[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (81)

Search Parameters:
Keywords = NOAA AVHRR

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
10 pages, 959 KiB  
Communication
One Algorithm to Rule Them All? Defining Best Strategy for Land Surface Temperature Retrieval from NOAA-AVHRR Afternoon Satellites
by Yves Julien, José A. Sobrino and Juan-Carlos Jiménez-Muñoz
Remote Sens. 2024, 16(15), 2720; https://doi.org/10.3390/rs16152720 - 25 Jul 2024
Viewed by 330
Abstract
The NOAA-AVHRR (National Oceanographic and Atmospheric Administration–Advanced Very High-Resolution Radiometer) archive includes data from 1981 onwards, which allow for estimating land surface temperature (LST), a key parameter for the study of global warming as well as surface characterization. However, algorithms for LST retrieval [...] Read more.
The NOAA-AVHRR (National Oceanographic and Atmospheric Administration–Advanced Very High-Resolution Radiometer) archive includes data from 1981 onwards, which allow for estimating land surface temperature (LST), a key parameter for the study of global warming as well as surface characterization. However, algorithms for LST retrieval were developed before the latest sensors and were based on more reduced atmospheric datasets. Here, we present 50 novel sets of coefficients for an LST retrieval algorithm from NOAA-AVHRR sensors, to which we added one historical methodology, which we validate against historical in situ as well as independent satellite data. This validation shows that the historical algorithm performs surprisingly well, with an in situ RMSE below 1.5 K and a quasi-null bias when compared with independent satellite data. A couple of the novel algorithms also perform within expectations (errors below 1.5 K), so any of these could be used for the complete processing of the AVHRR dataset. In our case, considering consistency with previous works, we opt for the use of the historical algorithm, now also tested for more recent periods. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of the satellite validation data through 1981–2022. The number of validation data points for each satellite are in parentheses.</p>
Full article ">
38 pages, 11952 KiB  
Article
NOAA MODIS SST Reanalysis Version 1
by Olafur Jonasson, Alexander Ignatov, Boris Petrenko, Victor Pryamitsyn and Yury Kihai
Remote Sens. 2023, 15(23), 5589; https://doi.org/10.3390/rs15235589 - 30 Nov 2023
Cited by 1 | Viewed by 1130
Abstract
The first NOAA full-mission reanalysis (RAN1) of the sea surface temperature (SST) from the two Moderate Resolution Imaging Spectroradiometers (MODIS) onboard Terra (24 February 2000–present) and Aqua (4 July 2002–present) was performed. The dataset was produced using the NOAA Advanced Clear-Sky Processor for [...] Read more.
The first NOAA full-mission reanalysis (RAN1) of the sea surface temperature (SST) from the two Moderate Resolution Imaging Spectroradiometers (MODIS) onboard Terra (24 February 2000–present) and Aqua (4 July 2002–present) was performed. The dataset was produced using the NOAA Advanced Clear-Sky Processor for Ocean (ACSPO) enterprise SST system from Collection 6.1 brightness temperatures (BTs) in three MODIS thermal emissive bands centered at 3.7, 11, and 12 µm with a spatial resolution of 1 km at nadir. In the initial stages of reprocessing, several instabilities in the MODIS SST time series were observed. In particular, Terra SSTs and corresponding BTs showed three ‘steps’: two on 30 October 2000 and 2 July 2001 (due to changes in the MODIS operating mode) and one on 25 April 2020 (due to a change in its nominal blackbody temperature, BBT, from 290 to 285 K). Additionally, spikes up to several tenths of a kelvin were observed during the quarterly warm-up/cool-down (WUCD) exercises, when the Terra MODIS BBT was varied. Systematic gradual drifts of ~0.025 K/decade were also seen in both Aqua and Terra SSTs over their full missions due to drifting BTs. These calibration instabilities were mitigated by debiasing MODIS BTs using the time series of observed minus modeled (‘O-M’) BTs. The RAN1 dataset was evaluated via comparisons with various in situ SSTs. The data meet the NOAA specifications for accuracy (±0.2 K) and precision (0.6 K), often by a wide margin, in a clear-sky ocean domain of 19–21%. The long-term SST drift is typically less than 0.01 K/decade for all MODIS SSTs, except for the daytime ‘subskin’ SST, for which the drift is ~0.02 K/decade. The MODIS RAN1 dataset is archived at NOAA CoastWatch and updated monthly in a delayed mode with a latency of two months. Additional archival with NASA JPL PO.DAAC is being discussed. Full article
(This article belongs to the Special Issue VIIRS 2011–2021: Ten Years of Success in Earth Observations)
Show Figures

Figure 1

Figure 1
<p>Time series of the ΔLEXT (delta between the actual satellite LEXTs and their nominal 1:30 a.m./p.m. and 10:30 a.m./p.m. values). The three vertical dotted lines mark key dates: 1 January 2002 (approximate end of the Terra LEXT drift from 10:45 to 10:30 a.m./p.m.) and 27 February 2020/18 March 2021 (last Terra/Aqua orbit correction maneuvers). Note that the monthly mean (on the first day of the month; shown by symbols) LEXTs are calculated using the ‘pyorbital’ python package [<a href="#B21-remotesensing-15-05589" class="html-bibr">21</a>] as the average of all ascending nodes in a day using two line elements (TLEs) from <a href="http://celestrak.com" target="_blank">celestrak.com</a> (accessed on 17 May 2023).</p>
Full article ">Figure 2
<p>Monthly (June 2016) 0.1° aggregated maps of the mean sensitivity in ACSPO v2.80 Terra MODIS (<b>a</b>) daytime and (<b>b</b>) nighttime ‘subskin’ SSTs. During this time of year, the Northern Hemisphere atmosphere is warmer and moister compared to that of the Southern Hemisphere, resulting in a noticeably lower sensitivity north of the equator for the daytime SST. At night, the sensitivity is more uniform and closer to 1 due to the use of the more transparent MWIR band 20 centered at 3.7 µm.</p>
Full article ">Figure 3
<p>Aqua MODIS night ‘subskin’ SST imagery over Chesapeake Bay on 21 August 2023. (<b>a</b>) All sky SST imagery with no overlay. Land is rendered in brown. (<b>b</b>) ACSM mask applied (gray). (<b>c</b>) ACSM and front indicator (black curves) overlaid. The imagery is taken from the NOAA ACSPO Regional Monitor for the SST (ARMS) online system [<a href="#B30-remotesensing-15-05589" class="html-bibr">30</a>].</p>
Full article ">Figure 4
<p>Time series of the monthly aggregated number of MODIS RAN1 nighttime matchups against (<b>a</b>) DTMs and (<b>b</b>) TMs only. Data are from the NOAA SQUAM online system [<a href="#B34-remotesensing-15-05589" class="html-bibr">34</a>].</p>
Full article ">Figure 5
<p>Time series of the monthly aggregated nighttime MODIS RAN1−DTM SSTs: (<b>a</b>,<b>b</b>) global mean biases (accuracy); (<b>c</b>,<b>d</b>) corresponding standard deviations (SDs; precision); (<b>a</b>,<b>c</b>) ‘subskin’ and (<b>b</b>,<b>d</b>) <b>‘</b>depth’ SSTs. The corresponding temporal mean and standard deviation values for each time series are given by µ and σ, respectively. Data are taken from the NOAA SQUAM online system [<a href="#B34-remotesensing-15-05589" class="html-bibr">34</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>d</b>) The same as <a href="#remotesensing-15-05589-f005" class="html-fig">Figure 5</a>, but for the daytime SST.</p>
Full article ">Figure 7
<p>Twenty-four-hour aggregated mean biases of the Terra MODIS—CMC L4 foundation SST [<a href="#B37-remotesensing-15-05589" class="html-bibr">37</a>]. SST data are from 1 August 2019: (<b>a</b>,<b>c</b>) RAN1 ‘subskin’; (<b>b</b>,<b>d</b>) R2019 ‘skin’; (<b>a</b>,<b>b</b>) night; (<b>c</b>,<b>d</b>) day. A +0.17 K offset was added to the NASA ‘skin’ SST to facilitate its comparison with the ACSPO ‘subskin’ SST.</p>
Full article ">Figure 8
<p>Yearly (2019) aggregated histograms (binned at 0.1 K) of MODIS−DTM ‘subskin’ SSTs: (<b>a</b>,<b>b</b>) Terra; (<b>c</b>,<b>d</b>) Aqua; (<b>a</b>,<b>c</b>) night; (<b>b</b>,<b>d</b>) day. Each panel shows (blue) ACSPO ‘subskin’ SST and (red) NASA ‘skin’ SST + 0.17 K. The histograms are normalized with the total area under each curve = 1. For each histogram, the mean (µ) and standard deviation (σ) in Kelvin are listed in the legend.</p>
Full article ">Figure 9
<p>Time series of the monthly aggregated nighttime ΔT<sub>S</sub> = MODIS RAN1 − DTM SST mean biases: (<b>a</b>,<b>b</b>) night; (<b>c</b>,<b>d</b>) day; (<b>a</b>,<b>c</b>) ‘subskin’; (<b>b</b>,<b>d</b>) ‘depth’ SSTs. The seasonal signal was subtracted using the STL (Seasonal and Trend decomposition using Loess) algorithm [<a href="#B39-remotesensing-15-05589" class="html-bibr">39</a>]. The figure legends list the slopes and associated uncertainties (in units of K/decade) obtained using a linear least square fit with a 95% confidence interval. The lines correspond to linear fits.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>d</b>) Same as in <a href="#remotesensing-15-05589-f009" class="html-fig">Figure 9</a>, except against TMs instead of DTMs.</p>
Full article ">Figure 11
<p>(<b>a</b>) Twenty-four-hour aggregated ACSPO ‘subskin’—CMC L4 SST global mean biases; ‘O-M’ BT for bands (<b>b</b>) 20, (<b>c</b>) 22, and (<b>d</b>) 23 (blue circles) with and (red circles) without the mitigation of BTs. The two vertical dashed lines mark the dates when the Terra MODIS operating configuration changed from AA to BB (30 October 2000) and from BB to AA2 (2 July 2001). For easier viewing, the results are shifted vertically with the convention that time series with mitigation are centered on zero. For a further explanation, see the main text.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>f</b>) Monthly aggregated global mean biases of the ΔT<sub>S</sub> = ACSPO MODIS nighttime ‘subskin’ T<sub>SAT</sub> minus the (<b>a</b>,<b>b</b>) DTM, (<b>c</b>,<b>d</b>) TM, (<b>e</b>,<b>f</b>) AF T<sub>IS</sub>. (<b>g</b>,<b>h</b>) The corresponding ‘Aqua minus Terra’ double differences (DDs). Left panels (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>): Results with the original MODIS BTs. Right panels (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>): Results with the detrended MODIS BTs. To suppress noise, the DTM and TM time series were smoothed with a 7-month sliding window average, centered at current month. A 13-month window was used for the AFs, due to their two orders of magnitude fewer matchups.</p>
Full article ">Figure 13
<p>(<b>a</b>–<b>h</b>) Same as in <a href="#remotesensing-15-05589-f012" class="html-fig">Figure 12</a> but for daytime.</p>
Full article ">Figure 14
<p>(<b>a</b>–<b>c</b>): Monthly aggregated global mean nighttime Terra-Aqua <b>‘</b>O-M’ DDs for (<b>a</b>) MWIR and (<b>b</b>,<b>c</b>) LWIR bands used in ACSPO; (<b>d</b>): corresponding global mean difference in the Terra and Aqua nighttime ‘subskin’ ΔT<sub>S</sub>s. (Blue) detrended and (red) original MODIS BTs. Black and gray lines show the linear trends obtained from the original (uncorrected) MODIS ‘O-M’ DDs. Seasonal signals in Terra–Aqua ‘O-M’ DDs/ΔT<sub>S</sub>s due to different Aqua and Terra orbits, have been subtracted using the STL algorithm [<a href="#B39-remotesensing-15-05589" class="html-bibr">39</a>]. For easier viewing, the mean ‘O-M’ DDs/ΔT<sub>S</sub>s are offset vertically with the convention that the detrended results are centered at zero.</p>
Full article ">Figure 15
<p>(<b>a</b>–<b>d</b>) Same as in <a href="#remotesensing-15-05589-f014" class="html-fig">Figure 14</a> but for daytime.</p>
Full article ">Figure 16
<p>Nighttime monthly aggregated global mean MODIS-AATSR ΔT<sub>S</sub>s (June 2002–March 2012) and MODIS-VIIRS ΔT<sub>S</sub>s (February 2012–present) for (<b>a</b>) Terra; (<b>b</b>) Aqua. NPP/VIIRS [<a href="#B47-remotesensing-15-05589" class="html-bibr">47</a>] and MODIS data are ACSPO ‘subskin’ SSTs. AATSR is the ‘skin’ SST from the ESA CCI project [<a href="#B48-remotesensing-15-05589" class="html-bibr">48</a>,<a href="#B49-remotesensing-15-05589" class="html-bibr">49</a>]. Long-term/seasonal signals in ΔT<sub>S</sub>s (due to different satellite orbits, sensor spectral response functions, etc.) have been subtracted using the STL algorithm [<a href="#B39-remotesensing-15-05589" class="html-bibr">39</a>]. For easier viewing, the mean ΔT<sub>S</sub>s are offset vertically to center all detrended mean biases at zero.</p>
Full article ">Figure 17
<p>(<b>a</b>,<b>b</b>) The same as in <a href="#remotesensing-15-05589-f016" class="html-fig">Figure 16</a>, but for the daytime SST. See <a href="#remotesensing-15-05589-f016" class="html-fig">Figure 16</a>’s caption for more details.</p>
Full article ">Figure A1
<p>Time series of the monthly aggregated number of MODIS RAN1 SST matchups with AFs during the (<b>a</b>) day; and (<b>b</b>) night. Data are from the NOAA SQUAM online system [<a href="#B34-remotesensing-15-05589" class="html-bibr">34</a>].</p>
Full article ">Figure A2
<p>Times series of the monthly aggregated global night-time MODIS RAN1−AF SSTs: (<b>a</b>,<b>b</b>) mean biases; (<b>c</b>,<b>d</b>) corresponding standard deviations (SDs); (<b>a</b>,<b>c</b>) ‘subskin’; (<b>b</b>,<b>d</b>) ‘depth’ SST. Temporal means and standard deviations of the time series are given by µ and σ, respectively. Dates prior to 1 January 2003 are omitted due to the low NOBS. Data are taken from the NOAA SQUAM online system [<a href="#B34-remotesensing-15-05589" class="html-bibr">34</a>].</p>
Full article ">Figure A3
<p>(<b>a</b>–<b>d</b>) The same as in <a href="#remotesensing-15-05589-f0A2" class="html-fig">Figure A2</a>, but for daytime.</p>
Full article ">Figure A4
<p>(<b>a</b>) Terra MODIS BBT anomaly ΔT<sub>BB</sub> = T<sub>BB</sub> − T<sub>BB,nom</sub> (nominal BBT, T<sub>BB,nom</sub> = 290 K) during one WUCD exercise on 18–21 September 2009. Each data point represents an average ΔT<sub>BB</sub> with 10 min intervals. (<b>b</b>) Histogram showing the number of clear-sky MODIS-T L2P observations (NOBS) as a function of the BBT anomaly during all WUCD exercises from 3 July 2001 to 25 April 2020 (from the onset of AA2 configuration to the time when the nominal BBT was changed to 285 K). Observations with ΔT<sub>BB</sub> near 0 K are not shown in the histogram to avoid the saturation of the <span class="html-italic">y</span>-axis.</p>
Full article ">Figure A5
<p>Time series of 24 h aggregated global mean biases, ΔT<sub>S</sub> = ACSPO ‘subskin’—CMC L4 foundation SST bilinearly interpolated to the MODIS native grid: (blue squares) without and (red circles) with WUCD anomaly mitigation. The corresponding mean biases for NASA SSTs are also shown. A ±15 day running average was subtracted from the time series to remove slow seasonal variations and to normalize the <span class="html-italic">y</span>-axis to ~0 K to facilitate comparisons of the NASA ‘skin’ and ACSPO ‘subskin’ SST products. (<b>a</b>) Results for the year 2009, when the nominal BBT was set at 290 K and five WUCD exercises were performed (with beginning dates of 9 January, 3 April, 26 June, 18 September, and 11 December; denoted by vertical dashed lines). (<b>b</b>) Similar results, but for the year 2021, when the nominal BBT was set at 285 K and five WUCD exercises were performed (with beginning dates of 30 December 2020 and 24 March, 16 June, 8 September, and 1 December 2021; denoted by vertical dashed lines).</p>
Full article ">Figure A6
<p>‘O-M’ BTs stratified by the BBT anomaly for Terra MODIS (<b>top</b>) MWIR and (<b>bottom</b>) LWIR bands (<b>a</b>) 20 (3.7 µm), (<b>b</b>) 22 (3.9 µm), (<b>c</b>) 23 (4.0 µm), (<b>d</b>) 29 (8.6 µm), (<b>e</b>) 31 (11 µm), and (<b>f</b>) 32 (12 µm). Black dotted curves: clear-sky ‘O-M’ BTs from all WUCD cycles from 3 July 2001 to 25 April 2020 (last date before the MODIS nominal BBT was changed from 290 to 285 K). Red curves: model results using Equation (A1) with the coefficients listed in <a href="#remotesensing-15-05589-t0A2" class="html-table">Table A2</a>. See text for more details.</p>
Full article ">Figure A7
<p>(<b>a</b>–<b>f</b>) The same as in <a href="#remotesensing-15-05589-f0A6" class="html-fig">Figure A6</a>, except only for WUCD exercises after the switch to a nominal BBT of 285 K (26 April 2020–31 December 2022).</p>
Full article ">Figure A8
<p>Time series of the daily aggregated global mean biases (<b>a</b>) ΔT<sub>S</sub> = Terra ACSPO nighttime ‘subskin’ and NASA ‘skin’ SSTs minus the CMC L4 foundation SST. The results are shown with and without BT discontinuity mitigation. Terra MODIS ‘O-M’ BTs in bands (<b>b</b>) 20, (<b>c</b>) 22, and (<b>d</b>) 23. The two vertical dashed lines denote 22 April (start of the WUCD exercise) and 25 April (end of the WUCD exercise and first date with a nominal BBT of 285 K). For ease of viewing, the ACSPO SST and ‘O-M’ time series are shifted vertically by a constant value, such that the mean bias of the time series with BT discontinuity mitigation is centered at zero. The NASA SST time series were shifted, such that mean bias prior to 22 April is centered at zero. Note that the time series with mitigation have also been corrected for WUCD BT biases in addition to BT discontinuity mitigation.</p>
Full article ">
5 pages, 2204 KiB  
Proceeding Paper
The Impact of Spatial Resolution on Active Fire Monitoring Using Multispectral Satellite Imagery
by Andrea Gonnelli, Stefano Baronti, Roberto Carlà and Valentina Raimondi
Eng. Proc. 2023, 51(1), 30; https://doi.org/10.3390/engproc2023051030 - 6 Nov 2023
Viewed by 733
Abstract
Several studies have already evaluated the ability to identify wildfires and their characteristics by using NOAA-AVHRR and EO-MODIS images, which have adequate spectral bands for these targets, although with a spatial resolution limited to 1 km and a daily revisit time. In many [...] Read more.
Several studies have already evaluated the ability to identify wildfires and their characteristics by using NOAA-AVHRR and EO-MODIS images, which have adequate spectral bands for these targets, although with a spatial resolution limited to 1 km and a daily revisit time. In many cases, the latter features can limit the timely identification of a fire and the monitoring of its evolution. Conversely, sensors operating on geostationary platforms could acquire images within less than half an hour, yet still with a nominal spatial resolution of 1 km. In this study, we perform an analysis at different spatial resolutions of a sequence of OLI-Landsat-8 images referring to a natural fire that occurred near Massarosa, Tuscany, in July 2022. In particular, we investigate the potential of the SWIR bands, which are useful for monitoring high temperature wildfires. The results suggest that the use of sensors onboard a geostationary platform with relatively high nominal spatial resolution (of the order of 1 km) and frequent revisit time could enable the timely detection of fires and their monitoring. Full article
Show Figures

Figure 1

Figure 1
<p>Image of the study area located close to <span class="html-italic">Massarosa</span> (43°52′ N 10°20′ E), Italy.</p>
Full article ">Figure 2
<p>Landsat images: (<b>a</b>) image acquired before the fire (18 July 2022); (<b>b</b>) image acquired during the development of the fire (19 July 2022); (<b>c</b>) image acquired a few days after the end of the event (27 July 2022). RGB composition = (SWIR<sub>2200</sub>, NIR, SWIR<sub>1610</sub>).</p>
Full article ">Figure 3
<p>SWIR band radiance along a transect crossing the fire for different spatial resolutions: (<b>a</b>) SWIR band at 1610 nm and (<b>b</b>) SWIR band at 2200 nm. Radiance was normalized at 3σ<sub>B</sub>.</p>
Full article ">
16 pages, 3213 KiB  
Article
Remote Sensing Classification of Temperate Grassland in Eurasia Based on Normalized Difference Vegetation Index (NDVI) Time-Series Data
by Xuefeng Xu, Jiakui Tang, Na Zhang, Anan Zhang, Wuhua Wang and Qiang Sun
Sustainability 2023, 15(20), 14973; https://doi.org/10.3390/su152014973 - 17 Oct 2023
Viewed by 1114
Abstract
The Eurasian temperate grassland is the largest temperate grassland ecosystem and vegetation transition zone globally. The spatiotemporal distribution and changes of grassland types are vital for grassland monitoring and management. However, there is currently a lack of a unified classification method and standard [...] Read more.
The Eurasian temperate grassland is the largest temperate grassland ecosystem and vegetation transition zone globally. The spatiotemporal distribution and changes of grassland types are vital for grassland monitoring and management. However, there is currently a lack of a unified classification method and standard distribution map of Eurasian temperate grassland types. The Normalized Difference Vegetation Index (NDVI) from remote sensing data is commonly used in grassland monitoring. In this paper, the Accumulated Rate of NDVI Change Index (ARNCI) was proposed to characterize the annual NDVI trend of different temperate grassland types, and four transitional categories were introduced to account for the overlap between them. Based on survey data on the distribution of Eurasian temperate grassland types in the 1980s, the study area was divided into three sub-regions: Northern China, Central Asia, and Mongolia. Regionally, pixel-based ARNCI maps in the 1980s and 1990s were successfully calculated from using NOAA’s AVHRR NDVI time-series products. The ARNCI classification thresholds for different sub-regions were determined, and classification experiments and validation were conducted for each sub-region. The overall accuracies of grasslands types classification for Northern China, Central Asia, and Mongolia in the 1980s were 75.3%, 64.2%, and 84.6%, respectively, which demonstrated that there were variations in classification accuracy in the three sub-regions, and the overall performance was favorable. Finally, distribution maps of Eurasian temperate grassland types in the 1980s and 1990s were obtained, and the spatiotemporal changes of grassland types were analyzed and discussed. The ARNCI method is simple to operate and easy to obtain data, and it can be conveniently used in grassland type classification. The maps firstly address the lack of remote sensing classification maps of Eurasian temperate grassland types, and provide a promising tool for monitoring grassland degradation, management, and utilization. Full article
(This article belongs to the Special Issue Remote Sensing Monitoring of Resources and Ecological Environment)
Show Figures

Figure 1

Figure 1
<p>Coverage of temperate grassland types in Eurasia. The figure on the left shows the location of the study area on a global scale; the figure on the right presents the ground truth data in the 1980s.</p>
Full article ">Figure 2
<p>Monthly variation of NDVI in Inner Mongolia in 1989.</p>
Full article ">Figure 3
<p>Monthly mean NDVI of different grassland types in Northern China in 1989.</p>
Full article ">Figure 4
<p>ARNCI maps of Northern China in the 1980s and 1990s: (<b>a</b>) showing the ARNCI distribution of Northern China in the 1980s; and (<b>b</b>) showing the ARNCI distribution of Northern China in the 1990s.</p>
Full article ">Figure 5
<p>Pixel-based ARNCI statistic of different grassland types in Northern China in the 1980s.</p>
Full article ">Figure 6
<p>Maps of temperate grasslands types in Eurasia in the 1980s and 1990s: (<b>a</b>) 1980s; (<b>b</b>) 1990s.</p>
Full article ">Figure 7
<p>Eurasian temperate grassland areas and Sankey maps in the 1980s and 1990s. The letters (<b>a</b>,<b>b</b>,<b>c</b>) represent the cover area, area changes, and Sankey diagrams of grassland types, respectively, in different regions. The hyphenated numbers (1), (2), and (3) represent Northern China, Central Asia, and Mongolia, respectively.</p>
Full article ">
17 pages, 5704 KiB  
Article
Effects of Climate Variability and Human Activities on Vegetation Dynamics across the Qinghai–Tibet Plateau from 1982 to 2020
by Yiyang Liu, Yaowen Xie, Zecheng Guo and Guilin Xi
Remote Sens. 2023, 15(20), 4988; https://doi.org/10.3390/rs15204988 - 16 Oct 2023
Cited by 2 | Viewed by 1466
Abstract
In recent years, vegetation on the Qinghai–Tibet Plateau (QTP) has undergone significant greening. However, the causal factors underpinning this phenomenon, whether attributable to temperature fluctuations, precipitation patterns, or anthropogenic interventions, remain a subject of extensive scholarly debate. This study conducted a comprehensive analysis [...] Read more.
In recent years, vegetation on the Qinghai–Tibet Plateau (QTP) has undergone significant greening. However, the causal factors underpinning this phenomenon, whether attributable to temperature fluctuations, precipitation patterns, or anthropogenic interventions, remain a subject of extensive scholarly debate. This study conducted a comprehensive analysis of the evolving vegetation across the QTP. The National Oceanic and Atmospheric Administration Climate Data Record Advanced Very High Resolution Radiometer Normalized Vegetation Difference Index (NOAA CDR AVHRR NDVI) dataset was employed to elucidate the intricate relationship between climatic variables and human activities driving vegetative transformations. The findings were as follows: The NDVI on the QTP has exhibited a significant greening trend at a rate of 0.0013/a (per year). A minor decline, accounting for only 17.6% of grasslands, was observed, which was primarily concentrated in the northwestern and northern regions. Through residual analysis, climate change was found to be the predominant driver, explaining 70.6% of the vegetation variability across the plateau. Concurrently, noticeable trends in temperature and precipitation increases were observed on the QTP, with the southern region demonstrating improved sensitivity to precipitation alterations. In summary, these results substantiate that a confluence of climatic warming, enhanced moisture availability, and a reduction in livestock population collectively creates an environment conducive to enhanced vegetation vigor on the QTP. This study highlights the significance of acknowledging the dual influence of climate and human agency in shaping vegetative dynamics, which is a critical consideration for informed land management strategies and sustainable development initiatives on this ecologically pivotal plateau. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>The vegetation type (<b>a</b>) and DEM (<b>b</b>) on the QTP in China.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) Spatial distributions of correlation coefficient NOAA CDR AVHRR NDVI and MODIS NDVI, NOAA CDR AVHRR NDVI, and GIMMS3g NDVI.</p>
Full article ">Figure 3
<p>Annual NDVI (<b>a</b>) and NDVI in each season (<b>b</b>) from 1982 to 2020. The spring is March to May, the summer is June to August, the autumn is September to November, and the winter is December to February in the next year.</p>
Full article ">Figure 4
<p>Spatial distribution of interannual NDVI trends during 1982–2020 on the QTP. The upper panel shows NDVI trend for each 1° longitude bin, right panel shows the longitudinal gradient.</p>
Full article ">Figure 5
<p>The statistical trends of vegetation in different zones: (<b>a</b>) eco-climatic zones; (<b>b</b>) vegetation land covers (agriculture: all kinds of cropland, forest: broad-leaved and needle-leaved tree cover, grassland: mosaic herbaceous covered grassland and grassland, wetland: shrub or herbaceous covered water, settlement: urban, sparse vegetation: liches and sparse vegetation, bare: bare areas, water: water, other forest: tree cover with flooded); and (<b>c</b>) 500 m interval altitude zones, as well as statistical proportions of areas for different regions (<b>d</b>–<b>f</b>).</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) Spatial distribution of interannual <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> <mi>D</mi> <mi>V</mi> <mi>I</mi> </mrow> <mrow> <mi>h</mi> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> <mi>D</mi> <mi>V</mi> <mi>I</mi> </mrow> <mrow> <mi>c</mi> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> trends during 1982–2020. (<b>c</b>,<b>d</b>) Spatial distribution of relative contribution rates in HA and CC to the annual NDVI change on the QTP.</p>
Full article ">Figure 7
<p>(<b>a</b>,<b>b</b>) Contributions and contribution rates of CC and HA to NDVI change at the city scale.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>b</b>) Temporal trends during 1979–2018 of mean annual temperature and annual precipitation. (<b>c</b>,<b>d</b>) Spatial distributions of correlation coefficient between NDVI and mean annual temperature and annual precipitation on the QTP.</p>
Full article ">Figure 9
<p>Spatial distribution of interannual NDVI trends during 2000–2020 on the Naqu, from NOAA CDR AVHRR NDVI (<b>a</b>) and MODIS NDVI (<b>b</b>). The smaller plots at the top right of the figure show the slopes of the annual changes in NDVI for the same region.</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>b</b>) GDP and population density over the QTP in 2015. (<b>c</b>,<b>d</b>) Their increase rate during 1995–2019.</p>
Full article ">Figure 11
<p>(<b>a</b>) The population and the production of three industries over the QTP during 1970–2020. NDVI, livestock population, temperature, and precipitation of the Tibet (<b>b</b>,<b>c</b>) and Qinghai provinces (<b>e</b>,<b>f</b>) over time series. (<b>d</b>) The overview of the administrative boundary between Tibet and Qinghai provinces.</p>
Full article ">
20 pages, 9266 KiB  
Article
Retrieval of Sea Surface Skin Temperature from the High Resolution Picture Transmission Data of the National Oceanic and Atmospheric Administration Series Satellites
by Yan Chen, Liqin Qu, Zhuomin Li and Lei Guan
Remote Sens. 2023, 15(15), 3723; https://doi.org/10.3390/rs15153723 - 26 Jul 2023
Viewed by 961
Abstract
The High Resolution Picture Transmission (HRPT) data of the National Oceanic and Atmospheric Administration (NOAA) series meteorological satellites had been received by the SeaSpace ground station located at the Ocean University of China (OUC). Based on the atmospheric radiative transfer model, we obtained [...] Read more.
The High Resolution Picture Transmission (HRPT) data of the National Oceanic and Atmospheric Administration (NOAA) series meteorological satellites had been received by the SeaSpace ground station located at the Ocean University of China (OUC). Based on the atmospheric radiative transfer model, we obtained the NOAA-15/16/17/18/19 Advanced Very High Resolution Radiometer (AVHRR) sea surface skin temperature (SSTskin) data using the Bayesian cloud detection method and the optimal estimation (OE) sea surface temperature (SST) retrieval algorithm. Compared with the NOAA/AVHRR multi-channel SST data, the AVHRR SSTskin data have higher data accuracy. We also compared the AVHRR SSTskin with the buoy SST with spatial and temporal windows of 0.01° and 30 min. The daytime biases ranged from −0.32 °C (NOAA-16) to 0.08 °C (NOAA-17) with standard deviations (SDs) ranging from 0.36 °C (NOAA-18/ NOAA-19) to 0.60 °C (NOAA-16), and the nighttime biases ranged from −0.26 °C (NOAA-16) to −0.02 °C (NOAA-17) with SDs ranging from 0.33 °C (NOAA-19) to 0.60 °C (NOAA-16). The accuracy of all five satellite data during daytime and nighttime was significantly improved. These results show that the AVHRR SSTskin of NOAA series satellites is good and consistent in different periods, and the SSTskin data products with high spatial resolution and accuracy can be used for mesoscale and submesoscale marine applications. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>RTTOV simulated BTs of the 11 and 12 μm channels for NOAA/AVHRR: (<b>a</b>) NOAA-15 11 μm channel; (<b>b</b>) NOAA-15 12 μm channel; (<b>c</b>) NOAA-16 11μm channel; (<b>d</b>) NOAA-16 12 μm channel; (<b>e</b>) NOAA-17 11μm channel; (<b>f</b>) NOAA-17 12μm channel; (<b>g</b>) NOAA-18 11 μm channel; (<b>h</b>) NOAA-18 12 μm channel; (<b>i</b>) NOAA-19 11 μm channel; (<b>j</b>) NOAA-19 12 μm channel. (The gray areas indicate land, and the white areas indicate areas out of the satellite coverage.)</p>
Full article ">Figure 1 Cont.
<p>RTTOV simulated BTs of the 11 and 12 μm channels for NOAA/AVHRR: (<b>a</b>) NOAA-15 11 μm channel; (<b>b</b>) NOAA-15 12 μm channel; (<b>c</b>) NOAA-16 11μm channel; (<b>d</b>) NOAA-16 12 μm channel; (<b>e</b>) NOAA-17 11μm channel; (<b>f</b>) NOAA-17 12μm channel; (<b>g</b>) NOAA-18 11 μm channel; (<b>h</b>) NOAA-18 12 μm channel; (<b>i</b>) NOAA-19 11 μm channel; (<b>j</b>) NOAA-19 12 μm channel. (The gray areas indicate land, and the white areas indicate areas out of the satellite coverage.)</p>
Full article ">Figure 1 Cont.
<p>RTTOV simulated BTs of the 11 and 12 μm channels for NOAA/AVHRR: (<b>a</b>) NOAA-15 11 μm channel; (<b>b</b>) NOAA-15 12 μm channel; (<b>c</b>) NOAA-16 11μm channel; (<b>d</b>) NOAA-16 12 μm channel; (<b>e</b>) NOAA-17 11μm channel; (<b>f</b>) NOAA-17 12μm channel; (<b>g</b>) NOAA-18 11 μm channel; (<b>h</b>) NOAA-18 12 μm channel; (<b>i</b>) NOAA-19 11 μm channel; (<b>j</b>) NOAA-19 12 μm channel. (The gray areas indicate land, and the white areas indicate areas out of the satellite coverage.)</p>
Full article ">Figure 2
<p>NOAA/AVHRR 11 μm channel BTs before and after Bayesian cloud detection: (<b>a</b>) NOAA-15 before cloud detection; (<b>b</b>) NOAA-15 after cloud detection; (<b>c</b>) NOAA-16 before cloud detection; (<b>d</b>) NOAA-16 after cloud detection; (<b>e</b>) NOAA-17 before cloud detection; (<b>f</b>) NOAA-17 after cloud detection; (<b>g</b>) NOAA-18 before cloud detection; (<b>h</b>) NOAA-18 after cloud detection; (<b>i</b>) NOAA-19 before cloud detection; (<b>j</b>) NOAA-19 after cloud detection. (The gray areas indicate land, and the white areas indicate cloudy areas or areas out of satellite coverage.)</p>
Full article ">Figure 2 Cont.
<p>NOAA/AVHRR 11 μm channel BTs before and after Bayesian cloud detection: (<b>a</b>) NOAA-15 before cloud detection; (<b>b</b>) NOAA-15 after cloud detection; (<b>c</b>) NOAA-16 before cloud detection; (<b>d</b>) NOAA-16 after cloud detection; (<b>e</b>) NOAA-17 before cloud detection; (<b>f</b>) NOAA-17 after cloud detection; (<b>g</b>) NOAA-18 before cloud detection; (<b>h</b>) NOAA-18 after cloud detection; (<b>i</b>) NOAA-19 before cloud detection; (<b>j</b>) NOAA-19 after cloud detection. (The gray areas indicate land, and the white areas indicate cloudy areas or areas out of satellite coverage.)</p>
Full article ">Figure 3
<p>NOAA/AVHRR SST<sub>skin</sub> retrieved by the OE SST algorithm: (<b>a</b>) NOAA-15 on 9 October 2006, at 21:09; (<b>b</b>) NOAA-16 on 9 October 2006, at 19:00; (<b>c</b>) NOAA-17 on 9 October 2006, at 02:36; (<b>d</b>) NOAA-18 on 9 October 2006, at 05:38; (<b>e</b>) NOAA-19 on 9 October 2009, at 04:45. (The gray areas indicate land, and the white areas indicate cloudy areas or areas out of the satellite coverage).</p>
Full article ">Figure 4
<p>Histogram of biases between NOAA/AVHRR SST<sub>skin</sub> and buoy SST: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 4 Cont.
<p>Histogram of biases between NOAA/AVHRR SST<sub>skin</sub> and buoy SST: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 5
<p>Scatterplots of biases between NOAA/AVHRR SST<sub>skin</sub> and buoy SSTs: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 5 Cont.
<p>Scatterplots of biases between NOAA/AVHRR SST<sub>skin</sub> and buoy SSTs: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 6
<p>Dependency of NOAA/AVHRR SST<sub>skin</sub> minus buoy SSTs on satellite zenith angle: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 6 Cont.
<p>Dependency of NOAA/AVHRR SST<sub>skin</sub> minus buoy SSTs on satellite zenith angle: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 7
<p>Dependency of NOAA/AVHRR SST<sub>skin</sub> minus buoy SSTs on TCWV: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">Figure 7 Cont.
<p>Dependency of NOAA/AVHRR SST<sub>skin</sub> minus buoy SSTs on TCWV: (<b>a</b>) NOAA-15, daytime; (<b>b</b>) NOAA-15, nighttime; (<b>c</b>) NOAA-16, daytime; (<b>d</b>) NOAA-16, nighttime; (<b>e</b>) NOAA-17, daytime; (<b>f</b>) NOAA-17, nighttime; (<b>g</b>) NOAA-18, daytime; (<b>h</b>) NOAA-18, nighttime; (<b>i</b>) NOAA-19, daytime; (<b>j</b>) NOAA-19, nighttime.</p>
Full article ">
18 pages, 4726 KiB  
Article
Characteristics of NDVI Changes in the Altay Region from 1981 to 2018 and Their Relationship to Climatic Factors
by Yang Yan, Junhui Cheng, Yongkang Li, Jie Fan and Hongqi Wu
Land 2023, 12(3), 564; https://doi.org/10.3390/land12030564 - 26 Feb 2023
Cited by 2 | Viewed by 1959
Abstract
Vegetation growth and its response to climatic factors have become one of the most pressing issues in ecological research. However, no consensus has yet been reached on how to resolve this problem in arid areas with a high-elevation gradient and complex underlying surface. [...] Read more.
Vegetation growth and its response to climatic factors have become one of the most pressing issues in ecological research. However, no consensus has yet been reached on how to resolve this problem in arid areas with a high-elevation gradient and complex underlying surface. Here, NOAA CDR AVHRR NDVI V5 for 1981–2018 and China’s regional surface meteorological faction-driven datasets were used. General linear regression, the Mann-Kendall test and sliding t-test, Pearson correlations, and the Akaike information criterion (AIC), on a grid-scale, were applied to analyze the annual normalized difference vegetation index (NDVI) and its relationship with temperature and precipitation in the Altay region. Results revealed that the temporal trend of NDVI for most grid cells was non-significant. However, mountains, coniferous forests, grasslands, and meadows in the high-elevation zone displayed a slow increasing trend in NDVI. Further, NDVI was positively correlated with the mean annual temperature and total annual precipitation, the latter playing a more significant role. Yet, for desert and shrub vegetation and coniferous forest, their NDVI had insignificant negative correlations with the mean annual temperature. Hence, both the trends and drivers of NDVI of high elevation are highly complex. This study’s findings provide a reference for research on vegetation responses to climate change in arid areas having a high-elevation gradients and complex underlying surfaces. Full article
Show Figures

Figure 1

Figure 1
<p>Summary maps of the study area: (<b>a</b>) geographical location and distribution of (<b>b</b>) its geomorphic types and (<b>c</b>) vegetation types.</p>
Full article ">Figure 2
<p>Altay region elevation distribution: (<b>a</b>) distribution of geomorphic types, (<b>b</b>) distribution of vegetation types.</p>
Full article ">Figure 2 Cont.
<p>Altay region elevation distribution: (<b>a</b>) distribution of geomorphic types, (<b>b</b>) distribution of vegetation types.</p>
Full article ">Figure 3
<p>Analysis workflow chart.</p>
Full article ">Figure 4
<p>Spatial distribution of changes in the NDVI of the Altay region, from 1981 to 2018. From the perspective of geomorphic types, the trend for NDVI was of increasing significantly mainly in mountainous areas (48.1% of all grid cells). For the other three landforms, the majority of their changes in NDVI consisted of non-significant trends, for which the proportion of grid cells was 56.45% (plains) to 68.1% (hills) (<a href="#land-12-00564-t002" class="html-table">Table 2</a>).</p>
Full article ">Figure 5
<p>Histograms for slopes of NDVI change in the Altay region from 1981 to 2018: (<b>a</b>) significant increase and decrease in the NDVI of geomorphic types, (<b>b</b>) significant increase and decrease in the NDVI of vegetation types.</p>
Full article ">Figure 6
<p>Histograms of the abrupt-change year of the NDVI ofof the Altay region from 1981 to 2018 for (<b>a</b>) both trends in the NDVI; for geomorphic types showing (<b>b</b>) unimodal trend and (<b>c</b>) U-shaped trend in the NDVI; for vegetation types showing (<b>d</b>) unimodal trend and (<b>e</b>) U-shaped trend in the NDVI.</p>
Full article ">Figure 7
<p>Spatial distribution of changes in (<b>a</b>) annual mean temperature and (<b>b</b>) total annual precipitation in the Altay region, from 1981 to 2018.</p>
Full article ">Figure 8
<p>Spatial distribution for correlations of the NDVI with annual mean temperature and precipitation in the Altay region from 1981 to 2018: (<b>a</b>) between the NDVI and annual mean temperature, (<b>b</b>) between NDVI and total annual precipitation.</p>
Full article ">
6 pages, 1140 KiB  
Proceeding Paper
Spatial Dynamics of Tree Stand Disturbance under Siberian Silk Moth (Dendrolimus sibiricus) Impact in Central Siberia in 2016–2020 Based on Remote Sensing Data
by Evgenii I. Ponomarev, Andrey A. Goroshko, Evgeny G. Shvetsov, Nikita D. Yakimov, Pavel D. Tretyakov, Svetlana M. Sultson and Pavel V. Mikhaylov
Environ. Sci. Proc. 2022, 22(1), 4; https://doi.org/10.3390/IECF2022-13056 - 15 Oct 2022
Cited by 1 | Viewed by 782
Abstract
In this study, we have analyzed the spatial dynamics of the forests disturbed by Siberian Silk Moth (Dendrolimus sibiricus Tschetverikov (Lepidoptera: Lasiocampidae)) in Central Siberia and obtained model equations that fit these dynamics. We considered three sites that experienced silk [...] Read more.
In this study, we have analyzed the spatial dynamics of the forests disturbed by Siberian Silk Moth (Dendrolimus sibiricus Tschetverikov (Lepidoptera: Lasiocampidae)) in Central Siberia and obtained model equations that fit these dynamics. We considered three sites that experienced silk moth outbreaks in 1993–1996, 2015–2018, and 2018–2020 and used satellite data (NOAA/AVHRR, Terra/MODIS, Landsat/ETM/OLI), field data, a digital elevation model, and maps of predominant forests. Silk moth-disturbed areas were classified using NDVI, which was calculated for each 15-day period during the growing season (April–September). Time series of disturbed forest areas were obtained for three sites located in the Krasnoyarsk region (Central Siberia, Russia). Total damaged areas for these sites were 41, 430, and 470 thousand hectares. We obtained formalized descriptions for the temporal dynamics of the disturbed area. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of forests disturbed by silk moth, in (<bold>a</bold>) EnP for 2020; (<bold>b</bold>) in IrP for 2020.</p>
Full article ">Figure 2
<p>The dynamics of area disturbed by silk moth in EnP (<bold>a</bold>) and in IrP (<bold>b</bold>).</p>
Full article ">Figure 3
<p>The dynamics of the area disturbed by silk moth for EnP в 2015–2018 гг. (<bold>a</bold>) and IrP в 2018–2020 гг. (<bold>d</bold>), ASTER GDEM [<xref ref-type="bibr" rid="B26-environsciproc-22-00004">26</xref>] is shown as background. Model of the monthly distribution of the area disturbed by silk moth for EnP, which approximates the Terra/MODIS data with 14 days temporal resolution, where I corresponds to exponential growth (phase “I”) and II is the approximating of the final stage. For IrP (<bold>b</bold>), 1 is the experimental data, 2 is an approximation the maximum of the final area, 3 is the model solution for the phase of exponential area growth, and 4 is the model solution for the phase of decreasing rate of area growth (logarithmic approximation) and further saturation (<bold>e</bold>). Plot of the correlation field for experimental measurement data (based on satellite data) and model calculation (S<sub>Model</sub>) results for EnP (<bold>c</bold>) and for IrP (<bold>f</bold>).</p>
Full article ">
13 pages, 4569 KiB  
Article
Varying Responses of Vegetation Greenness to the Diurnal Warming across the Global
by Jie Zhao, Kunlun Xiang, Zhitao Wu and Ziqiang Du
Plants 2022, 11(19), 2648; https://doi.org/10.3390/plants11192648 - 8 Oct 2022
Cited by 9 | Viewed by 1372
Abstract
The distribution of global warming has been varying both diurnally and seasonally. Little is known about the spatiotemporal variations in the relationships between vegetation greenness and day- and night-time warming during the last decades. We investigated the global inter- and intra-annual responses of [...] Read more.
The distribution of global warming has been varying both diurnally and seasonally. Little is known about the spatiotemporal variations in the relationships between vegetation greenness and day- and night-time warming during the last decades. We investigated the global inter- and intra-annual responses of vegetation greenness to the diurnal asymmetric warming during the period of 1982–2015, using the normalized different vegetation index (NDVI, a robust proxy for vegetation greenness) obtained from the NOAA/AVHRR NDVI GIMMS3g dataset and the monthly average daily maximum (Tmax) and minimum temperature (Tmin) obtained from the gridded Climate Research Unit, University of East Anglia. Several findings were obtained: (1) The strength of the relationship between vegetation greenness and the diurnal temperature varied on inter-annual and seasonal timescales, indicating generally weakening warming effects on the vegetation activity across the global. (2) The decline in vegetation response to Tmax occurred mainly in the mid-latitudes of the world and in the high latitudes of the northern hemisphere, whereas the decline in the vegetation response to Tmin primarily concentrated in low latitudes. The percentage of areas with a significantly negative trend in the partial correlation coefficient between vegetation greenness and diurnal temperature was greater than that of the areas showing the significant positive trend. (3) The trends in the correlation between vegetation greenness and diurnal warming showed a complex spatial pattern: the majority of the study areas had undergone a significant declining strength in the vegetation greenness response to Tmax in all seasons and to Tmin in seasons except autumn. These findings are expected to have important implications for studying the diurnal asymmetry warming and its effect on the terrestrial ecosystem. Full article
Show Figures

Figure 1

Figure 1
<p>Temporal variations in the partial correlation coefficients between mean annual NDVI and the diurnal temperature (T<sub>max</sub> and T<sub>min</sub>) for each 17-year moving window across latitudes intervals. (<b>a</b>–<b>e</b> represents latitudes intervals at 60°~90° N, 30°~60° N, 0°~30° N, −30°~0° S, and −60°~0° S, respectively (N and S indicates the Northern and southern hemisphere, respectively). The x axis is the last year of the 17-year moving-window (for example, 1998 stands for a moving-window from 1982 to 1998, …, 2015 stands for a moving-window from 1999 to 2015). The Y axis is the partial correlation coefficients).</p>
Full article ">Figure 2
<p>Temporal variations in the partial correlation coefficients between mean NDVI and the diurnal temperature (T<sub>max</sub> and T<sub>min</sub>) in spring for each 17-year moving window across latitudes intervals. (<b>a</b>–<b>e</b> represents latitudes intervals at 60°~90° N, 30°~60° N, 0°~30° N, −30°~0° S, and −60°~0° S, respectively (N and S indicates the Northern and southern hemisphere, respectively). The x axis is the last year of the 17-year moving-window (for example, 1998 stands for a moving-window from 1982 to 1998,…, 2015 stands for a moving-window from 1999 to 2015). The Y axis is the partial correlation coefficients).</p>
Full article ">Figure 3
<p>Temporal variations in the partial correlation coefficients between mean NDVI and the diurnal temperature (T<sub>max</sub> and T<sub>min</sub>) in summer for each 17-year moving window across latitudes intervals. (<b>a</b>–<b>e</b> represents latitudes intervals at 60°~90° N, 30°~60° N, 0°~30° N, −30°~0° S, and −60°~0° S, respectively (N and S indicates the Northern and southern hemisphere, respectively). The x axis is the last year of the 17-year moving-window (for example, 1998 stands for a moving-window from 1982 to 1998,…, 2015 stands for a moving-window from 1999 to 2015). The Y axis is the partial correlation coefficients).</p>
Full article ">Figure 4
<p>Temporal variations in the partial correlation coefficients between mean NDVI and the diurnal temperature (T<sub>max</sub> and T<sub>min</sub>) in autumn for each 17-year moving window across latitudes intervals. (<b>a</b>–<b>e</b> represents latitudes intervals at 60°~90° N, 30°~60° N, 0°~30° N, −30°~0° S, and −60°~0° S, respectively (N and S indicates the Northern and southern hemisphere, respectively). The x axis is the last year of the 17-year moving-window (for example, 1998 stands for a moving-window from 1982 to 1998,…, 2015 stands for a moving-window from 1999 to 2015). The Y axis is the partial correlation coefficients).</p>
Full article ">Figure 5
<p>Temporal variations in the partial correlation coefficients between mean NDVI and the diurnal temperature (T<sub>max</sub> and T<sub>min</sub>) in winter for each 17-year moving window across latitudes intervals. (<b>a</b>–<b>e</b> represents latitudes intervals at 60°~90° N, 30°~60° N, 0°~30° N, −30°~0° S, and −60°~0° S, respectively (N and S indicates the Northern and southern hemisphere, respectively). The x axis is the last year of the 17-year moving-window (for example, 1998 stands for a moving-window from 1982 to 1998,…, 2015 stands for a moving-window from 1999 to 2015). The Y axis is the partial correlation coefficients).</p>
Full article ">Figure 6
<p>The response of vegetation greenness to the diurnal temperature. (<b>a</b>. spatial distribution of the temporal trend of the partial coefficients between mean annual NDVI and T<sub>max</sub>. <b>b</b>. spatial distribution of the temporal trend of the partial coefficients between mean annual NDVI and T<sub>min</sub>. <a href="#app1-plants-11-02648" class="html-app">Supplementary Table S1</a>).</p>
Full article ">Figure 7
<p>The response of vegetation greenness to the diurnal temperature in spring. (<b>a</b>. spatial distribution of the temporal trend of the partial coefficients between spring NDVI and T<sub>max</sub>. <b>b</b>. spatial distribution of the temporal trend of the partial coefficients between spring NDVI and T<sub>min</sub>. <a href="#app1-plants-11-02648" class="html-app">Supplementary Table S2</a>).</p>
Full article ">Figure 8
<p>The response of vegetation greenness to the diurnal temperature in summer. (<b>a</b>. spatial distribution of the temporal trend of the partial coefficients between summer NDVI and T<sub>max</sub>. <b>b</b>. spatial distribution of the temporal trend of the partial coefficients between summer NDVI and T<sub>min</sub>. <a href="#app1-plants-11-02648" class="html-app">Supplementary Table S3</a>).</p>
Full article ">Figure 9
<p>The response of vegetation greenness to the diurnal temperature in autumn. (<b>a</b>. spatial distribution of the temporal trend of the partial coefficients between autumn NDVI and T<sub>max</sub>. <b>b</b>. spatial distribution of the temporal trend of the partial coefficients between autumn NDVI and T<sub>min</sub>. <a href="#app1-plants-11-02648" class="html-app">Supplementary Table S4</a>).</p>
Full article ">Figure 10
<p>The response of vegetation greenness to the diurnal temperature in winter. (<b>a</b>. spatial distribution of the temporal trend of the partial coefficients between winter NDVI and T<sub>max</sub>. <b>b</b>. spatial distribution of the temporal trend of the partial coefficients between winter NDVI and T<sub>min</sub>. <a href="#app1-plants-11-02648" class="html-app">Supplementary Table S5</a>).</p>
Full article ">
31 pages, 3252 KiB  
Article
Multi-Sensor Sea Surface Temperature Products from the Australian Bureau of Meteorology
by Pallavi Devidas Govekar, Christopher Griffin and Helen Beggs
Remote Sens. 2022, 14(15), 3785; https://doi.org/10.3390/rs14153785 - 6 Aug 2022
Cited by 7 | Viewed by 2462
Abstract
Sea surface temperature (SST) products that can resolve fine scale features, such as sub-mesoscale eddies, ocean fronts and coastal upwelling, are increasingly in demand. In response to user requirements for gap-free, highest spatial resolution, best quality and highest accuracy SST data, the Australian [...] Read more.
Sea surface temperature (SST) products that can resolve fine scale features, such as sub-mesoscale eddies, ocean fronts and coastal upwelling, are increasingly in demand. In response to user requirements for gap-free, highest spatial resolution, best quality and highest accuracy SST data, the Australian Bureau of Meteorology (BoM) produces operational, real-time Multi-sensor SST level 3 products by compositing SST from Advanced Very-High-Resolution Radiometer (AVHRR) sensors on Meteorological Operational satellite (MetOp)-B and National Oceanic and Atmospheric Administration (NOAA) 18, along with SST from Visible Infrared Imaging Radiometer Suite (VIIRS) sensors on the Suomi National Polar-orbiting Partnership (Suomi NPP) and NOAA 20 polar-orbiting satellites for the Australian Integrated Marine Observing System (IMOS) project. Here we discuss our method to combine data from different sensors and present validation of the satellite-derived SST against in situ SST data. The Multi-sensor Level 3 Super Collated (L3S) SSTs exhibit significantly greater spatial coverage and improved accuracy compared with the pre-existing IMOS AVHRR-only L3S SSTs. When compared to the Geo Polar Blended level 4 analysis SST data over the Great Barrier Reef, Multi-sensor L3S SST differs by less than 1 °C while exhibiting a wider range of SSTs over the region. It shows more variability and restores small-scale features better than the Geo Polar Blended level 4 analysis SST data. The operational Multi-sensor L3S SST products are used as input for applications such as IMOS OceanCurrent and the BoM ReefTemp Next-Generation Coral Bleaching Nowcasting service and provide useful insight into the study of marine heatwaves and ocean upwelling in near-coastal regions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of processing methods to composite data from different satellite sensors to construct Multi-sensor L3S SST products.</p>
Full article ">Figure 2
<p>Modified quality level (QL) as per <a href="#sec2dot2dot5-remotesensing-14-03785" class="html-sec">Section 2.2.5</a> for the fv02 L3C-1day daytime file for (<b>a</b>) NOAA-18 AVHRR, (<b>b</b>) MetOp-B AVHRR, (<b>c</b>) NPP VIIRS and (<b>d</b>) N20 VIIRS for 12 December 2018, obtained from [<a href="#B45-remotesensing-14-03785" class="html-bibr">45</a>].</p>
Full article ">Figure 3
<p>Modified quality level (QL) as per <a href="#sec2dot2dot5-remotesensing-14-03785" class="html-sec">Section 2.2.5</a> for the fv02 L3C-1day night-time file for (<b>a</b>) NOAA-18 AVHRR, (<b>b</b>) MetOp-B AVHRR, (<b>c</b>) NPP VIIRS and (<b>d</b>) N20 VIIRS for 12 December 2018, obtained from [<a href="#B46-remotesensing-14-03785" class="html-bibr">46</a>].</p>
Full article ">Figure 4
<p>Bias -corrected fv02 L3C-01day, night only, monthly statistics, 1 February 2017–31 December 2018, median for (<b>a</b>) QL 4 and (<b>b</b>) QL 5, and standard deviation for (<b>c</b>) QL 4 and (<b>d</b>) QL 5, number of matchups for (<b>e</b>) QL 4 and (<b>f</b>) QL 5, when compared with drifting and tropical moored buoy’s SST. Note: Matchup thresholds: &lt;10 km distance and &lt;6 h time difference. SSTs were obtained from sensors on NOAA–18 (orange), NOAA–20 (blue), NPP (green) and MetOp-B (pink) satellites from [<a href="#B46-remotesensing-14-03785" class="html-bibr">46</a>].</p>
Full article ">Figure 5
<p>Sea surface temperatures for fv02 L3S–1day from AVHRR-only (NOAA-18) (<b>a</b>) day, (<b>c</b>) night and (<b>e</b>) day and night, and Multi–sensor (NOAA–18, MetOp–B, NPP, N20) (<b>b</b>) day, (<b>d</b>) night and (<b>f</b>) day and night for 12 December 2018. The data shown are for QL ≥ 3. Data were accessed from [<a href="#B48-remotesensing-14-03785" class="html-bibr">48</a>,<a href="#B49-remotesensing-14-03785" class="html-bibr">49</a>].</p>
Full article ">Figure 6
<p>Bias-corrected fv02 L3S-1day, night only, monthly statistics, February 2017 to December 2018, median for (<b>a</b>) AVHRR-only and (<b>b</b>) Multi-sensor, and standard deviation for (<b>c</b>) AVHRR-only and (<b>d</b>) Multi-sensor, number of matchups for (<b>e</b>) AVHRR-only and (<b>f</b>) Multi-sensor, when compared with drifting and tropical moored buoys. Note: Matchup thresholds: &lt;10 km distance and &lt;6 h time difference. Statistics relating to QL 5 L3S data are shown in green, QL 4 in blue and QL 3 in orange. Data were obtained from [<a href="#B48-remotesensing-14-03785" class="html-bibr">48</a>,<a href="#B49-remotesensing-14-03785" class="html-bibr">49</a>].</p>
Full article ">Figure 7
<p>Bias-corrected fv02 L3C-1day and L3S-1day, QL ≥ 3, night only, monthly statistics, February 2017 to December 2018, (<b>a</b>) median and (<b>b</b>) standard deviation for NOAA–15 (green line), NOAA–18 (aqua line), NOAA–19 (violet line), NPP (grey line),NOAA-20 (red line), MetOp-B (pink line), HRPT AVHRR L3S (thick blue line) and Multi-sensor L3S (thick orange line), when compared with drifting and tropical moored buoys. Note: Matchup thresholds: &lt;10 km distance and &lt;6 h time difference. Data were obtained from [<a href="#B48-remotesensing-14-03785" class="html-bibr">48</a>,<a href="#B49-remotesensing-14-03785" class="html-bibr">49</a>].</p>
Full article ">Figure 8
<p>Great Barrier Reef Marine Park Extent of Park and catchments. <a href="https://www.transparency.gov.au/annual-reports/great-barrier-reef-marine-park-authority/reporting-year/2019-20-4" target="_blank">https://www.transparency.gov.au/annual-reports/great-barrier-reef-marine-park-authority/reporting-year/2019-20-4</a>, accessed on 10 April 2022. This study considered the region outlined by the red boundary.</p>
Full article ">Figure 9
<p>GBR Marine Park SST stacked plot of coverage (%) for (<b>a</b>) NOAA AVHRR versus Multi-sensor for QL ≥ 3 and (<b>b</b>) Multi-sensor by quality. Data were sourced from [<a href="#B48-remotesensing-14-03785" class="html-bibr">48</a>,<a href="#B49-remotesensing-14-03785" class="html-bibr">49</a>].</p>
Full article ">Figure 10
<p>A comparison of IMOS Multi-sensor L3S and Geo-Polar Blended L4 over GBR Marine Park (<b>a</b>) average and range for Multi-sensor L3S (QL ≥ 3), (<b>b</b>) average and range for Geo-polar Blended SST L4 Analysis and (<b>c</b>) SST discrepancy, SST Multi-sensor Geo-polar Blend by quality. Data were sourced from [<a href="#B49-remotesensing-14-03785" class="html-bibr">49</a>,<a href="#B66-remotesensing-14-03785" class="html-bibr">66</a>].</p>
Full article ">Figure 11
<p>A comparison of IMOS Multi-sensor L3S and Geo-Polar Blended L4 with attributions of variability over GBR Marine Park. SST Anomaly (SST-SSTAARS) (<b>a</b>) standard deviation variation variability, (<b>b</b>) standard deviation ratio, (<b>c</b>) variability ratio with different filter radii and (<b>d</b>) allocation of residual variability. Where the red line denotes combined variability, the blue line denotes exponential variability (0.042 ± 0.004C<math display="inline"><semantics> <msup> <mrow/> <mn>2</mn> </msup> </semantics></math>) on a ≈ 8 km scale and the orange line denotes sampling variability ≈<math display="inline"><semantics> <mrow> <mo>(</mo> <mn>0.04</mn> <mo>±</mo> <mn>0.02</mn> <msup> <mi>C</mi> <mn>2</mn> </msup> <mo>)</mo> <mo>/</mo> </mrow> </semantics></math>(sample size), where these all add to the green line denoting a residual constant variability (skin to depth variability) of ≈<math display="inline"><semantics> <mrow> <mn>0.021</mn> <mo>±</mo> <mn>0.002</mn> <msup> <mi>C</mi> <mn>2</mn> </msup> </mrow> </semantics></math>. Data were sourced from [<a href="#B49-remotesensing-14-03785" class="html-bibr">49</a>,<a href="#B66-remotesensing-14-03785" class="html-bibr">66</a>].</p>
Full article ">
26 pages, 7602 KiB  
Article
JPSS VIIRS SST Reanalysis Version 3
by Olafur Jonasson, Alexander Ignatov, Victor Pryamitsyn, Boris Petrenko and Yury Kihai
Remote Sens. 2022, 14(14), 3476; https://doi.org/10.3390/rs14143476 - 20 Jul 2022
Cited by 6 | Viewed by 3264
Abstract
The 3rd full-mission reanalysis (RAN3) of global sea surface temperature (SST) with a 750 m resolution at nadir is available from VIIRS instruments flown onboard two JPSS satellites: NPP (February 2012–present) and N20 (January 2018–present). Two SSTs, ‘subskin’ (sensitive to skin SST) and [...] Read more.
The 3rd full-mission reanalysis (RAN3) of global sea surface temperature (SST) with a 750 m resolution at nadir is available from VIIRS instruments flown onboard two JPSS satellites: NPP (February 2012–present) and N20 (January 2018–present). Two SSTs, ‘subskin’ (sensitive to skin SST) and ‘depth’ (proxy for in situ SST at depth of 20 cm), were produced from brightness temperatures (BTs) in the VIIRS bands centered at 8.6, 11 and 12 µm during the daytime and an additional 3.7 µm band at night, using the NOAA Advanced Clear Sky Processor for Ocean (ACSPO) enterprise SST system. The RAN3 dataset is fully archived at NASA JPL PO.DAAC and NOAA CoastWatch, and routinely supplemented in near real time (NRT) with a latency of a few hours. Delayed mode (DM) processing with a 2 months latency follows NRT, resulting in a more uniform science quality SST record. This paper documents and evaluates the performance of the VIIRS RAN3 dataset. Comparisons with in situ SSTs from drifters and tropical moorings (D+TM) as well as Argo floats (AFs) (both available from the NOAA iQuam system) show good agreement, generally within the NOAA specifications for accuracy (±0.2 K) and precision (0.6 K), in a clear-sky domain covering 18–20% of the global ocean. The nighttime SSTs compare with in situ data more closely, as expected due to the reduced diurnal thermocline. The daytime SSTs are also generally within NOAA specs but show some differences between the (D+TM) and AF validations as well as residual drift on the order of −0.1 K/decade. BT comparisons between two VIIRSs and MODIS-Aqua show good consistency in the 3.7 and 12 µm bands. The 11 µm band, while consistent between NPP and N20, shows residual drift with respect to MODIS-Aqua. Similar analyses of the 8.6 µm band are inconclusive, as the performance of the MODIS band 29 centered at 8.6 µm is degraded and unstable in time and cannot be used for comparisons. Full article
(This article belongs to the Special Issue VIIRS 2011–2021: Ten Years of Success in Earth Observations)
Show Figures

Figure 1

Figure 1
<p>Monthly mean sensitivity (on a 0.1° × 0.1° latitude/longitude grid) for ACSPO V2.80 NPP VIIRS daytime ‘subskin’ SST for May 2020. In most cases, it is close to 1 (except in the tropical regions, and especially at slant VZAs, where atmospheric humidity is high, and the sensitivity can be as low as 0.8).</p>
Full article ">Figure 2
<p>ACSPO NPP ‘subskin’-(D+TM) SST for the full year of 2020, stratified by (<b>a</b>) latitude and (<b>b</b>) retrieved ‘subskin’ SST in RAN2 (produced with ACSPO V2.61) and RAN3 (produced with ACSPO V2.80). The number of observations (NOBS) in the RAN3 datasets is denoted by gray histograms. Bins with NOBS &lt; 5% of the median NOBS were omitted.</p>
Full article ">Figure 3
<p>ACSPO N20 VIIRS nighttime L2P SST imagery over the Chesapeake Bay on 13 September 2021. (<b>a</b>) ACSPO V2.80 all-sky ‘subskin’ SST (no mask applied); (<b>b</b>) ΔSST = ‘subskin’−CMC L4 SST (pixels with ΔSST &lt; −2 K are shown in black); (<b>c</b>) same as (<b>a</b>) but with the V2.80 mask applied (gray pixels); (<b>d</b>) same as (<b>c</b>) except the mask is now from V2.61. Note that over-screening of cold ΔSSTs is significantly reduced in V2.80.</p>
Full article ">Figure 4
<p>ACSPO NPP VIIRS nighttime L2P SST over the Chesapeake Bay on 14 November 2020: (<b>a</b>) all-sky imagery (no ACSM overlaid); (<b>b</b>) the same scene but with the ACSM V2.80 applied (gray overlay) and thermal fronts overlaid (in black). Land is rendered in brown. Images are from the NOAA ACSPO Regional Monitor for SST online system (ARMS) [<a href="#B19-remotesensing-14-03476" class="html-bibr">19</a>].</p>
Full article ">Figure 5
<p>Aggregated for year 2020 maps of nighttime N20 VIIRS ‘subskin’-(D+TM) SST. The figure is from the NOAA SST Quality Monitor (SQUAM) online system [<a href="#B18-remotesensing-14-03476" class="html-bibr">18</a>].</p>
Full article ">Figure 6
<p>Aggregated for year 2020 histograms of ACSPO-(D+TM) SST for N20: (<b>a</b>,<b>b</b>) night; (<b>c</b>,<b>d</b>) day; (<b>a</b>,<b>c</b>) ‘subskin’ SST (computed using global regression, GR); (<b>b</b>,<b>d</b>) ‘depth’ SST (computed using piecewise regression, PWR). Figures are from the NOAA SQUAM system [<a href="#B18-remotesensing-14-03476" class="html-bibr">18</a>].</p>
Full article ">Figure 7
<p>Times series of 24 h aggregated nighttime ACSPO-(D+TM) SSTs: (<b>a</b>,<b>b</b>) accuracy (global mean bias); (<b>c</b>,<b>d</b>) precision (corresponding SD); (<b>a</b>,<b>c</b>) ‘subskin’ SST (computed using global regression, GR); (<b>b</b>,<b>d</b>) ‘depth’ SST (computed using piecewise regression, PWR). Temporal means and SDs of the 24 h aggregated accuracy and precision are given by µ and σ, respectively. Data are from the NOAA SQUAM system [<a href="#B18-remotesensing-14-03476" class="html-bibr">18</a>].</p>
Full article ">Figure 8
<p>Panels (<b>a</b>–<b>d</b>) show the same results as in <a href="#remotesensing-14-03476-f007" class="html-fig">Figure 7</a> but for the daytime.</p>
Full article ">Figure 9
<p>Aggregated for year 2020 maps of nighttime N20 VIIRS ‘subskin’-AF SST. The figure is from the NOAA SST Quality Monitor (SQUAM) online system [<a href="#B18-remotesensing-14-03476" class="html-bibr">18</a>].</p>
Full article ">Figure 10
<p>Panels (<b>a</b>–<b>d</b>) show the same results as <a href="#remotesensing-14-03476-f006" class="html-fig">Figure 6</a> but against the AFs.</p>
Full article ">Figure 11
<p>Panels (<b>a</b>–<b>d</b>) show the same results as <a href="#remotesensing-14-03476-f007" class="html-fig">Figure 7</a> but for monthly aggregated statistics against AFs.</p>
Full article ">Figure 12
<p>Panels (<b>a</b>–<b>d</b>) show the same results as in <a href="#remotesensing-14-03476-f011" class="html-fig">Figure 11</a> but for daytime.</p>
Full article ">Figure 13
<p>Time series of monthly aggregated nighttime ACSPO−TM SST mean biases: (<b>a</b>,<b>b</b>) nighttime; (<b>c</b>,<b>d</b>) daytime; (<b>a</b>,<b>c</b>) ‘subskin’ SST (computed using global regression, GR); (<b>b</b>,<b>d</b>) ‘depth’ SST (computed using piecewise regression, PWR). For each time series, the slope (obtained using a linear least square fit) in units of K/decade is overlaid, along with the associated uncertainty obtained using a 95% confidence interval. Lines show the corresponding linear fits.</p>
Full article ">Figure 14
<p>Time series of monthly aggregated ‘day minus night’ DDs calculated from (<b>a</b>,<b>b</b>) ACSPO−(D+TM); (<b>c</b>,<b>d</b>) ACSPO−L4 ΔSSTs for (<b>a</b>,<b>c</b>) ‘subskin’; (<b>b</b>,<b>d</b>) ‘depth’ SSTs. Slopes (K/decade; obtained using a linear least squares fit) are provided, along with the associated uncertainties obtained using a 95% confidence interval. Lines show the corresponding linear fits.</p>
Full article ">Figure 15
<p>Panels (<b>a</b>–<b>h</b>) show monthly aggregated DDs of mean O−M BTs biases in thermal bands centered at: (<b>a</b>,<b>b</b>) 3.7 µm; (<b>c</b>,<b>d</b>) 8.6 µm; (<b>e</b>,<b>f</b>) 11 µm; (<b>g</b>,<b>h</b>) 12 µm; (<b>i</b>,<b>j</b>) corresponding DDs for satellite ‘subskin’ minus CMC L4 SST [<a href="#B37-remotesensing-14-03476" class="html-bibr">37</a>,<a href="#B47-remotesensing-14-03476" class="html-bibr">47</a>]). Left (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) and right (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) panels show nighttime and daytime results, respectively. Slopes (K/decade; from a linear least squares fit) are shown, along with their uncertainty obtained using a 95% confidence interval. Symbols represent monthly aggregated mean DDs and lines show the corresponding linear fits. Note that the CMC L4 SST is a model input into CRTM results presented in (<b>a</b>–<b>h</b>). Data are from the NOAA MICROS online system [<a href="#B38-remotesensing-14-03476" class="html-bibr">38</a>].</p>
Full article ">
17 pages, 4262 KiB  
Article
Retrieval of Fractional Snow Cover over High Mountain Asia Using 1 km and 5 km AVHRR/2 with Simulated Mid-Infrared Reflective Band
by Fangbo Pan, Lingmei Jiang, Zhaojun Zheng, Gongxue Wang, Huizhen Cui, Xiaonan Zhou and Jinyu Huang
Remote Sens. 2022, 14(14), 3303; https://doi.org/10.3390/rs14143303 - 8 Jul 2022
Cited by 4 | Viewed by 1546
Abstract
Accurate long-term snow-covered-area mapping is essential for climate change studies and water resource management. The NOAA AVHRR/2 provides a unique data source for long-term, large-spatial-scale monitoring of snow-covered areas at a daily scale. However, the value of AVHRR/2 in mapping snow-covered areas is [...] Read more.
Accurate long-term snow-covered-area mapping is essential for climate change studies and water resource management. The NOAA AVHRR/2 provides a unique data source for long-term, large-spatial-scale monitoring of snow-covered areas at a daily scale. However, the value of AVHRR/2 in mapping snow-covered areas is limited, due to its lack of a shortwave infrared band for snow/cloud discrimination. We simulated the reflectance in the 3.75 µm mid-infrared band with a radiative transfer model and then developed three fractional-snow-cover retrieval algorithms for AVHRR/2 imagery at 1 km and 5 km resolutions. These algorithms are based on the multiple endmember spectral mixture analysis algorithm (MESMA), snow index (SI) algorithm, and non-snow/snow two endmember model (TEM) algorithm. Evaluation and comparison of these algorithms were performed using 313 scenarios that referenced snow-cover maps from Landsat-5/TM imagery at 30 m resolution. For all the evaluation data, the MESMA algorithm outperformed the other two algorithms, with an overall accuracy of 0.84 (0.85) and an RMSE of 0.23 (0.21) at the 1 km (5 km) scale. Regarding the effect of land cover type, we found that the three AVHRR/2 fractional-snow-cover retrieval algorithms have good accuracy in bare land, grassland, and Himalayan areas; however, the accuracy decreases in forest areas due to the shading of snow by the canopy. Regarding the topographic effect, the accuracy evaluation indices showed a decreasing and then increasing trend as the elevation increased. The accuracy was worst in the 4000–5000 m range, which was due to the severe snow fragmentation in the High Mountain Asia region; the early AVHRR/2 sensors could not effectively monitor the snow cover in this region. In this study, by increasing the number of bands of AVHRR/2 1 km data for fractional-snow-cover retrieval, a good foundation for subsequent long time series kilometre- resolution snow-cover monitoring has been laid. Full article
(This article belongs to the Special Issue Remote Sensing for Mountain Vegetation and Snow Cover)
Show Figures

Figure 1

Figure 1
<p>The land cover type map of High Mountain Asia. The spatial extent of Landsat–5 Thematic Mapper (TM).</p>
Full article ">Figure 2
<p>The 3.75 μm band reflectivity scatter density plot for the results of this study compared to the official product results.</p>
Full article ">Figure 3
<p>Comparison of the results from the three fractional-snow-cover algorithms on the 1 km scale for High Mountain Asia (25 January 1993): (<b>a</b>) the false colour image of AVHRR/2 bands 1, 2, and 3; (<b>b</b>) FSC from AVHRR/2 using the MESMA algorithm; (<b>c</b>) FSC from AVHRR/2 using the SI algorithm; (<b>d</b>) FSC from AVHRR/2 using the TEM algorithm.</p>
Full article ">Figure 4
<p>Comparison of the results from the three fractional-snow cover algorithms on the 5 km scale for High Mountain Asia (25 January 1993): (<b>a</b>) the false colour image of AVHRR/2 Bands 1, 2, and 3; (<b>b</b>) FSC from AVHRR/2 using the MESMA algorithm; (<b>c</b>) FSC from AVHRR/2 using the SI algorithm; (<b>d</b>) FSC from AVHRR/2 using the TEM algorithm.</p>
Full article ">Figure 5
<p>Boxplots of accuracy evaluation results for three AVHRR/2 fractional-snow-cover retrieval algorithms at 1 km and 5 km scales, based on 313 Landsat-5/TM scenes: (<b>a</b>) the RMSE of the three AVHRR/2 FSC retrieval algorithms at 1 km scale; (<b>b</b>) the RMSE of the three AVHRR/2 FSC retrieval algorithms at 5 km scale; (<b>c</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 1 km scale; (<b>d</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 5 km scale; (<b>e</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km scale; (<b>f</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 5 km scale.</p>
Full article ">Figure 5 Cont.
<p>Boxplots of accuracy evaluation results for three AVHRR/2 fractional-snow-cover retrieval algorithms at 1 km and 5 km scales, based on 313 Landsat-5/TM scenes: (<b>a</b>) the RMSE of the three AVHRR/2 FSC retrieval algorithms at 1 km scale; (<b>b</b>) the RMSE of the three AVHRR/2 FSC retrieval algorithms at 5 km scale; (<b>c</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 1 km scale; (<b>d</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 5 km scale; (<b>e</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km scale; (<b>f</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 5 km scale.</p>
Full article ">Figure 6
<p>Accuracy evaluation results under different surface types for three AVHRR/2 fractional-snow-cover retrieval algorithms at 1 km and 5 km scales based on 313 Landsat-5/TM scenes: (<b>a</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales; (<b>b</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales; (<b>c</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales.</p>
Full article ">Figure 6 Cont.
<p>Accuracy evaluation results under different surface types for three AVHRR/2 fractional-snow-cover retrieval algorithms at 1 km and 5 km scales based on 313 Landsat-5/TM scenes: (<b>a</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales; (<b>b</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales; (<b>c</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales.</p>
Full article ">Figure 7
<p>Accuracy evaluation results at different altitudes for three AVHRR/2 fractional snow cover retrieval algorithms at 1 km and 5 km scales based on 313 Landsat–5/TM scenes: (<b>a</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales; (<b>b</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales; (<b>c</b>) the RMSE of the three AVHRR/2 FSC retrieval algorithms at 1 km and 5 km scales.</p>
Full article ">Figure 8
<p>Boxplots of accuracy evaluation results for the three AVHRR/2 FSC retrieval algorithms from 1 km aggregation to 5 km scale: (<b>a</b>) the RMSE of the three AVHRR/2 FSC retrieval algorithms at 1 km aggregation to 5 km scale; (<b>b</b>) the R<sup>2</sup> of the three AVHRR/2 FSC retrieval algorithms at 1 km aggregation to 5 km scale; (<b>c</b>) the OA of the three AVHRR/2 FSC retrieval algorithms at 1 km aggregation to 5 km scale.</p>
Full article ">
23 pages, 14191 KiB  
Article
AVHRR GAC Sea Surface Temperature Reanalysis Version 2
by Boris Petrenko, Victor Pryamitsyn, Alexander Ignatov, Olafur Jonasson and Yury Kihai
Remote Sens. 2022, 14(13), 3165; https://doi.org/10.3390/rs14133165 - 1 Jul 2022
Cited by 4 | Viewed by 1776
Abstract
The 40+ years-long sea surface temperature (SST) dataset from 4 km Global Area Coverage (GAC) data of the Advanced Very High-Resolution Radiometers (AVHRR/2s and/3s) flown onboard ten NOAA satellites (N07/09/11/12/14/15/16/17/18/19) has been created under the NOAA AVHRR GAC SST Reanalysis 2 (RAN2) Project. [...] Read more.
The 40+ years-long sea surface temperature (SST) dataset from 4 km Global Area Coverage (GAC) data of the Advanced Very High-Resolution Radiometers (AVHRR/2s and/3s) flown onboard ten NOAA satellites (N07/09/11/12/14/15/16/17/18/19) has been created under the NOAA AVHRR GAC SST Reanalysis 2 (RAN2) Project. The data were reprocessed with the NOAA Advanced Clear Sky Processor for Ocean (ACSPO) enterprise SST system. Two SST products are reported in the full ~3000 km AVHRR swath: ‘subskin’ (highly sensitive to true skin SST, but debiased with respect to in situ SST) and ‘depth’ (a closer proxy for in situ data, but with reduced sensitivity). The reprocessing methodology aims at close consistency of satellite SSTs with in situ SSTs, in an optimal retrieval domain. Long-term orbital and calibration trends were compensated by daily recalculation of regression coefficients using matchups with drifters and tropical moored buoys (supplemented by ships for N07/09), collected within limited time windows centered at the processed day. The nighttime Sun impingements on the sensor black body were mitigated by correcting the L1b calibration coefficients. The Earth view pixels contaminated with a stray light were excluded. Massive cold SST outliers caused by volcanic aerosols following three major eruptions were filtered out by a modified, more conservative ACSPO clear-sky mask. The RAN2 SSTs are available in three formats: swath L2P (144 10-min granules per 24 h interval) and two 0.02° gridded (uncollated L3U, also 144 granules/24 h; and collated L3C, two global maps per 24 h, one for day and one for the night). This paper evaluates the RAN2 SST dataset, with a focus on the L3C product and compares it with two other available AVHRR GAC L3C SST datasets, NOAA Pathfinder v5.3 and ESA Climate Change Initiative v2.1. Among the three datasets, the RAN2 covers the global ocean more completely and shows reduced regional and temporal biases, improved stability and consistency between different satellites, and in situ SSTs. Full article
(This article belongs to the Special Issue VIIRS 2011–2021: Ten Years of Success in Earth Observations)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Local equator crossing times (LEXTs) of the ascending half-orbits for the NOAA satellites processed in RAN2. (From the NOAA 3S system [<a href="#B23-remotesensing-14-03165" class="html-bibr">23</a>,<a href="#B24-remotesensing-14-03165" class="html-bibr">24</a>]).</p>
Full article ">Figure 2
<p>Time series of monthly numbers of in situ observations of different types. (From the NOAA <span class="html-italic">i</span>Quam system [<a href="#B18-remotesensing-14-03165" class="html-bibr">18</a>,<a href="#B19-remotesensing-14-03165" class="html-bibr">19</a>]).</p>
Full article ">Figure 3
<p>Monthly nighttime (<b>a</b>) biases, <span class="html-italic">µ</span>, and (<b>b</b>) SDs, <span class="html-italic">σ</span>, of ΔSST = ‘first guess’ − in situ SST. The inserts show mission-averaged means, <span class="html-italic">µ</span>(<span class="html-italic">µ</span>) and <span class="html-italic">µ</span>(<span class="html-italic">σ</span>), and corresponding SDs, <span class="html-italic">σ</span>(<span class="html-italic">µ</span>) and <span class="html-italic">σ</span>(<span class="html-italic">σ</span>). The single and double asterisks denote periods when different ‘first guess’ SSTs were used: (*) REF = CCI L4 and (**) REF = CMC L4 SST. Note that for N07 and N09, in situ = (S + D + TM), whereas for N11–N19, in situ = (D + TM).</p>
Full article ">Figure 4
<p>NOAA satellites and periods of data processing in RAN2, CCI, and PF datasets.</p>
Full article ">Figure 5
<p>Global nighttime N18 L3C ‘subskin’ and ‘skin’ SSTs on 1 Jan 2009, produced from RAN2, CCI, and PF L3C. (Note that +0.17 K was added to CCI and PF ‘skin’ SSTs.) Land is rendered in lighter grey and ocean data with missing SST in darker grey.</p>
Full article ">Figure 6
<p>Nighttime N18 L3C ‘subskin’ and ‘skin’ SSTs over the Gulf of Mexico and Caribbean Sea on 1 Jan 2009, produced from RAN2, CCI, and PF L3C. (Note that +0.17 K was added to CCI and PF ‘skin’ SSTs.) Land is rendered in lighter grey and ocean data with missing SST in darker grey.</p>
Full article ">Figure 7
<p>Time series of monthly nighttime Clear-Sky Ratios (CSRs) in RAN2 and PF L3C products. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 8
<p>Time series of monthly nighttime Clear-Sky Ratios (CSRs) in RAN2 and CCI L2P products. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 9
<p>Hovmöller diagrams (latitude vs. time) of nighttime N11 − (D + TM) SST (<b>a</b>) RAN2 ‘subskin’, (<b>b</b>) CCI ‘skin’ + 0.17 K; (<b>c</b>) PF ‘skin’ + 0.17 K. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 10
<p>Hovmöller diagrams (latitude vs. time) of nighttime N12 − (D + TM) SST (<b>a</b>) RAN2 ‘subskin’, (<b>b</b>) CCI ‘skin’ + 0.17 K. (Note that PF did not process N12.) (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 11
<p>Hovmöller diagrams (latitude vs. time) of nighttime N16 − (D + TM) SST (<b>a</b>) RAN2 ‘subskin’, (<b>b</b>) CCI ‘skin’ + 0.17 K; (<b>c</b>) PF ‘skin’ + 0.17 K. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 12
<p>Monthly numbers of matchups with (D + TM) for (<b>a</b>,<b>d</b>) RAN2, (<b>b</b>,<b>e</b>) CCI and (<b>c</b>,<b>f</b>) PF L3C SSTs, for (<b>a</b>–<b>c</b>) AVHRR/2 and (<b>d</b>–<b>f</b>) AVHRR/3 radiometers. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 13
<p>Monthly numbers of validation matchups with AFs for (<b>a</b>) RAN2, (<b>b</b>) CCI and (<b>c</b>) PF L3C SSTs, for AVHRR/3 radiometers. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 14
<p>Global monthly nighttime biases with respect to (D + TM) for (<b>a</b>,<b>b</b>) RAN2 ‘subskin’ SST, (<b>c</b>,<b>d</b>) CCI ‘skin’ + 0.17 K and (<b>e</b>,<b>f</b>) PF ‘skin’ SSTs + 0.17 K. (<b>a</b>,<b>c</b>,<b>e</b>) AVHRR/2s; (<b>b</b>,<b>d</b>,<b>f</b>) AVHRR/3s. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 15
<p>Global monthly nighttime SDs with respect to (D + TM) for (<b>a</b>,<b>b</b>) RAN2 ‘subskin’ SST, (<b>c</b>,<b>d</b>) CCI and (<b>e</b>,<b>f</b>) PF ‘skin’ SSTs. (<b>a</b>,<b>c</b>,<b>e</b>) AVHRR/2s; (<b>b</b>,<b>d</b>,<b>f</b>) AVHRR/3s. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 16
<p>Global monthly nighttime biases with respect to (D + TM) for (<b>a</b>,<b>b</b>) RAN2 ‘depth’ SST and (<b>c</b>,<b>d</b>) CCI (<b>a</b>,<b>c</b>) AVHRR/2s; (<b>b</b>,<b>d</b>) AVHRR/3s. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 17
<p>Global monthly nighttime SDs with respect to (D + TM) for (<b>a</b>,<b>b</b>) RAN2 ‘depth’ SST and (<b>c</b>,<b>d</b>) CCI (<b>a</b>,<b>c</b>) AVHRR/2s; (<b>b</b>,<b>d</b>) AVHRR/3s. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 18
<p>Time series of global monthly (<b>a</b>,<b>c</b>,<b>e</b>) biases and (<b>b</b>,<b>d</b>,<b>f</b>) SDs of (<b>a</b>,<b>b</b>) RAN2 ‘subskin’ SST, (<b>c</b>,<b>d</b>) CCI ‘skin’ SST + 0.17 K and (<b>e</b>,<b>f</b>) PF ‘skin’ SST +0.17 K with respect to AF. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 19
<p>Time series of global monthly (<b>a</b>,<b>c</b>) biases and (<b>b</b>,<b>d</b>) SDs of (<b>a</b>,<b>b</b>) RAN2 ‘depth’ and (<b>c</b>,<b>d</b>) CCI ‘depth’ SST with respect to AF. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">Figure 20
<p>Time series of monthly averaged day−night double differences in (<b>a</b>) RAN2 ‘subskin’, (<b>b</b>) CCI, and (<b>c</b>) PF ‘skin’ + 0.17 K SSTs. (Data are from the NOAA SQUAM system [<a href="#B21-remotesensing-14-03165" class="html-bibr">21</a>,<a href="#B22-remotesensing-14-03165" class="html-bibr">22</a>]).</p>
Full article ">
17 pages, 9876 KiB  
Article
Detection of Snow Cover from Historical and Recent AVHHR Data—A Thematic TIMELINE Processor
by Sebastian Rößler and Andreas J. Dietz
Geomatics 2022, 2(1), 144-160; https://doi.org/10.3390/geomatics2010009 - 18 Mar 2022
Cited by 3 | Viewed by 2831
Abstract
Global snow cover forms the largest and most transient part of the cryosphere in terms of area. On the local and regional scale, small changes can have drastic effects such as floods and droughts, and on the global scale is the planetary albedo. [...] Read more.
Global snow cover forms the largest and most transient part of the cryosphere in terms of area. On the local and regional scale, small changes can have drastic effects such as floods and droughts, and on the global scale is the planetary albedo. Daily imagery of snow cover forms the basis of long-term observation and analysis, and only optical sensors offer the necessary spatial and temporal resolution to address decadal developments and the impact of climate change on snow availability. The MODIS sensors have been providing this daily information since 2000; before that, only the Advanced Very High-Resolution Radiometer (AVHRR) from the National Oceanographic and Atmospheric Administration (NOAA) was suitable. In the TIMELINE project of the German Aerospace Center, the historic AVHRR archive in HRPT (High Resolution Picture Transmission) format is processed for the European area and, among other processors, one output is the thematic product ‘snow cover’ that will be made available in 1 km resolution since 1981. The snow detection is based on the Normalized Difference Snow Index (NDSI), which enables a direct comparison with the MODIS snow product. In addition to the NDSI, ERA5 re-analysis data on the skin temperature and other level 2 TIMELINE products are included in the generation of the binary snow mask. The AVHRR orbit segments are projected from the swath projection into LAEA Europe, aggregated into daily coverages, and from this, the 10-day and monthly snow covers are finally calculated. In this publication, the snow cover algorithm is presented, as well as the results of the first validations and possible applications of the final product. Full article
Show Figures

Figure 1

Figure 1
<p>Reflective properties of snow compared to vegetation. The spectral response functions (sensitivity) for the AVHRR bands 1 to 3a are shown, with channels 1 and 2 being available for all AVHRR sensors, but channel 3a only for the third generation (AVHRR/3) and on MetOP-A and -B.</p>
Full article ">Figure 2
<p><span class="html-italic">NDSI</span> thresholds for detecting snow. The area was expanded based on the <span class="html-italic">NDVI</span> values for the “snow in the forest” class.</p>
Full article ">Figure 3
<p>Flowchart of the TIMELINE Snow Mask Processor. The central <span class="html-italic">NDSI</span>-based snow recognition is highlighted with red color.</p>
Full article ">Figure 4
<p>Tiling scheme of the TIMELINE area in ETRS89-extended/LAEA Europe projection.</p>
Full article ">Figure 5
<p>Example of a Level 2 snow cover product in satellite swath projection from 8 January 2009 (recording time 09:15:00–09:26:26 UTC); Platform NOAA-17; Sensor AVHRR/3. (<b>A</b>) RGB composite from the Level 1b bands 3a, 2, 1; (<b>B</b>) from band 1 and band 3a calculated <span class="html-italic">NDSI</span> (gradient of black, −1, towards white, +1); (<b>C</b>) snowmask created according to the sheme presented in <a href="#geomatics-02-00009-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 6
<p>Example of a daily Level 3 product for 15 January 2009 (mid-winter).</p>
Full article ">Figure 7
<p>Example of a Daily Level 3 product for 22 February 2009 (maximum snow coverage).</p>
Full article ">Figure 8
<p>Boxplots showing the variance of classification accuracy indicators during the first 120 days of the test year 2009.</p>
Full article ">Figure 9
<p>The development of classification accuracy indicators over the months January to April 2009 (i.e., the first 120 days of the year).</p>
Full article ">Figure 10
<p>Daily composite of the classification accuracy for 13 January 2009–the date with the fewest matches between MODIS and AVHRR (TNR: 47.5%, PPV: 34.25%, and ACC: 58.67%).</p>
Full article ">Figure 11
<p>Daily composite of the classification accuracy for 28 April 2009—the date with the best matches between MODIS and AVHRR in terms of TNR (95.67%) and ACC (95.58%).</p>
Full article ">Figure 12
<p>Daily composite of the classification accuracy for 11 April 2009—the date with the best matches between MODIS and AVHRR in terms of PPV (70.54%).</p>
Full article ">
23 pages, 11756 KiB  
Article
Detection of Multidecadal Changes in Vegetation Dynamics and Association with Intra-Annual Climate Variability in the Columbia River Basin
by Andrew B. Whetten and Hannah J. Demler
Remote Sens. 2022, 14(3), 569; https://doi.org/10.3390/rs14030569 - 25 Jan 2022
Cited by 3 | Viewed by 2551
Abstract
Remotely-sensed Leaf Area Index (LAI) is a useful metric for assessing changes in vegetation cover and greeness over time and space. Satellite-derived LAI measurements can be used to assess these intra- and inter-annual vegetation dynamics and how they correlate with changing regional and [...] Read more.
Remotely-sensed Leaf Area Index (LAI) is a useful metric for assessing changes in vegetation cover and greeness over time and space. Satellite-derived LAI measurements can be used to assess these intra- and inter-annual vegetation dynamics and how they correlate with changing regional and local climate conditions. The detection of such changes at local and regional levels is challenged by the underlying continuity and extensive missing values of high-resolution spatio-temporal vegetation data. Here, the feasibility of functional data analysis methods was evaluated to improve the exploration of such data. In this paper, an investigation of multidecadal variation in LAI is conducted in the Columbia River Watershed, as detected by NOAA Advanced Very High-Resolution Radiometer (AVHRR) satellite imaging. The inter- and intra-annual correlation of LAI with temperature and precipitation were then investigated using data from the European Centre for Medium-Range Weather Forecasts global atmospheric re-analysis (ERA-Interim) in the period 1996–2017. A functional cluster analysis model was implemented to identify regions in the Columbia River Watershed that exhibit similar long-term greening trends. Across this region, a multidecadal trend toward earlier and higher annual LAI peaks was detected, and strong correlations were found between earlier and higher LAI peaks and warmer temperatures in late winter and early spring. Although strongly correlated to LAI, maximum temperature and precipitation do not demonstrate a similar strong multidecadal trend over the studied time period. The modeling approach is proficient for analyzing tens or hundreds of thousands of sampled sites without parallel processing or high-performance computing (HPC). Full article
Show Figures

Figure 1

Figure 1
<p>Flow diagram of the analysis of the LAI CDR and ERA-Interim data products in the CRB region.</p>
Full article ">Figure 2
<p>The boundary of the Columbia River watershed is represented on the map by the light yellow line.</p>
Full article ">Figure 3
<p>Illustration of the LAI smoothing process. (<b>Upper</b>) Raw LAI (blue) and spline smoothed LAI (orange) are overlayed for Site X (42.325 N–109.775 W). The spline model retains the functional structure of the raw LAI recording while filtering out anomaly/false recordings. (<b>Lower Left</b>) Spline smoothed LAI for adjacent sites: Site X and Site Y (42.325 N–109.825 W). The Spearman correlation is 0.988. Site X Elevation = 2133.8 m and Site Y Elevation = 2168.8 m, and both sites are classified as scrub/shrub locations. (<b>Lower Right</b>) Spline smoothed LAI for Site X and Site Z (47.425 N–118.225 W). The Spearman correlation is 0.769. Site Z Elevation = 690.5 m and is classified as an agriculture location [<a href="#B23-remotesensing-14-00569" class="html-bibr">23</a>,<a href="#B24-remotesensing-14-00569" class="html-bibr">24</a>]. The correlation in the lower left figure is higher because the annual phenological changes are more synchronized between these sites than between sites in the lower right plot.</p>
Full article ">Figure 4
<p>Temporal correlation of maximum annual LAI from 1996 to 2017 discretized to Greening, Browning, and Neither using a significance level of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.15</mn> </mrow> </semantics></math> [<a href="#B32-remotesensing-14-00569" class="html-bibr">32</a>].</p>
Full article ">Figure 5
<p>The functional cluster analysis of the pairwise correlation matrix of the 27,191 B-spline smoothed LAI profiles [<a href="#B32-remotesensing-14-00569" class="html-bibr">32</a>].</p>
Full article ">Figure 6
<p>Inter-annual regional average weekly maximum LAI, maximum temperature, and cumulative annual precipitation profiles. The curves are colored on a gradient scale where purple curves are closer to 1996 and gold curves are closer to 2017, and a blue transition color in the intermediate years. Noticeable time-dependent changes in LAI were identified where no clear trend in temperature and precipitation is visually observed.</p>
Full article ">Figure 7
<p>LAI functional principal components results for the first principal component of each cluster. Upper plots solid lines are the mean annual profiles and the (+) markers denote the trend line for the first component added to the mean function with an appropriate scaling factor, and the (☐) markers denote the trend line for the first component subtracted from the mean function with the same scaling factor. The lower plots display the annual scores of the first component as a time series from 1996 to 2017. Years with scores greater than zero are characterized by the (+) trend line and years with scores less than zero are characterized by the (☐) trend line.</p>
Full article ">Figure 8
<p>Weekly maximum temperature first functional principal component results for the first principal component of each cluster.</p>
Full article ">Figure 9
<p>Weekly maximum temperature second functional principal component results for the second principal component of each cluster.</p>
Full article ">Figure 10
<p>Weekly average precipitation first functional principal component results for the first principal component of each cluster.</p>
Full article ">Figure 11
<p>LAI vs. max. temperature functional canonical correlation results for the first pair of canonical weight functions. The correlation between site attributes is listed above each columns of plots. The (+) markers denote the trend line for the first weight function (for either the LAI weight function or the maximum temperature weight function) added to the mean function with an appropriate scaling factor, and the (☐) markers denote the trend line for the first weight function subtracted from the mean function with the same scaling factor. The strength of the correlation between LAI and maximum temperature is characterized by examining the pair of (+) profiles or (☐) profiles across (for each site attribute).</p>
Full article ">Figure 12
<p>LAI vs. cumulative precipitation functional canonical correlation results for the first pair of canonical weight functions. The correlation between site attributes is listed above each columns of plots. The (+) markers denote the trend line for the first weight function (for either the LAI weight function or the precipitation weight function) added to the mean function with an appropriate scaling factor, and the (☐) markers denote the trend line for the first weight function subtracted from the mean function with the same scaling factor. The strength of the correlation between LAI and precipitation is characterized by examining the pair of (+) profiles or (☐) profiles across (for each site attribute).</p>
Full article ">Figure A1
<p>Distribution of site elevations by Cluster. We conducted one-way anova testing and post-hoc Tukey adjusted comparison testing of cluster means, and we detected highly significant (&lt;0.0001) differences in the distribution of elevation across clusters.</p>
Full article ">
Back to TopTop