[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

Journals

Article Types

Countries / Regions

Search Results (36)

Search Parameters:
Keywords = red SIF

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
28 pages, 4053 KiB  
Article
Simulating High-Resolution Sun-Induced Chlorophyll Fluorescence Image of Three-Dimensional Canopy Based on Photon Mapping
by Yaotao Luo, Donghui Xie, Jianbo Qi, Guangjian Yan and Xihan Mu
Remote Sens. 2024, 16(20), 3783; https://doi.org/10.3390/rs16203783 - 11 Oct 2024
Viewed by 842
Abstract
The remote sensing of sun-induced chlorophyll fluorescence (SIF) is an emerging technique with immense potential for terrestrial vegetation sciences. However, the interpretation of fluorescence data is often hindered by the complexity of observed land surfaces. Therefore, advanced remote sensing models, particularly physically based [...] Read more.
The remote sensing of sun-induced chlorophyll fluorescence (SIF) is an emerging technique with immense potential for terrestrial vegetation sciences. However, the interpretation of fluorescence data is often hindered by the complexity of observed land surfaces. Therefore, advanced remote sensing models, particularly physically based simulations, are critical to accurately interpret SIF data. In this work, we propose a three-dimensional (3D) radiative transfer model that employs the Monte Carlo ray-tracing technique to simulate the excitation and transport of SIF within plant canopies. This physically based approach can quantify the various radiative processes contributing to the observed SIF signal with high fidelity. The model’s performance is rigorously evaluated by comparing the simulated SIF spectra and angular distributions to field measurements, as well as conducting systematic comparisons with an established radiative transfer model. The results demonstrate the proposed model’s ability to reliably reproduce the key spectral and angular characteristics of SIF, with the coefficient of determination (R2) exceeding 0.98 and root mean square error (RMSE) being less than 0.08 mW m−2 sr−1 nm−1 for both the red and far-red fluorescence peaks. Furthermore, the model’s versatile representation of canopy structures, enabled by the decoupling of radiation and geometry, is applied to study the impact of 3D structure on SIF patterns. This capability makes the proposed model a highly attractive tool for investigating SIF distributions in realistic, heterogeneous canopy environments. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagrams of the two steps of the proposed model. (<b>a</b>) In the first step, the point representation of the radiation is built through photon tracing, where photons with weight <math display="inline"><semantics> <mi>α</mi> </semantics></math> are emitted from light sources and recorded upon interacting with surfaces within the scene. (<b>b</b>) In the second step, the routine solutions for the radiation transfer problem are adopted to solve the light transport equation, given the emitted fluorescence <math display="inline"><semantics> <msub> <mi>F</mi> <mi>e</mi> </msub> </semantics></math>. The total SIF detected by a sensor can be divided into the direct emission from leaves and the portion that is scattered. The flux contributed by <math display="inline"><semantics> <mi>α</mi> </semantics></math> measures the amount of light that hits a surface over a finite area (represented by the area of a circle at <span class="html-italic">q</span> with radius <span class="html-italic">r</span> here) from all directions. The irradiance <math display="inline"><semantics> <msub> <mi>E</mi> <mrow> <mi>b</mi> <mo>/</mo> <mi>f</mi> </mrow> </msub> </semantics></math> integrates the light arriving at a single point <span class="html-italic">q</span> over the whole hemisphere on the backward or the forward side.</p>
Full article ">Figure 2
<p>Generated abstract scenes representing (<b>a</b>) homogeneous canopy (S2) and (<b>b</b>) row canopy (S3), respectively.</p>
Full article ">Figure 3
<p>The components’ spectra of the abstract scenes.</p>
Full article ">Figure 4
<p>Soybean leaf reflectance and transmittance spectra measured in the field.</p>
Full article ">Figure 5
<p>The components’ measured spectra of the corn scene.</p>
Full article ">Figure 6
<p>The simulated corn canopy based on the in situ measured data.</p>
Full article ">Figure 7
<p>The spectra comparisons of the total SIF (F), the emission part (Fem), and the scattering part (Fsc) in the nadir or hotspot direction for two homogeneous canopies simulated by the proposed model (dashed line) and the SCOPE model (solid line). (<b>a</b>) LAI = 0.9, nadir; (<b>b</b>) LAI = 2.1, nadir; (<b>c</b>) LAI = 0.9, hotspot; and (<b>d</b>) LAI = 2.1, hotspot.</p>
Full article ">Figure 8
<p>Polar plots of fluorescence generated by the proposed model (<b>left</b> panels: <b>a</b>,<b>d</b>,<b>g</b>) and the SCOPE model (<b>middle</b> panels: <b>b</b>,<b>e</b>,<b>h</b>), as well as their differences (<b>right</b> panels: <b>c</b>,<b>f</b>,<b>i</b>), including the total SIF at 685 nm (<b>a</b>,<b>b</b>), the total SIF at 740 nm (<b>d</b>,<b>e</b>), and the scattering part at 740 nm (<b>g</b>,<b>h</b>) for the homogeneous LAI = 2.1 scene. The sun zenith angle is at 30°, and the sun azimuth angle is at 140°.</p>
Full article ">Figure 9
<p>The angular distributions of the total SIF at 685 nm (F685) and 740 nm (F740), along with its emission part at 740 nm (Fem740) and scattering part (Fsc740) along the solar principal plane are simulated for two homogeneous canopies. These distributions are represented by a dashed line for the proposed model and a solid line for the SCOPE model. Negative-view zenith angles denote the forward direction, while positive-view zenith angles represent the backward direction. (<b>a</b>) LAI = 0.9; (<b>b</b>) LAI = 2.1.</p>
Full article ">Figure 10
<p>The SIF spectra for the soybean, reconstructed from measured data, along with simulation by the proposed model.</p>
Full article ">Figure 11
<p>The SIF distributions of the row scene (solid line) and the homogeneous scene (dashed line) simulated using the proposed model. The total SIF, the emission part, and the scattering part are denoted by F, Fem, and Fsc, respectively. (<b>a</b>) The spectra in the nadir direction. (<b>b</b>) The angular distributions along the PP at 685 nm and 740 nm. Negative-view zenith angles denote the forward direction, while positive-view zenith angles represent the backward direction.</p>
Full article ">Figure 12
<p>Polar plots of fluorescence simulated by the the proposed model model for the row scene (left panels) and the homogeneous LAI = 2.1 scene (right panels): the total SIF at 685 nm (<b>a</b>,<b>b</b>) and 740 nm (<b>c</b>,<b>d</b>), and the scattering part at 740 nm (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 13
<p>Subplots illustrating the effects of spatial resolution aggregation (shown in the captions of the subplots) overlaid onto the simulated image of the row scene at 740 nm. (<b>a</b>) The original SIF image. (<b>b</b>–<b>e</b>) Different levels of aggregation in simulated images.</p>
Full article ">Figure 14
<p>Close-up of the natural color image of the row scene in the nadir view.</p>
Full article ">Figure 15
<p>Examples of SIF imaging. (<b>a</b>) A natural color image of the realistic corn scene. (<b>b</b>) The SIF image of the realistic corn scene at 685 nm with the layout of the samples (markers). Several samples are taken as examples to show the SIF imaging data. (<b>c</b>) The SIF spectra of the samples.</p>
Full article ">Figure 16
<p>An ideal diffuse reflectance target is attached to the sample. (<b>a</b>) Sample (3). (<b>b</b>) Sample (4).</p>
Full article ">Figure 17
<p>Examples of simulated SIF images at 685 nm at 10:30 under different light conditions of the realistic corn scene. (<b>a</b>) SIF image under a clear sky. (<b>b</b>) SIF image on a slightly hazy day. (<b>c</b>) The SIF spectra of the samples.</p>
Full article ">Figure 18
<p>The diurnal variations of the PAR and the SIF at 685 nm.</p>
Full article ">Figure 19
<p>The SIF versus the PAR in a day.</p>
Full article ">Figure 20
<p>Examples of the diurnal variations of the SIF at 685 nm. (<b>a</b>) The layout of the samples (markers) in the realistic corn scene. The samples are taken as examples to show the diurnal variations of the SIF. (<b>b</b>) The diurnal variations of the SIF at 685 nm.</p>
Full article ">Figure 21
<p>Examples of SIF images at 685 nm (every 1 h from 10:00 to 15:00) from the simulation of the diurnal variation of the SIF of the realistic corn scene.</p>
Full article ">
17 pages, 1208 KiB  
Article
First-Principles Linear Combination of Atomic Orbitals Calculations of K2SiF6 Crystal: Structural, Electronic, Elastic, Vibrational and Dielectric Properties
by Leonid L. Rusevich, Mikhail G. Brik, Denis Gryaznov, Alok M. Srivastava, Ilya Chervyakov, Guntars Zvejnieks, Dmitry Bocharov and Eugene A. Kotomin
Materials 2024, 17(19), 4865; https://doi.org/10.3390/ma17194865 - 2 Oct 2024
Viewed by 687
Abstract
The results of first-principles calculations of the structural, electronic, elastic, vibrational, dielectric and optical properties, as well as the Raman and infrared (IR) spectra, of potassium hexafluorosilicate (K2SiF6; KSF) crystal are discussed. KSF doped with manganese atoms (KSF:Mn4+ [...] Read more.
The results of first-principles calculations of the structural, electronic, elastic, vibrational, dielectric and optical properties, as well as the Raman and infrared (IR) spectra, of potassium hexafluorosilicate (K2SiF6; KSF) crystal are discussed. KSF doped with manganese atoms (KSF:Mn4+) is known for its ability to function as a phosphor in white LED applications due to the efficient red emission from Mn⁴⁺ activator ions. The simulations were performed using the CRYSTAL23 computer code within the linear combination of atomic orbitals (LCAO) approximation of the density functional theory (DFT). For the study of KSF, we have applied and compared several DFT functionals (with emphasis on hybrid functionals) in combination with Gaussian-type basis sets. In order to determine the optimal combination for computation, two types of basis sets and four different functionals (three advanced hybrid—B3LYP, B1WC, and PBE0—and one LDA functional) were used, and the obtained results were compared with available experimental data. For the selected basis set and functional, the above-mentioned properties of KSF were calculated. In particular, the B1WC functional provides us with a band gap of 9.73 eV. The dependencies of structural, electronic and elastic parameters, as well as the Debye temperature, on external pressure (0–20 GPa) were also evaluated and compared with previous calculations. A comprehensive analysis of vibrational properties was performed for the first time, and the influence of isotopic substitution on the vibrational frequencies was analyzed. IR and Raman spectra were simulated, and the calculated Raman spectrum is in excellent agreement with the experimental one. Full article
(This article belongs to the Section Materials Simulation and Design)
Show Figures

Figure 1

Figure 1
<p>Calculated crystal structure and crystallographic unit cell (36 atoms) of KSF. K atoms—violet balls, Si—blue, F—grey. SiF<sub>6</sub> octahedra are highlighted. The cube drawn with black lines represents a unit cell.</p>
Full article ">Figure 2
<p>Projected and total electronic DOSs of KSF calculated with the hybrid B1WC functional and “TZVP_2012 basis set” at zero external pressure. Contributions of two K, one Si and six F atoms are shown. The zero-energy value corresponds to the Fermi level. (<b>a</b>) Top of valence band (negative values of abscissa axis); (<b>b</b>) bottom of conducting band (band gap is 9.73 eV). The number of Legendre polynomials used for the DOS expansion into series is 12 [<a href="#B21-materials-17-04865" class="html-bibr">21</a>].</p>
Full article ">Figure 3
<p>Dependences of the lattice constant (<b>a</b>) and the interatomic distances Si–F (right vertical axis) and K–F (left vertical axis) (<b>b</b>) on the external pressure with corresponding fitting (dotted lines). Calculations were performed with “TZVP_2012 basis set” and the B1WC functional.</p>
Full article ">Figure 4
<p>Dependence of the KSF band gap value on the external pressure with corresponding fitting (orange dotted line). “TZVP_2012 basis set” and the B1WC functional.</p>
Full article ">Figure 5
<p>One−phonon IR absorbance spectrum of KSF simulated with the B1WC functional and “TZVP_2012 basis set”.</p>
Full article ">Figure 6
<p>Raman spectrum of KSF simulated with the B1WC functional and “TZVP_2012 basis set”.</p>
Full article ">Figure 7
<p>Real Re[<span class="html-italic">ε</span>(<span class="html-italic">ν</span>)] and imaginary Im[<span class="html-italic">ε</span>(<span class="html-italic">ν</span>)] parts of KSF relative permittivity. “TZVP_2012 basis set” and the B1WC functional.</p>
Full article ">Figure 8
<p>Real Re[<span class="html-italic">n</span>(<span class="html-italic">ν</span>)] and imaginary Im[<span class="html-italic">n</span>(<span class="html-italic">ν</span>)] parts of KSF refractive index. “TZVP_2012 basis set” and the B1WC functional.</p>
Full article ">
19 pages, 9732 KiB  
Article
Improved Methods for Retrieval of Chlorophyll Fluorescence from Satellite Observation in the Far-Red Band Using Singular Value Decomposition Algorithm
by Kewei Zhu, Mingmin Zou, Shuli Sheng, Xuwen Wang, Tianqi Liu, Yongping Cheng and Hui Wang
Remote Sens. 2024, 16(18), 3441; https://doi.org/10.3390/rs16183441 - 17 Sep 2024
Viewed by 932
Abstract
Solar-induced chlorophyll fluorescence (SIF) is highly correlated with photosynthesis and can be used for estimating gross primary productivity (GPP) and monitoring vegetation stress. The far-red band of the solar Fraunhofer lines (FLs) is close to the strongest SIF emission peak and is unaffected [...] Read more.
Solar-induced chlorophyll fluorescence (SIF) is highly correlated with photosynthesis and can be used for estimating gross primary productivity (GPP) and monitoring vegetation stress. The far-red band of the solar Fraunhofer lines (FLs) is close to the strongest SIF emission peak and is unaffected by chlorophyll absorption, making it suitable for SIF intensity retrieval. In this study, we propose a retrieval window for far-red SIF, significantly enhancing the sensitivity of data-driven methods to SIF signals near 757 nm. This window introduces a weak O2 absorption band based on the FLs window, allowing for better separation of SIF signals from satellite spectra by altering the shape of specific singular vectors. Additionally, a frequency shift correction algorithm based on standard non-shifted reference spectra is proposed to discuss and eliminate the influence of the Doppler effect. SIF intensity retrieval was achieved using data from the GOSAT satellite, and the retrieved SIF was validated using GPP, enhanced vegetation index (EVI) from the MODIS platform, and published GOSAT SIF products. The validation results indicate that the SIF products provided in this study exhibit higher fitting goodness with GPP and EVI at high spatiotemporal resolutions, with improvements ranging from 55% to 129%. At low spatiotemporal resolutions, the SIF product provided in this study shows higher consistency with EVI and GPP spatially. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Singular vectors in the forward model and the state vector of Fs within two retrieval windows: (<b>a</b>) FLs band, (<b>b</b>) joint retrieval for FLs-O<sub>2</sub> absorption bands.</p>
Full article ">Figure 2
<p>The identical set of spectra before and after wavenumber correction. For the sake of clarity, only a limited portion of the FTS-Band1 spectrum is displayed.</p>
Full article ">Figure 3
<p>Scatter plot and goodness-of-fit (R<sup>2</sup>), Pearson correlation coefficient (P) of monthly SIF products with GPP and VI for January 2019 at 0.1° spatial resolution.</p>
Full article ">Figure 4
<p>Goodness-of-fit (GOF) of the two SIF products with GPP and EVI at 0.1° spatial resolution and monthly scale from January 2018 to June 2020. The shaded part corresponds to December 2018 when SIF retrieves too few results.</p>
Full article ">Figure 5
<p>Pearson correlation coefficients (<span class="html-italic">p</span>-values) of the two SIF products with GPP and EVI at 0.1° spatial resolution and monthly scale from January 2018 to June 2020. The shaded part corresponds to December 2018 when SIF retrieves too few results.</p>
Full article ">Figure 6
<p>Scatter plot and goodness-of-fit (R), Pearson correlation coefficient (P) of 2019 annual SIF products with GPP and VI at 0.1° spatial resolution.</p>
Full article ">Figure 7
<p>Scatter plot and goodness-of-fit (R), Pearson correlation coefficient (P) of 2019 annual SIF products with GPP and VI at 2° spatial resolution.</p>
Full article ">Figure 8
<p>Intensity distribution of the two 2019 annual mean SIF products and 2019 annual mean EVI, GPP in 0.1° grid units. To facilitate observation, morphological dilation was applied to the SIF intensity distribution images.</p>
Full article ">Figure 9
<p>Intensity distribution of the two 2019 annual solar-normalized SIF products and annual mean EVI, GPP in 2° grid units.</p>
Full article ">Figure 10
<p>Maps of the mean 2019 annual intensity distribution for the two SIF products at 2° grid cells and of the mean annual results for EVI at 0.1° grid cells. The color–intensity relationship is the same for each row of subplots. The maps contain four regions: Northern South America, the United States and southern Canada, Western Europe, and southern Africa.</p>
Full article ">Figure 11
<p>Slope and GOF of the linear fit between the SIF retrieval results from the combined retrieval window (O<sub>2</sub> absorption and FLs bands) and FLs band alone with GPP/EVI.</p>
Full article ">Figure 12
<p>The first six singular vectors and the seventh to eighth singular vectors obtained from the spectra of the training set before and after frequency shift correction.</p>
Full article ">Figure 13
<p>Scatter plots and linear fitting results of retrieval outcomes with GPP and EVI before and after satellite spectral frequency shift correction in January 2019.</p>
Full article ">Figure 14
<p>The intensity distribution of the existing daily average SIF (<b>a</b>) and the proposed daily average SIF (<b>b</b>) is presented for the entire year of 2019 under 10.5 Km × 10.5 Km spatial resolution. Additionally, the intensity distribution of monthly MODIS Enhanced Vegetation Index (EVI) (<b>c</b>) under 0.5° grid cells and annual GPP (<b>d</b>) under 500 m SIN grid is displayed for the entire year of 2019.</p>
Full article ">
20 pages, 2254 KiB  
Article
Monitoring of Wheat Stripe Rust Using Red SIF Modified by Pseudokurtosis
by Xia Jing, Qixing Ye, Bing Chen, Bingyu Li, Kaiqi Du and Yiyang Xue
Agronomy 2024, 14(8), 1698; https://doi.org/10.3390/agronomy14081698 - 1 Aug 2024
Viewed by 686
Abstract
Red solar-induced chlorophyll fluorescence (SIFB) is closely related to the photosynthetically active radiation absorbed by chlorophyll. The scattering and reabsorption of SIFB by the vegetation canopy significantly change the spectral intensity and shape of SIF, which affects the relationship between [...] Read more.
Red solar-induced chlorophyll fluorescence (SIFB) is closely related to the photosynthetically active radiation absorbed by chlorophyll. The scattering and reabsorption of SIFB by the vegetation canopy significantly change the spectral intensity and shape of SIF, which affects the relationship between SIF and crop stress. To address this, we propose a method of modifying SIFB using SIF spectral shape characteristic parameters to reduce this influence. A red pseudokurtosis (PKB) parameter that can characterize spectral shape features was calculated using full-spectrum SIF data. On this basis, we analyzed the photosynthetic physiological mechanism of PKB and found that it significantly correlates with both the fraction of photosynthetically active radiation absorbed by chlorophyll(fPARchl) and the red SIF escape rate (fesc680); thus, it is closely related to the scattering and reabsorption of SIFB by the vegetation canopy. Consequently, we constructed an expression of PKB to modify SIFB. To evaluate the modified SIFB (MSIFB) in monitoring the severity of wheat stripe rust, we analyzed the correlations between SIFB, MSIFB, SIFB-VIs (a fusion of the vegetation index and SIFB), and MSIFB-VIs (a fusion of the vegetation index and MSIFB) with the severity level (SL), respectively. The results show that the correlation between MSIFB and the severity of wheat stripe rust increased by an average of 25.6% and at least 16.95% compared with that for SIFB. In addition, we constructed remote sensing monitoring models for wheat stripe rust using linear regression methods, with SIFB, MSIFB, SIFB-VIs, and MSIFB-VIs as independent variables. PKB significantly improves the accuracy and robustness of models based on SIFB and its fusion index SIFB-VIs in the constructed testing set. The R-value between the predicted SL and the measured SL of the remote sensing monitoring model for wheat stripe rust was established using MSIFB-VIs as the independent variable, and it was improved by an average of 39.49% compared with the model using SIFB-VIs. The RMSE was reduced by an average of 18.22%. Therefore, the SIFB modified by PKB can weaken the effects of chlorophyll reabsorption and canopy architecture on SIFB and improve the ability of SIFB to detect stress information. Full article
(This article belongs to the Section Pest and Disease Management)
Show Figures

Figure 1

Figure 1
<p>Overview of the natural disease field area; the green dots depict sampling sites.</p>
Full article ">Figure 2
<p>Emission peak range: the shaded area is the SIF<sub>B</sub> emission peak.</p>
Full article ">Figure 3
<p>Relative absorption rates at different chlorophyll contents.</p>
Full article ">Figure 4
<p>The correlation coefficient between PK<sub>B</sub> and photosynthetic parameters.</p>
Full article ">Figure 5
<p>The relationship between SIF<sub>B</sub> and SL before and after PK<sub>B</sub> modification is investigated in the following experiments: (<b>a</b>) Plot disease field experiment (n = 45); (<b>b</b>) Natural disease field experiment (n = 253). Additionally, the relationship between SIF<sub>A</sub> and SL is also examined in: (<b>c</b>) Plot disease field experiment (n = 45); (<b>d</b>) Natural disease field experiment (n = 253). Note: All scatter plots exhibit extremely significant correlation (<span class="html-italic">p</span> &lt; 0.001), with the different color lines representing the regression lines.</p>
Full article ">Figure 6
<p>The relationship between SIF<sub>B</sub>-N and SL, both before and after PK<sub>B</sub> modification, is investigated in the following experiments: (<b>a</b>) Plot disease field experiment (n = 45); (<b>b</b>) Natural disease field experiment (n = 253). Additionally, the relationship between SIF<sub>A</sub>-N and SL is also examined in separate experiments: (<b>c</b>) Plot disease field experiment (n = 45); (<b>d</b>) Natural disease field experiment (n = 253). The different color lines in the scatter plots denote the regression lines.</p>
Full article ">Figure 7
<p>The relationship between SIF<sub>B</sub>-M and SL, both before and after PK<sub>B</sub> modification, is investigated in the following experiments: (<b>a</b>) Plot disease field experiment (n = 45); (<b>b</b>) Natural disease field experiment (n = 253). Additionally, the relationship between SIF<sub>A</sub>-M and SL is also examined in: (<b>c</b>) Plot disease field experiment (n = 45); (<b>d</b>) Natural disease field experiment (n = 253). The different color lines in the scatter plots represent the regression lines for the respective relationships.</p>
Full article ">Figure 8
<p>The relationship between SIF<sub>B</sub>-NM and SL, before and after PK<sub>B</sub> modification, is explored in the following experiments: (<b>a</b>) Plot disease field experiment with a sample size of 45; (<b>b</b>) Natural disease field experiment with a sample size of 253. Additionally, the relationship between SIF<sub>A</sub>-NM and SL is also examined in: (<b>c</b>) Another plot disease field experiment with a sample size of 45; (<b>d</b>) A separate natural disease field experiment with a sample size of 253. The different color lines in the scatter plots indicate the regression lines for the corresponding relationships.</p>
Full article ">Figure 9
<p>Model accuracy verification. The different color lines denote the regression.</p>
Full article ">Figure 10
<p>Mean and standard deviation plots showing the growth dynamic of SL (<b>a</b>), SIF<sub>B</sub> (<b>b</b>), MSIF<sub>B</sub> (<b>c</b>), and C<sub>ab</sub> (<b>d</b>) for winter wheat.</p>
Full article ">Figure 11
<p>Accuracy of the F-SFM algorithm: The red line denotes the regression line, with the accompanying shaded area representing the 95% confidence interval around the regression estimate.</p>
Full article ">Figure 12
<p>VI compensation for MSIF<sub>B</sub>.</p>
Full article ">Figure 13
<p>A box plot displaying the median, 0th percentile (minimum), 25th percentile (first quartile), 75th percentile (third quartile), and 100th percentile (maximum) of the LAI across different growth periods. The outliers in the data are represented by individual points outside the whiskers.</p>
Full article ">
13 pages, 4101 KiB  
Article
Phosphor Ceramic Composite for Tunable Warm White Light
by Ross A. Osborne, Nerine J. Cherepy, Peter S. Bleier, Romain M. Gaume and Stephen A. Payne
Materials 2024, 17(13), 3187; https://doi.org/10.3390/ma17133187 - 29 Jun 2024
Viewed by 816
Abstract
Composite phosphor ceramics for warm white LED lighting were fabricated with K2SiF6:Mn4+ (KSF) as both a narrowband red phosphor and a translucent matrix in which yellow-emitting Y3Al5O12:Ce3+ (YAG) particles were dispersed. [...] Read more.
Composite phosphor ceramics for warm white LED lighting were fabricated with K2SiF6:Mn4+ (KSF) as both a narrowband red phosphor and a translucent matrix in which yellow-emitting Y3Al5O12:Ce3+ (YAG) particles were dispersed. The emission spectra of these composites under blue LED excitation were studied as a function of YAG loading and thickness. Warm white light with a color temperature of 2716 K, a high CRI of 92.6, and an R9 of 77.6 was achieved. A modest improvement in the thermal conductivity of the KSF ceramic of up to 9% was observed with the addition of YAG particles. In addition, a simple model was developed for predicting the emission spectra based on several parameters of the composite ceramics and validated with the experimental results. The emission spectrum can be tuned by varying the dopant concentrations, thickness, YAG loading, and YAG particle size. This work demonstrates the utility of KSF/YAG composite phosphor ceramics as a means of producing warm white light, which are potentially suitable for higher-drive applications due to their increased thermal conductivity and reduced droop compared with silicone-dispersed phosphor powders. Full article
Show Figures

Figure 1

Figure 1
<p>Synthesis process for composite phosphor ceramics.</p>
Full article ">Figure 2
<p>(<b>a</b>) KSF powder and (<b>b</b>) YAG powder used to make the phosphor ceramics.</p>
Full article ">Figure 3
<p>XRD patterns of the raw powders and ceramic samples with 2.5, 5, and 7.5 wt% YAG loading fractions, along with the powder diffraction reference patterns for the KSF (PDF file 04-006-8962) and YAG (PDF file 04-007-2667) crystal phases.</p>
Full article ">Figure 4
<p>Energy-dispersive X-ray spectroscopy maps of the polished surfaces of samples with 2.5, 4, 5, and 7.5 wt% YAG loading fractions (pictures of samples at left).</p>
Full article ">Figure 5
<p>Backscatter electron micrographs of a fracture surface of (<b>a</b>) KSF-only ceramic and (<b>b</b>) composite ceramic, and (<b>c</b>) close-up of YAG particles in a composite ceramic.</p>
Full article ">Figure 6
<p>(<b>Top</b>) Pictures of the samples under room lights and 450 nm excitation. (<b>Bottom</b>) A diffuse light rod sitting on top of the samples irradiated from beneath with a blue LED.</p>
Full article ">Figure 7
<p>Experimental and modeled emission spectra of the samples excited by a blue LED with various YAG loading fractions at different thicknesses. The spectral contributions attributed to the LED, YAG:Ce, and KSF:Mn are labeled in the top left panel.</p>
Full article ">Figure 8
<p>CIE 1931 color coordinate plot of the samples along with the blackbody radiation curve as a function of their (<b>a</b>) YAG loading (thickness held constant at 0.7 mm) and (<b>b</b>) thickness (YAG loading held constant at 2.5 wt%).</p>
Full article ">Figure 9
<p>Calculated correlated color temperature (CCT), color rendering index (CRI), R9, and efficacy of the samples.</p>
Full article ">Figure 10
<p>Thermal conductivity as a function of temperature for various loading fractions of YAG in a Mn:KSF matrix.</p>
Full article ">
13 pages, 4662 KiB  
Article
In Situ Synthesis of CsPbX3/Polyacrylonitrile Nanofibers with Water-Stability and Color-Tunability for Anti-Counterfeiting and LEDs
by Yinbiao Shi, Xiaojia Su, Xiaoyan Wang and Mingye Ding
Polymers 2024, 16(11), 1568; https://doi.org/10.3390/polym16111568 - 1 Jun 2024
Viewed by 948
Abstract
Inorganic CsPbX3 (X = Cl, Br, I) perovskite quantum dots (PQDs) have attracted widespread attention due to their excellent optical properties and extensive application prospects. However, their inherent structural instability significantly hinders their practical application despite their outstanding optical performance. To enhance [...] Read more.
Inorganic CsPbX3 (X = Cl, Br, I) perovskite quantum dots (PQDs) have attracted widespread attention due to their excellent optical properties and extensive application prospects. However, their inherent structural instability significantly hinders their practical application despite their outstanding optical performance. To enhance stability, an in situ electrospinning strategy was used to synthesize CsPbX3/polyacrylonitrile composite nanofibers. By optimizing process parameters (e.g., halide ratio, electrospinning voltage, and heat treatment temperature), all-inorganic CsPbX3 PQDs have been successfully grown in a polyacrylonitrile (PAN) matrix. During the electrospinning process, the rapid solidification of electrospun fibers not only effectively constrained the formation of large-sized PQDs but also provided effective physical protection for PQDs, resulting in the improvement in the water stability of PQDs by minimizing external environmental interference. Even after storage in water for over 100 days, the PQDs maintained approximately 93.5% of their photoluminescence intensity. Through the adjustment of halogen elements, the as-obtained composite nanofibers exhibited color-tunable luminescence in the visible light region, and based on this, a series of multicolor anti-counterfeiting patterns were fabricated. Additionally, benefiting from the excellent water stability and optical performance, the CsPbBr3/PAN composite film was combined with red-emitting K2SiF6:Mn4+ (KSF) on a blue LED (460 nm), producing a stable and efficient WLED device with a color temperature of around 6000 K and CIE coordinates of (0.318, 0.322). These results provide a general approach to synthesizing PQDs/polymer nanocomposites with excellent water stability and multicolor emission, thereby promoting their practical applications in multifunctional optoelectronic devices and advanced anti-counterfeiting. Full article
(This article belongs to the Special Issue New Advances in Polymer Electrospun Fibers)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>,<b>b</b>) Schematic of the one-step single-axis electrospinning setup to fabricate the perovskite light-emitting nanofibers. Photographs of as-synthesized CsPbX<sub>3</sub>/PAN samples: (<b>c</b>) CsPbCl<sub>3</sub>/PAN, (<b>d</b>) CsPbBr<sub>1.5</sub>Cl<sub>1.5</sub>/PAN, (<b>e</b>) CsPbBr<sub>3</sub>/PAN, (<b>f</b>) CsPbBr<sub>1.5</sub>I<sub>1.5</sub>/PAN, (<b>g</b>) CsPbI<sub>3</sub>/PAN.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) SEM images of pure PAN nanofibers. (<b>c</b>,<b>d</b>) SEM images of CsPbBr<sub>3</sub>/PAN (CsBr:PbBr<sub>2</sub> is 1:1) composite nanofibers. (<b>e</b>) TEM images of CsPbBr<sub>3</sub>/PAN composite nanofibers. EDS mapping for C, Pb and Br elements. (<b>f</b>,<b>g</b>) Fluorescent field images and bright-field images of CsPbBr<sub>3</sub>/PAN composite nanofibers recorded by LSCM, (<b>h</b>) XRD patterns of CsPbBr<sub>3</sub>/PAN composite nanofibers.</p>
Full article ">Figure 3
<p>CsPbBr<sub>3</sub>/PAN composite nanofibers: (<b>a</b>) UV-vis absorption and PL spectra. (<b>b</b>) Different ratios of CsBr and PbBr<sub>2</sub>. (<b>c</b>) Different voltage. (<b>d</b>) Thermostability under 423 K. (<b>e</b>) Temporal evolution of PL intensity in room-temperature water. (<b>f</b>) Long-term storage stability. (<b>g</b>) Photos of CsPbBr<sub>3</sub>/PAN composite nanofibers dipped in water for 1, 3, 5, 7, 14, 21, 97 and 104 days.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) SEM images of CsPbX<sub>3</sub>/PAN composite nanofibers: (<b>a</b>) CsPbCl<sub>3</sub>. (<b>b</b>) CsPbBr<sub>1.5</sub>Cl<sub>1.5</sub>. (<b>c</b>) CsPbBr<sub>1.5</sub>I<sub>1.5</sub>. (<b>d</b>) CsPbI<sub>3</sub>. (<b>e</b>) XRD patterns of CsPbX<sub>3</sub>/PAN composite nanofibers.</p>
Full article ">Figure 5
<p>UV-vis absorption and PL spectra of CsPbX<sub>3</sub>/PAN composite nanofibers (<b>a</b>) CsPbCl<sub>3</sub>/PAN. (<b>b</b>) CsPbBr<sub>1.5</sub>Cl<sub>1.5</sub>/PAN. (<b>c</b>) CsPbBr<sub>1.5</sub>I<sub>1.5</sub>/PAN. (<b>d</b>) CsPbI<sub>3</sub>/PAN.</p>
Full article ">Figure 6
<p>CsPbX<sub>3</sub>/PAN composite nanofibers with different halogen constitutions: (<b>a</b>) PL spectra. (<b>b</b>) Digital photographs under UV excitation (λ = 365 nm). (<b>c</b>) PL decay curves. (<b>d</b>) CIE coordinate and color gamut.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) Demonstration of the application of a multicolor fluorescent patterns for anti-counterfeiting. (<b>d</b>,<b>e</b>) Flexibility on display.</p>
Full article ">Figure 8
<p>(<b>a</b>) Simplified structure of WLED. (<b>b</b>) EL spectrum (inset shows a digital camera image of the working WLED). (<b>c</b>) CIE coordinate and color gamut of WLED.</p>
Full article ">
20 pages, 9140 KiB  
Article
The Afternoon/Morning Ratio of Tower-Based Solar-Induced Chlorophyll Fluorescence Can Be Used to Monitor Drought in a Chinese Cork Oak Plantation
by Qingmei Pan, Xiangfen Cheng, Meijun Hu, Linqi Liu, Xin Wang, Jinsong Zhang, Zhipeng Li, Wenwen Yuan and Xiang Gao
Remote Sens. 2024, 16(11), 1897; https://doi.org/10.3390/rs16111897 - 24 May 2024
Viewed by 865
Abstract
Monitoring drought stress is crucial for estimating productivity and assessing the health status of forest ecosystems under global climate change. Solar-induced chlorophyll fluorescence (SIF) is mechanistically coupled to photosynthesis and has advantages over greenness-based vegetation indices in detecting drought. In recent years, SIF [...] Read more.
Monitoring drought stress is crucial for estimating productivity and assessing the health status of forest ecosystems under global climate change. Solar-induced chlorophyll fluorescence (SIF) is mechanistically coupled to photosynthesis and has advantages over greenness-based vegetation indices in detecting drought. In recent years, SIF has commonly been used in monitoring drought stress in crop ecosystems. However, the response of tower-based SIF to drought stress in forest ecosystems remains unclear. In this study, we investigated the potential of tower-based SIF to monitor drought, which was quantified using the plant water stress index (PWSI) in a Chinese cork oak plantation. The results show the negative effect of drought on SIF, and afternoon depression of SIF emission under drought stress was observed. Canopy SIF (F) exhibited a nonlinear relationship with PWSI, while the quantum yield of SIF (ΦF) exhibited a significant linear relationship with PWSI at 687 nm and 760 nm (ΦF687: R2 = 0.90; ΦF760: R2 = 0.85). Incident radiation (PAR) and canopy structure affected the response of SIF to drought stress, with PAR as the main factor causing the nonlinear relationship between F and PWSI. Afternoon depression was described as the afternoon/morning ratio (AMR). AMRF and AMRΦF exhibited a negative linear response to PWSI. AMRF was less affected than F by PAR and canopy structures, and AMRΦF was more physiologically representative than ΦF. Moreover, AMRΦF was sensitive to VPD and REW, and it might be a good indicator of drought. Red SIF was more sensitive to drought than far-red SIF, as the R2 of PWSI with AMRΦF687 (R2 = 0.89) was higher than that with AMRΦF687 (R2 = 0.84). These results highlight the potential of tower-based SIF, especially red SIF, for drought monitoring in a plantation, and consideration of the physiological diurnal variation in SIF under drought stress is crucial for improving the accuracy of drought stress monitoring. Full article
Show Figures

Figure 1

Figure 1
<p>Seasonal variations in 2020 and 2021 for (<b>a</b>) plant water stress index (PWSI), (<b>b</b>) normalized difference vegetation index (NDVI), (<b>c</b>) canopy solar-induced chlorophyll fluorescence in the red band (F<sub>687</sub>, mWm<sup>−2</sup>nm<sup>−1</sup>sr<sup>−1</sup>), (<b>d</b>) canopy solar-induced chlorophyll fluorescence in the far-red band (F<sub>760</sub>, mWm<sup>−2</sup>nm<sup>−1</sup>sr<sup>−1</sup>), (<b>e</b>) photosynthetically active radiation (PAR, umolm<sup>−2</sup>s<sup>−1</sup>), (<b>f</b>) red reflectance of the vegetation (Redv), (<b>g</b>) near-infrared reflectance of the vegetation (NIRv), (<b>h</b>) SIF quantum yield in the red band (ΦF<sub>687</sub>), and (<b>i</b>) SIF quantum yield in the far-red band (ΦF<sub>760</sub>). Gray dots indicate half-hourly observed value, red rings indicate daily averages of 8:00–17:00, and red curves indicate 8 day moving averages.</p>
Full article ">Figure 2
<p>Pearson correlation coefficient heatmap. Correlation coefficients of SIF(F) and its quantum yield (ΦF) with PWSI at (<b>a</b>) daily timescales and (<b>b</b>) half-hourly timescales. Correlation coefficients between canopy structural parameters and PWSI at (<b>c</b>) daily timescales and (<b>d</b>) half-hour timescales. *** represents significance levels below 0.001 (<span class="html-italic">p</span> &lt; 0.001), ** represents significance levels below 0.01 (<span class="html-italic">p</span> &lt; 0.01), and * represents significance levels below 0.05 (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Daily mean PWSI, REDv, and NIRv in (<b>a</b>–<b>c</b>), as well as diurnal patterns for PAR, F<sub>687</sub>, F<sub>760</sub>, ΦF<sub>687</sub>, and ΦF<sub>760</sub> for the four observation days (DOY136, DOY138, DOY146, and DOY148 in 2020) in (<b>d</b>–<b>h</b>). Means followed by the same letter were not significantly different at <span class="html-italic">p</span> ≤ 0.05 according to Tukey’s HSD test in (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 4
<p>Percentage of decline for DOY138, DOY146, and DOY148 relative to DOY136 in the morning, noon, and afternoon for (<b>a</b>) F<sub>687</sub>, (<b>b</b>) F<sub>760</sub>, (<b>c</b>) ΦF<sub>687</sub>, and (<b>d</b>) ΦF<sub>760</sub>.</p>
Full article ">Figure 5
<p>Relationships between PWSI and SIF, and afternoon depression (AMR) of SIF. The PWSI of the X-axis is depicted as PWSIbin. The top row shows the daily average, and the bottom row shows AMR. The error bars indicate the standard deviation (SD). Red straight lines indicate linear and nonlinear regression. It should be noted that the black straight dashed line indicates linear regression at PWSI &lt; 0.42; the green straight dotted and dashed line indicates linear regression at PWSI &gt; 0.42 in subfigures (<b>a</b>,<b>b</b>).</p>
Full article ">Figure 6
<p>Correlation coefficients of F and AMR<sub>F</sub> with (<b>a</b>) radiation (PAR) and (<b>b</b>) structural (REDv or NIRv) factors. *** represents significance levels below 0.001 (<span class="html-italic">p</span> &lt; 0.001), and * represents significance levels below 0.05 (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Relationship between PWSI and SIF excluding radiation for (<b>a</b>,<b>c</b>) and canopy structure for (<b>b</b>,<b>d</b>). The error bars indicate the standard deviation (SD). Straight lines indicate linear regression.</p>
Full article ">Figure 8
<p>Correlation coefficients between PWSI and F, AMR<sub>F</sub>. Pearson correlation coefficients (r) and partial correlation coefficients exclude PAR for PWSI and F in (<b>a</b>), and Pearson correlation coefficients (r) and partial correlation coefficients exclude canopy structural factors (REDv or NIRv) for PWSI and F, AMR<sub>F</sub> in (<b>b</b>). Red-filled bars indicate Pearson correlation coefficients, and grey twill-filled bars indicate partial correlation coefficients. *** represents significance levels below 0.001 (<span class="html-italic">p</span> &lt; 0.001), ** represents significance levels below 0.01 (<span class="html-italic">p</span> &lt; 0.01), and * represents significance levels below 0.05 (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>The response of AMR<sub>ΦF</sub> and daily mean ΦF to VPD for (<b>a</b>,<b>b</b>) and to REW for (<b>c</b>,<b>d</b>). The shadows indicate the standard deviation (SD). Straight lines indicate linear regression. Blue shows AMR and red shows ΦF.</p>
Full article ">
24 pages, 11080 KiB  
Article
Prediction of Open Woodland Transpiration Incorporating Sun-Induced Chlorophyll Fluorescence and Vegetation Structure
by Sicong Gao, William Woodgate, Xuanlong Ma and Tanya M. Doody
Remote Sens. 2024, 16(1), 143; https://doi.org/10.3390/rs16010143 - 28 Dec 2023
Cited by 1 | Viewed by 1332
Abstract
Transpiration (T) represents plant water use, while sun-induced chlorophyll fluorescence (SIF) emitted during photosynthesis, relates well to gross primary production. SIF can be influenced by vegetation structure, while uncertainties remain on how this might impact the relationship between SIF and T, especially for [...] Read more.
Transpiration (T) represents plant water use, while sun-induced chlorophyll fluorescence (SIF) emitted during photosynthesis, relates well to gross primary production. SIF can be influenced by vegetation structure, while uncertainties remain on how this might impact the relationship between SIF and T, especially for open and sparse woodlands. In this study, a method was developed to map T in riverine floodplain open woodland environments using satellite data coupled with a radiative transfer model (RTM). Specifically, we used FluorFLiES, a three-dimensional SIF RTM, to simulate the full spectrum of SIF for three open woodland sites with varying fractional vegetation cover. Five specific SIF bands were selected to quantify their correlation with field measured T derived from sap flow sensors. The coefficient of determination of the simulated far-red SIF and field measured T at a monthly scale was 0.93. However, when comparing red SIF from leaf scale to canopy scale to predict T, performance declined by 24%. In addition, varying soil reflectance and understory leaf area index had little effect on the correlation between SIF and T. The method developed can be applied regionally to predict tree water use using remotely sensed SIF datasets in areas of low data availability or accessibility. Full article
(This article belongs to the Special Issue Monitoring Ecohydrology with Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Site locations of Mallee Cliffs <span class="html-italic">E. largiflorens</span> sparse site (MC<sub>BB_S</sub>), Mallee Cliffs <span class="html-italic">E. largiflorens</span> moderately sparse site (MC<sub>BB_MS</sub>), and Lindsay Forest <span class="html-italic">E. camaldulensis</span> open woodland site (LF<sub>RRG_OW</sub>) in the Murray–Darling Basin (MDB). Pictures in the first row were taken by site cameras.</p>
Full article ">Figure 2
<p>Flowchart of the simulations. Step 1 (S1)—determine 3-D vegetation structure from LiDAR data. Step 2 (S2)—Sentinel-2 LAI (LAI<sub>S2</sub>) partitioned into canopy (LAI<sub>c</sub>) and understory LAI (LAI<sub>u</sub>). Step 3 (S3)—retrieve leaf area density (LAD) and six bioparameters (BP), including leaf mesophyll structure, leaf chlorophyll a + b content, total carotenoid content, senescence material, leaf dry matter content, and leaf water content, using a two-step inversion method with two lookup tables (LUT<sub>LAD</sub> and LUT<sub>BP</sub>). Step 4 (S4)—use FluorFLiES model with vegetation structure and bioparameters to simulate the full spectrum sun-induced chlorophyll fluorescence (SIF). Step 5 (S5)—statistical analysis conducted on the relationship between simulated SIF and field measured transpiration (T). Height (H); radius (R). H and R mean height and radius. Simulated LAI (LAI<sub>F</sub>), LAD (LAD<sub>F</sub>), and fraction of absorbed photosynthetically active radiation (FAPAR<sub>F</sub>) with the FLiES model. Leaf reflectance (ref); transmittance (tran); soil reflectance (R<sub>soil</sub>); excitation-fluorescence matrices (EF-matrix); photosynthetically active radiation (PAR).</p>
Full article ">Figure 3
<p>Scatterplot of transpiration (T) against (<b>a</b>) sun-induced chlorophyll fluorescence (SIF) emission at 685 nm (SIF<sub>red</sub>), (<b>b</b>) 699 nm (SIF<sub>valley</sub>), (<b>c</b>) 720 nm (SIF<sub>water vapour</sub>), (<b>d</b>) 740 nm (SIF<sub>NIR</sub>), (<b>e</b>) 760 nm (SIF<sub>O2-A</sub>) at hourly, daily, and monthly scale. All <span class="html-italic">p</span>-values &lt; 0.05. Blue colour represents the hourly data, green colour represents the daily data, and red colour represents the monthly data.</p>
Full article ">Figure 4
<p>Scatterplot of transpiration (T) against sun-induced chlorophyll fluorescence (SIF) emissions of selected SIF bands from total leaves, canopy, and understory components. All <span class="html-italic">p</span>-values &lt; 0.05. SIF<sub>red</sub> is SIF emission at 685 nm; SIF<sub>valley</sub> is 699 nm; SIF<sub>water vapour</sub> is 720 nm; SIF<sub>NIR</sub> is 740 nm; SIF<sub>O2-A</sub> is 760 nm. SIF<sub>total</sub> is the total SIF emission at the leaf level, SIF<sub>canopy</sub> is the SIF emission from the canopy, SIF<sub>understory</sub> is the SIF emission from understory vegetation.</p>
Full article ">Figure 5
<p>Time series of sun-induced chlorophyll fluorescence (SIF) simulations (685 and 740 nm) and transpiration (T) measurements at (<b>a</b>) Mallee Cliffs E. largiflorens sparse site (MCBB_S), (<b>b</b>) Mallee Cliffs <span class="html-italic">E. largiflorens</span> moderately sparse site (MCBB_MS), and (<b>c</b>) Lindsay Forest <span class="html-italic">E. camaldulensis</span> open woodland site (LFRRG_OW).</p>
Full article ">Figure 6
<p>Coefficient of determinations of correlations between sun-induced chlorophyll fluorescence (SIF) and transpiration (T) at the hourly scale with defined understory leaf area index (LAI<sub>u</sub>) for each site. MC<sub>BB_S</sub> is Mallee Cliffs <span class="html-italic">E. largiflorens</span> sparse site, MC<sub>BB_MS</sub> is Mallee Cliffs <span class="html-italic">E. largiflorens</span> moderately sparse site, and LF<sub>RRG_OW</sub> is Lindsay Forest <span class="html-italic">E. camaldulensis</span> open woodland site. SIF<sub>red</sub> is SIF emission at 685 nm; SIF<sub>valley</sub> is 699 nm; SIF<sub>water vapour</sub> is 720 nm; SIF<sub>NIR</sub> is 740 nm; SIF<sub>O2-A</sub> is 760 nm.</p>
Full article ">Figure 7
<p>Coefficient of determination of selected sun-induced chlorophyll fluorescence (SIF) bands and transpiration (T) at the hourly scale for various leaf area index (LAI) conditions. SIF<sub>red</sub> is SIF emission at 685 nm; SIF<sub>valley</sub> is 699 nm; SIF<sub>water vapour</sub> is 720 nm; SIF<sub>NIR</sub> is 740 nm; SIF<sub>O2-A</sub> is 760 nm.</p>
Full article ">Figure 8
<p>Coefficient of determination of sun-induced chlorophyll fluorescence (SIF) and transpiration (T) at the hourly scale with defined temperature (TEMP) categories. MC<sub>BB_S</sub> is Mallee Cliffs <span class="html-italic">E. largiflorens</span> sparse site, MC<sub>BB_MS</sub> is Mallee Cliffs <span class="html-italic">E. largiflorens</span> moderately sparse site, and LF<sub>RRG_OW</sub> is Lindsay Forest <span class="html-italic">E. camaldulensis</span> open woodland site. SIF<sub>red</sub> is SIF emission at 685 nm; SIF<sub>valley</sub> is 699 nm; SIF<sub>water vapour</sub> is 720 nm; SIF<sub>NIR</sub> is 740 nm; SIF<sub>O2-A</sub> is 760 nm. “*” denotes <span class="html-italic">p</span>-value is more than 0.05.</p>
Full article ">Figure 9
<p>Coefficient of determinations of sun-induced chlorophyll fluorescence (SIF) and transpiration (T) at the hourly scale with various daily rainfall (mm day<sup>−1</sup>) conditions. MC<sub>BB_S</sub> is Mallee Cliffs <span class="html-italic">E. largiflorens</span> sparse site, MC<sub>BB_MS</sub> is Mallee Cliffs <span class="html-italic">E. largiflorens</span> moderately sparse site, and LF<sub>RRG_OW</sub> is Lindsay Forest <span class="html-italic">E. camaldulensis</span> open woodland site. SIF<sub>red</sub> is SIF emission at 685 nm; SIF<sub>valley</sub> is 699 nm; SIF<sub>water vapour</sub> is 720 nm; SIF<sub>NIR</sub> is 740 nm; SIF<sub>O2-A</sub> is 760 nm. “*” denotes <span class="html-italic">p</span>-value is more than 0.05.</p>
Full article ">
25 pages, 8268 KiB  
Article
Using Sentinel-2-Based Metrics to Characterize the Spatial Heterogeneity of FLEX Sun-Induced Chlorophyll Fluorescence on Sub-Pixel Scale
by Nela Jantol, Egor Prikaziuk, Marco Celesti, Itza Hernandez-Sequeira, Enrico Tomelleri, Javier Pacheco-Labrador, Shari Van Wittenberghe, Filiberto Pla, Subhajit Bandopadhyay, Gerbrand Koren, Bastian Siegmann, Tarzan Legović, Hrvoje Kutnjak and M. Pilar Cendrero-Mateo
Remote Sens. 2023, 15(19), 4835; https://doi.org/10.3390/rs15194835 - 5 Oct 2023
Cited by 1 | Viewed by 2412
Abstract
Current and upcoming Sun-Induced chlorophyll Fluorescence (SIF) satellite products (e.g., GOME, TROPOMI, OCO, FLEX) have medium-to-coarse spatial resolutions (i.e., 0.3–80 km) and integrate radiances from different sources into a single ground surface unit (i.e., pixel). However, intrapixel heterogeneity, i.e., different soil and vegetation [...] Read more.
Current and upcoming Sun-Induced chlorophyll Fluorescence (SIF) satellite products (e.g., GOME, TROPOMI, OCO, FLEX) have medium-to-coarse spatial resolutions (i.e., 0.3–80 km) and integrate radiances from different sources into a single ground surface unit (i.e., pixel). However, intrapixel heterogeneity, i.e., different soil and vegetation fractional cover and/or different chlorophyll content or vegetation structure in a fluorescence pixel, increases the challenge in retrieving and quantifying SIF. High spatial resolution Sentinel-2 (S2) data (20 m) can be used to better characterize the intrapixel heterogeneity of SIF and potentially extend the application of satellite-derived SIF to heterogeneous areas. In the context of the COST Action Optical synergies for spatiotemporal SENsing of Scalable ECOphysiological traits (SENSECO), in which this study was conducted, we proposed direct (i.e., spatial heterogeneity coefficient, standard deviation, normalized entropy, ensemble decision trees) and patch mosaic (i.e., local Moran’s I) approaches to characterize the spatial heterogeneity of SIF collected at 760 and 687 nm (SIF760 and SIF687, respectively) and to correlate it with the spatial heterogeneity of selected S2 derivatives. We used HyPlant airborne imagery acquired over an agricultural area in Braccagni (Italy) to emulate S2-like top-of-the-canopy reflectance and SIF imagery at different spatial resolutions (i.e., 300, 20, and 5 m). The ensemble decision trees method characterized FLEX intrapixel heterogeneity best (R2 > 0.9 for all predictors with respect to SIF760 and SIF687). Nevertheless, the standard deviation and spatial heterogeneity coefficient using k-means clustering scene classification also provided acceptable results. In particular, the near-infrared reflectance of terrestrial vegetation (NIRv) index accounted for most of the spatial heterogeneity of SIF760 in all applied methods (R2 = 0.76 with the standard deviation method; R2 = 0.63 with the spatial heterogeneity coefficient method using a scene classification map with 15 classes). The models developed for SIF687 did not perform as well as those for SIF760, possibly due to the uncertainties in fluorescence retrieval at 687 nm and the low signal-to-noise ratio in the red spectral region. Our study shows the potential of the proposed methods to be implemented as part of the FLEX ground segment processing chain to quantify the intrapixel heterogeneity of a FLEX pixel and/or as a quality flag to determine the reliability of the retrieved fluorescence. Full article
Show Figures

Figure 1

Figure 1
<p>Study area, Braccagni, Italy. Map of the area on the left was produced using Sentinel-2 RGB bands (B4-B3-B2).</p>
Full article ">Figure 2
<p>Workflow diagram. (<b>a</b>) HyPlant reflectance image (4.5 × 4.5 m) was aggregated to mimic S2 spectral and spatial resolution (13 bands and 20 × 20 m ~ S2-R<sub>20</sub>). At the same time, HyPlant fluorescence products were spatially aggregated to 5 × 5 m resolution (SIF<sub>687,5</sub> and SIF<sub>760,5</sub>) (see data preparation section). * For the ensemble decision trees method, SIF was additionally aggregated to 300 × 300 m. (<b>b</b>) Synthetic S2-R<sub>20</sub> bands were used to obtain the biophysical traits (S2-BT<sub>20</sub>) and vegetation indices (S2-VI<sub>20</sub>), which were later used to characterize the spatial heterogeneity of SIF (<a href="#remotesensing-15-04835-t001" class="html-table">Table 1</a> and <a href="#remotesensing-15-04835-t002" class="html-table">Table 2</a>). (<b>c</b>) The Structural Similarity Index Measure (SSIM) was implemented to filter the input data (i.e., S2 bands, VIs, BT) used in the study. (<b>d</b>) To determine the spatial heterogeneity of a FLEX pixel, a 300 × 300 m grid was applied to the S2 synthetic (S2-R<sub>20</sub>, S2-BT<sub>20</sub> and S2-VI<sub>20</sub>) and SIF (SIF<sub>687,5</sub> and SIF<sub>760,5</sub>) resampled images. Each FLEX pixel potentially contained 15 × 15 S2 pixels and 60 × 60 SIF 5 × 5 m pixels. (<b>e</b>) Different heterogeneity methods (see methods to characterize sun-induced chlorophyll fluorescence heterogeneity section) were applied to the S2 and HyPlant SIF products using the 300 × 300 FLEX grid defined in step (<b>d</b>). A FLEX heterogeneity product was obtained for each S2 predictor (S2-R<sub>20</sub>, S2-BT<sub>20</sub> and S2-VI<sub>20</sub>) and SIF reference data (SIF<sub>687,5</sub> and SIF<sub>760,5</sub>). (<b>f</b>) Finally, we compared S2 vs. SIF heterogeneity products using linear regression (see models’ performance section).</p>
Full article ">Figure 3
<p>Tukey’s test applied to the Structural Similarity Index Measure (SSIM) used to measure the similarity between two normalized images (SIF<sub>760,20</sub>, SIF<sub>687,20</sub>, and respective Sentinel-2 predictors). Dashed vertical lines indicate the similarity threshold of ±0.1 SSIM. Bands with SSIM values above this threshold for both SIF<sub>760</sub> and SIF<sub>687</sub> were used for further analysis.</p>
Full article ">Figure 4
<p>Scene classification maps: (<b>a</b>) Map produced using supervised classification with 5 classes; (<b>b</b>) Map produced using k-means algorithm with 15 classes.</p>
Full article ">Figure 5
<p>Dataset imagery: (<b>a</b>) Sun-induced fluorescence with 5 × 5 m resolution; (<b>b</b>) Vegetation indices maps from Sentinel-2 data NIRv, NDVI, EVI, ChlRE, MSI with 20 × 20 m resolution; (<b>c</b>) Biophysical traits maps for LCC, fAPAR, fCover and LAI at 20 × 20 m resolution; (<b>d</b>) Reflectance bands maps from Sentinel-2 data B6, B7, B8, B8A, B9 with 20 × 20 m resolution. Values in maps are shown as 2nd and 98th percentiles of the raster band values. Lower values are shown in blue, higher values in red.</p>
Full article ">Figure 6
<p>(<b>a</b>) Histogram of normalized values for sun-induced fluorescence; (<b>b</b>) Vegetation indices; (<b>c</b>) Biophysical traits; (<b>d</b>) Reflectance bands. Notice that all the distributions (<b>b</b>–<b>d</b>) are skewed, compared to (<b>a</b>).</p>
Full article ">Figure 7
<p>Square of Pearson’s correlation coefficient between reference <span class="html-italic">SIF</span><sub>760</sub> and <span class="html-italic">SIF</span><sub>687</sub> heterogeneity and predictors’ heterogeneity: Sentinel-2 derived vegetation indices, biophysical traits, reflectance bands and their stacks using the following methods: (<b>a</b>) Ensemble decision trees; (<b>b</b>) Standard deviation; (<b>c</b>) Local Moran’s I; (<b>d</b>) Spatial heterogeneity coefficient using scene classification with 5 classes (SCL-5); (<b>e</b>) Spatial heterogeneity coefficient using scene classification with 15 classes (SCL-15); (<b>f</b>) Normalized entropy. *** <span class="html-italic">p</span>-value ≤ 0.001; ** <span class="html-italic">p</span>-value ≤ 0.01; * <span class="html-italic">p</span>-value ≤ 0.05, ns—<span class="html-italic">p</span>-value &gt; 0.05.</p>
Full article ">Figure 8
<p>The number of times a pixel (pixel ID shown as numbers next to pixels) was considered an outlier (top 6 RMSE) by ensemble decision trees, spatial heterogeneity coefficient with SCL-15 and standard deviation methods using the most important predictors from each category (i.e., NIRv—vegetation index category, fAPAR—biophysical trait category, B7—reflectance band category) for (<b>a</b>) SIF<sub>760,300</sub>, (<b>b</b>) SIF<sub>687,300</sub> and (<b>c</b>) the sum of counts for both SIF<sub>760,300</sub> and SIF<sub>687,300</sub>. The maximum possible count for (<b>a</b>,<b>b</b>) is 9 (three models, three predictors), for (<b>c</b>) 18.</p>
Full article ">Figure 9
<p>Top two outlier pixels (ID 166 and 124) with input data for (<b>a</b>,<b>h</b>) SIF<sub>760,5</sub>; (<b>b</b>,<b>i</b>) SIF<sub>687,5</sub>; (<b>c</b>,<b>j</b>) NIRv; (<b>d</b>,<b>k</b>) fAPAR; (<b>e</b>,<b>l</b>) B7; (<b>f</b>,<b>m</b>) SCL-5 classes; (<b>g</b>,<b>n</b>) SCL-15 classes. SIF units are W m<sup>−2</sup> sr<sup>−1</sup> µm<sup>−1</sup>, NIRv and fAPAR are unitless; and B7 is in sr<sup>−1</sup>. Classes for SCL-5 are water (cyan), cropland (olive), pasture (green), and other (gray), as in <a href="#remotesensing-15-04835-f004" class="html-fig">Figure 4</a>. Classes for SCL-15 are discrete values and represent spectral rather than land cover classes.</p>
Full article ">Figure 10
<p>Heterogeneity maps (300 × 300 m) for standard deviation, ensemble decision trees and spatial heterogeneity coefficient SCL-15 methods; (<b>a</b>,<b>d</b>,<b>g</b>) reference SIF<sub>760</sub>; (<b>b</b>,<b>e</b>,<b>h</b>) best predictor NIRv; (<b>c</b>,<b>f</b>,<b>i</b>) scatter plots with lowest (green circle) and highest (red circle) heterogeneity pixels highlighted.</p>
Full article ">Figure 11
<p>Highest and lowest heterogeneity pixels from best-performing models (standard deviation, spatial heterogeneity coefficient using 15 classes, ensemble decision trees) with input data shown for (<b>a</b>,<b>d</b>) SIF<sub>760,5</sub>; (<b>b</b>,<b>e</b>) NIRv<sub>20</sub>; (<b>c</b>,<b>f</b>) scene classification map 15 classes.</p>
Full article ">Figure A1
<p>Heterogeneity maps (300 × 300 m) for the standard deviation method: (<b>a</b>) Reference SIF<sub>687,5</sub>; (<b>b</b>) Best predictor NIRv; (<b>c</b>) Scatter plot with lowest (red circle) and highest (green circle) heterogeneity pixels highlighted; (<b>d</b>) 5 m pixel with high heterogeneity for SIF<sub>687,5</sub>; (<b>e</b>) 20 m pixel with high heterogeneity for NIRv; (<b>f</b>) Scene classification with 15 classes for a pixel with high heterogeneity; (<b>g</b>) 5 m pixel with low heterogeneity for SIF<sub>687,5</sub>; (<b>h</b>) 20 m pixel with low heterogeneity for NIRv; (<b>i</b>) Scene classification with 15 classes for a pixel with low heterogeneity.</p>
Full article ">Figure A2
<p>Heterogeneity maps (300 × 300 m) for the ensemble decision trees method: (<b>a</b>) Reference SIF687,5; (<b>b</b>) Best predictor NIRv; (<b>c</b>) Scatter plot with lowest (red circle) and highest (green circle) heterogeneity pixels highlighted; (<b>d</b>) 5 m pixel with high heterogeneity for SIF<sub>687,5</sub>; (<b>e</b>) 20 m pixel with high heterogeneity for NIRv; (<b>f</b>) Scene classification with 15 classes for a pixel with high heterogeneity; (<b>g</b>) 5 m pixel with low heterogeneity for SIF<sub>687,5</sub>; (<b>h</b>) 20 m pixel with low heterogeneity for NIRv; (<b>i</b>) Scene classification with 15 classes for a pixel with low heterogeneity.</p>
Full article ">Figure A3
<p>Heterogeneity maps (300 × 300 m) for the spatial heterogeneity coefficient method: (<b>a</b>) Reference SIF<sub>687,5</sub>; (<b>b</b>) Best predictor NIRv; (<b>c</b>) Scatter plot with lowest (red circle) and highest (green circle) heterogeneity pixels highlighted; (<b>d</b>) 5 m pixel with high heterogeneity for SIF<sub>687,5</sub>; (<b>e</b>) 20 m pixel with high heterogeneity for NIRv; (<b>f</b>) Scene classification with 15 classes for a pixel with high heterogeneity; (<b>g</b>) 5 m pixel with low heterogeneity for SIF<sub>687,5</sub>; (<b>h</b>) 20 m pixel with low heterogeneity for NIRv; (<b>i</b>) Scene classification with 15 classes for a pixel with low heterogeneity.</p>
Full article ">
10 pages, 3028 KiB  
Article
Highly Efficient and Stable CsPbBr3-Alginic Acid Composites for White Light-Emitting Diodes
by Muyi Wang, Song Wang, Renjie Chen, Mengmeng Zhu, Yunpeng Liu, Haojie Ding, Jun Ren, Tongtong Xuan and Huili Li
Coatings 2023, 13(6), 1062; https://doi.org/10.3390/coatings13061062 - 7 Jun 2023
Cited by 4 | Viewed by 1780
Abstract
All-inorganic perovskite nanocrystals (NCs) have attractive potential for applications in display and lighting fields due to their special optoelectronic properties. However, they still suffer from poor water and thermal stability. In this work, green CsPbBr3-alginic acid (CsPbBr3-AA) perovskite composites [...] Read more.
All-inorganic perovskite nanocrystals (NCs) have attractive potential for applications in display and lighting fields due to their special optoelectronic properties. However, they still suffer from poor water and thermal stability. In this work, green CsPbBr3-alginic acid (CsPbBr3-AA) perovskite composites were synthesized by an in situ hot-injection process which showed a high photoluminescence quantum yield (PLQY) of 86.43% and improved moisture and thermal stability. Finally, white light-emitting diodes (WLEDs) were fabricated by combining the green CsPbBr3-AA perovskite composites with red K2SiF6:Mn4+ phosphors and blue InGaN LED chips. The WLEDs show a relatively high luminous efficacy of 36.4 lm/W and a wide color gamut (124% of the National Television System Committee). These results indicate that the green CsPbBr3-AA perovskite composites have great potential applications in backlight displays. Full article
(This article belongs to the Collection Feature Papers of Coatings for Energy Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns of CsPbBr<sub>3</sub> NCs and CsPbBr<sub>3</sub>-AA composites. (<b>b</b>) FTIR spectra of AA, CsPbBr<sub>3</sub> NCs and the CsPbBr<sub>3</sub>-AA composite. (<b>c</b>) XPS spectra of CsPbBr<sub>3</sub> NCs, AA, and CsPbBr<sub>3</sub>-AA composites. (<b>d</b>) High-resolution XPS spectra of O 1 s of AA and CsPbBr<sub>3</sub>-AA composites. (<b>e</b>) Schematic diagram of CsPbBr<sub>3</sub>-AA composites.</p>
Full article ">Figure 2
<p>TEM and the corresponding histogram of the particle size distribution (<b>b</b>) of the CsPbBr<sub>3</sub> NCs (<b>a</b>) and CsPbBr<sub>3</sub>-AA composite (<b>b</b>). HRTEM images of the CsPbBr<sub>3</sub> NCs (<b>c</b>) and CsPbBr<sub>3</sub>-AA composite (<b>d</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) PL spectra of pure CsPbBr<sub>3</sub> NCs and CsPbBr<sub>3</sub>-AA composites. (<b>b</b>) Absorption spectra of AA, pure CsPbBr<sub>3</sub> NCs, and the CsPbBr<sub>3</sub>-AA composite. (<b>c</b>) PLQYs of CsPbBr<sub>3</sub> NCs and CsPbBr<sub>3</sub>-AA composites with different concentrations of AA. (<b>d</b>) PL decay curves of CsPbBr<sub>3</sub> pure NCs and CsPbBr<sub>3</sub>-AA composites.</p>
Full article ">Figure 4
<p>(<b>a</b>) Relative PL intensity of the pure CsPbBr<sub>3</sub> NCs and CsPbBr<sub>3</sub>-AA contact angle of CsPbBr<sub>3</sub> NCs (<b>b</b>) and CsPbBr<sub>3</sub>-AA (<b>c</b>) with water.</p>
Full article ">Figure 5
<p>Photographs of (<b>a</b>) unlighted and (<b>b</b>) lighted of the WLEDs operated at 3.7 V and 20 mA. (<b>c</b>) EL spectrum of the WLEDs at 20 mA and 3.7 V, as well as (<b>d</b>) the CIE coordinates.</p>
Full article ">
11 pages, 1919 KiB  
Article
Study on Local-Structure Symmetrization of K2XF6 Crystals Doped with Mn4+ Ions by First-Principles Calculations
by Mega Novita, Sigit Ristanto, Ernawati Saptaningrum, Slamet Supriyadi, Dian Marlina, Ferdy Semuel Rondonuwu, Alok Singh Chauhan, Benjamin Walker, Kazuyoshi Ogasawara, Michal Piasecki and Mikhail G. Brik
Materials 2023, 16(11), 4046; https://doi.org/10.3390/ma16114046 - 29 May 2023
Cited by 5 | Viewed by 1774
Abstract
The crystals of Mn4+-activated fluorides, such as those of the hexafluorometallate family, are widely known for their luminescence properties. The most commonly reported red phosphors are A2XF6: Mn4+ and BXF6: Mn4+ fluorides, where [...] Read more.
The crystals of Mn4+-activated fluorides, such as those of the hexafluorometallate family, are widely known for their luminescence properties. The most commonly reported red phosphors are A2XF6: Mn4+ and BXF6: Mn4+ fluorides, where A represents alkali metal ions such as Li, Na, K, Rb, Cs; X=Ti, Si, Ge, Zr, Sn, B = Ba and Zn; and X = Si, Ge, Zr, Sn, and Ti. Their performance is heavily influenced by the local structure around dopant ions. Many well-known research organizations have focused their attention on this area in recent years. However, there has been no report on the effect of local structural symmetrization on the luminescence properties of red phosphors. The purpose of this research was to investigate the effect of local structural symmetrization on the polytypes of K2XF6 crystals, namely Oh-K2MnF6, C3v-K2MnF6, Oh-K2SiF6, C3v-K2SiF6, D3d-K2GeF6, and C3v-K2GeF6. These crystal formations yielded seven-atom model clusters. Discrete Variational Xα (DV-Xα) and Discrete Variational Multi Electron (DVME) were the first principles methods used to compute the Molecular orbital energies, multiplet energy levels, and Coulomb integrals of these compounds. The multiplet energies of Mn4+ doped K2XF6 crystals were qualitatively reproduced by taking lattice relaxation, Configuration Dependent Correction (CDC), and Correlation Correction (CC) into account. The 4A2g4T2g (4F) and 4A2g4T1g (4F) energies increased when the Mn-F bond length decreased, but the 2Eg4A2g energy decreased. Because of the low symmetry, the magnitude of the Coulomb integral became smaller. As a result, the decreasing trend in the R-line energy could be attributed to a decreased electron–electron repulsion. Full article
(This article belongs to the Special Issue Glasses and Ceramics for Luminescence Applications)
Show Figures

Figure 1

Figure 1
<p>The crystal structure of K<sub>2</sub>XF<sub>6</sub> (X = Mn, Si, or Ge) with (<b>a</b>) Cubic structure and space group <math display="inline"><semantics> <mrow> <mi>F</mi> <mi>m</mi> <mover accent="true"> <mn>3</mn> <mo>¯</mo> </mover> <mi>m</mi> </mrow> </semantics></math>, (<b>b</b>) Rhombohedral structure with space group <math display="inline"><semantics> <mrow> <mi>P</mi> <mover accent="true"> <mn>3</mn> <mo>¯</mo> </mover> <mi>m</mi> <mn>1</mn> </mrow> </semantics></math>, and (<b>c</b>) Hexagonal structure with space group <math display="inline"><semantics> <mrow> <mi>P</mi> <mn>63</mn> <mi>m</mi> <mi>c</mi> </mrow> </semantics></math> as seen from the <span class="html-italic">c</span> axis. The seven atoms represent clusters with (<b>d</b>) <span class="html-italic">O<sub>h</sub></span>, (<b>e</b>) <span class="html-italic">D</span><sub>3<span class="html-italic">d</span></sub>, and (<b>f</b>) <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub> symmetry, with the Mn<sup>4+</sup> ion in the core.</p>
Full article ">Figure 2
<p>Molecular orbital energies of <span class="html-italic">O<sub>h</sub></span>-K<sub>2</sub>MnF<sub>6</sub>, <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>-K<sub>2</sub>MnF<sub>6</sub>, <span class="html-italic">O<sub>h</sub></span>-K<sub>2</sub>SiF<sub>6</sub>, <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>-K<sub>2</sub>SiF<sub>6</sub>, <span class="html-italic">D</span><sub>3<span class="html-italic">d</span></sub>-K<sub>2</sub>GeF<sub>6</sub>, and <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>-K<sub>2</sub>GeF<sub>6</sub> doped with Mn<sup>4+</sup>. The Valence Band (VB) is represented by black solid lines. The Conduction Band (CB) is shown by the black dashed lines. The <span class="html-italic">t</span><sub>2<span class="html-italic">g</span></sub> levels are indicated by the solid red lines, while the <span class="html-italic">e<sub>g</sub></span> levels are indicated by the dashed blue lines.</p>
Full article ">Figure 3
<p>Pure K<sub>2</sub>MnF<sub>6</sub> and K<sub>2</sub>SiF<sub>6</sub>: Mn<sup>4+</sup> multiplet energy diagrams. Additionally, demonstrated is the impact of CDC, CC, and lattice relaxation. The left side of each column explains the calculation using <span class="html-italic">O<sub>h</sub></span>-symmetric clusters, while the right side describes the calculation using <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>-symmetric clusters. Black and red lines denote the doublet and quartet states, respectively. When the lower symmetry (<span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>) was used, these states were further divided into the <span class="html-italic">a</span> (dashed lines) and <span class="html-italic">e</span> (solid lines) categories. There are the doublet states <sup>2</sup>E,.<sup>2</sup>T<sub>2,</sub> and <sup>2</sup>T<sub>1,</sub> as well as the quartet states <sup>4</sup>T<sub>2</sub> and <sup>4</sup>T<sub>1a</sub>. The <sup>4</sup>A<sub>2</sub> is the ground state. The absorption occurred during the electronic transitions from the ground <sup>4</sup>A<sub>2</sub> state to <sup>4</sup>T<sub>2</sub> and <sup>4</sup>T<sub>1a</sub> states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the <sup>2</sup>E state to the ground <sup>4</sup>A<sub>2</sub> state (R-line), as illustrated by the red arrow. More information can be found in the text.<div class="html-table-p">Pure K<sub>2</sub>MnF<sub>6</sub> and K<sub>2</sub>SiF<sub>6</sub>: Mn<sup>4+</sup> multiplet energy diagrams. Additionally, demonstrated is the impact of CDC, CC, and lattice relaxation. The left side of each column explains the calculation using <span class="html-italic">O<sub>h</sub></span>-symmetric clusters, while the right side describes the calculation using <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>-symmetric clusters. Black and red lines denote the doublet and quartet states, respectively. When the lower symmetry (<span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub>) was used, these states were further divided into the <span class="html-italic">a</span> (dashed lines) and <span class="html-italic">e</span> (solid lines) categories. There are the doublet states <sup>2</sup>E,.<sup>2</sup>T<sub>2,</sub> and <sup>2</sup>T<sub>1,</sub> as well as the quartet states <sup>4</sup>T<sub>2</sub> and <sup>4</sup>T<sub>1a</sub>. The <sup>4</sup>A<sub>2</sub> is the ground state. The absorption occurred during the electronic transitions from the ground <sup>4</sup>A<sub>2</sub> state to <sup>4</sup>T<sub>2</sub> and <sup>4</sup>T<sub>1a</sub> states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the <sup>2</sup>E state to the ground <sup>4</sup>A<sub>2</sub> state (R-line), as illustrated by the red arrow. More information can be found in the text.</p>
Full article ">Figure 4
<p>K<sub>2</sub>GeF<sub>6</sub>: Mn<sup>4+</sup> multiplet energy diagrams. Additionally demonstrated is the impact of corrections, including CDC, CC, and lattice relaxation. A calculation using clusters with <span class="html-italic">D</span><sub>3<span class="html-italic">d</span></sub> symmetry is described on the left side of each column, while a calculation using clusters with <span class="html-italic">C</span><sub>3<span class="html-italic">v</span></sub> symmetry is described on the right side. The <span class="html-italic">O<sub>h</sub></span> symmetry notations, in this instance, were borrowed. Black and red lines denote the doublet and quintet states, respectively; dashed (<span class="html-italic">a</span> level) and solid lines (<span class="html-italic">e</span> level) denote the multiplet splitting. There are the doublet states <sup>2</sup>E,.<sup>2</sup>T<sub>2,</sub> and <sup>2</sup>T<sub>1,</sub> as well as the quartet states <sup>4</sup>T<sub>2</sub> and <sup>4</sup>T<sub>1a</sub>. The <sup>4</sup>A<sub>2</sub> is the ground state. The absorption occurred during the electronic transitions from the ground <sup>4</sup>A<sub>2</sub> state to <sup>4</sup>T<sub>2</sub> and <sup>4</sup>T<sub>1a</sub> states (U- and Y-band, respectively), as illustrated by the green and blue arrows. The emission, on the other hand, happened as an electronic transition from the <sup>2</sup>E state to the ground <sup>4</sup>A<sub>2</sub> state (R-line), as illustrated by the red arrow. More information can be found in the text.</p>
Full article ">
15 pages, 3070 KiB  
Article
Exploring the Sensitivity of Solar-Induced Chlorophyll Fluorescence at Different Wavelengths in Response to Drought
by Shan Xu, Zhigang Liu, Shuai Han, Zhuang Chen, Xue He, Huarong Zhao and Sanxue Ren
Remote Sens. 2023, 15(4), 1077; https://doi.org/10.3390/rs15041077 - 16 Feb 2023
Cited by 8 | Viewed by 2452
Abstract
Due to the mechanistic coupling between solar-induced chlorophyll fluorescence (SIF) and photosynthesis, SIF has an advantage over greenness-based vegetation indices in detecting drought. Since photosystem I (PSI) contributes very little to red SIF, red SIF is assumed to be more responsive to environmental [...] Read more.
Due to the mechanistic coupling between solar-induced chlorophyll fluorescence (SIF) and photosynthesis, SIF has an advantage over greenness-based vegetation indices in detecting drought. Since photosystem I (PSI) contributes very little to red SIF, red SIF is assumed to be more responsive to environmental stress than far-red SIF. However, in addition to affecting photosynthesis, drought also has an impact on vegetation chlorophyll concentration and thus affects the reabsorption process of red SIF. When these responses are entangled, the sensitivity of SIF in the red and far-red regions in response to drought is not yet clear. In this study, we conducted a water stress experiment on maize in the field and measured the upward and downward leaf SIF spectra by a spectrometer assembled with a leaf clip. Simultaneously, leaf-level active fluorescence was measured with a pulse-amplified modulation (PAM) fluorometer. We found that SIF, after normalization by photosynthetically active radiation (PAR) and dark-adapted minimal fluorescence (Fo), is a better estimation of SIF yield. By comparing the wavelength-dependent link between SIF yield and nonphotochemical quenching (NPQ) across the range of 660 to 800 nm, the results show that red SIF and far-red SIF have different sensitivities in response to drought. SIF yield in the far-red region has a strong and stable correlation with NPQ. Drought not only reduces red SIF due to photosynthetic regulation, but it also increases red SIF by reducing chlorophyll content (weakening the reabsorption effect). The co-existence of these two contradictory effects makes the red SIF of leaf level unable to reliably indicate NPQ. In addition, the red:far-red ratio of downward SIF and the ratio between the downward SIF and upward SIF at the red peak can be good indicators of chlorophyll content. These findings can help to interpret SIF variations in remote sensing techniques and fully exploit SIF information in red and far-red regions when monitoring plant water stress. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of measuring spectroscopy and PAM fluorescence on the leaf scale. A FluoWat leaf clip was used to measure reflectance (r), transmittance (t), and sun-induced fluorescence (SIF) in the visible and near-infrared wavelength range (350–1100 nm) by placing a fiber optic either in upward or downward position. After placing the shortpass to restrict incoming PAR to visible wavelengths up to 650 nm, upward and downward SIF are measured. A PAM-2500 fluorometer was used to measure active fluorescence near the clip.</p>
Full article ">Figure 2
<p>Differences in the key physiological parameters in the three water treatments (mean ± std, n = 3) measured on 21 August and 25 August. Four key parameters are the maximum quantum yield of PSII photochemistry (<math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mi>v</mi> </msub> <mo>/</mo> <msub> <mi>F</mi> <mi>m</mi> </msub> </mrow> </semantics></math>), nonphotochemical quenching (NPQ), operating quantum yield of photochemistry in PSII (<math display="inline"><semantics> <mrow> <mrow> <mi mathvariant="sans-serif">Φ</mi> <mi mathvariant="normal">P</mi> </mrow> <mo stretchy="false">)</mo> </mrow> </semantics></math>, and steady state active fluorescence (<math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mi>t</mi> </msub> </mrow> </semantics></math>) normalized to dark-adapted fluorescence (<math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mi>o</mi> </msub> </mrow> </semantics></math>). C, D1 and D2 indicate the control, moderate drought, and severe drought plots, respectively.</p>
Full article ">Figure 3
<p>(<b>A</b>,<b>B</b>) Two forms of total SIF yield <math display="inline"><semantics> <mrow> <mfrac> <mrow> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mi>λ</mi> </msub> <mi>t</mi> <mi>o</mi> <mi>t</mi> </mrow> <mrow> <mi>A</mi> <mi>P</mi> <mi>A</mi> <mi>R</mi> </mrow> </mfrac> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mfrac> <mrow> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mi>λ</mi> </msub> <mi>t</mi> <mi>o</mi> <mi>t</mi> </mrow> <mrow> <mi>P</mi> <mi>A</mi> <mi>R</mi> <mo>×</mo> <msub> <mi>F</mi> <mi>o</mi> </msub> </mrow> </mfrac> </mrow> </semantics></math> spectra (660~800 nm) for all leaf samples; colored lines represent the nonphotochemical quenching (NPQ) observed in each leaf. (<b>C</b>,<b>D</b>) Coefficient of determination (<math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> </mrow> </semantics></math>) for all wavelengths of <math display="inline"><semantics> <mrow> <mi>SIF</mi> </mrow> </semantics></math> yield against NPQ. The coefficient of determination represents the performance of the linear regression.</p>
Full article ">Figure 4
<p>(<b>A</b>,<b>B</b>) The upward and downward SIF yield (<math display="inline"><semantics> <mrow> <mfrac> <mrow> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mi>λ</mi> </msub> <mo>↑</mo> </mrow> <mrow> <mi>P</mi> <mi>A</mi> <mi>R</mi> <mo>×</mo> <msub> <mi>F</mi> <mi>o</mi> </msub> </mrow> </mfrac> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mfrac> <mrow> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mi>λ</mi> </msub> <mo>↓</mo> </mrow> <mrow> <mi>P</mi> <mi>A</mi> <mi>R</mi> <mo>×</mo> <msub> <mi>F</mi> <mi>o</mi> </msub> </mrow> </mfrac> </mrow> </semantics></math>) spectra (660~800 nm); colored lines represent the nonphotochemical quenching (NPQ) observed in each leaf. (<b>C</b>,<b>D</b>) The coefficient of determination (<math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> </mrow> </semantics></math>) for all wavelengths of SIF yield against NPQ. The coefficient of determination represents the linear regression performance.</p>
Full article ">Figure 5
<p>Linear relationships between the red (687 nm): far-red (760 nm) SIF ratio and NPQ for downward (<b>A</b>), upward (<b>B</b>) and total SIF (<b>C</b>). Linear relationships between the red (687 nm): far-red (760 nm) SIF ratio and fPAR for downward (<b>D</b>), upward (<b>E</b>) and total SIF (<b>F</b>). Here, total SIF is calculated as the sum of downward and upward SIF, and fPAR is used to represent the chlorophyll content of leaves.</p>
Full article ">Figure 6
<p>Downward SIF (<math display="inline"><semantics> <mrow> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mi>λ</mi> </msub> <mo>↓</mo> </mrow> </semantics></math>) and upward SIF (<math display="inline"><semantics> <mrow> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mi>λ</mi> </msub> <mo>↑</mo> </mrow> </semantics></math>) ratios for the entire SIF spectrum from 660 nm to 800 nm; colored lines represent the fPAR observed in each leaf. Here, fPAR was used to represent chlorophyll a and b content. The right panel shows the linear relationship between the ratio of downward and upward SIF and fPAR at 687 nm and 760 nm, respectively.</p>
Full article ">Figure 7
<p>From left to right panel, relationship between the SIF yield at 687 nm before correction, SIF yield at 687 nm after correction and SIF yield at 760 nm and NPQ.</p>
Full article ">
19 pages, 4637 KiB  
Article
Exploring the Ability of Solar-Induced Chlorophyll Fluorescence for Drought Monitoring Based on an Intelligent Irrigation Control System
by Wenhui Zhao, Jianjun Wu, Qiu Shen, Jianhua Yang and Xinyi Han
Remote Sens. 2022, 14(23), 6157; https://doi.org/10.3390/rs14236157 - 5 Dec 2022
Cited by 7 | Viewed by 1804
Abstract
Drought is one of the most devastating disasters and a serious constraint on agricultural development. The reflectance-based vegetation indices (VIs), such as Normalized Difference Vegetation Index (NDVI), have been widely used for drought monitoring, but there is a lag in the response of [...] Read more.
Drought is one of the most devastating disasters and a serious constraint on agricultural development. The reflectance-based vegetation indices (VIs), such as Normalized Difference Vegetation Index (NDVI), have been widely used for drought monitoring, but there is a lag in the response of VIs to the changes of photosynthesis induced by drought. Solar-induced chlorophyll fluorescence (SIF) is closely related to photosynthesis of vegetation and can capture changes induced by drought timely. This study investigated the capability of SIF for drought monitoring. An intelligent irrigation control system (IICS) utilizing the Internet of Things was designed and constructed. The soil moisture of the experiment plots was controlled at 60–80% (well-watered, T1), 50–60% (mild water stress, T2), 40–50% (moderate water stress, T3) and 30–40% (severe water stress, T4) of the field water capacity using the IICS based on data collected by soil moisture sensors. Meanwhile, SIF, NDVI, Normalized Difference Red Edge (NDRE) and Optimized Soil Adjusted Vegetation Index (OSAVI) were collected for a long time series using an automated spectral monitoring system. The differences in the responses of SIF, NDVI, NDRE and OSAVI to different drought intensities were fully analyzed. This study illustrates that the IICS can realize precise irrigation management strategies and the construction of regulated deficit irrigation treatments. SIF significantly decreased under mild stress, while NDVI, NDRE and OSAVI only significantly decreased under moderate and severe stress, indicating that SIF is more sensitive to drought. This study demonstrates the excellent ability of SIF for drought monitoring and lays the foundation for the future application of SIF in agricultural drought monitoring. Full article
(This article belongs to the Special Issue Remote Sensing for Agricultural Water Management (RSAWM))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The location of experimental station.</p>
Full article ">Figure 2
<p>Design of intelligent irrigation control system (IICS). The IICS consists of a soil moisture monitoring system and an automatic irrigation system.</p>
Full article ">Figure 3
<p>(<b>a</b>) Soil moisture monitoring system; (<b>b</b>,<b>c</b>) are the components of the soil moisture monitoring system. It mainly includes solar panel, data collector and soil moisture sensors. The sensors were installed in each plot at depths of 10, 20, 50 and 80 cm below the surface.</p>
Full article ">Figure 4
<p>Hardware (<b>a</b>) and software (<b>b</b>) of automatic irrigation system (AIS). Each data collector controls two pots (A and B).</p>
Full article ">Figure 5
<p>The automatic spectral monitoring system. The system consists of QEpro spectrometer, electronic switch, optical fiber, cosine corrector and so on. T1, T2, T3 and T4 represent well-watered, mild water stress, moderate water stress and severe water stress, respectively.</p>
Full article ">Figure 6
<p>Seasonal variation of soil moisture under different water stresses. The soil moisture was represented by the average soil moisture at depths of 10, 20 and 50 cm. DAP means days after planting.</p>
Full article ">Figure 7
<p>Comparison of the average soil moisture under different water stresses. The soil moisture was represented by the average soil moisture at depths of 10, 20 and 50 cm. The average was calculated using the values of soil moisture from 176 DAP to 240 DAP. DAP means days after planting.</p>
Full article ">Figure 8
<p>Seasonal variation of soil moisture at different depths in T1. DAP means days after planting.</p>
Full article ">Figure 9
<p>Seasonal variations of the relative water content (RWC) under different water stresses. Each point and bar indicate the mean value ± standard deviation (SD). DAP means days after planting.</p>
Full article ">Figure 10
<p>Seasonal variations of biomass under different water stresses. Each point and bar indicate the mean value ± standard deviation (SD). DAP means days after planting. Values with different letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 11
<p>The responds of (<b>a</b>) SIF, (<b>b</b>) the NDVI, (<b>c</b>) NDRE, (<b>d</b>) OSAVI and (<b>e</b>) SM under different water stresses. Values with different letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05. The hollow blocks in the figure represent the average values, which were calculated using the data collected from 177 DAP to 223 DAP.</p>
Full article ">Figure 12
<p>The seasonal changes of (<b>a</b>) SIF, (<b>b</b>) NDVI, (<b>c</b>) NDRE and (<b>d</b>) OSAVI under different water stresses. The red, yellow, blue and green solid circles represent well-watered, mild stress, moderate stress and severe stress, respectively. All values are averaged from 9:00 to 16:00. DAP means days after planting.</p>
Full article ">
19 pages, 12043 KiB  
Article
HyScreen: A Ground-Based Imaging System for High-Resolution Red and Far-Red Solar-Induced Chlorophyll Fluorescence
by Huaiyue Peng, Maria Pilar Cendrero-Mateo, Juliane Bendig, Bastian Siegmann, Kelvin Acebron, Caspar Kneer, Kari Kataja, Onno Muller and Uwe Rascher
Sensors 2022, 22(23), 9443; https://doi.org/10.3390/s22239443 - 2 Dec 2022
Cited by 4 | Viewed by 2330
Abstract
Solar-induced chlorophyll fluorescence (SIF) is used as a proxy of photosynthetic efficiency. However, interpreting top-of-canopy (TOC) SIF in relation to photosynthesis remains challenging due to the distortion introduced by the canopy’s structural effects (i.e., fluorescence re-absorption, sunlit-shaded leaves, etc.) and sun–canopy–sensor geometry (i.e., [...] Read more.
Solar-induced chlorophyll fluorescence (SIF) is used as a proxy of photosynthetic efficiency. However, interpreting top-of-canopy (TOC) SIF in relation to photosynthesis remains challenging due to the distortion introduced by the canopy’s structural effects (i.e., fluorescence re-absorption, sunlit-shaded leaves, etc.) and sun–canopy–sensor geometry (i.e., direct radiation infilling). Therefore, ground-based, high-spatial-resolution data sets are needed to characterize the described effects and to be able to downscale TOC SIF to the leafs where the photosynthetic processes are taking place. We herein introduce HyScreen, a ground-based push-broom hyperspectral imaging system designed to measure red (F687) and far-red (F760) SIF and vegetation indices from TOC with single-leaf spatial resolution. This paper presents measurement protocols, the data processing chain and a case study of SIF retrieval. Raw data from two imaging sensors were processed to top-of-canopy radiance by dark-current correction, radiometric calibration, and empirical line correction. In the next step, the improved Fraunhofer line descrimination (iFLD) and spectral-fitting method (SFM) were used for SIF retrieval, and vegetation indices were calculated. With the developed protocol and data processing chain, we estimated a signal-to-noise ratio (SNR) between 50 and 200 from reference panels with reflectance from 5% to 95% and noise equivalent radiance (NER) of 0.04 (5%) to 0.18 (95%) mW m2 sr1 nm1. The results from the case study showed that non-vegetation targets had SIF values close to 0 mW m2 sr1 nm1, whereas vegetation targets had a mean F687 of 1.13 and F760 of 1.96 mW m2 sr1 nm1 from the SFM method. HyScreen showed good performance for SIF retrievals at both F687 and F760; nevertheless, we recommend further adaptations to correct for the effects of noise, varying illumination and sensor optics. In conclusion, due to its high spatial resolution, Hyscreen is a promising tool for investigating the relationship between leafs and TOC SIF as well as their relationship with plants’ photosynthetic capacity. Full article
(This article belongs to the Section Sensing and Imaging)
Show Figures

Figure 1

Figure 1
<p>Overview of the HyScreen system: (<b>a</b>) HyScreen mounted on a movable scaffolding with components noted in numbers. Legend: 1. fluorescence sensor (FLUO), 2. visible and near-infrared sensor (VNIR), 3. power and control units (PCUs) and data acquisition computers (DACs) of the two sensors, 4. displays, and 5. linear axis. (<b>b</b>) HyScreen mounted on the mobile field phenotyping platform.</p>
Full article ">Figure 2
<p>Vegetation and non-vegetation target ROIs used for the HyScreen case study. The 5 and 20% reflectance panels were used for measuring downwelling radiance and empirical line correction. The other targets were used to test the performance of HyScreen: banana leaf, weeping fig leaf, substrate, and brick. The scanning direction along the track was from left-to-right, while spatial pixels across the track are in the vertical direction. The sun zenith angle and azimuth angles were 43.35 and 192.12°, respectively, (yellow arrows). The arrow in the bottom right corner is pointing north.</p>
Full article ">Figure 3
<p>Flowchart of the HyScreen processing chain consisting of the fluorescence sensor (FLUO) and the visible and near-infrared sensor (VNIR) modules divided into four clusters: (I) raw data preparation, (II) from raw data to top-of-canopy downwelling and upwelling radiance and apparent reflectance, (III) vegetation indices and (IV) solar-induced chlorophyll fluorescence (SIF) retrieval. Spatial resolution is represented by full width at half maximum (FWHM).</p>
Full article ">Figure 4
<p>(<b>a</b>) Offset of at-sensor upwelling radiance across the FLUO module’s spectral range calculated with the empirical line method (ELM) described in <a href="#sec2dot4dot2-sensors-22-09443" class="html-sec">Section 2.4.2</a>. Each value corresponds to the intercept from the linear equation fitted to the 5 and 20% reference panels. (<b>b</b>) ratio of offset radiance to downwelling radiance.</p>
Full article ">Figure 5
<p>Signal-to-noise ratio (SNR) and noise-equivalent-radiance (NER) of HyScreen’s FLUO module from 670–780 nm derived from Lambertian reference panels with reflectances of 5% (red), 20% (green), 50% (orange) and 95% (blue): (<b>a</b>) provides information on the SNR, and (<b>b</b>) shows the NERs of four Lambertian reference panels. The solid lines represent mean values of SNR or NER of across-track samples from regions of interest (ROIs), and the light-colored areas illustrate their corresponding standard deviations.</p>
Full article ">Figure 6
<p>Means and standard deviations of top-of-canopy (TOC) upwelling radiance of different vegetation and artificial targets recorded by the HyScreen (<b>a</b>) VNIR and (<b>b</b>) FLUO modules. (<b>a</b>,<b>b</b>) share the same color legend shown in (<b>b</b>). The colored lines represent averaged spectra of all pixels of a target covered by an ROI, while the shaded areas represent corresponding standard derivations.</p>
Full article ">Figure 7
<p>Means and standard deviations of top-of-canopy (TOC) apparent reflectance of different vegetation and artificial targets recorded by the HyScreen (<b>a</b>) VNIR and (<b>b</b>) FLUO modules. The colored lines represent averaged spectra of all pixels of a target covered by an ROI, while the shaded areas represent corresponding standard derivations.</p>
Full article ">Figure 8
<p>Images of vegetation indices and SIFs at 687 nm and 760 nm: (<b>a</b>) true-color composite image of the measured targets; (<b>b</b>) normalized difference vegetation index (NDVI), highlighting the vegetation part; (<b>c</b>) transformed chlorophyll absorption in reflectance index (TCARI), indicating chlorophyll concentration; (<b>d</b>) photochemical reflectance index (PRI) image related to xanthophyll cycle of the NPQ process; (<b>e</b>–<b>h</b>) retrieved SIF images at 760 nm and 687 nm, respectively, from SFM and iFLD. The ROI of the weeping fig leaf is labeled by a red rectangle.</p>
Full article ">Figure A1
<p>Signal-to-noise ratios (SNRs) and noise-equivalent-radiances (NERs) of HyScreen’s fluorescence sensor (FLUO) module from 670–780 nm derived from Lambertian reference panels with 5% and 20% reflectance: (<b>a</b>) provides information on the SNRs, and (<b>b</b>) shows the NERs of two Lambertian reference panels. The solid lines represent mean values of SNRs or NERs of across-track samples, which are from regions of interest (ROIs), and the light-colored areas illustrate their corresponding standard deviations.</p>
Full article ">
23 pages, 2880 KiB  
Article
Different Responses of Solar-Induced Chlorophyll Fluorescence at the Red and Far-Red Bands and Gross Primary Productivity to Air Temperature for Winter Wheat
by Jidai Chen, Xinjie Liu, Guijun Yang, Shaoyu Han, Yan Ma and Liangyun Liu
Remote Sens. 2022, 14(13), 3076; https://doi.org/10.3390/rs14133076 - 26 Jun 2022
Cited by 1 | Viewed by 2178
Abstract
Solar-induced chlorophyll fluorescence (SIF) is closely related to the light-reaction process and has been recognized as a good indicator for tracking gross primary productivity (GPP). Nevertheless, it has not been widely examined how SIF and GPP respond to temperature. Here, we explored the [...] Read more.
Solar-induced chlorophyll fluorescence (SIF) is closely related to the light-reaction process and has been recognized as a good indicator for tracking gross primary productivity (GPP). Nevertheless, it has not been widely examined how SIF and GPP respond to temperature. Here, we explored the linkage mechanisms between SIF and GPP in winter wheat based on continuous measurements of canopy SIF (cSIF), GPP, and meteorological data. To separately explore the structural and physiological mechanisms underlying the SIF–GPP relationship, we studied the temperature responses of the estimated light use efficiency (LUEp), canopy-level chlorophyll fluorescence yield (cSIFyield) and photosystem-level chlorophyll fluorescence yield (ΦF) estimated using canopy-scale remote sensing measurements. We found that GPP, red canopy SIF (cSIF688) and far-red canopy SIF (cSIF760) all exhibited a decreasing trend during overwintering periods. However, GPP and cSIF688 showed relatively more obvious changes in response to air temperature (Ta) than cSIF760 did. In addition, the LUEp responded sensitively to Ta (the correlation coefficient, r = 0.83, p-value < 0.01). The cSIFyield_688 and ΦF_688 (ΦF at 688 nm) also exhibited significantly positive correlations with Ta (r > 0.7, p-value < 0.05), while cSIFyield_760 and ΦF_760 (ΦF at 760 nm) were weakly correlated with Ta (r < 0.3, p-value > 0.05) during overwintering periods. The results also show that LUEp was more sensitive to Ta than ΦF, which caused changes in the LUEp/ΦF ratio in response to Ta. By considering the influence of Ta, the GPP estimation based on the total SIF emitted at the photosystem level (tSIF) was improved (with R2 increased by more than 0.12 for tSIF760 and more than 0.05 for tSIF688). Therefore, our results indicate that the LUEp/ΦF ratio is affected by temperature conditions and highlights that the SIF–GPP model should consider the influence of temperature. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sample reabsorption and scattering inside leaves (<b>a</b>) and canopy (<b>c</b>). PAR is the photosynthetically active radiation. SIF is the solar-induced chlorophyll fluorescence. R represents the reflectance, and T is the transmittance. A represents the absorbed effects and E indicates the canopy scattering effects. Relative PSII or PSI fluorescence represents the relative magnitude of fluorescence emitted from PSII or PSI, respectively (<b>b</b>).</p>
Full article ">Figure 2
<p>Details about the XTS site. (<b>a</b>) Location of the XTS farm; (<b>b</b>) Landsat8 image of the XTS farm (the blue dotted line) on 3 August 2020. The location of the XTS tower is represented by the red rectangle, and the location of the eddy covariance system is represented by the orange circle; (<b>c</b>) the spectral measurement system and the approximate field of view (FOV, 25°) represented by the white ellipse.</p>
Full article ">Figure 3
<p>Seasonal variations in daily mean air temperature (<math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>a</mi> </msub> </mrow> </semantics></math>, (<b>a</b>)) The red dotted line in Panel (<b>a</b>) represents the biological zero. The seasonal changes of the photosynthetically active radiation (<math display="inline"><semantics> <mrow> <mi>P</mi> <mi>A</mi> <mi>R</mi> </mrow> </semantics></math>, (<b>b</b>)). The shadow represents the overwintering period. Local regression was used to smooth the data. The 95 percent confidence interval is indicated by the black shaded area.</p>
Full article ">Figure 4
<p>Seasonal dynamics of daily mean <math display="inline"><semantics> <mrow> <mi>G</mi> <mi>P</mi> <mi>P</mi> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>U</mi> <msub> <mi>E</mi> <mi>p</mi> </msub> </mrow> </semantics></math> (<b>b</b>) and <math display="inline"><semantics> <mrow> <mi>A</mi> <mi>P</mi> <mi>A</mi> <mi>R</mi> </mrow> </semantics></math> (<b>c</b>) are represented in the first row. The second row indicates the changes in canopy <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>d</b>), <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>e</b>), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">Φ</mi> <mrow> <mi>F</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>f</b>). The last row represents the temporal patterns of canopy <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>g</b>), <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>h</b>), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">Φ</mi> <mrow> <mi>F</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>i</b>). The blue shaded rectangle represents the overwintering period. Local regression was used to smooth the data. The black shaded area indicates the 95% confidence interval.</p>
Full article ">Figure 5
<p>Seasonal dynamics of daily mean <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>). The blue shaded rectangle represents the overwintering periods. Local regression was used to smooth the data. The black shaded area indicates the 95% confidence interval.</p>
Full article ">Figure 6
<p>The relationships of daily mean <math display="inline"><semantics> <mrow> <mi>G</mi> <mi>P</mi> <mi>P</mi> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>), <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>), WDRVI (<b>d</b>), <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <msub> <mi>d</mi> <mi>v</mi> </msub> </mrow> </semantics></math> (<b>e</b>), <math display="inline"><semantics> <mrow> <mi>N</mi> <mi>I</mi> <msub> <mi>R</mi> <mi>v</mi> </msub> </mrow> </semantics></math> (<b>f</b>), <math display="inline"><semantics> <mrow> <mi>f</mi> <mi>P</mi> <mi>A</mi> <mi>R</mi> </mrow> </semantics></math> (<b>g</b>), <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>h</b>), and <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>e</mi> <mi>s</mi> <mi>c</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>i</b>) with air temperature (<math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>a</mi> </msub> </mrow> </semantics></math> ). The data were averaged over intervals of 2 °C, and the error bars indicate the standard deviation. The 95% confidence levels for prediction are represented by the grey-shaded zones. * represents a significance level of 0.05 and ** represents a significance level of 0.01.</p>
Full article ">Figure 7
<p>The relationships of daily mean <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>U</mi> <msub> <mi>E</mi> <mi>p</mi> </msub> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>), <math display="inline"><semantics> <mrow> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>), <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">Φ</mi> <mrow> <mi>F</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>d</b>), and <math display="inline"><semantics> <mrow> <msub> <mi mathvariant="sans-serif">Φ</mi> <mrow> <mi>F</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>e</b>) with air temperature (<math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>a</mi> </msub> </mrow> </semantics></math> ). The data were averaged over intervals of 2 °C, and the error bars indicate the standard deviation. The 95% confidence levels for prediction are represented by the grey-shaded zones. * represents a significance level of 0.05 and ** represents a significance level of 0.01.</p>
Full article ">Figure 8
<p>The relationships of daily mean <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>U</mi> <msub> <mi>E</mi> <mi>p</mi> </msub> <mo>/</mo> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>U</mi> <msub> <mi>E</mi> <mi>p</mi> </msub> <mo>/</mo> <mi>c</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>), <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>U</mi> <msub> <mi>E</mi> <mi>p</mi> </msub> <mo>/</mo> <msub> <mi mathvariant="sans-serif">Φ</mi> <mrow> <mi>F</mi> <mo>_</mo> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>), and <math display="inline"><semantics> <mrow> <mi>L</mi> <mi>U</mi> <msub> <mi>E</mi> <mi>p</mi> </msub> <mo>/</mo> <msub> <mi mathvariant="sans-serif">Φ</mi> <mrow> <mi>F</mi> <mo>_</mo> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>d</b>) with air temperature (<math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>a</mi> </msub> </mrow> </semantics></math> ). The data were averaged over intervals of 2 °C, and the error bars indicate the standard deviation. The 95% confidence levels for prediction are represented by the grey-shaded zones. * represents a significance level of 0.05.</p>
Full article ">Figure 9
<p>The relationships of GPP with <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>688</mn> </mrow> </msub> <mo>∗</mo> <mi>f</mi> <mrow> <mo>(</mo> <mrow> <mi>T</mi> <mi>a</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> (<b>b</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>), and <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>760</mn> </mrow> </msub> <mo>∗</mo> <mi>f</mi> <mrow> <mo>(</mo> <mrow> <msub> <mi>T</mi> <mi>a</mi> </msub> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> (<b>d</b>) based on half-hour measurements. The blue dotted line is the 1:1 line. The red line is the fitted line.</p>
Full article ">Figure 10
<p>The relationships of GPP with <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>688</mn> </mrow> </msub> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>688</mn> </mrow> </msub> <mo>∗</mo> <mi>f</mi> <mrow> <mo>(</mo> <mrow> <msub> <mi>T</mi> <mi>a</mi> </msub> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> (<b>b</b>), <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>760</mn> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>), and <math display="inline"><semantics> <mrow> <mi>t</mi> <mi>S</mi> <mi>I</mi> <msub> <mi>F</mi> <mrow> <mn>760</mn> </mrow> </msub> <mo>∗</mo> <mi>f</mi> <mrow> <mo>(</mo> <mrow> <mi>T</mi> <mi>a</mi> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> (<b>d</b>) based on daily mean data. The blue dotted line is the 1:1 line. The red line is the fitted line.</p>
Full article ">
Back to TopTop