[go: up one dir, main page]

Next Issue
Volume 13, March-2
Previous Issue
Volume 13, February-2
 
 
remotesensing-logo

Journal Browser

Journal Browser

Remote Sens., Volume 13, Issue 5 (March-1 2021) – 214 articles

Cover Story (view full-size image): Sea surface temperature (SST) is a fundamental property of the ocean surface and one of the first ocean variables to be studied using satellites. Quantifying the accuracy and precision of satellite SST data requires comparing it with in-situ data. In nearshore coastal waters this is not well known, owing to a lack of in-situ data. Here, we compare a Smartfin, a surfboard fin designed to measure ocean temperature in the nearshore, with an infrared SST autonomous radiometer (ISAR) and an underway oceanographic temperature sensor (UOTS) on an expedition through the Atlantic Ocean. We found a mean absolute difference between Smartfin and UOTS of ­0.06 K and Smartfin and ISAR of 0.12 K. Differences were related to sampling depth and environmental variability. Results add confidence to the use of Smartfin as a tool for satellite validation. 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:
21 pages, 9309 KiB  
Article
Examining Relationships between Heat Requirement of Remotely Sensed Green-Up Date and Meteorological Indicators in the Hulun Buir Grassland
by Jian Guo, Xiuchun Yang, Fan Chen, Jianming Niu, Sha Luo, Min Zhang, Yunxiang Jin, Ge Shen, Ang Chen, Xiaoyu Xing, Dong Yang and Bin Xu
Remote Sens. 2021, 13(5), 1044; https://doi.org/10.3390/rs13051044 - 9 Mar 2021
Cited by 2 | Viewed by 3132
Abstract
The accumulation of heat and moderate precipitation are the primary factors that are used by grasslands to trigger a green-up date. The accumulated growing degree-days (AGDD) requirement over the preseason is an important indicator of the response of grassland spring phenology to climate [...] Read more.
The accumulation of heat and moderate precipitation are the primary factors that are used by grasslands to trigger a green-up date. The accumulated growing degree-days (AGDD) requirement over the preseason is an important indicator of the response of grassland spring phenology to climate change. This study adopted the Normalized Difference Phenology Index (NDPI), which derived from the Moderate Resolution Imaging Spectroradiometer (MODIS), to extract annual green-up dates in the Hulun Buir grassland in China between 2001–2015. Our analysis indicated that the range (standard deviation) and trend for the green-up date were DOY (day of year) 104 to DOY 144 (10.6 days) and −2.0 days per decade. Nine point two percent of the study area had significant (p < 0.05) changes in AGDD requirements. The partial correlations between the AGDD requirements and chilling days (67.04%, pixels proportion) were negative and significant (p < 0.05). The partial correlations between the AGDD requirement and precipitation (28.87%) were positive and significant (p < 0.05). Finally, the partial correlation between the AGDD requirement and insolation (97.65%) were positive and significant (p < 0.05). The results of this study could reveal the response of vegetation to climate warming and contribute to improving the phenological mechanism model of different grassland types in future research. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>The multiyear annual average temperature during 1979–2015.</p>
Full article ">Figure 2
<p>The multiyear annual average precipitation during 1979–2015.</p>
Full article ">Figure 3
<p>Administrative divisions and grassland classification in Hulun Buir. In this figure, county boundaries were used to divide the different districts. The name of each administrative district is given in bold. The blank areas represent non-grassland areas. The grassland classification dataset is based on a 1:1,000,000 Vegetation Atlas of China [<a href="#B49-remotesensing-13-01044" class="html-bibr">49</a>].</p>
Full article ">Figure 4
<p>Time series reconstruction results under the double-logistic base normalized difference phenology index (NDPI) (<b>a</b>) and normalized difference vegetation index (NDVI) (<b>b</b>).</p>
Full article ">Figure 5
<p>Flow chart of the key indicator computations and statistical analysis. NDPI: normalized difference phenology index. AGDD: accumulated growing degree-days. CD: chilling days.</p>
Full article ">Figure 6
<p>Mean green-up dates between 2001 and 2015.</p>
Full article ">Figure 7
<p>Standard deviations of the green-up dates.</p>
Full article ">Figure 8
<p>Green-up dates in the different types of grasslands.</p>
Full article ">Figure 9
<p>Spatial pattern of the temporal trends in green-up dates (in days per decade) between 2001 and 2015. The inset shown at the top-right of the figure indicates pixels with a significant (<span class="html-italic">p</span> &lt; 0.05) increase (red) or decrease (green). The middle-left inset shows the frequency distribution of trends corresponding to the values indicated by the map legend.</p>
Full article ">Figure 10
<p>Spatial pattern of the temporal trends in accumulated growing degree-days requirements (in °C-days year<sup>–1</sup>) between 2001 and 2015. The top-right inset indicates pixels with a significant (<span class="html-italic">p</span> &lt; 0.05) increase (red) or decrease (blue). The middle-left inset shows the frequency distribution of trends corresponding to the values indicated by the map legend.</p>
Full article ">Figure 11
<p>Spatial pattern of the temporal trends in (<b>a</b>) chilling days in days year<sup>−1</sup>, (<b>b</b>) precipitation in mm year<sup>−1</sup>, and (<b>c</b>) insolation in W m<sup>–2</sup> year<sup>–1</sup>. The top-right insets indicate pixels with a significant (<span class="html-italic">p</span> &lt; 0.05) increase (blue in (<b>a</b>,<b>b</b>) and red in (<b>c</b>)) or decrease (red in (<b>a</b>,<b>b</b>) and blue in (<b>c</b>)). The middle-left insets show the frequency distributions of trends corresponding to the values indicated by the map legends.</p>
Full article ">Figure 12
<p>Spatial patterns of the interannual partial correlations between the accumulated growing degree-days requirement and chilling days (<b>a</b>), precipitation (<b>b</b>), and insolation (<b>c</b>). Partial correlation coefficient values of ±0.68, ±0.55, and ±0.48 correspond to significance at <span class="html-italic">p</span> = 0.01, <span class="html-italic">p</span> = 0.05, and <span class="html-italic">p</span> = 0.10, respectively. Top-left insets show the frequency distributions of the correlation coefficients corresponding to values indicated by the map legends.</p>
Full article ">Figure 13
<p>The average annual chilling days (CD) during 2001–2015.</p>
Full article ">Figure 14
<p>Spatial patterns of the interannual partial correlations between the accumulated growing degree-days requirements and chilling days (<b>a</b>), precipitation (<b>b</b>), and insolation (<b>c</b>) under the 0 °C threshold 30 days before the green-up date. Partial correlation coefficient values of ±0.68, ±0.55, and ±0.48 correspond to significance at <span class="html-italic">p</span> = 0.01, <span class="html-italic">p</span> = 0.05, and <span class="html-italic">p</span> = 0.10, respectively. Top-left insets show the frequency distributions of the correlation coefficients corresponding to the values indicated by the map legends.</p>
Full article ">Figure 15
<p>Spatial patterns of the interannual partial correlations between the accumulated growing degree-days requirement and chilling days (<b>a</b>), precipitation (<b>b</b>), and insolation (<b>c</b>) under the 5 °C threshold 60 days before the green-up date. Partial correlation coefficient values of ±0.68, ±0.55, and ±0.48 correspond to significance at <span class="html-italic">p</span> = 0.01, <span class="html-italic">p</span> = 0.05, and <span class="html-italic">p</span> = 0.10, respectively. Top-left insets show the frequency distributions of the correlation coefficients corresponding to the values indicated by the map legends.</p>
Full article ">
27 pages, 5648 KiB  
Article
Deriving Water Quality Parameters Using Sentinel-2 Imagery: A Case Study in the Sado Estuary, Portugal
by Giulia Sent, Beatriz Biguino, Luciane Favareto, Joana Cruz, Carolina Sá, Ana Inés Dogliotti, Carla Palma, Vanda Brotas and Ana C. Brito
Remote Sens. 2021, 13(5), 1043; https://doi.org/10.3390/rs13051043 - 9 Mar 2021
Cited by 48 | Viewed by 7537
Abstract
Monitoring water quality parameters and their ecological effects in transitional waters is usually performed through in situ sampling programs. These are expensive and time-consuming, and often do not represent the total area of interest. Remote sensing techniques offer enormous advantages by providing cost-effective [...] Read more.
Monitoring water quality parameters and their ecological effects in transitional waters is usually performed through in situ sampling programs. These are expensive and time-consuming, and often do not represent the total area of interest. Remote sensing techniques offer enormous advantages by providing cost-effective systematic observations of a large water system. This study evaluates the potential of water quality monitoring using Sentinel-2 observations for the period 2018–2020 for the Sado estuary (Portugal), through an algorithm intercomparison exercise and time-series analysis of different water quality parameters (i.e., colored dissolved organic matter (CDOM), chlorophyll-a (Chl-a), suspended particulate matter (SPM), and turbidity). Results suggest that Sentinel-2 is useful for monitoring these parameters in a highly dynamic system, however, with challenges in retrieving accurate data for some of the variables, such as Chl-a. Spatio-temporal variability results were consistent with historical data, presenting the highest values of CDOM, Chl-a, SPM and turbidity during Spring and Summer. This work is the first study providing annual and seasonal coverage with high spatial resolution (10 m) for the Sado estuary, being a key contribution for the definition of effective monitoring programs. Moreover, the potential of remote sensing methodologies for continuous water quality monitoring in transitional systems under the scope of the European Water Framework Directive is briefly discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the Sado estuary in the Portuguese territory and of the 8 sampling stations within the estuary. (<b>A</b>) (blue box) and (<b>B</b>) (red box) are the outermost and innermost areas of the estuary, respectively.</p>
Full article ">Figure 2
<p>Processing chain applied to Sentinel-2 imagery for water quality parameters retrieval. ρw refers to the water-leaving reflectance obtained from the different AC processors: Acolite, Polymer, and AC-C2RCC.</p>
Full article ">Figure 3
<p>Ternary plot of absorption at 443 nm of CDOM (<span class="html-italic">a</span><sub>CDOM</sub>, N = 33), non-algal particles (<span class="html-italic">a</span><sub>nap</sub>, N = 33) and phytoplankton (<span class="html-italic">a</span><sub>phy</sub>, N = 33) produced using the in situ data considered for the algorithms intercomparison exercise.</p>
Full article ">Figure 4
<p>Sentinel-2 derived <math display="inline"><semantics> <mrow> <msub> <mi>ρ</mi> <mi>w</mi> </msub> </mrow> </semantics></math> spectra for the different processors under investigation (Polymer: solid, C2RCC: dotted and ACOLITE: dashed lines). Mean and standard deviation (bars) of the match-ups available for all the processors at stations in the outer (A) and inner basin (B) are shown with blue and red lines, respectively.</p>
Full article ">Figure 5
<p>Taylor diagrams (top) and Target plots (bottom) showing the performances of the different combinations of AC processors and bio-optical algorithms respect to in situ observations. For CDOM, SPM, and turbidity, only algorithms that presented APD &lt; 100% were considered. For the names reference, the first letter refers to the AC processor (a (Acolite), p (polymer), c (C2RCC)) followed by the bio-optical algorithm and band involved.</p>
Full article ">Figure 6
<p>Match-ups of the best performing AC plus bio-optical algorithm of the four WQ parameters under investigation. Statistics shown correspond to regions A and B together, but they are identified with different colors. The full statistics of the match-ups results are presented in <a href="#remotesensing-13-01043-t004" class="html-table">Table 4</a>. (Please note the different scales).</p>
Full article ">Figure 7
<p>Time series of S2-MSI-derived and in situ water quality parameters in the two regions of the estuary, A (blue) and B (red). The time-series considered the average of the pixels corresponding to the sampling stations included in each of the areas. Please note the log-scale in the Chl-<span class="html-italic">a</span> plots. It should be noted that the in situ variability does not consider the tidal effect since the in situ samplings were always performed at high tide.</p>
Full article ">Figure 8
<p>Inter-annual seasonal variability of S2-MSI-derived and in situ water quality parameters in the two regions of the estuary, A (blue) and B (red). The time-series considered the average of the pixels corresponding to the sampling stations included in each of the areas. Please note the log-scale in the Chl-<span class="html-italic">a</span> plots. It should be noted that satellite data include a mixture of tidal conditions while the in situ was always collected at high tide.</p>
Full article ">Figure 9
<p>Seasonal averages of the four water quality parameters for the period between March 2018 and March 2020. Averages of <span class="html-italic">a</span><sub>CDOM</sub>, Chl-<span class="html-italic">a</span>, SPM, and turbidity were performed via the best performing processing chain selected during the algorithms intercomparison exercise (<a href="#sec3dot1-remotesensing-13-01043" class="html-sec">Section 3.1</a>). Please note that no bottom corrections were applied.</p>
Full article ">Figure 10
<p>(<b>A</b>) S2A RGB image and (<b>B</b>) SPM map of the Sado estuary acquired on 19 February 2020. Inside the red box, the sediment plume produced by the dredging activities is clearly visible. The SPM map (<b>B</b>) was produced through the best performing processing chain selected for SPM retrieval (AC-C2RCC in combination with Nechad algorithm at 740 nm). Please note that no bottom reflection correction was implemented.</p>
Full article ">
22 pages, 5559 KiB  
Article
Advanced Dual-Satellite Method for Detection of Low Stratus and Fog near Japan at Dawn from FY-4A and Himawari-8
by Jung-Hyun Yang, Jung-Moon Yoo and Yong-Sang Choi
Remote Sens. 2021, 13(5), 1042; https://doi.org/10.3390/rs13051042 - 9 Mar 2021
Cited by 9 | Viewed by 3089
Abstract
The detection of low stratus and fog (LSF) at dawn remains limited because of their optical features and weak solar radiation. LSF could be better identified by simultaneous observations of two geostationary satellites from different viewing angles. The present study developed an advanced [...] Read more.
The detection of low stratus and fog (LSF) at dawn remains limited because of their optical features and weak solar radiation. LSF could be better identified by simultaneous observations of two geostationary satellites from different viewing angles. The present study developed an advanced dual-satellite method (DSM) using FY-4A and Himawari-8 for LSF detection at dawn in terms of probability indices. Optimal thresholds for identifying the LSF from the spectral tests in DSM were determined by the comparison with ground observations of fog and clear sky in/around Japan between April to November of 2018. Then the validation of these thresholds was carried out for the same months of 2019. The DSM essentially used two traditional single-satellite tests for daytime such as the 0.65-μm reflectance (R0.65), and the brightness temperature difference between 3.7 μm and 11 μm (BTD3.7-11); in addition to four more tests such as Himawari-8 R0.65 and BTD13.5-8.5, the dual-satellite stereoscopic difference in BTD3.7-11 (ΔBTD3.7-11), and that in the Normalized Difference Snow Index (ΔNDSI). The four were found to show very high skill scores (POD: 0.82 ± 0.04; FAR, 0.10 ± 0.04). The radiative transfer simulation supported optical characteristics of LSF in observations. The LSF probability indices (average POD: 0.83, FAR: 0.10) were constructed by a statistical combination of the four to derive the five-class probability values of LSF occurrence in a grid. The indices provided more details and useful results in LSF spatial distribution, compared to the single satellite observations (i.e., R0.65 and/or BTD3.7-11) of either LSF or no LSF. The present DSM could apply for remote sensing of environmental phenomena if the stereoscopic viewing angle between two satellites is appropriate. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Viewing zenith angles (VZAs) of two geostationary satellites (Himawari-8 and FY-4A) available for near-simultaneous observations of the low stratus and fog (LSF) at dawn near Japan. The VZA difference between the satellites is 40.4° at the Himawari-8 nadir point.</p>
Full article ">Figure 2
<p>Locations of 31 meteorological stations near Japan used for LSF verification during 2018–2019. The stations are located in Japan except for the 31st Ulleung station in South Korea.</p>
Full article ">Figure 3
<p>Flow diagram for LSF detection in terms of probability index (PI) near Japan at dawn during the period of March to November of 2018–2019, based on near-simultaneous satellite observations of Himawari-8 and FY-4A. The meaning of the acronyms in the diagram is explained in <a href="#remotesensing-13-01042-t0A1" class="html-table">Table A1</a>.</p>
Full article ">Figure 4
<p>Average values of eight satellite-observed variables with their standard deviation (±1σ) for two weather conditions (black for LSF and red for clear sky) at dawn during the control period of 2018 near Japan. (<b>a</b>) Brightness temperatures (BTD<sub>13.5-8.5</sub>, ΔBTD<sub>3.7-11</sub>, BTD<sub>3.7-11</sub>, and LST-BT<sub>11</sub>). (<b>b</b>) Reflectances (or NDSI) of ΔNDSI, NDSI, R<sub>0.65</sub>, and ΔR<sub>0.65</sub>.</p>
Full article ">Figure 5
<p>Scatter diagrams of (<b>a</b>) BTD<sub>13.5-8.5</sub>, (<b>b</b>) ΔBTD<sub>3.7-11</sub>, (<b>c</b>) ΔNDSI, and (<b>d</b>) R<sub>0.65</sub> with respect to LST-BT<sub>11</sub> for the weather phenomena of LSF (red circle) and clear sky (blue triangle) at dawn during 2018 near Japan. The values in the ordinate of the gray-shaded bands indicate the LSF threshold ranges for each variable. The satellite-observed value of one positive standard deviation (i.e., +1σ; upper boundary) for LSF detection is used to remove the middle or high-level clouds without accompanying LSF, except for ΔNDSI. The negative value of -1σ (lower boundary) is needed for ΔNDSI of <a href="#remotesensing-13-01042-f005" class="html-fig">Figure 5</a>c based on the results shown in <a href="#remotesensing-13-01042-f004" class="html-fig">Figure 4</a>b.</p>
Full article ">Figure 6
<p>RTM LSF simulation in the domain of (<b>a</b>) BTD<sub>13.5-8.5</sub> vs. LST-BT<sub>11</sub>, (<b>b</b>) ΔBTD<sub>3.7-11</sub> vs. LST-BT<sub>11</sub>, (<b>c</b>) ΔNDSI vs. LST-BT<sub>11</sub>, and (<b>d</b>) R<sub>0.67</sub> vs. LST-BT<sub>11</sub> at dawn during 2018 near Japan under the different LSF conditions. The yellow and blue symbols in the figures denote the fog layer at 0–1 km and 0–2 km without higher clouds. Middle clouds of water/ice at 4–6 km height are shown in red and navy, respectively. Green asterisk means higher cloud (8–10 km) above the 0–2km fog layer. The average values of the Himawari-8 angles (i.e., SZA, RAA, and VZA) for 176 LSF cases were used as RTM input. The RTM conditions are described in <a href="#remotesensing-13-01042-t0A2" class="html-table">Table A2</a>. The satellite-observed mean values are shown for LSF (black asterisk) and clear sky (gray-cross). The clear-sky values from the observations and simulations (pink asterisk) are depicted in black circles. Some simulated data out of range in the domain were not shown for a comparison with the observations of <a href="#remotesensing-13-01042-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 6 Cont.
<p>RTM LSF simulation in the domain of (<b>a</b>) BTD<sub>13.5-8.5</sub> vs. LST-BT<sub>11</sub>, (<b>b</b>) ΔBTD<sub>3.7-11</sub> vs. LST-BT<sub>11</sub>, (<b>c</b>) ΔNDSI vs. LST-BT<sub>11</sub>, and (<b>d</b>) R<sub>0.67</sub> vs. LST-BT<sub>11</sub> at dawn during 2018 near Japan under the different LSF conditions. The yellow and blue symbols in the figures denote the fog layer at 0–1 km and 0–2 km without higher clouds. Middle clouds of water/ice at 4–6 km height are shown in red and navy, respectively. Green asterisk means higher cloud (8–10 km) above the 0–2km fog layer. The average values of the Himawari-8 angles (i.e., SZA, RAA, and VZA) for 176 LSF cases were used as RTM input. The RTM conditions are described in <a href="#remotesensing-13-01042-t0A2" class="html-table">Table A2</a>. The satellite-observed mean values are shown for LSF (black asterisk) and clear sky (gray-cross). The clear-sky values from the observations and simulations (pink asterisk) are depicted in black circles. Some simulated data out of range in the domain were not shown for a comparison with the observations of <a href="#remotesensing-13-01042-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 7
<p>Statistical verification of (<b>a</b>) POD minus FAR, (<b>b</b>) POD, (<b>c</b>) FAR, and (<b>d</b>) CSI with respect to the LSF threshold values of eight satellite-observed variables during the experimental period of 2019. The verification was performed based on the total ground-observations of 171 fog and 110 clear-sky occurrences at dawn near Japan.</p>
Full article ">Figure 8
<p>(<b>a</b>) Same as <a href="#remotesensing-13-01042-f004" class="html-fig">Figure 4</a>a except for the period of 2019. (<b>b</b>) Same as <a href="#remotesensing-13-01042-f004" class="html-fig">Figure 4</a>b except for the period of 2019.</p>
Full article ">Figure 9
<p>The number of cases that have passed the threshold test of the four satellite-observed variables (i.e., BTD<sub>13.5-8.5</sub> in gray rectangle, ΔNDSI in orange, ΔBTD<sub>3.7-11</sub> in blue, and R<sub>0.65</sub> in green) for detection of (<b>a</b>) LSF and (<b>b</b>) clear sky at dawn during 2019 near Japan. The zones of red (Class 1 for the possibility of the weather phenomena), yellow (Class 2), blue (Class 3), and white (Class 4) are defined in <a href="#remotesensing-13-01042-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 10
<p>Spatial distributions of fog probability at dawn (06:00 LST) on 9 July 2019, near Japan from detection methods of (<b>a</b>) R<sub>0.65</sub> threshold, (<b>b</b>) BTD<sub>3.7-11</sub> threshold, and (<b>c</b>) LSF PI in this study. (<b>d</b>) BT<sub>11</sub> distribution for cloud-top temperatures, and (<b>e</b>) SYNOP map for station-observed fog occurrences. Fog occurrences at the stations (pink triangles in <a href="#remotesensing-13-01042-f009" class="html-fig">Figure 9</a>a,b) were reported by ASOS.</p>
Full article ">Figure 11
<p>(<b>a</b>) Same as <a href="#remotesensing-13-01042-f010" class="html-fig">Figure 10</a>a except for the date and time (05:30 LST on 4 July 2019), (<b>b</b>) Same as <a href="#remotesensing-13-01042-f010" class="html-fig">Figure 10</a>b except for the date and time. (<b>c</b>) Same as <a href="#remotesensing-13-01042-f010" class="html-fig">Figure 10</a>c except for the date and time. (<b>d</b>) Same as <a href="#remotesensing-13-01042-f010" class="html-fig">Figure 10</a>d except for the date and time, and no SYNOP map (not available at the time).</p>
Full article ">
15 pages, 5141 KiB  
Article
The Horizontal Distribution of Branch Biomass in European Beech: A Model Based on Measurements and TLS Based Proxies
by César Pérez-Cruzado, Christoph Kleinn, Paul Magdon, Juan Gabriel Álvarez-González, Steen Magnussen, Lutz Fehrmann and Nils Nölke
Remote Sens. 2021, 13(5), 1041; https://doi.org/10.3390/rs13051041 - 9 Mar 2021
Cited by 7 | Viewed by 3277
Abstract
Forest biomass is currently among the most important and most researched target variables in forest monitoring. The common approach of observing individual tree biomass in forest inventory is to assign the total tree biomass to the dimensionless point of the tree position. However, [...] Read more.
Forest biomass is currently among the most important and most researched target variables in forest monitoring. The common approach of observing individual tree biomass in forest inventory is to assign the total tree biomass to the dimensionless point of the tree position. However, the tree biomass, in particular in the crown, is horizontally distributed above the crown projection area. This horizontal distribution of individual tree biomass (HBD) has not attracted much attention—but if quantified, it can improve biomass estimation and help to better represent the spatial distribution of forest fuel. In this study, we derive a first empirical model of the branch HBD for individual trees of European beech (Fagus sylvatica L.). We destructively measured 23 beech trees to derive an empirical model for the branch HBD. We then applied Terrestrial Laser Scanning (TLS) to a subset of 17 trees to test a simple point cloud metric predicting the branch HBD. We observed similarities between a branch HBD and commonly applied taper functions, which inspired our HBD model formulations. The models performed well in representing the HBD both for the measured biomass, and the TLS-based metric. Our models may be used as first approximations to the HBD of individual trees—while our methodological approach may extend to trees of different sizes and species. Full article
(This article belongs to the Special Issue 3D Point Clouds in Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Illustration of (<b>A</b>) the measurements made for each sample tree before felling, and (<b>B</b>) for individual branch measurements. <span class="html-italic">dbh</span> = diameter at breast height, <span class="html-italic">h</span> = tree height, <span class="html-italic">bh</span> = branch height at the stem, <span class="html-italic">bd</span> = branch base diameter, <span class="html-italic">bl</span> = branch length, and <span class="html-italic">α</span> = vertical branch angle.</p>
Full article ">Figure 2
<p>Observed horizontal branch biomass density over relative crown radii (<span class="html-italic">rcr</span>) for the 23 destructively sampled beech trees.</p>
Full article ">Figure 3
<p>A subset of 17 sample trees was carefully felled and re-erected (<b>A</b>) in an open space in (<b>B</b>) a tailor-made tree holder. These sample trees were then scanned (by TLS) using the scan and target locations as illustrated in (<b>C</b>).</p>
Full article ">Figure 4
<p>(<b>A</b>) Merged single tree point cloud. (<b>B</b>) Point cloud with stem axis removed. (<b>C</b>) after noise removal.</p>
Full article ">Figure 5
<p>Schematic representation of the branches in the crown, projected onto a horizontal plane, where we assumed an ellipsoidal shape of the crown projection with the semi-axes centered at the stem axis and thus defining four quadrants (I–IV). The returns in the surface of the stem were manually removed.</p>
Full article ">Figure 6
<p>Schematic representation of the computation of the standardized composite histogram (SCH) from the point cloud in the horizontal plane, conducted for each quadrant separately and subsequently combined (see text for explanation of histograms (<b>A</b>–<b>E</b>)).</p>
Full article ">Figure 7
<p>Observed horizontal branch biomass distribution for the 23 destructively sampled beech trees (<b>grey lines</b>); the distribution of the average tree obtained using a non-parametric local regression (LOESS) with a smoothing span <span class="html-italic">α</span> = 0.75 (<b>red line</b>); and the fitted Max and Burkhart [<a href="#B35-remotesensing-13-01041" class="html-bibr">35</a>] segmented model (<b>black line</b>).</p>
Full article ">Figure 8
<p>Residual variability of horizontal branch biomass distribution per relative crown radius (<span class="html-italic">rcr</span>) class, where the dashed blue line corresponds to zero. The width of the radius classes is 0.05, and the midpoints are indicated by ticks on the <span class="html-italic">x</span>-axis. The same 20 classes of <span class="html-italic">rcr</span> have been considered as for <a href="#remotesensing-13-01041-f002" class="html-fig">Figure 2</a>, <a href="#remotesensing-13-01041-f007" class="html-fig">Figure 7</a> and <a href="#remotesensing-13-01041-f009" class="html-fig">Figure 9</a>.</p>
Full article ">Figure 9
<p>Standardized composite histogram (<span class="html-italic">SCH</span>) of the crown hits from TLS for the 17 scanned sample trees (grey lines). The red line represents the mean values obtained by LOESS.</p>
Full article ">Figure 10
<p>Comparison of the empirical horizontal branch biomass distribution (black) with (<b>A</b>) the original Standardized Composite Histogram obtained from Terrestrial Laser Scanning data (red) and (<b>B</b>) the Standardized Composite Histogram obtained by TLS data adjusted with the empirical data. The solid lines represent the mean local regression trends (smoothing span <span class="html-italic">α</span> = 0.75) and the shaded areas represent the 95% confidence intervals.</p>
Full article ">
30 pages, 13124 KiB  
Article
Applying Close Range Non-Destructive Techniques for the Detection of Conservation Problems in Rock-Carved Cultural Heritage Sites
by William Frodella, Mikheil Elashvili, Daniele Spizzichino, Giovanni Gigli, Akaki Nadaraia, Giorgi Kirkitadze, Luka Adikashvili, Claudio Margottini, Nikoloz Antidze and Nicola Casagli
Remote Sens. 2021, 13(5), 1040; https://doi.org/10.3390/rs13051040 - 9 Mar 2021
Cited by 17 | Viewed by 5062
Abstract
Rock-carved cultural heritage sites are often developed in slopes formed by weak rocks, which due to their peculiar lithological, geotechnical, and morpho-structural features are characterized by excellent carvability, which at the same time makes them prone to weathering, deterioration, and slope instability issues. [...] Read more.
Rock-carved cultural heritage sites are often developed in slopes formed by weak rocks, which due to their peculiar lithological, geotechnical, and morpho-structural features are characterized by excellent carvability, which at the same time makes them prone to weathering, deterioration, and slope instability issues. In this context the use of advanced close-range nondestructive techniques, such as Infrared Thermography (IRT) and Unmanned Aerial vehicle-based Digital Photogrammetry (UAV-DP) can be profitably used for the rapid detection of conservation issues (e.g., open fractures, unstable ledges-niches, water seepage and moisture) that can lead to slope instability phenomena. These techniques, when combined with traditional methods (e.g., field surveys, laboratory analysis), can provide fundamental data (such as 3D maps of the kinematic mechanisms) to implement a site-specific and interdisciplinary approach for the sustainable protection and conservation of such fragile cultural heritage sites. In this paper some examples of conservation problems in several rupestrian sites characterized by different geological contexts, from the mountainous regions of Georgia to the ancient city of Petra in Jordan, are presented, with the aim of evaluating the potential of the proposed integrated approach. The final aim is to provide conservators, practitioners, and local authorities with a useful, versatile, and low-cost methodology, to be profitably used in the protection and conservation strategies of rock-carved sites. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map with the location of the analyzed rock-carved cultural heritage sites (<b>a</b>); locations at a higher scale: Vanis Kvabebi (<b>b</b>); David Gareja area (<b>c</b>); Uplistsikhe (<b>d</b>); The Monastery of Petra (<b>e</b>).</p>
Full article ">Figure 2
<p>Vanis Kvabebi geological-geomorphological features: true color UAV-DP photogrammetric digital models: nadiral view (<b>a</b>); frontal view of the eastern main cliff of the monastery showing the outcropping lithologies of the Godertzi formation and the 2017 rock fall deposits (<b>b</b>); welded tuffs (dashed red line) (<b>c</b>); lower chapel with ancient rockfall boulder (<b>d</b>); upper chapel (<b>e</b>).</p>
Full article ">Figure 3
<p>The David Gareja area: structural monocline (<b>a</b>); outcropping lithologies (silts and clays with thick sand layer (<b>b</b>) coarse-grained sands-conglomerates (<b>c</b>); Lavra monastery (<b>d</b>); Examples of frescoed cave chapels in the monasteries of Udabno (<b>e</b>) and Sabereebi (<b>f</b>).</p>
Full article ">Figure 4
<p>The Uplistsikhe cave town complex: top of the rocky terrace with the brick church (<b>a</b>); examples of cave systems and defensive walls (<b>b</b>); the unstable rock cliff bordering the site (red dashed circles highlight April 2019 detachments, Areas 1–2, the black circle the one from July 2019, Area 3) (<b>c</b>); examples of sandstone wind weathering (corrasion) (<b>d</b>).</p>
Full article ">Figure 5
<p>The Monastery of Petra: rock mass on the right side of the monument showing the area of interest (red rectangle) and bedding of the sandstones (<b>a</b>); pictures of the area of interest during the surveys of April 2011 (<b>b</b>) and June 2014 (<b>c</b>).</p>
Full article ">Figure 6
<p>Structural setting and kinematic analysis at Vanis Kvabebi: identified sets (<b>a</b>); kinematic analysis for planar (<b>b</b>) and wedge (<b>c</b>) failures; 2017 rock fall deposits (<b>d</b>); subvertical fracture where creek channel erosion has concentrated (<b>e</b>); welded tuff portion affected by wedge failures (<b>f</b>).</p>
Full article ">Figure 7
<p>Vanis Kvabebi surface water run off model: (<b>a</b>) 14 cm resolution DSM with streams categorized by catchment area (ovals highlight details in b-e; stream legend: L = left cliff; F = front cliff; R = right cliff; CH = channels); (<b>b</b>) top-view of the lower chapel area; (<b>c</b>) Large niche created by erosional features of the J2 set Master Joint; (<b>d</b>) erosional channel in correspondence of J3 fracture; (<b>e</b>) stream gullies on the slope of the ravine opposite the monastery.</p>
Full article ">Figure 8
<p>Vanis Kvabebi IRT data: mosaicked surface temperature map of the main sector of the eastern cliff (<b>a</b>) (the white rectangle represents the Area of Interest 1 = AOI 1 shown in (<b>c</b>)); vertical surface temperature profile of the cliff (<b>b</b>) (Li1 in <b>a</b>); detail of the 2017 collapse sector (AOI 1) (<b>c</b>); surface temperature map of the cliff’s southern sector (<b>d</b>).</p>
Full article ">Figure 9
<p>Results of the 3D kinematic analysis. Wedge failure = 48% (<b>a</b>); planar failure = 31% (<b>b</b>); free fall = 31% (white ovals highlighting the detected unstable niches) (<b>c</b>); GKI up to 62% (<b>d</b>).</p>
Full article ">Figure 10
<p>Outcomes of the field surveys in the Natlismcemeni monastery: local geological features (<b>a</b>) and structural discontinuities (<b>b</b>). Main instability phenomena detected in the field surveys and corresponding 2D kinematic analysis: planar (<b>c</b>,<b>g</b>) and wedge (<b>d</b>,<b>h</b>) failures; topples (<b>e</b>,<b>i</b>). Open fracture (<b>f</b>).</p>
Full article ">Figure 11
<p>Outcomes of the IRT surveys of the Natlismcemeni monastery: IRT mosaicked image (2075 × 505 pixels with 7.1 cm pixel size) showing the detected AOI 2–3 (white rectangles) and the warm thermal anomalies (dashed white ovals) (<b>a</b>); AOI 2 (1160 × 480-pixel image size) (<b>b</b>) and AOI 3 (2120 × 590-pixel image size) (<b>c</b>), both having 1.3 cm spatial resolution; 3D kinematic analysis (GKI index up to 64%) (<b>d</b>).</p>
Full article ">Figure 12
<p>The Sabereebi monastery geological features: cliff face showing the chambers and the detached block in 2018 (<b>a</b>); stratigraphic sequence (<b>b</b>); the detachment sector in 2020, showing the collapsed pillar (<b>c</b>); 2D kinematic analysis: wedge failure mechanism (<b>d</b>); planar failure mechanism (<b>e</b>); direct toppling (<b>f</b>).</p>
Full article ">Figure 13
<p>Sabereebi IRT data: Mosaicked surface temperature map (950 × 480 pixels with 6.5 cm resolution (<b>a</b>); vertical and horizontal surface temperature profiles Li1 (<b>b</b>) and Li2 (<b>c</b>). Classified surface temperature maps in correspondence of AOI 4 (<b>d</b>) and AOI 5 (<b>e</b>).</p>
Full article ">Figure 14
<p>UAV-DP high resolution DEM derived products: slope map (<b>a</b>) (the monastery complex is located by the red rectangle); drainage pattern map (<b>b</b>) (the monastery complex is located by the yellow rectangle); UAV close-up on the detachment sector showing light brown colored mud drips on the dark brown sandstones (<b>c</b>).</p>
Full article ">Figure 15
<p>Structural data in the site of Uplistsikhe: main discontinuity sets (<b>a</b>); 2D kynematic analysis for planar (<b>b</b>) and wedge failures (<b>c</b>); 3D kynematic analysis (GKI index up to 49.9%) (<b>d</b>), showing the areas affected by rockfalls (Areas 1–2 in the red ovals), and the AOI highlighted in <a href="#remotesensing-13-01040-f016" class="html-fig">Figure 16</a> (black square).</p>
Full article ">Figure 16
<p>Outcomes of the IRT surveys: mosaicked thermograms (2800 × 850 pixel) (<b>a</b>) showing open fractures along J4 (white dashed ovals) and AOI 6 highlighted in b, d; detail of the cold thermal anomaly (<b>b</b>) with spots showing the warm thermal anomalies where the corrasion cavities develop (Bx = box plot showing min., max., avg. temperature); modelled ephemeral drainage network of the site (<b>c</b>) (red arrow highlights the ephemeral stream developing the cold thermal anomaly in <b>b</b>); correspondent optical image of b (<b>d</b>).</p>
Full article ">Figure 17
<p>Petra Monastery IRT data: mosaicked thermal image of the low-left corner of the monastery (680 × 240 pixels; corresponding optical image in <a href="#remotesensing-13-01040-f005" class="html-fig">Figure 5</a>c) (<b>a</b>); close-up single thermograms (320 × 240 pixels): AOI 5 (<b>b</b>–<b>d</b>) and AOI 6 (<b>e</b>–<b>h</b>).</p>
Full article ">Figure 18
<p>Vanis Kvabebi mitigation measures: hydrographic model (<b>a</b>) showing the watershed surfaces, the rock cliff areas (red rectangles), the area where the water drainage system (shown in <b>b</b>) is planned (dashed black square), and the Javakheti Plateau divide (dashed green line), (map of the planned water diverting wall (dashed red line) (<b>b</b>).</p>
Full article ">Figure 19
<p>Nonstructural measures, such as displacement monitoring systems, planned in David Gareja: Robotized Total Station (<b>a</b>); ground based interferometric synthetic aperture radar system (<b>b</b>). The scaffolding constructed for the anchoring phase in Uplistsikhe (<b>c</b>)<b>.</b></p>
Full article ">Figure 20
<p>Rainfall data in Petra between September 2010 and 2014 [<a href="#B90-remotesensing-13-01040" class="html-bibr">90</a>].</p>
Full article ">
22 pages, 5217 KiB  
Article
Improvement of Spatial Interpolation of Precipitation Distribution Using Cokriging Incorporating Rain-Gauge and Satellite (SMOS) Soil Moisture Data
by Bogusław Usowicz, Jerzy Lipiec, Mateusz Łukowski and Jan Słomiński
Remote Sens. 2021, 13(5), 1039; https://doi.org/10.3390/rs13051039 - 9 Mar 2021
Cited by 14 | Viewed by 3882
Abstract
Precipitation data provide a crucial input for examining hydrological issues, including watershed management and mitigation of the effects of floods, drought, and landslides. However, they are collected frequently from the scarce and often insufficient network of ground-based rain-gauge stations to generate continuous precipitation [...] Read more.
Precipitation data provide a crucial input for examining hydrological issues, including watershed management and mitigation of the effects of floods, drought, and landslides. However, they are collected frequently from the scarce and often insufficient network of ground-based rain-gauge stations to generate continuous precipitation maps. Recently, precipitation maps derived from satellite data have not been sufficiently linked to ground-based rain gauges and satellite-derived soil moisture to improve the assessment of precipitation distribution using spatial statistics. Kriging methods are used to enhance the estimation of the spatial distribution of precipitations. The aim of this study was to assess two geostatistical methods, ordinary kriging (OK) and ordinary cokriging (OCK), and one deterministic method (i.e., inverse distance weighting (IDW)) for improved spatial interpolation of quarterly and monthly precipitations in Poland and near-border areas of the neighbouring countries (~325,000 or 800,000 km2). Quarterly precipitation data collected during a 5-year period (2010–2014) from 113–116 rain-gauge stations located in the study area were used. Additionally, monthly precipitations in the years 2014–2017 from over 400 rain-gauge stations located in Poland were used. The spatiotemporal data on soil moisture (SM) from the Soil Moisture and Ocean Salinity (SMOS) global satellite (launched in 2009) were used as an auxiliary variable in addition to precipitation for the OCK method. The predictive performance of the spatial distribution of precipitations was the best for OCK for all quarters, as indicated by the coefficient of determination (R2 = 0.944–0.992), and was less efficient (R2 = 0.039–0.634) for the OK and IDW methods. As for monthly precipitation, the performance of OCK was considerably higher than that of IDW and OK, similarly as with quarterly precipitation. The performance of all interpolation methods was better for monthly than for quarterly precipitations. The study indicates that SMOS data can be a valuable source of auxiliary data in the cokriging and/or other multivariate methods for better estimation of the spatial distribution of precipitations in various regions of the world. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of rain-gauge stations in Poland and neighbouring countries and SMOS pixel (<b>A</b>). Map was created using Google Earth (v. 7.3.2.5776). Google, proprietary software, <a href="https://www.google.com/earth/" target="_blank">https://www.google.com/earth/</a>. The background maps from Google Maps (<a href="https://www.google.com/maps/@51.1367607,20.6545385,5.5z" target="_blank">https://www.google.com/maps/@51.1367607,20.6545385,5.5z</a>), accessed 12 April 2017, and <a href="https://pl.wikipedia.org/wiki/Plik:Poland_location_map_white.svg" target="_blank">https://pl.wikipedia.org/wiki/Plik:Poland_location_map_white.svg</a> (<b>B</b>). Elevation map of Poland, n.p.m.—above sea level (<b>C</b>), <a href="https://pl.wikipedia.org/wiki/Mapa_hipsometryczna#/media/Plik:Poland-hipsometric_map.jpg" target="_blank">https://pl.wikipedia.org/wiki/Mapa_hipsometryczna#/media/Plik:Poland-hipsometric_map.jpg</a>. The background maps were modified using Microsoft Office PowerPoint 2019.</p>
Full article ">Figure 2
<p>Mean, minimal (Min), and maximal (Max) values for annual rainfall (<b>A</b>) and soil moisture data (<b>B</b>) with standard deviations in Poland and neighbouring countries for the study period 2010–2014 and for monthly rainfall (<b>C</b>) and soil moisture data (<b>D</b>) in Poland for the study period 2014–2017.</p>
Full article ">Figure 2 Cont.
<p>Mean, minimal (Min), and maximal (Max) values for annual rainfall (<b>A</b>) and soil moisture data (<b>B</b>) with standard deviations in Poland and neighbouring countries for the study period 2010–2014 and for monthly rainfall (<b>C</b>) and soil moisture data (<b>D</b>) in Poland for the study period 2014–2017.</p>
Full article ">Figure 3
<p>Spatial distribution of monthly rainfall (2D maps) in Poland estimated by the ordinary cokriging interpolation method for selected months in the years 2014 and 2016 (1° = approximately 100 km). Maps were created using Gamma Design Software GS+10 [<a href="#B60-remotesensing-13-01039" class="html-bibr">60</a>].</p>
Full article ">Figure 3 Cont.
<p>Spatial distribution of monthly rainfall (2D maps) in Poland estimated by the ordinary cokriging interpolation method for selected months in the years 2014 and 2016 (1° = approximately 100 km). Maps were created using Gamma Design Software GS+10 [<a href="#B60-remotesensing-13-01039" class="html-bibr">60</a>].</p>
Full article ">
20 pages, 17023 KiB  
Article
The Effect of Forest Mask Quality in the Wall-to-Wall Estimation of Growing Stock Volume
by Elia Vangi, Giovanni D’Amico, Saverio Francini, Francesca Giannetti, Bruno Lasserre, Marco Marchetti, Ronald E. McRoberts and Gherardo Chirici
Remote Sens. 2021, 13(5), 1038; https://doi.org/10.3390/rs13051038 - 9 Mar 2021
Cited by 21 | Viewed by 3673
Abstract
Information about forest cover and its characteristics are essential in national and international forest inventories, monitoring programs, and reporting activities. Two of the most common forest variables needed to support sustainable forest management practices are forest cover area and growing stock volume (GSV [...] Read more.
Information about forest cover and its characteristics are essential in national and international forest inventories, monitoring programs, and reporting activities. Two of the most common forest variables needed to support sustainable forest management practices are forest cover area and growing stock volume (GSV m3 ha−1). Nowadays, national forest inventories (NFI) are complemented by wall-to-wall maps of forest variables which rely on models and auxiliary data. The spatially explicit prediction of GSV is useful for small-scale estimation by aggregating individual pixel predictions in a model-assisted framework. Spatial knowledge of the area of forest land is an essential prerequisite. This information is contained in a forest mask (FM). The number of FMs is increasing exponentially thanks to the wide availability of free auxiliary data, creating doubts about which is best-suited for specific purposes such as forest area and GSV estimation. We compared five FMs available for the entire area of Italy to examine their effects on the estimation of GSV and to clarify which product is best-suited for this purpose. The FMs considered were a mosaic of local forest maps produced by the Italian regional forest authorities; the FM produced from the Copernicus Land Monitoring System; the JAXA global FM; the hybrid global FM produced by Schepaschencko et al., and the FM estimated from the Corine Land Cover 2006. We used the five FMs to mask out non-forest pixels from a national wall-to-wall GSV map constructed using inventory and remotely sensed data. The accuracies of the FMs were first evaluated against an independent dataset of 1,202,818 NFI plots using four accuracy metrics. For each of the five masked GSV maps, the pixel-level predictions for the masked GSV map were used to calculate national and regional-level model-assisted estimates. The masked GSV maps were compared with respect to the coefficient of correlation (ρ) between the estimates of GSV they produced (both in terms of mean and total of GSV predictions within the national and regional boundaries) and the official NFI estimates. At the national and regional levels, the model-assisted GSV estimates based on the GSV map masked by the FM constructed as a mosaic of local forest maps were closest to the official NFI estimates with ρ = 0.986 and ρ = 0.972, for total and mean GSV, respectively. We found a negative correlation between the accuracies of the FMs and the differences between the model-assisted GSV estimates and the NFI estimate, demonstrating that the choice of the FM plays an important role in GSV estimation when using the model-assisted estimator. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The study area with the distribution of the national forest inventory (NFI) plots colored by growing stock volume (GSV) expressed in m<sup>3</sup> ha<sup>−1</sup>. On the right, a detail of the distribution of sample points used in the study within the NFI 1 x 1 km grid where the third-phase NFI plots (<a href="#sec2dot2dot1-remotesensing-13-01038" class="html-sec">Section 2.2.1</a>) are depicted in blue and the Inventario dell’Uso delle Terre in Italia (IUTI) points (<a href="#sec2dot2dot2-remotesensing-13-01038" class="html-sec">Section 2.2.2</a>) in white.</p>
Full article ">Figure 2
<p>Distribution of Landsat 7 ETM+ images per month, divided by acquisition years.</p>
Full article ">Figure 3
<p>Confusion matrices of each forest mask.</p>
Full article ">Figure 4
<p>Comparison of four accuracy metrics among the FMs, calculated at regional level (NUTS2).</p>
Full article ">Figure 5
<p>Growing stock map of Italy generated with random forests model. GSV in m<sup>3</sup> ha<sup>−</sup><sup>1</sup>. On the right, a detail of the GSV map masked with the five forest masks.</p>
Full article ">
26 pages, 7972 KiB  
Article
A New Cycle-Slip Repair Method for Dual-Frequency BDS Against the Disturbances of Severe Ionospheric Variations and Pseudoranges with Large Errors
by Dehai Li, Yaming Dang, Yunbin Yuan and Jinzhong Mi
Remote Sens. 2021, 13(5), 1037; https://doi.org/10.3390/rs13051037 - 9 Mar 2021
Cited by 8 | Viewed by 2071
Abstract
Many Beidou navigation satellite system (BDS) receivers or boards provide dual-frequency measurements to conduct precise positioning and navigation for low-power consumption. Cycle-slip processing is a primary work to guarantee consistent, precise positioning with the phase data. However, the cycle-slip processing of BDS dual-frequency [...] Read more.
Many Beidou navigation satellite system (BDS) receivers or boards provide dual-frequency measurements to conduct precise positioning and navigation for low-power consumption. Cycle-slip processing is a primary work to guarantee consistent, precise positioning with the phase data. However, the cycle-slip processing of BDS dual-frequency phases still follows with those of existing GPS methods. For single-satellite data, cycle-slip detection (CSD) with the geometry-free phase (GF) is disturbed by severe ionospheric delay variations, while CSD or cycle-slip repair (CSR) with the Melbourne–Wubbena combination (MW) must face the risk of the tremendous disturbance from large pseudorange errors. To overcome the above limitations, a new cycle-slip repair method for BDS dual-frequency phases (BDCSR) is proposed: (1) An optimal model to minimize the variance of the cycle-slip calculation was established to the dual-frequency BDS, after correcting the ionospheric variation with a reasonable and feasible way. (2) Under the BDS dual-frequency condition, a discrimination function was built to exclude the adverse disturbance from the pseudorange errors on the CSR, according to the rankings of the absolute epoch-difference GFs calculated by the searched cycle-slip candidates after correcting the ionospheric variation. Subsequently, many compared CSR tests were implemented in conditions of low and medium elevations during strong geomagnetic storms. Comparisons from the results of different methods show that: (1) The variations of ionospheric delays are intolerable in the cycle-slip calculation during the geomagnetic storm, and the tremendous influence from the ionospheric variation should be corrected before calculating the cycle-slip combination with the BDS dual-frequency data. (2) Under the condition of real dual-frequency BDS data during the geomagnetic storm, the actual success rate of the conventional dual-frequency CSR (CDCSR) by employing the optimized combinations, but absenting from the discrimination function, is lower than that of BDCSR by about 2%; The actual success rate of the CSD with MW (MWCSD), is lower than that of BDCSR by about 2%. (3) After adding gross errors of 0.7 m to all real epoch-difference pseudoranges epoch-by-epoch, results of CDCSR and MWCSD showed many errors. However, BDCSR achieved a higher actual success rate than those of CDCSR and MWCSD, about 43% and 16%, respectively, and better performance of refraining the disturbance of large pseudorange error on the cycle-slip determination was achieved in the BDCSR methodology. Full article
(This article belongs to the Special Issue Positioning and Navigation in Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The first-order and second-order differences of the ionospheric delays on the high-dynamic medium earth orbit (MEO) Satellite 11 during the geomagnetic storm, which was calculated by the Beidou navigation satellite system (BDS) dual-frequency data, started from UTC 8 May 2016 00:00:00 with an interval of 30 s at the CUT0 station.</p>
Full article ">Figure 2
<p>The process flow of the cycle-slip repair (CSR) method for BDS dual-frequency phases (BDCSR).</p>
Full article ">Figure 3
<p>The cycle-slip combinations of WICSR without correcting the ionospheric variations and those of the BDCSR methodology after correcting the ionospheric variations, where the BDS dual-frequency pseudoranges and phases without cycle slip were collected from the GEO Satellite 2 during the geomagnetic storm, started from UTC 8 May 2016 00:00:00 with an interval of 30 s at the CUT0 station.</p>
Full article ">Figure 4
<p>The original cycle slips without correcting the ionospheric variations (WICSR) and those of the BDCSR methodology after correcting the ionospheric variations, where the BDS dual-frequency pseudoranges and phases without cycle slip were collected from the GEO Satellite 2 during the geomagnetic storm, started from UTC 8 May 2016 00:00:00 with an interval of 30 s at the CUT0 station.</p>
Full article ">Figure 5
<p>The cycle-slip combinations of WICSR without correcting the ionospheric variations and those of the BDCSR methodology after correcting the ionospheric variations, where the BDS dual-frequency pseudoranges and phases without cycle slip were collected from the MEO Satellite 11 during the geomagnetic storm, started from UTC 8 May 2016 00:00:00 with an interval of 30 s at the CUT0 station.</p>
Full article ">Figure 6
<p>The original cycle slips of WICSR without correcting the ionospheric variations and those of the BDCSR methodology after correcting the ionospheric variations, where the BDS dual-frequency pseudoranges and phases without cycle slip were collected from the MEO Satellite 11 during the geomagnetic storm, started from UTC 8 May 2016 00:00:00 with an interval of 30 s at the CUT0 station.</p>
Full article ">Figure 7
<p>The original cycle slips of WICSR without correcting the ionospheric variations and those of the BDCSR methodology after correcting the ionospheric variations, where the BDS dual-frequency pseudoranges and phases without cycle slip were collected from the IGSO Satellite 6 during the geomagnetic storm, started from UTC 8 May 2016 00:00:00 with an interval of 30 s at the CUT0 station.</p>
Full article ">Figure 8
<p>Results of CDCSR before selection and BDCSR after selection with the discrimination function from the GEO Satellite 2 under a fine condition with the medium elevations during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 9
<p>Results of CDCSR before selection and BDCSR after selection with the discrimination function from MEO Satellite 11 in a weak condition of low elevations during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 10
<p>Sequences of the MW cycle-slip combinations from MEO Satellite 11 in a weak condition of low elevation during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 11
<p>Results of CDCSR before selection and BDCSR after selection with the discrimination function from IGSO Satellite 6 in a weak condition of low elevation during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 12
<p>Results of conventional dual-frequency CSR (CDCSR) before selection and BDCSR after selection via the discrimination function in the case of appending gross errors of 0.7 m on all real epoch-difference pseudoranges from MEO Satellite 11 under a weak condition of low elevations during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 13
<p>Floating values of the BDCSR methodology’s cycle-slip combination in the case of appending gross errors of 0.7 m on all real epoch-difference pseudoranges from MEO Satellite 11 under a weak condition of low elevations during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 14
<p>Floating values of the Melbourne–Wubbena combination (MW)’s cycle-slip combination in the case of appending gross errors of 0.7 m on all real epoch-difference pseudoranges from MEO Satellite 11 under a weak condition of low elevation during the geomagnetic storm, where the phase data without cycle slip was collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 15
<p>Results of the cycle-slip combinations and original cycle slip supplied by the BDCSR methodology from MEO Satellite 11 in a weak condition of low elevations during the geomagnetic storm, where the known cycle slips of (1,0) were inserted epoch-by-epoch in the phase data collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 16
<p>Results of the cycle-slip combinations and original cycle slip supplied by the BDCSR methodology from MEO Satellite 11 in a weak condition of low elevations during the geomagnetic storm, where the known cycle slips of (1,1) were inserted epoch-by-epoch in the phase data collected at the CUT0 station started from UTC 8 May 2016 00:00:00 with an interval of 30 s.</p>
Full article ">Figure 17
<p>The differentials of geometry-free phases, provided by the unrepaired data and the repaired data with the BDCSR methodology and CDCSR in the case of normal pseudorange data and phase data with slips of (1,1) cycles on (B1,B2), where the observation data were collected by the CUT0 started from UTC 8 May 2016 00:00:00.</p>
Full article ">
20 pages, 4222 KiB  
Article
UAV-Based Estimate of Snow Cover Dynamics: Optimizing Semi-Arid Forest Structure for Snow Persistence
by Adam Belmonte, Temuulen Sankey, Joel Biederman, John Bradford, Scott Goetz and Thomas Kolb
Remote Sens. 2021, 13(5), 1036; https://doi.org/10.3390/rs13051036 - 9 Mar 2021
Cited by 11 | Viewed by 3865
Abstract
Seasonal snow cover in the dry forests of the American West provides essential water resources to both human and natural systems. The structure of trees and their arrangement across the landscape are important drivers of snow cover distribution across these forests, varying widely [...] Read more.
Seasonal snow cover in the dry forests of the American West provides essential water resources to both human and natural systems. The structure of trees and their arrangement across the landscape are important drivers of snow cover distribution across these forests, varying widely in both space and time. We used unmanned aerial vehicle (UAV) multispectral imagery and Structure-from-Motion (SfM) models to quantify rapidly melting snow cover dynamics and examine the effects of forest structure shading on persistent snow cover in a recently thinned ponderosa pine forest. Using repeat UAV multispectral imagery (n = 11 dates) across the 76 ha forest, we first developed a rapid and effective method for identifying persistent snow cover with 90.2% overall accuracy. The SfM model correctly identified 98% (n = 1280) of the trees, when compared with terrestrial laser scanner validation data. Using the SfM-derived forest structure variables, we then found that canopy shading associated with the vertical and horizontal metrics was a significant driver of persistent snow cover patches (R2 = 0.70). The results indicate that UAV image-derived forest structure metrics can be used to accurately predict snow patch size and persistence. Our results provide insight into the importance of forest structure, specifically canopy shading, in the amount and distribution of persistent seasonal snow cover in a typical dry forest environment. An operational understanding of forest structure effects on snow cover will help drive forest management that can target snow cover dynamics in addition to forest health. Full article
(This article belongs to the Special Issue Ecohydrological Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>An overview of the study site highlighting the locations and extents of the UAV Structure-from-Motion (SfM) data, terrestrial laser scanner (TLS) data, and field-based validation data. (<b>A)</b> shows an example of the field-measured and TLS validated plots. (<b>B)</b> shows the distribution of all field-measured and TLS validated plots within the SfM data extent, which includes both thinned and unthinned portions of the forest. A majority of our ground-based measurements are distributed across the larger, thinned portion of the study site. (C) shows the UAV and TLS instruments used in this study.</p>
Full article ">Figure 2
<p>UAV multispectral image analysis workflow. First, the original multispectral bands and the calculated indices were stacked into a single raster image, one for each image date (n = 11) and classified using a supervised random forest model to generate binary snow/non-snow classes. From the binary classification from each image date, the pixels with 10 or 11 snow-covered days (out of 11) were grouped to create the boundaries of individual persistent snow patches.</p>
Full article ">Figure 3
<p>Calculating the individual tree metrics for the final analysis involved using both the field-measured and terrestrial laser scanner (TLS) point cloud-derived datasets. Individual trees detected in the TLS point cloud were compared to field-measured trees for an omission and commission analysis. Their crown height (CH), average crown diameter (CD), and crown base heights (CBH) were compared to assess the accuracy of the TLS point cloud-derived estimates. The same comparison was made between matching trees in the TLS point cloud and Structure-from-Motion (SfM)-derived point cloud datasets, with the only difference being the addition of crown volume (CV).</p>
Full article ">Figure 4
<p>A persistent snow patch and the trees influencing it at the three different ranges: 1, 1.5, and 2 times the height (CH) of the tree. A tree was included within each respective tree shading influence area (TSIA), if its CH multiplied by 1, 1.5, or 2 was extended into the snow patch and if the bearing of its location (XY) to the patch was within the range of daily solar angles.</p>
Full article ">Figure 5
<p>The percent of the treated and untreated regions of the study site that are classified SCA for each UAV snow-series image following a storm. This data is grouped by storm, and partitioned by forest treatment condition to illustrate the reductions in SCA within and across the storms. SCA patterns following storm events show a greater SCA reduction in the treated portion of the study site than in the untreated portion. Evident also are the relatively similar initial amounts of classified SCA on the first image date after a storm in the treated portion, contrasting the initial amounts in the untreated portion.</p>
Full article ">Figure 6
<p>Each UAV image (n = 11) was classified into snow/non-snow pixels (<b>A</b>) and the resulting rasters were assembled into a single composite, in which persistent snow patches were identified (<b>B</b>). The orthomosaic image used as the base imagery is from storm 2, day 3 (1 May 2019) and its binary snow classification (<b>A</b>) was used for the classification accuracy assessment because it contained a nearly even snow/non-snow (49/51%) area distribution.</p>
Full article ">Figure 7
<p>Variable importance scores calculated for the forest structure metric predictor variables used in fitting the final MARS model framework. The variable importance score ranges from 0 to 100, with higher scores indicating that the predictor was more influential in model construction (n = 500). The chart depicts the frequency with which each forest structure metric was within the respective range of the importance score. The data are partitioned by tree shading influence area (TSIA) size to better illustrate the interaction with forest structure metric. Overall, the most influential forest structure metric predictors are canopy cover (CC), mean solar radiation footprint (SRF) value, and mean trees per hectare (TPH).</p>
Full article ">Figure 8
<p>A collection of the simple linear relationships reflecting each model’s predictive ability for the TSIA 1.5. Each line represents the prediction accuracy of a single model, and the red points are the value pairs for the predicted vs. observed persistent snow patch sizes (m<sup>2</sup>). The mean performance statistics for the entire set of prediction models in TSIA 1.5 is R<sup>2</sup> = 0.70 and a RMSE = 267 m<sup>2</sup>.</p>
Full article ">Figure 9
<p>A comparison of trees and their shadows with different vertical structure metrics, namely CBH:CH and CV. The perceived difference in forest canopy shading between trees with high CBH:CH and low CV values (Panel <b>A)</b> are compared to those trees with overall larger crowns (Panel <b>B)</b>. These canopy shading differences were expected to result in vertical forest structure metrics being significant in the modeling of persistent snow patch size.</p>
Full article ">
20 pages, 5290 KiB  
Article
Remote Soil Moisture Measurement from Drone-Borne Reflectance Spectroscopy: Applications to Hydroperiod Measurement in Desert Playas
by Joseph S. Levy and Jessica T. E. Johnson
Remote Sens. 2021, 13(5), 1035; https://doi.org/10.3390/rs13051035 - 9 Mar 2021
Cited by 8 | Viewed by 4048
Abstract
The extent, timing, and magnitude of soil moisture in wetlands (the hydropattern) is a primary physical control on biogeochemical processes in desert environments. However, determining playa hydropatterns is challenged by the remoteness of desert basin sites and by the difficulty in determining soil [...] Read more.
The extent, timing, and magnitude of soil moisture in wetlands (the hydropattern) is a primary physical control on biogeochemical processes in desert environments. However, determining playa hydropatterns is challenged by the remoteness of desert basin sites and by the difficulty in determining soil moisture from remotely sensed data at fine spatial and temporal scales (hundreds of meters to kilometers, and hours to days). Therefore, we developed a new, reflectance-based soil moisture index (continuum-removed water index, or CRWI) that can be determined via hyperspectral imaging from drone-borne platforms. We compared its efficacy at remotely determining soil moisture content to existing hyperspectral and multispectral soil moisture indices. CRWI varies linearly with in situ soil moisture content (R2 = 0.89, p < 0.001) and is comparatively insensitive to soil clay content (R2 = 0.4, p = 0.01), soil salinity (R2 = 0.82, p < 0.001), and soil grain size distribution (R2 = 0.67, p < 0.001). CRWI is negatively correlated with clay content, indicating it is not sensitive to hydrated mineral absorption features. CRWI has stronger correlation with surface soil moisture than other hyperspectral and multispectral indices (R2 = 0.69, p < 0.001 for WISOIL at this site). Drone-borne reflectance measurements allow monitoring of soil moisture conditions at the Alvord Desert playa test site over hectare-scale soil plots at measurement cadences of minutes to hours. CRWI measurements can be used to determine surface soil moisture at a range of desert sites to inform management decisions and to better reveal ecosystem processes in water-limited environments. Full article
(This article belongs to the Special Issue Application of Hyperspectral Data in Ecological Environment)
Show Figures

Figure 1

Figure 1
<p>Context map of the Alvord Desert and Alvord Hot Springs (AHS). (<b>a</b>) Regional setting showing the full Alvord Desert playa and context location map in Oregon, USA. Portion of Landsat 8 image LC08_L1TP_043030_20190816_20190902. (<b>b</b>) Overview of the AHS discharge plume. Planet image 20190814_182056. (<b>c</b>) Overview of the AHS plume sampling site. Green points denote ground sampling transect points. Orthoimage produced from drone-derived color imagery.</p>
Full article ">Figure 2
<p>Comparison between ground spectra and airborne reflectance of plume soils measured using the same spectrometer. Note, ground spectra were collected at ~10:00 local time; airborne spectra were collected 11:37 local time.</p>
Full article ">Figure 3
<p>Schematic illustration of the drone-borne spectrometer system. Drone and camera system in flight.</p>
Full article ">Figure 4
<p>Ground truth data summary for the two sampling transects crossing the AHS plume from dry soil north of the plume, through the plume thalweg, to the south margin. Net GWC is net gravimetric water content (soil water mass per gram of soil), VWC is volumetric water content (cm<sup>3</sup> of water per cm<sup>3</sup> of soil) measured via 5TE probe. Extract EC is soil electrical conductivity converted to ppm (parts per million) solutes. CRWI and WISOIL are two reflectance spectroscopic soil moisture indices (see text).</p>
Full article ">Figure 5
<p>Sediment grain size distribution for ground control samples at the Alvord Hot Springs plume discharge site.</p>
Full article ">Figure 6
<p>Calibration curve for relating ground truth soil moisture content to ground truth continuum removed water index (CRWI). <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Relationship between continuum removed water index (CRWI) and ground control sample clay content. <span class="html-italic">p</span> = 0.01.</p>
Full article ">Figure 8
<p>Continuum removed water index (CRWI) versus residual water content. Samples that have enhanced residual water content after air drying showed lower CRWI in the field. <span class="html-italic">p</span> &lt; 0.001</p>
Full article ">Figure 9
<p>Continuum removed water index (CRWI) versus soil extract TDS. More saline samples have lower CRWI measured in the field. <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>Ground control sampling point water content versus continuum removed water index (CRWI). In this formulation, soil net GWC can be predicted using drone-derived CRWI. Error bars show 95% confidence interval for GWC determined via CRWI analysis.</p>
Full article ">Figure 11
<p>CRWI-derived soil moisture for each of the three completed overflights. Color composite panel shows just the color ortho-image of the Alvord hot spring plume. Panels AHS 4, 5, and 7 show gridded data for three drone sorties.</p>
Full article ">Figure 12
<p>Ground control sampling point water content versus ground-measured WISOIL index. <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 13
<p>Ground control point net gravimetric water content (GWC) versus CRWI-derived GWC at the ground sampling points. Linear fit show is for AHS 7, R<sup>2</sup> = 0.42, <span class="html-italic">p</span> = 0.017.</p>
Full article ">Figure 14
<p>(<b>a</b>) Scale-location plot and (<b>b</b>) QQ-plot for the CRWI-net GWC predictive model.</p>
Full article ">Figure 15
<p>CRWI-derived net GWC values versus ground surface elevation at AHS5 measurement points. Soil moisture is highest at the plume center at the upslope end of the plume, and decreases off the main plume axis and downslope.</p>
Full article ">
18 pages, 2878 KiB  
Article
Dynamics of Vibrio cholerae in a Typical Tropical Lake and Estuarine System: Potential of Remote Sensing for Risk Mapping
by Abdulaziz Anas, Kiran Krishna, Syamkumar Vijayakumar, Grinson George, Nandini Menon, Gemma Kulk, Jasmin Chekidhenkuzhiyil, Angelo Ciambelli, Hridya Kuttiyilmemuriyil Vikraman, Balu Tharakan, Abdul Jaleel Koovapurath Useph, Elizabeth Goult, Jithin Vengalil, Trevor Platt and Shubha Sathyendranath
Remote Sens. 2021, 13(5), 1034; https://doi.org/10.3390/rs13051034 - 9 Mar 2021
Cited by 17 | Viewed by 5012
Abstract
Vibrio cholerae, the bacterium responsible for the disease cholera, is a naturally-occurring bacterium, commonly found in many natural tropical water bodies. In the context of the U.N. Sustainable Development Goals (SDG) targets on health (Goal 3), water quality (Goal 6), life under [...] Read more.
Vibrio cholerae, the bacterium responsible for the disease cholera, is a naturally-occurring bacterium, commonly found in many natural tropical water bodies. In the context of the U.N. Sustainable Development Goals (SDG) targets on health (Goal 3), water quality (Goal 6), life under water (Goal 14), and clean water and sanitation (Goal 6), which aim to “ensure availability and sustainable management of water and sanitation for all”, we investigated the environmental reservoirs of V. cholerae in Vembanad Lake, the largest lake in Kerala (India), where cholera is endemic. The response of environmental reservoirs of V. cholerae to variability in essential climate variables may play a pivotal role in determining the quality of natural water resources, and whether they might be safe for human consumption or not. The hydrodynamics of Vembanad Lake, and the man-made barrier that divides the lake, resulted in spatial and temporal variability in salinity (1–32 psu) and temperature (23 to 36 °C). The higher ends of this salinity and temperature ranges fall outside the preferred growth conditions for V. cholerae reported in the literature. The bacteria were associated with filtered water as well as with phyto- and zooplankton in the lake. Their association with benthic organisms and sediments was poor to nil. The prevalence of high laminarinase and chitinase enzyme expression (more than 50 µgmL−1 min−1) among V. cholerae could underlie their high association with phyto- and zooplankton. Furthermore, the diversity in the phytoplankton community in the lake, with dominance of genera such as Skeletonema sp., Microcystis sp., Aulacoseira sp., and Anabaena sp., which changed with location and season, and associated changes in the zooplankton community, could also have affected the dynamics of the bacteria in the lake. The probability of presence or absence of V. cholerae could be expressed as a function of chlorophyll concentration in the water, which suggests that risk maps for the entire lake can be generated using satellite-derived chlorophyll data. In situ observations and satellite-based extrapolations suggest that the risks from environmental V. cholerae in the lake can be quite high (with probability in the range of 0.5 to 1) everywhere in the lake, but higher values are encountered more frequently in the southern part of the lake. Remote sensing has an important role to play in meeting SDG goals related to health, water quality and life under water, as demonstrated in this example related to cholera. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area showing the sampling locations in the Vembanad lake. Stations are ranked based on the number of phytoplankton samples in which environmental <span class="html-italic">Vibrio cholerae</span> was detected, relative to the total number of samples examined, with extreme risk (black) corresponding to 80 to 100% of samples being positive, high risk (red) with 60–80% of samples being positive and seasonally risky (blue) in which 40–60% of samples showed the presence of <span class="html-italic">V. cholerae</span>. Locations of Kochi city, industrial areas and tourism spots mentioned in the main text are also labelled.</p>
Full article ">Figure 2
<p>(<b>A</b>) Spatio-temporal variation of salinity in the brackish-water (stations 1–6) and fresh-water (stations 7–13) regions of the Vembanad Lake; and (<b>B</b>) the growth of pathogenic isolate of <span class="html-italic">V. cholerae</span> at different salinities.</p>
Full article ">Figure 3
<p>In situ observations in the 13 stations (blue) and the lake-averaged satellite (red) estimate of the temporal variation of temperature in the Vembanad Lake during the study.</p>
Full article ">Figure 4
<p>Distribution of <span class="html-italic">Vibrio cholerae</span> in sediment (red), macrobenthos (purple), micro-zooplankton (ash), micro-phytoplankton (green) and filtered water (blue) samples during wet (<b>A</b>) and dry (<b>B</b>) seasons in the Vembanad Lake.</p>
Full article ">Figure 5
<p>Spatio-temporal variation of chlorophyll in the Vembanad Lake measured using (<b>A</b>) in situ and (<b>B</b>) satellite remote sensing. In (<b>A</b>), the lower half of the figure (stations 1–6) constitute the brackish-water region of the lake, and the upper half (stations 7–13) constitute the fresh-water region of the lake. In (<b>B</b>), dark grey represents missing data. Land is delineated as light grey and ocean as light blue.</p>
Full article ">Figure 6
<p>The relationship between satellite chlorophyll and probability of finding environmental <span class="html-italic">Vibrio cholerae</span> in the (<b>A</b>) brackish water (BW) and (<b>B</b>) fresh water (FW) regions in the Vembanad Lake. Note that a number of observations are used to generate each of the points in the graph, and the number of observations used are indicated against each point. (<b>C</b>) Representative satellite image (for 23 February 2019) of chlorophyll and (<b>D</b>) the corresponding risk map of environmental <span class="html-italic">V. cholerae</span> incidence in the lake, generated from the relationships shown in (<b>A</b>,<b>B</b>) for the two parts of the lake.</p>
Full article ">Figure 7
<p>Dominant groups of micro-phytoplankton in the brackish water (BW) and freshwater (FW) regions of the Vembanad Lake during dry and wet seasons.</p>
Full article ">
18 pages, 1723 KiB  
Article
Evaluating Multi-Sensors Spectral and Spatial Resolutions for Tree Species Diversity Prediction
by Enoch Gyamfi-Ampadu, Michael Gebreslasie and Alma Mendoza-Ponce
Remote Sens. 2021, 13(5), 1033; https://doi.org/10.3390/rs13051033 - 9 Mar 2021
Cited by 11 | Viewed by 3964
Abstract
Forests contribute significantly to terrestrial biodiversity conservation. Monitoring of tree species diversity is vital due to climate change factors. Remote sensing imagery is a means of data collection for predicting diversity of tree species. Since various sensors have different spectral and spatial resolutions, [...] Read more.
Forests contribute significantly to terrestrial biodiversity conservation. Monitoring of tree species diversity is vital due to climate change factors. Remote sensing imagery is a means of data collection for predicting diversity of tree species. Since various sensors have different spectral and spatial resolutions, it is worth comparing them to ascertain which could influence the accuracy of prediction of tree species diversity. Hence, this study evaluated the influence of the spectral and spatial resolutions of PlanetScope, RapidEye, Sentinel 2 and Landsat 8 images in diversity prediction based on the Shannon diversity index (H′), Simpson diversity Index (D1) and Species richness (S). The Random Forest regression was applied for the prediction using the spectral bands of the sensors as variables. The Sentinel 2 was the best image, producing the highest coefficient of determination (R2) under both the Shannon Index (R2 = 0.926) and the Species richness (R2 = 0.923). Both the Sentinel and RapidEye produced comparable higher accuracy for the Simpson Index (R2 = 0.917 and R2 = 0.915, respectively). The PlanetScope was the second-accurate for the Species richness (R2 = 0.90), whiles the Landsat 8 was the least accurate for the three diversity indices. The outcomes of this study suggest that both the spectral and spatial resolutions influence prediction accuracies of satellite imagery. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the study area. Note: A is the Nkandla forest reserve, B is the map of South Africa indicating the location of KZN province and the forest and C is the map of Africa indicating the location of South Africa.</p>
Full article ">Figure 2
<p>Scatter plot for the Shannon Index prediction. (<b>A</b>) is for Sentinel 2, (<b>B</b>) is for RapidEye, (<b>C</b>) is for PlanetScope, and (<b>D</b>) is for Landsat 8. The blue line is the line of best fit and the dashed line is the 1:1 line as shown on the individual plots.</p>
Full article ">Figure 3
<p>Scatter plot for the Simpson Index prediction. (<b>A</b>) is for Sentinel 2, (<b>B</b>) is for RapidEye, (<b>C</b>) is for PlanetScope, and (<b>D</b>) is for Landsat 8. The blue line is the line of best fit and the dashed line is the 1:1 line as shown on the individual plots.</p>
Full article ">Figure 4
<p>Scatter plot for the Species richness predictions. (<b>A</b>) is for Sentinel 2, (<b>B</b>) is for RapidEye, (<b>C</b>) is for PlanetScope, and (<b>D</b>) is for Landsat 8. The blue line is the line of best fit and the dashed line is the 1:1 line as shown on the individual plots.</p>
Full article ">
23 pages, 23408 KiB  
Article
The Decrease in Lake Numbers and Areas in Central Asia Investigated Using a Landsat-Derived Water Dataset
by Xianghong Che, Min Feng, Qing Sun, Joseph O. Sexton, Saurabh Channan and Jiping Liu
Remote Sens. 2021, 13(5), 1032; https://doi.org/10.3390/rs13051032 - 9 Mar 2021
Cited by 16 | Viewed by 3835
Abstract
Although Central Asia has a strong continental climate with a constant moisture deficit and low relative humidity, it is covered by thousands of lakes that are critical to the sustainability of ecosystems and human welfare in the region. Vulnerability to climate change and [...] Read more.
Although Central Asia has a strong continental climate with a constant moisture deficit and low relative humidity, it is covered by thousands of lakes that are critical to the sustainability of ecosystems and human welfare in the region. Vulnerability to climate change and anthropogenic activities have contributed to dramatic inter-annual and seasonal changes of the lakes. In this study, we explored the high spatio–temporal dynamics of the lakes of Central Asia using the terraPulse™ monthly Landsat-derived surface water extent dataset from 2000 to 2015 and the HydroLAKES dataset. The results identified 9493 lakes and significant linear decreasing trends were identified for both the number (rate: −85 lakes/year, R2: 0.69) and area (rate: −1314.1 km2/year, R2: 0.84) of the lakes in Central Asia between 2000 and 2015. The decrease rate in lake area accounted for 1.41% of the total lake area. About 75% of the investigated lakes (7142 lakes), mainly located in the Kazakh steppe (especially in the north) and the Badghyz and Karabil semi-desert terrestrial ecological zones, experienced a decrease in the water area. Lakes with increasing water area were mainly distributed in the Northern Tibetan Plateau–Kunlun Mountains alpine desert and Qaidam Basin semi-desert zones in the east-south corner of Central Asia. The possible driving factors of lake decreases in Central Asia were explored for the Aral Sea and Tengiz Lake on yearly and monthly time scales. The Aral Sea showed the greatest decrease in the summer months because of increased evaporation and massive irrigation, while the largest decrease for Tengiz Lake was observed in early spring and was linked to decreasing snowmelt. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The extent of Central Asia overlayed with a digital elevation model (DEM), glaciers, main rivers, and vectorized lakes from the HydroLAKES dataset [<a href="#B37-remotesensing-13-01032" class="html-bibr">37</a>].</p>
Full article ">Figure 2
<p>The graphical iteration processing for lake identification.</p>
Full article ">Figure 3
<p>Power-law frequency distribution of lake size in Central Asia.</p>
Full article ">Figure 4
<p>Number and area of lakes within seven terrestrial ecological zones in Central Asia [<a href="#B35-remotesensing-13-01032" class="html-bibr">35</a>].</p>
Full article ">Figure 5
<p>The change trend of the number of lakes in Central Asia.</p>
Full article ">Figure 6
<p>The yearly change (line) and intra-year fluctuation (blue bar) of the total lake area (the variables Max, Mean, and Min were calculated as the maximum, average, and minimum water area over the seven months within one year; the intra-annual fluctuation is the standard deviation of water areas over the seven months within one year).</p>
Full article ">Figure 7
<p>Spatial distribution of yearly change rate of each lake over seven terrestrial ecological zones in Central Asia as <a href="#remotesensing-13-01032-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>Water occurrence distribution of water bodies for the Aral Sea (<b>a</b>) and Tengiz Lake (<b>b</b>) during the 16-year study period.</p>
Full article ">Figure 9
<p>Yearly and intra-annual area changes for Aral Sea and Tengiz Lake from 2000 to 2015.</p>
Full article ">Figure 10
<p>Monthly dynamics of water bodies for the Aral Sea over 16-year study period.</p>
Full article ">Figure 11
<p>Monthly dynamics of water bodies for Tengiz Lake over the 16-year study period.</p>
Full article ">Figure 12
<p>Correlations between lake area and temperature (<b>a</b>), snowmelt (<b>b</b>), surface runoff (<b>c</b>), evaporation (<b>d</b>) and precipitation (<b>e</b>).</p>
Full article ">Figure 13
<p>Water level change of the North and South Aral Seas [<a href="#B61-remotesensing-13-01032" class="html-bibr">61</a>].</p>
Full article ">Figure 14
<p>Comparison between yearly lake area and yearly temperature and summer precipitation from 2000 to 2014 for Tengiz Lake Basin.</p>
Full article ">Figure 15
<p>Yearly variations in Snow Water Equivalent (SWE) for the months of January, February, and March and their linear trends for Tengiz Lake Basin.</p>
Full article ">
18 pages, 753 KiB  
Article
Towards a Topographically-Accurate Reflection Point Prediction Algorithm for Operational Spaceborne GNSS Reflectometry—Development and Verification
by Lucinda King, Martin Unwin, Jonathan Rawlinson, Raffaella Guida and Craig Underwood
Remote Sens. 2021, 13(5), 1031; https://doi.org/10.3390/rs13051031 - 9 Mar 2021
Cited by 4 | Viewed by 3354
Abstract
GNSS Reflectometry (GNSS-R), a method of remote sensing using the reflections from satellite navigation systems, was initially envisaged for ocean wind speed sensing. In recent times there has been significant interest in the use of GNSS-R for sensing land parameters such as soil [...] Read more.
GNSS Reflectometry (GNSS-R), a method of remote sensing using the reflections from satellite navigation systems, was initially envisaged for ocean wind speed sensing. In recent times there has been significant interest in the use of GNSS-R for sensing land parameters such as soil moisture, which has been identified as an Essential Climate Variable (ECV). Monitoring objectives for ECVs set by the Global Climate Observing System (GCOS) organisation include a reduction in data gaps from spaceborne sources. GNSS-R can be implemented on small, relatively cheap platforms and can enable the launch of constellations, thus reducing such data gaps in these important datasets. However in order to realise operational land sensing with GNSS-R, adaptations are required to existing instrumentation. Spaceborne GNSS-R requires the reflection points to be predicted in advance, and for land sensing this means the effect of topography must be considered. This paper presents an algorithm for on-board prediction of reflection points over the land, allowing generation of DDMs on-board as well as compression and calibration. The algorithm is tested using real satellite data from TechDemoSat-1 in a software receiver with on-board constraints being considered. Three different resolutions of Digital Elevation Model are compared. The algorithm is shown to perform better against the operational requirements of sensing land parameters than existing methods and is ready to proceed to flight testing. Full article
(This article belongs to the Special Issue Applications of GNSS Reflectometry for Earth Observation)
Show Figures

Figure 1

Figure 1
<p>Example DDMs from different Earth surfaces. (<b>a</b>) A typical DDM collected over the ocean, with the characteristic horseshoe shape showing the spreading of power over the surface. The peak power is located on the DDM cross-hairs as the specular point has been predicted with sufficient accuracy by the quasi-spherical model. (<b>b</b>) A DDM from a dataset over low-elevation mountains in South Sudan (approximate elevation 500 m). The cross-hairs represent the predicted reflection point, showing the effect of topography on the location of the peak reflected power in the DDM.</p>
Full article ">Figure 2
<p>Diagram with simplified geometry demonstrating the influence of topography on specular reflection point prediction. The transmitter is considered to be far enough removed (20,200 km altitude for GPS) that the incoming direction of the incident signals at two nearby points can be assumed parallel. This diagram assumes that the elevated surface is flat, which is not often the case, but is a simplification to allow the most basic form of the algorithm to be implemented (see <a href="#sec2-remotesensing-13-01031" class="html-sec">Section 2</a>). Reproduced from [<a href="#B10-remotesensing-13-01031" class="html-bibr">10</a>] with permission.</p>
Full article ">Figure 3
<p>These DDMs demonstrate the impact of topography on the DDM and the impact of compressing the DDM using windowing. The red box represents an 8 × 8 pixel window. In this case, without the topography algorithm the peak power data would be lost as it falls outside the window. The surface elevation where this DDM was collected was approximately 400 m. (<b>a</b>) Land DDM generated with quasi-spherical algorithm. (<b>b</b>) The same DDM generated with the topography algorithm.</p>
Full article ">Figure 4
<p>Diagram of the baseline version of the Topographically Accurate Reflection Point Prediction algorithm (TARPP v1).</p>
Full article ">Figure 5
<p>The results of Analysis 1a for an example PRN from each test dataset—plots of peak power offset from the centre of the DDM. (<b>a</b>) TT1, PRN 21; (<b>b</b>) TT2, PRN 15; (<b>c</b>) TT3, PRN 05; (<b>d</b>) TT4, PRN 20; (<b>e</b>) TT5, PRN 08; (<b>f</b>) TT6, PRN 11.</p>
Full article ">Figure 6
<p>Histograms, generated using the combined data of all DDMs from all six datasets, showing the offset of the peak power pixel from the centre of the DDM. The three DEM resolutions are shown and compared with the quasi-spherical model in (<b>a</b>–<b>d</b>) shows the three DEM resolutions directly compared to each other. All three resolutions have a similar performance and thus the histograms nearly overlay each other, and so appear green. (<b>a</b>) 1 km DEM; (<b>b</b>) 5 km DEM; (<b>c</b>) 10 km DEM; (<b>d</b>) DEM Comparison.</p>
Full article ">
16 pages, 4027 KiB  
Article
Sentinel-1 Polarimetry to Map Apple Orchard Damage after a Storm
by Samuele De Petris, Filippo Sarvia, Michele Gullino, Eufemia Tarantino and Enrico Borgogno-Mondino
Remote Sens. 2021, 13(5), 1030; https://doi.org/10.3390/rs13051030 - 9 Mar 2021
Cited by 20 | Viewed by 3878
Abstract
Climate change increases extreme whether events such as floods, hailstorms, or storms, which can affect agriculture, causing damages and economic loss within the agro-food sector. Optical remote sensing data have been successfully used in damage detections. Cloud conditions limit their potential, especially while [...] Read more.
Climate change increases extreme whether events such as floods, hailstorms, or storms, which can affect agriculture, causing damages and economic loss within the agro-food sector. Optical remote sensing data have been successfully used in damage detections. Cloud conditions limit their potential, especially while monitoring floods or storms that are usually related to cloudy situations. Conversely, data from the Polarimetric Synthetic Aperture Radar (PolSAR) are operational in all-weather conditions and are sensitive to the geometrical properties of crops. Apple orchards play a key role in the Italian agriculture sector, presenting a cultivation system that is very sensitive to high-wind events. In this work, the H-α-A polarimetric decomposition technique was adopted to map damaged apple orchards with reference to a stormy event that had occurred in the study area (NW Italy) on 12 August 2020. The results showed that damaged orchards have higher H (entropy) and α (alpha angle) values compared with undamaged ones taken as reference (Mann–Whitney one-tailed test U = 14,514, p < 0.001; U = 16604, p < 0.001 for H and α, respectively). By contrast, A (anisotropy) values were significantly lower for damaged orchards (Mann–Whitney one-tailed test U = 8616, p < 0.001). Based on this evidence, the authors generated a map of potentially storm-damaged orchards, assigning a probability value to each of them. This map is intended to support local funding restoration policies by insurance companies and local administrations. Full article
(This article belongs to the Special Issue Remote Sensing in Agriculture: State-of-the-Art)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>An apple orchard (cultivar “Gala”) with hail nets uprooted by the storm on 12 August 2020. At the bottom, many mature apples can be noted, suggesting the economic loss caused by the storm.</p>
Full article ">Figure 2
<p>Italian regions (light gray) and the Piemonte region (dark gray). (Red) The AOI includes the Saluzzo, Manta, Lagnasco, and Verzuolo municipalities (reference frame: WGS84 UTM32N).</p>
Full article ">Figure 3
<p>Parcels belonging to the training and test sets. Colors (see legend) define the state of the surveyed parcel (damaged/undamaged). Reference frame is WGS84 UTM 32N.</p>
Full article ">Figure 4
<p>The adopted workflow. All steps were managed in SNAP ESA v. 7.0.</p>
Full article ">Figure 5
<p>Boxplots of H-α-A distributions for UTFs and DTFs. The boxplot values are from bottom to top, respectively, 5th, 25th, 50th—cross is mean value—75th, and 95th percentiles. (<b>a</b>) Entropy pixel distribution; (<b>b</b>) alpha angle pixel distribution; (<b>c</b>) anisotropy pixel distribution.</p>
Full article ">Figure 6
<p>A CP map of apple orchards in the AOI (Reference frame: WGS84 UTM32N).</p>
Full article ">Figure 7
<p>(<b>a</b>) DM binary classification of the OM in the AOI (reference frame is WGS84 UTM 32N); (<b>b</b>) bar chart representing mean and 1 SEM of the CP for DTFs and UTFs.</p>
Full article ">Figure 8
<p>A sketch representing orchard condition before (<b>a</b>) and after (<b>b</b>) the storm. In (<b>a</b>) the pattern row/inter-row is well defined; (<b>b</b>) after the storm, apple tree uprooting occurred, altering the row/inter-row pattern, and crowns covering the ground increased volumetric scattering.</p>
Full article ">
11 pages, 2123 KiB  
Technical Note
The Temporal Variation of Optical Depth in the Candidate Landing Area of China’s Mars Mission (Tianwen-1)
by Zhencheng Tang, Jianjun Liu, Xing Wang, Xin Ren, Wei Yan and Wangli Chen
Remote Sens. 2021, 13(5), 1029; https://doi.org/10.3390/rs13051029 - 9 Mar 2021
Cited by 4 | Viewed by 2687
Abstract
The atmospheric dust is an important factor in the evolution of the Martian climate and has a major impact on the scientific exploration of the Martian lander or rover and its payload. This paper used remote sensing images to calculate atmospheric optical depth [...] Read more.
The atmospheric dust is an important factor in the evolution of the Martian climate and has a major impact on the scientific exploration of the Martian lander or rover and its payload. This paper used remote sensing images to calculate atmospheric optical depth that characterizes the spatial distribution of the atmospheric dust of Mars. The optical depth calculated by the images of the High Resolution Imaging Science Experiment (HiRISE) in the inspection area of the Spirit rover had a similar temporal variation to the optical depth directly measured by the Spirit rover from the sunlight decay. We also used the HiRISE images to acquire the seasonal variation of optical depths in the candidate landing area of China’s Mars Mission (Tianwen-1). The results have shown that the seasonal pattern of the optical depth in the candidate landing area is consistent with the dust storm sequences in this area. After Tianwen-1 enters the orbit around Mars, the images collected by the Moderate Resolution Imaging Camera (MoRIC), and the High Resolution Imaging Camera (HiRIC) can be used to study the atmospheric optical depth in the candidate landing area, providing reference for the safe landing and operation of the lander and rover. Full article
(This article belongs to the Special Issue Planetary Remote Sensing: Chang’E-4/5 and Mars Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The schematic diagram of the research area and the footprints of images. (<b>a</b>) The red polygon represents the candidate landing area of China’s Mars Mission (Tianwen-1) [<a href="#B24-remotesensing-13-01029" class="html-bibr">24</a>], the blue rectangle represents the study area (part of the candidate landing area), and the yellow rectangle represents the inspection area of the Spirit rover. The base map is global color mosaic of the Mars by Viking Orbiter [<a href="#B25-remotesensing-13-01029" class="html-bibr">25</a>]; (<b>b</b>) The footprints (purple rectangle) of images in the inspection area of the Spirit rover; (<b>c</b>) The footprints (purple rectangle) of images in the candidate landing area.</p>
Full article ">Figure 2
<p>The schematic diagram of the selection of the shadowed and the adjacent sunlit areas (The image is part of the HiRISE image ESP_013855_1650. The red line indicates shadowed areas, the yellow line indicates flat sunlit areas, and the green line indicates the sunlit areas with obvious slope).</p>
Full article ">Figure 3
<p>The distribution of optical depth at different solar longitudes in MY29 (the blue error bars represents the optical depth of the inspection area of the Spirit rover calculated by the shadow method, the red error bars represents the optical depth divided by the correction coefficient, the black line represents the optical depth measured by the data of the Spirit rover [<a href="#B16-remotesensing-13-01029" class="html-bibr">16</a>], and green dots represents the number of dust storm activity recorded in the Mars Dust Activity Database (MDAD) at 0°–30°S in MY29 [<a href="#B28-remotesensing-13-01029" class="html-bibr">28</a>]).</p>
Full article ">Figure 4
<p>The retrieved optical depth at different solar longitude (the blue error bars represents the optical depth in the candidate landing area calculated by the shadow method, red error bars represent the optical depth divided by the correction coefficient, and green dots represents the number of dust storm activity recorded in the Mars Dust Activity Database (MDAD) at 10°–40°N [<a href="#B28-remotesensing-13-01029" class="html-bibr">28</a>]).</p>
Full article ">
4 pages, 3528 KiB  
Reply
Reply to Comment on “Comparison of Cloud Cover Detection Algorithms on Sentinel–2 Images of the Amazon Tropical Forest”
by Alber Hamersson Sanchez, Michelle Cristina A. Picoli, Gilberto Camara, Pedro R. Andrade, Michel Eustaquio D. Chaves, Sarah Lechler, Anderson R. Soares, Rennan F. B. Marujo, Rolf Ezequiel O. Simões, Karine R. Ferreira and Gilberto R. Queiroz
Remote Sens. 2021, 13(5), 1028; https://doi.org/10.3390/rs13051028 - 9 Mar 2021
Viewed by 2252
Abstract
In their comments about our paper, the authors remark on two issues regarding our results relating to the MACCS-ATCOR Joint Algorithm (MAJA). The first relates to the sub-optimal performance of this algorithm under the conditions of our tests, while the second corresponds to [...] Read more.
In their comments about our paper, the authors remark on two issues regarding our results relating to the MACCS-ATCOR Joint Algorithm (MAJA). The first relates to the sub-optimal performance of this algorithm under the conditions of our tests, while the second corresponds to an error in our interpretation of MAJA’s bit mask. To answer the first issue, we acknowledge MAJA’s capacity to improve its performance as the number of images increases with time. However, in our paper, we used the images we had available at the time we wrote our paper. Regarding the second issue, we misread the MAJA’s bit mask and mistakenly labelled shadows as clouds. We regret our error and here we present the updated tables and images. We corrected our estimation and, consequently, there is an increment in MAJA’s accuracy in the detection of clouds and cloud shadows. However, these increments are not enough to change the conclusion of our original paper. Full article
Show Figures

Figure 1

Figure 1
<p>Updated Figure 3d [<a href="#B1-remotesensing-13-01028" class="html-bibr">1</a>]. Clouds detected on the Sentinel–2A image T19LFK of 7 May 2018. The color picture (<b>a</b>) uses bands 4, 3, and 2; (<b>b</b>) MAJA cloud mask.</p>
Full article ">
15 pages, 4355 KiB  
Technical Note
Measuring the Service Capacity of Public Facilities Based on a Dynamic Voronoi Diagram
by Haifu Cui, Liang Wu, Sheng Hu and Rujuan Lu
Remote Sens. 2021, 13(5), 1027; https://doi.org/10.3390/rs13051027 - 9 Mar 2021
Cited by 5 | Viewed by 3119
Abstract
The supply–demand relationship of urban public service facilities is the key to measuring a city’s service level and quality, and a balanced supply–demand relationship is an important indicator that reflects the optimal allocation of resources. To address the problem presented by the unbalanced [...] Read more.
The supply–demand relationship of urban public service facilities is the key to measuring a city’s service level and quality, and a balanced supply–demand relationship is an important indicator that reflects the optimal allocation of resources. To address the problem presented by the unbalanced distribution of educational resources, this paper proposes a dynamic Voronoi diagram algorithm with conditional constraints (CCDV). The CCDV method uses the Voronoi diagram to divide the plane so that the distance from any position in each polygon to the point is shorter than the distance from the polygon to the other points. In addition, it can overcome the disadvantage presented by the Voronoi diagram’s inability to use the nonspatial attributes of the point set to precisely constrain the boundary range; the CCDV method can dynamically plan and allocate according to the school’s capacity and the number of students in the coverage area to maintain a balance between supply and demand and achieve the optimal distribution effect. By taking the division of school districts in the Bao’an District, Shenzhen, as an example, the method is used to obtain a school district that matches the capacity of each school, and the relative error between supply and demand fluctuates only from −0.1~0.15. According to the spatial distribution relationship between schools and residential areas in the division results, the schools in the Bao’an District currently have an unbalanced distribution in some areas. A comparison with the existing school district division results shows that the school district division method proposed in this paper has advantages. Through a comprehensive analysis of the accessibility of public facilities and of the balance of supply and demand, it is shown that school districts based on the CCDV method can provide a reference for the optimal layout of schools and school districts. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flow chart of the CCDV method.</p>
Full article ">Figure 2
<p>Schematic diagram of the CCDV method for iterative division of school districts.</p>
Full article ">Figure 3
<p>Distribution of elementary schools and the school-age population in the existing school districts in the Bao’an District, Shenzhen.</p>
Full article ">Figure 4
<p>School district division results.</p>
Full article ">Figure 5
<p>Correspondence between schools and school-age children in the community unit.</p>
Full article ">Figure 6
<p>The relative error of school supply and demand in the school districts based on different methods.</p>
Full article ">
16 pages, 17744 KiB  
Article
Meridional Changes in Satellite Chlorophyll and Fluorescence in Optically-Complex Coastal Waters of Northern Patagonia
by Sebastián I. Vásquez, María Belén de la Torre, Gonzalo S. Saldías and Aldo Montecinos
Remote Sens. 2021, 13(5), 1026; https://doi.org/10.3390/rs13051026 - 9 Mar 2021
Cited by 5 | Viewed by 2754
Abstract
Northern Patagonia is one of the largest estuarine systems worldwide. It is characterized by complex geography, including islands, peninsulas, channels, and fjords. Here, the Inner Sea of Chiloé (ISC) is the largest estuarine system extending about 230 km in the meridional direction. Phytoplankton’s [...] Read more.
Northern Patagonia is one of the largest estuarine systems worldwide. It is characterized by complex geography, including islands, peninsulas, channels, and fjords. Here, the Inner Sea of Chiloé (ISC) is the largest estuarine system extending about 230 km in the meridional direction. Phytoplankton’s long-term dynamics and the main physical drivers of their variability are not well understood yet. Time-space fluctuations of Chlorophyll-a (Chl-a) and Chlorophyll fluorescence (nFLH) within the ISC and their association with meteorological and oceanographic processes were analyzed using high resolution (1000 m) satellite data (2003–2019). Our results revealed a meridional Chl-a and nFLH gradient along the ISC, with higher concentrations north of the Desertores islands where the topography promotes a semi-closed system with estuarine characteristics yearlong. Satellite Chl-a and nFLH were characterized by asynchronous seasonal cycles (nFLH peaks in fall) that differed from the southern ISC where the maximum Chl-a and nFLH occurs in spring-summer. The adjacent coastal ocean influences the southern ISC, and thus, the Chl-a and nFLH variability correlated well with the seasonal variation of meridional winds. The northern ISC was clearly influenced by river discharges, which can bias the Chl-a retrievals, decoupling the annual cycles of Chl-a and nFLH. In situ data from a buoy in Seno Reloncaví reaffirmed this bias in satellite Chl-a and a higher correlation with nFLH, by which the construction of a local Chl-a algorithm for northern Patagonia is essential. Full article
(This article belongs to the Special Issue Remote Sensing of Inland Waters and Their Catchments)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Map of northern Patagonia. The Inner Sea of Chiloé (ISC) is enclosed in a red box and shown in (<b>b</b>,<b>c</b>). Black circles correspond to the location of oceanographic profiles shown in Figure 8. PR corresponds to the location of the Puelo River. The blue circle corresponds to the location of the oceanographic buoy in Seno Reloncavi. The average values of Moderate Resolution Imaging Spectroradiometer (MODIS) Chlorophyll-a (Chl-a) and Chlorophyll fluorescence (nFLH) (2003–2019) are color-coded in (<b>a</b>–<b>c</b>), respectively.</p>
Full article ">Figure 2
<p>Seasonal climatologies of (upper panels) CCMP satellite wind vectors and (lower panels) AVISO satellite geostrophic velocities (2003–2019) for (<b>a</b>,<b>e</b>) spring (October, November, December), (<b>b</b>,<b>f</b>) summer (January, February, March), (<b>c</b>,<b>g</b>) fall (April, May, June), and (<b>d</b>,<b>h</b>) winter (July, August, September). The colors in (<b>a</b>–<b>d</b>) represent the wind speed. For reference of locations, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>a.</p>
Full article ">Figure 3
<p>Seasonal climatologies of (upper panels) Chl-a and (lower panels) nFLH computed from all daily MODIS-Aqua images (2003–2019) for (<b>a</b>,<b>e</b>) spring (October, November, December), (<b>b</b>,<b>f</b>) summer (January, February, March), (<b>c</b>,<b>g</b>) fall (April, May, Jun), and (<b>d</b>,<b>h</b>) winter (July, August, September). For reference of locations, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>a.</p>
Full article ">Figure 4
<p>Temporal variability of Chl-a (blue curves) and nFLH (red curves) at 4 locations (R1–R4, shown as blue boxes in (<b>a</b>) along the Inner Sea of Chiloé (ISC). The interannual time series are presented in (<b>b</b>), whereas the corresponding annual cycles are shown in (<b>c</b>); r-values correspond to linear correlation coefficients. The average Chl-a is color-coded in (<b>a</b>). Notice that the annual cycles start from June in (<b>c</b>). For reference of locations, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>a.</p>
Full article ">Figure 5
<p>(<b>a</b>) Local fractional variance (LFV) spectra for Chl-a (blue curves) and nFLH (red curves). The values on the right hand side y-axis shows the confidence levels (%). Original time series (black lines) and reconstructed annual cycle (blue and red lines) for Chl-a and nFLH are shown in (<b>b</b>,<b>c</b>), respectively, for the location specified with a green asterisk in (<b>d</b>,<b>f</b>). The spatial field of explained annual variance and the corresponding phase is presented for Chl-a (<b>d</b>,<b>e</b>) and nFLH (<b>f</b>,<b>g</b>). For reference of locations, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>a.</p>
Full article ">Figure 6
<p>Time series of mean (<b>a</b>) meridional wind stress and (<b>b</b>) Chl-a and nFLH at Boca del Guafo (region between 73<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>–74<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>W and 43<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>–44<math display="inline"><semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics></math>S). (<b>c</b>) Chl-a and nFLH concentration as function of meridional wind stress. Spatial fields of linear correlation coefficient between meridional wind stress at Boca del Guafo and Chl-a (<b>d</b>–<b>f</b>) and nFLH (<b>e</b>–<b>g</b>). The Inner Sea of Chiloé (ISC) is enclosed in a blue box and shown in (<b>f</b>,<b>g</b>) for better visualization. For reference of locations, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>a.</p>
Full article ">Figure 7
<p>Comparison between near-surface (1 m) In situ fluorescence (7-day averages) and (<b>a</b>,<b>d</b>) nFLH and (<b>b</b>,<b>e</b>) Chl-a. r-values correspond to linear correlation coefficients. Puelo river discharge and its annual cycle is presented in (<b>c</b>,<b>f</b>), respectively. For reference of the buoy location, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>a.</p>
Full article ">Figure 8
<p>Mean vertical profiles of: (<b>a</b>,<b>f</b>) Brunt-Väisälä frequency, (<b>b</b>,<b>g</b>) Salinity, (<b>c</b>,<b>h</b>) phosphate, (<b>d</b>,<b>i</b>) nitrate, and (<b>e</b>,<b>j</b>) silicate. Top (bottom) panels correspond to profiles north (south) of Desertores Islands (see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>) for the winter of 2004 (blue) and the winter of 2006 (red). For reference of the location of the oceanographic stations, see <a href="#remotesensing-13-01026-f001" class="html-fig">Figure 1</a>b,c.</p>
Full article ">
23 pages, 2945 KiB  
Article
Bio-Inspired Hybridization of Artificial Neural Networks: An Application for Mapping the Spatial Distribution of Soil Texture Fractions
by Ruhollah Taghizadeh-Mehrjardi, Mostafa Emadi, Ali Cherati, Brandon Heung, Amir Mosavi and Thomas Scholten
Remote Sens. 2021, 13(5), 1025; https://doi.org/10.3390/rs13051025 - 8 Mar 2021
Cited by 40 | Viewed by 4052
Abstract
Soil texture and particle size fractions (PSFs) are a critical characteristic of soil that influences most physical, chemical, and biological properties of soil; furthermore, reliable spatial predictions of PSFs are crucial for agro-ecological modeling. Here, series of hybridized artificial neural network (ANN) models [...] Read more.
Soil texture and particle size fractions (PSFs) are a critical characteristic of soil that influences most physical, chemical, and biological properties of soil; furthermore, reliable spatial predictions of PSFs are crucial for agro-ecological modeling. Here, series of hybridized artificial neural network (ANN) models with bio-inspired metaheuristic optimization algorithms such as a genetic algorithm (GA-ANN), particle swarm optimization (PSO-ANN), bat (BAT-ANN), and monarch butterfly optimization (MBO-ANN) algorithms, were built for predicting PSFs for the Mazandaran Province of northern Iran. In total, 1595 composite surficial soil samples were collected, and 64 environmental covariates derived from terrain, climatic, remotely sensed, and categorical datasets were used as predictors. Models were tested using a repeated 10-fold nested cross-validation approach. The results indicate that the hybridized ANN methods were far superior to the reference approach using ANN with a backpropagation training algorithm (BP-ANN). Furthermore, the MBO-ANN approach was consistently determined to be the best approach and yielded the lowest error and uncertainty. The MBO-ANN model improved the predictions in terms of RMSE by 20% for clay, 10% for silt, and 24% for sand when compared to BP-ANN. The physiographical units, soil types, geology maps, rainfall, and temperature were the most important predictors of PSFs, followed by the terrain and remotely sensed data. This study demonstrates the effectiveness of bio-inspired algorithms for improving ANN models. The outputs of this study will support and inform sustainable soil management practices, agro-ecological modeling, and hydrological modeling for the Mazandaran Province of Iran. Full article
(This article belongs to the Special Issue Remote Sensing for Soil Mapping and Monitoring)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Localization of the Mazandaran province in Iran (<b>a</b>); and spatial distribution of soil sampling locations within the Mazandaran Province (<b>b</b>).</p>
Full article ">Figure 2
<p>Flowchart diagram of the ANN training using metaheuristic optimization algorithms. α and β represent the weights and biases, respectively.</p>
Full article ">Figure 3
<p>Distribution of soil texture classes for all samples based on the USDA soil texture triangle. (Cl: clay; SiCl: silty clay; SiClLo: silty clay loam; SaCl: sandy clay; SaClLo: sandy clay loam; ClLo: clay loam; Si: silt; SiLo: silt loam; Lo: loam; Sa: sand; LoSa: loamy sand; SaLo: sandy loam).</p>
Full article ">Figure 4
<p>The relative influence of environmental covariates for predicting clay, silt, and sand contents using the MBO-ANN algorithm. (Refer to <a href="#remotesensing-13-01025-t0A1" class="html-table">Table A1</a> for covariate codes).</p>
Full article ">Figure 5
<p>Continues maps produced using the MBO-ANN model for the Mazandaran Province of: surficial clay content (<b>a</b>); surficial sand content (<b>b</b>); and surficial silt content (<b>c</b>); and textual classes (<b>d</b>).</p>
Full article ">
23 pages, 7492 KiB  
Article
Bottom Detection from Backscatter Data of Conventional Side Scan Sonars through 1D-UNet
by Jun Yan, Junxia Meng and Jianhu Zhao
Remote Sens. 2021, 13(5), 1024; https://doi.org/10.3390/rs13051024 - 8 Mar 2021
Cited by 19 | Viewed by 3886
Abstract
As widely applicated in many underwater research fields, conventional side-scan sonars require the sonar height to be at the seabed for geocoding seabed images. However, many interference factors, including compensation with unknown gains, suspended matters, etc., would bring difficulties in bottom detection. Existing [...] Read more.
As widely applicated in many underwater research fields, conventional side-scan sonars require the sonar height to be at the seabed for geocoding seabed images. However, many interference factors, including compensation with unknown gains, suspended matters, etc., would bring difficulties in bottom detection. Existing methods need manual parameter setups or to use postprocessing methods, which limits automatic and real-time processing in complex situations. To solve this problem, a one-dimensional U-Net (1D-UNet) model for sea bottom detection of side-scan data and the bottom detection and tracking method based on 1D-UNet are proposed in this work. First, the basic theory of sonar bottom detection and the interference factors is introduced, which indicates that deep learning of the bottom is a feasible solution. Then, a 1D-UNet model for detecting the sea bottom position from the side-scan backscatter strength sequences is proposed, and the structure and implementation of this model are illustrated in detail. Finally, the bottom detection and tracking algorithms of a single ping and continuous pings are presented on the basis of the proposed model. The measured side-scan sonar data in Meizhou Bay and Bayuquan District were selected in the experiments to verify the model and methods. The 1D-UNet model was first trained and applied with the side-scan data in Meizhou Bay. The training and validation accuracies were 99.92% and 99.77%, respectively, and the sea bottom detection accuracy of the training survey line was 99.88%. The 1D-UNet model showed good robustness to the interference factors of bottom detection and fully real-time performance in comparison with other methods. Moreover, the trained 1D-UNet model is used to process the data in the Bayuquan District for proving model generality. The proposed 1D-UNet model for bottom detection has been proven effective for side-scan sonar data and also has great potentials in wider applications on other types of sonars. Full article
(This article belongs to the Special Issue Deep Learning for Radar and Sonar Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Operation and one-ping port-side (starboard-side) backscatter strength sequence of the side-scan sonar. (<b>a</b>) shows the beam pattern and the sound propagation of a side-scan sonar; (<b>b</b>) shows a backscatter strength sequence of a port-side (starboard-side) ping affected by many factors.</p>
Full article ">Figure 2
<p>The flow chart of the proposed one-dimensional (1D)-UNet model.</p>
Full article ">Figure 3
<p>Sampling (input and output) for the 1D-UNet model. (<b>a</b>) is the raw backscatter strength sequence of a ping port-side data. (<b>b</b>) shows normalization and resizing of the sequence. (<b>c</b>) shows the corresponding sea bottom probabilities of each backscatter sample.</p>
Full article ">Figure 4
<p>Structure of the proposed 1D-UNet for bottom detection from backscatter strength data of the conventional side-scan sonar.</p>
Full article ">Figure 5
<p>Bottom detection result for ping data obtained by using the trained 1D-UNet model.</p>
Full article ">Figure 6
<p>Procedure of bottom detection and tracking of successive side-scan pings.</p>
Full article ">Figure 7
<p>Experimental regions of side-scan sonar data. (<b>a</b>) is the Meizhou Bay region; (<b>b</b>) is the Bayuquan District.</p>
Full article ">Figure 8
<p>Sampling for input and target data from the measured side-scan data. (<b>a</b>) is the selected survey line for the training; (<b>b</b>) is the waterfall image constructed using the backscatter strength data; (<b>c</b>) shows the known sea bottom result of the survey line; (<b>d</b>) is the input sequence from a port-side ping data; and (<b>e</b>) shows the corresponding target sequence that represents the bottom probabilities.</p>
Full article ">Figure 9
<p>Training and validation accuracies and losses of 1D-UNet in 20 epochs.</p>
Full article ">Figure 10
<p>Bottom detection results obtained by using the trained 1D-UNet model. (<b>a</b>) shows the detected sea bottom positions of the port-side and starboard-side pings on the side-scan waterfall image; (<b>b</b>) shows the comparison and differences between the bottom position labels with the port-side and starboard-side predicted results; (<b>c</b>) shows the histograms and fitted probability distribution function (PDF) curves of the position difference data from (<b>b</b>).</p>
Full article ">Figure 11
<p>Sea bottom detection and tracking of other survey lines in Meizhou Bay. Panels (<b>a</b>–<b>e</b>) are the corresponding bottom detection results represented on the waterfall image of five survey lines.</p>
Full article ">Figure 12
<p>Bottom detection results of five different side-scan ping data processed by the last peak method, 1D-CNN, and 1D-UNet. Panels (<b>a</b>–<b>e</b>) show five different backscatter strength sequences, and <span class="html-italic">i</span>1 to <span class="html-italic">i</span>3 (<span class="html-italic">i</span> = a, b, …, e) are the detected bottom results using the last peak, 1D-CNN, and 1D-UNet, respectively.</p>
Full article ">Figure 13
<p>Real-time experimental results of 10,000 successive pings using 1D-CNN and 1D-UNet. These results were processed by the computer with AMD R5-2600X CPU and GTX-2070 GPU. (<b>a</b>) shows the bottom detection results of these pings; (<b>b</b>) shows the corresponding time costs of each ping using 1D-CNN and 1D-UNet; (<b>c</b>) shows the histogram and PDF curves of the time costs of 1D-CNN and 1D-UNet.</p>
Full article ">Figure 14
<p>Model validation using side-scan data in Bayuquan. (<b>a</b>) shows the survey track lines, and the red lines were selected for validation of the 1D-UNet model. (<b>b</b>–<b>d</b>) are the bottom detection and tracking results of these three selected survey lines. (<b>b1</b>), (<b>c1</b>) and (<b>d1</b>) are the bottom detection result of three backscatter strength sequences from these three lines.</p>
Full article ">Figure 15
<p>Backscatter strength sequence with very low signal to noise ratio.</p>
Full article ">Figure 16
<p>Backscatter strength sequence with a very large suspended object in the water column region.</p>
Full article ">Figure 17
<p>Apply the 1D-UNet model for bottom detection from backscatter data of multibeam echosounders. (<b>a</b>,<b>b</b>) show two multibeam pings containing water column backscatter data. The backscatter strengths of selected beams in blue rectangles are processed using 1D-UNet, and the corresponding bottom detection results are obtained at the right.</p>
Full article ">
3 pages, 4851 KiB  
Comment
Comment on “Comparison of Cloud Cover Detection Algorithms on Sentinel-2 Images of the Amazon Tropical Forest”
by Olivier Hagolle and Jerome Colin
Remote Sens. 2021, 13(5), 1023; https://doi.org/10.3390/rs13051023 - 8 Mar 2021
Cited by 3 | Viewed by 2476
Abstract
In their recent study, Sanchez et al. compared various cloud detection methods applied to Sentinel-2, specifically on images acquired over the Amazonian region, known for its frequent cloud cover. Comparison of cloud screening methods for optical satellite images is a complex task, which [...] Read more.
In their recent study, Sanchez et al. compared various cloud detection methods applied to Sentinel-2, specifically on images acquired over the Amazonian region, known for its frequent cloud cover. Comparison of cloud screening methods for optical satellite images is a complex task, which must take several parameters into account, such as the definition of a cloud, which can differ according to the methods, the different coding of the cloud and shadow masks, the possible dilation of masks, and also the way the method must be used to perform in nominal conditions. We found that the otherwise serious and useful comparison of cloud masks by Sanchez et al. is not fair to the real performances of MAJA cloud detection, for two reasons: (i) two thirds of the images used in the comparison were acquired before the launch of Sentinel-2B satellite, when the revisit of the Sentinel-2 mission was 20 days instead of five days for the nominal conditions of the mission, and (ii) there is an error in the understanding of how MAJA cloud masks are coded which also probably artificially degraded the results of MAJA as compared to the other methods. Full article
Show Figures

Figure 1

Figure 1
<p>Equivalent to Sanchez et al. Figure 3 with: (<b>a</b>) the SENTINEL-2A L1C image of 7 May 2018 and (<b>b</b>) the MAJA cloud mask with a proper representation of binary codes. Here, we aggregated bit-mask combinations to match the classes as shown in Sanchez et al.</p>
Full article ">
18 pages, 2839 KiB  
Article
Accuracy Analysis of GNSS Hourly Ultra-Rapid Orbit and Clock Products from SHAO AC of iGMAS
by Qinming Chen, Shuli Song and Weili Zhou
Remote Sens. 2021, 13(5), 1022; https://doi.org/10.3390/rs13051022 - 8 Mar 2021
Cited by 19 | Viewed by 2657
Abstract
With the development of the global navigation satellite system(GNSS), the hourly ultra-rapid products of GNSS are attracting more attention due to their low latency and high accuracy. A new strategy and method was applied by the Shanghai Astronomical Observatory (SHAO) Analysis Center (AC) [...] Read more.
With the development of the global navigation satellite system(GNSS), the hourly ultra-rapid products of GNSS are attracting more attention due to their low latency and high accuracy. A new strategy and method was applied by the Shanghai Astronomical Observatory (SHAO) Analysis Center (AC) of the international GNSS Monitoring and Assessment Service (iGMAS) for generating 6-hourly and 1-hourly GNSS products, which mainly include the American Global Positioning System (GPS), the Russian Global’naya Navigatsionnaya Sputnikova Sistema (GLONASS), the European Union’s Galileo, and the Chinese BeiDou navigation satellite system (BDS). The 6-hourly and 1-hourly GNSS orbit and clock ultra-rapid products included a 24-h observation session which is determined by 24-h observation data from global tracking stations, and a 24-h prediction session which is predicted from the observation session. The accuracy of the 1-hourly orbit product improved about 1%, 31%, 13%, 11%, 23%, and 9% for the observation session and 18%, 43%, 45%, 34%, 53%, and 15% for the prediction session of GPS, GLONASS, Galileo, BDS Medium Earth Orbit (MEO), Inclined Geosynchronous Orbit (IGSO), and GEO orbit, when compared with reference products with high accuracy from the International GNSS service (IGS).The precision of the 1-hourly clock products can also be seen better than the 6-hourly clock products. The accuracy and precision of the 6-hourly and 1-hourly orbit and clock verify the availability and reliability of the hourly ultra-rapid products, which can be used for real-time or near-real-time applications, and show encouraging prospects. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of the global tracking stations used to generate hourly ultra-rapid orbit and clock global navigation satellite system (GNSS) products. The blue circles represent the stations, which can track American Global Positioning System (GPS) satellites; the green circles represent the stations tracking Ruassian Global’naya Navigatsionnaya Sputnikova Sistema (GLONASS); the yellow circles represent the stations tracking European Union’s Galileo, and the red circles represent the stations tracking Chinese BeiDou navigation satellite system (BDS) satellites.</p>
Full article ">Figure 2
<p>Slide window for hourly ultra-rapid obit and clock products generation.</p>
Full article ">Figure 3
<p>Data processing flow chart of the combined serial and parallel threads (CSPT) method for hourly orbit and clock products.</p>
Full article ">Figure 4
<p>Computational efficiency of 6H and 1H products for GPS/GLONASS/Galileo/BDS. The 6H and 1H are on behalf of 6-hourly and 1-hourly products, respectively.</p>
Full article ">Figure 5
<p>Average root mean square (RMS) in along, cross, and radial directions of differences between 6-hourlyorbit products generated by the new and old method and IGS/MGEX final products for 24-h observation session and the 1st–8th hour prediction session.</p>
Full article ">Figure 6
<p>Average RMS of 6-hourly GNSS ultra-rapid orbit products with respect to IGS/MGEX final products for 24-h observation session and 1st–8th prediction session. (<b>a</b>) shows RMS of the MEO orbit for GPS/GLONASS/Galileo, and MEO/IGSO orbit for BDS, (<b>b</b>) displays RMS of GEO orbit for BDS for the observation session and prediction session.</p>
Full article ">Figure 7
<p>Average RMS of 1-hourly GNSS ultra-rapid orbit products with respect to IGS/MGEX final products for 24-h observation session and 1st–2nd prediction session. (<b>a</b>) shows RMS of the MEO orbit for GPS/GLONASS/Galileo, and MEO/IGSO for BDS, (<b>b</b>) displays RMS of GEO orbit for BDS for the observation session and prediction session.</p>
Full article ">Figure 8
<p>Average SD of 6-hourly (<b>a</b>) and 1-hourly (<b>b</b>) GNSS clock versus reference products from IGS/MGEX.</p>
Full article ">Figure 8 Cont.
<p>Average SD of 6-hourly (<b>a</b>) and 1-hourly (<b>b</b>) GNSS clock versus reference products from IGS/MGEX.</p>
Full article ">Figure 9
<p>Average RMS of 6-hourly (<b>a</b>) and 1-hourly (<b>b</b>) GNSS clock versus reference products from IGS/MGEX.</p>
Full article ">Figure 9 Cont.
<p>Average RMS of 6-hourly (<b>a</b>) and 1-hourly (<b>b</b>) GNSS clock versus reference products from IGS/MGEX.</p>
Full article ">
18 pages, 8636 KiB  
Article
Evaluation of Three Different Machine Learning Methods for Object-Based Artificial Terrace Mapping—A Case Study of the Loess Plateau, China
by Hu Ding, Jiaming Na, Shangjing Jiang, Jie Zhu, Kai Liu, Yingchun Fu and Fayuan Li
Remote Sens. 2021, 13(5), 1021; https://doi.org/10.3390/rs13051021 - 8 Mar 2021
Cited by 20 | Viewed by 3127
Abstract
Artificial terraces are of great importance for agricultural production and soil and water conservation. Automatic high-accuracy mapping of artificial terraces is the basis of monitoring and related studies. Previous research achieved artificial terrace mapping based on high-resolution digital elevation models (DEMs) or imagery. [...] Read more.
Artificial terraces are of great importance for agricultural production and soil and water conservation. Automatic high-accuracy mapping of artificial terraces is the basis of monitoring and related studies. Previous research achieved artificial terrace mapping based on high-resolution digital elevation models (DEMs) or imagery. As a result of the importance of the contextual information for terrace mapping, object-based image analysis (OBIA) combined with machine learning (ML) technologies are widely used. However, the selection of an appropriate classifier is of great importance for the terrace mapping task. In this study, the performance of an integrated framework using OBIA and ML for terrace mapping was tested. A catchment, Zhifanggou, in the Loess Plateau, China, was used as the study area. First, optimized image segmentation was conducted. Then, features from the DEMs and imagery were extracted, and the correlations between the features were analyzed and ranked for classification. Finally, three different commonly-used ML classifiers, namely, extreme gradient boosting (XGBoost), random forest (RF), and k-nearest neighbor (KNN), were used for terrace mapping. The comparison with the ground truth, as delineated by field survey, indicated that random forest performed best, with a 95.60% overall accuracy (followed by 94.16% and 92.33% for XGBoost and KNN, respectively). The influence of class imbalance and feature selection is discussed. This work provides a credible framework for mapping artificial terraces. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Artificial terraces in Ansai, the Loess Plateau, China, which have almost returned to grassland after the implement of “Grain for Green” project. Erosion can be observed between the ridges of the terraces due to the lack of maintenance.</p>
Full article ">Figure 2
<p>Study area and data. (<b>a</b>,<b>b</b>) show the location of the Loess Plateau, China, and the study area. (<b>c</b>) Digital elevation models (DEMs) of the study area, and (<b>d</b>) WorldView-3 imagery of the study area with ground truth.</p>
Full article ">Figure 3
<p>Details of terraces in the study area. (<b>a</b>) Ortho-image of a subset (white polygon) in the study area; (<b>b</b>–<b>f</b>) show the panchromatic band, elevation, slope, shaded relief, and curvature of the subset, respectively.</p>
Full article ">Figure 4
<p>Workflow of the proposed method.</p>
Full article ">Figure 5
<p>Examples of selected features for terrace mapping.</p>
Full article ">Figure 6
<p>Segmentation result and training samples.</p>
Full article ">Figure 7
<p>Mapping results based on three machine learning (ML) classifiers.</p>
Full article ">Figure 8
<p>Estimates of the class ratio (R) in the training stage that triggers a balance of Table 10. fold random replicate runs. The gray margins represent the corresponding standard deviations.</p>
Full article ">
28 pages, 86897 KiB  
Article
Hydrocarbon Pollution Detection and Mapping Based on the Combination of Various Hyperspectral Imaging Processing Tools
by Véronique Achard, Pierre-Yves Foucher and Dominique Dubucq
Remote Sens. 2021, 13(5), 1020; https://doi.org/10.3390/rs13051020 - 8 Mar 2021
Cited by 15 | Viewed by 4516
Abstract
Oil extraction and transportation may lead to small or large scale accidental spills, whether at sea or on land. Detecting these spills is a major problem that can be addressed by means of hyperspectral images and specific processing methods. In this work, several [...] Read more.
Oil extraction and transportation may lead to small or large scale accidental spills, whether at sea or on land. Detecting these spills is a major problem that can be addressed by means of hyperspectral images and specific processing methods. In this work, several cases of onshore oil spills are studied. First, a controlled experiment was carried out: four boxes containing soil or sand mixed with crude oil or gasoil were deployed on the ONERA site near Fauga, France, and were overflown by HySpex hyperspectral cameras. Owing to this controlled experiment, different detection strategies were developed and tested, with a particular focus on the most automated methods requiring the least supervision. The methods developed were then applied to two very different cases: mapping of the shoreline contaminated due to the explosion of the Deepwater Horizon (DWH) platform based on AVIRIS images (AVIRIS: Airborne Visible/InfraRed Imaging Spectrometer), and detection of a tar pit on a former oil exploration site. The detection strategy depends on the type of oil, light or heavy, recently or formerly spilled, and on the substrate. In the first case (controlled experiment), the proposed methods included spectral index calculations, anomaly detection and spectral unmixing. In the case of DWH, spectral indices were computed and the unmixing method was tested. Finally, to detect the tar pit, a strategy based on anomaly detection and spectral indices was applied. In all the cases studied, the proposed methods were successful in detecting and mapping the oil pollution. Full article
(This article belongs to the Special Issue Monitoring Soil Contamination by Remote Sensors)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Color composition of the proof of concept (POC) SWIR image. The red circle indicates boxes of pure sand and soil, boxes of hydrocarbon (HC) mixtures are in the yellow circle and plastic materials are in the white circles. (<b>b</b>) Boxes containing sand, soil, and hydrocarbons mixed with soil or sand. The table gives the volume percentage of hydrocarbon in the mixture. The white bucket contains crude oil. (<b>c</b>) Mean reflectance spectra of the six boxes extracted from the HySpex image. The blue and red vertical lines correspond to the HC absorption features.</p>
Full article ">Figure 2
<p>Composition of the synthetic image and simulated image (RGB). In (<b>a</b>) the red, orange, and yellow squares refers to the location of crude oil mixtures with a gradation from red to yellow corresponding to decreasing percentage values of crude oil in the pixels from 100% to 25%. The blue squares correspond to gasoil mixtures with the same gradation from dark to light blue. Yellow background is sand, the green one is vegetation and the marron one, soil, with a vertical gradation to symbolize the relative percentage of two types of background in the pixels. The lower part of the image is real patches of vegetation extracted from the image. (<b>b</b>) is the RGB extract of the simulated image</p>
Full article ">Figure 3
<p>(<b>a</b>) RGB Airborne Visible/InfraRed Imaging Spectrometer (AVIRIS) image over Bay Jimmy; (<b>b</b>) RGB HySpex image over the tar pit (yellow outline) and a picture of the tar pit.</p>
Full article ">Figure 4
<p>Illustration of the area index (in blue).</p>
Full article ">Figure 5
<p>Receiver Operating Characteristic (ROC) curves computed for LRX with (<b>a</b>) IWR = 5 and (<b>b</b>) IWR = 11; (<b>c</b>) Area1700 with vegetation rejection; (<b>d</b>) Area1700 applied to LRX results (IWR11) without vegetation rejection; (<b>e</b>) CRX with 30 classes; (<b>f</b>) CRX with 40 classes.</p>
Full article ">Figure 6
<p>Endmembers extracted with Orthogonal Subspace Projection (OSP) applied to the entire image, with vegetation rejection test (Case OSPv1).</p>
Full article ">Figure 7
<p>Endmembers associated with the HC-based mixtures compared with the true HC-mixture spectra (cases OSPv1 and OSPv2).</p>
Full article ">Figure 8
<p>Abundance map (OSPv1) of HC+soil (<b>a</b>), crude oil+sand (<b>b</b>) and gasoil+sand (<b>c</b>). A horizontal (marked by the arrow) profile of the abundance is drawn. Red triangles indicate the true values.</p>
Full article ">Figure 9
<p>Reconstruction error (percentage).</p>
Full article ">Figure 10
<p>(<b>a</b>) HySpex SWIR POC2 image. The HC mixtures are in the yellow ellipse, (<b>b</b>) Area1700, (<b>c</b>) LRX (0.01% detection), (<b>d</b>) Area1700 computed on (<b>c</b>) mask, (<b>e</b>) CRX (0.01% detection), (<b>f</b>) Area1700 computed on (<b>e</b>) mask.</p>
Full article ">Figure 11
<p>Area1700 (<b>a</b>), Aire2300 (<b>b</b>) and KHI (<b>c</b>). Histograms are adjusted on the entire images (linear stretch with min = 0, and max = 0.5 for Area1700 and Area2300, and max = 0.02 for KHI).</p>
Full article ">Figure 11 Cont.
<p>Area1700 (<b>a</b>), Aire2300 (<b>b</b>) and KHI (<b>c</b>). Histograms are adjusted on the entire images (linear stretch with min = 0, and max = 0.5 for Area1700 and Area2300, and max = 0.02 for KHI).</p>
Full article ">Figure 12
<p>Oiled shorelines mapped from the same Aviris image [<a href="#B15-remotesensing-13-01020" class="html-bibr">15</a>].</p>
Full article ">Figure 13
<p>Environmental Response Management Application (ERMA), 17 September 2010—SCAT (Shoreline Cleanup and Assessment Technique) oiling ground observations [<a href="#B45-remotesensing-13-01020" class="html-bibr">45</a>].</p>
Full article ">Figure 14
<p>Spectral reflectance of the 3rd, 14th, 17th, and 5th endmembers.</p>
Full article ">Figure 15
<p>Abundance maps of the 3rd (<b>a</b>), 5th (<b>b</b>), 14th (<b>c</b>) and 17th (<b>d</b>) endmembers. Histograms are adjusted on the entire images (linear stretch [0, 0.5]). Pixels with the maximum abundance values (equal to 1) are at the centers of the red squares.</p>
Full article ">Figure 16
<p>Spectral reflectance of the tar pit acquired with HySpex and measured in the laboratory using an ASD FieldSpec spectroradiometer.</p>
Full article ">Figure 17
<p>RGB HySpex image (<b>a</b>). Normalized Area1700 with modified upper wavelength (<b>b</b>).</p>
Full article ">Figure 18
<p>(<b>a</b>) Anomaly detection scores using LRX (Bottom left, the entire image; top, the area in the red square; bottom right, a zoom in on the tar pit area); (<b>b</b>) Anomaly detection scores using CRX (bottom left, the whole image; top, the area in the red square; bottom right, a zoom in on the tar pit area) [<a href="#B46-remotesensing-13-01020" class="html-bibr">46</a>].</p>
Full article ">Figure 19
<p>RGB HySpex image (<b>a</b>), Normalized Area1700 applied to the most anomalous 1/1000 pixels (<b>b</b>).</p>
Full article ">
18 pages, 33199 KiB  
Article
Exploring the Variation Trend of Urban Expansion, Land Surface Temperature, and Ecological Quality and Their Interrelationships in Guangzhou, China, from 1987 to 2019
by Jianhui Xu, Yi Zhao, Caige Sun, Hanbin Liang, Ji Yang, Kaiwen Zhong, Yong Li and Xulong Liu
Remote Sens. 2021, 13(5), 1019; https://doi.org/10.3390/rs13051019 - 8 Mar 2021
Cited by 23 | Viewed by 2981
Abstract
This study explored the model of urban impervious surface (IS) density, land surface temperature (LST), and comprehensive ecological evaluation index (CEEI) from urban centers to suburbs. The interrelationships between these parameters in Guangzhou from 1987 to 2019 were analyzed using time-series Landsat-5 TM [...] Read more.
This study explored the model of urban impervious surface (IS) density, land surface temperature (LST), and comprehensive ecological evaluation index (CEEI) from urban centers to suburbs. The interrelationships between these parameters in Guangzhou from 1987 to 2019 were analyzed using time-series Landsat-5 TM (Thematic Mapper), Landsat-8 OLI (Operational Land Imager), and TIRS (Thermal Infrared Sensor) images. The urban IS densities were calculated in concentric rings using time-series IS fractions, which were used to construct an inverse S-shaped urban IS density function to depict changes in urban form and the spatio-temporal dynamics of urban expansion from the urban center to the suburbs. The results indicated that Guangzhou experienced expansive urban growth, with the patterns of urban spatial structure changing from a single-center to a multi-center structure over the past 32 years. Next, the normalized LST and CEEI in each concentric ring were calculated, and their variation trends from the urban center to the suburbs were modeled using linear and nonlinear functions, respectively. The results showed that the normalized LST had a gradual decreasing trend from the urban center to the suburbs, while the CEEI showed a significant increasing trend. During the 32-year rapid urban development, the normalized LST difference between the urban center and suburbs increased gradually with time, and the CEEI significantly decreased. This indicated that rapid urbanization significantly expanded the impervious surface areas in Guangzhou, leading to an increase in the LST difference between urban centers and suburbs and a deterioration in ecological quality. Finally, the potential interrelationships among urban IS density, normalized LST, and CEEI were also explored using different models. This study revealed that rapid urbanization has produced geographical convergence between several ISs, which may increase the risk of the urban heat island effect and degradation of ecological quality. Full article
(This article belongs to the Special Issue Understanding Urban Systems Using Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area: (<b>a</b>) Guangdong–Hong Kong–Macao Greater Bay Area (GBA), (<b>b</b>) Guangzhou city, and (<b>c</b>) the city center and the rings in Guangzhou.</p>
Full article ">Figure 2
<p>The time-series impervious surface (IS) fraction in (<b>a</b>) 1987, (<b>b</b>) 1999, (<b>c</b>) 2009, and (<b>d</b>) 2019.</p>
Full article ">Figure 3
<p>The urban IS densities in concentric rings with the distance to the urban center in 1987, 1999, 2009, and 2019.</p>
Full article ">Figure 4
<p>Fitting of the urban ISF density data with the inverse S-shape function in (<b>a</b>) 1987, (<b>b</b>) 1999, (<b>c</b>) 2009, and (<b>d</b>) 2019.</p>
Full article ">Figure 5
<p>The normalized land surface temperature in concentric rings with the distance to the urban center in 1987, 1999, 2009, and 2019.</p>
Full article ">Figure 6
<p>The fitting curve of normalized land surface temperature as a function of the urban center distance in (<b>a</b>) 1987, (<b>b</b>) 1999, (<b>c</b>) 2009, and (<b>d</b>) 2019.</p>
Full article ">Figure 7
<p>The comprehensive ecological evaluation index (CEEI) in concentric rings with the distance to the urban center in 1987, 1999, 2009, and 2019.</p>
Full article ">Figure 8
<p>The fitting curve of CEEI as a function of the urban center distance in (<b>a</b>) 1987, (<b>b</b>) 1999, (<b>c</b>) 2009, and (<b>d</b>) 2019.</p>
Full article ">Figure 9
<p>The regression models between urban IS densities versus normalized LST in each concentric ring in (<b>a</b>) 1987, (<b>b</b>) 1999, (<b>c</b>) 2009, and (<b>d</b>) 2019.</p>
Full article ">Figure 10
<p>The regression models between urban IS densities versus CEEI in each concentric ring in (<b>a</b>) 1987, (<b>b</b>) 1999, (<b>c</b>) 2009, and (<b>d</b>) 2019.</p>
Full article ">
25 pages, 23534 KiB  
Article
Performance Comparison of Machine Learning Models for Annual Precipitation Prediction Using Different Decomposition Methods
by Chao Song and Xiaohong Chen
Remote Sens. 2021, 13(5), 1018; https://doi.org/10.3390/rs13051018 - 8 Mar 2021
Cited by 14 | Viewed by 3255
Abstract
It has become increasingly difficult in recent years to predict precipitation scientifically and accurately due to the dual effects of human activities and climatic conditions. This paper focuses on four aspects to improve precipitation prediction accuracy. Five decomposition methods (time-varying filter-based empirical mode [...] Read more.
It has become increasingly difficult in recent years to predict precipitation scientifically and accurately due to the dual effects of human activities and climatic conditions. This paper focuses on four aspects to improve precipitation prediction accuracy. Five decomposition methods (time-varying filter-based empirical mode decomposition (TVF-EMD), robust empirical mode decomposition (REMD), complementary ensemble empirical mode decomposition (CEEMD), wavelet transform (WT), and extreme-point symmetric mode decomposition (ESMD) combined with the Elman neural network (ENN)) are used to construct five prediction models, i.e., TVF-EMD-ENN, REMD-ENN, CEEMD-ENN, WT-ENN, and ESMD-ENN. The variance contribution rate (VCR) and Pearson correlation coefficient (PCC) are utilized to compare the performances of the five decomposition methods. The wavelet transform coherence (WTC) is used to determine the reason for the poor prediction performance of machine learning algorithms in individual years and the relationship with climate indicators. A secondary decomposition of the TVF-EMD is used to improve the prediction accuracy of the models. The proposed methods are used to predict the annual precipitation in Guangzhou. The subcomponents obtained from the TVF-EMD are the most stable among the four decomposition methods, and the North Atlantic Oscillation (NAO) index, the Nino 3.4 index, and sunspots have a smaller influence on the first subcomponent (Sc-1) than the other subcomponents. The TVF-EMD-ENN model has the best prediction performance and outperforms traditional machine learning models. The secondary decomposition of the Sc-1 of the TVF-EMD model significantly improves the prediction accuracy. Full article
(This article belongs to the Special Issue Artificial Intelligence in Remote Sensing of Atmospheric Environment)
Show Figures

Figure 1

Figure 1
<p>Implementation steps of the time-varying filter-based empirical mode decomposition (TVF-EMD) [<a href="#B20-remotesensing-13-01018" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>The technical route of WT.</p>
Full article ">Figure 3
<p>The schematic diagram of the Elman neural network (ENN).</p>
Full article ">Figure 4
<p>The region of Guangzhou.</p>
Full article ">Figure 5
<p>Annual precipitation in Guangzhou.</p>
Full article ">Figure 6
<p>The precipitation decomposition result of TVF-EMD. Tn represents the nth subcomponent obtained from TVF-EMD.</p>
Full article ">Figure 7
<p>The absolute prediction error of TVF-EMD-ENN. TEn represents the absolute prediction error of the nth subcomponent obtained from TVF-EMD-ENN.</p>
Full article ">Figure 8
<p>The precipitation decomposition result of robust empirical mode decomposition (REMD). Rn represents the nth subcomponent obtained from REMD.</p>
Full article ">Figure 9
<p>The absolute prediction error of REMD-ENN. REn represents the absolute prediction error of the nth subcomponent obtained from REMD-ENN.</p>
Full article ">Figure 10
<p>The precipitation decomposition result of ensemble empirical mode decomposition (CEEMD). Cn represents the nth subcomponent obtained from CEEMD.</p>
Full article ">Figure 11
<p>The absolute prediction error of CEEMD-ENN. CEn represents the absolute prediction error of the nth subcomponent obtained from CEEMD-ENN.</p>
Full article ">Figure 12
<p>The precipitation decomposition result of the WT. Xn represents the nth subcomponent obtained from the WT.</p>
Full article ">Figure 13
<p>The absolute prediction error of WT-ENN. XEn represents the absolute prediction error of the nth subcomponent obtained from WT-ENN.</p>
Full article ">Figure 14
<p>The precipitation decomposition result of ESMD. En represents the nth subcomponent obtained from ESMD.</p>
Full article ">Figure 15
<p>The absolute prediction error of ESMD-ENN. EEn represents the absolute prediction error of the nth subcomponent obtained by ESMD-ENN.</p>
Full article ">Figure 16
<p>The wavelet transform coherence (WTC) between Sc-1 and the sunspot data (<b>A</b>–<b>D</b>), North Atlantic oscillation (NAO) (<b>E</b>–<b>H</b>), and the Nino 3.4 index (<b>I</b>–<b>L</b>). (<b>A</b>) represents the relationship between Sc-1 of REMD and the sunspots, (<b>B</b>) represents the relationship between SC-1 of CEEMD and the sunspots, (<b>C</b>) represents the relationship between Sc-1 of the WT and the sunspots, and (<b>D</b>) represents the relationship between Sc-1 of ESMD and the sunspots. (<b>E</b>–<b>H</b>) and (<b>I</b>–<b>L</b>) reflect the corresponding relationships between the Sc-1 of the same models with the NAO index and the Nino 3.4 index, respectively. The period is measured in years. Thick contours denote the 5% significance levels. The pale regions denote the cone of influence (COI), where edge effects might distort the results. The colors denote the strength of the wavelet transform.</p>
Full article ">Figure 17
<p>The WTC between Sc-1 of TVF-EMD and the sunspots, NAO index, and Nino 3.4 index. (<b>a</b>) represents the relationship between the SC-1 and the sunspots, (<b>b</b>) represents the relationship between the Sc-1 and the NAO index, and (<b>c</b>) represents the relationship between the Sc-1 and the Nino 3.4 index.</p>
Full article ">Figure 18
<p>The variance contribution rate (VCR) and Pearson correlation coefficient (PCC) values of the five decomposition methods. The black area indicates that this item does not exist</p>
Full article ">Figure 19
<p>The relative prediction errors of the five models. Sc-n represents the nth subcomponent obtained from the decomposition method. Ap=Annual precipitation</p>
Full article ">Figure 20
<p>The flowchart for the secondary decomposition of Sc-1.</p>
Full article ">Figure 21
<p>The improvement in the relative error of the Sc-1 and annual precipitation prediction after TVF − EMD decomposition. CENN = CEEMD − ENN; CTEENN = CEEMD − TVF − EMD − ENN; EENN = ESMD − ENN; ETENN = ESMD − TVF − EMD − ENN; RENN = REMD − ENN; RTEENN = REMD − TVF − EMD − ENN; WTENN = WT − ENN; WTEENN = WT − TVF − EMD − ENN.</p>
Full article ">Figure 22
<p>The secondary decomposition of Sc-1 improved the annual precipitation prediction of the four models. CENN = CEEMD − ENN; CTEENN = CEEMD − TVF − EMD − ENN; EENN = ESMD − ENN; ETENN = ESMD − TVF − EMD − ENN; RENN = REMD − ENN; RTEENN = REMD − TVF − EMD − ENN; WTENN = WT − ENN; WTEENN = WT − TVF − EMD − ENN.</p>
Full article ">
27 pages, 17831 KiB  
Article
Spatiotemporal Variability of Mesoscale Eddies in the Indonesian Seas
by Zhanjiu Hao, Zhenhua Xu, Ming Feng, Qun Li and Baoshu Yin
Remote Sens. 2021, 13(5), 1017; https://doi.org/10.3390/rs13051017 - 8 Mar 2021
Cited by 14 | Viewed by 3096
Abstract
Mesoscale eddies are ubiquitous in the world ocean and well researched both globally and regionally, while their properties and distributions across the whole Indonesian Seas are not yet fully understood. This study investigates for the first time the spatiotemporal variations and generation mechanisms [...] Read more.
Mesoscale eddies are ubiquitous in the world ocean and well researched both globally and regionally, while their properties and distributions across the whole Indonesian Seas are not yet fully understood. This study investigates for the first time the spatiotemporal variations and generation mechanisms of mesoscale eddies across the whole Indonesian Seas. Eddies are detected from altimetry sea level anomalies by an automatic identification algorithm. The Sulu Sea, Sulawesi Sea, Maluku Sea and Banda Sea are the main eddy generation regions. More than 80% of eddies are short-lived with a lifetime below 30 days. The properties of eddies exhibit high spatial inhomogeneity, with the typical amplitudes and radiuses of 2–6 cm and 50–160 km, respectively. The most energetic eddies are observed in the Sulawesi Sea and Seram Sea. Eddies feature different seasonal cycles between anticyclonic and cyclonic eddies in each basin, especially given that the average latitude of the eddy centroid has inverse seasonal variations. About 48% of eddies in the Sulawesi Sea are highly nonlinear, which is the case for less than 30% in the Sulu Sea and Banda Sea. Instability analysis is performed using high-resolution model outputs from Bluelink Reanalysis to assess mechanisms of eddy generation. Barotropic instability of the mean flow dominates eddy generation in the Sulu Sea and Sulawesi Sea, while baroclinic instability is slightly more in the Maluku Sea and Banda Sea. Full article
(This article belongs to the Special Issue Remote Sensing Applications in Ocean Observation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of the upper-ocean circulation in the Indonesian Seas. Color shading is the bathymetry from ETOPO1. The solid line is the 200 m isobath. KS, LS and OS represent the Karimata Strait, Lombok Strait and Ombai Strait, respectively. MC and NECC represent the Mindanao Current and North Equatorial Countercurrent, respectively. Four dashed boxes are the Sulu Box, Sulawesi Box, Maluku Box and Banda Box from north to south, respectively.</p>
Full article ">Figure 2
<p>Spatial pattern of empirical orthogonal function mode 1 of sea surface height from (<b>a</b>) AVISO and (<b>b</b>) BRAN. (<b>c</b>) Principal component of empirical orthogonal function mode 1 from AVISO (black line) and BRAN (red line). The gray shadings in (<b>c</b>) represent El Niño events.</p>
Full article ">Figure 3
<p>Spatial pattern of empirical orthogonal function mode 2 of sea surface height from (<b>a</b>) AVISO and (<b>b</b>) BRAN. (<b>c</b>) Principal component of empirical orthogonal function mode 2 from AVISO (black line) and BRAN (red line).</p>
Full article ">Figure 4
<p>Spatial distribution of the numbers of identified (<b>a</b>) anticyclonic (AE) and (<b>b</b>) cyclonic (CE) eddies in the Indonesian Seas over the 1993–2018 period.</p>
Full article ">Figure 5
<p>Spatial distribution of eddy genesis and decay events over the 1993–2018 period: (<b>a</b>) AE genesis events, (<b>b</b>) CE genesis events, (<b>c</b>) AE decay events and (<b>d</b>) CE decay events. The unit is the number of events.</p>
Full article ">Figure 6
<p>Spatial distribution of propagation velocity vectors (black arrows, unit: cm/s) for (<b>a</b>) AE and (<b>b</b>) CE over the 1993–2018 period. The color shading represents the standard deviation (STD) of propagation azimuths relative to west at each 0.25° × 0.25° grid (units: degree).</p>
Full article ">Figure 7
<p>Upper-tail cumulative histograms of the eddy lifespan in the (<b>a</b>) Sulu Sea, (<b>b</b>) Sulawesi Sea, (<b>c</b>) Maluku Sea and (<b>d</b>) Banda Sea. The red and blue lines in each panel correspond to AE and CE, respectively.</p>
Full article ">Figure 8
<p>Spatial distribution (color spotted) of mean eddy properties over the 1993–2018 period: (<b>a</b>) AE amplitude, (<b>b</b>) CE amplitude, (<b>c</b>) AE radius and (<b>d</b>) CE radius. A value in each grid is averaged from eddies centered within this grid. The units of amplitude and radius are cm and km, respectively.</p>
Full article ">Figure 9
<p>Spatial distribution (color spotted) of mean eddy properties over the 1993–2018 period: (<b>a</b>) the common logarithm of mean AE EKE, (<b>b</b>) the common logarithm of mean CE EKE, (<b>c</b>) mean AE vorticity and (<b>d</b>) mean CE vorticity. A value in each grid is averaged from eddies centered within this grid. The units of EKE and vorticity are cm<sup>2</sup>/s<sup>2</sup> and 10<sup>−6</sup>s<sup>−1</sup>, respectively.</p>
Full article ">Figure 10
<p>Seasonal cycles of eddy properties (black bars are for CE and white bars are for AE) in the Sulu Sea: (<b>a</b>) number of eddies, (<b>b</b>) amplitude, (<b>c</b>) radius, (<b>d</b>) central latitude, (<b>e</b>) mean EKE and (<b>f</b>) mean vorticity. The units of amplitude, radius, mean EKE and mean vorticity correspond to cm, km, cm<sup>2</sup>/s<sup>2</sup> and 10<sup>−6</sup>s<sup>−1</sup>.</p>
Full article ">Figure 11
<p>Seasonal cycles of eddy properties (black bars are for CE and white bars are for AE) in the Sulawesi Sea: (<b>a</b>) number of eddies, (<b>b</b>) amplitude, (<b>c</b>) radius, (<b>d</b>) central latitude, (<b>e</b>) mean EKE and (<b>f</b>) mean vorticity. The units of amplitude, radius, mean EKE and mean vorticity correspond to cm, km, cm<sup>2</sup>/s<sup>2</sup> and 10<sup>−6</sup>s<sup>−1</sup>.</p>
Full article ">Figure 12
<p>Seasonal cycles of eddy properties (black bars are for CE and white bars are for AE) in the Maluku Sea: (<b>a</b>) number of eddies, (<b>b</b>) amplitude, (<b>c</b>) radius and (<b>d</b>) central latitude. The units of amplitude and radius correspond to cm and km.</p>
Full article ">Figure 13
<p>Seasonal cycles of eddy properties (black bars are for CE and white bars are for AE) in the Banda Sea: (<b>a</b>) number of eddies, (<b>b</b>) amplitude, (<b>c</b>) radius, (<b>d</b>) central latitude, (<b>e</b>) mean EKE and (<b>f</b>) mean vorticity. The units of amplitude, radius, mean EKE and mean vorticity correspond to cm, km, cm<sup>2</sup>/s<sup>2</sup> and 10<sup>−6</sup>s<sup>−1</sup>.</p>
Full article ">Figure 14
<p>Upper-tail cumulative histograms of <span class="html-italic">U</span>/<span class="html-italic">c</span> in the (<b>a</b>) Sulu Sea, (<b>b</b>) Sulawesi Sea and (<b>c</b>) Banda Sea. The red and blue lines in each panel correspond to AE and CE, respectively. Additionally, the vertical dotted lines in each panel indicate that the value of <span class="html-italic">U</span>/<span class="html-italic">c</span> is equal to 1.</p>
Full article ">Figure 15
<p>Seasonal variation of volume integral barotropic conversion rate (BTR), baroclinic conversion rate (BCR) and Kelvin-Helmholtz conversion rate (KHR) in the (<b>a</b>) Sulu Box, (<b>b</b>) Sulawesi Box, (<b>c</b>) Maluku Box and (<b>d</b>) Banda Box, respectively. The unit is m<sup>5</sup>/s<sup>3</sup>.</p>
Full article ">Figure 16
<p>Spatial distribution of eddy genesis events dominated by (<b>a</b>) barotropic instability and (<b>b</b>) baroclinic instability. The unit is the number of eddy genesis events.</p>
Full article ">Figure A1
<p>Mean currents (blue arrows) averaged above 300 m depth in (<b>a</b>) January, (<b>b</b>) April, (<b>c</b>) July and (<b>d</b>) October from the BRAN outputs. The solid line is the 200 m isobath. Location of the four subregions in the Sulu Sea, Sulawesi Sea, Maluku Sea and Banda Sea. The boxes are (<b>a</b>) Sulu Box, (<b>b</b>) Banda Box, (<b>c</b>) Sulawesi Box and (<b>d</b>) Maluku Box, respectively. For regions shallower than 300 m, the average is for the whole water column.</p>
Full article ">Figure A2
<p>Spatial distribution of eddy genesis and decay events over the 1994–2015 period from AVISO (the first row) and BRAN (the second row): (<b>a</b>) genesis number from AVISO, (<b>b</b>) decay number from AVISO, (<b>c</b>) genesis number from BRAN and (<b>d</b>) decay number from BRAN. The unit is the number of events.</p>
Full article ">Figure A3
<p>Trajectories of same eddies from AVISO and BRAN in the (<b>a</b>) Sulu Sea, (<b>b</b>) Sulawesi Sea and (<b>c</b>) Banda Sea. The red line represents the trajectory of an eddy observed in AVISO. The blue line represents the trajectory of the corresponding eddy simulated by BRAN. The dots represent the initial locations of each trajectory.</p>
Full article ">
18 pages, 35894 KiB  
Article
Improving the Estimation of Weighted Mean Temperature in China Using Machine Learning Methods
by Zhangyu Sun, Bao Zhang and Yibin Yao
Remote Sens. 2021, 13(5), 1016; https://doi.org/10.3390/rs13051016 - 8 Mar 2021
Cited by 33 | Viewed by 3298
Abstract
As a crucial parameter in estimating precipitable water vapor from tropospheric delay, the weighted mean temperature (Tm) plays an important role in Global Navigation Satellite System (GNSS)-based water vapor monitoring techniques. However, the rigorous calculation of Tm requires vertical [...] Read more.
As a crucial parameter in estimating precipitable water vapor from tropospheric delay, the weighted mean temperature (Tm) plays an important role in Global Navigation Satellite System (GNSS)-based water vapor monitoring techniques. However, the rigorous calculation of Tm requires vertical profiles of temperature and water vapor pressure that are difficult to acquire in practice. As a result, empirical models are widely used but have limited accuracy. In this study, we use three machine learning methods, i.e., random forest (RF), backpropagation neural network (BPNN), and generalized regression neural network (GRNN), to improve the estimation of empirical Tm in China. The basic idea is to use the high-quality radiosonde observations estimated Tm to calibrate and optimize the empirical Tm through machine learning methods. Validating results show that the three machine learning methods improve the Tm accuracy by 37.2%, 32.6%, and 34.9% compared with the global pressure and temperature model 3 (GPT3). In addition to the overall accuracy improvement, the proposed methods also mitigate the accuracy variations in space and time, guaranteeing evenly high accuracy. This study provides a new idea to estimate Tm, which could potentially contribute to the GNSS meteorology. Full article
(This article belongs to the Special Issue Remote Sensing of Land Surface and Earth System Modelling)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Topography of the study area and geographical distribution of the radiosonde stations. The red points indicate the location of the radiosonde stations.</p>
Full article ">Figure 2
<p>Structure of backpropagation neural network (BPNN).</p>
Full article ">Figure 3
<p>Structure of generalized regression neural network (GRNN).</p>
Full article ">Figure 4
<p>Root mean square errors versus (<b>a</b>) the number of trees, (<b>b</b>) the neuron number in the hidden layer, and (<b>c</b>) the spread value derived from the 10-fold cross-validation technique.</p>
Full article ">Figure 5
<p>Scatter plots of estimated the weighted mean temperature (<span class="html-italic">T<sub>m</sub></span>) against observed <span class="html-italic">T<sub>m</sub></span> for different models: (<b>a</b>) RF-model fitting; (<b>b</b>) RF-cross-validation; (<b>c</b>) BPNN-model fitting; (<b>d</b>) BPNN-cross-validation; (<b>e</b>) GRNN-model fitting; (<b>f</b>) GRNN-cross-validation. The dashed line is the 1:1 line. RF: random forest; BPNN: backpropagation neural network; GRNN: generalized regression neural network.</p>
Full article ">Figure 6
<p>Scatter plots of estimated <span class="html-italic">T<sub>m</sub></span> against observed <span class="html-italic">T<sub>m</sub></span> for GPT3 model. The dashed line is the 1:1 line. GPT3: global pressure and temperature model 3.</p>
Full article ">Figure 7
<p>Spatial distribution of the root mean square errors (RMSEs) at each radiosonde station for (<b>a</b>) GPT3 (global pressure and temperature model 3), (<b>b</b>) RF (random forest), (<b>c</b>) BPNN (backpropagation neural network), and (<b>d</b>) GRNN (generalized regression neural network).</p>
Full article ">Figure 8
<p>Root mean square errors of different models at different latitude bands.</p>
Full article ">Figure 9
<p>Root mean square errors of different models at different height layers.</p>
Full article ">Figure 10
<p>Root mean square errors of different models at different days.</p>
Full article ">Figure 11
<p><span class="html-italic">T<sub>m</sub></span> time series estimated from different models and radiosonde observations at the station CHM00057083 (34.71°N, 113.65°E).</p>
Full article ">Figure 12
<p>Power spectrums of <span class="html-italic">T<sub>m</sub></span> residuals from (<b>a</b>) GPT3 (global pressure and temperature model 3), (<b>b</b>) RF (random forest), (<b>c</b>) BPNN (backpropagation neural network), and (<b>d</b>) GRNN (generalized regression neural network) models at the station CHM00057083 (34.71°N, 113.65°E).</p>
Full article ">
23 pages, 5469 KiB  
Article
Integration of Sentinel-3 OLCI Land Products and MERRA2 Meteorology Data into Light Use Efficiency and Vegetation Index-Driven Models for Modeling Gross Primary Production
by Fengji Zhang, Zhijiang Zhang, Yi Long and Ling Zhang
Remote Sens. 2021, 13(5), 1015; https://doi.org/10.3390/rs13051015 - 8 Mar 2021
Cited by 1 | Viewed by 2649
Abstract
Accurately and reliably estimating total terrestrial gross primary production (GPP) on a large scale is of great significance for monitoring the carbon cycle process. The Sentinel-3 satellite provides the OLCI FAPAR and OTCI products, which possess a higher spatial and temporal resolution than [...] Read more.
Accurately and reliably estimating total terrestrial gross primary production (GPP) on a large scale is of great significance for monitoring the carbon cycle process. The Sentinel-3 satellite provides the OLCI FAPAR and OTCI products, which possess a higher spatial and temporal resolution than MODIS products. However, few studies have focused on using LUE models and VI-driven models based on the Sentinel-3 satellites to estimate GPP on a large scale. The purpose of this study is to evaluate the performance of Sentinel-3 OLCI FAPAR and OTCI products combined with meteorology reanalysis data in estimating GPP at site and regional scale. Firstly, we integrated OLCI FAPAR and meteorology reanalysis data into the MODIS GPP algorithm and eddy covariance light use efficiency (EC-LUE) model (GPPMODIS-GPP and GPPEC-LUE, respectively). Then, we combined OTCI and meteorology reanalysis data with the greenness and radiation (GR) model and vegetation index (VI) model (GPPGR and GPPVI, respectively). Lastly, GPPMODIS-GPP, GPPEC-LUE, GPPGR, and GPPVI were evaluated against the eddy covariance flux data (GPPEC) at the site scale and MODIS GPP products (GPPMOD17) at the regional scale. The results showed that, at the site scale, GPPMODIS-GPP and GPPEC-LUE agreed well with GPPEC for the US-Ton site, with R2 = 0.73 and 0.74, respectively. The performance of GPPGR and GPPVI varied across different biome types. Strong correlations were obtained across deciduous broadleaf forests, mixed forests, grasslands, and croplands. At the same time, there are overestimations and underestimations in croplands, evergreen needleleaf forests and deciduous broadleaf forests. At the regional scale, the annual mean and maximum daily GPPMODIS-GPP and GPPEC-LUE agreed well with GPPMOD17 in 2017 and 2018, with R2 > 0.75. Overall, the above findings demonstrate the feasibility of using Sentinel-3 OLCI FAPAR and OTCI products combined with meteorology reanalysis data through LUE and VI-driven models to estimate GPP, and fill in the gaps for the large-scale evaluation of GPP via Sentinel-3 satellites. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographical location of AmeriFlux sites and research areas employed in this study.</p>
Full article ">Figure 2
<p>Scatter plots of daily Modern-Era Retrospective analysis for Research and Applications (MERRA2) meteorology reanalysis data against daily site meteorology data: (<b>a</b>) T2M_MIN; (<b>b</b>) T2M_MEAN; (<b>c</b>) IPAR; and (<b>d</b>) VPD from different biome types.</p>
Full article ">Figure 3
<p>Scatter plots between GPP<sub>MODISGPP</sub>, GPP<sub>EC-LUE</sub>, GPP<sub>MOD17,</sub> and GPP<sub>EC</sub> for all sites in 2017–2018. The left column refers to the MODIS-GPP model, the middle column represents the EC-LUE model, and the right column represents the MODIS-GPP products.</p>
Full article ">Figure 4
<p>Temporal dynamics of GPP<sub>MODIS-GPP</sub>, GPP<sub>EC-LUE</sub>, and GPP<sub>EC</sub> for all sites in 2017–2018. (<b>a</b>): evergreen needleleaf forest site, (<b>b</b>) deciduous broadleaf forest site, (<b>c</b>) mixed forest site, (<b>d</b>) closed shrubland site, (<b>e</b>) open shrubland site, (<b>f</b>) woody savannas site, (<b>g</b>) grassland site, (<b>h</b>) cropland site.</p>
Full article ">Figure 5
<p>Scatter plots between GPP<sub>GR</sub> and GPP<sub>EC</sub> for all sites in 2017–2018. (<b>a</b>–<b>h</b>) represent the site the same as <a href="#remotesensing-13-01015-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>Scatter plots between GPP<sub>VI</sub> and GPP<sub>EC</sub> for all sites in 2017–2018. (<b>a</b>–<b>h</b>) represent the site the same as <a href="#remotesensing-13-01015-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 7
<p>GPP<sub>GR</sub>, GPP<sub>VI</sub>, and GPP<sub>EC</sub> temporal dynamics for all sites in 2017–2018. (<b>a</b>–<b>h</b>) represent the site the same as <a href="#remotesensing-13-01015-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>Spatial distribution of the annual mean GPP from the MODIS-GPP algorithm (<b>a</b>,<b>d</b>), EC-LUE model (<b>b</b>,<b>e</b>), and MODIS-GPP products (<b>c</b>,<b>f</b>) in 2017 and 2018 and the maximum daily GPP from the MODIS-GPP algorithm (<b>g</b>,<b>j</b>), EC-LUE model (<b>h</b>,<b>k</b>), and MODIS-GPP products (<b>i</b>,<b>l</b>) in 2017 and 2018.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop