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

Journals

Article Types

Countries / Regions

Search Results (81)

Search Parameters:
Keywords = FLUXNET

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 15431 KiB  
Article
Warming Diminishes the Day–Night Discrepancy in the Apparent Temperature Sensitivity of Ecosystem Respiration
by Nan Li, Guiyao Zhou, Mayank Krishna, Kaiyan Zhai, Junjiong Shao, Ruiqiang Liu and Xuhui Zhou
Plants 2024, 13(23), 3321; https://doi.org/10.3390/plants13233321 - 26 Nov 2024
Viewed by 585
Abstract
Understanding the sensitivity of ecosystem respiration (ER) to increasing temperature is crucial to predict how the terrestrial carbon sink responds to a warming climate. The temperature sensitivity of ER may vary on a diurnal basis but is poorly understood due to the paucity [...] Read more.
Understanding the sensitivity of ecosystem respiration (ER) to increasing temperature is crucial to predict how the terrestrial carbon sink responds to a warming climate. The temperature sensitivity of ER may vary on a diurnal basis but is poorly understood due to the paucity of observational sites documenting real ER during daytime at a global scale. Here, we used an improved flux partitioning approach to estimate the apparent temperature sensitivity of ER during the daytime (E0,day) and nighttime (E0,night) derived from multiyear observations of 189 FLUXNET sites. Our results demonstrated that E0,night is significantly higher than E0,day across all biomes, with significant seasonal variations in the day–night discrepancy in the temperature sensitivity of ER (ΔE0 = E0,night/E0,day) except for evergreen broadleaf forest and savannas. Such seasonal variations in ΔE0 mainly result from the effect of temperature and the seasonal amplitude of NDVI. We predict that future warming will decrease ΔE0 due to the reduced E0,night by the end of the century in most regions. Moreover, we further find that disregarding the ΔE0 leads to an overestimation of annual ER by 10~80% globally. Thus, our study highlights that the divergent temperature dependencies between day- and nighttime ER should be incorporated into Earth system models to improve predictions of carbon–climate change feedback under future warming scenarios. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of 189 flux sites from FLUXNET 2015. (<b>a</b>) Location of each site on the latitude and longitude grid; (<b>b</b>) Location of each site according to the classic Whittaker Biome Classification based on the climate variables of mean annual temperature and precipitation. CRO, cropland; DBF, deciduous broadleaf forest; EBF, evergreen broadleaf forest; ENF, evergreen needleleaf forest; MF, mixed forest; GRA, grassland; SAV, savannas; SH, shrubland; WET, wetland.</p>
Full article ">Figure 2
<p>Day–night discrepancy in the temperature sensitivity of ER (E<sub>0</sub>). (<b>a</b>). Density plot of the daytime (E<sub>0,day</sub>) and nighttime (E<sub>0,night</sub>) temperature sensitivity of ER, which were estimated from DT-RH flux partitioning algorithms. (<b>b</b>). The difference in daytime and nighttime E<sub>0</sub> among nine ecosystem types. *** <span class="html-italic">p</span> &lt; 0.001; CRO, cropland; DBF, deciduous broadleaf forest; EBF, evergreen broadleaf forest; ENF, evergreen needleleaf forest; MF, mixed forest; GRA, grassland; SAV, savannas; SH, shrubland; WET, wetland.</p>
Full article ">Figure 3
<p>Average monthly daytime (E<sub>0,day</sub>) and nighttime (E<sub>0,night</sub>) temperature sensitivity of ER and their discrepancies (ΔE<sub>0</sub> = E<sub>0,night</sub>/E<sub>0,day</sub>) among different vegetation types. (<b>a</b>–<b>c</b>) Mean monthly ΔE<sub>0</sub>, E<sub>0,day</sub> and E<sub>0,night</sub> in DBF, EBF, ENF and MF, respectively. (<b>d</b>–<b>f</b>) Mean monthly ΔE<sub>0</sub>, E<sub>0,day</sub> and E<sub>0,night</sub> in CRO, GRA, SAV, SH and WET, respectively. If ΔE<sub>0</sub> &gt; 1, it means that the E<sub>0</sub> during the day is smaller than at night, and vice versa. CRO, cropland; DBF, deciduous broadleaf forest; EBF, evergreen broadleaf forest; ENF, evergreen needleleaf forest; MF, mixed forest; GRA, grassland; SAV, savannas; SH, shrubland; WET, wetland.</p>
Full article ">Figure 4
<p>The distribution of the seasonal amplitude (ASV) in the E<sub>0,night</sub>/E<sub>0,day</sub> ratio (ΔE<sub>0,ASV</sub>) across the temperature–precipitation space (<b>a</b>) and latitude (<b>b</b>). ΔE<sub>0,ASV</sub> was calculated as the difference between highest and lowest monthly E<sub>0,night</sub>/E<sub>0,day</sub> ratios from January to December. Each climatic bin is 4 °C (temperature) by 1 mm day<sup>−1</sup> (precipitation) in panel a. CRO, cropland; DBF, deciduous broadleaf forest; EBF, evergreen broadleaf forest; ENF, evergreen needleleaf forest; MF, mixed forest; GRA, grassland; SAV, savannas; SH, shrubland; WET, wetland.</p>
Full article ">Figure 5
<p>The relative contribution of environmental and plant variables to the seasonal amplitude (ASV) of E<sub>0,night</sub>/E<sub>0,day</sub>. (<b>a</b>,<b>b</b>) Principle component analysis of environmental and plant variables for calculating their relative contributions on the seasonal amplitude (ASV) of E<sub>0,night</sub>/E<sub>0,day</sub>. Contribution of environmental and plant variables to Dim 1 and Dim 2 of the principal component analysis in panel b, and the red dashed line represents the average contribution. (<b>c</b>) Results from random forest analysis, showing the relative importance of the various drivers in ASV of E<sub>0,night</sub>/E<sub>0,day</sub>. Variable importance represents the increase in mean error (computed on the out-of-bag data) across trees when a predictor is permuted. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. MAT, mean annual temperature; Tair, monthly air temperature; Tnight, nighttime air temperature; Tday, daytime air temperature; Tmax, monthly maximum air temperature; Tmin, monthly minimum air temperature; DTR, diurnal temperature range; PAR, solar radiation; VPD, vapor pressure deficit; SPEI, standardized precipitation ET index; LAI, leaf area index; NDVI, normalized difference vegetation index; SIF, sun-induced chlorophyll fluorescence; LNC, leaf nitrogen concentration; LPC, leaf phosphorus concentration; LN/P, LNC/LNP; LDMC, leaf dry matter content; SLA, specific leaf area; SWC, surface volumetric soil moisture; SOC, soil organic carbon; PFT, plant functional types. Tnight<sub>ASV</sub>, Tair<sub>ASV</sub>, Tday<sub>ASV</sub>, SPEI<sub>ASV</sub>, LAI<sub>ASV</sub>, PAR<sub>ASV</sub>, VPD<sub>ASV</sub>, Precipitation<sub>ASV</sub>, and NDVI<sub>ASV</sub> are the seasonal amplitude of Tnight, Tair, Tday, SPEI, LAI, PAR, VPD, precipitation, and NDVI, respectively, from January to December.</p>
Full article ">Figure 6
<p>Global patterns of the daytime (E<sub>0,day</sub>) and nighttime (E<sub>0,night</sub>) temperature sensitivity of ER and their discrepancies (ΔE<sub>0</sub> = E<sub>0,night</sub>/E<sub>0,day</sub>) based on historical data and future projections. (<b>a</b>–<b>c</b>) The figures shows the global patterns of E<sub>0,night</sub>, E<sub>0,day</sub>, and ΔE<sub>0</sub> in 2015. (<b>d</b>–<b>i</b>) The figures shows the global patterns of E<sub>0,night</sub>, E<sub>0,day</sub>, and ΔE<sub>0</sub> in 2100 under two future corresponding shared socioeconomic pathway-representative concentration pathway (SSP-RCP) scenarios (<b>d</b>–<b>f</b>, SSP1-RCP2.6; <b>g</b>–<b>i</b>, SSP5-RCP8.5).</p>
Full article ">Figure 7
<p>Historical and future ER values estimated from DT-RH and CMIP6 models. (<b>a</b>,<b>d</b>) The figures show the global patterns of ER in 2015 derived from DT-RH and CMIP6 models, respectively. (<b>b</b>,<b>c</b>,<b>e</b>,<b>f</b>) The figures show the global patterns of ER in 2100 derived from DT-RH and CMIP6 models under two future corresponding shared socioeconomic pathway-representative concentration pathway (SSP-RCP) scenarios.</p>
Full article ">
17 pages, 2917 KiB  
Article
Sensitivity and Uncertainty Analysis of the GeeSEBAL Model Using High-Resolution Remote-Sensing Data and Global Flux Site Data
by Shunjun Hu, Changyan Tian and Ping Jiao
Water 2024, 16(20), 2978; https://doi.org/10.3390/w16202978 - 18 Oct 2024
Viewed by 673
Abstract
Actual evapotranspiration (ETa) is an important component of the surface water cycle. The geeSEBAL model is increasingly being used to estimate ETa using high-resolution remote-sensing data (Landsat 4/5/7/8). However, due to surface heterogeneity, there is significant uncertainty. By optimizing [...] Read more.
Actual evapotranspiration (ETa) is an important component of the surface water cycle. The geeSEBAL model is increasingly being used to estimate ETa using high-resolution remote-sensing data (Landsat 4/5/7/8). However, due to surface heterogeneity, there is significant uncertainty. By optimizing the quantile values of the reverse-modelling automatic calibration algorithm (CIMEC) endpoint-component selection algorithm under extreme conditions through 212 global flux sites, we obtained the optimized quantile values of 11 vegetation types of cold- and hot-pixel endpoint components (Ts and NDVI). Based on the observation data of the global FLUXNET tower, the sensitivity of 20 parameters in the improved geeSEBAL model was determined through Sobol’s sensitivity analysis. Among them, the parameters dT and SAVI,hot were confirmed as the most sensitive parameters of the algorithm. Subsequently, we used the differential evolution Markov chain (DE-MC) method to analyse the uncertainty of the parameters in the geeSEBAL model used the posterior distribution of the parameters to modify the sensitive parameter values or ranges in the improved geeSEBAL model and to simulate the daily ETa. The results indicate that by analysing the end element components of the geeSEBAL model (Ts and NDVI), quantile numerical optimization and parameter optimization can be performed. Compared with the original algorithm, the improved geeSEBAL model has significantly improved simulation performance, as shown by higher R2 values, higher NSE values, smaller bias values, and lower RMSE values. The most suitable values of the predefined parameter Zoh were determined, and the reanalysis of meteorological data inputs (relative humidity (RH), temperature (T), wind speed (WS), and net radiation (Rn)) was also found to be an important source of uncertainty for the accurate estimation of ETa. This study indicates that optimizing the quantiles and key parameters of the model end component has certain potential for further improving the accuracy of the geeSEBAL model based on high-resolution remote-sensing data in estimating the ETa for various vegetation types. Full article
(This article belongs to the Special Issue Agricultural Water-Land-Plant System Engineering)
Show Figures

Figure 1

Figure 1
<p>Framework diagram of geeSEBAL model based on the CIMEC algorithm.</p>
Full article ">Figure 2
<p>Spatial distribution of available global flux sites in the FLUXNET2015 dataset (review number: GS (2021) 6375).</p>
Full article ">Figure 3
<p>Changes in the quantile value q of end-element components on cold and hot pixels Δ<span class="html-italic">T</span><sub>s</sub>, Δ<span class="html-italic">ET</span><sub>a</sub>, Δ<span class="html-italic">NDVI,</span> and Δ<span class="html-italic">H</span> sensitivity: the high and low lines in the graph represent the maximum and minimum changes, and the points represent the average changes. (<b>a</b>) Δ<span class="html-italic">T</span> − <span class="html-italic">T</span><sub>s,q</sub>, (<b>b</b>) Δ<span class="html-italic">ET</span><sub>a</sub> − <span class="html-italic">T</span><sub>s,q</sub>, (<b>c</b>) Δ<span class="html-italic">NDVI</span> − <span class="html-italic">NDVI</span><sub>,q</sub> and (<b>d</b>) Δ<span class="html-italic">H</span> − <span class="html-italic">T</span><sub>s,q</sub>.</p>
Full article ">Figure 4
<p>Sensitivity of the Sobol first-order and total-order parameters for different vegetation types.</p>
Full article ">Figure 5
<p>First-order (Si) (<b>a</b>) and full-order (Sti) (<b>b</b>) sensitivity analyses of the SEBAL model parameters based on Sobol’s method. The red dotted line and the black dotted line represent 0.1 and 0, respectively, and the parameters above the red dashed line are extremely sensitive parameters, and the parameters above the black dashed line are sensitive parameters.</p>
Full article ">Figure 6
<p>Posterior distribution of the geeSEBAL model parameters for various vegetation types.</p>
Full article ">Figure 6 Cont.
<p>Posterior distribution of the geeSEBAL model parameters for various vegetation types.</p>
Full article ">Figure 7
<p>ERA5 Meteorological input evaluation station verification Taylor chart relative humidity <span class="html-italic">RH</span> (<b>a</b>), temperature <span class="html-italic">T</span> (<b>b</b>), wind speed <span class="html-italic">WS</span> (<b>c</b>), and net radiation <span class="html-italic">R</span><sub>n</sub> (<b>d</b>).</p>
Full article ">
21 pages, 13840 KiB  
Article
Estimating Forest Gross Primary Production Using Machine Learning, Light Use Efficiency Model, and Global Eddy Covariance Data
by Zhenkun Tian, Yingying Fu, Tao Zhou, Chuixiang Yi, Eric Kutter, Qin Zhang and Nir Y. Krakauer
Forests 2024, 15(9), 1615; https://doi.org/10.3390/f15091615 - 13 Sep 2024
Viewed by 1140
Abstract
Forests play a vital role in atmospheric CO2 sequestration among terrestrial ecosystems, mitigating the greenhouse effect induced by human activity in a changing climate. The LUE (light use efficiency) model is a popular algorithm for calculating terrestrial GPP (gross primary production) based [...] Read more.
Forests play a vital role in atmospheric CO2 sequestration among terrestrial ecosystems, mitigating the greenhouse effect induced by human activity in a changing climate. The LUE (light use efficiency) model is a popular algorithm for calculating terrestrial GPP (gross primary production) based on physiological mechanisms and is easy to implement. Different versions have been applied for many years to simulate the GPP of different ecosystem types at regional or global scales. For estimating forest GPP using different approaches, we implemented five LUE models (EC-LUE, VPM, GOL-PEM, CASA, and C-Fix) in forests of type DBF, EBF, ENF, and MF, using the FLUXNET2015 dataset, remote sensing observations, and Köppen–Geiger climate zones. We then fused these models to additionally improve the ability of the GPP estimation using an RF (random forest) and an SVM (support vector machine). Our results indicated that under a unified parameterization scheme, EC-LUE and VPM yielded the best performance in simulating GPP variations, followed by GLO-PEM, CASA, and C-fix, while MODIS also demonstrated reliable GPP estimation ability. The results of the model fusion across different forest types and flux net sites indicated that the RF could capture more GPP variation magnitudes with higher R2 and lower RMSE than the SVM. Both RF and SVM were validated using cross-validation for all forest types and flux net sites, showing that the accuracy of the GPP simulation could be improved by the RF and SVM by 28% and 27%. Full article
(This article belongs to the Section Forest Ecology and Management)
Show Figures

Figure 1

Figure 1
<p>Köppen–Geiger climate zones and 45 FLUXNET2015 forest sites (red triangles) distribution. Köppen–Geiger climate symbols are listed in <a href="#app1-forests-15-01615" class="html-app">Table S2 in the Supporting Information File</a>.</p>
Full article ">Figure 2
<p>Workflow of GPP estimation through the integration of LUE models based on ground measurements, remote sensing observations, and Köppen–Geiger climate zones.</p>
Full article ">Figure 3
<p>The Taylor diagrams for site-derived GPP and LUE models/machine learning estimates at the 45 FLUXNET2015 sites. The dotted circular lines which connect the X and Y axes denote <span class="html-italic">SD</span>. The dotted radial lines represent R. The brown curves are <span class="html-italic">RMSD</span> compared to the referenced site’s GPP.</p>
Full article ">Figure 4
<p>The <span class="html-italic">R</span><sup>2</sup> (<b>a</b>), <span class="html-italic">RMSE</span> (<b>b</b>), and <span class="html-italic">RPE</span> (<b>c</b>) of 5 single models, MODIS, SVM, and RF across the DBF, EBF, ENF, and MF.</p>
Full article ">Figure 5
<p>The scatter plots of <span class="html-italic">R</span><sup>2</sup>, <span class="html-italic">RMSE</span>, and <span class="html-italic">RPE</span> across the DBF between site-derived GPP and the estimates from the LUE models, MODIS, SVM, and RF.</p>
Full article ">Figure 6
<p>The probability distribution of errors from the LUE models, MODIS, SVM, and RF.</p>
Full article ">Figure 7
<p>The <span class="html-italic">AIC</span> (<b>a</b>) and <span class="html-italic">BIC</span> (<b>b</b>) of the LUE models, MODIS, SVM, and RF across the DBF, EBF, ENF, and MF.</p>
Full article ">Figure 8
<p>Daily FLUXNET2015’s GPP (black dots), the best LUE model-estimated GPP (EC-LUE, line in blue), and the best fusion method-estimated GPP (RF, line in orange) at 4 sites: DE-Hai of DBF (<b>a</b>), AU-Tum of EBF (<b>b</b>), US-Blo of ENF (<b>c</b>), and BE-Bra of MF (<b>d</b>).</p>
Full article ">Figure 9
<p>Boxplot of performance of <span class="html-italic">R</span><sup>2</sup> (<b>a</b>), <span class="html-italic">RMSE</span> (<b>b</b>), and <span class="html-italic">AIC</span> (<b>c</b>) of LUE and machine learning methods across 45 FLUXNET2015 sites.</p>
Full article ">
17 pages, 6630 KiB  
Article
Interpreting Controls of Stomatal Conductance across Different Vegetation Types via Machine Learning
by Runjia Xue, Wenjun Zuo, Zhaowen Zheng, Qin Han, Jingyan Shi, Yao Zhang, Jianxiu Qiu, Sheng Wang, Yan Zhu, Weixing Cao and Xiaohu Zhang
Water 2024, 16(16), 2251; https://doi.org/10.3390/w16162251 - 9 Aug 2024
Viewed by 1538
Abstract
Plant stomata regulate transpiration (T) and CO2 assimilation, essential for the water–carbon cycle. Quantifying how environmental factors influence stomatal conductance will provide a scientific basis for understanding the vegetation–atmosphere water–carbon exchange process and water use strategies. Based on eddy covariance [...] Read more.
Plant stomata regulate transpiration (T) and CO2 assimilation, essential for the water–carbon cycle. Quantifying how environmental factors influence stomatal conductance will provide a scientific basis for understanding the vegetation–atmosphere water–carbon exchange process and water use strategies. Based on eddy covariance and hydro-metrological observations from FLUXNET sites with four plant functional types and using three widely applied methods to estimate ecosystem T from eddy covariance data, namely uWUE, Perez-Priego, and TEA, we quantified the regulation effect of environmental factors on canopy stomatal conductance (Gs). The environmental factors considered here include radiation (net radiation and solar radiation), water (soil moisture, relative air humidity, and vapor pressure deficit), temperature (air temperature), and atmospheric conditions (CO2 concentration and wind speed). Our findings reveal variation in the influence of these factors on Gs across biomes, with air temperature, relative humidity, soil water content, and net radiation being consistently significant. Wind speed had the least influence. Incorporating the leaf area index into a Random Forest model to account for vegetation phenology significantly improved model accuracy (R2 increased from 0.663 to 0.799). These insights enhance our understanding of the primary factors influencing stomatal conductance, contributing to a broader knowledge of vegetation physiology and ecosystem functioning. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of the study sites.</p>
Full article ">Figure 2
<p>Histogram of the importance of <span class="html-italic">G</span><sub>s</sub> characteristics at each site. The y-axis starts at −0.5 because the importance and error bars exceed 0 for some sites, and the negative y indicates that the feature hurts the RF prediction results.</p>
Full article ">Figure 2 Cont.
<p>Histogram of the importance of <span class="html-italic">G</span><sub>s</sub> characteristics at each site. The y-axis starts at −0.5 because the importance and error bars exceed 0 for some sites, and the negative y indicates that the feature hurts the RF prediction results.</p>
Full article ">Figure 3
<p>Model accuracy of before (Benchmark) and after adding LAI (LAI). The darker colors in the panels correspond to higher values. (<b>a</b>) presents R<sup>2</sup> results for the training set, while (<b>b</b>) displays R<sup>2</sup> results for the test set. (<b>c</b>) shows RMSE outcomes for the training set, and (<b>d</b>) outlines RMSE results for the test set.</p>
Full article ">Figure A1
<p>Histogram of the importance of <span class="html-italic">G</span><sub>s</sub> characteristics at each site (adding LAI). The y-axis starts at −0.5 because the importance and error bars exceed 0 for some sites, and the negative y indicates that the feature hurts the RF prediction results.</p>
Full article ">Figure A1 Cont.
<p>Histogram of the importance of <span class="html-italic">G</span><sub>s</sub> characteristics at each site (adding LAI). The y-axis starts at −0.5 because the importance and error bars exceed 0 for some sites, and the negative y indicates that the feature hurts the RF prediction results.</p>
Full article ">
25 pages, 19977 KiB  
Article
Different Vegetation Covers Leading to the Uncertainty and Consistency of ET Estimation: A Case Study Assessment with Extended Triple Collocation
by Xiaoxiao Li, Huaiwei Sun, Yong Yang, Xunlai Sun, Ming Xiong, Shuo Ouyang, Haichen Li, Hui Qin and Wenxin Zhang
Remote Sens. 2024, 16(13), 2484; https://doi.org/10.3390/rs16132484 - 6 Jul 2024
Viewed by 1204
Abstract
Accurate and reliable estimation of actual evapotranspiration (AET) is essential for various hydrological studies, including drought prediction, water resource management, and the analysis of atmospheric–terrestrial carbon exchanges. Gridded AET products offer potential for application in ungauged areas, but their uncertainties may be significant, [...] Read more.
Accurate and reliable estimation of actual evapotranspiration (AET) is essential for various hydrological studies, including drought prediction, water resource management, and the analysis of atmospheric–terrestrial carbon exchanges. Gridded AET products offer potential for application in ungauged areas, but their uncertainties may be significant, making it difficult to identify the best products for specific regions. While in situ data directly estimate gridded ET products, their applicability is limited in ungauged areas that require FLUXNET data. This paper employs an Extended Triple Collocation (ETC) method to estimate the uncertainty of Global Land Evaporation Amsterdam Model (GLEAM), Famine Early Warning Systems Network (FLDAS), and Maximum Entropy Production (MEP) AET product without requiring prior information. Subsequently, a merged ET product is generated by combining ET estimates from three original products. Furthermore, the study quantifies the uncertainty of each individual product across different vegetation covers and then compares three original products and the Merged ET with data from 645 in situ sites. The results indicate that GLEAM covers the largest area, accounting for 39.1% based on the correlation coefficient criterion and 39.9% based on the error variation criterion. Meanwhile, FLDAS and MEP exhibit similar performance characteristics. The merged ET derived from the ETC method demonstrates the ability to mitigate uncertainty in ET estimates in North American (NA) and European (EU) regions, as well as tundra, forest, grassland, and shrubland areas. This merged ET could be effectively utilized to reduce uncertainty in AET estimates from multiple products for ungauged areas. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of this study.</p>
Full article ">Figure 2
<p>Mean annual global actual evapotranspiration and latitudinal distribution from three gridded products. (<b>a</b>–<b>c</b>) show the mean annual evapotranspiration from (<b>a</b>) GLEAM, (<b>b</b>) FLDAS, and (<b>c</b>) MEP. Note that the AET data for Greenland in GLEAM and FLDAS are missing.</p>
Full article ">Figure 3
<p>The correlation coefficient between AET product and the unknown truth estimated by ETC for (<b>a</b>) GLEAM, (<b>b</b>) FLDAS, and (<b>c</b>) MEP.</p>
Full article ">Figure 4
<p>The best-performing product according to (<b>a</b>) the correlation coefficient and (<b>b</b>) the error variation in each grid.</p>
Full article ">Figure 5
<p>The error variation (err, mm<sup>2</sup>/day<sup>2</sup>) and the estimated correlation coefficient (rho2) from ETC method between the truth and AET products (GLEAM, FLDAS, MEP) of different vegetation covers.</p>
Full article ">Figure 6
<p>Mean global annual actual evapotranspiration and latitudinal distribution from merged ET.</p>
Full article ">Figure 7
<p>In situ comparison of three gridded products and merged ET. Maps showing the difference in the Kling–Gupta efficiency (KGE) metric between three gridded products and merged ET, calculated using observations at flux tower sites in different geographical zones: North America (NA), Europe (EU), Asia (AS), and rest of the world (RW). (<b>a</b>–<b>c</b>) show the KGE difference between (<b>a</b>) GLEAM, (<b>b</b>) FLDAS, and (<b>c</b>) MEP compared to merged ET. Blue (red) tones indicate an improvement (degradation) in merged ET compared to the respective gridded products.</p>
Full article ">Figure 8
<p>KGE of three gridded products and merged ET compared to in situ data under different vegetation covers.</p>
Full article ">Figure 9
<p>Multi-year average monthly mean of three products and merged ET under different vegetation covers.</p>
Full article ">Figure 10
<p>Comparison of the physical accuracy of GLEAM, FLDAS, MEP, and merged ET with flux tower measurements, assessed by (<b>a</b>) the coefficient of determination (R2), (<b>b</b>) the mean absolute error (MAE, mm/mon), (<b>c</b>) the root mean square error (RMSE, mm/mon), and (<b>d</b>) the similarity indicator (SI).</p>
Full article ">Figure A1
<p>Classification of vegetation cover in the world, with six covers of vegetation and spatial distribution of the flux towers used in the study.</p>
Full article ">Figure A2
<p>In situ comparison of merged ET. Maps showing the Kling–Gupta efficiency (KGE) metric between merged ET and flux tower sites data in different geographical zones: North America (NA), Europe (EU), Asia (AS), and rest of the world (RW). Blue (red) tones indicate accuracy (inaccuracy) in merged ET compared to in situ data.</p>
Full article ">Figure A3
<p>The weight of GLEAM product in evapotranspiration merging.</p>
Full article ">Figure A4
<p>The weight of FLDAS product in evapotranspiration merging.</p>
Full article ">Figure A5
<p>The weight of FLDAS product in evapotranspiration merging.</p>
Full article ">Figure A6
<p>Global monthly means of GLEAM, FLDAS, MEP, and merged ET from 2003 to 2018.</p>
Full article ">Figure A7
<p>Seasonal means of GLEAM, FLDAS, MEP, and merged ET in 2003–2018. (<b>a</b>–<b>d</b>) The seasonal annual mean of (<b>a</b>) spring, (<b>b</b>) summer, (<b>c</b>) autumn, and (<b>d</b>) winter.</p>
Full article ">
17 pages, 10012 KiB  
Article
Arctic Sea Ice Albedo Estimation from Fengyun-3C/Visible and Infra-Red Radiometer
by Xiaohui Sun and Lei Guan
Remote Sens. 2024, 16(10), 1719; https://doi.org/10.3390/rs16101719 - 12 May 2024
Viewed by 1521
Abstract
The sea ice albedo can amplify global climate change and affect the surface energy in the Arctic. In this paper, the data from Visible and Infra-Red Radiometer (VIRR) onboard Fengyun-3C satellite are applied to derive the Arctic sea ice albedo. Two radiative transfer [...] Read more.
The sea ice albedo can amplify global climate change and affect the surface energy in the Arctic. In this paper, the data from Visible and Infra-Red Radiometer (VIRR) onboard Fengyun-3C satellite are applied to derive the Arctic sea ice albedo. Two radiative transfer models, namely, 6S and FluxNet, are used to simulate the reflectance and albedo in the shortwave band. Clear sky sea ice albedo in the Arctic region (60°~90°N) from 2016 to 2019 is derived through the physical process, including data preprocessing, narrowband to broadband conversion, anisotropy correction, and atmospheric correction. The results are compared with aircraft measurements and AVHRR Polar Pathfinder-Extended (APP-x) albedo product and OLCI MPF product. The bias and standard deviation of the difference between VIRR albedo and aircraft measurements are −0.040 and 0.071, respectively. Compared with APP-x product and OLCI MPF product, a good consistency of albedo is shown. And analyzed together with melt pond fraction, an obvious negative relationship can be seen. After processing the 4-year data, an obvious annual trend can be observed. Due to the influence of snow on the ice surface, the average surface albedo of the Arctic in March and April can reach more than 0.8. Starting in May, with the ice and snow melting and melt ponds forming, the albedo drops rapidly to 0.5–0.6. Into August, the melt ponds begin to freeze and the surface albedo increases. Full article
Show Figures

Figure 1

Figure 1
<p>The process flow of the retrieval algorithm.</p>
Full article ">Figure 2
<p>Cloud detection tree.</p>
Full article ">Figure 3
<p>Clear sky reflectance of channel 1 after cloud detection on 6 June 2017.</p>
Full article ">Figure 4
<p>Spectral reflectance curves processed from MOSAIC.</p>
Full article ">Figure 5
<p>Clear sky broadband reflectance at the top of atmosphere on 6 June 2017.</p>
Full article ">Figure 6
<p>Clear sky albedo at the top of atmosphere after anisotropy correction on 6 June 2017.</p>
Full article ">Figure 7
<p>Relationship between the top of atmospheric albedo and surface albedo varied with the total column ozone (<b>a</b>), aerosol optical depth (<b>b</b>), and total column water vapor (<b>c</b>) (original condition: the solar zenith angle is 60°, the aerosol optical depth is 0.25, the total column ozone is 6.96 g/m<sup>2</sup>, and the total column water vapor is 1.0 g/cm<sup>2</sup>).</p>
Full article ">Figure 8
<p>Clear sky sea ice albedo on 6 June 2017.</p>
Full article ">Figure 9
<p>Space distribution of aircraft measurement data matched up with VIRR albedo.</p>
Full article ">Figure 10
<p>Scatterplot (<b>a</b>) and frequency (<b>b</b>) distribution of the VIRR albedo and aircraft measurements.</p>
Full article ">Figure 11
<p>Daily average broadband albedo scatterplots of retrieval (blue lines) and APP-x product (orange lines) in 2016 (<b>a</b>), 2017 (<b>b</b>), 2018 (<b>c</b>), and 2019 (<b>d</b>).</p>
Full article ">Figure 12
<p>Daily average broadband albedo scatterplots of retrieval (blue lines) and OLCI albedo (orange lines) and melt pond fraction product (green lines) in 2017 (<b>a</b>), 2018 (<b>b</b>), and 2019 (<b>c</b>).</p>
Full article ">Figure 13
<p>Monthly average sea ice albedo map in (<b>a</b>) March, (<b>b</b>) April, (<b>c</b>) May, (<b>d</b>) June, (<b>e</b>) July, and (<b>f</b>) August, from 2016 to 2019.</p>
Full article ">
17 pages, 6792 KiB  
Technical Note
Retrieval of Surface Energy Fluxes Considering Vegetation Changes and Aerosol Effects
by Lijuan Chen, Haishan Chen, Xinguan Du and Ren Wang
Remote Sens. 2024, 16(4), 668; https://doi.org/10.3390/rs16040668 - 13 Feb 2024
Cited by 1 | Viewed by 1424
Abstract
The exchange of moisture and energy between the land and the atmosphere plays a crucial role in terrestrial hydrological cycle and climate change. However, existing studies on the retrieval of surface water and heat flux tend to overlook the dynamic changes in surface [...] Read more.
The exchange of moisture and energy between the land and the atmosphere plays a crucial role in terrestrial hydrological cycle and climate change. However, existing studies on the retrieval of surface water and heat flux tend to overlook the dynamic changes in surface vegetation and atmospheric aerosols, which directly affect surface energy and indirectly alter various meteorological factors, including cloud, precipitation, and temperature. In this study, we assess the machine-learning retrieval method for surface fluxes that takes into account vegetation changes and aerosol effects, using FLUXNET observations and remote sensing data to retrieve latent heat flux (LE) and sensible heat flux (H). We constructed four sets of deep neural network models: (a) The first set considers only meteorological factors, (b) the second set considers meteorological factors and aerosols, (c) the third set considers meteorological factors and vegetation changes, and (d) the fourth set comprehensively considers meteorological factors, aerosols, and vegetation changes. All model performances were evaluated using statistical indicators. ERA5 reanalysis and remote sensing data were used to drive the models and retrieve daily H and LE. The retrieved results were validated against ground observation sites that were not involved in model training or the FLUXCOM product. The results show that the model that considers meteorological factors, aerosols, and vegetation changes has the smallest errors and highest correlation for retrieving H and LE (RH = 0.85, RMSEH = 24.88; RLE = 0.88, RMSELE = 22.25). The ability of the four models varies under different vegetation types. In terms of seasons, the models that consider meteorological factors and vegetation changes, as well as those that comprehensively consider meteorological factors, aerosols, and vegetation changes, perform well in retrieving the surface fluxes. As for spatial distribution, when atmospheric aerosols are present in the region, the model that considers both meteorological factors and aerosols retrieves higher values of H compared to the model that considers only meteorological factors, while the LE values are relatively lower. The model that considers meteorological factors and vegetation changes, as well as the model that comprehensively considers meteorological factors, aerosols, and vegetation changes, retrieves lower values in most regions. Through the validation of independent observation sites and FLUXCOM products, we found that the model, considering meteorological factors, aerosols, and vegetation changes, was generally more accurate in the retrieval of surface fluxes. This study contributes to improving the retrieval and future prediction accuracy of surface fluxes in a changing environment. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Spatial distribution of flux tower observation stations selected in the study.</p>
Full article ">Figure 2
<p>Scatter density diagram of ANN model validation at daily scale (<b>a</b>,<b>b</b>) represent the first model considering only meteorological factors; (<b>c</b>,<b>d</b>) represent the second model considering meteorological factors and aerosols; (<b>e</b>,<b>f</b>) represent the third model considering meteorological factors and vegetation variation; (<b>g</b>,<b>h</b>) represent the fourth model comprehensively considering meteorological factors, aerosols, and vegetation variation. Number of sample points 6372. The black solid line is the best linear fit line and the grey dashed line is the 1:1 fit line.</p>
Full article ">Figure 3
<p>R and RMSE of four models under different vegetation type covers. (<b>a</b>) Correlation of H predicted by the four models with observed data. (<b>b</b>) Correlation of LE predicted by the four models with observed data. (<b>c</b>) Errors between H predicted by the four models and the observed data. (<b>d</b>) Errors between LE predicted by the four models and the observed data. (The first model only considers meteorological factors; the second model considers both meteorological factors and aerosols; the third model considers the changes in meteorological factors and vegetation; and the fourth model comprehensively considers meteorological factors, aerosols, and vegetation changes).</p>
Full article ">Figure 4
<p>R and RMSE for the four models in different seasons. (<b>a</b>) Correlation of H predicted by the four models with observed data. (<b>b</b>) Correlation of LE predicted by the four models with observed data. (<b>c</b>) Errors between H predicted by the four models and the observed data. (<b>d</b>) Errors between LE predicted by the four models and the observed data.</p>
Full article ">Figure 5
<p>Spatial distribution of AOD and LAI in China and Europe on 1 June 2018 ((<b>a</b>,<b>b</b>) show the spatial distribution of AOD, (<b>c</b>,<b>d</b>) show the spatial distribution of LAI).</p>
Full article ">Figure 6
<p>Spatial distribution of H and LE for the four model retrievals on 1 June 2018 in parts of Europe. ((<b>a</b>–<b>d</b>) are the H values retrieved by the first to fourth models. (<b>e</b>–<b>h</b>) are the LE retrieved by the first to the fourth models. The first model considers only meteorological factors. The second model considers meteorological factors and aerosols. The third model considers meteorological factors and vegetation changes, and the fourth model combines meteorological factors, aerosols, and vegetation changes).</p>
Full article ">Figure 7
<p>Spatial distribution of H and LE for four model retrievals on 1 June 2018 in China. ((<b>a</b>–<b>d</b>) are the first to fourth model retrieved H. (<b>e</b>–<b>h</b>) are the first to fourth model retrieved LE).</p>
Full article ">Figure 8
<p>Validation results for H and LE at regional scales (The first model: (<b>a</b>,<b>b</b>) considering only meteorological factors. The second model: (<b>c</b>,<b>d</b>) considering meteorological factors and aerosols. The third model: (<b>e</b>,<b>f</b>) considering meteorological factors and vegetation changes. The verification results of (<b>g</b>,<b>h</b>) are comprehensive considerations of three factors: meteorological factors, aerosols, and vegetation changes). The solid red line is the line of best linear fit.</p>
Full article ">Figure 9
<p>The validation results of the four sets of models and FLUXCOM product at the corresponding location of this site are as follows.(The first model: (<b>a</b>,<b>b</b>) considering only meteorological factors. The second model: (<b>c</b>,<b>d</b>) considering meteorological factors and aerosols. The third model: (<b>e</b>,<b>f</b>) considering meteorological factors and vegetation changes. The verification results of (<b>g</b>,<b>h</b>) are comprehensive considerations of three factors: meteorological factors, aerosols, and vegetation changes). The solid red line is the line of best linear fit.</p>
Full article ">
31 pages, 10443 KiB  
Article
The Response of Daily Carbon Dioxide and Water Vapor Fluxes to Temperature and Precipitation Extremes in Temperate and Boreal Forests
by Daria Gushchina, Maria Tarasova, Elizaveta Satosina, Irina Zheleznova, Ekaterina Emelianova, Ravil Gibadullin, Alexander Osipov and Alexander Olchev
Climate 2023, 11(10), 206; https://doi.org/10.3390/cli11100206 - 12 Oct 2023
Cited by 1 | Viewed by 2266
Abstract
Forest ecosystems in the mid-latitudes of the Northern Hemisphere are significantly affected by frequent extreme weather events. How different forest ecosystems respond to these changes is a major challenge. This study aims to assess differences in the response of daily net ecosystem exchange [...] Read more.
Forest ecosystems in the mid-latitudes of the Northern Hemisphere are significantly affected by frequent extreme weather events. How different forest ecosystems respond to these changes is a major challenge. This study aims to assess differences in the response of daily net ecosystem exchange (NEE) of CO2 and latent heat flux (LE) between different boreal and temperate ecosystems and the atmosphere to extreme weather events (e.g., anomalous temperature and precipitation). In order to achieve the main objective of our study, we used available reanalysis data and existing information on turbulent atmospheric fluxes and meteorological parameters from the global and regional FLUXNET databases. The analysis of NEE and LE responses to high/low temperature and precipitation revealed a large diversity of flux responses in temperate and boreal forests, mainly related to forest type, geographic location, regional climate conditions, and plant species composition. During the warm and cold seasons, the extremely high temperatures usually lead to increased CO2 release in all forest types, with the largest response in coniferous forests. The decreasing air temperatures that occur during the warm season mostly lead to higher CO2 uptake, indicating more favorable conditions for photosynthesis at relatively low summer temperatures. The extremely low temperatures in the cold season are not accompanied by significant NEE anomalies. The response of LE to temperature variations does not change significantly throughout the year, with higher temperatures leading to LE increases and lower temperatures leading to LE reductions. The immediate response to heavy precipitation is an increase in CO2 release and a decrease in evaporation. The cumulative effect of heavy precipitations is opposite to the immediate effect in the warm season and results in increased CO2 uptake due to intensified photosynthesis in living plants under sufficient soil moisture conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Number of grid cells in the area north of 40° N of the ERA5 reanalysis where the trend turning point was detected for air temperature (<b>a</b>) and precipitation (<b>b</b>) time series.</p>
Full article ">Figure 2
<p>Number of trend turning points detected in each grid cell for the period 1979–2021 for air temperature (<b>a</b>) and precipitation (<b>b</b>) north of 40° N.</p>
Full article ">Figure 3
<p>The difference between the late 20th century (1980–2000) and the early 21st century (2001–2021) in the number of months with the temperature exceeding the 95% quantile (color) for the warm period—April–September (<b>a</b>) and the cold period—October–March (<b>b</b>). Overlaid stations are the FLUXNET stations located in the areas of maximum changes in extreme temperature (dots), which were selected for further analysis.</p>
Full article ">Figure 4
<p>The difference between the late 20th (1980–2000) and early 21st (2001–2021) centuries in the number of months with precipitation exceeding the 95% quantile (color) for the warm period—April–September (<b>a</b>) and the cold period—October–March (<b>b</b>). Overlaid stations are the FLUXNET stations located in the areas of maximum changes in extreme precipitation (dots), which were selected for further analysis.</p>
Full article ">Figure 5
<p>The FLUXNET stations of the Northern Hemisphere with vegetation types according to the ecosystem classification used in the FLUXNET archive.</p>
Full article ">Figure 6
<p>The forest stations of global and regional FLUXNET networks selected for the analysis of the flux response to extreme weather conditions. Maps of the monitoring stations are overlaid with maps of the Köppen climate classification (<b>a</b>) and vegetation types (<b>b</b>) according to the ecosystem classification used in the FLUXNET archive. The color of the circles correspond to the forest types, the number inside the circle—to the number of the monitoring station (<a href="#climate-11-00206-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 7
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>,<b>c</b>) and low (<b>b</b>,<b>d</b>) temperatures during the warm (<b>a</b>,<b>b</b>) and cold (<b>c</b>,<b>d</b>) seasons in evergreen needleleaf forests (see text for details). The number of cases where the anomalies of temperature and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 8
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>) and low (<b>b</b>) temperatures during the warm season in mixed and evergreen broadleaf forests (see text for details). The number of cases where the anomalies of temperature and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 9
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>) and low (<b>b</b>) temperatures during the warm season in deciduous broadleaf and needleleaf forests (see text for details). The number of cases where the anomalies of temperature and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 10
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>) and low (<b>b</b>) temperatures during the cold season in mixed evergreen broadleaf, deciduous broadleaf, and needleleaf forests (see text for details). The number of cases where the anomalies of temperature and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 11
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high precipitation during the warm (<b>a</b>) and cold (<b>b</b>) seasons in evergreen needleleaf forests (see text for details). The number of cases where the anomalies of precipitation and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 12
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high precipitation during the warm (<b>a</b>,<b>b</b>) and cold (<b>c</b>) seasons in mixed and evergreen broadleaf forests (<b>a</b>), deciduous broadleaf and needleleaf forests (<b>b</b>), and mixed deciduous broadleaf and needleleaf forests (<b>c</b>) (see text for details). The number of cases where the anomalies of precipitation and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 13
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>,<b>c</b>) and low (<b>b</b>,<b>d</b>) API during warm (<b>a</b>,<b>b</b>) and cold (<b>c</b>,<b>d</b>) seasons in evergreen needleleaf forests (see text for details). The number of cases where the anomalies of API and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 14
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>) and low (<b>b</b>) API during the warm season in mixed and evergreen broadleaf forests (see text for details). The number of cases where the anomalies of API and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 15
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>) and low (<b>b</b>) API during the warm season in deciduous broadleaf and needleleaf forests (see text for details). The number of cases where the anomalies of API and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">Figure 16
<p>The percentage of days when NEE and LE anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>) and low (<b>b</b>) API during the cold season in mixed evergreen broadleaf, deciduous broadleaf, and needleleaf forests (see text for details). The number of cases where the anomalies of API and NEE/LE were observed on the same day is shown above the bars.</p>
Full article ">
17 pages, 6188 KiB  
Article
Evaluation of Ecosystem Water Use Efficiency Based on Coupled and Uncoupled Remote Sensing Products for Maize and Soybean
by Lingxiao Huang, Meng Liu and Na Yao
Remote Sens. 2023, 15(20), 4922; https://doi.org/10.3390/rs15204922 - 12 Oct 2023
Cited by 5 | Viewed by 1319
Abstract
Accurate quantification of ecosystem water use efficiency (eWUE) over agroecosystems is crucial for managing water resources and assuring food security. Currently, the uncoupled Moderate Resolution Imaging Spectroradiometer (MODIS) product is the most widely applied dataset for simulating local, regional, and global eWUE across [...] Read more.
Accurate quantification of ecosystem water use efficiency (eWUE) over agroecosystems is crucial for managing water resources and assuring food security. Currently, the uncoupled Moderate Resolution Imaging Spectroradiometer (MODIS) product is the most widely applied dataset for simulating local, regional, and global eWUE across different plant functional types. However, it has been rarely investigated as to whether the coupled product can outperform the uncoupled product in eWUE estimations for specific C4 and C3 crop species. Here, the eWUE as well as gross primary production (GPP) and evapotranspiration (ET) from the uncoupled MODIS product and the coupled Penman–Monteith–Leuning version 2 (PMLv2) product were evaluated against the in-situ observations on eight-day and annual scales (containing 1902 eight-day and 61 annual samples) for C4 maize and C3 soybean at the five cropland sites from the FLUXNET2015 and AmeriFlux datasets. Our results show the following: (1) For GPP estimates, the PMLv2 product showed paramount improvements for C4 maize and slight improvements for C3 soybean, relative to the MODIS product. (2) For ET estimates, both products performed similarly for both crop species. (3) For eWUE estimates, the coupled PMLv2 product achieved higher-accuracy eWUE estimates than the uncoupled MODIS product at both eight-day and annual scales. Taking the result at an eight-day scale for example, compared to the MODIS product, the PMLv2 product could reduce the root mean square error (RMSE) from 2.14 g C Kg−1 H2O to 1.36 g C Kg−1 H2O and increase the coefficient of determination (R2) from 0.06 to 0.52 for C4 maize, as well as reduce the RMSE from 1.33 g C Kg−1 H2O to 0.89 g C Kg−1 H2O and increase the R2 from 0.05 to 0.49 for C3 soybean. (4) Despite the outperformance of the PMLv2 product in eWUE estimations, both two products failed to differentiate C4 and C3 crop species in their model calibration and validation processes, leading to a certain degree of uncertainties in eWUE estimates. Our study not only provides an important reference for applying remote sensing products to derive reliable eWUE estimates over cropland but also indicates the future modification of the current remote sensing models for C4 and C3 crop species. Full article
Show Figures

Figure 1

Figure 1
<p>Satellite images over a grid area of 500 m × 500 m centered at five cropland sites accessed through Google Maps.</p>
Full article ">Figure 2
<p>Scatter plots of the eight-day observed GPP and RS-based GPP from the PMLv2 and MOD17 products for C4 maize and C3 soybean at the five cropland sites. The statistical metrics with green, orange, and black colors represent the performance of the PMLv2 and MOD17 products for maize, soybean, and both species, respectively.</p>
Full article ">Figure 3
<p>Scatter plots of the eight-day observed ET and RS-based ET from the PMLv2 and MOD16 products for C4 maize and C3 soybean at the five cropland sites. The statistical metrics with green, orange, and black colors represent the performance of the PMLv2 and MOD17 products for maize, soybean, and both species, respectively.</p>
Full article ">Figure 4
<p>Time series of the eight-day observed GPP and RS-based GPP from the PMLv2 and MOD16 products at the five cropland sites. The years with white backgrounds represent that C4 maize was planted this year. The years with grey backgrounds represent that C3 soybean was planted this year.</p>
Full article ">Figure 5
<p>Time series of the eight-day observed ET and RS-based ET from the PMLv2 and MOD16 products at the five cropland sites. The years with white backgrounds represent that C4 maize was planted this year. The years with grey backgrounds represent that C3 soybean was planted this year.</p>
Full article ">Figure 6
<p>Scatter plots of the eight-day observed eWUE and RS-based eWUE from the PMLv2 and MODIS products for C4 maize and C3 soybean at the five cropland sites. The PMLv2 eWUE was calculated as the ratio of GPP and ET from the PMLv2 product. The MODIS eWUE was calculated as the ratio of GPP from the MOD17 product and ET from the MOD16 product.</p>
Full article ">Figure 7
<p>Time series of the eight-day observed eWUE and RS-based eWUE from the PMLv2 and MODIS products at the five cropland sites. The PMLv2 eWUE was calculated as the ratio of GPP and ET from the PMLv2 product. The MODIS eWUE was calculated as the ratio of GPP from the MOD17 product and ET from the MOD16 product. The years with white backgrounds represent that C4 maize was planted this year. The years with grey backgrounds represent that C3 soybean was planted this year.</p>
Full article ">Figure 8
<p>Scatter plots of the annual observed and RS-based GPP, ET, and eWUE at the five cropland sites from the PMLv2 products (<b>left</b>) and MODIS products (<b>right</b>) for C4 maize and C3 soybean. The PMLv2 eWUE was calculated as the ratio of GPP and ET from the PMLv2 product. The MODIS eWUE was calculated as the ratio of GPP from the MOD17 product and ET from the MOD16 product.</p>
Full article ">
28 pages, 11039 KiB  
Article
High-Resolution Estimation of Methane Emissions from Boreal and Pan-Arctic Wetlands Using Advanced Satellite Data
by Yousef A. Y. Albuhaisi, Ype van der Velde, Richard De Jeu, Zhen Zhang and Sander Houweling
Remote Sens. 2023, 15(13), 3433; https://doi.org/10.3390/rs15133433 - 6 Jul 2023
Cited by 2 | Viewed by 2972
Abstract
This paper investigates the use of soil moisture data from satellites and a hydrological model as inputs to a simplified CH4 emission model (MeSMOD) for estimating CH4 emissions from boreal and pan-Arctic regions between 2015 and 2021. MeSMOD is calibrated using [...] Read more.
This paper investigates the use of soil moisture data from satellites and a hydrological model as inputs to a simplified CH4 emission model (MeSMOD) for estimating CH4 emissions from boreal and pan-Arctic regions between 2015 and 2021. MeSMOD is calibrated using FLUXNET—CH4 sites and the predictive performance is evaluated using several metrics, including the Nash-Sutcliffe efficiency (NSE). Using satellite soil moisture with 100 m resolution, MeSMOD has the highest performance (NSE = 0.63) compared with using satellite soil moisture of 10 km and hydrological model soil moisture of 10 km and 50 km (NSE = 0.59, 0.56, and 0.53, respectively) against site-level CH4 flux. This study has upscaled the estimates to the pan-Arctic region using MeSMOD, resulting in comparable mean annual estimates of CH4 emissions using satellite soil moisture of 10 km (33 Tg CH4 yr−1) and hydrological model soil moisture of 10 km (39 Tg CH4 yr−1) compared with previous studies using random forest technique for upscaling (29.5 Tg CH4 yr−1), LPJ-wsl process model (30 Tg CH4 yr−1), and CH4 CAMS inversion (34 Tg CH4 yr−1). MeSMOD has also accurately captured the high methane emissions observed by LPJ-wsl and CAMS in 2016 and 2020 and effectively caught the interannual variability of CH4 emissions from 2015 to 2021. The study emphasizes the importance of using high-resolution satellite soil moisture data for accurate estimation of CH4 emissions from wetlands, as these data directly reflect soil moisture conditions and lead to more reliable estimates. The approach adopted in this study helps to reduce errors and improve our understanding of wetlands’ role in CH4 emissions, ultimately reducing uncertainties in global CH4 budgets. Full article
(This article belongs to the Section Biogeosciences Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The CLCC wetland map, representing the wetland fraction in 2020, for the study domain.</p>
Full article ">Figure 2
<p>Schematic overview of the PCR-GLOBWB model. Figure adapted from Sutanudjaja et al. (2018) [<a href="#B20-remotesensing-15-03433" class="html-bibr">20</a>].</p>
Full article ">Figure 3
<p>Simplified flowchart of Planet data processing. Raw satellite brightness temperature data are processed via downscaling and are run through the retrieval algorithm land parameter retrieval model (LPRM) before being linked to land surface variables, such as land surface temperature (LST), vegetation optical depth (VOD), and the soil dielectric constant, which is related to soil moisture (SM). This figure is adapted from De Jeu et al. [<a href="#B58-remotesensing-15-03433" class="html-bibr">58</a>].</p>
Full article ">Figure 4
<p>The Taylor diagram shown in this figure is a graphical representation of the performance of different soil moisture (SM) values in simulating CH<sub>4</sub> emissions compared with all FLUXNET—CH<sub>4</sub> sites used in the study. The angle between the blue line and the x-axis represents the correlation coefficient (r) between the observed and simulated CH<sub>4</sub> emissions, indicating how well the simulations capture the variability in the observations. The red dot on the x-axis represents the observations. The distance from the origin (0,0) to the red symbols in the chart represents the standard deviation, while the dashed black half circles represent the root-mean-square deviation (RMSE). The red symbols in the chart represent the simulated values. For more information about the performance of the simulations at individual sites, refer to <a href="#remotesensing-15-03433-f0A1" class="html-fig">Figure A1</a>a,b. The value on the red quarter circle has the same standard deviation as the observations.</p>
Full article ">Figure 5
<p>Group-1 CH<sub>4</sub> site emission measurements (black) in comparison to simulated emissions using different SM inputs for April 2015 to December 2018 and soil surface temperature (grey).</p>
Full article ">Figure 6
<p>Mean annual CH<sub>4</sub> wetland emissions estimated by upscaling FLUXNET—CH<sub>4</sub> EC data using (<b>a</b>) Planet satellite SSM 10 km (2015–2018); (<b>b</b>) PCR-GLOBWB hydrological HSM 10 km (2015–2018); (<b>c</b>) CH<sub>4</sub> from CAMS inversion (2015–2018); (<b>d</b>–<b>f</b>) RF-PEATMAP, RF-DYPTOP, and RF-GLWD models using three different wetland maps (2013–2014); (<b>g</b>) WetCHARTs extended ensemble (mean of all models) (2001–2015); and (<b>h</b>) LPJ-wsl (2015–2018).</p>
Full article ">Figure 7
<p>Seasonal variation of CH<sub>4</sub> of the upscaled MeSMOD emissions in comparison to different models covering the same study area.</p>
Full article ">Figure 8
<p>(<b>a</b>) HB and (<b>b</b>) WSL emission anomalies compared with North American and Eurasian anomalies, respectively; (<b>c</b>) HB and (<b>d</b>) WSL contribution ratio to the CH<sub>4</sub> emission annual total budget for North America and Eurasia, respectively.</p>
Full article ">Figure 9
<p>Study area HSM_10 km and SSM_10 km mean seasonal interannual variability for the years 2015–2021.</p>
Full article ">Figure 10
<p>Monthly anomalies of the mean SM from Planet and PCR-GLOBWB in comparison to WTD expressed in units of SD.</p>
Full article ">Figure A1
<p>(<b>a</b>) Detailed Taylor diagram for group 1 sites for CH<sub>4</sub> simulations using different SM in comparison to observations. The cosine (blue) of the angle from the x-axis is the correlation coefficient between observed and simulated CH<sub>4</sub> emission, which determines how well the simulated CH<sub>4</sub> emission captures the daily interannual variability of the observed CH<sub>4</sub> emission from FLUXNET—CH<sub>4</sub> sites. The red dot on the x-axis represents observations. The radial distance from the origin (0,0) is the standard deviation. The dashed green half circles represent RMSE. The red symbols floating in the chart are the simulations. (<b>b</b>) Same as <a href="#remotesensing-15-03433-f0A1" class="html-fig">Figure A1</a>a, but for group 2 sites.</p>
Full article ">Figure A1 Cont.
<p>(<b>a</b>) Detailed Taylor diagram for group 1 sites for CH<sub>4</sub> simulations using different SM in comparison to observations. The cosine (blue) of the angle from the x-axis is the correlation coefficient between observed and simulated CH<sub>4</sub> emission, which determines how well the simulated CH<sub>4</sub> emission captures the daily interannual variability of the observed CH<sub>4</sub> emission from FLUXNET—CH<sub>4</sub> sites. The red dot on the x-axis represents observations. The radial distance from the origin (0,0) is the standard deviation. The dashed green half circles represent RMSE. The red symbols floating in the chart are the simulations. (<b>b</b>) Same as <a href="#remotesensing-15-03433-f0A1" class="html-fig">Figure A1</a>a, but for group 2 sites.</p>
Full article ">Figure A2
<p>Group-2 CH<sub>4</sub> sites emission in comparison to simulated emissions using different SM model inputs. The years vary depending on site data availability.</p>
Full article ">Figure A3
<p>(<b>a</b>) Wetland map used in LPJ-wsl simulations; (<b>b</b>) the CLCC wetland map used in this study. The dashed hexagons show the locations of the wetland deviation from each other.</p>
Full article ">
21 pages, 3331 KiB  
Article
Effects of Extreme Temperature and Precipitation Events on Daily CO2 Fluxes in the Tropics
by Daria Gushchina, Maria Tarasova, Elizaveta Satosina, Irina Zheleznova, Ekaterina Emelianova, Elena Novikova and Alexander Olchev
Climate 2023, 11(6), 117; https://doi.org/10.3390/cli11060117 - 25 May 2023
Cited by 1 | Viewed by 2229
Abstract
The effects of anomalous weather conditions (such as extreme temperatures and precipitation) on CO2 flux variability in different tropical ecosystems were assessed using available reanalysis data, as well as information about daily net CO2 fluxes from the global FLUXNET database. A [...] Read more.
The effects of anomalous weather conditions (such as extreme temperatures and precipitation) on CO2 flux variability in different tropical ecosystems were assessed using available reanalysis data, as well as information about daily net CO2 fluxes from the global FLUXNET database. A working hypothesis of the study suggests that the response of tropical vegetation can differ depending on local geographical conditions and intensity of temperature and precipitation anomalies. The results highlighted the large diversity of CO2 flux responses to the fluctuations of temperature and precipitation in tropical ecosystems that may differ significantly from some previously documented relationships (e.g., higher CO2 emission under the drier and hotter weather, higher CO2 uptake under colder and wetter weather conditions). They showed that heavy precipitation mainly leads to the strong intensification of mean daily CO2 release into the atmosphere at almost all stations and in all types of study biomes. For the majority of considered tropical ecosystems, the intensification of daily CO2 emission during cold and wet weather was found, whereas the ecosystems were predominantly served as CO2 sinks from the atmosphere under hot/dry conditions. Such disparate responses suggested that positive and negative temperature and precipitation anomalies influence Gross Primary Production (GPP) and Ecosystem Respiration (ER) rates differently that may result in various responses of Net Ecosystem Exchanges (NEE) of CO2 to external impacts. Their responses may also depend on various local biotic and abiotic factors, including plant canopy age and structure, plant biodiversity and plasticity, soil organic carbon and water availability, surface topography, solar radiation fluctuation, etc. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the monitoring stations with the type of biome indicated by the color (see the map legend) and the available period of observations (indicated in the brackets).</p>
Full article ">Figure 2
<p>The time series of daily net CO<sub>2</sub> flux (NEE) anomaly and daily temperature anomalies (<b>a</b>–<b>c</b>), daily precipitation amount (<b>d</b>,<b>e</b>) at flux monitoring station AU-Dry (Dry River) for the period January–July 2012. The days when the daily CO<sub>2</sub> flux anomalies were greater (lower) than 1 STD of CO<sub>2</sub> time series for this station are marked by red (blue) dots. The red (blue) shading is applied for the periods when the temperature exceeds the upper (lower) threshold: (<b>a</b>) 1 STD (−1 STD); (<b>b</b>) 90% (10%); (<b>c</b>) 95% (5%) PDF quantile. The red shaded column (blue triangle) is applied for the days when the precipitation daily amount exceeds the upper (lower) threshold: (<b>d</b>) 90% (10%); (<b>e</b>) 95% (5%) PDF quantile.</p>
Full article ">Figure 3
<p>The percentage of the days when daily net CO<sub>2</sub> flux (NEE) anomalies greater than 1 STD occurred simultaneously with extremely high (<b>a</b>), low (<b>c</b>) temperatures, and extremely high (<b>b</b>) and low (<b>d</b>) precipitation in savannas, grasslands, and permanent wetlands (see the text for the details).</p>
Full article ">Figure 4
<p>The same as <a href="#climate-11-00117-f003" class="html-fig">Figure 3</a> but for tropical evergreen, deciduous needle leaf, tropical rain, and dry or seasonal forests.</p>
Full article ">Figure 5
<p>The percentage of the days when temperature anomalies (<b>a</b>,<b>c</b>) and daily precipitation rates (<b>b</b>,<b>d</b>), exceeding thresholds, occurred simultaneously with extremely positive (<b>a</b>,<b>b</b>) and negative (<b>c</b>,<b>d</b>) daily net CO<sub>2</sub> flux (NEE) anomalies in savannas, grasslands, and permanent wetlands (see the text for the details).</p>
Full article ">Figure 6
<p>The same as <a href="#climate-11-00117-f005" class="html-fig">Figure 5</a> but for tropical evergreen, deciduous needle leaf, tropical rain, and dry or seasonal forests.</p>
Full article ">Figure 7
<p>The percentage of the days with daily net CO<sub>2</sub> flux (NEE) anomalies, exceeding STD, observed during the periods of extremely hot/wet (<b>a</b>), cold/wet (<b>b</b>), hot/dry (<b>c</b>), and cold/dry (<b>d</b>) conditions.</p>
Full article ">
21 pages, 9328 KiB  
Article
Shifted Global Vegetation Phenology in Response to Climate Changes and Its Feedback on Vegetation Carbon Uptake
by Husheng Fang, Moquan Sha, Yichun Xie, Wenjuan Lin, Dai Qiu, Jiangguang Tu, Xicheng Tan, Xiaolei Li and Zongyao Sha
Remote Sens. 2023, 15(9), 2288; https://doi.org/10.3390/rs15092288 - 26 Apr 2023
Cited by 4 | Viewed by 2732
Abstract
Green vegetation plays a vital role in energy flows and matter cycles in terrestrial ecosystems, and vegetation phenology may not only be influenced by, but also impose active feedback on, climate changes. The phenological events of vegetation such as the start of season [...] Read more.
Green vegetation plays a vital role in energy flows and matter cycles in terrestrial ecosystems, and vegetation phenology may not only be influenced by, but also impose active feedback on, climate changes. The phenological events of vegetation such as the start of season (SOS), end of season (EOS), and length of season (LOS) can respond to climate changes and affect gross primary productivity (GPP). Here, we coupled satellite remote sensing imagery with FLUXNET observations to systematically map the shift of SOS, EOS, and LOS in global vegetated area, and explored their response to climate fluctuations and feedback on GPP during the last two decades. The results indicated that 11.5% of the global vegetated area showed a significantly advanced trend in SOS, and that only 5.2% of the area presented significantly delayed EOS during the past two decades, resulting in significantly prolonged LOS in 12.6% of the vegetated area. The climate factors, including seasonal temperature and precipitation, attributed to the shifts in vegetation phenology, but with high spatial and temporal difference. LOS was positively and significantly correlated with GPP in 20.2% of the total area, highlighting that longer LOS is likely to promote vegetation productivity. The feedback on GPP from the shifted vegetation phenology may serve as an adaptation mechanism for terrestrial ecosystems to mitigate global warming through improved carbon uptake from the atmosphere. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Global distribution of vegetation types and locations of FLUXNET (Flux) sites, where land cover type products from MCD12Q1 are reclassified into five groups, including forest(s) (evergreen needleleaf forests, evergreen broadleaf forests, deciduous needleleaf forests, deciduous broadleaf forests, and mixed forests), shrub (open shrublands and closed shrublands), savannas (woody savannas and savannas), grass (grasslands), and crop (croplands and cropland/natural vegetation mosaics).</p>
Full article ">Figure 2
<p>Typical curve fitting for GPP and vegetation phenology extraction from remotely sensed imagery in the Northern Hemisphere. (<b>a</b>) Time-series remotely-sensed imageries showing time-series changes in vegetation photosynthesis levels, (<b>b</b>) vegetation growth status (GPP) for pixel Pi, (<b>c</b>) smoothed curve, and (<b>d</b>) extracted phenological dates using FOD.</p>
Full article ">Figure 3
<p>Validation of the remote-sensing-derived phenological metrics, on a yearly basis, against the FLUXNET observations through indicators of (<b>a</b>) root mean square error (RMSE), (<b>b</b>) mean error (ME), (<b>c</b>) scatter of SOS, and (<b>d</b>) scatter of EOS.</p>
Full article ">Figure 3 Cont.
<p>Validation of the remote-sensing-derived phenological metrics, on a yearly basis, against the FLUXNET observations through indicators of (<b>a</b>) root mean square error (RMSE), (<b>b</b>) mean error (ME), (<b>c</b>) scatter of SOS, and (<b>d</b>) scatter of EOS.</p>
Full article ">Figure 4
<p>Mean and standard deviation (std) for SOS, EOS, and LOS during 2001–2020, with the mean value aggregated along the latitude gradients at intervals of one degree (see left inset figures). Areas with a lack of data or/and low reliability (no/low obvious seasonality, std ≥ 60 days) are masked out. The critical latitudes at 30° in both the Northern Hemisphere and Southern Hemisphere segment the global vegetation into three distinct parts, the north sector over 30°N northward (North), the south sector over 30°S southward (South), and the middle between them (Middle). Some hotspot regions are labeled, including the southern part of North America, Southern Europe, and Australia in sub-figure (<b>a</b>–<b>d</b>), as well as southeast of the United States, west of Europe, south of China, and Japan in sub-figure (<b>e</b>–<b>f</b>).</p>
Full article ">Figure 4 Cont.
<p>Mean and standard deviation (std) for SOS, EOS, and LOS during 2001–2020, with the mean value aggregated along the latitude gradients at intervals of one degree (see left inset figures). Areas with a lack of data or/and low reliability (no/low obvious seasonality, std ≥ 60 days) are masked out. The critical latitudes at 30° in both the Northern Hemisphere and Southern Hemisphere segment the global vegetation into three distinct parts, the north sector over 30°N northward (North), the south sector over 30°S southward (South), and the middle between them (Middle). Some hotspot regions are labeled, including the southern part of North America, Southern Europe, and Australia in sub-figure (<b>a</b>–<b>d</b>), as well as southeast of the United States, west of Europe, south of China, and Japan in sub-figure (<b>e</b>–<b>f</b>).</p>
Full article ">Figure 5
<p>Trend analysis for the vegetation phenological events during 2001–2020, with the inset figure showing the statistics (percentage of area) from the Sen’s slope trend analysis on the phenological metrics (SOS, EOS, and LOS) as advance/delay and significant advance/delay (advance sig and delay sig) in (<b>a</b>,<b>d</b>), grow season length shortened (short) or prolonged (prolong) and significantly shortened/prolonged (short sig and prolong sig) in (<b>g</b>), and MK significance test on the Sen’s slope (at the probability level of 0.05) with labels’ significance (Sig) and no significance (No sig) in (<b>b</b>,<b>e</b>,<b>h</b>). The spatially averaged values are 0.19, 0.06, and 0.24 days/year for SOS (<b>a</b>), EOS (<b>d</b>), and LOS (<b>g</b>), respectively. Uncertainties reflected by one sigma confidence interval in the trend are given in (<b>c</b>,<b>f</b>,<b>i</b>).</p>
Full article ">Figure 5 Cont.
<p>Trend analysis for the vegetation phenological events during 2001–2020, with the inset figure showing the statistics (percentage of area) from the Sen’s slope trend analysis on the phenological metrics (SOS, EOS, and LOS) as advance/delay and significant advance/delay (advance sig and delay sig) in (<b>a</b>,<b>d</b>), grow season length shortened (short) or prolonged (prolong) and significantly shortened/prolonged (short sig and prolong sig) in (<b>g</b>), and MK significance test on the Sen’s slope (at the probability level of 0.05) with labels’ significance (Sig) and no significance (No sig) in (<b>b</b>,<b>e</b>,<b>h</b>). The spatially averaged values are 0.19, 0.06, and 0.24 days/year for SOS (<b>a</b>), EOS (<b>d</b>), and LOS (<b>g</b>), respectively. Uncertainties reflected by one sigma confidence interval in the trend are given in (<b>c</b>,<b>f</b>,<b>i</b>).</p>
Full article ">Figure 5 Cont.
<p>Trend analysis for the vegetation phenological events during 2001–2020, with the inset figure showing the statistics (percentage of area) from the Sen’s slope trend analysis on the phenological metrics (SOS, EOS, and LOS) as advance/delay and significant advance/delay (advance sig and delay sig) in (<b>a</b>,<b>d</b>), grow season length shortened (short) or prolonged (prolong) and significantly shortened/prolonged (short sig and prolong sig) in (<b>g</b>), and MK significance test on the Sen’s slope (at the probability level of 0.05) with labels’ significance (Sig) and no significance (No sig) in (<b>b</b>,<b>e</b>,<b>h</b>). The spatially averaged values are 0.19, 0.06, and 0.24 days/year for SOS (<b>a</b>), EOS (<b>d</b>), and LOS (<b>g</b>), respectively. Uncertainties reflected by one sigma confidence interval in the trend are given in (<b>c</b>,<b>f</b>,<b>i</b>).</p>
Full article ">Figure 6
<p>Double-peak patterns in the histogram of the correlation coefficient between SOS and temperature for different vegetation types and all types combined, with the statistics (percentage of area) for positive (pos) and negative (neg) correlation (and significant correlation) (<b>a</b>–<b>f</b>). The x-axis was divided into 200 bins for the correlation coefficient located between −1 and 0 (negative correlation) and 0 to 1 (positive correlation), and the y-axis shows the percentage of the count within each bin.</p>
Full article ">Figure 7
<p>Double-peak patterns in the histogram of the correlation coefficient between EOS and temperature for different vegetation types and all types combined, with the statistics (percentage of area) for positive (pos) and negative (neg) correlation (and significant correlation) (<b>a</b>–<b>f</b>). The x-axis was divided into 200 bins for the correlation coefficient located between −1 and 0 (negative correlation), and 0 to 1 (positive correlation), and the y-axis shows the percentage of the count within each bin.</p>
Full article ">Figure 7 Cont.
<p>Double-peak patterns in the histogram of the correlation coefficient between EOS and temperature for different vegetation types and all types combined, with the statistics (percentage of area) for positive (pos) and negative (neg) correlation (and significant correlation) (<b>a</b>–<b>f</b>). The x-axis was divided into 200 bins for the correlation coefficient located between −1 and 0 (negative correlation), and 0 to 1 (positive correlation), and the y-axis shows the percentage of the count within each bin.</p>
Full article ">Figure 8
<p>Correlation coefficient between GPP and LOG, with the inset figure showing the proportion of the statistics (percentage of area) for positive (pos) and negative (neg) correlation and significant correlation (pos/neg sig). The labeled numbers in the figure, #1 for the Midwest of the United States, #2 for the south of central Asia, and #3 for Inner Mongolia in China and South Mongolia, exemplify three key regions showing a significantly negative correlation between LOS and GPP.</p>
Full article ">Figure 9
<p>Comparison of the temporal trends of annual GPP and LOS for the significant positive correlation area, which accounts for 20.2% of the total vegetated area (see <a href="#remotesensing-15-02288-t001" class="html-table">Table 1</a>).</p>
Full article ">
15 pages, 3541 KiB  
Article
Comprehensive Effects of Atmosphere and Soil Drying on Stomatal Behavior of Different Plant Types
by Zhi Xu, Ye Tian, Zhiwu Liu and Xinran Xia
Water 2023, 15(9), 1675; https://doi.org/10.3390/w15091675 - 25 Apr 2023
Cited by 4 | Viewed by 3017
Abstract
The soil water supply and atmospheric humidity conditions are crucial in controlling plants’ stomatal behavior and water use efficiency. When there is water stress caused by an increase in saturated water vapor pressure (VPD) and a decrease in soil water content (SWC), plants [...] Read more.
The soil water supply and atmospheric humidity conditions are crucial in controlling plants’ stomatal behavior and water use efficiency. When there is water stress caused by an increase in saturated water vapor pressure (VPD) and a decrease in soil water content (SWC), plants tend to close stomata to reduce water loss. This affects the gross primary productivity (GPP) and evapotranspiration (ET), subsequently leading to changes in water use efficiency (WUE) and carbon use efficiency (CUE) in plants. However, land–atmosphere interactions mean that water vapor in the atmosphere and soil moisture content causing water stress for plants are closely related. This study aims to compare and estimate the effects of VPD and SWC on the carbon cycle and water cycle for different plant functional types. Based on the fluxnet2015 dataset from around the world, the WUE and CUE of five plant functional types (PFTs) were estimated under varying levels of VPD and SWC. The results showed that high VPD and low SWC limit the stomatal conductance (Gs) and gross primary productivity (GPP) of plants. However, certain types of vegetation (crops, broad-leaved forests) could partially offset the negative effects of high VPD with higher SWC. Notably, higher SWC could even alleviate limitations and partially promote the increase in GPP and net primary production (NPP) with increasing VPD. WUE and CUE were directly affected by Gs and productivity. In general, the increase in VPD in the five PFTs was the dominant factor in changing WUE and CUE. The impact of SWC limitations on CUE was minimal, with an overall impact of only −0.05μmol/μmol on the four PFTs. However, the CUE of savanna plants changed differently from the other four PFTs. The rise in VPD dominated the changes in CUE, and there was an upward trend as SWC declined, indicating that the increase in VPD and decrease in SWC promote the increase in the CUE of savanna plants to some extent. Full article
(This article belongs to the Special Issue Challenges of Hydrological Drought Monitoring and Prediction)
Show Figures

Figure 1

Figure 1
<p>Global distribution of 36 flux stations (5 PFTs are included: CRO, DBF, ENF, GRA, and SAV).</p>
Full article ">Figure 2
<p>Ranges of four variables (<b>a</b>) VPD, (<b>b</b>) SWC, (<b>c</b>) SWin, and (<b>d</b>) Ta for five PFT (CRO, DBF, ENF, GRA, and SAV) sites. The upper, middle, and lower horizontal lines represent the 75th percentile, median, and 25th percentile, respectively. The red dot represents the mean value, and the blue ‘+’ sign represents an abnormal value.</p>
Full article ">Figure 3
<p>The influence of Gs, GPP, and NPP on five PFTs ((<b>a</b>–<b>c</b>) CRO, (<b>d</b>–<b>f</b>) DBF, (<b>g</b>–<b>i</b>) ENF, (<b>j</b>–<b>l</b>) GRA, (<b>m</b>–<b>o</b>) SAV) when VPD and SWC are decoupled. The six colors represent different percentile ranges of soil water content (0–15%, 15–30%, 30–50%, 50–70%, 70–90%, and 90–100%). The interval of each point on VPD is 0.2 kpa.</p>
Full article ">Figure 4
<p>Changes in Gs (<b>a</b>–<b>e</b>), GPP (<b>f</b>–<b>j</b>), and NPP (<b>k</b>–<b>o</b>) affected by VPD and SWC for five PFTs (CRO, DBF, ENF, GRA, SAV). The color in the figure indicates the abnormality of Gs, GPP, CUE, and WUE for each VPD and SWC percentile. For different vegetation types, the difference between these variables and the average value of the multi-year growing season is calculated as the abnormal value on the hourly scale.</p>
Full article ">Figure 5
<p>The WUE and CUE of five PFTs (CRO (<b>a</b>,<b>b</b>), DBF (<b>c</b>,<b>d</b>), ENF (<b>e</b>,<b>f</b>), GRA (<b>g</b>,<b>h</b>), SAV (<b>i</b>,<b>j</b>)) when VPD and SWC are decoupled. The six colors represent different percentile ranges of soil water content (0–15%, 15–30%, 30–50%, 50–70%, 70–90%, and 90–100%). The interval of each point on VPD is 0.2 kpa.</p>
Full article ">Figure 6
<p>Changes in WUE (<b>a</b>–<b>e</b>) and CUE (<b>f</b>–<b>j</b>) affected by VPD and SWC for five PFTs (CRO, DBF, ENF, GRA, SAV).</p>
Full article ">Figure 7
<p>Changes in (<b>a</b>) Gs, (<b>b</b>) GPP, (<b>c</b>) NPP, (<b>d</b>) WUE and (<b>e</b>) CUE for PFTs (CRO, DBF, ENF, GRA, SAV) under the guidance of increasing VPD (VPD|SWC) and decreasing SWC (SWC|VPD), respectively. The horizontal line represents the median, the point represents the average, and the ‘+’ sign represents the abnormal value.</p>
Full article ">
12 pages, 1715 KiB  
Communication
Photosynthetically Active Radiation and Foliage Clumping Improve Satellite-Based NIRv Estimates of Gross Primary Production
by Iolanda Filella, Adrià Descals, Manuela Balzarolo, Gaofei Yin, Aleixandre Verger, Hongliang Fang and Josep Peñuelas
Remote Sens. 2023, 15(8), 2207; https://doi.org/10.3390/rs15082207 - 21 Apr 2023
Cited by 2 | Viewed by 2549
Abstract
Monitoring gross primary production (GPP) is necessary for quantifying the terrestrial carbon balance. The near-infrared reflectance of vegetation (NIRv) has been proven to be a good predictor of GPP. Given that radiation powers photosynthesis, we hypothesized that (i) the addition of photosynthetic photon [...] Read more.
Monitoring gross primary production (GPP) is necessary for quantifying the terrestrial carbon balance. The near-infrared reflectance of vegetation (NIRv) has been proven to be a good predictor of GPP. Given that radiation powers photosynthesis, we hypothesized that (i) the addition of photosynthetic photon flux density (PPFD) information to NIRv would improve estimates of GPP and that (ii) a further improvement would be obtained by incorporating the estimates of radiation distribution in the canopy provided by the foliar clumping index (CI). Thus, we used GPP data from FLUXNET sites to test these possible improvements by comparing the performance of a model based solely on NIRv with two other models, one combining NIRv and PPFD and the other combining NIRv, PPFD and the CI of each vegetation cover type. We tested the performance of these models for different types of vegetation cover, at various latitudes and over the different seasons. Our results demonstrate that the addition of daily radiation information and the clumping index for each vegetation cover type to the NIRv improves its ability to estimate GPP. The improvement was related to foliage organization, given that the foliar distribution in the canopy (CI) affects radiation distribution and use and that radiation drives productivity. Evergreen needleleaf forests are the vegetation cover type with the greatest improvement in GPP estimation after the addition of CI information, likely as a result of their greater radiation constraints. Vegetation type was more determinant of the sensitivity to PPFD changes than latitude or seasonality. We advocate for the incorporation of PPFD and CI into NIRv algorithms and GPP models to improve GPP estimates. Full article
(This article belongs to the Special Issue Remote Sensing Applications for the Biosphere)
Show Figures

Figure 1

Figure 1
<p>Sites used in this study.</p>
Full article ">Figure 2
<p>Relationship between the measured GPP and the GPP estimated from (i) NIRv, (ii) NIRv + PPFD and (iii) NIRv + PPFD + CI, for the whole dataset. CI used in the model was the CI per vegetation cover type from He et al. [<a href="#B24-remotesensing-15-02207" class="html-bibr">24</a>]. Data correspond to daily mean data from 26 sites from 2000 to 2014. The red line corresponds to the linear regression and the black line to the 1:1 line. Different colours identify the vegetation cover types.</p>
Full article ">Figure 3
<p>Relationship between the clumping index and the variation in R<sup>2</sup> (∆R<sup>2</sup>) and RMSE (∆RMSE) between the ground-measured GPP and the satellite-estimated GPP after adding PPFD information to the NIRv model (GPPnirppfd vs. GPPnirv) in black for the different studied sites (<a href="#remotesensing-15-02207-t001" class="html-table">Table 1</a>) (CI calculated as the average annual value per site from the CI product LIS-CI-A1), and in red for the different vegetation cover types (CI for vegetation cover type as in He et al. [<a href="#B24-remotesensing-15-02207" class="html-bibr">24</a>]): evergreen needleleaf forest (ENF), evergreen broadleaf forest (EBF), deciduous broadleaf forest (DBF), mixed forest (MF), open shrubland (OSH), grassland (GRA), woody savanna (WSA) and wetland (WET).</p>
Full article ">Figure 4
<p>Variation in absolute bias between the ground-measured GPP and the satellite-estimated GPP after adding biome CI information to the NIR and PPFD model (GPPnirppfdci vs. GPPnirppfd) for the different vegetation cover types.</p>
Full article ">
15 pages, 4994 KiB  
Article
Meteorological Influences on Short-Term Carbon-Water Relationships in Two Forests in Subtropical China
by Jihua Pan, Jane Liu, Mengmiao Yang and Rong Wang
Atmosphere 2023, 14(3), 457; https://doi.org/10.3390/atmos14030457 - 24 Feb 2023
Viewed by 1646
Abstract
Carbon and water fluxes in ecosystems are tightly coupled by gas diffusion through stomata. However, carbon–water (C–W) relationships vary largely across time scales, vegetation types, and regions. Subtropical forests in China play an important role in the global carbon and water cycles, yet [...] Read more.
Carbon and water fluxes in ecosystems are tightly coupled by gas diffusion through stomata. However, carbon–water (C–W) relationships vary largely across time scales, vegetation types, and regions. Subtropical forests in China play an important role in the global carbon and water cycles, yet studies of C–W relationships in this region remain limited. Here, we investigated summer-time C–W relationships in this region at two subtropical sites: the evergreen broadleaved forest at Dinghushan (23.17° N, 112.53° E, 300 m) and the evergreen coniferous forest at Qianyanzhou (26.74° N, 115.06° W, 106 m), using the flux tower data from the FLUXNET2015. The C–W relationship was examined using two measures. The first was daily water use efficiency (WUE), which is the ratio of daily gross primary productivity (GPP) to evapotranspiration (ET). The second was the correlation coefficient (r) of hourly GPP and ET. Our analysis showed that the daily WUE in the two forests ranged over 4–14 mg CO2 per g H2O, higher in the coniferous forest than in the broadleaved forest. The mean values of r for hourly C–W coupling were similar at the two forests, being 0.5–0.6, which suggests asynchronous diurnal variations in GPP and ET. Both daily WUE and r were modulated by meteorological conditions. In general, high radiation, air temperature, and humidity can reduce WUE at both sites. For the broadleaved forest, the most influential factor on WUE was VPD, followed by radiation, while in the coniferous forest, VPD, air temperature, and radiation were almost equally important. For hourly C–W coupling, VPD plays a significant role. The drier the air is, the weaker the coupling in the two forests. The daily WUE and hourly C–W coupling reflect the C–W relationship from different perspectives. Both showed the strongest response to VPD but with different sensitivity. Full article
(This article belongs to the Special Issue Forests and Climate Interactions)
Show Figures

Figure 1

Figure 1
<p>The frequency of summer daytime shortwave radiation (SR, (<b>a</b>)), air temperature (Ta, (<b>b</b>)), vapor pressure deficit (VPD, (<b>c</b>)), and soil water content (SWC, (<b>d</b>)) in the Dinghushan evergreen broadleaved forest (red solid line) and Qianyanzhou evergreen coniferous forest (cyan solid line). The dashed lines show the mean value of the meteorological variables at Dinghushan (red dashed line) and Qianyanzhou (cyan dashed line) during summer daytime; the specific values are presented in <a href="#atmosphere-14-00457-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 2
<p>An example of daily variations in daytime water use efficiency (WUE<sub>dt</sub>, (<b>a</b>), blue hollow squares) and diurnal carbon–water coupling (C–W coupling, (<b>b</b>), cyan hollow squares) in July 2003 at Qianyanzhou, along with meteorological conditions. The solid lines in yellow, red, cyan, and blue indicate the daily means of shortwave radiation (SR), air temperature (T<sub>a</sub>), vapor pressure deficit (VPD), and soil water content (SWC), respectively, during the daytime (6 a.m.–6 p.m.).</p>
Full article ">Figure 3
<p>Histogram of WUE<sub>dt</sub> in the broadleaved forest at Dinghushan and in the coniferous forest at Qianyanzhou on summer no-rain days over 2003–2005.</p>
Full article ">Figure 4
<p>Variation in WUE<sub>dt</sub> with meteorological conditions (SR, T<sub>a</sub>, VPD, and SWC) in the broadleaved forest at Dinghushan (<b>a</b>–<b>d</b>) and in the coniferous forest at Qianyanzhou (<b>e</b>–<b>h</b>), where (<b>a</b>,<b>c</b>) are WUE<sub>dt</sub> variations with SR, (<b>b</b>,<b>f</b>) are for T<sub>a</sub>, (<b>c</b>,<b>g</b>) are for VPD, (<b>d</b>,<b>h</b>) are for SWC. The means and standard deviations of WUE<sub>dt</sub> varying with meteorological conditions are marked with dots and error bars, respectively.</p>
Full article ">Figure 5
<p>Regression coefficients of meteorological conditions (SR, T<sub>a</sub>, VPD, and SWC) with WUE<sub>dt</sub> in the broadleaved forest at Dinghushan and in the coniferous forest at Qianyanzhou. “*” indicates that <span class="html-italic">r</span> is at the 90% significance level (<span class="html-italic">p</span> &lt; 0.10), and “***” at the 99% level (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Histogram of diurnal C–W coupling in the broadleaved forest at Dinghushan and in the coniferous forest at Qianyanzhou on summer no-rain days over 2003–2005.</p>
Full article ">Figure 7
<p>Diurnal C–W coupling varying with meteorological conditions (SR, T<sub>a</sub>, VPD, and SWC) in the broadleaved forest at Dinghushan (<b>a</b>–<b>d</b>) and in the coniferous forest at Qianyanzhou (<b>e</b>–<b>h</b>), where (<b>a</b>,<b>c</b>) are WUE<sub>dt</sub> variations with SR, (<b>b</b>,<b>f</b>) are for T<sub>a</sub>, (<b>c</b>,<b>g</b>) are for VPD, (<b>d</b>,<b>h</b>) are for SWC. The means and standard deviations of C–W coupling varying with meteorological conditions are marked with dots and error bars, respectively.</p>
Full article ">Figure 8
<p>Regression coefficients of meteorological conditions (SR, T<sub>a</sub>, VPD, and SWC) with diurnal C–W coupling in the broadleaved forest at Dinghushan and in the coniferous forest at Qianyanzhou. “*” indicates that <span class="html-italic">r</span> is at the 90% significance level (<span class="html-italic">p</span> &lt; 0.10), and “***” at the 99% level (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 9
<p>The daily GPP and ET relationship in the broadleaved forest at Dinghushan (<b>a</b>) and in the coniferous forest at Qianyanzhou (<b>b</b>) with different values of WUE<sub>dt</sub> in colors. The slope of each regression line is listed in different colors. All values are on summer no-rain days over 2003–2005.</p>
Full article ">Figure 10
<p>The hourly GPP and ET relationship in the broadleaved forest at Dinghushan with different C–W couplings (<span class="html-italic">r</span>): (<b>a</b>) 0.0–0.2, (<b>b</b>) 0.2–0.4, (<b>c</b>) 0.4–0.6, (<b>d</b>) 0.6–0.8, and (<b>e</b>) 0.8–1.0. Coefficient of determination R<sup>2</sup> is listed in each panel. All values are on summer no-rain days over 2003–2005.</p>
Full article ">Figure 11
<p>The hourly GPP and ET relationship in the coniferous forest at Qianyanzhou with different C–W couplings (<span class="html-italic">r</span>): (<b>a</b>) 0.0–0.2, (<b>b</b>) 0.2–0.4, (<b>c</b>) 0.4–0.6, (<b>d</b>) 0.6–0.8, and (<b>e</b>) 0.8–1.0. Coefficient of determination R<sup>2</sup> is listed in each panel. All values are on summer no-rain days over 2003–2005.</p>
Full article ">
Back to TopTop