Nothing Special   »   [go: up one dir, main page]

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
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
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
remove_circle_outline
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,896)

Search Parameters:
Keywords = near infrared

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 4608 KiB  
Article
Transmission Spectroscopy Along the Transit of Venus: A Proxy for Exoplanets Atmospheric Characterization
by Alexandre Branco, Pedro Machado, Olivier Demangeon, Tomás Azevedo Silva, Sarah A. Jaeggli, Thomas Widemann and Paolo Tanga
Atmosphere 2024, 15(12), 1431; https://doi.org/10.3390/atmos15121431 - 28 Nov 2024
Abstract
We present an analysis of high-resolution, near-infrared (NIR) spectra relative to the solar transit of Venus of 5–6 June 2012, as observed with the Facility Infrared Spectropolarimeter (FIRS) at the Dunn Solar Telescope in New Mexico. These observations offer the unique opportunity to [...] Read more.
We present an analysis of high-resolution, near-infrared (NIR) spectra relative to the solar transit of Venus of 5–6 June 2012, as observed with the Facility Infrared Spectropolarimeter (FIRS) at the Dunn Solar Telescope in New Mexico. These observations offer the unique opportunity to probe the upper layers (above ∼84 km in altitude) of a thick, CO2-dominated atmosphere with the transmission spectroscopy technique—a proxy for future studies of highly-irradiated atmospheres of Earth-sized exoplanets. We were able to directly observe absorption lines from the two most abundant CO2 isotopologues, and from the main isotopologue of CO in the retrieved spectrum of Venus. Furthermore, we performed a cross-correlation analysis of the transmission spectrum using transmission templates generated with petitRADTRANS. With the cross-correlation technique, it was possible to confirm detections of both CO2 isotopologues and CO. Additionally, we retrieved a tentative cross-correlation signal for O3 on Venus. We demonstrate the feasibility of high-resolution, ground-based observations to study the chemical inventory of planetary atmospheres, employing techniques commonly used in exoplanet characterization. Full article
(This article belongs to the Section Planetary Atmospheres)
Show Figures

Figure 1

Figure 1
<p>Intensity raster map with averaged spectral intensity represented in log-scale. Gray dashed lines delimit the region containing the planetary limb. These represent two concentric ellipses centered at (X, Y) = (57″, 39.75″), marked with a red cross. Two vertical dashed lines were included to outline the region of the limb where the atmospheric aureole is observed. Two solid gray lines delimit the location of the solar limb.</p>
Full article ">Figure 2
<p>Intensity raster map of the aureole region aligned by the highest intensity pixel in each slice, with averaged spectral intensity represented in log-scale. The red dashed lines delimit a 4-pixel tall window of interest, defining the region of interest pertaining to the atmospheric aureole. The gray dashed line defines an out-of-interest area separated by 13 pixels from the top of our region of interest and spanning until the top of each slice. Out-of-interest spectra were used to correct for the presence of stray light in aureole spectra.</p>
Full article ">Figure 3
<p>Average transmission spectrum of Venus extracted from the atmospheric aureole as observed during the solar transit of 2012 (in gray). For each slice, an average spectrum of the Venus limb was calculated and divided by an average spectrum of the background contained in that same slice. The presented spectrum corresponds to the average of the division products over all slices after being continuum-corrected. Synthetic transmission spectra for CO<sub>2</sub> and CO isotopologues were generated with petitRADTRANS and are shown for comparison. The absorption lines identified upon visual inspection of the observed spectrum have been marked: <sup>12</sup>C<sup>16</sup>O<sub>2</sub> (yellow), <sup>13</sup>C<sup>16</sup>O<sub>2</sub> (red), <sup>12</sup>C<sup>16</sup>O (dark blue).</p>
Full article ">Figure 4
<p>Cross-correlation functions for (<b>a</b>) <sup>12</sup>C<sup>16</sup>O<sub>2</sub>, (<b>b</b>) <sup>13</sup>C<sup>16</sup>O<sub>2</sub>, (<b>c</b>) <sup>12</sup>C<sup>16</sup>O (dark blue line). The best-fit Gaussian profiles are shown for each CCF (red dashed line). All panels show the CCFs resulting from the self cross-correlation of the templates (light orange area), which are scaled arbitrarily.</p>
Full article ">Figure 5
<p>Cross-correlation function for <sup>16</sup>O<sub>3</sub> (dark blue line) along with the best-fit Gaussian profile (red dashed line). The CCF resulting from the self cross-correlation of the template is shown as a light orange area, arbitrarily scaled for comparison.</p>
Full article ">Figure A1
<p>Transmission templates used in cross-correlation analysis for each chemical species. The y-axis displays the statistical weight assigned to each spectral wavelength, which is used to calculate the cross-correlation function. This function acts as a weighted average of the relative spectral intensity observed across the entire wavelength range, for distinct radial velocity offsets of the template. The results we present for each species use an isothermal PT profile at 175 K.</p>
Full article ">
9 pages, 543 KiB  
Article
Comparing Images from Near-Infrared Light Reflection and Bitewing Radiography to Detect Proximal Caries in Primary Teeth
by Aviv Shmueli, Avia Fux-Noy, Esti Davidovich, Diana Ram and Moti Moskovitz
Children 2024, 11(12), 1455; https://doi.org/10.3390/children11121455 - 28 Nov 2024
Abstract
Objectives: The present prospective study aimed to compare near-infrared light reflection (NIRI) and bitewing radiographs (BWR) images to detect proximal caries in primary teeth. Methods: 71 children underwent routine BWR, and scans were performed using an intra-oral scanner (iTero Element 5D, Align Technology, [...] Read more.
Objectives: The present prospective study aimed to compare near-infrared light reflection (NIRI) and bitewing radiographs (BWR) images to detect proximal caries in primary teeth. Methods: 71 children underwent routine BWR, and scans were performed using an intra-oral scanner (iTero Element 5D, Align Technology, Tempe, AZ, USA), including a near-infrared light source (850 nm) and sensor. Five specialist pediatric dentists examined the NIRI and BWR images. Results: The average participant age was 7.8 years. A total of 1004 proximal surfaces of primary molars and canines were examined, 209 carious lesions were detected on BWR, and 227 on NIRI. Comparison between all carious lesions detected on BWR and NIRI: Sensitivity (53.6%); Specificity (85.5%); Positive Predictive Value (PPV) (49.3%); Negative Predictive Value (NPV) (87.5%). Comparison between carious lesions involving the DEJ detected on BWR and at any level in NIRI: Sensitivity (61%); Specificity (83.4%), PPV (36.6%); NPV (93.2%). Comparison between enamel-only carious lesions detected on BWR and all lesions detected using NIRI: Sensitivity (44.8%); Specificity (85.5%); PPV (20.7%); NPV (94.8%). Conclusions: No additional diagnostic information can be gleaned from BWR if initial caries lesions in the enamel are not detected by clinical examination or in images from a NIRI scanner, making BWR unnecessary. Full article
(This article belongs to the Section Pediatric Dentistry & Oral Medicine)
Show Figures

Figure 1

Figure 1
<p>Box-plot Agreement Analysis between examiners.</p>
Full article ">
29 pages, 4490 KiB  
Article
Genotypic Influences on Actuators of Aerobic Performance in Tactical Athletes
by Martin Flück, Christian Protte, Marie-Noëlle Giraud, Thomas Gsponer and Alain Dössegger
Genes 2024, 15(12), 1535; https://doi.org/10.3390/genes15121535 - 28 Nov 2024
Viewed by 75
Abstract
Background: This study examines genetic variations in the systemic oxygen transport cascade during exhaustive exercise in physically trained tactical athletes. Research goal: To update the information on the distribution of influence of eleven polymorphisms in ten genes, namely ACE (rs1799752), AGT (rs699), MCT1 [...] Read more.
Background: This study examines genetic variations in the systemic oxygen transport cascade during exhaustive exercise in physically trained tactical athletes. Research goal: To update the information on the distribution of influence of eleven polymorphisms in ten genes, namely ACE (rs1799752), AGT (rs699), MCT1 (rs1049434), HIF1A (rs11549465), COMT (rs4680), CKM (rs8111989), TNC (rs2104772), PTK2 (rs7460 and rs7843014), ACTN3 (rs1815739), and MSTN (rs1805086)—on the connected steps of oxygen transport during aerobic muscle work. Methods: 251 young, healthy tactical athletes (including 12 females) with a systematic physical training history underwent exercise tests, including standardized endurance running with a 12.6 kg vest. Key endurance performance metrics were assessed using ergospirometry, blood sampling, and near-infrared spectroscopy of knee and ankle extensor muscles. The influence of gene polymorphisms on the above performance metrics was analyzed using Bayesian analysis of variance. Results: Subjects exhibited good aerobic fitness (maximal oxygen uptake (VO2max): 4.3 ± 0.6 L min−1, peak aerobic power: 3.6 W ± 0.7 W kg−1). Energy supply-related gene polymorphisms rs1799752, rs4680, rs1049434, rs7843014, rs11549465, and rs8111989 did not follow the Hardy–Weinberg equilibrium. Polymorphisms in genes that regulate metabolic and contractile features were strongly associated with variability in oxygen transport and metabolism, such as body mass-related VO2 (rs7843014, rs2104772), cardiac output (rs7460), total muscle hemoglobin content (rs7460, rs4680), oxygen saturation in exercised muscle (rs1049434), and respiration exchange ratio (rs7843014, rs11549465) at first or secondary ventilatory thresholds or VO2max. Moderate influences were found for mass-related power output. Conclusions: The posterior distribution of effects from genetic modulators of aerobic metabolism and muscle contractility mostly confirmed prior opinions in the direction of association. The observed genetic effects of rs4680 and rs1049434 indicate a crucial role of dopamine- and lactate-modulated muscle perfusion and oxygen metabolism during running, suggesting self-selection in Swiss tactical athletes. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Visualized hypotheses. Composite drawing of the research hypothesis of genetic influences on oxygen transport and aerobic performance. (<b>Left</b>) Assessed elements of the oxygen transport cascade during muscle work. (<b>Right</b>) Color-coded listing of the hypothesized effects of gene polymorphism on the assessed parameters of the oxygen transport cascade (prior knowledge). Numbers denote the mean-centered differences (step size) between major and minor alleles for the hypothesized effects of a respective gene polymorphism. Light gray cells indicate hypothesized influences of the intensity of exercise or prior exercise/warm-up. Darkly highlighted cells denote those where the prior available information was contradictory. Empty cells denote instances where no prior information was indicated to formulate a specific hypothesis. Abbreviations: ACE, angiotensin-converting enzyme; ACTN3, alpha actinin-3; ADP, adenosine diphosphate; aePerf, aerobic performance; AGT, angiotensinogen; ATP, adenosine triphosphate; anae Perf_res, anaerobic power reserve; CKM, muscle-type creatine kinase; CO<sub>2</sub>, carbon dioxide; COMT, catechol-O-methyltransferase; GAS, gastrocnemius (medialis) muscle; HIF1A, hypoxia-inducible factor 1 alpha; m, minor variant of a gene polymorphism; M, major variant of gene polymorphism; MCT1, monocarboxylate transporter 1; MSTN, myostatin; PTK2, protein tyrosine kinase 2 (or focal adhesion kinase); O<sub>2</sub>, oxygen; rsid, identifier of gene polymorphism; SmO<sub>2</sub>, muscle oxygen saturation; TNC, tenascin-C; tHb, total hemoglobin concentration; VAS, vastus lateralis muscle.</p>
Full article ">Figure 2
<p>Physiological characteristics during loaded, graded exercise. Rain cloud line plots of individual values for assessed characteristics of oxygen transport (Y-axes) during the different phases (X-axes) of loaded and graded running exercise into exhaustion for the studied 251 subjects. Cases with missing data for a phase were excluded from the display. (<b>A</b>–<b>H</b>) time (<b>A</b>), performance (<b>B</b>), VO<sub>2</sub> (<b>C</b>), Q′ (<b>D</b>), tHb in VAS (<b>E</b>), SmO<sub>2</sub> in VAS (<b>F</b>), SmO<sub>2</sub> in GAS (<b>G</b>), and RER (<b>H</b>). Abbreviations: max, maximal values; stop + 2 min, 120 s into recovery after the cessation of running; VO<sub>2</sub>, oxygen uptake; VT1, ventilatory threshold 1; VT2, ventilatory threshold 2; Perf, power; Q′, cardiac output.</p>
Full article ">Figure 3
<p>Examples of the identified genotype effects on body mass-related oxygen uptake. (<b>A</b>–<b>C</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) and individual values (circles) for the influence of rs2104772 on VO<sub>2</sub> at VT2 (<b>A</b>) and VO<sub>2</sub>max (<b>B</b>), and rs7843014 at VT1 (<b>C</b>). Y-axes resume the identity of the respective response variable and applicable unit, while the X-axes indicate the respective genotypes of the addressed gene polymorphism. Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5; **, 30.0 ≥ BF10 &gt; 10.0; ***, BF10 &gt; 30.0. Abbreviations: VO2@VT1, VO<sub>2</sub> at VT1; VO2@VT2, VO<sub>2</sub> at VT2.</p>
Full article ">Figure 4
<p>Examples of the identified genotype effects on cardiac output. (<b>A</b>–<b>C</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) with individual values (circles) for the influence of rs7460 (<b>A</b>), rs1815739 (<b>B</b>), and rs8111989 (<b>C</b>) on overall cardiac output. Y- and X-axes resume the identity of the respective response variable, the applicable unit, and the respective genotype. Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5.</p>
Full article ">Figure 5
<p>Examples of the identified genotype effects on total hemoglobin concentration of vastus lateralis muscle. (<b>A</b>–<b>F</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) with individual values (circles) for the influence of rs4680 on overall tHb in VAS (<b>A</b>), tHb in VAS at VT2 (<b>B</b>), and tHb in VAS at VO<sub>2</sub>max (<b>C</b>), as well as rs2104772 (<b>D</b>), rs7460 (<b>E</b>), and rs1049413 (<b>F</b>) on overall tHb in VAS. Y- and X-axes resume the identity of the respective response variable, the applicable unit, and the respective genotype. Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5: **, 30.0 ≥ BF10 &gt; 10.0; ***, BF10 &gt; 30.0. Abbreviations: tHb@VT2, tHb at VT2; tHb@VO2max, tHb at VO<sub>2</sub>max.</p>
Full article ">Figure 6
<p>Examples of the identified genotype effects on respiration exchange ratio. (<b>A</b>–<b>D</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) with individual values (circles) for the influence of rs11549465 (<b>A</b>) and rs7843014 (<b>B</b>) on RER at VT2, and rs2104772 (<b>C</b>) and rs7843014 (<b>D</b>) on RER at VO<sub>2</sub>max. Y- and X-axes resume the identity of the response variable, the applicable unit, and the respective genotype. Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5; **, 30.0 ≥ BF10 &gt; 10.0. Abbreviations: RER@VO2max; RER at VO<sub>2</sub>max; RER@VT2; RER at VT2.</p>
Full article ">Figure 7
<p>Examples of the identified genotype effects on oxygen saturation in vastus lateralis muscle. (<b>A</b>–<b>E</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) with individual values (circles) for the influence of rs4680 (<b>A</b>) and rs1049434 (<b>B</b>) on SmO<sub>2</sub> in VAS at the start of exercise, as well as rs1815739 on SmO<sub>2</sub> in VAS at VO<sub>2</sub>max (<b>C</b>), and rs4680 (<b>D</b>) and rs1049434 (<b>E</b>) SmO<sub>2</sub> in VAS at VT1. Y- and X-axes resume the identity of the response variable and the applicable unit, and the respective genotype. Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5; **, 30.0 ≥ BF10 &gt; 10.0; ***, BF10 &gt; 30.0. Abbreviations: SmO2@start, SmO<sub>2</sub> at start VAS; SmO2@VO2max, SmO<sub>2</sub> at VO<sub>2</sub>max; SmO2@VT1, SmO<sub>2</sub> at VT1.</p>
Full article ">Figure 8
<p>Examples of the identified genotype effects on oxygen saturation in gastrocnemius medialis muscle. (<b>A</b>–<b>D</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) with individual values (circles) for the influence of rs7460 (<b>A</b>) and rs1049434 (<b>B</b>) on SmO<sub>2</sub> in GAS at the start of exercise, as well as rs11549465 on SmO<sub>2</sub> in GAS at VT1 (<b>C</b>) and rs1049434 on SmO<sub>2</sub> in GAS at VO<sub>2</sub>max (<b>D</b>). Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5; ***, BF10 &gt; 30.0.</p>
Full article ">Figure 9
<p>Examples of the identified genotype effects on power output. (<b>A</b>–<b>G</b>) Box plots (line: median; cross: mean; box: data from first to third quartile; whiskers: ±1.5 × interquartile range) with individual values (circles) for the influence of rs11549465 on performance at VO<sub>2</sub>max (<b>A</b>), rs7460 (<b>B</b>), and rs1815739 (<b>C</b>) on performance at VT1, as well as rs11549465 (<b>D</b>) and rs2104772 (<b>E</b>) on performance at VT2, as well as rs699 (<b>F</b>) and rs1815739 (<b>G</b>) on the anaerobic power reserve. Y- and X-axes resume the identity of the response variable, the applicable unit, and the respective genotype. Respective Bayes factors (BF10) for post hoc effects are given as follows: *, 10.0 ≥ BF10 &gt; 2.5. Abbreviations: Perf@VO2max, power at VO<sub>2</sub>max; Perf@VT1, power at VT1; Perf@VT2, power at VT2.</p>
Full article ">Figure 10
<p>Sites of interaction of assessed gene polymorphism with metabolic and contractile processes involved in energy provision and force development by muscle fibers. Arrows point to demonstrated points of influence of anatomical/biochemical processes by the connected gene polymorphisms. Drawing of the cellular makeup of skeletal muscle in muscle fibers and capillaries and the embedded myofibrillar and mitochondrial organelles, as well as biochemical processes for fueling energetic requirements during physical work by means of the aerobic combustion of organic substrates. Arrows indicate sites of demonstrated influence of the eleven studied gene polymorphisms. We refer to the other illustrations and tables regarding the abbreviations.</p>
Full article ">
11 pages, 498 KiB  
Article
Using Vital Signs for the Early Prediction of Necrotizing Enterocolitis in Preterm Neonates with Machine Learning
by Rosa Verhoeven, Thijmen Kupers, Celina L. Brunsch, Jan B. F. Hulscher and Elisabeth M. W. Kooi
Children 2024, 11(12), 1452; https://doi.org/10.3390/children11121452 - 28 Nov 2024
Viewed by 111
Abstract
Background/Objectives: Necrotizing enterocolitis (NEC), a devastating neonatal gastrointestinal disease mostly seen in preterm infants, lacks accurate prediction despite known risk factors. This hinders the possibility of applying targeted preventive therapies. This study explores the use of vital signs, including cerebral and splanchnic oxygenation, [...] Read more.
Background/Objectives: Necrotizing enterocolitis (NEC), a devastating neonatal gastrointestinal disease mostly seen in preterm infants, lacks accurate prediction despite known risk factors. This hinders the possibility of applying targeted preventive therapies. This study explores the use of vital signs, including cerebral and splanchnic oxygenation, measured with near-infrared spectroscopy in early NEC prediction. Methods: Several machine learning algorithms were trained on data from very preterm patients (<30 weeks gestational age). Time Series FeatuRe Extraction on the basis of scalable hypothesis tests (TSFRESH) extracted significant features from the vital signs of the first 5 postnatal days. We present the F1-scores and area under the precision-recall curve (AUC-PR) of the models. The contribution of separate vital signs to the selected TSFRESH features was also determined. Results: Among 267 patients, 32 developed NEC Bell’s stage > 1. Using a 1:4 NEC:control ratio, support vector machine and logistic regression predicted NEC better than extreme gradient boosting regarding the F1-score (0.82, 0.82, 0.76, resp., p = 0.001) and AUC-PR (0.82, 0.83, 0.77, resp., p < 0.001). Splanchnic and cerebral oxygenation contributed most to the prediction (40.1% and 24.8%, resp.). Conclusions: Using vital signs, we predicted NEC in the first 5 postnatal days with an F1-score up to 0.82. Splanchnic and cerebral oxygenation were the most contributing vital predictors. This pioneering effort in early NEC prediction using vital signs underscores the potential for targeted preventive measures and also emphasizes the need for additional data in future studies. Full article
(This article belongs to the Section Pediatric Neonatology)
Show Figures

Figure 1

Figure 1
<p>Patient inclusion. Patients with gestational age &lt; 30 weeks, who survived for more than 7 days without NEC, and whose vital signs were available for the first 5 days after birth are included. Patients with Bell’s stage IIa or higher are categorized as NEC.</p>
Full article ">
9 pages, 4121 KiB  
Communication
Correction of Multispectral Singlet Oxygen Luminescent Dosimetry (MSOLD) for Tissue Optical Properties in Photofrin-Mediated Photodynamic Therapy
by Weibing Yang, Madelyn Johnson, Baozhu Lu, Dennis Sourvanos, Hongjing Sun, Andreea Dimofte, Vikas Vikas, Theresa M. Busch, Robert H. Hadfield, Brian C. Wilson and Timothy C. Zhu
Antioxidants 2024, 13(12), 1458; https://doi.org/10.3390/antiox13121458 - 28 Nov 2024
Viewed by 217
Abstract
The direct detection of singlet-state oxygen (1O2) constitutes the holy grail dosimetric method for type-II photodynamic therapy (PDT), a goal that can be quantified using multispectral singlet oxygen near-infrared luminescence dosimetry (MSOLD). The optical properties of tissues, specifically their [...] Read more.
The direct detection of singlet-state oxygen (1O2) constitutes the holy grail dosimetric method for type-II photodynamic therapy (PDT), a goal that can be quantified using multispectral singlet oxygen near-infrared luminescence dosimetry (MSOLD). The optical properties of tissues, specifically their scattering and absorption coefficients, play a crucial role in determining how the treatment and luminescence light are attenuated. Variations in these properties can significantly impact the spatial distribution of the treatment light and hence the generation of singlet oxygen and the detection of singlet oxygen luminescence signals. In this study, we investigated the impact of varying optical properties on the detection of 1O2 luminescence signals during Photofrin-mediated PDT in tissue-mimicking phantoms. For comparison, we also conducted Monte Carlo (MC) simulations under the same conditions. The experimental and simulations are substantially equivalent. This study advances the understanding of MSOLD during PDT. Full article
(This article belongs to the Section ROS, RNS and RSS)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>). Schematic illustration of the experimental set-up in which the laser strikes the cuvette along the z-axis and the detection fibers with NA = 0.5 are placed at 20 ± 2° to the laser beam with the fiber tip 4 ± 0.5 mm above the surface, the pink color represents the Photofrin solution, while the black color represents the black phantom. (<b>b</b>) Schematic illustration of the Monte Carlo simulation configuration, mirroring the experimental set-up, the red region represents the liquid phantom, while the blue region represents air.</p>
Full article ">Figure 2
<p>The extinction coefficient spectrum of Photofrin around the excitation (<b>a</b>) and <sup>1</sup>O<sub>2</sub> luminescence emission (<b>b</b>) wavelengths. The former is consistent with the values in the literature [<a href="#B25-antioxidants-13-01458" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) The absorption spectrum of the black ink from 600 nm to 1000 nm. (<b>b</b>) The calculated absorption coefficient at 632 nm as a function of the concentration, together with a linear fit with the measurements. (<b>c</b>,<b>d</b>) Corresponding measurements from 1200 to 1350 nm and at 1270 nm.</p>
Full article ">Figure 4
<p>(<b>a</b>) An example of the measured light fluence rate per unit source strength (Φ/S) at 661 nm versus the distance along the catheter from a point source: fitted <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msubsup> <mo>=</mo> <mn>8.5</mn> <mtext> </mtext> <msup> <mrow> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>. (<b>b</b>) The reduced scattering coefficient from (<b>a</b>) at 661 nm for Intralipid as a function of the concentration and the corresponding linear fit.</p>
Full article ">Figure 5
<p>Singlet oxygen luminescence spectrum of Photofrin, measured in a turbid phantom with <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo>=</mo> <mn>1.0</mn> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msubsup> <mo>=</mo> <mn>15</mn> <mo> </mo> <msup> <mrow> <mi mathvariant="normal">c</mi> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>, together with the calculated component spectra.</p>
Full article ">Figure 6
<p>Measured singlet oxygen luminescence spectra with 10-point smoothing and the corresponding spectra obtained by SVD fitting. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> ranging from 0.6 to 1.5 cm<sup>−1</sup> at 632 nm (ink + Photofrin contributions) with <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mo>′</mo> </msubsup> </mrow> </semantics></math> fixed at 15 cm<sup>−1</sup>. (<b>b</b>) Corresponding spectra with <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mo>′</mo> </msubsup> </mrow> </semantics></math> = 5–40 cm<sup>−1</sup> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = 0.5 cm<sup>−1</sup> at 632 nm.</p>
Full article ">Figure 7
<p>Dependence of the normalized <sup>1</sup>O<sub>2</sub> signal on the scattering coefficient at 632 nm, comparing the experimental data (points) and the Monte Carlo simulations (solid lines), normalized to the values of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = 0.8 cm<sup>−1</sup> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>s</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msubsup> </mrow> </semantics></math> = 15 cm<sup>−1</sup>.</p>
Full article ">
21 pages, 2657 KiB  
Systematic Review
Enhancing Glioblastoma Resection with NIR Fluorescence Imaging: A Systematic Review
by Hadeel M. Mansour, Siddharth Shah, Tania M. Aguilar, Mohammed Abdul-Muqsith, Gabriel S. Gonzales-Portillo and Ankit I. Mehta
Cancers 2024, 16(23), 3984; https://doi.org/10.3390/cancers16233984 - 27 Nov 2024
Viewed by 195
Abstract
Glioblastoma (GB) is among the most aggressive and difficult-to-treat brain tumors, with a median survival of only 12–15 months despite maximal treatments, including surgery, radiotherapy, and chemotherapy. Extensive surgical resection improves survival in glioblastoma patients; however, achieving complete resection is often hindered by [...] Read more.
Glioblastoma (GB) is among the most aggressive and difficult-to-treat brain tumors, with a median survival of only 12–15 months despite maximal treatments, including surgery, radiotherapy, and chemotherapy. Extensive surgical resection improves survival in glioblastoma patients; however, achieving complete resection is often hindered by limitations in neurosurgical guidance technologies for accurate tumor margin detection. Recent advancements in fluorescence-guided surgery (FGS) and imaging techniques have significantly enhanced the precision and extent of glioblastoma resections. This study evaluates the impact of NIR fluorescence imaging on tumor visualization, surgical precision, cost-effectiveness, and patient survival. A systematic review of PubMed, Scopus, Google Scholar, and Embase was conducted to identify studies on the role of NIR fluorescence in glioblastoma surgery. A total of 135 studies were included, comprising 10 reviews, three clinical studies, 10 randomized controlled trials (RCTs), 10 preclinical studies, and four case reports, all focused on NIR fluorescence imaging in glioblastoma surgery. The findings indicate that NIR fluorescence imaging significantly improves tumor visualization, resulting in an 18–22% increase in gross total resection (GTR) rates in clinical studies. NIR fluorescence provides continuous real-time feedback, minimizing repeat imaging, reducing operational costs, and increasing GTR. These improvements contribute to better patient outcomes, including extended progression-free survival, improved overall survival, and reduced postoperative neurological deficits. This review underscores the potential of NIR imaging to establish a new standard for intraoperative glioblastoma management. Full article
(This article belongs to the Section Cancer Causes, Screening and Diagnosis)
Show Figures

Figure 1

Figure 1
<p>Various imaging techniques used in glioblastoma management. (<b>A</b>) CT scan, highlighting tumor necrosis and edema. (<b>B</b>) MRI, providing detailed visualization of the tumor’s structure. (<b>C</b>) PET scan, revealing increased metabolic activity in glioblastoma. (<b>D</b>) 5-ALA fluorescence under NIR light, offering real-time feedback and enhanced visualization of tumor margins during surgery.</p>
Full article ">Figure 2
<p>Setup for NIR-guided surgery.</p>
Full article ">Figure 3
<p>Absorbance and emission spectra of NIR-II fluorescence.</p>
Full article ">Figure 4
<p>PRISMA flowchart.</p>
Full article ">Figure 5
<p>Comparison of gross total resection across different imaging modalities in GB.</p>
Full article ">
15 pages, 16548 KiB  
Article
Short-Wave Infrared (SWIR) Imaging for Robust Material Classification: Overcoming Limitations of Visible Spectrum Data
by Hanbin Song, Sanghyeop Yeo, Youngwan Jin, Incheol Park, Hyeongjin Ju, Yagiz Nalcakan and Shiho Kim
Appl. Sci. 2024, 14(23), 11049; https://doi.org/10.3390/app142311049 - 27 Nov 2024
Viewed by 287
Abstract
This paper presents a novel approach to material classification using short-wave infrared (SWIR) imaging, aimed at applications where differentiating visually similar objects based on material properties is essential, such as in autonomous driving. Traditional vision systems, relying on visible spectrum imaging, struggle to [...] Read more.
This paper presents a novel approach to material classification using short-wave infrared (SWIR) imaging, aimed at applications where differentiating visually similar objects based on material properties is essential, such as in autonomous driving. Traditional vision systems, relying on visible spectrum imaging, struggle to distinguish between objects with similar appearances but different material compositions. Our method leverages SWIR’s distinct reflectance characteristics, particularly for materials containing moisture, and demonstrates a significant improvement in accuracy. Specifically, SWIR data achieved near-perfect classification results with an accuracy of 99% for distinguishing real from artificial objects, compared to 77% with visible spectrum data. In object detection tasks, our SWIR-based model achieved a mean average precision (mAP) of 0.98 for human detection and up to 1.00 for other objects, demonstrating its robustness in reducing false detections. This study underscores SWIR’s potential to enhance object recognition and reduce ambiguity in complex environments, offering a valuable contribution to material-based object recognition in autonomous driving, manufacturing, and beyond. Full article
28 pages, 4077 KiB  
Article
Inter-Sensor Level 1 Radiometric Comparisons Using Deep Convective Clouds
by Louis Rivoire, Sébastien Clerc, Bahjat Alhammoud, Frédéric Romand and Nicolas Lamquin
Remote Sens. 2024, 16(23), 4445; https://doi.org/10.3390/rs16234445 - 27 Nov 2024
Viewed by 238
Abstract
To evaluate the radiometric performance of top-of-atmosphere reflectance images, Deep Convective Clouds (DCCs) can be used as temporally, spatially and spectrally stable targets. The DCCs method has been developed more than 20 years ago and applied recently to Sentinel-2 and Sentinel-3 sensors. In [...] Read more.
To evaluate the radiometric performance of top-of-atmosphere reflectance images, Deep Convective Clouds (DCCs) can be used as temporally, spatially and spectrally stable targets. The DCCs method has been developed more than 20 years ago and applied recently to Sentinel-2 and Sentinel-3 sensors. In this paper, among other developments, we built a new methodology upon those existing by using the bootstrap method and spectral band adjustment factors computed with the Hyper-Spectral Imager (HSI) from the Environmental Mapping and Analysis Program (EnMAP). This methodology is applied to the two Multi-Spectral Imager (MSI) instruments onboard Sentinel-2A and 2B, but also the two Operational Land Imager (OLI) instruments onboard Landsat 8 and 9, from visible wavelength at 442 nm to shortwave-infrared at 2200 nm, using images with a ground resolution spanning from 10 m to 60 m. The results demonstrate the good inter-calibration of MSI units A and B, which are within one percent of relative difference on average between January 2022 and June 2024 for all visible, near-infrared and shortwave-infrared bands, except for the band at 1375 nm for which saturation prevents the use of the method. Similarly, OLI and OLI-2 are found to have a relative difference on the same period lower than one percent for all 30 m resolution bands. Evaluation of the relative difference between the MSI sensors and the OLI sensors with the DCCs method gives values lower than three percent. Finally, these validation results are compared to those obtained with Pseudo-Invariant Calibration Sites (PICSs) over Libya-4: an agreement better than two percent is found between the DCCs and PICSs methods. Full article
Show Figures

Figure 1

Figure 1
<p>DCC detection steps for band B02 at 492 nm in product S2A_MSIL1C_20240505T144731_ N0510_R139_T20NKG_20240505T181400. (<b>a</b>) Raw Sentinel-2A TOA reflectance at 10 m resolution. (<b>b</b>) Subsampled TOA reflectance at 60 m resolution. (<b>c</b>) Subsampled TOA reflectance with detection thresholds applied using bands B8A and B10. (<b>d</b>) TOA reflectance with small DCC clusters removed. (<b>e</b>) TOA reflectance with morphological dilation applied to the DCC mask. (<b>f</b>) Top-Of-DCC reflectance after atmospheric correction has been applied.</p>
Full article ">Figure 2
<p>MSI-A histogram of a DCC in band 9 at 945 nm for all detectors in product S2A_ MSIL1C_20240505T144731_N0510_R139_T20NKG_20240505T181400.</p>
Full article ">Figure 3
<p>Sum of all MSI-A DCC histograms obtained in May 2024 along with the skewed Gaussian distribution fit and the associated first point of inflexion, mode and second point of inflexion.</p>
Full article ">Figure 4
<p>Distributions obtained with the bootstrap method (N equals 20 for readability) applied to MSI-A DCC histograms available in May 2024, with the second point of inflexion of each distribution.</p>
Full article ">Figure 5
<p>Example of DCC observed by EnMAP at 468 nm and used in this study. Product identifier is ENMAP-HSI-L1CDT0000060165_12-2024-02-01T08:30:08.938.</p>
Full article ">Figure 6
<p>Spectral response functions of Sentinel-2 MSI-A, MSI-B, Landsat 8 OLI and Lansat 9 OLI-2 band 3 at 560 nm at their highest resolution available and sampled at the resolution of EnMAP bands. (<b>a</b>) Sentinel-2 MSI-A and MSI-B. (<b>b</b>) Landsat 8 OLI and Lansat 9 OLI-2.</p>
Full article ">Figure 7
<p>Mean DCC spectrum measured with 15 EnMAP products, with Sentinel-2A bands overimposed. Error bars are a measure of the uncertainty in terms of the standard deviation of the mean.</p>
Full article ">Figure 8
<p>Location of products containing DCC pixels for Sentinel-2 units A and B, Landsat 8 and Landsat 9, in August 2023.</p>
Full article ">Figure 9
<p>Temporal evolution of the number of products containing DCC pixels for Sentinel-2 units A and B, Landsat 8 and Landsat 9, between January 2022 and June 2024.</p>
Full article ">Figure 10
<p>Temporal evolution of the reflectance indicators for Sentinel-2 units A and B, Landsat 8 and Landsat 9, between January 2022 and June 2024, for bands centered at 492 nm. (<b>a</b>) represents the temporal evolution of the absolute reflectance indicators while (<b>b</b>) represents the temporal evolution of the relative differences of a reflectance indicator of a sensor with that of Sentinel-2A. SBAFs are applied in (<b>a</b>,<b>b</b>); in (<b>a</b>) SBAFs are applied taking Sentinel-2A as reference. Error bars are a measure of the uncertainty in terms of the standard deviation of the mean. Dashed lines represent one and three percent relative difference with the reference.</p>
Full article ">Figure 11
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 864 nm.</p>
Full article ">Figure 12
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 1610 nm.</p>
Full article ">Figure 13
<p>Sentinel-2A image and associated histogram of a DCC over Barbados in band 10 on the 01/07/2024 (product S2A_MSIL1C_20240701T143751_N0510_R096_T20PRV_20240701T175847). Saturation over the DCC can be seen on the right-hand part of (<b>a</b>). (<b>a</b>) Sentinel-2A image. (<b>b</b>) Histogram of reflectances.</p>
Full article ">Figure 14
<p>Sentinel-2B over Sentinel-2A mean relative differences for the thirteen Sentinel-2 bands, averaged between January 2022 and June 2024; SBAFs are applied. Relative difference at 1375 nm is impacted by saturation. Error bars are a measure of the statistical dispersion of the relative difference over the averaging period. Dashed lines represent one and three percent relative difference with the reference.</p>
Full article ">Figure 15
<p>Landsat 9 over Landsat 8 mean relative differences for eight common bands, averaged between January 2022 and June 2024; SBAFs are applied. Error bars are a measure of the statistical dispersion of the relative difference over the averaging period. Dashed lines represent one and three percent relative difference with the reference.</p>
Full article ">Figure 16
<p>Sentinel-2B, Landsat 8 and Landsat 9 over Sentinel-2A mean relative differences for eight common bands, averaged between January 2022 and June 2024; SBAFs are applied. Relative differences at 1375 nm are impacted by MSI saturation. Error bars are a measure of the statistical dispersion of the relative difference over the averaging period. Dashed lines represent one and three percent relative difference with the reference.</p>
Full article ">Figure 17
<p>Mean Sentinel-2B, Landsat 8 and Landsat 9 over Sentinel-2A relative differences for bands at 442 nm, 492 nm, 560 nm, 665 nm and 864 nm, averaged between January 2022 and December 2023, for the PICSs method and the DCCs method. Error bars are a measure of the statistical dispersion of the relative difference over the averaging period. Dashed lines represent one and three percent relative difference with the reference.</p>
Full article ">Figure 18
<p>Relative differences of Sentinel-2B reflectance indicator over that of Sentinel-2A, for four geographical zones corresponding to strips of latitude and four months spanning over 2023. SBAFs are applied. Error bars are a measure of the uncertainty in terms of the standard deviation of the mean. Dashed lines represent one and three percent relative difference with the reference.</p>
Full article ">Figure A1
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 442 nm.</p>
Full article ">Figure A2
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 560 nm.</p>
Full article ">Figure A3
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 665 nm.</p>
Full article ">Figure A4
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 704 nm, with only Sentinel-2A and Sentinel-2B.</p>
Full article ">Figure A5
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 740 nm, with only Sentinel-2A and Sentinel-2B.</p>
Full article ">Figure A6
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 780 nm, with only Sentinel-2A and Sentinel-2B.</p>
Full article ">Figure A7
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 833 nm, with only Sentinel-2A and Sentinel-2B.</p>
Full article ">Figure A8
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 945 nm, with only Sentinel-2A and Sentinel-2B.</p>
Full article ">Figure A9
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 1375 nm; y-axis scale increased to view error bars.</p>
Full article ">Figure A10
<p>Same as <a href="#remotesensing-16-04445-f010" class="html-fig">Figure 10</a> for bands centred at 2200 nm.</p>
Full article ">
13 pages, 3283 KiB  
Article
Laser Emission at 675 nm: Molecular Counteraction of the Aging Process
by Lorenzo Notari, Laura Pieri, Francesca Cialdai, Irene Fusco, Chiara Risaliti, Francesca Madeddu, Stefano Bacci, Tiziano Zingoni and Monica Monici
Biomedicines 2024, 12(12), 2713; https://doi.org/10.3390/biomedicines12122713 - 27 Nov 2024
Viewed by 311
Abstract
Background/Objectives: Many lasers applied in skin rejuvenation protocols show emissions with wavelengths falling in the red or near-infrared (NIR) bands. To obtain further in vitro data on the potential therapeutic benefits regarding rejuvenation, we employed a 675 nm laser wavelength on cultured human [...] Read more.
Background/Objectives: Many lasers applied in skin rejuvenation protocols show emissions with wavelengths falling in the red or near-infrared (NIR) bands. To obtain further in vitro data on the potential therapeutic benefits regarding rejuvenation, we employed a 675 nm laser wavelength on cultured human dermal fibroblasts to understand the mechanisms involved in the skin rejuvenation process’s signaling pathways by analyzing cytoskeletal proteins, extracellular matrix (ECM) components, and membrane integrins. Methods: Normal human dermal fibroblasts (NHDFs) were irradiated with a 675 nm laser 24 h after seeding, and immunofluorescence microscopy and Western blotting were applied. Results: The results demonstrate that the laser treatment induces significant changes in human dermal fibroblasts, affecting cytoskeleton organization and the production and reorganization of ECM molecules. The cell response to the treatment appears to predominantly involve paxillin-mediated signaling pathways. Conclusions: These changes suggest that laser treatment can potentially improve the structure and function of skin tissue, with interesting implications for treating skin aging. Full article
(This article belongs to the Special Issue Photodynamic Therapy (3rd Edition))
Show Figures

Figure 1

Figure 1
<p>Effects of laser treatment on vimentin (<b>upper panels</b>), F-actin (phalloidin) (<b>middle panels</b>), and alpha-SMA (<b>lower panels</b>) using immunofluorescence analysis. Left: control samples; right: laser-treated samples.</p>
Full article ">Figure 2
<p>Effects of laser treatment on vinculin using immunofluorescence and Western blot analysis. Left: control samples; right: laser-treated samples. Western blot bands of control vs. laser-treated samples and graph showing the bands’ signal intensities (<span class="html-italic">p</span> = 0.001). The used symbol of the asterisk (*) in the figure indicates statistically significant results compared to the control, with <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Effects of laser treatment on paxillin using immunofluorescence and Western blot analysis. Left: control samples; right: laser-treated samples. Western blot bands of control vs. laser-treated samples and graph showing the bands’ signal intensities (<span class="html-italic">p</span> = 0.045). The used symbol of the asterisk (*) in the figure indicates statistically significant results compared to the control, with <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Effects of laser treatment on collagen III using immunofluorescence and Western blot analysis. Left: control samples; right: laser-treated samples. Western blot bands of control vs. laser-treated samples and graph showing the bands’ signal intensities (<span class="html-italic">p</span> = 0.008). The used symbol of the asterisk (*) in the figure indicates statistically significant results compared to the control, with <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Effects of laser treatment on fibronectin using immunofluorescence and Western blot analysis. Left: control samples; right: laser-treated samples. Western blot bands of control vs. laser-treated samples and graph showing the bands’ signal intensities (<span class="html-italic">p</span> = 0.034). The used symbol of the asterisk (*) in the figure indicates statistically significant results compared to the control, with <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Effects of laser treatment on MMP1 using immunofluorescence and Western blot analysis. Left: control samples; right: laser-treated samples. Western blot bands of control vs. laser-treated samples and graph showing bands’ signal intensity (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">
18 pages, 4198 KiB  
Article
Solar-Induced Chlorophyll Fluorescence-Based GPP Estimation and Analysis of Influencing Factors for Xinjiang Vegetation
by Cong Xue, Mei Zan, Yanlian Zhou, Kunyu Li, Jia Zhou, Shunfa Yang and Lili Zhai
Forests 2024, 15(12), 2100; https://doi.org/10.3390/f15122100 - 27 Nov 2024
Viewed by 176
Abstract
With climate change and the intensification of human activity, drought event frequency has increased, affecting the Gross Primary Production (GPP) of terrestrial ecosystems. Accurate estimation of the GPP and in-depth exploration of its response mechanisms to drought are essential for understanding ecosystem stability [...] Read more.
With climate change and the intensification of human activity, drought event frequency has increased, affecting the Gross Primary Production (GPP) of terrestrial ecosystems. Accurate estimation of the GPP and in-depth exploration of its response mechanisms to drought are essential for understanding ecosystem stability and developing strategies for climate change adaptation. Combining remote sensing technology and machine learning is currently the mainstream method for estimating the GPP in terrestrial ecosystems, which can eliminate the uncertainty of model parameters and errors in input data. This study employed extreme gradient boosting, random forest (RF), and light use efficiency models. Additionally, we integrated solar-induced chlorophyll fluorescence (SIF), near-infrared reflectance of vegetation, and the leaf area index (LAI) to construct various GPP estimation models. The standardised precipitation evapotranspiration index (SPEI) was utilised at various timescales to analyse the relationship between the GPP and SPEI during dry years. Moreover, the potential pathways and coefficients of environmental factors that influence GPP were explored using structural equation modelling. Our key findings include the following: (1) the model combining the SIF and RF algorithms exhibits higher accuracy and applicability in estimating vegetation GPP in the arid zone of Xinjiang, with an overall accuracy (MODIS R2) of 0.775; (2) the vegetation in Xinjiang had different response characteristics to different timescales of drought, in which the optimal timescale for GPP to respond to drought was 9 months, with a mean correlation coefficient of 0.244 between grass land GPP and SPEI09, indicating high sensitivity; (3) using structural equation modelling, we found that temperature and precipitation can affect GPP both directly and indirectly through LAI. This study provides a reliable tool for estimating the GPP in Xinjiang, and its methodology and conclusions are important references for similar environments. In addition, this study bridges the research gap in drought response to GPP at different timescales, and the potential influence mechanism of natural factors on GPP provides a scientific basis for early warning of drought and ecosystem management. Further validation using a longer time series is required to confirm the robustness of the model. Full article
(This article belongs to the Section Forest Ecology and Management)
Show Figures

Figure 1

Figure 1
<p>Summary of the research region.</p>
Full article ">Figure 2
<p>Technical framework.</p>
Full article ">Figure 3
<p>Performance evaluation of different models for different vegetation types (red represents RF, blue represents XGboost, and green represents LUE).</p>
Full article ">Figure 4
<p>GPP<sub>RF</sub> and GPP<sub>MODIS</sub> accuracy verification.</p>
Full article ">Figure 5
<p>(<b>a</b>) Trend changes in SPEI12 before and after standardised anomalies (2001–2020); (<b>b</b>) trend changes in mean GOSIF values during the growing season (2009–2011) (the red box represents the most anomalous year, 2009).</p>
Full article ">Figure 6
<p>Spatial distribution of GPP during the growing season (May–October) of 2009.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>e</b>) Spatial distribution of GPP in response to SPEI at different timescales in 2009; (<b>f</b>) correlation coefficients between vegetation GPP and SPEI09 in 2009.</p>
Full article ">Figure 8
<p>Path analysis of the impact of PRE, TEM, LEI, PAR, and FAPAR on GPP (red and blue indicate positive and negative influences, respectively; arrow thickness represents the degree of influence, where the thicker the arrow, the greater the influence. *** represent <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">
13 pages, 7776 KiB  
Communication
Moisture Content Vegetation Seasonal Variability Based on a Multiscale Remote Sensing Approach
by Filippe L. M. Santos, Gonçalo Rodrigues, Miguel Potes, Flavio T. Couto, Maria João Costa, Susana Dias, Maria José Monteiro, Nuno de Almeida Ribeiro and Rui Salgado
Remote Sens. 2024, 16(23), 4434; https://doi.org/10.3390/rs16234434 - 27 Nov 2024
Viewed by 188
Abstract
Water content is one of the most critical characteristics in plant physiological development. Therefore, this information is a crucial factor in determining the water stress conditions of vegetation, which is essential for assessing the wildfire risk and land management decision-making. Remote sensing can [...] Read more.
Water content is one of the most critical characteristics in plant physiological development. Therefore, this information is a crucial factor in determining the water stress conditions of vegetation, which is essential for assessing the wildfire risk and land management decision-making. Remote sensing can be vital for obtaining information over large, limited access areas with global coverage. This is important since conventional techniques for collecting vegetation water content are expensive, time-consuming, and spatially limited. This work aims to evaluate the vegetation live fuel moisture content (LFMC) seasonal variability using a multiscale remote sensing approach, particularly on rockroses, the Cistus ladanifer species, a Western Mediterranean basin native species with wide spatial distribution, over the Herdade da Mitra at the University of Évora, Portugal. This work used four dataset sources, collected monthly between June 2022 and July 2023: (i) Vegetation samples used to calculate the LFMC; (ii) Vegetation reflectance spectral signature using the portable spectroradiometer FieldSpec HandHeld-2 (HH2); (iii) Multispectral optical imagery obtained from the Multispectral Instrument (MSI) sensor onboard the Sentinel-2 satellite; and (iv) Multispectral optical imagery derived from a camera onboard an Unmanned Aerial Vehicle Phantom 4 Multispectral (P4M). Several temporal analyses were performed based on datasets from different sensors and on their intercomparison. Furthermore, the Random Forest (RF) classifier, a machine learning model, was used to estimate the LFMC considering each sensor approach. MSI sensor presented the best results (R2 = 0.94) due to the presence of bands on the Short-Wave Infrared Imagery region. However, despite having information only in the Visible and Near Infrared spectral regions, the HH2 presents promising results (R2 = 0.86). This suggests that by combining these spectral regions with a RF classifier, it is possible to effectively estimate the LFMC. This work shows how different spatial scales, from remote sensing observations, affect the LFMC estimation through machine learning techniques. Full article
Show Figures

Figure 1

Figure 1
<p>Study area: Herdade da Mitra site, Évora (black triangle). The red dots indicate the locations where vegetation samples used in this study were collected.</p>
Full article ">Figure 2
<p>Meteorological variables over the study area: monthly mean air temperature (black), monthly accumulated precipitation (in blue) and monthly mean relative humidity (grey), whereas monthly LFMC (green dots) for the period between January 2022 and July 2023.</p>
Full article ">Figure 3
<p>NDVI maps over the study area obtained from P4M measurements for each fieldwork (the date is indicated in each image).</p>
Full article ">Figure 4
<p>(<b>a</b>) Vegetation spectral signature obtained during the fieldwork campaigns derived from HH2 sensor. (<b>b</b>) Anomaly between the reflectance spectral signature for each date and the average reflectance spectral signature.</p>
Full article ">Figure 5
<p>Spectral reflectance considering different sensors: HH2 (black), MSI (green) and P4M (blue).</p>
Full article ">Figure 6
<p>RF model evaluation between observations and predicted for LFMC values based on (<b>a</b>) HH2, (<b>b</b>) MSI and (<b>c</b>) P4M sensors. The red line denotes a 1:1 relationship.</p>
Full article ">
13 pages, 1076 KiB  
Article
Fruit Sorting Based on Maturity Reduces Internal Disorders in Vapor Heat-Treated ‘B74’ Mango
by Amit Khanal, Muhammad Asad Ullah, Priya Joyce, Neil White, Andrew Macnish, Eleanor Hoffman, Donald Irving, Richard Webb and Daryl Joyce
Horticulturae 2024, 10(12), 1257; https://doi.org/10.3390/horticulturae10121257 - 27 Nov 2024
Viewed by 268
Abstract
Postharvest internal disorders (IDs) in mango fruit present a significant challenge to the industry, with their underlying causes still unclear. This study investigated the relationship between fruit maturity and the susceptibility of vapor heat-treated (VHT) ‘B74’ mangoes to IDs in three experiments. In [...] Read more.
Postharvest internal disorders (IDs) in mango fruit present a significant challenge to the industry, with their underlying causes still unclear. This study investigated the relationship between fruit maturity and the susceptibility of vapor heat-treated (VHT) ‘B74’ mangoes to IDs in three experiments. In the first experiment, fruit were categorized into three maturity groups based on dry matter content (DMC): <15%, 15–17%, and >17%, using a handheld near-infrared device. Half of the fruit in each group underwent VHT, while the remainder were untreated controls. Flesh cavity with white patches (FCWP) was the only disorder observed exclusively in VHT fruit. The incidence and severity of FCWP was significantly higher (p < 0.05) in fruit with <15% DMC, with 12.4% incidence and a severity score of 0.2 on a 0–3 scale (0: healthy and 3: severely affected), compared to more mature fruit. In the second experiment, the fruits were harvested at early and late maturity stages, with average DMC values of 14.5% and 17.4%, respectively. The fruit was subjected to no VHT, VHT, and VHT following a 12 h pre-conditioning period at 37 ± 1 °C. Consistent with the first experiment, FCWP was observed only in VHT fruit, with early-harvested fruit displaying a significantly higher (p < 0.05) FCWP incidence (26.9%) and severity (0.3) compared to late-harvested fruit (8.3% incidence and 0.1 severity). Pre-conditioning significantly reduced FCWP, particularly in early-harvested fruit. In the third experiment, fruit maturity sorted based on density was assessed, followed by VHT and simulated sea freight under controlled (CA) and ambient atmospheres. Fruit density did not effectively differentiate maturity considering DMC as a maturity indicator. Storage conditions significantly reduced (p < 0.05) flesh browning incidence from 71.1% under ambient conditions to 33.3% under CA. This study highlights fruit maturity as a key factor in the susceptibility of ‘B74’ mangoes to postharvest IDs following VHT. Therefore, sorting fruit based on DMC at harvest or at the packing facility prior to VHT serves as a valuable decision support for reducing IDs in VHT fruit. Further research will explore advanced technologies to enable rapid and efficient fruit sorting based on DMC. Full article
(This article belongs to the Special Issue Postharvest Physiology of Horticultural Crops)
Show Figures

Figure 1

Figure 1
<p>Starch iodine staining for starch after VHT in less mature (starch hydrolysis not initiated; <b>left</b>), optimum maturity (starch hydrolysis initiated in the middle of the flesh; <b>middle</b>), and more mature (advanced stage of starch hydrolysis as indicated by yellow flesh; <b>right</b>) fruit. Starch stains black.</p>
Full article ">Figure 2
<p>Starch iodine staining affirming disruption of starch hydrolysis in fruit expressing flesh cavity with white patches. White lesions in flesh after VHT (<b>left</b>) are starchy as confirmed by black starch-iodine staining (<b>right</b>).</p>
Full article ">Figure 3
<p>Starch iodine staining for starch in mango fruit flesh with (starch hydrolysis initiated represented by yellow flesh inside red circle; (<b>right</b>)) and without (starch hydrolysis not initiated; (<b>left</b>)) temperature conditioning of 37 ± 1 °C prior to vapor heat treatment (VHT).</p>
Full article ">
11 pages, 1037 KiB  
Article
Muscle Metabolism During Multiple Muscle Stimulation Using an Affordable Equipment
by Samantha Ye, Sydney Stetter and Kevin K. McCully
J. Funct. Morphol. Kinesiol. 2024, 9(4), 248; https://doi.org/10.3390/jfmk9040248 - 26 Nov 2024
Viewed by 403
Abstract
Background/Objectives: Previous studies have shown that neuromuscular electrical stimulation (NMES), while expensive, can provide some of the health benefits of exercise to people who cannot exercise their legs normally. The aim of this study was to quantify the increases in muscle metabolism [...] Read more.
Background/Objectives: Previous studies have shown that neuromuscular electrical stimulation (NMES), while expensive, can provide some of the health benefits of exercise to people who cannot exercise their legs normally. The aim of this study was to quantify the increases in muscle metabolism in four muscles of the legs of able-bodied individuals with NMES. Methods: Healthy college-aged students were tested. NMES of four muscle groups was performed with inexpensive stimulators and reusable tin foil electrodes. The biceps femoris, vastus lateralis, medial gastrocnemius, and tibialis anterior muscles on one leg were stimulated for ten minutes with twitch stimulations at the highest comfortable stimulation current. Muscle metabolism was measured using the slope of oxygen consumption measured with near-infrared spectroscopy (NIRS) during 5 s of cuff ischemia. Results: Initial studies found fold increases in muscle metabolism above rest of 8.9 ± 8.6 for the vastus lateralis, 7.9 ± 11.9 for the biceps femoris, 6.6 ± 7.8 for the medial gastrocnemius, and 4.9 ± 3.9 for the tibialis anterior. Some participants were able to obtain large increases in muscle metabolism, while other participants had lower increases. Conclusions: The ability to produce large increases in metabolism has the potential to allow NMES to replace or augment exercise to improve health in people who cannot otherwise exercise. The devices used were inexpensive and could be adapted for easy use by a wide range of individuals. Full article
Show Figures

Figure 1

Figure 1
<p>The experimental setup for measurements on the four muscles using near-infrared spectroscopy (NIRS). Aluminum foil electrodes were placed proximally and distally. The black sleeve is an inflatable blood pressure cuff, placed on the upper thigh, for rapid cuff inflation. The participant’s foot rests on the pedal and receives support from a thin pillow placed below the knee.</p>
Full article ">Figure 2
<p>Representative example of the vastus lateralis muscle oxygen saturation during rest, resting arterial occlusions, 5 min neuromuscular electrical stimulation exercises, and end-exercise recovery. A 5 s arterial occlusion is performed after each period of electrical stimulation to determine the relative oxygen levels.</p>
Full article ">Figure 3
<p>(<b>a</b>) Average muscle oxygen saturation for all four muscles throughout the procedure. Values are means (STDEV). Statistical differences were found for all comparisons to rest in this figure; (<b>b</b>) average metabolic rates for all four muscles during the different conditions. The normalized value is calculated as a ratio of the resting slope and the stimulation slope. Values are means (STDEV). * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>A histogram of average increases in mVO2 for 5 and 10 min for the four muscle groups. The target goal of an 8–12-fold increase is shown by the yellow shaded area.</p>
Full article ">
11 pages, 2365 KiB  
Article
Non-Destructive Detection of Pesticide-Treated Baby Leaf Lettuce During Production and Post-Harvest Storage Using Visible and Near-Infrared Spectroscopy
by Dimitrios S. Kasampalis, Pavlos I. Tsouvaltzis and Anastasios S. Siomos
Sensors 2024, 24(23), 7547; https://doi.org/10.3390/s24237547 - 26 Nov 2024
Viewed by 241
Abstract
The market demand for baby leaf lettuce is constantly increasing, while safety has become one of the most important traits in determining consumer preference driven by human health hazards concerns. In this study, the performance of visible and near-infrared (vis/NIR) spectroscopy was tested [...] Read more.
The market demand for baby leaf lettuce is constantly increasing, while safety has become one of the most important traits in determining consumer preference driven by human health hazards concerns. In this study, the performance of visible and near-infrared (vis/NIR) spectroscopy was tested in discriminating pesticide-free against pesticide-treated lettuce plants. Two commercial fungicides (mancozeb and fosetyl-al) and two insecticides (deltamethrin and imidacloprid) were applied as spray solutions at the recommended rates on baby leaf lettuce plants. Untreated-control plants were sprayed with water. Reflectance data in the wavelength range 400–2500 nm were captured on leaf samples until harvest on the 10th day upon pesticide application, as well as after 4 and 8 days during post-harvest storage at 5 °C. In addition, biochemical components in leaf tissue were also determined during storage, such as antioxidant enzymes’ activities (peroxidase [POD], catalase [CAT], and ascorbate peroxidase [APX]), along with malondialdehyde [MDA] and hydrogen peroxide [H2O2] content. Partial least square discriminant analysis (PLSDA) combined with feature-selection techniques was implemented, in order to classify baby lettuce tissue into pesticide-free or pesticide-treated ones. The genetic algorithm (GA) and the variable importance in projection (VIP) scores identified eleven distinct regions and nine specific wavelengths that exhibited the most significant effect in the detection models, with most of them in the near-infrared region of the electromagnetic spectrum. According to the results, the classification accuracy of discriminating pesticide-treated against non-treated lettuce leaves ranged from 94% to 99% in both pre-harvest and post-harvest periods. Although there were no significant differences in enzyme activities or H2O2, the MDA content in pesticide-treated tissue was greater than in untreated ones, implying that the chemical spray application probably induced a stress response in the plant that was disclosed with the reflected energy. In conclusion, vis/NIR spectroscopy appears as a promising, reliable, rapid, and non-destructive tool in distinguishing pesticide-free from pesticide-treated lettuce products. Full article
(This article belongs to the Section Chemical Sensors)
Show Figures

Figure 1

Figure 1
<p>Classification rate (%) of pesticide-free and pesticide-treated baby lettuce leaves based on reflectance spectra data (340–2500 nm) within each day of pre-harvest production or postharvet storage, as well as average means for the whole period upon pooling the data of all individual days.</p>
Full article ">Figure 2
<p>Spectra reflectance (%) of pesticide-free (blue line) and pesticide-treated (red line) baby lettuce leaves in the vis-NIR part (340–2500 nm) as average means for the whole period upon pooling the data captured in all individual days. The eleven green areas represent the parts of the spectrum that exhibited the most significant effect on the partial least squares discrimination analysis classifier and were detected using the genetic algorithm (GA).</p>
Full article ">Figure 3
<p>The variable importance in projection scores (VIP) in the vis/NIR part (340–2500 nm), which represents the individual effect of each wavelength on the partial least squares discrimination analysis classifier. The vertical green lines correspond to the wavelengths with the highest VIP scores. The red dot line corresponds to the lowest limit above which a wavelength exhibits a significant effect in the discriminant analysis algorithm.</p>
Full article ">Figure 4
<p>Classification rate (%) of pesticide-free and pesticide-treated baby lettuce leaves based on the reflectance spectra data at 377, 517, 689, 959, 994, 1361, 1390, 1875, and 2177 nm that were selected using the VIP scores analysis, within each day of pre-harvest production or post-harvest storage, as well as average for the whole period upon pooling the data captured in all individual days.</p>
Full article ">
23 pages, 15560 KiB  
Article
Surface Modification of Gold Nanorods (GNRDs) Using Double Thermo-Responsive Block Copolymers: Evaluation of Self-Assembly and Stability of Nanohybrids
by Jesús E. Márquez-Castro, Angel Licea-Claverie, Carlos Guerrero-Sánchez and Eugenio R. Méndez
Polymers 2024, 16(23), 3293; https://doi.org/10.3390/polym16233293 - 26 Nov 2024
Viewed by 334
Abstract
A series of copolymers containing a thermo-responsive biocompatible first block of poly[di(ethylene glycol) methyl ether methacrylate)-co-(oligo(ethylene glycol) methyl ether methacrylate], P(DEGMA-co-OEGMA) were chain-extended to incorporate either poly(N-isopropylacrylamide), PNIPAAm or poly(N-isopropylacrylamide-co-butyl acrylate), P(NIPAAm-co-BA) as [...] Read more.
A series of copolymers containing a thermo-responsive biocompatible first block of poly[di(ethylene glycol) methyl ether methacrylate)-co-(oligo(ethylene glycol) methyl ether methacrylate], P(DEGMA-co-OEGMA) were chain-extended to incorporate either poly(N-isopropylacrylamide), PNIPAAm or poly(N-isopropylacrylamide-co-butyl acrylate), P(NIPAAm-co-BA) as second thermo-responsive block using reversible addition–fragmentation chain transfer (RAFT) polymerization. P(DEGMA-co-OEGMA)-b-PNIPAAm copolymers showed two response temperatures at 33 and 43 °C in an aqueous solution forming stable aggregates at 37 °C. In contrast, P(DEGMA-co-OEGMA)-b-P(NIPAAm-co-BA) copolymers showed aggregation below room temperature due to the shift in response temperature provoked by the presence of hydrophobic butyl acrylate (BA) units, and shrinkage upon heating up to body temperature, while maintaining the second response temperature above 40 °C. The terminal trithiocarbonate group of the block copolymers was modified to a thiol functionality and used to stabilize gold nanorods (GNRDs) via the “grafting to” approach. The Localized Surface Plasmon Resonance (LSPR) absorption band of GNRDs with an aspect ratio of 3.9 (length/diameter) was located at 820 nm after surface grafting with block copolymers showing a hydrodynamic diameter of 160 nm at 37 °C. On the other hand, the stability of the P(DEGMA-co-OEGMA)-b-PNIPAAm@GNRDs and P(DEGMA-co-OEGMA)-b-P(NIPAAm-co-BA)@GNRDs nanohybrids was monitored for 8 days; where the LSPR absorption band did not shift or show any broadening. Aqueous dispersed nanohybrids were irradiated with a near-infrared laser (300 mW), where the temperature of the surroundings increased 16 °C after 16 min, where conditions for no precipitation were determined. These tailored temperature-responsive nanohybrids represent interesting candidates to develop drug nanocarriers for photo-thermal therapies. Full article
(This article belongs to the Special Issue State-of-the-Art Polymer Science and Technology in Mexico)
Show Figures

Figure 1

Figure 1
<p><sup>1</sup>H-NMR spectrum (400 MHz) in CDCl<sub>3</sub> of the P(DEGMA-<span class="html-italic">co</span>-OEGMA) copolymer P2.</p>
Full article ">Figure 2
<p>Analysis of <span class="html-italic">T<sub>cp</sub></span> of aqueous solutions of P(DEGMA-<span class="html-italic">co</span>-OEGMA) copolymers (inserted pictures correspond to sample P(DEGMA<sub>70%</sub>-<span class="html-italic">co</span>-OEGMA<sub>30%</sub>), P3.</p>
Full article ">Figure 3
<p><sup>1</sup>H-NMR spectra (400 MHz) in CDCl<sub>3</sub> of block copolymers: (<b>a</b>) P(DEGMA-<span class="html-italic">co</span>-OEGMA)-<span class="html-italic">b</span>-PNIPAAm (P2-2), (<b>b</b>) P(DEGMA-<span class="html-italic">co</span>-OEGMA)-<span class="html-italic">b</span>-P(NIPAAm-<span class="html-italic">co</span>-BA) (P2-3). Colors of polymeric structures correspond to <a href="#polymers-16-03293-sch001" class="html-scheme">Scheme 1</a> and Figure 9c.</p>
Full article ">Figure 4
<p>Analysis of size change in aqueous media and size distribution at different temperatures for block copolymers (1 mg mL<sup>−1</sup>): (<b>a</b>,<b>b</b>) P(DEGMA-<span class="html-italic">co</span>-OEGMA)<sub>37%</sub>-<span class="html-italic">b</span>-PNIPAAm<sub>63%</sub>; (<b>c</b>,<b>d</b>) P(DEGMA-<span class="html-italic">co</span>-OEGMA)<sub>48%</sub>-<span class="html-italic">b</span>-P(NIPAAm<sub>43%</sub>-<span class="html-italic">co</span>-BA<sub>9%</sub>).</p>
Full article ">Figure 5
<p>(<b>a</b>) UV-vis spectra (c = 1 mg mL<sup>−1</sup> in ethanol), and (<b>b</b>) SEC traces in DMF for P(DEGMA-<span class="html-italic">co</span>-OEGMA)<sub>37%</sub>-<span class="html-italic">b</span>-PNIPAAm<sub>63%</sub> (<span class="html-italic">M</span><sub>n</sub> = 32,600 g mol<sup>−1</sup>, <span class="html-italic">Ð</span> = 1.09) and P(DEGMA-<span class="html-italic">co</span>-OEGMA)<sub>37%</sub>-<span class="html-italic">b</span>-PNIPAAm<sub>63%</sub>-SH (<span class="html-italic">M</span><sub>n</sub> = 35,800 g mol<sup>−1</sup>, <span class="html-italic">Ð</span> = 1.11).</p>
Full article ">Figure 6
<p>Characterization of the synthesized GNRDs featuring an aspect ratio of 3.98: (<b>a</b>) UV-Vis spectra with and without CTAB; (<b>b</b>) TEM micrograph of GNRDs dispersed after removing CTAB; (<b>c</b>) UV-Vis spectra of the block copolymer@GNRDs nanohybrids in an aqueous medium.</p>
Full article ">Figure 7
<p>Characterization of P(EGMA-<span class="html-italic">co</span>-OEGMA)<sub>37%</sub>-<span class="html-italic">b</span>-PNIPAAm<sub>63%</sub>@GNRDs nanohybrid: (<b>a</b>) UV-Vis spectra of dispersions monitored for 8 days at room temperature, (<b>b</b>) Representative TEM micrographs of block copolymer@GNRDs in dry state (as stained with uranyl acetate).</p>
Full article ">Figure 8
<p>Size distributions of P(EGMA-<span class="html-italic">co</span>-OEGMA)<sub>37%</sub>-<span class="html-italic">b</span>-PNIPAAm<sub>63%</sub>@GNRDs nanohybrids in an aqueous medium at different temperatures: (<b>a</b>) 25 °C; (<b>b</b>) 27 °C; (<b>c</b>) 40 °C; (<b>d</b>) self-assembly behavior of the nanohybrid in an aqueous medium upon increasing temperature; (<b>e</b>) schematic representation of the self-assembly behavior. Colors of polymeric structures correspond to <a href="#polymers-16-03293-sch001" class="html-scheme">Scheme 1</a>.</p>
Full article ">Figure 9
<p>Size distributions of block copolymers@GNRDs nanohybrids: (<b>a</b>) at 25 °C; (<b>b</b>) at 37 °C; (<b>c</b>) schematic representation of the behavior of the block copolymer@GNRDs nanohybrids in the temperature range between 25 and 37 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) Evolution of size vs. temperature for different block copolymer@GNRDs nanohybrids in an aqueous medium; (<b>b</b>) UV-Vis absorption spectra of P(EGMA-<span class="html-italic">co</span>-OEGMA)<sub>37%</sub>-<span class="html-italic">b</span>-PNIPAAm<sub>63%</sub>@GNRDs in an aqueous medium at 25, 37 and 40 °C.</p>
Full article ">Figure 11
<p>(<b>a</b>) Surrounding temperature of block copolymer@GNRDs and GNRDs aqueous dispersions vs. NIR-irradiation time (at 795 nm, 300 mW); (<b>b</b>) TGA thermograms of the investigated block copolymer@GNRDs.</p>
Full article ">Scheme 1
<p>Schematic illustration of the preparation of P(DEGMA-<span class="html-italic">co</span>-OEGMA)-<span class="html-italic">b</span>-PNIPAAm@GNRDs nanohybrids.</p>
Full article ">Scheme 2
<p>Synthesis route for P(DEGMA-<span class="html-italic">co</span>-OEGMA) copolymers vía RAFT polymerization.</p>
Full article ">Scheme 3
<p>Schematic representation of the synthesis route of P(DEGMA-<span class="html-italic">co</span>-OEGMA)-<span class="html-italic">b</span>-PNIPAAm and P(DEGMA-<span class="html-italic">co</span>-OEGMA)-<span class="html-italic">b</span>-P(NIPAAm-<span class="html-italic">co</span>-BA) block copolymers via RAFT polymerization. Colors of polymeric structures correspond to <a href="#polymers-16-03293-sch001" class="html-scheme">Scheme 1</a> and Figure 9c.</p>
Full article ">Scheme 4
<p>Schematic representation of the preparation of thiol-terminated block copolymers via aminolysis of trithiocarbonates. Colors of polymeric structures correspond to <a href="#polymers-16-03293-sch001" class="html-scheme">Scheme 1</a> and Figure 9c.</p>
Full article ">
Back to TopTop