AR event <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </semantics></math> location. The calculations are shown for one of the five African domains (11–12<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>N, 8<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>W–12<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>E) and for every of the 10 heterogeneity ranges. The blue points indicate AR events with an earlier onset time, i.e., at 15 LST. All the other AR events are indicated by gray dots. Boundary layer parameters shown are (from top to bottom) mean Relative Humidity between 850 and 600 hPa (RH), Total column Precipitable Water (TPW), the Humidity Index (HI), Lifting Condensation Level (LCL), Boundary Layer Height (BLH), Convective Triggering Potential (CTP) and Convective Available Potential Energy (CAPE).</p>
Full article ">Figure 6
<p>Preference of early onset AR events to occur over particular antecedent soil moisture and boundary layer conditions. (<b>a</b>–<b>c</b>) Difference in the Joint Probability Distribution (JPD) of absolute soil moisture (at 13:30 LST) and each boundary layer variable (at 12 LST) between AR events with an earlier onset time (i.e., 15 LST) and a later onset time (i.e., 18–24 LST). The plots refer to the events occurring over extreme heterogeneous and dry soils (<math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> Range #1) only. Accordingly, red shading indicates a higher probability of early onset AR events to occur over dry soil and moist boundary layer conditions as compared to AR events with later onset time (and vice versa for gray shading). (<b>d</b>–<b>f</b>) Difference in the JPD of absolute soil moisture (at 13:30 LST) and each boundary layer variable (at 12 LST) between early AR events that fall over extremely heterogeneous landscapes (Range #1) and those that fall in less heterogeneous states (Ranges #2–10). Accordingly, red shading indicates a higher probability of early onset AR events from the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> Range #1 to occur over dry soil and moist boundary layer conditions as compared to early onset AR events from Ranges # 2–10. All plots show the mean relationship over the five small domains illustrated in <a href="#remotesensing-10-00969-f001" class="html-fig">Figure 1</a>a. Red crosses place the lowest 10th percentile (i.e., the strongest) negative soil moisture gradients from all five domains in the corresponding parameter space. The selected boundary layer parameters from left to right are: Relative Humidity (RH, %) (averaged between 850 and 600 hPa), Lifting Condensation Level (LCL, hPa) and Boundary Layer Height (BLH, hPa).</p>
Full article ">Figure A1
<p>Correspondence of the number of rain pixels per day in TRMM-3B42 and CMORPH-v1.0 with Morning accumulated Rain (MR) or Afternoon accumulated Rain (AR) exceeding a threshold of 1 mm/9 h (<b>a</b>,<b>b</b>) or 0 mm/9 h (<b>c</b>,<b>d</b>).</p>
Full article ">Figure A2
<p>(<b>a</b>) Probability histogram of AR system size estimated for the TRMM-3B42 and CMORPH-v1.0 datasets. The AR size parameter here is calculated as a percentage of AR pixels within an event box of 9 × 9 pixels; the 9 × 9 deg box is commensurate with the 90th percentile of all possible meso-scale convective system size values, i.e., 200,000 km<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math> [<a href="#B72-remotesensing-10-00969" class="html-bibr">72</a>], and therefore should be well suited for capturing the range of possible AR areas. (<b>b</b>) Spearman’s rank correlation between the soil moisture gradient and the onset time of afternoon rainfall calculated for each of the ten LocS ranges for either the TRMM-3B42 (red) or CMORPH-v1.0 (blue) dataset. Correlation is first estimated in every 5 × 5 degree grid box as in <a href="#remotesensing-10-00969-f004" class="html-fig">Figure 4</a> and then averaged across the domain.</p>
Full article ">Figure A3
<p>Number of identified afternoon rain events aggregated over 1 × 1<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math> grid boxes (gray shading) and applied orography mask at the 0.25<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math> horizontal resolution (red dots).</p>
Full article ">Figure A4
<p>Correspondence between the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> value and the spatial gradient in absolute soil moisture estimated in the same <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </semantics></math> locations. Color shading indicates the magnitude of absolute soil moisture in the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </semantics></math> location only. The value of Spearman’s rank correlation coefficient calculated between the <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>o</mi> <mi>c</mi> <mi>S</mi> </mrow> </semantics></math> value and the spatial gradient is given.</p>
Full article ">Figure A5
<p>Same as <a href="#remotesensing-10-00969-f006" class="html-fig">Figure 6</a>a–c, but for every African domain separately. Note, that pressure levels on the <span class="html-italic">y</span>-axis of LCL and BLH plots have a reverse order.</p>
Full article ">Figure A6
<p>Same as <a href="#remotesensing-10-00969-f006" class="html-fig">Figure 6</a>d–f, but for every African domain separately. Note, that pressure levels on the <span class="html-italic">y</span>-axis of LCL and BLH plots have a reverse order.</p>
Full article ">
20 pages, 4745 KiB  
Article
The Use of Massive Deformation Datasets for the Analysis of Spatial and Temporal Evolution of Mauna Loa Volcano (Hawai’i)
by Susi Pepe, Luca D’Auria, Raffaele Castaldo, Francesco Casu, Claudio De Luca, Vincenzo De Novellis, Eugenio Sansosti, Giuseppe Solaro and Pietro Tizzani
Remote Sens. 2018, 10(6), 968; https://doi.org/10.3390/rs10060968 - 17 Jun 2018
Cited by 12 | Viewed by 6434
Abstract
In this work, we exploited large DInSAR and GPS datasets to create a 4D image of the magma transfer processes at Mauna Loa Volcano (Island of Hawai’i) from 2005 to 2015. The datasets consist of 23 continuous GPS time series and 307 SAR [...] Read more.
In this work, we exploited large DInSAR and GPS datasets to create a 4D image of the magma transfer processes at Mauna Loa Volcano (Island of Hawai’i) from 2005 to 2015. The datasets consist of 23 continuous GPS time series and 307 SAR images acquired from ascending and descending orbits by ENVISAT (ENV) and COSMO-SkyMed (CSK) satellites. Our results highlight how the joint use of SAR data acquired from different orbits (thus with different look angles and wavelengths), together with deformation data from GPS networks and geological information can significantly improve the constraints on the geometry and location of the sources responsible for the observed deformation. The analysis of these datasets has been performed by using an innovative method that allows building a complex source configuration. The results suggest that the deformation pattern observed from 2005 to 2015 has been controlled by three deformation sources: the ascent of magma along a conduit, the opening of a dike and the slip along the basal decollement. This confirms that the intrusion of the magma within a tabular system (rift dikes) may trigger the sliding of the SE portion of the volcanic edifice along the basal decollement. This case study confirms that it is now possible to exploit large geodetic datasets to improve our knowledge of volcano dynamics. The same approach could also be easily applied in other geodynamical contexts such as geothermal reservoirs and regions with complex tectonics. Full article
(This article belongs to the Special Issue Remote Sensing of Tectonic Deformation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Hawai’i map. (<b>A</b>) Hawai’i Island. The main volcanoes are indicated; the triangles coincide with Mouna Loa, Mauna Kea and Kilauea summits. The red line represents the projection of rift system on the surface. (<b>B</b>) Subsurface geological elements, as suggested by [<a href="#B9-remotesensing-10-00968" class="html-bibr">9</a>,<a href="#B10-remotesensing-10-00968" class="html-bibr">10</a>].</p>
Full article ">Figure 2
<p>GPS data: (<b>A</b>) GPS network on the Mauna Loa Volcano; and (<b>B</b>) GPS data availability for each station between 2003 and 2015. The horizontal displacements of GPS stations are shown in red (by the Nevada Geodetic Laboratory of the Hawai’i monitoring network).</p>
Full article ">Figure 3
<p>Multiplatform DInSAR data. Ascending (red traces) and Descending (green traces) for: ENVISAT (<b>A</b>,<b>B</b>); and COSMO-SkyMed (CSK) (<b>C</b>) satellites; and (<b>D</b>) temporal distribution of images acquired from each satellites. The ID numbers reported in (<b>D</b>) represent the track numbers of the considered orbits. The ENVISAT and CSK temporal coverage are indicated with light blue and yellow, respectively. Moreover, the blue vertical line indicates the reference time <span class="html-italic">t</span><sub>0</sub>.</p>
Full article ">Figure 4
<p>LOS mean velocity maps: (<b>A</b>–<b>C</b>) ENVISAT ascending with different swath from I2 (about 18°) to I4 (about 28°); (<b>D</b>–<b>G</b>) ENVISAT descending with different swath from I1 (about 15°) to I7 (about 38°); and (<b>H</b>–<b>I</b>) COSMO-SkyMed ascending and descending orbits with look angle of about 51° and 67°, respectively.</p>
Full article ">Figure 5
<p>Tridimensional representation of the inversion results. The top panel represents the geometry of the best-fit model configuration (Model 1, <a href="#remotesensing-10-00968-t002" class="html-table">Table 2</a>). The color scale (on the right) indicates the volumetric variation of the pipe and the dike system. The scale for the slip vectors along the decollement is indicated on the top panel. The panels in the mid and bottom rows are relative to four selected periods. All the depths are relative to the Mauna Loa summit.</p>
Full article ">Figure 6
<p>Comparison between data and model: DInSAR LOS time series. The plot represents the LOS deformation time series for the pixels of each dataset closest the MLSP GPS station. Black crosses represent the data, while red circles represent the model.</p>
Full article ">Figure 7
<p>Comparison between data and model: GPS vertical components. GPS vertical components are indicated with black lines, the model with red lines, and the labels are the GPS stations.</p>
Full article ">Figure 8
<p>Comparison between data and model: GPS horizontal components. GPS horizontal displacements are indicated by black lines, model results by red lines, and GPS stations are indicated by blue dots. The scale of GPS displacements has an exaggeration of 10<sup>5</sup>.</p>
Full article ">Figure 9
<p>Volume changes variation of the sources. Time series of the volume changes for the pipe (red line) and for the dike systems along the rift (blue line) versus the seismic moment released by the slip along the decollement (black line).</p>
Full article ">Figure 10
<p>Volumetric variation of the pipe source as a function of the time and depth. Horizontal black lines separate the four segments of the modeled conduit system. The vertical lines mark the intervals used for the interpolation. The depth is relative to the ground surface.</p>
Full article ">Figure 11
<p>Volumetric variation along the dike system for four considered time intervals. The dike system has been unwrapped: the side on the left of the pipe is the southern branch, while the other is the eastern branch. The depth is relative to the ground surface.</p>
Full article ">Figure 12
<p>Volume variation along the dike system as a function of the time and length. Horizontal black lines mark the separation between 20 segments of the modeled system. The vertical lines mark the intervals used for the interpolation. The dashed black line marks the position of the central pipe.</p>
Full article ">
16 pages, 6854 KiB  
Article
Investigation of Precipitable Water Vapor Obtained by Raman Lidar and Comprehensive Analyses with Meteorological Parameters in Xi’an
by Yufeng Wang, Liu Tang, Jing Zhang, Tianle Gao, Qing Wang, Yuehui Song and Dengxin Hua
Remote Sens. 2018, 10(6), 967; https://doi.org/10.3390/rs10060967 - 17 Jun 2018
Cited by 8 | Viewed by 4630
Abstract
To evaluate the potential of Raman lidar observations for measuring precipitable water vapor (PWV), PWV variations and distribution characteristics were investigated in Xi’an (34.233°N, 108.911°E), and its comparisons with meteorological parameters were also analysed. Comparisons of lidar PWV with radiosonde PWV verified the [...] Read more.
To evaluate the potential of Raman lidar observations for measuring precipitable water vapor (PWV), PWV variations and distribution characteristics were investigated in Xi’an (34.233°N, 108.911°E), and its comparisons with meteorological parameters were also analysed. Comparisons of lidar PWV with radiosonde PWV verified the ability and accuracy of using Raman lidars for PWV measurements. The diurnal and monthly variation trends in PWV in different layers are first discussed via the statistical analysis of lidar data from November 2013 to July 2016; different proportions of PWV were found in different layers, and the PWV in each layer presented a slight diurnal change trend and consistent seasonal variation, which was relatively rich in summer, less so in spring and autumn, and relatively deficient in winter. Furthermore, correlation analyses between lidar PWV and meteorological parameters are explored. Water vapor pressure and surface temperature revealed the same inter-seasonal oscillation of PWV, with a correlation coefficient of ~0.90. However, incomplete synchronization was found between PWV and relative humidity and precipitation parameters. Higher humidity appeared in the late summer and the beginning of autumn of each year, which was also the case for precipitation and precipitation efficiency. In addition, atmospheric water vapor density profiles and the obtained PWV by Raman lidar are discussed employing a rainfall case, and a comprehensive analysis with meteorological parameters is undertaken. The intensifying characteristics of vertical change in water vapor and the accumulation of PWV in the lower troposphere can be captured by lidar before the onset of rainfall. In contrast to the obvious diurnal change trend, such meteorological parameters as relative humidity, water vapor pressure, and dew-point temperature difference are accompanied with stable trends with a change rate of close to 0 in the rainfall processes; they also show high correlated variations with lidar PWV. Thus, with the advantage of lidar detection, investigation of water vapor profiles and PWV by Raman lidar, and the comprehensive correlation analyses with synchronic meteorological parameters can prove to be good indications of rainfall. Full article
(This article belongs to the Special Issue Remote Sensing of Atmospheric Properties)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of the Raman-Mie lidar system: DM-dichroic mirror; BS-beam splitter; IF-interference filter; L-lens.</p>
Full article ">Figure 2
<p>The measured Range-corrected signal of lidar returns, the retrieved water vapor density profile, and comparisons with radiosonde data taken at 20:00 CST on 16 November 2013.</p>
Full article ">Figure 3
<p>Scatter plot of correlation between Lidar PWV and radiosonde PWV.</p>
Full article ">Figure 4
<p>THI plot of water vapor mixing ratio profiles during 20:30–06:00 LST on 16 November 2016.</p>
Full article ">Figure 5
<p>Temporal and spatial variations in PWV; (<b>a</b>) PWV in each layer; (<b>b</b>) its proportion in each layer.</p>
Full article ">Figure 6
<p>Seasonal variations in the total PWV from September 2013 to July 2016.</p>
Full article ">Figure 7
<p>Seasonal variations in PWV in different layers from September 2013 to July 2016.</p>
Full article ">Figure 8
<p>Variations between PWV and atmospheric temperature, vapor pressure and humidity. (<b>a</b>) temperature; (<b>b</b>) vapor pressure; (<b>c</b>) relative humidity.</p>
Full article ">Figure 9
<p>Correlations between PWV and atmospheric temperature, vapor pressure and humidity. (<b>a</b>) temperature; (<b>b</b>) vapor pressure; (<b>c</b>) relative humidity.</p>
Full article ">Figure 10
<p>The seasonal variation trends in PWV and ground meteorological elements.</p>
Full article ">Figure 11
<p>The seasonal variations in averaged PWV, precipitation and precipitation efficiency.</p>
Full article ">Figure 12
<p>Several representative atmospheric water vapor density profiles during the rainfall process of 1–10 October 2015.</p>
Full article ">Figure 13
<p>Time series of the lidar PWV and surface water vapour density during 1–10 October 2015.</p>
Full article ">Figure 14
<p>Time series of meteorological parameters and the corresponding change rates during rainfall case. (<b>a</b>) dew-point temperature difference; (<b>b</b>) vapor pressure; (<b>c</b>) relative humidity.</p>
Full article ">Figure 15
<p>Correlation analysis between PWV and meterological parameters during the rainfall case.</p>
Full article ">
20 pages, 7050 KiB  
Article
Estimation of Water Level Changes of Large-Scale Amazon Wetlands Using ALOS2 ScanSAR Differential Interferometry
by Ning Cao, Hyongki Lee, Hahn Chul Jung and Hanwen Yu
Remote Sens. 2018, 10(6), 966; https://doi.org/10.3390/rs10060966 - 17 Jun 2018
Cited by 25 | Viewed by 5795
Abstract
Differential synthetic aperture radar (SAR) interferometry (DInSAR) has been successfully used to estimate water level changes (∂h/∂t) over wetlands and floodplains. Specifically, amongst ALOS PALSAR datasets, the fine-beam stripmap mode has been mostly implemented to estimate ∂h/∂t due to its availability of multitemporal [...] Read more.
Differential synthetic aperture radar (SAR) interferometry (DInSAR) has been successfully used to estimate water level changes (∂h/∂t) over wetlands and floodplains. Specifically, amongst ALOS PALSAR datasets, the fine-beam stripmap mode has been mostly implemented to estimate ∂h/∂t due to its availability of multitemporal images. However, the fine-beam observation mode provides limited swath coverage to study large floodplains and wetlands, such as the Amazon floodplains. Therefore, for the first time, this paper demonstrates that ALOS2 ScanSAR data can be used to estimate the large-scale ∂h/∂t in Amazon floodplains. The basic procedures and challenges of DInSAR processing with ALOS2 ScanSAR data are addressed and final ∂h/∂t maps are generated based on the Satellite with ARgos and ALtiKa (SARAL) altimetry’s reference data. This study reveals that the local ∂h/∂t patterns of Amazon floodplains are spatially complex with highly interconnected floodplain channels, but the large-scale (with 350 km swath) ∂h/∂t patterns are simply characterized by river water flow directions. Full article
(This article belongs to the Special Issue Imaging Geodesy and Infrastructure Monitoring)
Show Figures

Figure 1

Figure 1
<p>Study areas of Amazon floodplains with ALOS PALSAR fine-beam wrapped differential interferograms acquired between April and July 2010. Each fringe represents about 12 cm of water level changes in line-of-sight (LOS) direction. Black and white rectangles are the coverage of the ALOS PALSAR ScanSAR and ALOS2 PALSAR2 ScanSAR datasets used in this study.</p>
Full article ">Figure 2
<p>ScanSAR wrapped differential interferograms. (<b>a</b>) ALOS PALSAR ScanSAR with acquisition dates of 20090522–20090707; (<b>b</b>) ALOS2 ScanSAR with acquisition dates of 20150628–20150726.</p>
Full article ">Figure 3
<p>Flowchart of ALOS2 ScanSAR DInSAR processing.</p>
Full article ">Figure 4
<p>Wrapped differential interferograms of 20150517–20150628 pair before (<b>a</b>) and after (<b>b</b>) baseline error removal.</p>
Full article ">Figure 5
<p>Ionospheric correction results. (<b>a</b>) Original differential interferogram; (<b>b</b>) MAI interferogram; (<b>c</b>) Ionospheric phase screen; (<b>d</b>) Corrected interferogram.</p>
Full article ">Figure 6
<p>(<b>a</b>) A segment of wrapped differential interferogram of pair 20150628–20150726; (<b>b</b>) Unwrapped result without masking; (<b>c</b>) Unwrapped result with masking. The black dots are the reference point (with phase set to be zero) for phase unwrapping.</p>
Full article ">Figure 7
<p>SRTM DEM of the study area. Blue lines indicate segments of the SARAL altimetry tracks used in this study.</p>
Full article ">Figure 8
<p>Water levels measured by SARAL altimetry for year 2015. (<b>a</b>–<b>d</b>) indicates the floodplains of the four rivers. Labels (1)–(16) indicate the 16 segments of the altimetry tracks as shown in <a href="#remotesensing-10-00966-f007" class="html-fig">Figure 7</a>. For every segment, the altimetry water levels are averaged.</p>
Full article ">Figure 9
<p>Final wrapped differential interferograms. Three white rectangles in (<b>a</b>) are zoomed in as shown in <a href="#remotesensing-10-00966-f010" class="html-fig">Figure 10</a>. The time span between master and slave dates for the four interferograms are 42, 42, 42, and 28 days, respectively.</p>
Full article ">Figure 10
<p>Differential interferograms for three zoomed-in areas (<b>i</b>), (<b>ii</b>), and (<b>iii</b>) labeled in <a href="#remotesensing-10-00966-f009" class="html-fig">Figure 9</a>. SRTM DEMs are also included in (<b>a</b>).</p>
Full article ">Figure 11
<p>Final DInSAR ∂h/∂t maps. The black lines in (<b>a</b>) are profiles along the floodplains which will be examined for further analysis.</p>
Full article ">Figure 12
<p>Comparison of DInSAR ∂h/∂t with altimetry ∂h/∂t. The parameters p and r2 represent the slopes of the least-square fitted lines and <span class="html-italic">R</span><sup>2</sup> values, respectively.</p>
Full article ">Figure 13
<p>DInSAR ∂h/∂t along the center profile of the mainstem floodplain shown in <a href="#remotesensing-10-00966-f011" class="html-fig">Figure 11</a>a.</p>
Full article ">Figure 14
<p>DInSAR ∂h/∂t along the center profile of the Japura floodplain shown in <a href="#remotesensing-10-00966-f011" class="html-fig">Figure 11</a>a.</p>
Full article ">Figure 15
<p>DInSAR ∂h/∂t along the center profile of the Jutai floodplain shown in <a href="#remotesensing-10-00966-f011" class="html-fig">Figure 11</a>a.</p>
Full article ">Figure 16
<p>DInSAR ∂h/∂t along the center profile of the Jurua floodplain shown in <a href="#remotesensing-10-00966-f011" class="html-fig">Figure 11</a>a.</p>
Full article ">
14 pages, 3264 KiB  
Article
The Heterogeneity of Air Temperature in Urban Residential Neighborhoods and Its Relationship with the Surrounding Greenspace
by Yuguo Qian, Weiqi Zhou, Xiaofang Hu and Fan Fu
Remote Sens. 2018, 10(6), 965; https://doi.org/10.3390/rs10060965 - 16 Jun 2018
Cited by 23 | Viewed by 4830
Abstract
The thermal environment in residential areas is directly related to the living quality of residents. Therefore, it is important to understand thermal heterogeneity and ways to regulate temperature in residential neighborhoods. We investigated the spatial heterogeneity and temporal dynamics of air temperatures in [...] Read more.
The thermal environment in residential areas is directly related to the living quality of residents. Therefore, it is important to understand thermal heterogeneity and ways to regulate temperature in residential neighborhoods. We investigated the spatial heterogeneity and temporal dynamics of air temperatures in 20 residential neighborhoods within the 5th ring road of Beijing, China. We further explored how the variations in air temperature were related to the patterns of the surrounding greenspace at different scales. We found that: (1) large air temperature differences existed among residential neighborhoods, with hourly maximum differences in air temperature reaching 5.30 °C on hot summer days; (2) not only the percentage but also the spatial configuration (e.g., edge density) of greenspace affected the local air temperature; and (3) the effects of spatial greenspace patterns on air temperature were scale dependent and varied by season. For example, increasing the proportion of greenspace in surrounding areas within a 100-m radius and increasing the edge density within radii from 500 to 1000 m could lower air temperatures in summer but not affect air temperatures in winter. In addition, decreasing the edge density of greenspaces within a 100-m radius of the surrounding areas would lead to an increase in air temperature in winter but not affect the air temperature in summer. These results extend our understanding of thermal environments and their relationships with greenspace patterns at the microscale (i.e., residential neighborhoods). They also provide useful information for urban planners to optimize greenspace patterns under better thermal conditions at the neighborhood scale. Full article
(This article belongs to the Special Issue Remote Sensing of Urban Ecology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area showing the: (<b>a</b>) spatial distribution of the 20 residential neighborhoods, with an elevation above sea level (EASL) ranging from 41 to 61 m, and the corresponding land cover map; (<b>b</b>) field air temperature measurements using HOBO loggers covered by solar radiation shields; and (<b>c</b>) circles with different radii around the monitoring sites.</p>
Full article ">Figure 2
<p>Hourly air temperature variations over 20 neighborhoods in summer.</p>
Full article ">Figure 3
<p>Hourly air temperature variations over the 20 neighborhoods in winter.</p>
Full article ">Figure 4
<p>Diurnal temperature ranges (DTRs) for air temperature in the 20 neighborhoods.</p>
Full article ">Figure 5
<p>The hourly maximum differences in air temperature (MD) and the hourly standard deviation of air temperature (SD) in summer.</p>
Full article ">Figure 6
<p>The hourly maximum differences in air temperature (MD) and the hourly standard deviation of air temperature (SD) in winter.</p>
Full article ">
13 pages, 2289 KiB  
Article
Performance Evaluation of Single-Label and Multi-Label Remote Sensing Image Retrieval Using a Dense Labeling Dataset
by Zhenfeng Shao, Ke Yang and Weixun Zhou
Remote Sens. 2018, 10(6), 964; https://doi.org/10.3390/rs10060964 - 16 Jun 2018
Cited by 127 | Viewed by 9966 | Correction
Abstract
Benchmark datasets are essential for developing and evaluating remote sensing image retrieval (RSIR) approaches. However, most of the existing datasets are single-labeled, with each image in these datasets being annotated by a single label representing the most significant semantic content of the image. [...] Read more.
Benchmark datasets are essential for developing and evaluating remote sensing image retrieval (RSIR) approaches. However, most of the existing datasets are single-labeled, with each image in these datasets being annotated by a single label representing the most significant semantic content of the image. This is sufficient for simple problems, such as distinguishing between a building and a beach, but multiple labels and sometimes even dense (pixel) labels are required for more complex problems, such as RSIR and semantic segmentation.We therefore extended the existing multi-labeled dataset collected for multi-label RSIR and presented a dense labeling remote sensing dataset termed "DLRSD". DLRSD contained a total of 17 classes, and the pixels of each image were assigned with 17 pre-defined labels. We used DLRSD to evaluate the performance of RSIR methods ranging from traditional handcrafted feature-based methods to deep learning-based ones. More specifically, we evaluated the performances of RSIR methods from both single-label and multi-label perspectives. These results demonstrated the advantages of multiple labels over single labels for interpreting complex remote sensing images. DLRSD provided the literature a benchmark for RSIR and other pixel-based problems such as semantic segmentation. Full article
(This article belongs to the Special Issue Deep Learning for Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of single-label and multi-label RSIR.</p>
Full article ">Figure 2
<p>Example images and corresponding labeling results.</p>
Full article ">Figure 3
<p>The precision-recall curves for single-label and multi-label RSIR.</p>
Full article ">Figure 4
<p>The results of CNN features for each class in DLRSD.</p>
Full article ">
22 pages, 9275 KiB  
Article
A Seismic Capacity Evaluation Approach for Architectural Heritage Using Finite Element Analysis of Three-Dimensional Model: A Case Study of the Limestone Hall in the Ming Dynasty
by Siliang Chen, Shaohua Wang, Chen Li, Qingwu Hu and Hongjun Yang
Remote Sens. 2018, 10(6), 963; https://doi.org/10.3390/rs10060963 - 15 Jun 2018
Cited by 12 | Viewed by 5630
Abstract
A lot of architectural heritage in China are urgently in need to carry out seismic assessment for further conservation. In this paper, a seismic capacity evaluation approach for architectural heritage using finite element analysis with precision three-dimensional data was proposed. The Limestone Hall [...] Read more.
A lot of architectural heritage in China are urgently in need to carry out seismic assessment for further conservation. In this paper, a seismic capacity evaluation approach for architectural heritage using finite element analysis with precision three-dimensional data was proposed. The Limestone Hall of Shaanxi Province was taken as an example. First, low attitude unmanned aerial vehicle photogrammetry and a close-range photogrammetry camera were used to collect multiple view images to obtain the precision three-dimensional current model of the Limestone. Second, the dimensions of internal structures of Limestone Hall are obtained by means of structural analysis; re-establishing the ideal model of Limestone Hall based on the modeling software. Third, a finite element analysis was conducted to find out the natural frequency and seismic stress in various conditions with the 3D model using ANSYS software. Finally, the seismic capacity analysis results were comprehensively evaluated for the risk assessment and simulation. The results showed that for architectural heritage with a multilayer structure, utilizing photogrammetric surveying and mapping, 3D software modeling, finite element software simulation, and seismic evaluation for simulation was feasible where the precision of the modeling and parameters determine the accuracy of the simulation. The precise degree of the three-dimensional model, the accurate degree of parameter measurement and estimation, the setting of component attributes in the finite element model and the strategy of finite element analysis have an important effect on the result of seismic assessment. The main body structure of the Limestone Hall could resist an VII-degree earthquake at most, and the ridge of the second floor could not resist a V-degree earthquake due to unsupported conditions. The maximum deformation of the Limestone Hall during the earthquake occurred in the tabia layer below the second roof. Full article
(This article belongs to the Special Issue Advances in Remote Sensing for Archaeological Heritage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>General layout orthophoto of the Limestone Hall.</p>
Full article ">Figure 2
<p>Full view of the Limestone Hall and direction of the southern entrance.</p>
Full article ">Figure 3
<p>Detail of the structure of the Limestone Hall (<b>a</b>) the arch in the south of Limestone hall, (<b>b</b>) interior wall and inclined beam of Limestone hall, and (<b>c</b>) the octagonal dome in the interior of the Limestone Hall.</p>
Full article ">Figure 4
<p>Tabialayer exposed from internal stone damage of the Limestone Hall.</p>
Full article ">Figure 5
<p>Digitization and earthquake resistance analysis flow diagram of architectural heritage with multilayer structures.</p>
Full article ">Figure 6
<p>Mapping and modeling flow diagram of architectural heritage with multilayer structures.</p>
Full article ">Figure 7
<p>Layout diagram of the external control points of the Limestone Hall.</p>
Full article ">Figure 8
<p>Measurable model established by smart3d software (<b>a</b>) Measurement for external model; (<b>b</b>) Measurement for the internal model.</p>
Full article ">Figure 9
<p>Structure analysis of architectural heritage with multilayer structure and flow diagram of model reconstruction.</p>
Full article ">Figure 10
<p>Structural layer analysis of limestone hall and dimension measurement of all layers. (<b>a</b>) Out-wall crack in the north of limestone hall; (<b>b</b>) Architectural structure seen from crack of north wall; (<b>c</b>) Architectural structure seen from south of inside limestone hall.</p>
Full article ">Figure 11
<p>Construction process diagram of the Limestone Hall.</p>
Full article ">Figure 12
<p>Profile analysis diagram of the Limestone Hall with a five-layer structure.</p>
Full article ">Figure 13
<p>Finite element modeling and earthquake resistance analysis process diagram of architectural heritage with multilayer structure.</p>
Full article ">Figure 14
<p>Finite element model of limestone hall established in ANSYS. (<b>a</b>) Mesh diagram of external structure of limestone hall; (<b>b</b>) Mesh diagram of internal structure of Limestone Hall.</p>
Full article ">Figure 15
<p>Displacement nephogram and stress nephogram of the Limestone Hall with an ideal status under three loads of gravity, earthquake, snow, and the combined load.</p>
Full article ">Figure 16
<p>Damage conditions of limestone hall at present. (<b>a</b>) Ridge damage in the south and roof sedimentation; (<b>b</b>) Northern roof sedimentation and crack on the wall.</p>
Full article ">Figure 17
<p>Elevation diagram of the current situation of the Limestone Hall (currentstatus).</p>
Full article ">Figure 18
<p>Separable model, elevation diagram, and section diagram of the Limestone Hall obtained by recovery (ideal status).</p>
Full article ">Figure 19
<p>The first 10-order intrinsic frequency and vibration model diagram of the Limestone Hall under ideal status.</p>
Full article ">Figure 20
<p>Displacement diagram and stress diagram of the Limestone Hall under ideal status with earthquake accelerations of 0.15 g and 0.2 g.</p>
Full article ">Figure 21
<p>Displacement diagram and stress diagram of the Limestone Hall under current status with earthquake accelerations of 0.15 g and 0.2 g.</p>
Full article ">
25 pages, 8668 KiB  
Article
Winter Wheat Production Estimation Based on Environmental Stress Factors from Satellite Observations
by Juan Sui, Qiming Qin, Huazhong Ren, Yuanheng Sun, Tianyuan Zhang, Jiandong Wang and Shihong Gong
Remote Sens. 2018, 10(6), 962; https://doi.org/10.3390/rs10060962 - 15 Jun 2018
Cited by 21 | Viewed by 5003
Abstract
The rapid and accurate estimation of wheat production at a regional scale is crucial for national food security and sustainable agricultural development. This study developed a new gross primary productivity (GPP) estimation model (denoted as the [ACPM]), based on the effects of light, [...] Read more.
The rapid and accurate estimation of wheat production at a regional scale is crucial for national food security and sustainable agricultural development. This study developed a new gross primary productivity (GPP) estimation model (denoted as the [ACPM]), based on the effects of light, heat, soil moisture, and nitrogen content (N) on the light-use efficiency of winter wheat. The ACPM model used the quantic additivity of the environmental factors to improve the minimum form or multiple multiplication form in the previous model and thus characterized the joint effects of heat, soil moisture, and N on crop photosynthesis performance. The key parameters (i.e., light) were determined from the photosynthetically active radiation product of the Himawari-8 sensor and the fraction of photosynthetically active radiation product of Moderate Resolution Imaging Spectroradiometer (MODIS). The heat was determined from the land temperature products of MODIS. The soil moisture was obtained from the inversion using a visible and shortwave infrared drought index (VSDI), whereas the N stress of winter wheat was detected using the newly developed modified ratio vegetation index (MRVI), which could accurately obtain the spatiotemporal distribution of the leaf chlorophyll content of winter wheat. The ACPM and two other previous models (named the GPP1 and GPP2 models) were applied on the Himawari-8 and MODIS images in Hengshui City. The evaluation results, based on the ground measurement, indicated that the ACPM models exhibited the best estimate of dry aboveground biomass (DAM) and the wheat yield in Hengshui City, with errors of <10% and <12% for the DAM and yield, respectively. Considering the easy operation of the ACPM model and the accessibility of the corresponding satellite images, the Agriculture Crop Photosynthesis Model (ACPM) can be expected to provide information on the winter wheat shortfalls and surplus ahead of the availability of official statistical data. 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>The location of the Yucheng experimental station, study area, and experimental plots in the study area as follows: (<b>a</b>) the map of the Yucheng experimental station and study area; (<b>b</b>) experimental plots in study area (exp2); and (<b>c</b>) the scheme of the sub-plots.</p>
Full article ">Figure 2
<p>Winter wheat sowing area of Hengshui City with the computed percent of the wheat value (% of a pixel occupied) at 1 km scale in 2017.</p>
Full article ">Figure 3
<p>The schema of this research procedure for the winter wheat production estimation based on the environmental stress factors from the satellite observations. MRVI—modified ratio vegetation index; exp1—experiment 1; exp2—experiment 2; DAM—dry aboveground biomass; GPP—gross primary productivity; MODIS—Moderate Resolution Imaging Spectroradiometer; VSDI—visible and shortwave infrared drought index; WDRVI—wide dynamic range vegetation index; GNDVI—green normalized difference vegetation index.</p>
Full article ">Figure 4
<p>Relationship between the leaf C<sub>ab</sub> and leaf total nitrogen content (TN) of winter wheat in exp1 at Yucheng Station.</p>
Full article ">Figure 5
<p>The winter wheat canopy reflectance and leaf C<sub>ab</sub> with different N rates during the winter wheat growth period in exp1 at Yucheng Station, as follows: the measured canopy spectral reflectance on (<b>a</b>) 4 April 2011; (<b>b</b>) 12 April 2011; (<b>c</b>) 30 April 2011; and (<b>d</b>) 11 May 2011, and (<b>e</b>) was the measured leaf C<sub>ab</sub>.</p>
Full article ">Figure 5 Cont.
<p>The winter wheat canopy reflectance and leaf C<sub>ab</sub> with different N rates during the winter wheat growth period in exp1 at Yucheng Station, as follows: the measured canopy spectral reflectance on (<b>a</b>) 4 April 2011; (<b>b</b>) 12 April 2011; (<b>c</b>) 30 April 2011; and (<b>d</b>) 11 May 2011, and (<b>e</b>) was the measured leaf C<sub>ab</sub>.</p>
Full article ">Figure 6
<p>The relationship between the canopy reflectance and leaf C<sub>ab</sub> of winter wheat in exp1 at Yucheng Station, as follows: (<b>a</b>) the reflectance at the red band and leaf C<sub>ab</sub>; (<b>b</b>) the reflectance at the near-infrared (NIR) band and leaf C<sub>ab</sub>; (<b>c</b>) the reflectance at the blue band and leaf C<sub>ab</sub>; (<b>d</b>) the reflectance at the green band and leaf C<sub>ab</sub>; (<b>e</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>blue</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msup> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>green</mi> </mrow> </msub> <mo>−</mo> <msub> <mi>ρ</mi> <mrow> <mi>blue</mi> </mrow> </msub> </mrow> <mo>)</mo> </mrow> </mrow> <mn>2</mn> </msup> </mrow> </mrow> </mrow> </semantics></math> and the leaf C<sub>ab</sub>; (<b>f</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>green</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msup> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>green</mi> </mrow> </msub> <mo>−</mo> <msub> <mi>ρ</mi> <mrow> <mi>blue</mi> </mrow> </msub> </mrow> <mo>)</mo> </mrow> </mrow> <mn>2</mn> </msup> </mrow> </mrow> </mrow> </semantics></math> and the leaf C<sub>ab</sub>; (<b>g</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>red</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msup> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>red</mi> </mrow> </msub> <mo>−</mo> <msub> <mi>ρ</mi> <mrow> <mi>green</mi> </mrow> </msub> </mrow> <mo>)</mo> </mrow> </mrow> <mn>2</mn> </msup> </mrow> </mrow> </mrow> </semantics></math> and the leaf C<sub>ab</sub>; and (<b>h</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>green</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msup> <mrow> <mrow> <mo>(</mo> <mrow> <msub> <mi>ρ</mi> <mrow> <mi>red</mi> </mrow> </msub> <mo>−</mo> <msub> <mi>ρ</mi> <mrow> <mi>green</mi> </mrow> </msub> </mrow> <mo>)</mo> </mrow> </mrow> <mn>2</mn> </msup> </mrow> </mrow> </mrow> </semantics></math> and the leaf C<sub>ab</sub>.</p>
Full article ">Figure 7
<p>Relationship between VIs and leaf C<sub>ab</sub> in pure winter wheat pixels of exp2 in Hengshui City, as follows: (<b>a</b>) MRVI and C<sub>ab</sub>; (<b>b</b>) <span class="html-italic">ρ</span><sub>NIR</sub>/<span class="html-italic">ρ</span><sub>blue</sub> and C<sub>ab</sub>; (<b>c</b>) RVI2 and C<sub>ab</sub>; (<b>d</b>) SIPI and C<sub>ab</sub>; and (<b>e</b>) MSR and C<sub>ab</sub>.</p>
Full article ">Figure 7 Cont.
<p>Relationship between VIs and leaf C<sub>ab</sub> in pure winter wheat pixels of exp2 in Hengshui City, as follows: (<b>a</b>) MRVI and C<sub>ab</sub>; (<b>b</b>) <span class="html-italic">ρ</span><sub>NIR</sub>/<span class="html-italic">ρ</span><sub>blue</sub> and C<sub>ab</sub>; (<b>c</b>) RVI2 and C<sub>ab</sub>; (<b>d</b>) SIPI and C<sub>ab</sub>; and (<b>e</b>) MSR and C<sub>ab</sub>.</p>
Full article ">Figure 8
<p>Relationship between VIs and leaf C<sub>ab</sub> in all winter wheat pixels (pure and mixed pixels) of exp2 in Hengshui City: (<b>a</b>) MRVI and C<sub>ab</sub>; (<b>b</b>) <span class="html-italic">ρ</span><sub>NIR</sub>/<span class="html-italic">ρ</span><sub>blue</sub> and C<sub>ab</sub>; (<b>c</b>) RVI2 and C<sub>ab</sub>; (<b>d</b>) SIPI and C<sub>ab</sub>; (<b>e</b>) MSR and C<sub>ab</sub>.</p>
Full article ">Figure 9
<p>Distribution of the simulated DAM of the winter wheat with a 1 km resolution at the senescence stage, in Hengshui City in 2017. Simulated DAM based on the (<b>a</b>) GPP1 model, (<b>b</b>) GPP2 model, and (<b>c</b>) Agriculture Crop Photosynthesis Model (ACPM) model.</p>
Full article ">Figure 10
<p>Validation of the estimated winter wheat DAM based on the GPP1, GPP2, and ACPM models in Hengshui City in 2017, as follows: (<b>a</b>) GPP1 model in the pure pixels of the experimental plots in exp2; (<b>b</b>) GPP1 model in the pure and mixed pixels of the experimental plots in exp2; (<b>c</b>) GPP2 model in the pure pixels of the experimental plots in exp2; (<b>d</b>) GPP2 model in the pure and mixed pixels of the experimental plots in exp2; (<b>e</b>) ACPM model in the pure pixels of the experimental plots in exp2; and (<b>f</b>) ACPM model in the pure and mixed pixels of the experimental plots in exp2; e is the error.</p>
Full article ">Figure 10 Cont.
<p>Validation of the estimated winter wheat DAM based on the GPP1, GPP2, and ACPM models in Hengshui City in 2017, as follows: (<b>a</b>) GPP1 model in the pure pixels of the experimental plots in exp2; (<b>b</b>) GPP1 model in the pure and mixed pixels of the experimental plots in exp2; (<b>c</b>) GPP2 model in the pure pixels of the experimental plots in exp2; (<b>d</b>) GPP2 model in the pure and mixed pixels of the experimental plots in exp2; (<b>e</b>) ACPM model in the pure pixels of the experimental plots in exp2; and (<b>f</b>) ACPM model in the pure and mixed pixels of the experimental plots in exp2; e is the error.</p>
Full article ">Figure 11
<p>Distribution of the simulated yield of the winter wheat with a1 km resolution in Hengshui City in 2017. Simulated yield based on the (<b>a</b>) GPP1 model, (<b>b</b>) GPP2 model, and (<b>c</b>) ACPM model.</p>
Full article ">Figure 11 Cont.
<p>Distribution of the simulated yield of the winter wheat with a1 km resolution in Hengshui City in 2017. Simulated yield based on the (<b>a</b>) GPP1 model, (<b>b</b>) GPP2 model, and (<b>c</b>) ACPM model.</p>
Full article ">Figure 12
<p>Validation of the estimated winter wheat yield based on the GPP1, GPP2, and ACPM models at the harvest in Hengshui City in 2017, as follows: (<b>a</b>) GPP1 model in the pure pixels of the experimental plots in exp2; (<b>b</b>) GPP1 model in the pure and mixed pixels of the experimental plots in exp2; (<b>c</b>) GPP2 model in the pure pixels of the experimental plots in exp2; (<b>d</b>) GPP2 model in the pure and mixed pixels of the experimental plots in exp2; (<b>e</b>) ACPM model in the pure pixels of the experimental plots in exp2; and (<b>f</b>) ACPM model in the pure and mixed pixels of the experimental plots in exp2; e is the error.</p>
Full article ">
17 pages, 5763 KiB  
Article
Temporal and Spatial Characteristics of EVI and Its Response to Climatic Factors in Recent 16 years Based on Grey Relational Analysis in Inner Mongolia Autonomous Region, China
by Dong He, Guihua Yi, Tingbin Zhang, Jiaqing Miao, Jingji Li and Xiaojuan Bie
Remote Sens. 2018, 10(6), 961; https://doi.org/10.3390/rs10060961 - 15 Jun 2018
Cited by 37 | Viewed by 5633
Abstract
The Inner Mongolia Autonomous Region (IMAR) is a major source of rivers, catchment areas, and ecological barriers in the northeast of China, related to the nation’s ecological security and improvement of the ecological environment. Therefore, studying the response of vegetation to climate change [...] Read more.
The Inner Mongolia Autonomous Region (IMAR) is a major source of rivers, catchment areas, and ecological barriers in the northeast of China, related to the nation’s ecological security and improvement of the ecological environment. Therefore, studying the response of vegetation to climate change has become an important part of current global change research. Since existing studies lack detailed descriptions of the response of vegetation to different climatic factors using the method of grey correlation analysis based on pixel, the temporal and spatial patterns and trends of enhanced vegetation index (EVI) are analyzed in the growing season in IMAR from 2000 to 2015 based on moderate resolution imaging spectroradiometer (MODIS) EVI data. Combined with the data of air temperature, relative humidity, and precipitation in the study area, the grey relational analysis (GRA) method is used to study the time lag of EVI to climate change, and the study area is finally zoned into different parts according to the driving climatic factors for EVI on the basis of lag analysis. The driving zones quantitatively show the characteristics of temporal and spatial differences in response to different climatic factors for EVI. The results show that: (1) The value of EVI generally features in spatial distribution, increasing from the west to the east and the south to the north. The rate of change is 0.22/10°E from the west to the east, 0.28/10°N from the south to the north; (2) During 2000–2015, the EVI in IMAR showed a slightly upward trend with a growth rate of 0.021/10a. Among them, the areas with slight and significant improvement accounted for 21.1% and 7.5% of the total area respectively, ones with slight and significant degradation being 24.6% and 4.3%; (3) The time lag analysis of climatic factors for EVI indicates that vegetation growth in the study area lags behind air temperature by 1–2 months, relative humidity by 1–2 months, and precipitation by one month respectively; (4) During the growing season, the EVI of precipitation driving zone (21.8%) in IMAR is much larger than that in the air temperature driving zone (8%) and the relative humidity driving zone (11.6%). The growth of vegetation in IMAR generally has the closest relationship with precipitation. The growth of vegetation does not depend on the change of a single climatic factor. Instead, it is the result of the combined action of multiple climatic factors and human activities. Full article
(This article belongs to the Special Issue Earth Observations for Addressing Global Challenges)
Show Figures

Figure 1

Figure 1
<p>The Digital Elevation Model (DEM) and meteorological station distribution (<b>a</b>) and the vegetation types (<b>b</b>) in IMAR. (A—Hulunbuir; B—Hinggan League; C—Xilingol League; D—Chifeng; E—Tongliao; F—Ulanqab League; G—Baotou; H—Hohhot; I—Bayannur League; J—Erdos; K—Wuhai; L—Alxa League).</p>
Full article ">Figure 2
<p>The EVI time series data before (<b>a</b>) and after (<b>b</b>) HANTS processing.</p>
Full article ">Figure 3
<p>Inter-annual change of EVI in the study area from 2000 to 2015.</p>
Full article ">Figure 4
<p>Spatial distribution of multi-year mean of EVI in the growing season in the study area from 2000 to 2015. (<b>a</b>) Longitude and altitude statistics; (<b>b</b>) Latitude and altitude statistics.</p>
Full article ">Figure 5
<p>EVI trend (<b>a</b>) and EVI change (<b>b</b>) in the study area from 2000 to 2015.</p>
Full article ">Figure 6
<p>The GRGs of growing season EVI to air temperature, relative humidity and precipitation in February–June, March–July, April–August, and May–September.</p>
Full article ">Figure 7
<p>Inter-annual change of air temperature (<b>a</b>); relative humidity (<b>b</b>); and precipitation (<b>c</b>) in IMAR from 2000 to 2015.</p>
Full article ">Figure 8
<p>GRG between EVI and air temperature (<b>a</b>); relative humidity (<b>b</b>); and precipitation (<b>c</b>).</p>
Full article ">Figure 9
<p>Areas of precipitation drive, air temperature drive, and relative humidity drive in the study area.</p>
Full article ">
13 pages, 2269 KiB  
Article
Variability of Microwave Scattering in a Stochastic Ensemble of Measured Rain Drops
by Francisco J. Tapiador, Raúl Moreno, Andrés Navarro, Alfonso Jiménez, Enrique Arias and Diego Cazorla
Remote Sens. 2018, 10(6), 960; https://doi.org/10.3390/rs10060960 - 15 Jun 2018
Cited by 2 | Viewed by 4059
Abstract
While it has been proved that multiple scattering in the microwave frequencies has to be accounted for in precipitation retrieval algorithms, the effects of the random arrangements of drops in space has seldom been investigated. The fact is, a single rain drop size [...] Read more.
While it has been proved that multiple scattering in the microwave frequencies has to be accounted for in precipitation retrieval algorithms, the effects of the random arrangements of drops in space has seldom been investigated. The fact is, a single rain drop size distribution (RDSD) corresponds with many actual 3D distributions of those rain drops and each of those may a priori absorb and scatter radiation in a different way. Each spatial configuration is equivalent to any other in terms of the RDSD function, but not in terms of radiometric characteristics, both near and far from field, because of changes in the relative phases among the particles. Here, using the T-matrix formalism, we investigate the radiometric variability of two ensembles of 50 different 3D, stochastically-derived configurations from two consecutive measured RDSDs with 30 and 31 drops, respectively. The results show that the random distribution of drops in space has a measurable but apparently small effect in the scattering calculations with the exception of the asymmetry factor. Full article
(This article belongs to the Special Issue Remote Sensing of Precipitation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Disdrometers used to measure the Rain Drop Size Distribution (RDSD). (<b>Left</b>) Dual, orthogonal setup or autonomous disdrometers used to calculate the spatial variability of rain at hectometre resolution [<a href="#B37-remotesensing-10-00960" class="html-bibr">37</a>]. (<b>Right</b>) Calibration array of 18 disdrometers used to measure the small-scale variability of the RDSD [<a href="#B38-remotesensing-10-00960" class="html-bibr">38</a>].</p>
Full article ">Figure 2
<p>Measured RDSDs (<b>A</b>,<b>B</b>) used to build the two cases explored in this paper. Diameters are in mm. The fit is for a three-parameter gamma.</p>
Full article ">Figure 3
<p>A sample of the different randomly-generated spatial distributions of raindrops consistent with one of the measured RDSDs of <a href="#remotesensing-10-00960-f002" class="html-fig">Figure 2</a>. The diameters of the raindrops are exaggerated. Dimensions are in centimeters.</p>
Full article ">Figure 4
<p>Variability of the total extinction for an ensemble of 50 3D configurations of the (<b>A</b>,<b>B</b>) RDSDs in <a href="#remotesensing-10-00960-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>Variability of the total absorption for an ensemble of 50 3D configurations of the (<b>A</b>,<b>B</b>) RDSDs in <a href="#remotesensing-10-00960-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 6
<p>Variability of the scattering efficiency for an ensemble of 50 3D configurations of the (<b>A</b>,<b>B</b>) RDSDs in <a href="#remotesensing-10-00960-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 7
<p>Variability of the asymmetry parameter for an ensemble of 50 3D configurations of the (<b>A</b>,<b>B</b>) RDSDs in <a href="#remotesensing-10-00960-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 8
<p>Spread of the dependence of the phase functions, <span class="html-italic">S</span><sub>11</sub> and <span class="html-italic">S</span><sub>21</sub>, on the scattering angle for all the member of the ensemble.</p>
Full article ">
21 pages, 19221 KiB  
Article
The Temperature Vegetation Dryness Index (TVDI) Based on Bi-Parabolic NDVI-Ts Space and Gradient-Based Structural Similarity (GSSIM) for Long-Term Drought Assessment Across Shaanxi Province, China (2000–2016)
by Ying Liu and Hui Yue
Remote Sens. 2018, 10(6), 959; https://doi.org/10.3390/rs10060959 - 15 Jun 2018
Cited by 43 | Viewed by 7586
Abstract
Traditional NDVI-Ts space is triangular or trapezoidal, but Liu et al. (2015) discovered that the NDVI-Ts space was bi-parabolic when the study area was covered with low biomass vegetation. Moreover, the numerical value of the indicator was considered in most of [...] Read more.
Traditional NDVI-Ts space is triangular or trapezoidal, but Liu et al. (2015) discovered that the NDVI-Ts space was bi-parabolic when the study area was covered with low biomass vegetation. Moreover, the numerical value of the indicator was considered in most of the study when the drought conditions in the space domain were evaluated. In addition, quantitatively assessing the spatial-temporal changes of the drought was not enough. In this study, first, we used MODIS NDVI and Ts data to reexamine if the NDVI-Ts space with “time” and a single pixel domain is bi-parabolic in the Shaanxi province of China, which is vegetated with low biomass to high biomass. This is compared with the triangular NDVI-Ts space and one of the well-known drought indexes called the temperature-vegetation index (TVX). The results demonstrated that dry and wet edges exhibited a parabolic shape again in scatter plots of Ts and NDVI in the Shaanxi province, which was linear in the triangular NDVI-Ts space. The Temperature Vegetation Dryness Index (TVDIc) was obtained from bi-parabolic NDVI-Ts andTVDIt was obtained from the triangular NDVI-Ts space and TVX were compared with 10-cm depth relative soil moisture. By estimating the 10-cm depth soil moisture, TVDIc was better than TVDIt, which were all apparently better than TVX. Second, combined with MODIS data, the drought conditions of the study area were assessed by TVDIc between 2000 to 2016. Spatially, the drought in the Shaanxi Province between 2000 to 2016 were mainly distributed in the northwest, North Shaanxi, and the North and East Guanzhong plain. The drought area of the Shaanxi province accounted for 31.95% in 2000 and 27.65% in 2016, respectively. Third, we quantitatively evaluated the variation of the drought status by using Gradient-based Structural Similarity (GSSIM) methods. The area of the drought conditions significantly changed and moderately changed at 5.34% and 40.22%, respectively, between 2000 and 2016. Finally, the possible reasons for drought change were discussed. The change of precipitation, temperature, irrigation, destruction or betterment of vegetation, and the enlargement of opening mining, etc., can lead to the variations of drought. Full article
(This article belongs to the Section Biogeosciences Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The study area. (<b>a</b>) Location of Shaanxi province in China and (<b>b</b>) the land cover types of Shaanxi province (<a href="https://search.earthdata.nasa.gov/" target="_blank">https://search.earthdata.nasa.gov/</a>).</p>
Full article ">Figure 2
<p>The flow chart of constructing bi-parabolic NDVI-T<sub>s</sub> space.</p>
Full article ">Figure 3
<p>Sketch map of TVDI in bi-parabolic NDVI-T<sub>s</sub> space.</p>
Full article ">Figure 4
<p>Scatter plots of dry and wet edges in NDVI-T<sub>s</sub> space between 2000 to 2016. (<b>a</b>) 2000, (<b>b</b>) 2001, (<b>c</b>) 2002, (<b>d</b>) 2003, (<b>e</b>) 2004, (<b>f</b>) 2005, (<b>g</b>) 2006, (<b>h</b>) 2007, (<b>i</b>) 2008, (<b>j</b>) 2009, (<b>k</b>) 2010, (<b>l</b>) 2011, (<b>m</b>) 2012, (<b>n</b>) 2013, (<b>o</b>) 2014, (<b>p</b>) 2015, (<b>q</b>) 2016.</p>
Full article ">Figure 5
<p>Spatiotemporal variation of drought in Shaanxi province between 2000 to 2016. (<b>a</b>) 2000, (<b>b</b>) 2001, (<b>c</b>) 2002, (<b>d</b>) 2003, (<b>e</b>) 2004, (<b>f</b>) 2005, (<b>g</b>) 2006, (<b>h</b>) 2007, (<b>i</b>) 2008, (<b>j</b>) 2009, (<b>k</b>) 2010, (<b>l</b>) 2011, (<b>m</b>) 2012, (<b>n</b>) 2013, (<b>o</b>) 2014, (<b>p</b>) 2015, (<b>q</b>) 2016.</p>
Full article ">Figure 6
<p>Spatio-temperal distribution of GSSIM in Shaanxi province. (<b>a</b>) 2000–2005; (<b>b</b>) 2005–2011; (<b>c</b>) 2011–2016; (<b>d</b>) 2000–2016.</p>
Full article ">Figure 7
<p>The location of samples. (<b>a</b>) GSSIM between 2008 and 2015, (<b>b</b>) Samples of D, E, F, and G, (<b>c</b>) Samples of A, B, and C.</p>
Full article ">Figure 8
<p>Samples of mutation (A, B, C, D) and moderate variation (E, F, G). (<b>a</b>,<b>b</b>,<b>e</b>,<b>f</b>) drought was eased, (<b>c</b>,<b>d</b>,<b>g</b>) drought was aggravated.</p>
Full article ">Figure 9
<p>Comparison of TVDI<sub>c</sub>, TVDI<sub>t</sub>, and TVX with 10 cm soil moisture. (<b>a</b>) TVDI<sub>c</sub> against 10 cm soil moisture during 7 April 2013–14 April 2013, (<b>b</b>) TVDI<sub>t</sub> against 10 cm soil moisture during 7 April 2013–14 April 2013, (<b>c</b>) TVX against 10 cm soil moisture during 7 April 2013–14 April 2013, (<b>d</b>) TVDI<sub>c</sub> against 10 cm soil moisture during 15 April 2013–22 April 2013, (<b>e</b>) TVDI<sub>t</sub> against 10 cm soil moisture during 15 April 2013–22 April 2013, (<b>f</b>) TVX against 10 cm soil moisture during 15 April 2013–22 April 2013, (<b>g</b>) TVDI<sub>c</sub> against 10 cm soil moisture during 23 April 2013–30 April 2013, (<b>h</b>) TVDI<sub>t</sub> against 10 cm soil moisture during 23 April 2013–30 April 2013, (<b>i</b>) TVX against 10 cm soil moisture during 23 April 2013–30 April 2013.</p>
Full article ">Figure 10
<p>Comparison of scatter plots of bi-parabolic and triangular NDVI-T<sub>s</sub> space. (<b>a</b>) Scatter plots of bi-parabolic NDVI-T<sub>s</sub> space during 7 April 2013–14 April 2013, (<b>b</b>) scatter plots of triangular NDVI-T<sub>s</sub> space during 7 April 2013–14 April 2013, (<b>c</b>) scatter plots of bi-parabolic NDVI-T<sub>s</sub> space during 15 April 2013–22 April 2013, (<b>d</b>) acatter plots of triangular NDVI-T<sub>s</sub> space during 15 April 2013–22 April 2013, (<b>e</b>) scatter plots of bi-parabolic NDVI-T<sub>s</sub> space during 23 April 2013–30 April 2013, and (<b>f</b>) scatter plots of triangular NDVI-T<sub>s</sub> space during 23 April 2013–30 April 2013.</p>
Full article ">Figure 11
<p>Comparison of drought conditions from TVDI<sub>c</sub>, TVDI<sub>t</sub>, and TVX, respectively. (<b>a</b>) Drought conditions from TVDI<sub>c</sub> between 7 April 2013–14 April 2013. (<b>b</b>) Drought conditions from TVDI<sub>t</sub> between 7 April 2013–14 April 2013. (<b>c</b>) Drought conditions from TVX between 7 April 2013–14 April 2013. (<b>d</b>) Drought conditions from TVDI<sub>c</sub> between 15 April 2013–22 April 2013. (<b>e</b>) Drought conditions from TVDI<sub>t</sub> between 15 April 2013–22 April 2013. (<b>f</b>) Drought conditions from TVX between 15 April 2013–22 April 2013. (<b>g</b>) TVDI<sub>c</sub> between 23 April, 2013–30 April 2013. (<b>h</b>) Drought conditions from TVDI<sub>t</sub> between 23 April 2013–30 April 2013. (<b>i</b>) Drought conditions from TVX between 23 April 2013–30 April 2013.</p>
Full article ">Figure 12
<p>Comparison of GSSIM and linear regression analysis. (<b>a</b>) GSSIM between 2000 and 2016. (<b>b</b>) Significant linear regression slope values for trends derived from TVDI observations during 2000–2016 (reliable at 5% significance level).</p>
Full article ">Figure 13
<p>The correlation coefficient classification map between TVDI and meteorological factors. (<b>a</b>) Precipitation. (<b>b</b>) Annual average temperature. (<b>c</b>) Temperature anomaly.</p>
Full article ">
19 pages, 7228 KiB  
Article
Influences of Environmental Loading Corrections on the Nonlinear Variations and Velocity Uncertainties for the Reprocessed Global Positioning System Height Time Series of the Crustal Movement Observation Network of China
by Peng Yuan, Zhao Li, Weiping Jiang, Yifang Ma, Wu Chen and Nico Sneeuw
Remote Sens. 2018, 10(6), 958; https://doi.org/10.3390/rs10060958 - 15 Jun 2018
Cited by 22 | Viewed by 4881
Abstract
Mass redistribution of the atmosphere, oceans, and terrestrial water storage generates crustal displacements which can be predicted by environmental loading models and observed by the Global Positioning System (GPS). In this paper, daily height time series of 235 GPS stations derived from a [...] Read more.
Mass redistribution of the atmosphere, oceans, and terrestrial water storage generates crustal displacements which can be predicted by environmental loading models and observed by the Global Positioning System (GPS). In this paper, daily height time series of 235 GPS stations derived from a homogeneously reprocessed Crustal Movement Observation Network of China (CMONOC) and corresponding loading displacements predicted by the Deutsche GeoForschungsZentrum (GFZ) are compared to assess the effects of loading corrections on the nonlinear variations of GPS time series. Results show that the average root mean square (RMS) of vertical displacements due to atmospheric, nontidal oceanic, hydrological, and their combined effects are 3.2, 0.6, 2.7, and 4.0 mm, respectively. Vertical annual signals of loading and GPS are consistent in amplitude but different in phase systematically. The average correlation coefficient between loading and GPS height time series is 0.6. RMS of the GPS height time series are reduced by 20% on average. Moreover, an investigation of 208 CMONOC stations with observing time spans of ~4.6 years shows that environmental loading corrections lead to an overestimation of the GPS velocity uncertainty by about 1.4 times on average. Nevertheless, by using a common mode component filter through principal component analysis, the dilution of velocity precision due to environmental loading corrections can be compensated. Full article
(This article belongs to the Special Issue Remote Sensing of Tectonic Deformation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Geographical distribution and time spans of the selected 235 CMONOC stations used in this paper. The CMONOC-I (<b>square</b>) and CMONOC-II (<b>circle</b>), two subnetworks of the CMONOC, are composed of 27 and 208 stations, respectively. The blue triangles indicate 34 coastal stations located within 100 km distance of coast. The green inverted triangles indicate six selected representative stations.</p>
Full article ">Figure 2
<p>RMS of the vertical displacements due to ATML, NTOL, HYDL, and their sum (SumL), predicted with GFZ loading products for the CMONOC stations.</p>
Full article ">Figure 3
<p>Vertical displacement time series due to ATML (<b>red</b>), NTOL (<b>green</b>), HYDL (<b>blue</b>), and SumL (<b>orange</b>) predicted with GFZ loading products and GPS height time series (<b>black</b>) for the selected six CMONOC stations. The ATML, NTOL, and HYDL time series are shifted upward by 60, 45, and 30 mm, respectively.</p>
Full article ">Figure 4
<p>Vertical annual amplitudes and phases of the ATML (<b>a</b>,<b>b</b>), NTOL (<b>c</b>,<b>d</b>), HYDL (<b>e</b>,<b>f</b>), and SumL (<b>g</b>,<b>h</b>) time series predicted with GFZ loading products for the CMONOC stations. The annual phase is the day of year when an annual signal reaches its maximum.</p>
Full article ">Figure 5
<p>Vertical annual amplitudes and phases of the <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> (<b>a</b>,<b>b</b>) and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>,<b>d</b>) time series for the CMONOC stations. The annual phase is the day of year when an annual signal reaches its maximum.</p>
Full article ">Figure 6
<p>Comparison of vertical annual amplitudes between the <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> and SumL time series. The gray dashed line indicates perfect match. The black line indicates the best fit line estimated with a weighted total least squares algorithm.</p>
Full article ">Figure 7
<p>Phase differences between the vertical annual signals of SumL and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> time series (<b>left panel</b>) and its histogram (<b>right panel</b>).</p>
Full article ">Figure 8
<p>Annual amplitude reduction ratio in percentage for the height time series of the CMONOC stations with loading corrections from GFZ (<b>left panel</b>) and its histogram (<b>right panel</b>).</p>
Full article ">Figure 9
<p>Correlation coefficients between <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> and SumL time series for the vertical component of the CMONOC stations (<b>left panel</b>) and its histogram (<b>right panel</b>).</p>
Full article ">Figure 10
<p>RMS reduction ratio in percentage for the vertical component of the CMONOC stations with loading corrections from GFZ (<b>left panel</b>) and its histogram (<b>right panel</b>).</p>
Full article ">Figure 11
<p>Stacked power spectra of the vertical displacements due to ATML, NTOL, HYDL, and SumL for the selected 208 CMONOC-II stations. The stacked power spectrum of SumL has been shifted upward for clarity (by a factor of 100). Vertical gray lines indicate the annual and semiannual oscillations.</p>
Full article ">Figure 12
<p>Stacked power spectra of the solutions of <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math>, CMC-filtered <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math>, and CMC-filtered <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math> time series for the selected 208 CMONOC-II stations. The stacked power spectra for the <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math> time series have been shifted upward for clarity (by a factor of 10). Vertical gray lines indicate the annual and semiannual oscillations.</p>
Full article ">Figure 13
<p>Velocity differences and dilution of velocity precision for the <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math> (<b>a</b>,<b>b</b>), CMC-filtered <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>,<b>d</b>), and CMC-filtered <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math> (<b>e</b>,<b>f</b>) time series with respect to the <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> </mrow> </semantics></math> time series for the CMONOC-II stations.. All the velocities are estimated with PL + WN noise model.</p>
Full article ">Figure 14
<p>Interstation correlation coefficients for the residual <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="normal">H</mi> <mrow> <mi>before</mi> </mrow> </msub> <msub> <mrow> <mo> </mo> <mi>and</mi> <mo> </mo> <mi mathvariant="normal">H</mi> </mrow> <mrow> <mi>after</mi> </mrow> </msub> </mrow> </semantics></math> time series as a function of interstation distances in kilometers (<b>left panels</b>) and the corresponding histograms (<b>right panels</b>). The interstation correlation coefficients of the unfiltered and CMC-filtered time series are shown as blue and orange dot marks, respectively. Moreover, they are smoothed by a boxcar smoother with a full width of 50 km and shown as the blue and red lines, respectively.</p>
Full article ">
20 pages, 2079 KiB  
Article
Aerosol and Meteorological Parameters Associated with the Intense Dust Event of 15 April 2015 over Beijing, China
by Sheng Zheng and Ramesh P. Singh
Remote Sens. 2018, 10(6), 957; https://doi.org/10.3390/rs10060957 - 15 Jun 2018
Cited by 18 | Viewed by 4449
Abstract
The northeastern parts of China, including Beijing city, the capital of China, were hit by an intense dust storm on 15 April 2015. The present paper discusses aerosol and meteorological parameters associated with this dust storm event. The back trajectory clearly shows that [...] Read more.
The northeastern parts of China, including Beijing city, the capital of China, were hit by an intense dust storm on 15 April 2015. The present paper discusses aerosol and meteorological parameters associated with this dust storm event. The back trajectory clearly shows that the dust originated from Inner Mongolia, the border of China, and Mongolia regions. Pronounced changes in aerosol and meteorological parameters along the dust track were observed. High aerosol optical depth (AOD) with low Ångström exponent (AE) are characteristics of coarse-mode dominated dust particles in the wavelength range 440–870 nm during the dusty day. During dust storm, dominance of coarse aerosol concentrations is observed in the aerosol size distribution (ASD). The single scattering albedo (SSA) retrieved from AERONET station shows increase with higher wavelength on the dusty day, and is found to be higher compared to the days prior to and after the dust event, supported with high values of the real part and decrease in the imaginary part of the refractive index (RI). With regard to meteorological parameters, during the dusty day, CO volume mixing ratio (COVMR) is observed to decrease, from the surface up to mid-altitude, compared with the non-dusty days due to strong winds. O3 volume mixing ratio (O3VMR) enhances at the increasing altitudes (at the low-pressure levels), and decreases near the surface at the pressure levels 500–925 hPa during the dust event, compared with the non-dusty periods. An increase in the H2O mass mixing ratio (H2OMMR) is observed during dusty periods at the higher altitudes equivalent to the pressure levels 500 and 700 hPa. The mid-altitude relative humidity (RH) is observed to decrease at the pressure levels 700 and 925 hPa during sand storm days. With the onset of the dust storm event, the RH reduces at the surface level. Full article
(This article belongs to the Special Issue Remote Sensing of Atmospheric Properties)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Hourly variations of PM<sub>2.5</sub>, PM<sub>10</sub>, CO, and O<sub>3</sub> concentrations in Dongsi station, Beijing during the dust storm event.</p>
Full article ">Figure 2
<p>HYSPLIT 2 days back trajectory of air masses (dots) at three heights arriving at Beijing at 17:00 h on 15 April 2015 (local standard time), the yellow line represents the portion of CALIPSO orbit track on 14 April 2015 (UTC), the orange and pink box, respectively, represents the dust track passing through the city of Ulanqab and Zhangjiakou.</p>
Full article ">Figure 3
<p>MODIS, AOD, and AE in the two boxes located in the city of Ulanqab and Zhangjiakou.</p>
Full article ">Figure 4
<p>Hourly variations of PM<sub>2.5</sub>, PM<sub>10</sub>, and CO concentrations in the city of Ulanqab (<b>a</b>) and Zhangjiakou (<b>b</b>) on 15 April 2015 (local standard time).</p>
Full article ">Figure 5
<p>Total attenuated backscatter and depolarization ratio at 532 nm wavelength in the range of (20.00°N, 94.69°E)~(68.00°N, 115.88°E) on 14 April 2015 (UTC). The horizontal coordinate represents the latitude and longitude, the vertical coordinate represents the altitude. The letter “D” designates the dust layer. (<b>a</b>) Total attenuated backscatter, and the blue to white color bar represents the value of the total backscatter at 532 nm wavelength; (<b>b</b>) Depolarization ratio and the black to white color bar represent the value of the depolarization ratio at 532 nm wavelength.</p>
Full article ">Figure 6
<p>Diurnal (<b>a</b>) and daily (<b>b</b>) variations of AERONET AOD675 and Ångström exponent during dusty and non-dusty days. The red column represents AOD675, and the blue dot refers to Ångström exponent.</p>
Full article ">Figure 6 Cont.
<p>Diurnal (<b>a</b>) and daily (<b>b</b>) variations of AERONET AOD675 and Ångström exponent during dusty and non-dusty days. The red column represents AOD675, and the blue dot refers to Ångström exponent.</p>
Full article ">Figure 7
<p>Variation of aerosol size distributions during dusty and non-dusty days.</p>
Full article ">Figure 8
<p>Variations of total mode single scattering albedo during dust storm event.</p>
Full article ">Figure 9
<p>Variations of real (<b>a</b>) and imaginary (<b>b</b>) parts of refractive index during dusty and non-dusty days.</p>
Full article ">Figure 10
<p>CO volume mixing ratio (COVMR) during dusty and non-dusty days. The red line represents profile during the dusty day.</p>
Full article ">Figure 11
<p>O<sub>3</sub> volume mixing ratio on dusty day and non-dusty days. The red line represents vertical profile on dusty day.</p>
Full article ">Figure 12
<p>Relative humidity during dusty and non-dusty days. The red line represents profile during the dusty day.</p>
Full article ">Figure 13
<p>H<sub>2</sub>O mass mixing ratio during dusty and non-dusty days. The red line represents profile during the dusty day.</p>
Full article ">
12 pages, 3945 KiB  
Article
Temporal Variability of MODIS Phenological Indices in the Temperate Rainforest of Northern Patagonia
by Carlos Lara, Gonzalo S. Saldías, Alvaro L. Paredes, Bernard Cazelles and Bernardo R. Broitman
Remote Sens. 2018, 10(6), 956; https://doi.org/10.3390/rs10060956 - 15 Jun 2018
Cited by 18 | Viewed by 4692
Abstract
Western Patagonia harbors unique and sparsely studied terrestrial ecosystems that are threatened by land use changes and exposure to basin-scale climatic variability. We assessed the performance of two satellite vegetation indices derived from MODIS–Terra, EVI (Enhanced Vegetation Index) and NDVI (Normalized Difference Vegetation [...] Read more.
Western Patagonia harbors unique and sparsely studied terrestrial ecosystems that are threatened by land use changes and exposure to basin-scale climatic variability. We assessed the performance of two satellite vegetation indices derived from MODIS–Terra, EVI (Enhanced Vegetation Index) and NDVI (Normalized Difference Vegetation Index), over the northern and southern sectors of the Chiloé Island System (CIS) to advance our understanding of vegetation dynamics in the region. Then we examined their time-varying relationships with two climatic indices indicative of tropical and extratropical influence, the ENSO (El Niño–Southern Oscillation) and the Antarctic Oscillation (AAO) index, respectively. The 17-year time series showed that only EVI captured the seasonal pattern characteristic of temperate regions, with low (high) phenological activity during Autumn-Winter (Spring–Summer). NDVI saturated during the season of high productivity and failed to capture the seasonal cycle. Temporal patterns in productivity showed a weakened seasonal cycle during the past decade, particularly over the northern sector. We observed a non-stationary association between EVI and both climatic indices. Significant co-variation between EVI and the Niño–Southern Oscillation index in the annual band persisted from 2001 until 2008–2009; annual coherence with AAO prevailed from 2013 onwards and the 2009–2012 period was characterized by coherence between EVI and both climate indices over longer temporal scales. Our results suggest that the influence of large-scale climatic variability on local weather patterns drives phenological responses in the northern and southern regions of the CIS. The imprint of climatic variability on patterns of primary production across the CIS may be underpinned by spatial differences in the anthropogenic modification of this ecosystem, as the northern sector is strongly modified by forestry and agriculture. We highlight the need for field validation of satellite indices around areas of high biomass and high endemism, located in the southern sector of the island, in order to enhance the utility of satellite vegetation indices in the conservation and management of austral ecosystems. Full article
(This article belongs to the Special Issue Multitemporal Remote Sensing for Forestry)
Show Figures

Figure 1

Figure 1
<p>Study area: Chiloé Island System (CIS) coloured using the average EVI for all images over the 2001–2016 study period. Yellow and red polygons delimit the northern (yellow) and southern (red) regions described in our study. The color bar shows EVI values.</p>
Full article ">Figure 2
<p>Time series of (<b>a</b>) NDVI for northern (solid red line) and southern (solid blue line) region; and (<b>b</b>) EVI for northern (solid red line) and southern (solid blue line) region region of CIS; (<b>c</b>) Climatic indices used in this study: an ENSO index represented as the monthly time series of El Niño 3.4 (positive/negative, red/blue) and the monthly series of the Antarctic Oscillation (AAO; black line).</p>
Full article ">Figure 3
<p>Wavelet Power Spectrum for (<b>a</b>) Enhanced Vegetation Index (EVI) in northern region; (<b>b</b>) Normalized Difference Vegetation Index (NDVI) in northern region; (<b>c</b>) Enhanced Vegetation Index (EVI) in southern region and (<b>d</b>) Normalized Difference Vegetation Index (NDVI) in southern region. The left panels show the local wavelet power spectrum with increasing spectrum intensity from white to dark red; black dashed lines show the 95% statistically significant areas computed with adapted bootstrapping (see [<a href="#B34-remotesensing-10-00956" class="html-bibr">34</a>]); the dotted lines localize the maxima of spectra in time and period; the black curve indicates the cone of influence (region not influenced by edge effects). The right panels show the global power spectrum (blue line) with its threshold value of 95% CI (black line) (see [<a href="#B34-remotesensing-10-00956" class="html-bibr">34</a>]).</p>
Full article ">Figure 4
<p>Wavelet coherence between (<b>a</b>) EVI and ENSO in northern region; (<b>b</b>) EVI and AAO in northern region; (<b>c</b>) EVI and ENSO in southern region and (<b>d</b>) EVI and AAO in southern region. The colors coded for low coherence in white to high coherence in dark red. The blue dashed lines indicate the 95% and 90% statistically significant areas computed with adapted bootstrapping (see [<a href="#B34-remotesensing-10-00956" class="html-bibr">34</a>]). The cone of influence (black curve) indicates the region not influenced by edge effects.</p>
Full article ">
16 pages, 1956 KiB  
Article
A Novel Approach for the Short-Term Forecast of the Effective Cloud Albedo
by Isabel Urbich, Jörg Bendix and Richard Müller
Remote Sens. 2018, 10(6), 955; https://doi.org/10.3390/rs10060955 - 15 Jun 2018
Cited by 22 | Viewed by 10041
Abstract
The increasing use of renewable energies as a source of electricity has led to a fundamental transition of the power supply system. The integration of fluctuating weather-dependent energy sources into the grid already has a major impact on its load flows. As a [...] Read more.
The increasing use of renewable energies as a source of electricity has led to a fundamental transition of the power supply system. The integration of fluctuating weather-dependent energy sources into the grid already has a major impact on its load flows. As a result, the interest in forecasting wind and solar radiation with a sufficient accuracy over short time periods (<4 h) has grown. In this study, the short-term forecast of the effective cloud albedo based on optical flow estimation methods is investigated. The optical flow method utilized here is TV-L1 from the open source library OpenCV. This method uses a multi-scale approach to capture cloud motions on various spatial scales. After the clouds are displaced, the solar surface radiation will be calculated with SPECMAGIC NOW, which computes the global irradiation spectrally resolved from satellite imagery. Due to the high temporal and spatial resolution of satellite measurements, the effective cloud albedo and thus solar radiation can be forecasted from 5 min up to 4 h with a resolution of 0.05°. The validation results of this method are very promising, and the RMSE of the 30-min, 60-min, 90-min and 120-min forecast equals 10.47%, 14.28%, 16.87% and 18.83%, respectively. The paper gives a brief description of the method for the short-term forecast of the effective cloud albedo. Subsequently, evaluation results will be presented and discussed. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of the optical flow method. Two images of the effective cloud albedo serve as input for the TV-<math display="inline"><semantics> <msup> <mi>L</mi> <mn>1</mn> </msup> </semantics></math> method. The estimated cloud motion vectors are then applied to the latter of the consecutive images to extrapolate the cloud motion into the future.</p>
Full article ">Figure 2
<p>Example of the clear sky reflection <math display="inline"><semantics> <msub> <mi>ρ</mi> <mi>cs</mi> </msub> </semantics></math> (left) and effective cloud albedo CAL (right) for an 11 UTC slot in June 2005.</p>
Full article ">Figure 3
<p>Plots of the verification of the optical flow estimate with the method by Farnebäck (left) and the TV-<math display="inline"><semantics> <msup> <mi>L</mi> <mn>1</mn> </msup> </semantics></math> method (right) for a 15-min forecast. The area of Germany is marked by the red frame.</p>
Full article ">Figure 4
<p>Plots of the absolute bias of the effective cloud albedo against the forecast time for the cases of 30 September 2017 at 13:00 UTC (situation with convection behind a front over Germany) and 4 October 2017 at 12:00 UTC (stratiform situation). The unit of the absolute bias is %.</p>
Full article ">Figure 5
<p>A short-term forecast for 120 min of the effective cloud albedo for 30 September 2017 at 15:00 UTC can be seen in the left figure. The satellite image by MSG with the effective cloud albedo depicted for comparison is shown in the right image.</p>
Full article ">Figure 6
<p>A short-term forecast for 120 min of the effective cloud albedo for 4 October 2017 at 14:00 UTC can be seen in the left figure. The satellite image by MSG with the effective cloud albedo depicted for comparison is shown in the right image.</p>
Full article ">Figure 7
<p>Verification plots of the optical flow method for the cases of 30 September 2017 at 15:00 UTC and 4 October 2017 at 14:00 UTC. The absolute difference between the effective cloud albedo from satellite imagery and the effective cloud albedo from the short-term forecast for 120 min in % is shown.</p>
Full article ">
24 pages, 12146 KiB  
Article
Climate Sensitivity of High Arctic Permafrost Terrain Demonstrated by Widespread Ice-Wedge Thermokarst on Banks Island
by Robert H. Fraser, Steven V. Kokelj, Trevor C. Lantz, Morgan McFarlane-Winchester, Ian Olthof and Denis Lacelle
Remote Sens. 2018, 10(6), 954; https://doi.org/10.3390/rs10060954 - 15 Jun 2018
Cited by 75 | Viewed by 11298
Abstract
Ice-wedge networks underlie polygonal terrain and comprise the most widespread form of massive ground ice in continuous permafrost. Here, we show that climate-driven thaw of hilltop ice-wedge networks is rapidly transforming uplands across Banks Island in the Canadian Arctic Archipelago. Change detection using [...] Read more.
Ice-wedge networks underlie polygonal terrain and comprise the most widespread form of massive ground ice in continuous permafrost. Here, we show that climate-driven thaw of hilltop ice-wedge networks is rapidly transforming uplands across Banks Island in the Canadian Arctic Archipelago. Change detection using high-resolution WorldView images and historical air photos, coupled with 32-year Landsat reflectance trends, indicate broad-scale increases in ponding from ice-wedge thaw on hilltops, which has significantly affected at least 1500 km2 of Banks Island and over 3.5% of the total upland area. Trajectories of change associated with this upland ice-wedge thermokarst include increased micro-relief, development of high-centred polygons, and, in areas of poor drainage, ponding and potential initiation of thaw lakes. Millennia of cooling climate have favoured ice-wedge growth, and an absence of ecosystem disturbance combined with surface denudation by solifluction has produced high Arctic uplands and slopes underlain by ice-wedge networks truncated at the permafrost table. The thin veneer of thermally-conductive mineral soils strongly links Arctic upland active-layer responses to summer warming. For these reasons, widespread and intense ice-wedge thermokarst on Arctic hilltops and slopes contrast more muted responses to warming reported in low and subarctic environments. Increasing field evidence of thermokarst highlights the inherent climate sensitivity of the Arctic permafrost terrain and the need for integrated approaches to monitor change and investigate the cascade of environmental consequences. Full article
(This article belongs to the Special Issue Remote Sensing of Dynamic Permafrost Regions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematics showing ice wedges in high Arctic ice-cored terrain and how these landscapes are (<b>A</b>) modified over millennial time periods by solifluction and erosional processes, contributing to (<b>B</b>) heightened sensitivity to climate-driven top-down thaw. Landsat image composites (shortwave infrared (SWIR), near-infrared (NIR), Red = RGB) from 16 August 1986 and 10 August 2013 for an area on eastern Banks Island show widespread, textured decreases in reflectance resulting from the formation of ice-wedge melt ponds. Photographs show upland terrains dissected by ice-wedge polygons and influenced by slow diffusive denudation processes (solifluction) and rapid thaw driven processes [ice-wedge degradation, high centre polygon and trough pond development (Ci–Ciii), and thaw slumping]. Stratigraphic sections show a thin veneer of materials over wedge ice and massive segregated ice (<a href="#app1-remotesensing-10-00954" class="html-app">Figures S2–S4</a>). Also, see data in <a href="#app1-remotesensing-10-00954" class="html-app">Table S2</a> for thicknesses of soils over massive ice on Banks Island.</p>
Full article ">Figure 2
<p>Sentinel-2 10 m resolution mosaic (RGB = NIR, Red, Green) of Banks Island created using images from 19 July 2017. The numbers represent the locations of the eight high-resolution study areas.</p>
Full article ">Figure 3
<p>(<b>a</b>) Annual thawing degree-days (1959–2016) computed using Environment Canada weather station data from Sachs Harbour, Banks Island, where the long-term mean is 461 °C (orange circles indicate years when thaw depths were measured at Green Cabin); (<b>b</b>) Relationship between average thaw depth and thawing degree-days at the time of measurement during early summer 2010–2015 at Green Cabin, north-central Banks Island (73°13′51″N, 119°32′18″W). The square root of degree-days is plotted according to the Stefan equation for predicting thaw depth. Best-fit linear regression equations and coefficients of determination (r<sup>2</sup>) are shown on both plots.</p>
Full article ">Figure 4
<p>Illustration of the method used to delineate upland landscapes containing new ice-wedge melt ponds. Sentinel-2 Normalised Difference Vegetation Index (NDVI) values (<b>a</b>) Less than 0.3 were used to classify sparsely vegetated upland landscapes for analysis and exclude higher biomass, lowland areas shown as dark green in (<b>b</b>); Landsat SWIR trends are shown in (<b>c</b>) and negative SWIR trends within uplands (dark blue pixels) in (<b>d</b>) were used to delineate vector polygons of melt pond landscapes outlined in yellow. Each of these melt pond landscapes was then verified through the presence of melt ponds visible in the 10 m 2017 Sentinel-2 imagery (<b>e</b>) and a visible, textured decrease in SWIR reflectance based on a pair of early (~1985) and late (~2017) Landsat images (<b>f</b>,<b>g</b>).</p>
Full article ">Figure 5
<p>Time series of Landsat images (1.6 µm SWIR channel top-of-atmosphere (TOA) reflectance) centred at 72.96°N, 119.13°W showing progressive, textured decreases in reflectance in upland terrain due to expanding ice-wedge melt ponds (<b>a</b>–<b>f</b>). The bottom panel shows the 30-year Landsat SWIR trends with significant negative trends indicated as dark blue (<b>g</b>) and a comparison of 1985 Landsat and 2017 Sentinel-2 imagery (<b>h</b>,<b>i</b>) (RGB = NIR, Red, Green). The yellow polygons were derived from the island-wide delineation of new upland melt-pond landscapes that was based on the Landsat trends and Sentinel-2 imagery.</p>
Full article ">Figure 6
<p>(<b>a</b>) Time series of Landsat SWIR reflectance for the upland terrain shown in <a href="#remotesensing-10-00954-f004" class="html-fig">Figure 4</a> averaged across pixels with significant (<span class="html-italic">p</span> &lt; 0.05) negative trends (blue), the strongest 20% of the negative trends (orange), and stable upland pixels with no significant trends (grey). Significant Theil–Sen slopes (<span class="html-italic">p</span> &lt; 0.05) computed based on observation year for the 20 images are shown as dashed lines. To normalise for the effect of surface moisture from precipitation on SWIR reflectance, the average difference of significantly trended and stable pixels is shown for each Landsat date in (<b>b</b>); average SWIR reflectance for lakes in lowland terrain (including a ~30 m buffer) is shown for each date in (<b>c</b>).</p>
Full article ">Figure 7
<p>Upland landscapes across Banks Island where ice-wedge melt ponds have expanded since 1985, mapped using a semi-automated approach based on negative Landsat SWIR trends and 2017 Sentinel-2 imagery. The delineated melt pond landscapes are aggregated to represent percent coverage at 2 km resolution, and at this scale cover 12,092 km<sup>2</sup> or 17.3% of Banks Island. The inset shows the extent of Amundsen (Jesse) glacial till delineated in [<a href="#B47-remotesensing-10-00954" class="html-bibr">47</a>,<a href="#B48-remotesensing-10-00954" class="html-bibr">48</a>].</p>
Full article ">Figure 8
<p>Comparison of historical air photo (<b>a</b>); recent 0.5 m resolution WorldView-1 panchromatic satellite imagery (<b>b</b>); Landsat SWIR and SWIR trend imagery (significant negative trends are shown in dark blue) (<b>c</b>–<b>e</b>); Sentinel-2 10 m satellite imagery (RGB = NIR, Red, Green) (<b>f</b>); and ArcticDEM elevations (<b>g</b>) for a portion of high-resolution study area 2. Areas with continuous graminoid vegetation appear dark grey in the air photo (red in the Sentinel-2 image), while thermokarst lakes and ponds present a darker tone. WorldView-1 panchromatic imagery integrates both visible and near-infrared wavelengths, which causes this graminoid vegetation to appear brighter than in the air photo. Corresponding sets of images for the other seven high-resolution study areas are presented in <a href="#app1-remotesensing-10-00954" class="html-app">Figure S6</a>.</p>
Full article ">Figure 9
<p>Ground photos (<b>a</b>,<b>d</b>) and low-altitude air photos (<b>b</b>,<b>c</b>) of new melt ponds in upland terrain. Images were captured by the Government of the Northwest Territories on 23 July 2011 [<a href="#B29-remotesensing-10-00954" class="html-bibr">29</a>] at a location just west of study area 3 (72.102°N, 120.90°W). The indicated extents were measured using 0.5 m WorldView Imagery from 31 July 2017 and the 2 m ArcticDEM derived using 0.5 m WorldView imagery from 28 July 2012. Panel (<b>c</b>) shows recently submerged clumps of mountain avens (<span class="html-italic">Dryas integrifolia</span>), an evergreen dwarf shrub. Ice-wedge melt ponds on the eastern portion of Banks Island captured in late July 2011 are shown in (<b>e</b>,<b>f</b>).</p>
Full article ">Figure A1
<p>Average and standard deviation of Landsat channel TOA reflectances and NDVI for a sample of sub-pixel ice-wedge melt ponds (<span class="html-italic">n</span> = 152), upland ice-wedge polygons (<span class="html-italic">n</span> = 200), lowland wet productive tundra (<span class="html-italic">n</span> = 142), and lowland ponds (<span class="html-italic">n</span> = 123). Reflectance was sampled from a Landsat 8 image from 8 August 2013.</p>
Full article ">Figure A2
<p>Hexagonal bin density plot showing the relationship between Landsat 8 TOA SWIR reflectance and Landsat 7 TOA SWIR reflectance. The solid line shows the least-squares linear regression line, and the dashed line indicates a 1:1 relationship.</p>
Full article ">Figure A3
<p>Illustration of the processing steps used to characterise the relationship between the sub-pixel pond area and Landsat SWIR reflectance. Melt ponds extracted using image segmentation and thresholding are shown in cyan.</p>
Full article ">Figure A4
<p>Hexagonal bin density plot showing the relationship between the Landsat SWIR TOA reflectance and the melt pond area aggregated to 90-m resolution (<span class="html-italic">n</span> = 6206). The linear regression equation indicates that a 1% change in SWIR reflectance corresponds to a 5% (~45 m<sup>2</sup>) change in the sub-pixel pond area.</p>
Full article ">
22 pages, 7999 KiB  
Article
Side-Scan Sonar Image Mosaic Using Couple Feature Points with Constraint of Track Line Positions
by Jianhu Zhao, Xiaodong Shang and Hongmei Zhang
Remote Sens. 2018, 10(6), 953; https://doi.org/10.3390/rs10060953 - 15 Jun 2018
Cited by 13 | Viewed by 6028
Abstract
To obtain large-scale seabed surface image, this paper proposes a side-scan sonar (SSS) image mosaic method using couple feature points (CFPs) with constraint of track line positions. The SSS geocoded images are firstly used to form a coarsely mosaicked one and the overlapping [...] Read more.
To obtain large-scale seabed surface image, this paper proposes a side-scan sonar (SSS) image mosaic method using couple feature points (CFPs) with constraint of track line positions. The SSS geocoded images are firstly used to form a coarsely mosaicked one and the overlapping areas between adjacent strip images can be determined based on geographic information. Inside the overlapping areas, the feature point (FP) detection and registration operation are adopted for both strips. According to the detected CFPs and track line positions, an adjustment model is established to accommodate complex local distortions as well as ensure the global stability. This proposed method effectively solves the problem of target ghosting or dislocation and no accumulated errors arise in the mosaicking process. Experimental results show that the finally mosaicked image correctly reflects the object distribution, which is meaningful for understanding and interpreting seabed topography. Full article
(This article belongs to the Special Issue Advances in Undersea Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The procedure of image fusion using a multiresolution fusing method.</p>
Full article ">Figure 2
<p>Determination and segment of the overlapping area.</p>
Full article ">Figure 3
<p>The flow diagram of SSS image mosaic process.</p>
Full article ">Figure 4
<p>Four SSS geocoded images. (I) (II) (III) and (IV) are four SSS images with 50% overlapping</p>
Full article ">Figure 5
<p>The mosaicked image after: the geocoding mosaicking (<b>a</b>); and the refined adjustment (<b>b</b>).</p>
Full article ">Figure 6
<p>The CFPs detected in three segmented overlapping areas (<b>a</b>–<b>c</b>); no CFPs are detected in area (<b>d</b>); the S-L means the starboard side of left SSS image; the S-R means the starboard side of the right SSS image; the blue arrows mean the track line directions.</p>
Full article ">Figure 7
<p>The distribution of CFPs and corresponding track line positions in area (A).</p>
Full article ">Figure 8
<p>Coordinate deviations of CFPs. “☆” and “▷” symbols denote the coordinate deviations in east–west and north–south directions between the reference image and transformed sensed image; “○” and “◇” symbols denote these between the reference image and raw sensed image.</p>
Full article ">Figure 9
<p>Coordinate deviations of the track line positions after adjacent image mosaics.</p>
Full article ">Figure 10
<p>The comparison of image matching results using the MI and CR and the proposed method. (I) and (II) are three image matching results using the MI and CR and the proposed method.</p>
Full article ">Figure 11
<p>Transformed sensed image with (<b>a</b>) and without (<b>b</b>) constraint of track line positions. The black lines denote raw positions and the red lines denote transformed positions.</p>
Full article ">Figure 12
<p>The mosaicked image using the proposed method.</p>
Full article ">Figure 13
<p>Coordinate deviations of CFPs. “☆” and “▷” symbols denote the coordinate deviations in east–west and north–south directions between the reference image and transformed sensed image; “○” and “◇” symbols denote these between the reference image and raw sensed image.</p>
Full article ">Figure 14
<p>Coordinate deviations of the track line positions after multi-strip image mosaics.</p>
Full article ">Figure 15
<p>The SSS image mosaic results using different mosaicking methods.</p>
Full article ">Figure 16
<p>Image registration operation for the raw SSS images (<b>a</b>) and radiation corrected ones (<b>b</b>).</p>
Full article ">Figure 17
<p>The schematic diagram to obtain the track line positions of towfish.</p>
Full article ">Figure 18
<p>The image matching results with different overlapping ratios: (<b>A</b>–<b>C</b>) are from <a href="#remotesensing-10-00953-f005" class="html-fig">Figure 5</a>; and (<b>D</b>,<b>E</b>) are from the SSS images measured in Bohai Sea.</p>
Full article ">Figure 18 Cont.
<p>The image matching results with different overlapping ratios: (<b>A</b>–<b>C</b>) are from <a href="#remotesensing-10-00953-f005" class="html-fig">Figure 5</a>; and (<b>D</b>,<b>E</b>) are from the SSS images measured in Bohai Sea.</p>
Full article ">Figure 19
<p>The performances of three interpolation methods to fill gap with different sizes.</p>
Full article ">
19 pages, 4014 KiB  
Article
Using Google Earth Engine to Map Complex Shade-Grown Coffee Landscapes in Northern Nicaragua
by Lisa C. Kelley, Lincoln Pitcher and Chris Bacon
Remote Sens. 2018, 10(6), 952; https://doi.org/10.3390/rs10060952 - 14 Jun 2018
Cited by 55 | Viewed by 15868
Abstract
Shade-grown coffee (shade coffee) is an important component of the forested tropics, and is essential to the conservation of forest-dependent biodiversity. Despite its importance, shade coffee is challenging to map using remotely sensed data given its spectral similarity to forested land. This paper [...] Read more.
Shade-grown coffee (shade coffee) is an important component of the forested tropics, and is essential to the conservation of forest-dependent biodiversity. Despite its importance, shade coffee is challenging to map using remotely sensed data given its spectral similarity to forested land. This paper addresses this challenge in three districts of northern Nicaragua, here leveraging cloud-based computing techniques within Google Earth Engine (GEE) to integrate multi-seasonal Landsat 8 satellite imagery (30 m), and physiographic variables (temperature, topography, and precipitation). Applying a random forest machine learning algorithm using reference data from two field surveys produced a 90.5% accuracy across ten classes of land cover, with an 82.1% and 80.0% user’s and producer’s accuracy respectively for shade-grown coffee. Comparing classification accuracies obtained from five datasets exploring different combinations of non-seasonal and seasonal spectral data as well as physiographic data also revealed a trend of increasing accuracy when seasonal data were included in the model and a significant improvement (7.8–20.1%) when topographical data were integrated with spectral data. These results are significant in piloting an open-access and user-friendly approach to mapping heterogeneous shade coffee landscapes with high overall accuracy, even in locations with persistent cloud cover. 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>Overview of the classification approach. Red shading indicates primary datasets used in this analysis. *<b>NS</b> refers to non-seasonal data (brightness, wetness, greenness, and temperature); <b>S</b> refers to multi-seasonal data (rainy/dry hot/dry cool brightness, wetness, greenness, and temperature); <b>T</b> refers to elevation, slope, and aspect data; and <b>P</b> refers to the correlation matrix between precipitation and normalized difference vegetation index (NDVI). Treatments are described in greater detail in <a href="#sec2dot8-remotesensing-10-00952" class="html-sec">Section 2.8</a>.</p>
Full article ">Figure 2
<p>Overview of the study area in northwest Nicaragua. (<b>A</b>) Nicaragua in a Central American regional context. (<b>B</b>) Primary climate zones (grey) with study areas noted using thick district boundary lines. The specific study districts included Esteli, Madriz, and Nueva Segovia. (<b>C</b>) Field photo of discussion about participatory mapping approaches in a structurally complex shade-coffee plot. (<b>D</b>) Field photo of “milpa” croplands adjacent to pine stands.</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>D</b>) Datasets used for land-cover classification. (<b>A</b>) Seasonal image composite spanning dry hot months (January–April), visualized using derived brightness, greenness, and wetness bands. (<b>B</b>) Seasonal image composite spanning dry cool months (November–December), visualized in natural color composite using red, green, and blue spectral bands. (<b>C</b>) Correlation matrix of precipitation and NDVI on a 30-day lag. (<b>D</b>) Hillshade derived from Shuttle Radar Topography Mission (SRTM) data. Data are presented for the three study districts. Coordinates from west to east: −86.8° to −85.8°. Coordinates from south to north: 12.8° to 14.1°</p>
Full article ">Figure 4
<p>Land-cover classifications as compared to data on tree canopy coverage. (<b>A</b>) Land cover classification produced by the full multi-seasonal and topographical data model (<b>S+TP</b>, <span class="html-italic">mtry</span> = 3, N = 500) with shade coffee depicted in bright red. (<b>B</b>) Estimated tree canopy coverage for the reference year 2010, available at 30-m resolution [<a href="#B77-remotesensing-10-00952" class="html-bibr">77</a>]. Coordinates from west to east: −86.8° to −85.8°. Coordinates from south to north: 12.8° to 14.1°</p>
Full article ">Figure 5
<p>Variable importance analyses for the full model (% change in mean accuracy, <b>S+TP</b>, <span class="html-italic">mtry</span> = 3, N = 500).</p>
Full article ">
17 pages, 5595 KiB  
Article
Investigation and Analysis of All-Day Atmospheric Water Vapor Content over Xi’an Using Raman Lidar and Sunphotometer Measurements
by Yufeng Wang, Liu Tang, Tianle Gao, Qing Wang, Chuan Lu, Yuehui Song and Dengxin Hua
Remote Sens. 2018, 10(6), 951; https://doi.org/10.3390/rs10060951 - 14 Jun 2018
Cited by 2 | Viewed by 4836
Abstract
All-day atmospheric water vapor content measurements determined by Raman lidar and a sunphotometer were combined to investigate the all-day variation characteristics in the water vapor distribution in Xi’an, China (34.233°N, 108.911°E). To enhance the daytime lidar performance, the wavelet threshold de-noising method is [...] Read more.
All-day atmospheric water vapor content measurements determined by Raman lidar and a sunphotometer were combined to investigate the all-day variation characteristics in the water vapor distribution in Xi’an, China (34.233°N, 108.911°E). To enhance the daytime lidar performance, the wavelet threshold de-noising method is used to filter out the strong solar background light, and effective denoised results are demonstrated with the following optimization: wavelet sym6, the improved threshold function, and the improved threshold selection. The denoised system signal-to-noise ratio (SNR) for the water vapor daytime measurement is validated, with an enhancement of ~3.4 times up to a height of 3 km compared to that of the original signal. The time series of the atmospheric water vapor mixing ratio profiles and the obtained precipitable water vapor (PWV) measured by Raman lidar are used to reveal the temporal and spatial variations in water vapor, and the comparisons with the total column water vapor content (TCWV) measured by a sunphotometer validate the daytime variation trend of the water vapor. All-day continuous observations clearly present a consistent variation trend in the water vapor between the sunphotometer and Raman lidar measurements. The correlation analysis between TCWV and PWV at the layers below 850 hPa and below 700 hPa yields a good positive correlation coefficient (>0.75), indicating that PWV determination in the bottom layer by Raman lidar can directly reflect the variations in the total water vapor content. Moreover, different diurnal variation trends in water vapor are also observed, that is, a downward trend from the afternoon to the night, or a tendency of being high in the morning and afternoon and low at noon, demonstrating the high temporal-spatial variation characteristics of water vapor and close correlation with weather changes. The results reflected and validated that the diurnal variation in water vapor is complicated and can be an indicator of the weather to a certain extent. Full article
(This article belongs to the Special Issue Remote Sensing of Atmospheric Properties)
Show Figures

Figure 1

Figure 1
<p>Schematic of the Raman lidar system for water vapor measurement. (SHG: Second-harmonic generation, THG: Third-harmonic generation, DM: Dichroic mirror, IF: Interference filter).</p>
Full article ">Figure 2
<p>Comparison of denoised SNR and RMSE results by wavelet bases with different lengths: (<b>a</b>) denoised SNR and (<b>b</b>) RMSE.</p>
Full article ">Figure 3
<p>Comparison of the denoised range-corrected square signal of the daytime water vapor Raman scattering signal by wavelet bases of different lengths.</p>
Full article ">Figure 4
<p>Comparison of denoised range-corrected square signals of daytime water vapor Raman scattering signal using different threshold selection methods.</p>
Full article ">Figure 5
<p>Comparison of the daytime lidar results before and after denoising. (<b>a</b>) range-corrected signal of water vapor and nitrogen Raman scattering; (<b>b</b>) water vapor mixing ratio profile; and (<b>c</b>) lidar signal to noise ratio (SNR) for the water vapor measurement.</p>
Full article ">Figure 6
<p>THI plot of water vapor mixing ratio during 15:00–22:00 CST on 15 September 2016, by Raman lidar.</p>
Full article ">Figure 7
<p>Variation trend in PWV in different layers by lidar at 15:00–22:00 CST on 15 September 2016.</p>
Full article ">Figure 8
<p>Time series of AODs at different wavelengths and column water vapor contents by sunphotometer during 15:00–18:00 on 15 September 2016.</p>
Full article ">Figure 9
<p>Correlation analysis of the total water vapor amount with PWV below 700 hPa and below 850 hPa.</p>
Full article ">Figure 10
<p>THI plot of water vapor mixing ratio from 6:00–24:00 on 22 September 2016.</p>
Full article ">Figure 11
<p>Variation trend in PWV in different layers by lidar at 6:00–24:00 CST on 22 September 2016.</p>
Full article ">Figure 12
<p>Time series of AODs at different wavelengths and column water vapor contents by sunphotometer at 6:00–18:00 on 22 September 2016.</p>
Full article ">Figure 13
<p>Correlation analysis of the total water vapor amount with PWV below 700 hPa and below 850 hPa.</p>
Full article ">
19 pages, 4301 KiB  
Article
Quantitative Estimation of Wheat Phenotyping Traits Using Ground and Aerial Imagery
by Zohaib Khan, Joshua Chopin, Jinhai Cai, Vahid-Rahimi Eichi, Stephan Haefele and Stanley J. Miklavcic
Remote Sens. 2018, 10(6), 950; https://doi.org/10.3390/rs10060950 - 14 Jun 2018
Cited by 35 | Viewed by 6967
Abstract
This study evaluates an aerial and ground imaging platform for assessment of canopy development in a wheat field. The dependence of two canopy traits, height and vigour, on fertilizer treatment was observed in a field trial comprised of ten varieties of spring wheat. [...] Read more.
This study evaluates an aerial and ground imaging platform for assessment of canopy development in a wheat field. The dependence of two canopy traits, height and vigour, on fertilizer treatment was observed in a field trial comprised of ten varieties of spring wheat. A custom-built mobile ground platform (MGP) and an unmanned aerial vehicle (UAV) were deployed at the experimental site for standard red, green and blue (RGB) image collection on five occasions. Meanwhile, reference field measurements of canopy height and vigour were manually recorded during the growing season. Canopy level estimates of height and vigour for each variety and treatment were computed by image analysis. The agreement between estimates from each platform and reference measurements was statistically analysed. Estimates of canopy height derived from MGP imagery were more accurate (RMSE = 3.95 cm, R2 = 0.94) than estimates derived from UAV imagery (RMSE = 6.64 cm, R2 = 0.85). In contrast, vigour was better estimated using the UAV imagery (RMSE = 0.057, R2 = 0.57), compared to MGP imagery (RMSE = 0.063, R2 = 0.42), albeit with a significant fixed and proportional bias. The ability of the platforms to capture differential development of traits as a function of fertilizer treatment was also investigated. Both imaging methodologies observed a higher median canopy height of treated plots compared with untreated plots throughout the season, and a greater median vigour of treated plots compared with untreated plots exhibited in the early growth stages. While the UAV imaging provides a high-throughput method for canopy-level trait determination, the MGP imaging captures subtle canopy structures, potentially useful for fine-grained analyses of plants. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The MGP with stereo imaging cameras and calibration target. (<b>b</b>) The UAV with the camera payload. (<b>c</b>) A perspective view of the trial site with ground control points magnified as insets. The red broken line shows the aerial path taken by the UAV.</p>
Full article ">Figure 2
<p>Regression analysis of canopy height estimates for (<b>a</b>) MGP imaging and (<b>b</b>) UAV imaging, relative to reference heights. (<b>c</b>) Distribution of height estimation errors with time, <math display="inline"> <semantics> <msub> <mi>t</mi> <mi>n</mi> </msub> </semantics> </math>.</p>
Full article ">Figure 3
<p>Canopy heights of treated and control plots at time point <math display="inline"> <semantics> <msub> <mi>t</mi> <mi>n</mi> </msub> </semantics> </math> as derived from (<b>a</b>) MGP imaging and (<b>b</b>) UAV imaging. The data shown summarize the results over the 60 plots: 10 varieties and three replicates for each treatment. (<b>c</b>) Difference between average canopy height of treated and untreated plots of each variety derived from MGP imagery (time axis scaled to actual duration).</p>
Full article ">Figure 4
<p>Regression analysis of canopy vigour estimates for (<b>a</b>) MGP imaging and (<b>b</b>) UAV imaging, relative to reference vigour. (<b>c</b>) Distribution of vigour estimation errors with time, <math display="inline"> <semantics> <msub> <mi>t</mi> <mi>n</mi> </msub> </semantics> </math>.</p>
Full article ">Figure 5
<p>Canopy vigour of treated and control plots at time point <math display="inline"> <semantics> <msub> <mi>t</mi> <mi>n</mi> </msub> </semantics> </math> as derived from (<b>a</b>) MGP imaging and (<b>b</b>) UAV imaging. The data shown summarize the results over the 60 plots: 10 varieties and three replicates for each treatment. (<b>c</b>) Difference between average canopy vigour of treated and untreated plots of each variety derived from UAV imagery (time axis scaled to actual duration).</p>
Full article ">Figure 6
<p>RGB, height and vigour of a wheat plot at time <math display="inline"> <semantics> <msub> <mi>t</mi> <mn>2</mn> </msub> </semantics> </math> and <math display="inline"> <semantics> <msub> <mi>t</mi> <mn>4</mn> </msub> </semantics> </math> derived from MGP imaging (top) and UAV imaging (bottom). For the purpose of visualization, the illustrated MGP images are a result of stitching the three partial images per plot using Image Composite Editor (Microsoft) software.</p>
Full article ">Figure A1
<p>Regression analysis of the canopy trait estimates using reduced resolution MGP images for (<b>a</b>) height and (<b>b</b>) vigour, relative to reference traits.</p>
Full article ">Figure A2
<p>Reduced resolution RGB, height and vigour of the wheat plot in <a href="#remotesensing-10-00950-f006" class="html-fig">Figure 6</a> derived from MGP imaging. For the purpose of visualization, the illustrated MGP images are a result of stitching the three partial images per plot using Image Composite Editor (Microsoft) software.</p>
Full article ">Figure A3
<p>Height distribution histogram of plants of a single plot derived from (<b>a</b>) an MGP image and (<b>b</b>) a UAV image.</p>
Full article ">Figure A4
<p>Error between the reference and observed heights from MGP imaging. Data points are the mean ± standard deviation of all plots at <math display="inline"> <semantics> <msub> <mi>t</mi> <mn>1</mn> </msub> </semantics> </math>.</p>
Full article ">Figure A5
<p>A sample rectified stereo image pair captured by the MGP system.</p>
Full article ">
17 pages, 4762 KiB  
Article
Assessing Single-Polarization and Dual-Polarization TerraSAR-X Data for Surface Water Monitoring
by Katherine Irwin, Alexander Braun, Georgia Fotopoulos, Achim Roth and Birgit Wessel
Remote Sens. 2018, 10(6), 949; https://doi.org/10.3390/rs10060949 - 14 Jun 2018
Cited by 20 | Viewed by 5844
Abstract
Three synthetic aperture radar (SAR) data classification methodologies were used to assess the ability of single-polarization and dual-polarization TerraSAR-X (TSX) data to classify surface water, including open water, ice, and flooded vegetation. Multi-polarization SAR observations contain more information than single-polarization SAR, but the [...] Read more.
Three synthetic aperture radar (SAR) data classification methodologies were used to assess the ability of single-polarization and dual-polarization TerraSAR-X (TSX) data to classify surface water, including open water, ice, and flooded vegetation. Multi-polarization SAR observations contain more information than single-polarization SAR, but the availability of multi-polarization data is much lower, which limits the temporal monitoring capabilities. The study area is a principally natural landscape centered on a seasonally flooding river, in which four TSX dual-co-polarized images were acquired between the months of April and June 2016. Previous studies have shown that single-polarization SAR is useful for analyzing surface water extent and change using grey-level thresholding. The H-Alpha–Wishart decomposition, adapted to dual-polarization data, and the Kennaugh Element Framework were used to classify areas of water and flooded vegetation. Although grey-level thresholding was able to identify areas of water and non-water, the percentage of seasonal change was limited, indicating an increase in water area from 8% to 10%, which is in disagreement with seasonal trends. The dual-polarization methods show a decrease in water over the season and indicate a decrease in flooded vegetation, which agrees with expected seasonal variations. When comparing the two dual-polarization methods, a clear benefit of the Kennaugh Elements Framework is the ability to classify change in the transition zones of ice to open water, open water to marsh, and flooded vegetation to land, using the differential Kennaugh technique. The H-Alpha–Wishart classifier was not able to classify ice, and misclassified fields and ice as water. Although single-polarization SAR was effective in classifying open water, the findings of this study confirm the advantages of dual-polarization observations, with the Kennaugh Element Framework being the best performing classification framework. Full article
(This article belongs to the Special Issue Ten Years of TerraSAR-X—Scientific Results)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of study area showing the town of Montpellier, southern extent of Lac-Simon, and stream (Ruisseau Schryer) flowing from A (Lac-Schryer) to B (Baie-de-l’Ours). Base map is provided by the Quebec Ministry of Energy and Natural Resources (MERN) showing urban areas (white), water (blue), forest (dotted green), low vegetation (tan), golf course (light green), and marshlands (dashed areas). Zoomed in orange box of Google Earth imagery from July 2017 shows a flood plain surrounding the stream entering Lac-Simon at B.</p>
Full article ">Figure 2
<p>Time series of Landsat 8 true colour optical images (red-green-blue (RGB): 4-3-2) from the following scenes: (<b>A</b>) 13 April 2016; (<b>B</b>) 29 April 2016; and (<b>C</b>) 16 June 2016.</p>
Full article ">Figure 3
<p>Processing workflow for single- and dual-polarization data to create three final models for each TerraSAR-X (TSX) scene.</p>
Full article ">Figure 4
<p>Images of parameters entropy (<b>A</b>) and alpha (<b>B</b>) for the TerraSAR-X scene from 2 April 2016.</p>
Full article ">Figure 5
<p>Four Kennaugh elements derived from the dual-pol TerraSAR-X image from 2 April 2016. (<b>A</b>) K<sub>0</sub>—the total intensity sum of HH plus VV; (<b>B</b>) K<sub>3</sub>—difference double-bounce minus surface scattering; (<b>C</b>) K<sub>4</sub>—difference HH minus VV intensity; (<b>D</b>) K<sub>7</sub>—phase shift between double-bounce and surface scattering mechanisms. Open water is represented by dark blue in (<b>A</b>) and grey in (<b>B</b>). Flooded vegetation is represented by grey and yellow in (<b>B</b>) and red and yellow in (<b>C</b>). Ice cover is represented by light blue in (<b>A</b>) and dark blue in (<b>B</b>). Inundated vegetation is shown in yellow/red in (<b>C</b>) and cyan in (<b>D</b>).</p>
Full article ">Figure 6
<p>Grey-level thresholding classified models showing water (black) and other (grey) for (<b>A</b>) 2 April 2016; (<b>B</b>) 24 April 2016; (<b>C</b>) 5 May 2016; and (<b>D</b>) 18 June 2016. Coloured boxes indicate example areas of temporal change: blue—ice melting; red—marshland dries out; yellow—areas of misclassification due to ice (<b>A</b>) and vegetation (<b>C</b>).</p>
Full article ">Figure 7
<p>H-Alpha–Wishart classified models showing water (black), flooded vegetation (blue), and other (grey) for (<b>A</b>) 2 April 2016; (<b>B</b>) 24 April 2016; (<b>C</b>) 5 May 2016; and (<b>D</b>) 18 June 2016. Coloured boxes indicate example areas of change: blue—golf course misclassified as water; red—marshland dries out; yellow—flooded vegetation decreases, and misclassification of fields.</p>
Full article ">Figure 8
<p>False colour composites of the processed Kennaugh elements, K<sub>3</sub>-K<sub>0</sub>-K<sub>4</sub>, from (<b>A</b>) 2 April, (<b>B</b>) 24 April, (<b>C</b>) 5 May, and (<b>D</b>) 18 June 2016. Open water appears in pink, ice in dark purple, flooded vegetation in white/light pink, and ‘other’ in green and blue.</p>
Full article ">Figure 9
<p>Graphs of the average of each class for the Kennaugh elements used to classify water (red), flooded vegetation (yellow), and other (black). (<b>A</b>) 2 April 2016; (<b>B</b>) 24 April 2016; (<b>C</b>) 5 May 2016; and (<b>D</b>) 18 June 2016.</p>
Full article ">Figure 10
<p>Kennaugh Element models classified showing water (black), flooded vegetation (blue), and other (grey) for (<b>A</b>) 2 April 2016; (<b>B</b>) 24 April 2016; (<b>C</b>) 5 May 2016; and (<b>D</b>) 18 June 2016. Coloured boxes indicate example areas of change: blue—golf course misclassified as water; red—marshland dries with time; yellow—flooded vegetation decrease, and misclassification of field areas; green—ice melting.</p>
Full article ">Figure 11
<p>False colour composite of the differential Kennaugh elements, K<sub>0</sub>-K<sub>3</sub>-K<sub>4</sub>, differenced between the 18 June 2016 scene and the 2 April 2016 scene. Red represents the change from flooded vegetation to land. Green represents the change from ice to open water. Yellow represents the change from open water to marshland.</p>
Full article ">Figure 12
<p>Graphs showing percent of water and percent of flooded vegetation classified through time for each of the classification methods. Lines: Blue diamond—unsupervised k-means classification on Kennaugh Elements, orange square—H-Alpha–Wishart unsupervised classification, and grey triangle—grey-level thresholding.</p>
Full article ">
22 pages, 4469 KiB  
Article
A Ship Detector Applying Principal Component Analysis to the Polarimetric Notch Filter
by Tao Zhang, Armando Marino, Huilin Xiong and Wenxian Yu
Remote Sens. 2018, 10(6), 948; https://doi.org/10.3390/rs10060948 - 14 Jun 2018
Cited by 13 | Viewed by 4314
Abstract
Ship detection using polarimetric synthetic aperture radar (PolSAR) data has attracted a lot of attention in recent years. Polarimetry can provide information regarding the scattering mechanisms of targets, which helps discriminate between ships and sea clutter. This enhancement is particularly valuable when we [...] Read more.
Ship detection using polarimetric synthetic aperture radar (PolSAR) data has attracted a lot of attention in recent years. Polarimetry can provide information regarding the scattering mechanisms of targets, which helps discriminate between ships and sea clutter. This enhancement is particularly valuable when we aim at detecting smaller vessels in rough sea states. This work exploits a ship detector called the Geometrical Perturbation-Polarimetric Notch Filter (GP-PNF), and it is aimed at improving its performance especially when less polarimetric images are available (e.g., dual-polarimetric data). The idea is to design a new polarimetric feature vector containing more features that are renowned to allow separation between ships and sea clutter. Then, a Principal Component Analysis (PCA) is further used to reduce the dimensionality of the new feature space. Experiments on four real Sentinel-1 datasets are carried out to demonstrate the validity of the proposed method and compare it against other ship detectors. Analyses of the experimental results show that the proposed algorithm can not only reduce the false alarms significantly, but also enhance the target-to-clutter ratio (TCR) so that it can more effectively detect weaker ships. Full article
(This article belongs to the Special Issue Remote Sensing of Target Detection in Marine Environment)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The flowchart of the proposed method.</p>
Full article ">Figure 2
<p>The four datasets. (<b>a</b>) the optical image of the first scene; (<b>b</b>) the VH image of the first scene; (<b>c</b>) the VH image of dataset A (Zoom in the green box of (<b>b</b>)); (<b>d</b>) the optical image of the second scene; (<b>e</b>) the VH image of the second scene; (<b>f</b>) the VH image of dataset B (Zoom in the green box of (<b>e</b>)); (<b>g</b>) the optical image of the third scene; (<b>h</b>) the VH image of dataset C; (<b>i</b>) the optical image of the fourth scene; (<b>j</b>) the VH image of dataset D. Orange circles indicate the ships. The yellow icons in (<b>g</b>,<b>i</b>) represent the locations of (<b>h</b>,<b>j</b>).</p>
Full article ">Figure 3
<p>Ship detection results of different methods on A. (<b>a</b>) VH; (<b>b</b>) VH<math display="inline"><semantics> <msub> <mrow/> <mi>filter</mi> </msub> </semantics></math>; (<b>c</b>) MTC; (<b>d</b>) DoD; (<b>e</b>) RS; (<b>f</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>g</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>; (<b>h</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>i</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 4
<p>Ship detection results of different methods on B. (<b>a</b>) VH; (<b>b</b>) VH<math display="inline"><semantics> <msub> <mrow/> <mi>filter</mi> </msub> </semantics></math>; (<b>c</b>) MTC; (<b>d</b>) DoD; (<b>e</b>) RS; (<b>f</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>g</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>; (<b>h</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>i</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>. Red circles indicate missing ships. Yellow dashed circles mean false alarms.</p>
Full article ">Figure 5
<p>DoD values. (<b>a</b>) The sample area of sea clutter; (<b>b</b>) S1. Note that the units of the colorbar are normalized.</p>
Full article ">Figure 6
<p>RS values. (<b>a</b>) The sample area of sea clutter; (<b>b</b>) S1. Note that the units of the colorbar are normalized.</p>
Full article ">Figure 7
<p><math display="inline"><semantics> <msub> <mi>P</mi> <mi>T</mi> </msub> </semantics></math> values of the sample sea area and S1. (<b>a</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: The values of sea clutter; (<b>b</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of S1; (<b>c</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>d</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of S1; (<b>e</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>f</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of S1; (<b>g</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>h</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of S1. Note that the units of the colorbar are normalized.</p>
Full article ">Figure 8
<p>TCR values of the 19 ships in B.</p>
Full article ">Figure 9
<p>ROC curves. (<b>a</b>) the ROC curves (<math display="inline"><semantics> <msub> <mi>P</mi> <mi>f</mi> </msub> </semantics></math> belongs to [0, 1]); (<b>b</b>) zoom in on the <math display="inline"><semantics> <msub> <mi>P</mi> <mi>f</mi> </msub> </semantics></math> ([0, 10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </semantics></math>]); (<b>c</b>) zoom in on the ROC curves of PNF<math display="inline"><semantics> <msub> <mrow/> <mi mathvariant="normal">X</mi> </msub> </semantics></math> methods.</p>
Full article ">Figure 10
<p>Ship detection results of different methods on C. (<b>a</b>) VH; (<b>b</b>) VH<math display="inline"><semantics> <msub> <mrow/> <mi>filter</mi> </msub> </semantics></math>; (<b>c</b>) MTC; (<b>d</b>) DoD; (<b>e</b>) RS; (<b>f</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>g</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>; (<b>h</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>i</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>. Red circles indicate missing ships. Yellow dashed circles mean false alarms.</p>
Full article ">Figure 11
<p><math display="inline"><semantics> <msub> <mi>P</mi> <mi>T</mi> </msub> </semantics></math> values of the sample sea area and T5. (<b>a</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>b</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of S1; (<b>c</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>d</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of S1; (<b>e</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>f</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>: the values of S1; (<b>g</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of sea clutter; (<b>h</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>: the values of S1. Note that the units of the colorbar are normalized.</p>
Full article ">Figure 12
<p>ROC curves. (<b>a</b>) the ROC curves (<math display="inline"><semantics> <msub> <mi>P</mi> <mi>f</mi> </msub> </semantics></math> belongs to [0, 1]); (<b>b</b>) zoom in on the <math display="inline"><semantics> <msub> <mi>P</mi> <mi>f</mi> </msub> </semantics></math> ([5 × 10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </semantics></math>, 10<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </semantics></math>]); (<b>c</b>) zoom in on the ROC curves of PNF<math display="inline"><semantics> <msub> <mrow/> <mi mathvariant="normal">X</mi> </msub> </semantics></math> methods.</p>
Full article ">Figure 13
<p>Ship detection results of different methods on D. (<b>a</b>) VH; (<b>b</b>) VH<math display="inline"><semantics> <msub> <mrow/> <mi>filter</mi> </msub> </semantics></math>; (<b>c</b>) MTC; (<b>d</b>) DoD; (<b>e</b>) RS; (<b>f</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>g</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>3</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>; (<b>h</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi mathvariant="normal">D</mi> </mrow> </msub> </semantics></math>; (<b>i</b>) PNF<math display="inline"><semantics> <msub> <mrow/> <mrow> <mn>6</mn> <mi>DPCA</mi> </mrow> </msub> </semantics></math>. Red circles indicate missing ships. Yellow dashed circles mean false alarms.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop