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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (369)

Search Parameters:
Keywords = CHM

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3510 KiB  
Article
Transcatheter Embolization of Systemic-to-Pulmonary Collaterals: A New Approach Using Concerto™ Helix Nylon-Fibered Microcoils
by Jochen Pfeifer, Martin Poryo, Anas Gheibeh, Axel Rentzsch and Hashim Abdul-Khaliq
J. Clin. Med. 2025, 14(1), 113; https://doi.org/10.3390/jcm14010113 - 28 Dec 2024
Viewed by 252
Abstract
Background: Systemic-to-pulmonary collaterals (SPCs) are common in congenital heart disease (CHD). Particularly in single ventricle anatomy and Fontan circulation, SPC can both complicate the postoperative course and lead to clinical deterioration in the long term. The treatment of SPC is controversial. The aim [...] Read more.
Background: Systemic-to-pulmonary collaterals (SPCs) are common in congenital heart disease (CHD). Particularly in single ventricle anatomy and Fontan circulation, SPC can both complicate the postoperative course and lead to clinical deterioration in the long term. The treatment of SPC is controversial. The aim of our study was (1) to retrospectively analyse patients who underwent SPC embolization using Concerto™ Helix nylon-fibred microcoils (CHMs) and (2) to describe the interventional technique. Methods: In this single-centre retrospective observational cohort study, we analysed clinical and imaging data of all patients who underwent transcatheter embolization of SPCs using CHMs from January 2016 to December 2023. Results: In 38 consecutive patients (65.8% male, median age 41 months, range 2–490), a total number of 141 CHMs had been implanted into 64 SPCs in 49 procedures. The majority were arterial SPCs (n = 59/64) originating from the thoracic aorta or its branches; 5/64 were veno-venous SPCs. Primary closure succeeded in all procedures. The CHM diameters ranged from 3 to 8 mm, with 5 mm being the most commonly used diameter. The mean coil/SPC ratio was 2.6 (range 1.3–5.3). CHM implantation was performed via four French sheaths. Both detachment and stable positioning were simple and safe. Neither non-target embolization nor coil migration occurred. One complication was a vascular injury with resulting extravasation of contrast medium. In 18/49 procedures (36.7%), coils other than CHMs or vascular plugs were additionally inserted into separate SPCs. Conclusions: CHMs are appropriate for SPC embolization in all age groups, including infants, with a low complication rate. The coils are particularly suitable for the closure of collaterals with a small diameter or tortuous course. They can be used in combination with other embolization devices to achieve comprehensive collateral closure. Full article
Show Figures

Figure 1

Figure 1
<p>Concerto™ Helix nylon-fibered microcoil (source: Concerto Coils Brochure, <a href="http://medtronic.com" target="_blank">medtronic.com</a>, with kind permission).</p>
Full article ">Figure 2
<p>Patients’ ventricular morphology (<b>A</b>) and operations preceding the interventions (<b>B</b>). ASO = atrial switch operation; LV = left ventricle; SCPA = partial cavo–pulmonary anastomosis; RV = right ventricle; TAC = truncus arteriosus communis; TCPA = total cavo–pulmonary anastomosis; TOF = tetralogy of Fallot; * Blalock-Taussig shunt or central aorto–pulmonary shunt.</p>
Full article ">Figure 3
<p>Distribution of the originating vessels of the systemic-to-pulmonary collaterals (SPCs). BV = brachiocephalic vein; LCCA = left common carotid artery; LSA = left subclavian artery; LIMA = left internal mammary artery; RCA = right coronary artery; RSA = right subclavian artery; RIMA = right internal mammary artery; SVC = superior vena cava.</p>
Full article ">Figure 4
<p>Size distribution of the used coils.</p>
Full article ">Figure 5
<p>Angiograms of three different procedures showing the embolization of arterial systemic-to-pulmonary collaterals (SPCs): SPCs originating from the right subclavian artery (<b>A</b>), embolized with three coils (7, 8, and 8 mm) (<b>B</b>); SPCs originating from the right subclavian artery (<b>C</b>), embolized with three coils (4, 4, and 5 mm) (<b>D</b>); and SPCs from the right internal mammary artery (<b>E</b>), embolized with two coils (both 5 mm) (<b>F</b>). Arrows indicate the stretched configuration of the distal coil portion.</p>
Full article ">Figure 6
<p>Angiograms of the embolization of a venous systemic-to-pulmonary collateral originating from the left brachiocephalic vein (<b>A</b>) using three microcoils (4, 7, and 8 mm) (<b>B</b>).</p>
Full article ">Figure 7
<p>Oxygen saturation before and after the embolization of the systemic-to-pulmonary collateral: no significant change.</p>
Full article ">Figure 8
<p>Angiograms of an embolization complicated by a vascular injury: depiction of an arterial systemic-to-pulmonary collateral (SPC) from the left subclavian artery (<b>A</b>); perivascular extravasation (indicated by arrows) following the insertion of the first coil (5 mm) (<b>B</b>); implantation of two more coils (6 and 7 mm) into the residual landing zone with complete closure of the SPC and cessation of extravasation (<b>C</b>). Previously implanted coils can be seen on the left side of the images.</p>
Full article ">
18 pages, 12334 KiB  
Article
Canopy Height Integration for Precise Forest Aboveground Biomass Estimation in Natural Secondary Forests of Northeast China Using Gaofen-7 Stereo Satellite Data
by Caixia Liu, Huabing Huang, Zhiyu Zhang, Wenyi Fan and Di Wu
Remote Sens. 2025, 17(1), 47; https://doi.org/10.3390/rs17010047 - 27 Dec 2024
Viewed by 340
Abstract
Accurate estimates of forest aboveground biomass (AGB) are necessary for the accurate tracking of forest carbon stock. Gaofen-7 (GF-7) is the first civilian sub-meter three-dimensional (3D) mapping satellite from China. It is equipped with a laser altimeter system and a dual-line array stereoscopic [...] Read more.
Accurate estimates of forest aboveground biomass (AGB) are necessary for the accurate tracking of forest carbon stock. Gaofen-7 (GF-7) is the first civilian sub-meter three-dimensional (3D) mapping satellite from China. It is equipped with a laser altimeter system and a dual-line array stereoscopic mapping camera, which enables it to synchronously generate full-waveform LiDAR data and stereoscopic images. The bulk of existing research has examined how accurate GF-7 is for topographic measurements of bare land or canopy height. The measurement of forest aboveground biomass has not received as much attention as it deserves. This study aimed to assess the GF-7 stereo imaging capability, displayed as topographic features for aboveground biomass estimation in forests. The aboveground biomass model was constructed using the random forest machine learning technique, which was accomplished by combining the use of in situ field measurements, pairs of GF-7 stereo images, and the corresponding generated canopy height model (CHM). Findings showed that the biomass estimation model had an accuracy of R2 = 0.76, RMSE = 7.94 t/ha, which was better than the inclusion of forest canopy height (R2 = 0.30, RMSE = 21.02 t/ha). These results show that GF-7 has considerable application potential in gathering large-scale high-precision forest aboveground biomass using a restricted amount of field data. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area (Shangzhi, Heilongjiang, China). (<b>a</b>) The location of the study area; (<b>b</b>) field plots over the GF-7 multispectral image on 20 August 2020 (R—red, G—green, B—blue).</p>
Full article ">Figure 2
<p>The procedure for calculating forest canopy height and biomass from GF-7 stereoscopic imagery.</p>
Full article ">Figure 3
<p>The August and November DSM and CHM. This figure shows only the DSM and CHM for the common regions between August and November, highlighted by the read box. (<b>a</b>,<b>b</b>) DSMs for August and November, respectively, and (<b>e</b>,<b>f</b>) show the larger detail plots in the red boxes. (<b>c</b>,<b>d</b>) CHMs for August and November, respectively, and (<b>g</b>,<b>h</b>) show the larger detail plots in the red boxes.</p>
Full article ">Figure 4
<p>The scatter plot illustrates the relationship between canopy heights predicted using a canopy height model and field-measured heights for two different time points: August and November. The light green points and corresponding regression line represent the August data, while the light blue points and their regression line represent the November data. The 1:1 line (grey dashed) indicates perfect concordance between predicted and measured heights.</p>
Full article ">Figure 5
<p>Feature importance scores for predicting AGB using random forest models under two scenarios: S1 and S2. Features are ranked in decreasing order of importance based on the mean decrease in mean squared error (MSE). The feature “class” refers to land cover classification data, distinguishing between forested and non-forested areas, derived from geographic national condition data.</p>
Full article ">Figure 6
<p>Predicted biomass maps in different scenarios: (<b>a</b>) for S1 scenario and (<b>b</b>) for S2 scenario. Detailed drawings of the red-framed area are shown in <a href="#remotesensing-17-00047-f008" class="html-fig">Figure 8</a>.</p>
Full article ">Figure 7
<p>Scatter plots depicting the relationship between predicted biomass (t/ha) and field-measured biomass (t/ha) for four different scenarios (S1, S2, S3, and S4) in 2020. Detailed scenario descriptions are provided in <a href="#remotesensing-17-00047-t002" class="html-table">Table 2</a>. (<b>a</b>) Scatter plot includes a regression line, with annotations displaying the regression equation, coefficient of determination (R<sup>2</sup>), and root mean square error (RMSE) to quantitatively assess model performance. (<b>b</b>) Residuals for each model’s prediction compared with field biomass. The results demonstrate incremental improvements in biomass prediction accuracy from S1 to S4, highlighting the significant impact of incorporating CHM and DSM data. Scenarios S2 and S3 show enhanced prediction accuracy due to the inclusion of detailed canopy height information.</p>
Full article ">Figure 8
<p>Biomass, spectrum, and canopy height spatial features at the same location. The biomass detail map on the far left shows the result of the S1 scenario, while the one on the right shows the result of the S2 scenario. The biomass ramp is consistent with that shown in <a href="#remotesensing-17-00047-f006" class="html-fig">Figure 6</a>, and the RGB channel denotes the real color channel display of GF-7. The CHM ramp is similar to those shown in <a href="#remotesensing-17-00047-f003" class="html-fig">Figure 3</a>. The last column shows land cover, with black indicating forested area and white representing non-forest land. Figure (<b>a</b>): the disturbance of water bodies and soil moisture on the river valley delta causes the forest vegetation spectra to be mistaken for bare soil and water bodies, which leads to an underestimating of biomass forecast based solely on spectral properties. The regional variability of forest species and height under various topographic circumstances is depicted in Figure (<b>b</b>). Because of the region’s eastern side’s relative flatness, low forest heights, and predominance of coniferous tree species, biomass estimations that consider CHM factors are more accurate in reflecting the real distribution. Figure (<b>c</b>): due to the overestimation of the height of the farmland vegetation caused by spectral features alone (August during the growing season), the farmland’s spectrum is viewed as being spectral like the forest. In agriculture, the average biomass is less than 5 t/ha, although biomass is more precisely anticipated because of the constraint that CHM is approximately 0 t/ha. A logging site is in the region of figure (<b>d</b>), with low biomass. A more accurate prediction of the biomass distribution is made by taking canopy height features into account.</p>
Full article ">
17 pages, 4217 KiB  
Article
Novel Splice-Altering Variants in the CHM and CACNA1F Genes Causative of X-Linked Choroideremia and Cone Dystrophy
by Anna R. Ridgeway, Ciara Shortall, Laura K. Finnegan, Róisín Long, Evan Matthews, Adrian Dockery, Ella Kopčić, Laura Whelan, Claire Kirk, Giuliana Silvestri, Jacqueline Turner, David J. Keegan, Sophia Millington-Ward, Naomi Chadderton, Emma Duignan, Paul F. Kenna and G. Jane Farrar
Genes 2025, 16(1), 25; https://doi.org/10.3390/genes16010025 - 27 Dec 2024
Viewed by 392
Abstract
Background: An estimated 10–15% of all genetic diseases are attributable to variants in noncanonical splice sites, auxiliary splice sites and deep-intronic variants. Most of these unstudied variants are classified as variants of uncertain significance (VUS), which are not clinically actionable. This study investigated [...] Read more.
Background: An estimated 10–15% of all genetic diseases are attributable to variants in noncanonical splice sites, auxiliary splice sites and deep-intronic variants. Most of these unstudied variants are classified as variants of uncertain significance (VUS), which are not clinically actionable. This study investigated two novel splice-altering variants, CHM NM_000390.4:c.941-11T>G and CACNA1F NM_005183.4:c.2576+4_2576+5del implicated in choroideremia and cone dystrophy (COD), respectively, resulting in significant visual loss. Methods: Next-generation sequencing was employed to identify the candidate variants in CHM and CACNA1F, which were confirmed using Sanger sequencing. Cascade analysis was undertaken when additional family members were available. Functional analysis was conducted by cloning genomic regions of interest into gateway expression vectors, creating variant and wildtype midigenes, which were subsequently transfected into HEK293 cells. RNA was harvested and amplified by RT-PCR to investigate the splicing profile for each variant compared to the wildtype. Novel variants were reclassified according to ACMG/AMP and ClinGen SVI guidelines. Results: Midigene functional analysis confirmed that both variants disrupted splicing. The CHM NM_000390.4:c.941-11T>G variant caused exon 8 skipping, leading to a frameshift and the CACNA1F NM_005183.4:c.2576+4_2576+5del variant caused a multimodal splice defect leading to an in-frame insertion of seven amino acids and a frameshift. With this evidence, the former was upgraded to likely pathogenic and the latter to a hot VUS. Conclusions: This study adds to the mutational spectrum of splicing defects implicated in retinal degenerations by identifying and characterising two novel variants in CHM and CACNA1F. Our results highlight the importance of conducting functional analysis to investigate the consequences of intronic splice-altering variants and the significance of reclassifying VUS to confirm a genetic diagnosis. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Pedigree trees for families (<b>A</b>–<b>C</b>). A = family A, B = family B and C = Family C. Affected individuals are shaded black and unaffected individuals are unshaded. Patient IDs are written within circles for females and squares for males. Probands are denoted using a black arrow. Each generation in the pedigree is denoted by I–III.</p>
Full article ">Figure 2
<p>Clinical imaging for Pt-1, Pt-3, Pt-4, and Pt-5 from Family A. (<b>A</b>) Left eye fundus of Pt-1. (<b>B</b>) Left eye fundus autofluorescence of Pt-3. (<b>C</b>) Left eye fundus of Pt-3. (<b>D</b>) Right eye fundus of Pt-4. (<b>E</b>) Left eye fundus of Pt-4. (<b>F</b>) Left eye fundus autofluorescence of Pt-5. (<b>G</b>) Left eye fundus of Pt-5.</p>
Full article ">Figure 3
<p>Functional analysis of the <span class="html-italic">CHM</span> NM_000390.4:c.941-11T&gt;G variant (V1) compared to wildtype (WT). (<b>A</b>) Illustrates the <span class="html-italic">CHM</span> gene in its antisense orientation with black arrows representing the primers used to amplify the region of interest. (<b>B</b>) Scores from <span class="html-italic">in silico</span> splice prediction tools for WT compared to V1. Green triangles represent splice acceptor sites. The green letter t represents WT and red g represents the c.941-11T&gt;G variant. (<b>C</b>) Schematic of the expression clone including exons 6–8 of the <span class="html-italic">CHM</span> gene transfected into HEK293 cells. (<b>D</b>) Agarose gel illustrating the WT (band 1 of 682bp), V1 (band 2 of 456bp), <span class="html-italic">RHO</span> exon 5 control for WT, V1 and HEK cell only control and β-actin control for WT, V1 and HEK cell only control. (<b>E.1</b>) Sanger sequence chromatogram from the WT purified gel product in part D band 1. (<b>E.2</b>). V1 Sanger sequence chromatogram from the V1 purified gel product in part D band 2 with the altered amino acid sequence in red text and red arrow signifying the point at which the nucleotide sequence is altered. (<b>E.3</b>). Illustration of the protein product that would result from skipping of exon 8 denoted by the altered amino acid sequence in red text.</p>
Full article ">Figure 4
<p>AlphaFold protein models and InterPro domains. (<b>A</b>) AlphaFold protein folding predictions for wildtype (WT) compared to the amino acid sequence produced as a result of the <span class="html-italic">CHM</span> c.941-11T&gt;G, p.Tyr315CysfsTer18 variant. The Tyr315 residue is green and Cys315 residue is red. Hydrogen bonds are denoted by the orange dashed line. (<b>B</b>) InterPro domains for wildtype compared to the amino acid sequence produced as a result of the <span class="html-italic">CHM</span> c.941-11T&gt;G, p.Tyr315CysfsTer18 variant.</p>
Full article ">Figure 5
<p>Functional analysis of the <span class="html-italic">CACNA1F</span> c.2576+4_2576+5del variant (V2) compared to wildtype (WT). (<b>A</b>) Illustrates the <span class="html-italic">CACNA1F</span> gene in its antisense orientation with black arrows representing the primers used to amplify the region of interest. (<b>B</b>) Scores from <span class="html-italic">in silico</span> splice prediction tools for WT compared to V2. Blue triangles represent splice donor sites. The green letters ag represent the WT nucleotide sequence and the red line represents the V2 nucleotide sequence as a result of the c.2576+4_2576+5del variant. (<b>C</b>) Schematic of the expression clone including exons 18–26 of the <span class="html-italic">CACNA1F</span> gene which was transfected into HEK293 cells. (<b>D</b>) Agarose gel illustrating the WT (band 1 of 418 bp), V2 (band 2 of 439 bp), V2 (band 3 of 365 bp), <span class="html-italic">RHO</span> exon 5 control for WT, V2 and HEK cell only control and β-actin control for WT, V2 and HEK cell only control. (<b>E</b>) Sanger sequence chromatograms from the V2 band 2 and V2 band 3 purified gel products. The altered amino acid sequence as a result of the c.2576+4_2575+5del variant is illustrated by the red text. The red arrow signifies the point at which the nucleotide sequence is altered.</p>
Full article ">Figure 6
<p>AlphaFold protein models and InterPro domains. (<b>A</b>) AlphaFold protein folding predictions for wildtype (WT) compared to the amino acid sequence produced as a result of the <span class="html-italic">CACNA1F</span> c.2576+4_2576+5del,p.Pro859_Leu860insCysAlaGlySerGlyArgGly or p.Val842AlafsTer31 protein change. The Pro859 residue is coloured green, Pro859_Leu860insCysAlaGlySerGlyArgGly residues are coloured purple and Ala842 residue is coloured white. The white arrows represent the region of the protein altered as a result of the protein change. (<b>B</b>) InterPro domains for wildtype compared to the amino acid sequences produced as a result of the <span class="html-italic">CACNA1F</span> c.2576+4_2576+5del,p.Pro859_Leu860insCysAlaGlySerGlyArgGly/p.Val842AlafsTer31 protein change.</p>
Full article ">
17 pages, 11444 KiB  
Article
Oxidative Stress, Inflammation and Altered Glucose Metabolism Contribute to the Retinal Phenotype in the Choroideremia Zebrafish
by Cécile Méjécase, Neelima Nair, Hajrah Sarkar, Pablo Soro-Barrio, Maria Toms, Sophia Halliday, Katy Linkens, Natalia Jaroszynska, Constance Maurer, Nicholas Owen and Mariya Moosajee
Antioxidants 2024, 13(12), 1587; https://doi.org/10.3390/antiox13121587 - 23 Dec 2024
Viewed by 413
Abstract
Reactive oxygen species (ROS) within the retina play a key role in maintaining function and cell survival. However, excessive ROS can lead to oxidative stress, inducing dysregulation of metabolic and inflammatory pathways. The chmru848 zebrafish models choroideremia (CHM), an X-linked chorioretinal dystrophy, [...] Read more.
Reactive oxygen species (ROS) within the retina play a key role in maintaining function and cell survival. However, excessive ROS can lead to oxidative stress, inducing dysregulation of metabolic and inflammatory pathways. The chmru848 zebrafish models choroideremia (CHM), an X-linked chorioretinal dystrophy, which predominantly affects the photoreceptors, retinal pigment epithelium (RPE), and choroid. In this study, we examined the transcriptomic signature of the chmru848 zebrafish retina to reveal the upregulation of cytokine pathways and glia migration, upregulation of oxidative, ER stress and apoptosis markers, and the dysregulation of glucose metabolism with the downregulation of glycolysis and the upregulation of the oxidative phase of the pentose phosphate pathway. Glucose uptake was impaired in the chmru848 retina using the 2-NBDG glucose uptake assay. Following the overexpression of human PFKM, partial rescue was seen with the preservation of photoreceptors and RPE and increased glucose uptake, but without modifying glycolysis and oxidative stress markers. Therapies targeting glucose metabolism in CHM may represent a potential remedial approach. Full article
(This article belongs to the Special Issue Antioxidants and Retinal Diseases—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Cytokine pathway is upregulated in <span class="html-italic">chm<sup>ru848</sup></span> zebrafish retina. <span class="html-italic">chm<sup>ru848</sup></span> zebrafish samples are indicated in magenta and wt samples are indicated in cyan.</p>
Full article ">Figure 2
<p>Oxidative (<b>a</b>), ER stress (<b>b</b>) and apoptosis (<b>c</b>) markers are upregulated in the <span class="html-italic">chm<sup>ru848</sup></span> zebrafish retina. <span class="html-italic">chm<sup>ru848</sup></span> zebrafish are indicated in magenta and wt samples are indicated in cyan.</p>
Full article ">Figure 3
<p>Glycolysis is downregulated in favour of the pentose phosphate pathway in <span class="html-italic">chm<sup>ru848</sup></span> zebrafish retina. (<b>a</b>,<b>b</b>) Glycolysis pathways are affected in <span class="html-italic">chm<sup>ru848</sup></span> retina. (<b>a</b>) <span class="html-italic">chm<sup>ru848</sup></span> zebrafish are indicated in magenta and wt samples are indicated in cyan. (<b>b</b>) Scheme adapted from <a href="https://www.wikipathways.org/instance/WP1356" target="_blank">https://www.wikipathways.org/instance/WP1356</a> (accessed on 18 December 2024) [<a href="#B27-antioxidants-13-01587" class="html-bibr">27</a>,<a href="#B28-antioxidants-13-01587" class="html-bibr">28</a>]. Significantly downregulated genes are indicated in blue, while upregulated genes are in red. (<b>c</b>) Pentose phosphate pathway is upregulated in the <span class="html-italic">chm<sup>ru848</sup></span> retina compared to wt.</p>
Full article ">Figure 4
<p><span class="html-italic">PFKM</span> overexpression improves retinal phenotype but not photoreceptor cell death at 5 dpf. (<b>a</b>) Human <span class="html-italic">PFKM</span> expression was detected 5 days post-injection in <span class="html-italic">chm<sup>ru848</sup></span> and wt zebrafish. (<b>b</b>) mRNA-injected <span class="html-italic">chm<sup>ru848</sup></span> showed a more preserved photoreceptor layer (white asterisk) with better retinal lamination and a thicker RPE layer compared to the uninjected mutants (<span class="html-italic">n</span> = 9 zebrafish per group). Scale bar: 50 μm; scale bar zoom in: 20 μm. (<b>c</b>) Cell viability is unchanged after <span class="html-italic">PFKM</span> overexpression in wt and <span class="html-italic">chm<sup>ru848</sup></span> zebrafish (n = 9 zebrafish per group). Scale bar = 50 μm. ONL: outer nuclear layer; INL: inner nuclear layer. (<b>d</b>,<b>e</b>) Expression of anti-apoptotic (<b>d</b>) and pro-apoptotic (<b>e</b>) markers were analysed using RT-qPCR at 5 dpf in zebrafish eyes, from uninjected and injected wt and <span class="html-italic">chm<sup>ru848</sup></span>. Data are expressed as mean ± SEM from n = 3 (with 20 eyes per group). Statistical significance was determined by one-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005; **** <span class="html-italic">p</span> &lt; 0.0001. “●”—wt; “▲”—wt + <span class="html-italic">PFKM</span>; “▼”—<span class="html-italic">chm<sup>ru848</sup></span>; “◆”—<span class="html-italic">chm<sup>ru848</sup></span> + <span class="html-italic">PFKM</span>.</p>
Full article ">Figure 5
<p><span class="html-italic">PFKM</span> overexpression rescues 2-NBDG uptake in photoreceptors in <span class="html-italic">chm<sup>ru848</sup></span> retina at 5 dpf. Scale bar = 50 μm. 2-NBDG is in green; nucleus stained with DAPI (blue). * <span class="html-italic">p</span> &lt; 0.05; **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p><span class="html-italic">PFKM</span> overexpression does not improve activity of the glycolysis pathway or oxidative stress 5 dpf in <span class="html-italic">chm<sup>ru848</sup></span> zebrafish. Expression of genes involved in the glycolysis pathway (<b>a</b>), in the pentose phosphate oxidative branch (<b>b</b>) and in oxidative stress (<b>c</b>) were analysed using RT-qPCR at 5dpf in zebrafish eyes, from uninjected and injected wt and <span class="html-italic">chm<sup>ru848</sup></span>. Data are expressed as mean ± SEM from n = 3 (with 20 eyes per group). Statistical significance was determined by One-way ANOVA. * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.005; *** <span class="html-italic">p</span> &lt; 0.001; **** <span class="html-italic">p</span> &lt; 0.0001. “●”—wt; “▲”—wt + <span class="html-italic">PFKM</span>; “▼”—<span class="html-italic">chm<sup>ru848</sup></span>; “◆”—<span class="html-italic">chm<sup>ru848</sup></span> + <span class="html-italic">PFKM</span>.</p>
Full article ">
24 pages, 3802 KiB  
Article
Performance of Individual Tree Segmentation Algorithms in Forest Ecosystems Using UAV LiDAR Data
by Javier Marcello, María Spínola, Laia Albors, Ferran Marqués, Dionisio Rodríguez-Esparragón and Francisco Eugenio
Drones 2024, 8(12), 772; https://doi.org/10.3390/drones8120772 - 19 Dec 2024
Viewed by 646
Abstract
Forests are crucial for biodiversity, climate regulation, and hydrological cycles, requiring sustainable management due to threats like deforestation and climate change. Traditional forest monitoring methods are labor-intensive and limited, whereas UAV LiDAR offers detailed three-dimensional data on forest structure and extensive coverage. This [...] Read more.
Forests are crucial for biodiversity, climate regulation, and hydrological cycles, requiring sustainable management due to threats like deforestation and climate change. Traditional forest monitoring methods are labor-intensive and limited, whereas UAV LiDAR offers detailed three-dimensional data on forest structure and extensive coverage. This study primarily assesses individual tree segmentation algorithms in two forest ecosystems with different levels of complexity using high-density LiDAR data captured by the Zenmuse L1 sensor on a DJI Matrice 300RTK platform. The processing methodology for LiDAR data includes preliminary preprocessing steps to create Digital Elevation Models, Digital Surface Models, and Canopy Height Models. A comprehensive evaluation of the most effective techniques for classifying ground points in the LiDAR point cloud and deriving accurate models was performed, concluding that the Triangular Irregular Network method is a suitable choice. Subsequently, the segmentation step is applied to enable the analysis of forests at the individual tree level. Segmentation is crucial for monitoring forest health, estimating biomass, and understanding species composition and diversity. However, the selection of the most appropriate segmentation technique remains a hot research topic with a lack of consensus on the optimal approach and metrics to be employed. Therefore, after the review of the state of the art, a comparative assessment of four common segmentation algorithms (Dalponte2016, Silva2016, Watershed, and Li2012) was conducted. Results demonstrated that the Li2012 algorithm, applied to the normalized 3D point cloud, achieved the best performance with an F1-score of 91% and an IoU of 83%. Full article
(This article belongs to the Section Drones in Agriculture and Forestry)
Show Figures

Figure 1

Figure 1
<p>Map of the protected natural areas of the Canary Islands and photographs of the National Parks of Caldera de Taburiente in La Palma (<b>top left</b>) and Garajonay in La Gomera (<b>bottom left</b>).</p>
Full article ">Figure 2
<p>Vegetation, true color imagery, and ground truth data for the parks of (<b>a</b>) Caldera de Taburiente and (<b>b</b>) Garajonay.</p>
Full article ">Figure 3
<p>Simplified methodology of the individual tree detection and crown delineation using LiDAR data for the extraction of forest parameters.</p>
Full article ">Figure 4
<p>Selected algorithms to compare the performance of individual tree detection and crown delineation methods (Software implementation is represented by red, blue, and yellow colors and discussed in <a href="#sec3-drones-08-00772" class="html-sec">Section 3</a>).</p>
Full article ">Figure 5
<p>Color composite and LiDAR data (3D cloud and examples of horizontal and vertical profiles) of: (<b>a</b>) Caldera de Taburiente and (<b>b</b>) Garajonay.</p>
Full article ">Figure 6
<p>DEMs generated by the different combinations of ground classification and interpolation algorithms: (<b>a</b>) DEMs and (<b>b</b>) Error Maps (green colors refer to lower errors, while red colors refer to higher errors).</p>
Full article ">Figure 7
<p>LiDAR models for Taburiente and Garajonay: (<b>a</b>) DEM, (<b>b</b>) CHM and (<b>c</b>) Normalized point cloud.</p>
Full article ">Figure 8
<p>Performance evaluation of tree detection algorithms: (<b>a</b>) LMF-LidR, (<b>b</b>) LMF-LIDAR360 and (<b>c</b>) CF.</p>
Full article ">Figure 9
<p>Reference seeds (red stars) with respect to the seeds detected by the algorithms (black dots): (<b>a</b>) LMF-LidR (ws = 6), (<b>b</b>) LMF-LIDAR360 (<span class="html-italic">σ</span> = 7) and (<b>c</b>) CF (C = 2).</p>
Full article ">Figure 10
<p>Segmentation results: (<b>a</b>) Reference segmentation (color fill) with respect to the segmentation algorithms (vector overlay), (<b>b</b>) precision, recall, and F1-score and (<b>c</b>) IoU.</p>
Full article ">Figure 11
<p>Individual tree segmentation for Caldera de Taburiente (<b>left</b>/<b>top</b>) and Garajonay (<b>right</b>/<b>bottom</b>) parks.</p>
Full article ">Figure 12
<p>Forest metrics of Taburiente: (<b>a</b>) height, (<b>b</b>) area, and (<b>c</b>) volume.</p>
Full article ">Figure 13
<p>Vertical profiles at the Garajonay National Park.</p>
Full article ">
8 pages, 2806 KiB  
Proceeding Paper
Constructing Rasterized Covariates from LiDAR Point Cloud Data via Structured Query Language
by Rory Pittman and Baoxin Hu
Proceedings 2024, 110(1), 1; https://doi.org/10.3390/proceedings2024110001 - 3 Dec 2024
Viewed by 379
Abstract
For point cloud data compiled over larger spatial domains, the rasterization of features is effectively streamlined by means of structured query language (SQL). This comprises enhanced control with filtering data and implementing specific metrics for summarization to derive environmental covariates. LiDAR (light detection [...] Read more.
For point cloud data compiled over larger spatial domains, the rasterization of features is effectively streamlined by means of structured query language (SQL). This comprises enhanced control with filtering data and implementing specific metrics for summarization to derive environmental covariates. LiDAR (light detection and ranging) point cloud data were analyzed via SQL to generate rasterized covariates of the digital terrain model (DTM), canopy height model (CHM), and a gap fraction for a boreal study region in Northern Ontario, Canada. These features, along with topographic covariates computed from the DTM, were later ascertained as important for subsequent tree species classification research. Full article
(This article belongs to the Proceedings of The 31st International Conference on Geoinformatics)
Show Figures

Figure 1

Figure 1
<p>SQL for creating a DSM, DTM, and CHM.</p>
Full article ">Figure 2
<p>SQL to compute a gap fraction, as well as a summary table.</p>
Full article ">Figure 3
<p>Digital terrain models (DTMs), both of 30 m spatial resolution, for a boreal study area centered around Cochrane, Ontario, Canada: (<b>a</b>) LiDAR-derived DTM; (<b>b</b>) Provincial DTM attained from Ontario MNR. Grid coordinates are in the NAD UTM 1983 Zone 17 N projection.</p>
Full article ">Figure 4
<p>LiDAR-derived covariates, both of 30 m spatial resolution, for a boreal study area centered around Cochrane, Ontario, Canada: (<b>a</b>) Digital surface model (DSM); (<b>b</b>) Canopy height model (CHM). Grid coordinates are in the NAD UTM 1983 Zone 17 N projection.</p>
Full article ">Figure 5
<p>LiDAR-derived gap fraction for a boreal study area centered around Cochrane, Ontario, Canada. Grid coordinates are in the NAD UTM 1983 Zone 17 N projection.</p>
Full article ">
19 pages, 6073 KiB  
Article
Effective UAV Photogrammetry for Forest Management: New Insights on Side Overlap and Flight Parameters
by Atman Dhruva, Robin J. L. Hartley, Todd A. N. Redpath, Honey Jane C. Estarija, David Cajes and Peter D. Massam
Forests 2024, 15(12), 2135; https://doi.org/10.3390/f15122135 - 2 Dec 2024
Viewed by 1090
Abstract
Silvicultural operations such as planting, pruning, and thinning are vital for the forest value chain, requiring efficient monitoring to prevent value loss. While effective, traditional field plots are time-consuming, costly, spatially limited, and rely on assumptions that they adequately represent a wider area. [...] Read more.
Silvicultural operations such as planting, pruning, and thinning are vital for the forest value chain, requiring efficient monitoring to prevent value loss. While effective, traditional field plots are time-consuming, costly, spatially limited, and rely on assumptions that they adequately represent a wider area. Alternatively, unmanned aerial vehicles (UAVs) can cover large areas while keeping operators safe from hazards including steep terrain. Despite their utility, optimal flight parameters to ensure flight efficiency and data quality remain under-researched. This study evaluated the impact of forward and side overlap and flight altitude on the quality of two- and three-dimensional spatial data products from UAV photogrammetry (UAV-SfM) for assessing stand density in a recently thinned Pinus radiata D. Don plantation. A contemporaneously acquired UAV laser scanner (ULS) point cloud provided reference data. The results indicate that the optimal UAV-SfM flight parameters are 90% forward and 85% side overlap at a 120 m altitude. Flights at an 80 m altitude offered marginal resolution improvement (2.2 cm compared to 3.2 cm ground sample distance/GSD) but took longer and were more error-prone. Individual tree detection (ITD) for stand density assessment was then applied to both UAV-SfM and ULS canopy height models (CHMs). Manual cleaning of the detected ULS tree peaks provided ground truth for both methods. UAV-SfM had a lower recall (0.85 vs. 0.94) but a higher precision (0.97 vs. 0.95) compared to ULS. Overall, the F-score indicated no significant difference between a prosumer-grade photogrammetric UAV and an industrial-grade ULS for stand density assessments, demonstrating the efficacy of affordable, off-the-shelf UAV technology for forest managers. Furthermore, in addressing the knowledge gap regarding optimal UAV flight parameters for conducting operational forestry assessments, this study provides valuable insights into the importance of side overlap for orthomosaic quality in forest environments. Full article
(This article belongs to the Special Issue Image Processing for Forest Characterization)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study site within NZ (insert), with the optimal orthomosaic overlaid on a mixed topographic/aerial image of the region. The stand boundary is indicated by the purple line, GCP locations are in orange, and the take-off location is in cyan.</p>
Full article ">Figure 2
<p>Image of a GCP coated with reflective material and painted in a high contrast pattern for identification in the ULS and UAV-SfM datasets.</p>
Full article ">Figure 3
<p>Images of the UAVs utilised in this study: (<b>a</b>) the DJI Phantom 4 Pro; (<b>b</b>) the DJI Matrice 300 RTK with a DJI L1 sensor. (<b>c</b>) Shows the flight crew operating UAVs from an MEWP to maintain VLOS.</p>
Full article ">Figure 4
<p>Examples of image artefacts encountered when annotating the orthomosaics: (<b>a</b>,<b>d</b>,<b>e</b>) “blurring” and “smudging” effects; (<b>b</b>,<b>d</b>) “tearing” or “breaking” discontinuities within the image; (<b>c</b>,<b>e</b>) “ghosting”, in which the displacement of ground and canopy pixels results in transparent tree canopies; (<b>d</b>,<b>f</b>) in some areas tree canopies were fragmented into smaller chunks.</p>
Full article ">Figure 5
<p>Bar graph plotting the percentage area of each orthomosaic free from artefacts for each flight plan, coloured by altitude. The missions are arranged by overlap with the lowest on the left and highest on the right.</p>
Full article ">Figure 6
<p>Comparison of flight time duration between different overlap flight missions, coloured by altitude.</p>
Full article ">Figure 7
<p>Correlations between (<b>a</b>) forward overlap and (<b>b</b>) side overlap with the area of each orthomosaic that was clear of artefacts. The teal line represents the linear model between variables, and point locations are jittered so that multiple points with the same value are visible.</p>
Full article ">Figure 8
<p>Calculated relief displacement for an object (e.g., a tree) of height (<span class="html-italic">h</span>) 20 m, within imagery captured at a flying height (<span class="html-italic">H</span>) of 80 m (purple vectors) or 120 m (yellow vectors) above ground level. The displacement vectors in the legend are scaled to an image distance of 100 mm. Relief displacement vectors are plotted on an image captured at ~80 m AGL and at image radial distances of 0, 100, 200, 300, 400, 500, 650, and 900 mm from the principal point, which for this illustration is assumed to coincide with the image centre.</p>
Full article ">Figure 9
<p>Demonstration of the greater impact of side (dashed purple line) than forward (dotted green line) overlap on movement of the near-nadir viewing region of an image (solid teal line). Movement values are based on P4 Pro camera at a height of 120 m, with an image footprint of 180 × 120 m.</p>
Full article ">Figure 10
<p>The effects of shadow and occlusions in the orthomosaics produced by flights flown at overlap of 90:85 at 80 m (<b>a</b>) and 90:90 at 80 m (<b>b</b>).</p>
Full article ">
10 pages, 6891 KiB  
Article
Oxidative Addition to Group 1 (K, Rb, Cs) Alumanyl Anions as a Route to o-Carboranyl (hydrido)aluminates
by Han-Ying Liu, Kyle G. Pearce, Michael S. Hill and Mary F. Mahon
Inorganics 2024, 12(12), 309; https://doi.org/10.3390/inorganics12120309 - 27 Nov 2024
Viewed by 673
Abstract
The kinetic stability provided by the sterically demanding {SiNDipp}2− dianion (SiNDipp = {CH2SiMe2NDipp}2; Dipp = 2,6-i-Pr2C6H3) is intrinsic to the isolation of not only the [...] Read more.
The kinetic stability provided by the sterically demanding {SiNDipp}2− dianion (SiNDipp = {CH2SiMe2NDipp}2; Dipp = 2,6-i-Pr2C6H3) is intrinsic to the isolation of not only the group 1 alumanyl reagents ([{SiNDipp}AlM]2; M = K, Rb, Cs) but also facilitates the completely selective oxidative addition of a C-H bond of 1,2-C2B10H12 to the aluminium centre. In each case, the resultant compounds comprise a four-coordinate o-carboranyl (hydrido)aluminate anion, [(SiNDipp)Al(H)(1,2-C2B10H11)], in which the carboranyl cage is bonded to aluminium by an Al-C σ bond. Although the anions further assemble as extended network structures based on Al-H∙∙∙M, B-H∙∙∙M, and C-H∙∙∙M interactions, each structure is unique due to the significant variation in M+ ionic radius as group 1 is descended. The potassium derivative crystallises as a one-dimensional polymer, its rubidium analogue is a dimer due to the polyhapto-sequestration of a molecule of benzene solvent within the alkali metal coordination sphere, and the caesium species is a two-dimensional assembly of hexameric aggregates. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Synthesis of [(BDI)Ae(<span class="html-italic">o</span>-C<sub>2</sub>B<sub>10</sub>H<sub>11</sub>)] (<b>I</b> and <b>II</b>) through the deprotonation of 1,2-C<sub>2</sub>B<sub>10</sub>H<sub>12</sub>; (<b>b</b>) the structure of compound <b>III</b>; (<b>c</b>) the structures of <b>IV<sup>M</sup></b> and their reactivity with terminal alkynes.</p>
Full article ">Figure 2
<p>Plot depicting the polymeric structure of <b>1</b>. Ellipsoids are shown at 30% probability. A molecule of benzene solvent and hydrogen atoms (H1 and those attached to C32, B2, and B3 excepted) have been omitted for clarity. Peripheral substituents are depicted as wireframes, also for visual ease. Selected bond lengths (Å) and angles (°): Al1-N1 1.8905(11), Al1-N2 1.8776(11), Al1-C31 2.0804(14), C31-C32 1.6530(18), C31-B1 1.733(2), C31-B2 1.7090(19), C31-B3 1.7260(19), C31-B4 1.744(2), C32-B1 1.691(2), C32-B4 1.696(2), C32-B8 1.696(2), C32-B9 1.703(2), N2-Al1-N1 113.32(5), N1-Al1-C31 114.25(5), and N2-Al1-C31 111.13(5). Symmetry operations: <sup>1</sup> <span class="html-italic">x</span>, 1 − <span class="html-italic">y</span>, ½ + <span class="html-italic">z</span>; <sup>2</sup> <span class="html-italic">x</span>, 1 − <span class="html-italic">y</span>, −½ + <span class="html-italic">z</span>.</p>
Full article ">Figure 3
<p>Plot depicting the dimeric structure of <b>2</b>. Ellipsoids are shown at 30% probability. Hydrogen atoms (H1 and those attached to C32, B4, B9, and B10 excepted) have been omitted for clarity. Some peripheral substituents are depicted as wireframes, also for visual ease. Selected bond lengths (Å) and angles (°): Al1-N1 1.8928(15), Al1-N2 1.8673(14), Al1-C31 2.1018(16), C31-C32 1.730(2), C31-B1 1.742(2), C31-B2 1.661(2), C31-B3 1.728(3), C31-B4 1.712(2), C32-B1 1.752(3), C32-B4 1.756(3), C32-B9 1.778(3), N2-Al1-N1 113.60(6), N1-Al1-C31 113.56(7), and N2-Al1-C31 110.88(6). Symmetry operations to generate equivalent atoms: <sup>1</sup> 1 − <span class="html-italic">x</span>, 1 − <span class="html-italic">y</span>, 1 − <span class="html-italic">z</span>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Plot depicting the structure of <b>3</b>. Ellipsoids are shown at 30% probability. Occluded benzene solvent, disordered atoms, and hydrogen atoms (H1 and those attached to C32 and selected B atoms excepted) have been omitted for clarity. Some peripheral substituents are depicted as wireframes, also for visual ease. Symmetry operations to generate equivalent atoms: <sup>1</sup> 2 − <span class="html-italic">x</span>, 2 − <span class="html-italic">y</span>, 2 − <span class="html-italic">z</span>. (<b>b</b>) Plot depicting a section of the hexacyclic motifs that dominate the two-dimensional sheets present in 3. Hydrogen atoms, except those attached to the carboranyl cage, have been removed and iso-propyl substituents are shown as wireframes for visual ease. Colour scheme adopted as for (<b>a</b>). Selected bond lengths (Å) and angles (°): Al1-N1 1.888(3), Al1-N2 1.867(3), Al1-C31 2.127(3), C31-C32 1.713(5), C31-B1 1.733(5), C31-B4 1.733(4), C31-B5 1.726(5), C32-B1 1.759(5), C32-B4 1.756(3), N2-Al1-N1 114.32(12), N1-Al1-C31 115.10(12), and N2-Al1-C31 110.19(12).</p>
Full article ">Scheme 1
<p>Synthesis of compounds <b>1</b>–<b>3</b>.</p>
Full article ">
14 pages, 4866 KiB  
Article
Retinal Patterns and the Role of Autofluorescence in Choroideremia
by Federica E. Poli, Robert E. MacLaren and Jasmina Cehajic-Kapetanovic
Genes 2024, 15(11), 1471; https://doi.org/10.3390/genes15111471 - 14 Nov 2024
Viewed by 668
Abstract
Background: Choroideremia is a monogenic inherited retinal dystrophy that manifests in males with night blindness, progressive loss of peripheral vision, and ultimately profound sight loss, commonly by middle age. It is caused by genetic defects of the CHM gene, which result in a [...] Read more.
Background: Choroideremia is a monogenic inherited retinal dystrophy that manifests in males with night blindness, progressive loss of peripheral vision, and ultimately profound sight loss, commonly by middle age. It is caused by genetic defects of the CHM gene, which result in a deficiency in Rab-escort protein-1, a key element for intracellular trafficking of vesicles, including those carrying melanin. As choroideremia primarily affects the retinal pigment epithelium, fundus autofluorescence, which focuses on the fluorescent properties of pigments within the retina, is an established imaging modality used for the assessment and monitoring of affected patients. Methods and Results: In this manuscript, we demonstrate the use of both short-wavelength blue and near-infrared autofluorescence and how these imaging modalities reveal distinct disease patterns in choroideremia. In addition, we show how these structural measurements relate to retinal functional measures, namely microperimetry, and discuss the potential role of these retinal imaging modalities in clinical practice and research studies. Moreover, we discuss the mechanisms underlying retinal autofluorescence patterns by imaging with a particular focus on melanin pigment. Conclusions: This could be of particular significance given the current progress in therapeutic options, including gene replacement therapy. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Multimodal retinal imaging and retinal sensitivity testing of the left eye in a patient with choroideremia. Panel (<b>A</b>): short-wavelength blue autofluorescence (BAF) imaging demonstrating a smooth zone centrally around the fovea (outlined in red), a mottled zone eccentric to this (arrow), and an atrophic area peripherally (asterisk). Panel (<b>B</b>): Near-infrared autofluorescence (NIR-AF) imaging. The red outline demonstrates the correspondence between the smooth zone on BAF and the area of homogenous NIR-AF signal. Panel (<b>C</b>): microperimetry map taken with the MAIA microperimeter (CenterVue, Padova, Italy). This shows better function closer to the fovea (green points), impaired but present function in the area corresponding to mottled BAF (yellow-orange points), and no function outside the island of BAF (black points). Panel (<b>D</b>): Colour fundus photograph. Panel (<b>E</b>): trans-foveal optical computed tomography (OCT) image showing preserved ellipsoid zone within the area of smooth BAF/homogenous NIR-AF (between the two red markers), disrupted ellipsoid zone in the area of mottled BAF/absent NIR-AF (between the red and green markers), and absent ellipsoid zone with retinal and choroidal degeneration in the peripheries (outside the green markers). BAF, NIR-AF and OCT images were taken with Heidelberg Spectralis, Heidelberg Engineering GmbH, Heidelberg, Germany.</p>
Full article ">Figure 2
<p>Multimodal retinal imaging and retinal sensitivity testing in two female carriers. Case 1: 20-year-old asymptomatic female carrier, VA 6/5 (colour fundus photograph, (<b>D</b>)). Case 2: 44-year-old affected female carrier with male pattern choroideremia, VA 6/9 (colour fundus photograph, (<b>I</b>)). Fundus autofluorescence imaging demonstrates early ‘salt and pepper’ mottled appearance (BAF) (<b>A</b>) and normal smooth zone (NIR-AF) (<b>B</b>) in Case 1, compared with more advanced mottling and areas of atrophy (BAF) (<b>F</b>) and near-absent autofluorescence smooth zone (NIR-AF) (<b>G</b>) in Case 2. Microperimetry maps (MAIA microperimeter, CenterVue, Padova, Italy) show near normal sensitivity in the asymptomatic carrier (<b>C</b>) but reduced sensitivity over the mottled zone with absent sensitivity over the atrophic areas in the affected carrier typical of male pattern choroideremia (<b>H</b>). OCT imaging shows an intact ellipsoid zone in Case 1 (<b>E</b>), whilst disruption of the ellipsoid zone is visible in Case 2 (<b>J</b>).</p>
Full article ">Figure 3
<p>A schematic representation of the Rab prenylation pathway and effect on RPE melanin. REP1 (Rab Escort Protein 1) binds the newly synthesized Rabs (Ras-related proteins) and mediates the addition of a geranylgeranyl diphosphate group to the Rab C-terminus, resulting in prenylation. Rab27a is thought to have a role in melanosome transport, and its correct functioning is required for the light-dependent movement of melanosomes into the apical processes within RPE cells.</p>
Full article ">Figure 4
<p>Examples of corresponding 30° short wavelength blue autofluorescence (BAF) images and 30° near-infrared autofluorescence (NIR-AF) images for four patients with choroideremia ranging from children to older individuals, with varying phenotypes and visual acuities. These demonstrate qualitative visual correspondence between areas with a smooth appearance on BAF and areas with preserved homogenous NIR-AF (top four panels). The bottom panel is an example of a young patient with excellent visual acuity but without smooth zones on BAF and NIR-AF despite a large residual island.</p>
Full article ">Figure 5
<p>Longitudinal changes in 30° BAF and 30° NIR-AF imaging over a five-year follow-up period demonstrate slow progression over time. Correspondence of preserved NIR-AF and BAF smooth pattern is maintained throughout follow-up, with a qualitative reduction in the overall retinal island on BAF, as well as the smooth region on BAF, and are of homogenous NIR-AF.</p>
Full article ">
26 pages, 16750 KiB  
Article
Assessment and Application of Multi-Source Precipitation Products in Cold Regions Based on the Improved SWAT Model
by Zhaoqi Tang, Yi Wang and Wen Chen
Remote Sens. 2024, 16(22), 4132; https://doi.org/10.3390/rs16224132 - 6 Nov 2024
Viewed by 793
Abstract
In hydrological modeling, the accuracy of precipitation data and the reflection of the model’s physical mechanisms are crucial for accurately describing hydrological processes. Identifying reliable data sources and exploring reasonable hydrological evolution mechanisms for hydrology and water resources research in high-altitude mountainous regions [...] Read more.
In hydrological modeling, the accuracy of precipitation data and the reflection of the model’s physical mechanisms are crucial for accurately describing hydrological processes. Identifying reliable data sources and exploring reasonable hydrological evolution mechanisms for hydrology and water resources research in high-altitude mountainous regions with sparse stations and limited data constitute a significant challenge and focus in the field of hydrology. This study focuses on the Yarkant River Basin in Xinjiang, which originates from glaciers and contains a substantial amount of meltwater runoff. A dynamic glacier melt module considering the synergistic effects of multiple meteorological factors was developed and integrated into the original Soil and Water Assessment Tool (SWAT) model. Four precipitation datasets (ERA5-land, MSWEP, CMA V2.0, and CHM-PRE) were selected to train the model, including remote sensing precipitation products and station-interpolated precipitation data. The applicability of the improved SWAT model and precipitation datasets in the source region of the Yarkant River was evaluated and analyzed using statistical indicators, hydrological characteristic values, and watershed runoff simulation effectiveness. The optimal dataset was further used to analyze glacier evolution characteristics in the basin. The results revealed the following: (1) The improved model fills the gap in glacier runoff simulation with respect to the original SWAT model, with the simulation results more closely aligning with the actual runoff variation patterns in the study area, better describing the meltwater runoff process. (2) CMA V2.0 precipitation data has the best applicability in the study area. This is specifically reflected in the rationality of the spatial and temporal distribution patterns of the inverted precipitation, the accuracy observed in capturing precipitation events and actual precipitation characteristics, the goodness of fit in driving hydrological models, and the observed precision in reflecting the composition of watershed runoff, all of which are superior to those pertaining to other precipitation products. (3) The glacier melt calculated using the improved SWAT model informed by CMA V2.0 shows that during the study period, the basin formed a pattern with a positive–negative glacier balance demarcation at 36.5° N, featuring melting at higher latitudes and accumulation at lower latitudes. The results of this study are of significant importance for hydrometeorological applications and hydrological and water resources research in this region. Full article
(This article belongs to the Special Issue Monitoring Cold-Region Water Cycles Using Remote Sensing Big Data)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of the headwaters of the upper Yarkant River Basin.</p>
Full article ">Figure 2
<p>Flow chart for this study.</p>
Full article ">Figure 3
<p>Daily runoff simulation results before and after model improvement ((<b>A1</b>,<b>A2</b>): full period; (<b>B1</b>,<b>B2</b>): summer flood season).</p>
Full article ">Figure 4
<p>Spatial distribution of annual precipitation from precipitation products. ((<b>A</b>): CMA V2.0; (<b>B</b>): CHM-PRE; (<b>C</b>): ERA5-land; (<b>D</b>): MSWEP).</p>
Full article ">Figure 5
<p>Distribution of annual precipitation by elevation from precipitation products.</p>
Full article ">Figure 6
<p>Box plot of monthly precipitation from precipitation products.</p>
Full article ">Figure 7
<p>Taylor diagram of observed precipitation and precipitation products at stations ((A): CMA V2.0; (B): CHM-PRE; (C): ERA5-land; (D): MSWEP).</p>
Full article ">Figure 8
<p>Daily runoff simulation results before and after model improvement brought about by using precipitation products (<b>A1</b>–<b>D1</b>: Simulation results of CMA V2.0, CHM-PRE, ERA5-land, MSWEP before model improvement; <b>A2</b>–<b>D2</b>: Simulation results of CMA V2.0, CHM-PRE, ERA5-land, MSWEP after model improvement).</p>
Full article ">Figure 9
<p>Monthly runoff simulation results before and after improving the model using precipitation products.</p>
Full article ">Figure 10
<p>Annual and multi-year average contributions of glacier meltwater, rainfall, and snowmelt runoff to total runoff.</p>
Full article ">Figure 11
<p>Spatial distribution of glacier evolution characteristics (<b>A1</b>–<b>C1</b>: Glacier melting volume, accumulation and mass balance within sub-basins; <b>A2</b>–<b>C2</b>: Bubble plot of glacier melting volume, accumulation and mass balance).</p>
Full article ">
21 pages, 7085 KiB  
Article
Space-Based Mapping of Pre- and Post-Hurricane Mangrove Canopy Heights Using Machine Learning with Multi-Sensor Observations
by Boya Zhang, Daniel Gann, Shimon Wdowinski, Chaohao Lin, Erin Hestir, Lukas Lamb-Wotton, Khandker S. Ishtiaq, Kaleb Smith and Yuepeng Li
Remote Sens. 2024, 16(21), 3992; https://doi.org/10.3390/rs16213992 - 28 Oct 2024
Cited by 1 | Viewed by 1092
Abstract
Coastal mangrove forests provide numerous ecosystem services, which can be disrupted by natural disturbances, mainly hurricanes. Canopy height (CH) is a key parameter for estimating carbon storage. Airborne Light Detection and Ranging (LiDAR) is widely viewed as the most accurate method for estimating [...] Read more.
Coastal mangrove forests provide numerous ecosystem services, which can be disrupted by natural disturbances, mainly hurricanes. Canopy height (CH) is a key parameter for estimating carbon storage. Airborne Light Detection and Ranging (LiDAR) is widely viewed as the most accurate method for estimating CH but data are often limited in spatial coverage and are not readily available for rapid impact assessment after hurricane events. Hence, we evaluated the use of systematically acquired space-based Synthetic Aperture Radar (SAR) and optical observations with airborne LiDAR to predict CH across expansive mangrove areas in South Florida that were severely impacted by Category 3 Hurricane Irma in 2017. We used pre- and post-Irma LiDAR-derived canopy height models (CHMs) to train Random Forest regression models that used features of Sentinel-1 SAR time series, Landsat-8 optical, and classified mangrove maps. We evaluated (1) spatial transfer learning to predict regional CH for both time periods and (2) temporal transfer learning coupled with species-specific error correction models to predict post-Irma CH using models trained by pre-Irma data. Model performance of SAR and optical data differed with time period and across height classes. For spatial transfer, SAR data models achieved higher accuracy than optical models for post-Irma, while the opposite was the case for the pre-Irma period. For temporal transfer, SAR models were more accurate for tall trees (>10 m) but optical models were more accurate for short trees. By fusing data of both sensors, spatial and temporal transfer learning achieved the root mean square errors (RMSEs) of 1.9 m and 1.7 m, respectively, for absolute CH. Predicted CH losses were comparable with LiDAR-derived reference values across height and species classes. Spatial and temporal transfer learning techniques applied to readily available spaceborne satellite data can enable conservation managers to assess the impacts of disturbances on regional coastal ecosystems efficiently and within a practical timeframe after a disturbance event. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Florida state boundary and Hurricane Irma track. (<b>b</b>) Level 3 mangrove classification map. (<b>c</b>) Level 4 mangrove species classification. (<b>d</b>) The 30 m G-LiHT footprint of pre-Irma CHM. (<b>e</b>–<b>p</b>) Zoom-in views for four representative sites.</p>
Full article ">Figure 2
<p>Timeline of pre-Irma (orange lines) and post-Irma (purple lines) observations separated by September 2017 Hurricane Irma (the dark vertical line).</p>
Full article ">Figure 3
<p>(<b>a</b>) Flow chart of data filtering. (<b>b</b>) Spatial and temporal transfer learning based on <span class="html-italic">DS2_pre</span> and <span class="html-italic">DS2_post</span> data that are separately used in the spatial transfer but collectively used in the temporal transfer learning. Blue and green rectangles are input variables; white and red rectangles are intermediate products; ovals indicate model processes; pink and orange rectangles are output products.</p>
Full article ">Figure 4
<p>(<b>a</b>) Boxplots for the reference pre- and post-Irma CHM from <span class="html-italic">DS2</span> dataset. Red line indicates the median value; the boxes represent the interquartile range between the first quartile (25th percentile) and the third quartile (75th percentile); the whiskers extend from the edges of the box to the smallest and largest values within 1.5 times the interquartile range. (<b>b</b>) Comparison of backscatter time series and optical observations using a representative pixel from each species. For each subplot, darker color represents pre-Irma values and lighter color post-Irma values. CH values are displayed in the last row.</p>
Full article ">Figure 5
<p>(<b>a</b>) Scatter plot of predicted and referenced CH for pre- and post-Irma from one of the cross-validation evaluation datasets. The yellow lines mark the least-square linear regression model. “BW” in the legend indicates “buttonwood” species. (<b>b</b>) Mean and standard deviation of variable importance of top ten variables using the mixed feature.</p>
Full article ">Figure 6
<p>Scatter plots of prediction error and percentage (Perc) error versus reference CH for both time periods by species. Red lines are the least-squared linear models.</p>
Full article ">Figure 7
<p>Predicted CH (<b>a</b>) pre-Irma, (<b>b</b>) post-Irma, and (<b>c</b>) CH loss (positive values indicate losses). (<b>d</b>) Comparison between mean and standard deviation of predicted and reference CH loss from evaluation datasets in cross-validation. Deep color represents predicted values and light color reference values. Missing data are due to no pixel samples. (<b>e</b>) Local maps of CH loss. White circles in (<b>e3</b>) indicate (<b>left</b>) the bank areas and (<b>right</b>) the boundary between white and red mangroves according to <a href="#remotesensing-16-03992-f001" class="html-fig">Figure 1</a>f.</p>
Full article ">Figure 8
<p>(<b>a</b>) Scatter plot of predicted and corrected CH versus reference post-Irma CH from a cross-validation dataset. (<b>b</b>) Mean and standard deviation of top ten ranking of variable importance using mixed features.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>l</b>) Local maps of post-Irma CH reference, corrected predictions, and errors. White circle outlines indicate areas with large errors. (<b>m</b>) Comparison of the corrected predictions and reference CH losses across pre-Irma canopy height and species classes. Deep color indicates predicted values and light color reference values.</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>b</b>) Pre-Irma CH maps for (<b>a</b>) remake of <a href="#remotesensing-16-03992-f007" class="html-fig">Figure 7</a>a and (<b>b</b>) Figure from Jamaluddin et al. (2024) [<a href="#B32-remotesensing-16-03992" class="html-bibr">32</a>]. (<b>c</b>) Remake of CH loss predictions from <a href="#remotesensing-16-03992-f007" class="html-fig">Figure 7</a>c. (<b>d</b>) CH loss predictions from Lagomasino et al. (2021) [<a href="#B10-remotesensing-16-03992" class="html-bibr">10</a>].</p>
Full article ">
24 pages, 20172 KiB  
Article
Estimation of Forest Above-Ground Biomass in the Study Area of Greater Khingan Ecological Station with Integration of Airborne LiDAR, Landsat 8 OLI, and Hyperspectral Remote Sensing Data
by Lu Wang, Yilin Ju, Yongjie Ji, Armando Marino, Wangfei Zhang and Qian Jing
Forests 2024, 15(11), 1861; https://doi.org/10.3390/f15111861 - 24 Oct 2024
Cited by 2 | Viewed by 1220
Abstract
Accurate estimation of forest above-ground biomass (AGB) is significant for understanding changes in global carbon storage and addressing climate change. This study focuses on 53 samples of natural forests at the Greater Khingan Ecological Station, exploring the potential of integrating Canopy Height Model [...] Read more.
Accurate estimation of forest above-ground biomass (AGB) is significant for understanding changes in global carbon storage and addressing climate change. This study focuses on 53 samples of natural forests at the Greater Khingan Ecological Station, exploring the potential of integrating Canopy Height Model (CHM) with multi-source remote sensing (RS) data—airborne LiDAR, Landsat 8 OLI, and hyperspectral data to estimate forest AGB. Firstly, RS features with strong horizontal and vertical correlation with the forests AGB are optimized by a partial least squares algorithm (PLSR). Then, multivariate linear stepwise regression (MLSR) and K-nearest neighbor with fast iterative features selection (KNN-FIFS) are applied to estimate forest AGB using seven different data combinations. Finally, the leave-one-out cross-validation method is selected for the validation of the estimation results. The results are as follows: (1) When forest AGB is estimated using a single data source, the inversion results of using LiDAR are better, with R2 = 0.76 and RMSE = 21.78 Mg/ha. (2) The estimation accuracy of two models showed obvious improvement after using fused CHM into RS information. The MLSR model showed the best performance, with R2 increased by 0.41 and RMSE decreased to 14.15 Mg/ha. (3) The estimation results based on the KNN-FIFS model using the combined data of LiDAR, CHM + Landsat 8 OLI, and CHM + Hyperspectral imaging were the best in this study, with R2 = 0.85 and RMSE = 18.17 Mg/ha. The results of the study show that fusing CHM into multi-spectral data and hyperspectral data can improve the estimation accuracy a lot; the forest AGB estimation accuracies of the multi-source RS data are better than the single data source. This study provides an effective method for estimating forest AGB using multi-source data integrated with CHM to improve estimation accuracy. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>An overview map of the study area. (<b>a</b>) Boundaries of China; (<b>b</b>) location of the study area; (<b>c</b>) sample plot distribution area; (<b>d</b>) distribution of forest AGB in sample plots.</p>
Full article ">Figure 2
<p>Technology roadmap for this study.</p>
Full article ">Figure 3
<p>Preprocessing results of RS data. The legend in (<b>a</b>) indicates that the CHM values in the study area range from 0 to 31.86 m; the legend in (<b>b</b>) indicates that Red represents the red band (B4), Green represents the green band (B3), and Blue represents the blue band (B2); the legend in (<b>c</b>) indicates the three principal components retained from the hyperspectral data after PCA.</p>
Full article ">Figure 4
<p>The fusion of CHM and optical RS data. In the legend, F indicates that the band is fused with CHM information. (<b>a</b>) Multispectral and CHM fusion images; (<b>b</b>) Hyperspectral and CHM fusion images.</p>
Full article ">Figure 5
<p>Importance ranking of feature VIP ≥ 0.8.</p>
Full article ">Figure 6
<p>Forest AGB estimation results (a—LiDAR; b—Landsat 8 OLI; c—hyperspectral imaging; d—CHM + Landsat 8 OLI; e—CHM + hyperspectral imaging; f—LiDAR and Landsat 8 OLI and hyperspectral imaging; g—LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">Figure 7
<p>Cross-validation results of MLSR estimation ((<b>a</b>) LiDAR; (<b>b</b>) Landsat 8 OLI; (<b>c</b>) hyperspectral imaging; (<b>d</b>) CHM + Landsat 8 OLI; (<b>e</b>) CHM + hyperspectral imaging; (<b>f</b>) LiDAR and Landsat 8 OLI and hyperspectral imaging; (<b>g</b>) LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">Figure 8
<p>Cross-validation results of KNN-FIFS estimation ((<b>a</b>) LiDAR; (<b>b</b>) Landsat 8 OLI; (<b>c</b>) hyperspectral imaging; (<b>d</b>) CHM + Landsat 8 OLI; (<b>e</b>) CHM + hyperspectral imaging; (<b>f</b>) LiDAR and Landsat 8 OLI and hyperspectral imaging; (<b>g</b>) LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">Figure 9
<p>The residual scatter plot of the MLSR model ((<b>a</b>) LiDAR; (<b>b</b>) Landsat 8 OLI; (<b>c</b>) hyperspectral imaging; (<b>d</b>) CHM + Landsat 8 OLI; (<b>e</b>) CHM + hyperspectral imaging; (<b>f</b>) LiDAR and Landsat 8 OLI and hyperspectral imaging; (<b>g</b>) LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">Figure 10
<p>The residual scatter plot of the KNN-FIFS model ((<b>a</b>) LiDAR; (<b>b</b>) Landsat 8 OLI; (<b>c</b>) hyperspectral imaging; (<b>d</b>) CHM + Landsat 8 OLI; (<b>e</b>) CHM + hyperspectral imaging; (<b>f</b>) LiDAR and Landsat 8 OLI and hyperspectral imaging; (<b>g</b>) LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">Figure 11
<p>Multiple linear stepwise regression forest AGB distribution map ((<b>a</b>) LiDAR; (<b>b</b>) Landsat 8 OLI; (<b>c</b>) hyperspectral imaging; (<b>d</b>) CHM + Landsat 8 OLI; (<b>e</b>) CHM + hyperspectral imaging; (<b>f</b>) LiDAR and Landsat 8 OLI and hyperspectral imaging; (<b>g</b>) LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">Figure 12
<p>KNN-FIFS forest AGB distribution map ((<b>a</b>) LiDAR; (<b>b</b>) Landsat 8 OLI; (<b>c</b>) hyperspectral imaging; (<b>d</b>) CHM + Landsat 8 OLI; (<b>e</b>) CHM + hyperspectral imaging; (<b>f</b>) LiDAR and Landsat 8 OLI and hyperspectral imaging; (<b>g</b>) LiDAR and CHM + Landsat 8 OLI and CHM + hyperspectral imaging).</p>
Full article ">
16 pages, 7227 KiB  
Article
Use of Chitosan–Iron Oxide Gels for the Removal of Cd2+ Ions from Aqueous Solutions
by Eduardo Mendizábal, Nely Ríos-Donato, Minerva Guadalupe Ventura-Muñoz, Rosaura Hernández-Montelongo and Ilse Paulina Verduzco-Navarro
Gels 2024, 10(10), 630; https://doi.org/10.3390/gels10100630 - 30 Sep 2024
Viewed by 809
Abstract
High-quality water availability is substantial for sustaining life, so its contamination presents a serious problem that has been the focus of several studies. The presence of heavy metals, such as cadmium, is frequently studied due to the increase in the contamination levels caused [...] Read more.
High-quality water availability is substantial for sustaining life, so its contamination presents a serious problem that has been the focus of several studies. The presence of heavy metals, such as cadmium, is frequently studied due to the increase in the contamination levels caused by fast industrial expansion. Cadmium ions were removed from aqueous solutions at pH 7.0 by chitosan–magnetite (ChM) xerogel beads and chitosan–FeO (ChF) xerogel beads in batch systems. Kinetic studies were best modeled by the Elovich model. The adsorption isotherms obtained showed an inflection point suggesting the formation of a second layer, and the BET model adjusted to liquid–solid systems was adequate for the description of the experimental data. Maximum uptake capacities of 36.97 ± 0.77 and 28.60 ± 2.09 mg Cd/g xerogel were obtained for ChM and ChF, respectively. The studied composites are considered promising adsorbent materials for removing cadmium ions from aqueous systems. Full article
(This article belongs to the Special Issue Gel-Based Materials: Preparations and Characterization (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM images with: 100× magnification of (<b>A</b>) ChM and (<b>D</b>) ChF, 400× magnification of (<b>B</b>) ChM and (<b>E</b>) ChF, and 1000× magnification of (<b>C</b>) ChM and (<b>F</b>) ChF.</p>
Full article ">Figure 2
<p>Potentiometric titration for the determination of the zero charge point of (<b>A</b>) chitosan, (<b>B</b>) ChM, and (<b>C</b>) ChF.</p>
Full article ">Figure 3
<p>Plots of the change in pH (ΔpH) as a function of pH<sub>i</sub> using the salt addition method: (<b>A</b>) ChM and (<b>B</b>) ChF.</p>
Full article ">Figure 4
<p>X-ray diffraction pattern of (<b>A</b>) magnetite and ChM and (<b>B</b>) FeO and ChF.</p>
Full article ">Figure 5
<p>FTIR spectra of (<b>A</b>) magnetite and ChM, (<b>B</b>) FeO and ChF.</p>
Full article ">Figure 6
<p>Experimental kinetic data of Cd sorption (•) using different initial cadmium concentrations: (<b>A</b>) 10 mg/L ChM, (<b>B</b>) 10 mg/L ChF, (<b>C</b>)100 mg/L ChM, and (<b>D</b>) 100 mg/L ChF. Predicted values by Elovich (<b><span style="color:fuchsia">—</span></b>), pseudo-first-order (<b><span style="color:red">—</span></b>) and pseudo-second-order (<b><span style="color:blue">—</span></b>) models.</p>
Full article ">Figure 7
<p>Cadmium adsorption onto (<b>A</b>) ChM and (<b>B</b>) ChF at pH 7.0 and 25 °C; experimental (-) and predictions by BET (<b><span style="color:red">—</span></b>), Freundlich (<b><span style="color:blue">—</span></b>) and Temkin (<b><span style="color:fuchsia">—</span></b>) models.</p>
Full article ">Figure 8
<p>Uptake capacity (•) and change in pH (<span class="html-fig-inline" id="gels-10-00630-i001"><img alt="Gels 10 00630 i001" src="/gels/gels-10-00630/article_deploy/html/images/gels-10-00630-i001.png"/></span>) behavior as a function of cadmium concentration at equilibrium conditions at an initial solution pH of 7.0 and 25 °C. (<b>A</b>) ChM, (<b>B</b>) ChF.</p>
Full article ">Figure 9
<p>Scheme for the adsorption of Cd<sup>2+</sup> onto either ChM or ChF. Interactions between chitosan and iron oxides (magnetite or FeO) in the composites are represented by (<span style="color:#0000CC">---</span>) and the interaction between the composites and Cd<sup>2+</sup> is represented by (<span style="color:red">---</span>).</p>
Full article ">Figure 10
<p>Cadmium uptake capacity of ChM (<span class="html-fig-inline" id="gels-10-00630-i002"><img alt="Gels 10 00630 i002" src="/gels/gels-10-00630/article_deploy/html/images/gels-10-00630-i002.png"/></span>) and ChF (<span class="html-fig-inline" id="gels-10-00630-i003"><img alt="Gels 10 00630 i003" src="/gels/gels-10-00630/article_deploy/html/images/gels-10-00630-i003.png"/></span>) as a function of salinity, expressed as NaCl molarity.</p>
Full article ">Figure 11
<p>Scheme for the experimental procedure used for the obtention of ChM and ChF xerogel beads.</p>
Full article ">
21 pages, 2395 KiB  
Review
Quantitative Changes and Transformation Mechanisms of Saponin Components in Chinese Herbal Medicines during Storage and Processing: A Review
by Yuhang Wu, Hui Zheng, Tao Zheng, Jiani Jiang, Yao Xu, Fan Jia, Kai He and Yong Yang
Molecules 2024, 29(18), 4486; https://doi.org/10.3390/molecules29184486 - 21 Sep 2024
Viewed by 1675
Abstract
Saponins are an important class of active components in Chinese herbal medicines (CHMs), which are present in large quantities in Ginseng Radix et Rhizoma, Notoginseng Radix et Rhizoma, Polygonati Rhizoma, etc., and have immune regulation, anti-tumor, anti-inflammatory, anti-cardiovascular disease, and [...] Read more.
Saponins are an important class of active components in Chinese herbal medicines (CHMs), which are present in large quantities in Ginseng Radix et Rhizoma, Notoginseng Radix et Rhizoma, Polygonati Rhizoma, etc., and have immune regulation, anti-tumor, anti-inflammatory, anti-cardiovascular disease, and hypoglycemic activities. Storage and processing are essential processes in the production process of CHMs which affect the stability of saponin components and then reduce the medicinal and economic value. Therefore, it is of great importance to investigate the effects of storage and processing conditions on the content of saponin components in CHMs. In this paper, the effects of various storage and processing factors, including temperature, pH, enzymes, meta lions, extraction methods, etc., on the saponin content of CHMs are investigated and the underlying mechanisms for the quantitative changes of saponin are summarized. These findings may provide technical guidance for the production and processing of saponin-rich CHMs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Some Chinese herbal medicines (CHMs) and their saponin components.</p>
Full article ">Figure 2
<p>The dissolution of saponins increased after cell wall-breaking treatment.</p>
Full article ">Figure 3
<p>Dioscin-glycosidase hydrolysis pathway of dioscin.</p>
Full article ">Figure 4
<p>Conversion pathway of gypenosides during heat treatment.</p>
Full article ">Figure 5
<p>Thermal conversion pathway of malonyl ginsenosides.</p>
Full article ">Figure 6
<p>Dehydration condensation pathway of ginsenoside Rg3.</p>
Full article ">
20 pages, 14699 KiB  
Article
The Early Prediction of Kimchi Cabbage Heights Using Drone Imagery and the Long Short-Term Memory (LSTM) Model
by Seung-hwan Go and Jong-hwa Park
Drones 2024, 8(9), 499; https://doi.org/10.3390/drones8090499 - 18 Sep 2024
Cited by 1 | Viewed by 770
Abstract
Accurate and timely crop growth prediction is crucial for efficient farm management and food security, particularly given challenges like labor shortages and climate change. This study presents a novel method for the early prediction of Kimchi cabbage heights using drone imagery and a [...] Read more.
Accurate and timely crop growth prediction is crucial for efficient farm management and food security, particularly given challenges like labor shortages and climate change. This study presents a novel method for the early prediction of Kimchi cabbage heights using drone imagery and a long short-term memory (LSTM) model. High-resolution drone images were used to generate a canopy height model (CHM) for estimating plant heights at various growth stages. Missing height data were interpolated using a logistic growth curve, and an LSTM model was trained on this time series data to predict the final height at harvest well before the actual harvest date. The model trained on data from 44 days after planting (DAPs) demonstrated the highest accuracy (R2 = 0.83, MAE = 2.48 cm, and RMSE = 3.26 cm). Color-coded maps visualizing the predicted Kimchi cabbage heights revealed distinct growth patterns between different soil types, highlighting the model’s potential for site-specific management. Considering the trade-off between accuracy and prediction timing, the model trained on DAP 36 data (MAE = 2.77 cm) was deemed most suitable for practical applications, enabling timely interventions in cultivation management. This research demonstrates the feasibility and effectiveness of integrating drone imagery, logistic growth curves, and LSTM models for the early and accurate prediction of Kimchi cabbage heights, facilitating data-driven decision-making in precision agriculture for improved crop management and yield optimization. Full article
Show Figures

Figure 1

Figure 1
<p>Study site overview: Location of the Kimchi cabbage testbed within the Republic of Korea (left) and detailed layout of the testbed at the National Institute of Agricultural Sciences (NAS) showing soil zones (A and B) and ground control point (GCP) locations (right).</p>
Full article ">Figure 2
<p>Workflow for early prediction of Kimchi cabbage heights using drone imagery and LSTM-based modeling.</p>
Full article ">Figure 3
<p>Geometric correction and coordinate system unification of drone imagery using GCPs.</p>
Full article ">Figure 4
<p>Schematic of generating a canopy height model (CHM) from time series digital surface models (DSMs) and a digital terrain model (DTM).</p>
Full article ">Figure 5
<p>Data preparation for LSTM model development: (<b>a</b>) logistic growth curve fitting using CHM data at key growth stages; and (<b>b</b>) dataset partitioning for training, validation, and testing.</p>
Full article ">Figure 6
<p>LSTM network architecture for Kimchi cabbage height prediction, illustrating the input data structure, hidden states, and the internal gating mechanism of an LSTM cell. The input data are a three-dimensional matrix with dimensions representing the number of Kimchi cabbages, time steps, and CHM for each day.</p>
Full article ">Figure 7
<p>Temporal visualization of Kimchi cabbage growth in Zone A (loam) and Zone B (sandy loam) using drone imagery: (<b>a</b>) RGB orthomosaics, (<b>b</b>) DSMs, and (<b>c</b>) CHMs at three key growth stages (DAPs 0, 36, and 71).</p>
Full article ">Figure 8
<p>Validation of Kimchi cabbage height estimation from drone imagery: (<b>a</b>) boxplot of field-measured plant heights, (<b>b</b>) scatter plot comparing field-measured heights with canopy heights derived from CHMs, and (<b>c</b>) temporal progression of CHM in Zones A and B.</p>
Full article ">Figure 9
<p>Kimchi cabbage growth dynamics: boxplots depicting daily canopy height (CHM) variations alongside logistic growth curves fitted to CHM data at key growth stages (DAPs 29, 36, and 44).</p>
Full article ">Figure 10
<p>Evaluation of LSTM models for Kimchi cabbage height prediction at different training stages: (<b>a</b>) DAP 29, (<b>b</b>) DAP 36, and (<b>c</b>) DAP 44. Scatter plots compare predicted heights against true heights for both Zone A (loam) and Zone B (sandy loam), with corresponding R<sup>2</sup>, MAE, and RMSE values.</p>
Full article ">Figure 11
<p>Spatiotemporal visualization of predicted Kimchi cabbage height in Zone A (loam) and Zone B (sandy loam): comparison of LSTM model-generated height maps at different growth stages (<b>a</b>) DAP 29, (<b>b</b>) DAP 36, and (<b>c</b>) DAP 44 with a (<b>d</b>) reference map based on field measurements at harvest (DAP 71).</p>
Full article ">
Back to TopTop