[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
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
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
remove_circle_outline

Search Results (7,594)

Search Parameters:
Keywords = pigmentation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 1843 KiB  
Article
Responses of the Allium cepa L. to Heavy Metals from Contaminated Soil
by Ocsana Opriș, Ildiko Lung, Katalin Gméling, Adina Stegarescu, Noémi Buczkó, Otilia Culicov and Maria-Loredana Soran
Plants 2024, 13(20), 2913; https://doi.org/10.3390/plants13202913 (registering DOI) - 17 Oct 2024
Abstract
Heavy metals can accumulate and migrate in soil environments and can negatively affect crops and consumers. Because an increased incidence of chronic diseases can be observed, food security has become a high-priority concern. In the present work, we evaluate the impact of heavy [...] Read more.
Heavy metals can accumulate and migrate in soil environments and can negatively affect crops and consumers. Because an increased incidence of chronic diseases can be observed, food security has become a high-priority concern. In the present work, we evaluate the impact of heavy metals on bioactive compounds and elemental content from onions. Plants were grown in the absence and presence of various concentrations of heavy metal salts (Pb, Mn, Cu, Zn, Ni and Cd). The influence of heavy metal salts on onions was evaluated by analyzing the content of bioactive compounds, antioxidant capacity, and elemental content. The variation of assimilatory pigments, total polyphenols content, and antioxidant capacity increased or decreased depending on the heavy metal added to the soil as well as on the amount added. Regarding the amount of bioactive compounds, it increased between 6.79 and 34.39% or decreased between 4.68 and 62.42%. The content of ten elements in plants was reported, as well as elemental mutual correlation and correlation of element content with biologically active compounds and antioxidant capacity. Full article
(This article belongs to the Special Issue In Vivo and In Vitro Studies on Heavy Metal Tolerance in Plants)
Show Figures

Figure 1

Figure 1
<p>The assimilatory pigments (chlorophyll <span class="html-italic">a</span>—CHL<span class="html-italic">a</span>; chlorophyll <span class="html-italic">b</span>—CHL<span class="html-italic">b</span>; carotenoids–CARO) content expressed as mg pigment/g fresh weight (FW) of <span class="html-italic">Allium cepa</span> L. (onion tails) extracts grown in the presence of heavy metals (Ni, Pb, Mn, Cu, Cd and Zn) at different concentrations compared to control onions. I—concentration under accepted limit, II—maximum accepted limit and III—above maximum accepted limit. These concentrations correspond as follows: Ni (15, 30, 75 mg kg<sup>−1</sup>); Pb (15, 30, 50 mg kg<sup>−1</sup>); Mn (675, 1350, 1500 mg kg<sup>−1</sup>); Cu (15, 30, 100 mg kg<sup>−1</sup>); Cd (0.75, 1.5, 3 mg kg<sup>−1</sup>); Zn (75, 150, 300 mg kg<sup>−1</sup>), a, b, c, d—the statistically considerable differences among groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Total polyphenols content expressed as mg gallic acid (GA)/g fresh weight (FW) of <span class="html-italic">Allium cepa</span> L. (onion tails) extracts grown in the presence of heavy metals at different concentrations (I—concentration under accepted limit, II—maximum accepted limit and III—above maximum accepted limit) compared to control onions. The heavy metal concentrations correspond as follows: Ni (15, 30, 75 mg kg<sup>−1</sup>); Pb (15, 30, 50 mg kg<sup>−1</sup>); Mn (675, 1350, 1500 mg kg<sup>−1</sup>); Cu (15, 30, 100 mg kg<sup>−1</sup>); Cd (0.75, 1.5, 3 mg kg<sup>−1</sup>); Zn (75, 150, 300 mg kg<sup>−1</sup>), a, b—the statistically considerable differences among groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Antioxidant capacity expressed as mM Trolox/g fresh weight (FW) of <span class="html-italic">Allium cepa</span> L. (onion tails) extracts grown in the presence of heavy metals at different concentrations (I—concentration under accepted limit, II—maximum accepted limit and III—above maximum accepted limit) compared to control onions. The heavy metal concentrations correspond as follows: Ni (15, 30, 75 mg kg<sup>−1</sup>); Pb (15, 30, 50 mg kg<sup>−1</sup>); Mn (675, 1350, 1500 mg kg<sup>−1</sup>); Cu (15, 30, 100 mg kg<sup>−1</sup>); Cd (0.75, 1.5, 3 mg kg<sup>−1</sup>); Zn (75, 150, 300 mg kg<sup>−1</sup>), a, b, c—the statistically considerable differences among groups (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Comparative diagram of the element (Ca, K, Br, Co, Fe, Na, Rb, Zn and Au) content (mg/g) in <span class="html-italic">Allium cepa</span> L. (onion tails) grown in the presence of heavy metals at different concentrations (I—concentration under accepted limit, II—maximum accepted limit and III—above maximum accepted limit) compared to control onions. The heavy metal concentrations correspond as follows: Ni (15, 30, 75 mg kg<sup>−1</sup>); Pb (15, 30, 50 mg kg<sup>−1</sup>); Mn (675, 1350, 1500 mg kg<sup>−1</sup>); Cu (15, 30, 100 mg kg<sup>−1</sup>); Cd (0.75, 1.5, 3 mg kg<sup>−1</sup>); Zn (75, 150, 300 mg kg<sup>−1</sup>).</p>
Full article ">
17 pages, 3569 KiB  
Article
A Cippus from Turris Libisonis: Evidence for the Use of Local Materials in Roman Painting on Stone in Northern Sardinia
by Roberta Iannaccone, Stefano Giuliani, Sara Lenzi, Matteo M. N. Franceschini, Silvia Vettori and Barbara Salvadori
Minerals 2024, 14(10), 1040; https://doi.org/10.3390/min14101040 (registering DOI) - 17 Oct 2024
Abstract
The ancient Roman town of Turris Libisonis was located on the northern coast of Sardinia and was known in the past as an important naval port. Located in the Gulf of Asinara, it was a Roman colony from the 1st century BCE and [...] Read more.
The ancient Roman town of Turris Libisonis was located on the northern coast of Sardinia and was known in the past as an important naval port. Located in the Gulf of Asinara, it was a Roman colony from the 1st century BCE and became one of the richest towns on the island. Among the archaeological finds in the area, the cippus exhibited in the Antiquarium Turritano is of great interest for its well-preserved traces of polychromy. The artefact dates back to the early Imperial Age and could have had a funerary or votive function. The artefact was first examined using a portable and non-invasive protocol involving multi-band imaging (MBI), portable X-ray fluorescence (p-XRF), portable FT-IR in external reflectance mode (ER FT-IR) and Raman spectroscopy. After this initial examination, a few microfragments were collected and investigated by optical microscopy (OM), X-ray powder diffraction (XRPD), Fourier-transform infrared spectroscopy in ATR mode (ATR FT-IR) and micro-ATR mode (μATR FT-IR) and Scanning Electron Microscopy/Energy Dispersive Spectroscopy (SEM-EDS) to improve our knowledge and characterize the materials and to determine their provenience. The results contribute to a better understanding of the provenance of materials and shed light on pigments on stone and their use outside the Italian peninsula and, in particular, Roman Sardinia. Full article
(This article belongs to the Special Issue Geomaterials and Cultural Heritage)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The position of the cippus in the Antiquarium of Porto Torres (SS) and the two sides analyzed: (<b>b</b>) side A; (<b>c</b>) side B (courtesy of Ministero della Cultura–Direzione Regionale Musei Sardegna).</p>
Full article ">Figure 2
<p>(<b>a</b>) Raking light detail of flaking areas; (<b>b</b>) High magnification image (60×) of the area in the red square; (<b>c</b>) SEM image in BSE from the sample (red point) and (<b>d</b>) EDS analysis results of a point on recrystallized salt.</p>
Full article ">Figure 3
<p>(<b>a</b>) Raman spectrum of point corresponding to (<b>a</b>) a black area and (<b>b</b>) a yellow area, respectively. On the right, the optical microscopic details of the points (60×). In (<b>b</b>) Raman spectrum of yellow area, in black, and goethite reference spectrum in yellow (RRUFF mineral database).</p>
Full article ">Figure 4
<p>(<b>a</b>) p-XRF spectra of point 8 in black, point 10 in dotted black and the reference background in red; (<b>b</b>) the locations of measured points 8 and 10 are shown (<b>c</b>) Raman spectra of green earth pigment at point 8 (in green) and point 10 (in gray).</p>
Full article ">Figure 5
<p>Microphotographs of thin sections of the carbonate rock of the cippus (<b>a</b>,<b>b</b>) and of the mortar covering the cippus (<b>c</b>,<b>d</b>) (using a polarized light microscope): (<b>a</b>,<b>c</b>) parallel nicols; (<b>b</b>,<b>d</b>) crossed nicols.</p>
Full article ">Figure 6
<p>Geological map of Porto Torres and its surrounding area. Modified from [<a href="#B46-minerals-14-01040" class="html-bibr">46</a>].</p>
Full article ">Figure 7
<p>Macro-photo of carbonate rock of the cippus. Various fossils were observed in the carbonatic rock of the cippus: (<b>a</b>) gastropods; (<b>b</b>) ammonites; (<b>c</b>) algae tallus.</p>
Full article ">Figure 8
<p>Fragment analyzed by SEM−EDS. (<b>a</b>) Optical microscope image at 50×; (<b>b</b>) backscattered details at 127×; and (<b>c</b>) EDS analysis results.</p>
Full article ">Figure 9
<p>(<b>a</b>) FT-IR micro-ATR spectrum obtained from the green sample using point analysis with a TE-MCT detector and (<b>b</b>) FT-IR FPA-ATR spectrum extracted from the chemical map shown above. The arrow indicates the position of the spectrum.</p>
Full article ">
14 pages, 7233 KiB  
Article
Facile Synthesis of Low-Dimensional and Mild-Alkaline Magnesium Carbonate Hydrate for Safe Multiple Protection of Paper Relics
by Yi Wang, Zirui Zhu, Jinhua Wang, Peng Liu, Xingxiang Ji, Hongbin Zhang and Yi Tang
Molecules 2024, 29(20), 4921; https://doi.org/10.3390/molecules29204921 - 17 Oct 2024
Abstract
Paper-based cultural relics inevitably face a variety of diseases such as acidification, yellowing, and strength loss during long-term preservation, where weakly alkaline inorganic materials play an important role in their deacidification treatments. In this work, by simply adjusting the supersaturation of crystal growing [...] Read more.
Paper-based cultural relics inevitably face a variety of diseases such as acidification, yellowing, and strength loss during long-term preservation, where weakly alkaline inorganic materials play an important role in their deacidification treatments. In this work, by simply adjusting the supersaturation of crystal growing solution without the use of any organic additives, one-dimensional (1D) and two-dimensional (2D) weakly alkaline materials—magnesium carbonate hydrates (MCHs)—were controllably synthesized. It is worth noting that the coatings of 1D/2D MCHs not only cause little change in chromatic aberration and water wettability, but also ensure their safety for alkali-sensitive pigments. Meanwhile, the deacidification, anti-aging, strength-enhancing, and flame-retardant effects of these materials have been tested on ancient book papers, all of which achieved good protective effects. In contrast, 1D MCH materials brought about significant enhancement in both mechanical strengths and flame-retardant effects, and the related effects were investigated. Based on this facile micromorphology control strategy, more low-dimensional nanomaterials are expected to be synthesized by design for the protection of paper-based relics, which will expand our knowledge on functional deacidification and protection mechanisms. Full article
(This article belongs to the Special Issue Chemical Conservation of Paper-Based Cultural Heritage)
25 pages, 1749 KiB  
Article
Exploration of the Bioactivity of Pigmented Extracts from Streptomyces Strains Isolated Along the Banks of the Guaviare and Arauca Rivers (Colombia)
by Aixa A. Sarmiento-Tovar, Sara J. Prada-Rubio, Juliana Gonzalez-Ronseria, Ericsson Coy-Barrera and Luis Diaz
Fermentation 2024, 10(10), 529; https://doi.org/10.3390/fermentation10100529 - 17 Oct 2024
Abstract
Pigments are chemical compounds that impart color through mechanisms such as absorption, reflection, and refraction. While traditional natural pigments are derived from plant and insect tissues, microorganisms, including bacteria, yeasts, algae, and filamentous fungi, have emerged as promising sources for pigment production. In [...] Read more.
Pigments are chemical compounds that impart color through mechanisms such as absorption, reflection, and refraction. While traditional natural pigments are derived from plant and insect tissues, microorganisms, including bacteria, yeasts, algae, and filamentous fungi, have emerged as promising sources for pigment production. In this study, we focused on pigment production by 20 Streptomyces isolates from our in-house actinobacteria strain collection, sourced from the Guaviare and Arauca Rivers in Colombia. The isolates were identified via 16S rRNA gene sequencing, and the bioactivities—including antioxidant, antibacterial, and cytotoxic properties—of their extracts obtained across four different culture media were assessed. Promising pigmented hydroalcoholic extracts demonstrating these bioactivities were further analyzed using LC-MS, leading to the annotation of a variety of pigment-related compounds. This study revealed that culture media significantly influenced both pigment production and bioactivity outcomes. Notably, anthraquinones, phenazines, and naphthoquinones were predominant pigment classes associated with cytotoxic and antimicrobial activities, while carotenoids were linked to antioxidant effects. For instance, S. murinus 4C171 produced various compounds exhibiting both cytotoxic and antioxidant activities. These findings highlighted a growth medium-dependent effect, as pigment production, coloration, and bioactivity outcomes were influenced by growth media. These results demonstrate the significant potential of Streptomyces isolates as sources of bioactive pigments for diverse applications. Full article
(This article belongs to the Special Issue Pigment Production in Submerged Fermentation: Second Edition)
22 pages, 3101 KiB  
Article
Optimized Proportioning Techniques and Roadway Performance Evaluation of Colored Asphalt Pavement Materials
by Silin Fan, Shaopeng Zheng, Jian Ma, Liangliang Chen, Xiao Li and Cheng Cheng
Sustainability 2024, 16(20), 8996; https://doi.org/10.3390/su16208996 - 17 Oct 2024
Abstract
This study systematically investigated the formulation optimization, performance evaluation, and practical application of epoxy-based composite materials for colored asphalt pavement. By conducting comprehensive experiments, we optimized the composition of epoxy-based composites, verifying their excellent bonding performance, good heat resistance, and UV aging resistance [...] Read more.
This study systematically investigated the formulation optimization, performance evaluation, and practical application of epoxy-based composite materials for colored asphalt pavement. By conducting comprehensive experiments, we optimized the composition of epoxy-based composites, verifying their excellent bonding performance, good heat resistance, and UV aging resistance under various temperature conditions. The key optimized component ratios were determined as a 1:1 blend of Type I and Type II epoxy resins, 30 phr of curing agent, 10 phr of toughening agent, 5 phr of diluent, 10% filler, 12% flame retardant, and 10% pigment. At the recommended dosage of 2.0 kg/m2 of epoxy binder, the composite structure exhibited the best reinforcement effect, improving low-temperature performance significantly. Compared to ordinary asphalt mixtures, the colored pavement composite structure showed superior mechanical strength, deformation capacity, high-temperature stability (dynamic stability approximately three times higher), and water stability (TSR values up to 95.5%). Furthermore, its fatigue life decay rate was significantly lower, with fatigue limit loading frequencies more than three times those of ordinary asphalt mixtures, demonstrating excellent fatigue resistance. This study provides strong technical support and a theoretical basis for the development and practical application of colored asphalt pavement. Full article
Show Figures

Figure 1

Figure 1
<p>Relationship between curing agent content and tensile properties.</p>
Full article ">Figure 2
<p>Relationship between the amount of toughener and tensile properties.</p>
Full article ">Figure 3
<p>Relationship between the amount of diluent and tensile properties and viscosity.</p>
Full article ">Figure 4
<p>The relationship between the amount of flame retardant and tensile properties.</p>
Full article ">Figure 5
<p>Dumbbell-shaped specimen after pouring.</p>
Full article ">Figure 6
<p>Tensile experiment.</p>
Full article ">Figure 7
<p>Results of epoxy binder pull-out tests with different paving amounts.</p>
Full article ">Figure 8
<p>Tensile and shear test results of epoxy binders at different curing temperatures.</p>
Full article ">Figure 9
<p>Rutting test results.</p>
Full article ">Figure 10
<p>Comparison of load stress between control group and color pavement fatigue test.</p>
Full article ">
22 pages, 3096 KiB  
Article
Ascophyllum nodosum Extract Improves Olive Performance Under Water Deficit Through the Modulation of Molecular and Physiological Processes
by Maria Celeste Dias, Rui Figueiras, Marta Sousa, Márcia Araújo, José Miguel P. Ferreira de Oliveira, Diana C. G. A. Pinto, Artur M. S. Silva and Conceição Santos
Plants 2024, 13(20), 2908; https://doi.org/10.3390/plants13202908 - 17 Oct 2024
Abstract
The olive tree is well adapted to the Mediterranean climate, but how orchards based on intensive practices will respond to increasing drought is unknown. This study aimed to determine if the application of a commercial biostimulant improves olive tolerance to drought. Potted plants [...] Read more.
The olive tree is well adapted to the Mediterranean climate, but how orchards based on intensive practices will respond to increasing drought is unknown. This study aimed to determine if the application of a commercial biostimulant improves olive tolerance to drought. Potted plants (cultivars Arbequina and Galega) were pre-treated with an extract of Ascophyllum nodosum (four applications, 200 mL of 0.50 g/L extract per plant), and were then well irrigated (100% field capacity) or exposed to water deficit (50% field capacity) for 69 days. Plant height, photosynthesis, water status, pigments, lipophilic compounds, and the expression of stress protective genes (OeDHN1—protective proteins’ dehydrin; OePIP1.1—aquaporin; and OeHSP18.3—heat shock proteins) were analyzed. Water deficit negatively affected olive physiology, but the biostimulant mitigated these damages through the modulation of molecular and physiological processes according to the cultivar and irrigation. A. nodosum benefits were more expressive under water deficit, particularly in Galega, promoting height (increase of 15%) and photosynthesis (increase of 34%), modulating the stomatal aperture through the regulation of OePIP1.1 expression, and keeping OeDHN1 and OeHSP18.3 upregulated to strengthen stress protection. In both cultivars, biostimulant promoted carbohydrate accumulation and intrinsic water-use efficiency (iWUE). Under good irrigation, biostimulant increased energy availability and iWUE in Galega. These data highlight the potential of this biostimulant to improve olive performance, providing higher tolerance to overcome climate change scenarios. The use of this biostimulant can improve the establishment of younger olive trees in the field, strengthen the plant’s capacity to withstand field stresses, and lead to higher growth and crop productivity. Full article
(This article belongs to the Special Issue Drought Responses and Adaptation Mechanisms in Plants)
Show Figures

Figure 1

Figure 1
<p>Leaf relative water content (RWC) (<b>A</b>,<b>B</b>) and plant height increment (<b>C</b>,<b>D</b>) in <span class="html-italic">O. europaea</span> plants of the treatments C (well-watered), BC (biostimulant + well-watered), S (water deficit), and BS (biostimulant + water deficit). Bars represent mean ± standard error (<span class="html-italic">n</span> = 5–10). The effect of the factor irrigation (I), factor biostimulant (B), and the interaction between the factor irrigation and biostimulant (I × B) are presented, and when the effect of each factor or the interaction is statistically significant (<span class="html-italic">p</span> ≤ 0.05), it appears in bold. Different letters indicate statistically significant differences (<span class="html-italic">p</span> ≤ 0.05). Significant differences among I × B refer to differences between C, BC, S, and BS treatments. Significant differences among the factor I refer to differences between 100% (treatments C and BC) and 50% irrigation (treatments S and BS). Significant differences among the factor B are shown in a chart at the top of the corresponding graph, and statistic letters refer to treatments without biostimulant (0: treatments C and S) and treatments with biostimulant (B: treatments BC and BS).</p>
Full article ">Figure 2
<p>Net CO<sub>2</sub>-assimilation rate (<b>A</b>,<b>B</b>), stomatal conductance (<b>C</b>,<b>D</b>), ratio of intercellular CO<sub>2</sub> and extracellular CO<sub>2</sub> concentration (Ci/Ca) (<b>E</b>,<b>F</b>), and intrinsic water-use efficiency (<b>G</b>,<b>H</b>) in <span class="html-italic">O. europaea</span> plants of the treatments C (well-watered), BC (biostimulant + well-watered), S (water deficit), and BS (biostimulant + water deficit). Bars represent mean ± standard error (<span class="html-italic">n</span> = 6–9). The effect of the factor irrigation (I), factor biostimulant (B), and the interaction between the factor irrigation and biostimulant (I × B) are presented, and when the effect of each factor or the interaction is statistically significant (<span class="html-italic">p</span> ≤ 0.05), it appears in bold. Different letters indicate statistically significant differences (<span class="html-italic">p</span> ≤ 0.05). Significant differences among I × B refer to differences between C, BC, S, and BS treatments. Significant differences among the factor I refer to differences between 100% (treatments C and BC) and 50% irrigation (treatments S and BS). Significant differences among the factor B are shown in a chart at the top of the corresponding graph, and statistic letters refer to treatments without biostimulant (0: treatments C and S) and treatments with biostimulant (B: treatments BC and BS).</p>
Full article ">Figure 3
<p>Maximum efficiency of PSII (F<sub>v</sub>/F<sub>m</sub>) (<b>A</b>,<b>B</b>), effective efficiency of PSII (Φ<sub>PSII</sub>) (<b>C</b>,<b>D</b>), efficiency of excitation energy capture by open PSII reaction centers (F<sub>v</sub>′/F<sub>m</sub>′) (<b>E</b>,<b>F</b>), photochemical quenching (qP) (<b>G</b>,<b>H</b>), and non-photochemical quenching (NPQ) (<b>I</b>,<b>J</b>) in <span class="html-italic">O. europaea</span> plants of the treatments C (well-watered), BC (biostimulant + well-watered), S (water deficit), and BS (biostimulant + water deficit). Bars represent mean ± standard error (<span class="html-italic">n</span> = 5–10). The effect of the factor irrigation (I), factor biostimulant (B), and the interaction between the factor irrigation and biostimulant (I × B) are presented, and when the effect of each factor or the interaction is statistically significant (<span class="html-italic">p</span> ≤ 0.05), it appears in bold. Different letters indicate statistically significant differences (<span class="html-italic">p</span> ≤ 0.05). Significant differences among I × B refer to differences between C, BC, S, and BS treatments. Significant differences among the factor I refer to differences between 100% (treatments C and BC) and 50% irrigation (treatments S and BS).</p>
Full article ">Figure 4
<p>Chlorophyll <span class="html-italic">a</span> (<b>A</b>,<b>B</b>) and <span class="html-italic">b</span> (<b>C</b>,<b>D</b>), and carotenoid (<b>E</b>,<b>F</b>) contents in <span class="html-italic">O. europaea</span> plants of the treatments C (well-watered), BC (biostimulant + well-watered), S (water deficit), and BS (biostimulant + water deficit). Bars represent mean ± standard error (<span class="html-italic">n</span> = 6–8). The effect of the factor irrigation (I), factor biostimulant (B), and the interaction between the factor irrigation and biostimulant (I × B) are presented, and when the effect of each factor or the interaction is statistically significant (<span class="html-italic">p</span> ≤ 0.05), it appears in bold. Different letters indicate statistically significant difference (<span class="html-italic">p</span> ≤ 0.05). Significant differences among I × B refer to differences between C, BC, S, and BS treatments. Significant differences among the factor I refer to differences between 100% (treatments C and BC) and 50% irrigation (treatments S and BS).</p>
Full article ">Figure 5
<p>Carbohydrate (<b>A</b>,<b>C</b>) and terpene (<b>B</b>,<b>D</b>) relative abundance (%) in <span class="html-italic">O. europaea</span> plants of the cultivar Arbequina (<b>A</b>,<b>B</b>) and Galega (<b>C</b>,<b>D</b>) in the treatments C (well-watered), BC (biostimulant + well-watered), S (water deficit), and BS (biostimulant + water deficit). Bars represent mean ± standard error (<span class="html-italic">n</span> = 3–4). The effect of the factor irrigation (I), factor biostimulant (B), and the interaction between the factor irrigation and biostimulant (I × B) are presented, and when the effect of each factor or the interaction is statistically significant (<span class="html-italic">p</span> ≤ 0.05), it appears in bold. Different letters indicate statistically significant difference (<span class="html-italic">p</span> ≤ 0.05). Significant differences among I × B refer to differences between C, BC, S, and BS treatments. Significant differences among the factor I refer to differences between 100% (treatments C and BC) and 50% (treatments S and BS) irrigation. For the case of <span class="html-small-caps">d</span>-(−)-tagatofuranose, gluconolactone, <span class="html-small-caps">d</span>-glucose, <span class="html-small-caps">d</span>-(+)-galactose, <span class="html-small-caps">d</span>-(+)-turanose, and <span class="html-small-caps">d</span>-erythrose in Arbequina and <span class="html-small-caps">d</span>-(−)-tagatofuranose, <span class="html-small-caps">d</span>-glucose, <span class="html-small-caps">d</span>-(+)-galactose, <span class="html-small-caps">d</span>-(+)-turanose, <span class="html-small-caps">d</span>-erythrose, <span class="html-small-caps">d</span>-mannitol, and myo-inositol in Galega, one-way ANOVA was performed and significant differences (<span class="html-italic">p</span> ≤ 0.05) are marked in bold and indicated by different letters. <span class="html-small-caps">d</span>-(−)-Tagato.: <span class="html-small-caps">d</span>-(−)-Tagatofuranose; Gluconolac.: Gluconolactone; <span class="html-small-caps">d</span>-(+)-Galac.—<span class="html-small-caps">d</span>-(+)-Galactose; <span class="html-small-caps">d</span>-(+)-Turan.: <span class="html-small-caps">d</span>-(+)-Turanose.</p>
Full article ">Figure 6
<p>Relative expression of dehydrins <span class="html-italic">OeDHN1I</span> (<b>A</b>,<b>D</b>), small heat shock proteins <span class="html-italic">OeHSP18.3</span> (<b>B</b>,<b>E</b>), and aquaporins <span class="html-italic">OePIP1.1</span> (<b>C</b>,<b>F</b>) in <span class="html-italic">O. europaea</span> plants of the treatments C (well-watered), BC (biostimulant + well-watered), S (water deficit), and BS (biostimulant + water deficit). Bars represents mean ± standard error (<span class="html-italic">n</span> = 4–8). The effect of the factor irrigation (I), factor biostimulant (B), and the interaction between the factor irrigation and biostimulant (I × B) are presented, and when the effect of each factor or the interaction is statistically significant (<span class="html-italic">p</span> ≤ 0.05), it appears in bold. Different letters indicate statistically significant difference (<span class="html-italic">p</span> ≤ 0.05). Significant differences among I × B refer to differences between C, BC, S, and BS treatments. Significant differences among the factor B refer to treatments without biostimulant (treatments C and S) and treatments with biostimulant (treatments BC and BS).</p>
Full article ">Figure 7
<p>Principal component analysis plot (<span class="html-italic">x</span>-axis—first component PC1; and <span class="html-italic">y</span>-axis—second component PC2) of the physiological, molecular, and metabolomic data in olive leaves from both cultivars. PC1 explains 36% of the variance, while PC2 explain 23%. Circles with different colors depict sample scores of the different treatments. Ac.: acid; Carot.: carotenoids; DHN: <span class="html-italic">OeDHN1</span>; E: transpiration rate; Galacto.: <span class="html-small-caps">d</span>-galactose; Gluco.: <span class="html-small-caps">d</span>-Glucose; gs: stomatal conductance; Height: heigh increment; HSP: <span class="html-italic">OeHSP18.3</span>; LCA: long-chain alkane; Mio-In.: myo-inositol; Neophyt.: neophytadiene; PIP: <span class="html-italic">OePIP1.1</span>; P<sub>n</sub>: net CO<sub>2</sub>-assimilation rate; Tagato.: <span class="html-small-caps">d</span>-(−)- Tagatofuranose; WUE: intrinsic water-use efficiency.</p>
Full article ">Figure 8
<p>Schematization of the experiment. FC, field capacity.</p>
Full article ">
14 pages, 5956 KiB  
Article
Development of a Macro X-ray Fluorescence (MA-XRF) Scanner System for In Situ Analysis of Paintings That Operates in a Static or Dynamic Method
by Renato P. de Freitas, Miguel A. de Oliveira, Matheus B. de Oliveira, André R. Pimenta, Valter de S. Felix, Marcelo O. Pereira, Elicardo A. S. Gonçalves, João V. L. Grechi, Fabricio L. e. Silva, Cristiano de S. Carvalho, Jonas G. R. S. Ataliba, Leandro O. Pereira, Lucas C. Muniz, Robson B. dos Santos and Vitor da S. Vital
Quantum Beam Sci. 2024, 8(4), 26; https://doi.org/10.3390/qubs8040026 - 17 Oct 2024
Abstract
This work presents the development of a macro X-ray fluorescence (MA-XRF) scanner system for in situ analysis of paintings. The instrument was developed to operate using continuous acquisitions, where the module with the X-ray tube and detector moves at a constant speed, dynamically [...] Read more.
This work presents the development of a macro X-ray fluorescence (MA-XRF) scanner system for in situ analysis of paintings. The instrument was developed to operate using continuous acquisitions, where the module with the X-ray tube and detector moves at a constant speed, dynamically collecting spectra for each pixel of the artwork. Another possible configuration for the instrument is static acquisitions, where the module with the X-ray tube and detector remains stationary to acquire spectra for each pixel. The work also includes the analytical characterization of the system, which incorporates a 1.00 mm collimator that allows for a resolution of 1.76 mm. Additionally, the study presents the results of the analysis of two Brazilian paintings using this instrument. The elemental maps obtained enabled the characterization of the pigments used in the creation of the artworks and materials used in restoration processes. Full article
(This article belongs to the Special Issue New Advances in Macro X-ray Fluorescence Applications)
Show Figures

Figure 1

Figure 1
<p>Mechanical project containing the module with X-ray tube and detector, which was integrated into the movement system.</p>
Full article ">Figure 2
<p>Complete mechanical design of the MA-XRF system; X-ray generator power supply (<b>a</b>).</p>
Full article ">Figure 3
<p>Fitting model developed from MA-XRF scanning data of the painting “<span class="html-italic">A Morta</span>”.</p>
Full article ">Figure 4
<p>Verification of the scanner’s spatial resolution for distances of 12 and 13 mm using the knife−edge method. (<b>A</b>): counts × positions for distance 12 mm; (<b>B</b>): 1st derivative for distance 12 mm; (<b>C</b>): counts × positions for distance 13 mm; (<b>D</b>): 1st derivative for distance 12 mm.</p>
Full article ">Figure 5
<p>(<b>A</b>) Fe-Kα photon counts collected over 30 min; (<b>B</b>) spectrum from standard sample collected during 1 s; (<b>C</b>) results of the tests of sensitivity (S); (<b>D</b>) limit detection (DL).</p>
Full article ">Figure 6
<p>The painting “São Paulo” (1741 mm × 712 mm), collection of the D. João VI Museum, Rio de Janeiro, Brazil. The red polygon region indicates the scanning area of the painting.</p>
Full article ">Figure 7
<p>The painting “<span class="html-italic">A Morta</span>” (504 mm × 612 mm), collection of the Victor Meirelles Museum, Santa Catarina. The red polygon region indicates the scanning area of the painting.</p>
Full article ">Figure 8
<p>Comparison of the maximum XRF spectra collected in the matrix of the paintings “São Paulo” and “<span class="html-italic">A Morta</span>”.</p>
Full article ">Figure 9
<p>Elemental maps of the painting “São Paulo”.</p>
Full article ">Figure 10
<p>Elemental maps of the painting “<span class="html-italic">A Morta</span>”.</p>
Full article ">
19 pages, 15466 KiB  
Article
Transcriptomic Analysis Reveals the Mechanism of Color Formation in the Peel of an Evergreen Pomegranate Cultivar ‘Danruo No.1’ During Fruit Development
by Xiaowen Wang, Chengkun Yang, Wencan Zhu, Zhongrui Weng, Feili Li, Yuanwen Teng, Kaibing Zhou, Minjie Qian and Qin Deng
Plants 2024, 13(20), 2903; https://doi.org/10.3390/plants13202903 - 17 Oct 2024
Abstract
Pomegranate (Punica granatum L.) is an ancient fruit crop that has been cultivated worldwide and is known for its attractive appearance and functional metabolites. Fruit color is an important index of fruit quality, but the color formation pattern in the peel of [...] Read more.
Pomegranate (Punica granatum L.) is an ancient fruit crop that has been cultivated worldwide and is known for its attractive appearance and functional metabolites. Fruit color is an important index of fruit quality, but the color formation pattern in the peel of evergreen pomegranate and the relevant molecular mechanism is still unknown. In this study, the contents of pigments including anthocyanins, carotenoids, and chlorophyll in the peel of ‘Danruo No. 1’ pomegranate fruit during three developmental stages were measured, and RNA-seq was conducted to screen key genes regulating fruit color formation. The results show that pomegranate fruit turned from green to red during development, with a dramatic increase in a* value, indicating redness and anthocyanins concentration, and a decrease of chlorophyll content. Moreover, carotenoids exhibited a decrease–increase accumulation pattern. Through RNA-seq, totals of 30, 18, and 17 structural genes related to anthocyanin biosynthesis, carotenoid biosynthesis and chlorophyll metabolism were identified from differentially expressed genes (DEGs), respectively. Transcription factors (TFs) such as MYB, bHLH, WRKY and AP2/ERF were identified as key candidates regulating pigment metabolism by K-means analysis and weighted gene co-expression network analysis (WGCNA). The results provide an insight into the theory of peel color formation in evergreen pomegranate fruit. Full article
(This article belongs to the Special Issue Recent Advances in Horticultural Plant Genomics)
Show Figures

Figure 1

Figure 1
<p>Coloration and pigment contents in ‘Danruo No.1’ pomegranate peel during fruit development. (<b>A</b>) Representative images of fruits at developmental stage 1 (S1), stage 2 (S2), and stage 3 (S3). (<b>B</b>) Fruit peel lightness (<span class="html-italic">L*</span> value). (<b>C</b>) Fruit peel <span class="html-italic">a*</span> value (higher value means redness and lower value means greenness). (<b>D</b>) Fruit peel <span class="html-italic">b*</span> value (higher value means yellowness and lower value means blueness). (<b>E</b>) Chlorophyll a content. (<b>F</b>) Chlorophyll b content. (<b>G</b>) Total chlorophyll content. (<b>H</b>) Anthocyanin content. (<b>I</b>) Carotenoid content. Each value represents the mean ± standard deviation of three biological replicates. Different letters above the bars indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) according to one-way analysis of variance (ANOVA) followed by Tukey test.</p>
Full article ">Figure 2
<p>Differentially expressed genes (DEGs) identification and KEGG analysis. Volcano plots of DEGs from S2 vs. S1 (<b>A</b>), S3 vs. S1 (<b>B</b>), and S3 vs. S2 (<b>C</b>). Horizontal coordinates indicate the fold change of gene expression between different groups, and vertical coordinates indicate the significance level of gene expression difference in the two groups. Red dots indicate upregulated genes, green dots indicate downregulated genes, and grey dots indicate insignificant genes. Top 20 metabolic pathways analyzed by KEGG enrichment for DEGs from S2 vs. S1 (<b>D</b>), S3 vs. S1 (<b>E</b>), and S3 vs. S2 (<b>F</b>). The pathways associated with pigments metabolism are highlighted in red color.</p>
Full article ">Figure 3
<p>Expression patterns of the DEGs involved in anthocyanins synthesis in pomegranate peel at developmental stage 1 (S1), stage 2 (S2), and stage 3 (S3). The color scale from green to red represents the fragments per kilobase of transcript per million of fragments mapped (FPKM) values, from low to high.</p>
Full article ">Figure 4
<p>Expression pattern of the DEGs involved in carotenoids synthesis in pomegranate peel at developmental stage 1 (S1), stage 2 (S2), and stage 3 (S3). The color scale from blue to red represents the fragments per kilobase of transcript per million of fragments mapped (FPKM) values from low to high.</p>
Full article ">Figure 5
<p>Expression pattern of the DEGs involved in chlorophyll biosynthesis and degradation in pomegranate peel at developmental stage 1 (S1), stage 2 (S2), and stage 3 (S3). The color scale from green to red represents the fragments per kilobase of transcript per million of fragments mapped (FPKM) values from low to high.</p>
Full article ">Figure 6
<p>Identification of transcription factors (TFs) regulating pigments metabolism in pomegranate peel during fruit development. (<b>A</b>) K-means analysis of DEGs identified from transcriptome sequencing. The expression profiles of genes in each cluster are represented in different colors, and the average expression levels of all genes in developmental stage 1 (S1), S2, and S3 are represented in black. (<b>B</b>) Weighted gene co-expression network analysis (WGCNA) of DEGs identified from transcriptome sequencing. Module-trait correlations and corresponding <span class="html-italic">p</span>-values in parentheses. The left panel shows the six modules with gene numbers. The color scale on the right shows the module-trait correlations from −1 (blue) to 1 (red). ‘Anthocyanin’, ‘Chlorophyll a’, ‘Chlorophyll b’, ‘Total chlorophyll’ and ‘Carotenoid’ represent the changes in corresponding substances’ concentrations. (<b>C</b>) Heatmap presenting the expression patterns of regulatory genes regulating pomegranate peel pigments metabolism during fruit development. (<b>D</b>) Correlation network between TFs’ expression and pigments’ contents; pink and blue circles represent positive and negative correlations, respectively. Purple, orange, and green lines representing the relation between TFs and anthocyanin, carotenoid, and chlorophyll, respectively.</p>
Full article ">Figure 7
<p>The expressions of seven genes in pomegranate peel at developmental stage 1 (S1), S2, and S3 from transcriptome data were examined by quantitative polymerase chain reaction (q-PCR). The expression levels obtained by RNA-seq and q-PCR are shown with a line chart and histogram, respectively. Data are presented as the mean ± standard deviation of three biological replicates. Different letters above the bars indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) according to one-way analysis of variance (ANOVA) followed by Tukey test. Data analyzed by qPCR (marked with gray letters) or RNA-seq (marked with red letters) were tested separately.</p>
Full article ">
13 pages, 9828 KiB  
Article
Examining Carotenoid Metabolism Regulation and Its Role in Flower Color Variation in Brassica rapa L.
by Guomei Liu, Liuyan Luo, Lin Yao, Chen Wang, Xuan Sun and Chunfang Du
Int. J. Mol. Sci. 2024, 25(20), 11164; https://doi.org/10.3390/ijms252011164 - 17 Oct 2024
Abstract
Carotenoids are vital organic pigments that determine the color of flowers, roots, and fruits in plants, imparting them yellow, orange, and red hues. This study comprehensively analyzes carotenoid accumulation in different tissues of the Brassica rapa mutant “YB1”, which exhibits altered flower and [...] Read more.
Carotenoids are vital organic pigments that determine the color of flowers, roots, and fruits in plants, imparting them yellow, orange, and red hues. This study comprehensively analyzes carotenoid accumulation in different tissues of the Brassica rapa mutant “YB1”, which exhibits altered flower and root colors. Integrating physiological and biochemical assessments, transcriptome profiling, and quantitative metabolomics, we examined carotenoid accumulation in the flowers, roots, stems, and seeds of YB1 throughout its growth and development. The results indicated that carotenoids continued to accumulate in the roots and stems of YBI, especially in its cortex, throughout plant growth and development; however, the carotenoid levels in the petals decreased with progression of the flowering stage. In total, 54 carotenoid compounds were identified across tissues, with 30 being unique metabolites. Their levels correlated with the expression pattern of 22 differentially expressed genes related to carotenoid biosynthesis and degradation. Tissue-specific genes, including CCD8 and NCED in flowers and ZEP in the roots and stems, were identified as key regulators of color variations in different plant parts. Additionally, we identified genes in the seeds that regulated the conversion of carotenoids to abscisic acid. In conclusion, this study offers valuable insights into the regulation of carotenoid metabolism in B. rapa, which can guide the selection and breeding of carotenoid-rich varieties. Full article
Show Figures

Figure 1

Figure 1
<p>Dynamics of the phenotype and total carotenoid content in the mutants. (<b>A</b>): YB1 flowers; (<b>B</b>): TY7 flowers; (<b>C</b>): comparison of total carotenoid content at different flowering stages; CB, CO and CA represent the bud stage, semi-open stage, and full bloom stage of TY7 petals, respectively; YB, YO, and YA represent the bud stage, semi-open stage, and full bloom stage of YB1 petals, respectively; (<b>D</b>): YB1 rhizomes; (<b>E</b>): TY7 rhizomes; (<b>F</b>): comparison of total carotenoids at different stages of rhizome fertility; CP and YP denote the TY7 and YB1 cortices, respectively, and CW and YW denote the TY7 and YB1 vascular bundles, respectively; November 2022 is referred to as the 11th, December 2022 as the 12th, January 2023 as the 1st, February 2023 as the 2nd, March 2023 as the 3rd, April 2023 as the 4th, and May 2023 as the 5th. Data are expressed as the mean of three biological replicates. Differences between the two varieties were considered statistically significant at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Identification and clustering analysis of carotenoid differential metabolites. (<b>A</b>): OPLS-DA supervised analysis; CK1, CK2, and CK3 denote the petal, rhizome, and seed samples of the TY7 variety, respectively; YB1, YB2, and YB3 denote the petal, rhizome, and seed samples of the YB1 variety, respectively; (<b>B</b>): metabolite Wayne plots; comparisons between CSM (TY7 seed) and YSM (YB1 seed); CRM (TY7 root) and YRM (YB1 root); and CFM (TY7 petal) and YFM (YB1 petal); (<b>C</b>): heatmap of carotenoid metabolite clustering in different tissues; CF1−1, CF1−2, and CF1−3 represent the three biological replicates of TY7 petal samples; CR2−1, CR2−2, and CR2−3 represent the three biological replicates of TY7 root samples; CS3−1, CS3−2, and CS3−3 represent the three biological replicates of TY7 seed samples; YF1−1, YF1−2, YF1−3 represent the three biological replicates of YB1 petal samples; YR2−1, YR2−2, and YR2−3 represent the three biological replicates of YB1 root samples; and YS3−1, YS3−2, and YS3−3 represent the three biological replicates of YB1 seed samples; (<b>D</b>): KEGG analysis of differential metabolites. Note: CF stands for TY7 flower, YF stands for YB1 flower, CR stands for TY7 rhizome, YR stands for YB1 rhizome, CS stands for TY7 seed, and YS denotes YB1 seed.</p>
Full article ">Figure 3
<p>Transcriptome analysis of different samples. (<b>A</b>): Wayne plots of differentially expressed genes (DEGs) in different tissues of the control and mutant plants; (<b>B</b>): transcriptome DEGs; CF_vs._YF denotes the comparison between petals of TY7 and YB1; CR_vs._YR denotes the comparison between the roots of TY7 and YB1; and CS_vs._YS denotes the comparison between the seeds of TY7 and YB1. (<b>C</b>): GO classification of DEGs. (<b>D</b>): KEGG pathway enrichment of the DEGs.</p>
Full article ">Figure 4
<p>Weighted gene co-expression network analysis of the genes associated with carotenoid metabolism. (<b>A</b>): Hierarchical clustering tree diagram of co-expressed genes in WGCNA, with each leaf corresponding to one gene, and the main branches from seven modules labeled in different colors; (<b>B</b>): relationship between modules and carotenoid metabolism-related DEGs, with each row representing one module. Each column represents the carotenoid biosynthesis-related DEGs; the value of each cell at the intersection of rows and columns represents the coefficient of correlation between the modules and carotenoid metabolism DEGs (shown on the right side of the color scale), whereas the value in parentheses in each cell represents the <span class="html-italic">p</span> value; (<b>C</b>): KEGG enrichment analysis of turquoise module DEGs; (<b>D</b>): KEGG enrichment analysis of green module DEGs; (<b>E</b>): KEGG enrichment analysis of yellow module DEGs.</p>
Full article ">Figure 5
<p>Pearson correlation analysis of DEGs with carotenoid differential metabolites (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Carotenoid regulatory networks in different tissues. Note: CF: TY7 petals; YF: YB1 petals; CR: TY7 rhizomes; YR: YB1 rhizomes; CS: TY7 seeds; YS: YB1 seeds; PDS: 15-cis-octahydroxylycopene desaturase; crtL2: lycopene e-cyclase; CYP97A3: β-cyclohydroxylase; crtZ: β-carotenoids 3-lightening enzyme; CCD8: carotenoid cleavage dioxygenase; NCED: 9-cis-epoxycarotenoid dioxygenase; ABA2: xanthoxin dehydrogenase; CYP707A: (+)−abscisic acid 8′-hydroxylase; ZEP, ABA1: zeaxanthin epoxidase. Orange color indicates upregulation and light blue color indicates downregulation.</p>
Full article ">Figure 7
<p>qRT-PCR assay for the differential expression profiles of genes in the seeds, petals, and roots of the control and mutant plants and transcriptome heat map. *** Significantly Note: CF: TY7 petals; YF: YB1 petals; CR: TY7 rhizomes; YR: YB1 rhizomes; CS: TY7 seeds; YS: YB1 seeds.</p>
Full article ">
21 pages, 7722 KiB  
Article
Transcriptomic Analysis During Olive Fruit Development and Expression Profiling of Fatty Acid Desaturase Genes
by Alicia Serrano, Judith García-Martín, Martín Moret, José Manuel Martínez-Rivas and Francisco Luque
Int. J. Mol. Sci. 2024, 25(20), 11150; https://doi.org/10.3390/ijms252011150 - 17 Oct 2024
Abstract
The olive fruit is a drupe whose development and ripening takes several months from flowering to full maturation. During this period, several biochemical and physiological changes occur that affect the skin color, texture, composition, and size of the mesocarp. The final result is [...] Read more.
The olive fruit is a drupe whose development and ripening takes several months from flowering to full maturation. During this period, several biochemical and physiological changes occur that affect the skin color, texture, composition, and size of the mesocarp. The final result is a fruit rich in fatty acids, phenolic compounds, tocopherols, pigments, sterols, terpenoids, and other compounds of nutritional interest. In this work, a transcriptomic analysis was performed using flowers (T0) and mesocarp tissue at seven different stages during olive fruit development and ripening (T1–T7) of the ‘Picual’ cultivar. A total of 1755 genes overexpressed at any time with respect to the flowering stage were further analyzed. These genes were grouped into eight clusters based on their expression profile. The gene enrichment analysis revealed the most relevant biological process of every cluster. Highlighting the important role of hormones at very early stages of fruit development (T1, Cluster 1), whereas genes involved in fatty acid biosynthesis were relevant throughout the fruit developmental process. Hence, genes coding for different fatty acid desaturase (SAD, FAD2, FAD3, FAD4, FAD5, FAD6, and FAD7) enzymes received special attention. In particular, 26 genes coding for different fatty acid desaturase enzymes were identified in the ‘Picual’ genome, contributing to the improvement of the genome annotation. The expression pattern of these genes during fruit development corroborated their role in determining fatty acid composition. Full article
(This article belongs to the Special Issue Genomic and Transcriptomic Analysis of Olive (Olea europaea L.))
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Developmental stages collected for RNAseq analysis. (<b>B</b>) PCA plot showing the expression differences among olive fruit developing samples. Samples collected in triplicate: T0: flowers at full bloom, T1: fruits at 15 days after full blooming (AFB), T2: fruits at 1 month AFB, T3: fruits at 2 months AFB, T4: fruits at 3 months AFB, T5: fruits at 4 months AFB, T6: fruits at 5 months AFB, and T7: fruits at 6 months AFB.</p>
Full article ">Figure 2
<p>Differentially expressed genes throughout fruit development owing to the flowering stage. T0: flowers at full bloom, T1: 15 days after full blooming (AFB), T2: 1 month AFB, T3: 2 months AFB, T4: 3 months AFB, T5: 4 months AFB, T6: 5 months AFB, and T7: 6 months AFB.</p>
Full article ">Figure 3
<p>Cluster A. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 4
<p>Cluster B. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 5
<p>Cluster C. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 6
<p>Cluster D. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 7
<p>Cluster E. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 8
<p>Cluster F. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 9
<p>Cluster G. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 10
<p>Cluster H. (<b>A</b>) Gene expression profile. Blue line represents the mean value of expression for the total genes in the group. Blue shadow represents the standard error of gene expression. Gray lines represent the expression of individual genes. The red horizontal line represents the threshold separating positive from negative expression levels. (<b>B</b>) The 20 most representative biological processes according to the FDR and fold enrichment values obtained in ShinyGO 0.80.</p>
Full article ">Figure 11
<p>Expression of coding genes for isoforms of SAD enzyme during olive fruit development.</p>
Full article ">Figure 12
<p>Expression of coding genes for microsomal FAD enzymes during olive fruit development.</p>
Full article ">Figure 13
<p>Expression of coding genes for plastidial membrane-bound FAD enzymes during olive fruit development.</p>
Full article ">
15 pages, 666 KiB  
Article
Comparative Analysis of AI Models for Atypical Pigmented Facial Lesion Diagnosis
by Alessandra Cartocci, Alessio Luschi, Linda Tognetti, Elisa Cinotti, Francesca Farnetani, Aimilios Lallas, John Paoli, Caterina Longo, Elvira Moscarella, Danica Tiodorovic, Ignazio Stanganelli, Mariano Suppa, Emi Dika, Iris Zalaudek, Maria Antonietta Pizzichetta, Jean Luc Perrot, Gabriele Cevenini, Ernesto Iadanza, Giovanni Rubegni, Harald Kittler, Philipp Tschandl and Pietro Rubegniadd Show full author list remove Hide full author list
Bioengineering 2024, 11(10), 1036; https://doi.org/10.3390/bioengineering11101036 - 17 Oct 2024
Abstract
Diagnosing atypical pigmented facial lesions (aPFLs) is a challenging topic for dermatologists. Accurate diagnosis of these lesions is crucial for effective patient management, especially in dermatology, where visual assessment plays a central role. Incorrect diagnoses can result in mismanagement, delays in appropriate interventions, [...] Read more.
Diagnosing atypical pigmented facial lesions (aPFLs) is a challenging topic for dermatologists. Accurate diagnosis of these lesions is crucial for effective patient management, especially in dermatology, where visual assessment plays a central role. Incorrect diagnoses can result in mismanagement, delays in appropriate interventions, and potential harm. AI, however, holds the potential to enhance diagnostic accuracy and provide reliable support to clinicians. This work aimed to evaluate and compare the effectiveness of machine learning (logistic regression of lesion features and patient metadata) and deep learning (CNN analysis of images) models in dermoscopy diagnosis and the management of aPFLs. This study involved the analysis of 1197 dermoscopic images of facial lesions excised due to suspicious and histologically confirmed malignancy, classified into seven classes (lentigo maligna—LM; lentigo maligna melanoma—LMM; atypical nevi—AN; pigmented actinic keratosis—PAK; solar lentigo—SL; seborrheic keratosis—SK; and seborrheic lichenoid keratosis—SLK). Image samples were collected through the Integrated Dermoscopy Score (iDScore) project. The statistical analysis of the dataset shows that the patients mean age was 65.5 ± 14.2, and the gender was equally distributed (580 males—48.5%; 617 females—51.5%). A total of 41.7% of the sample constituted malignant lesions (LM and LMM). Meanwhile, the benign lesions were mainly PAK (19.3%), followed by SL (22.2%), AN (10.4%), SK (4.0%), and SLK (2.3%). The lesions were mainly localised in the cheek and nose areas. A stratified analysis of the assessment provided by the enrolled dermatologists was also performed, resulting in 2445 evaluations of the 1197 images (2.1 evaluations per image on average). The physicians demonstrated higher accuracy in differentiating between malignant and benign lesions (71.2%) than in distinguishing between the seven specific diagnoses across all the images (42.9%). The logistic regression model obtained a precision of 39.1%, a sensitivity of 100%, a specificity of 33.9%, and an accuracy of 53.6% on the test set, while the CNN model showed lower sensitivity (58.2%) and higher precision (47.0%), specificity (90.8%), and accuracy (59.5%) for melanoma diagnosis. This research demonstrates how AI can enhance the diagnostic accuracy in complex dermatological cases like aPFLs by integrating AI models with clinical data and evaluating different diagnostic approaches, paving the way for more precise and scalable AI applications in dermatology, showing their critical role in improving patient management and the outcomes in dermatology. Full article
(This article belongs to the Special Issue Computer Vision and Machine Learning in Medical Applications)
Show Figures

Figure 1

Figure 1
<p>Examples of images for each diagnosis in the iDScore database. (<b>A</b>) Atypical nevi, (<b>B</b>) lentigo maligna, (<b>C</b>) pigmented actinic keratosis, (<b>D</b>) seborrheic keratosis, (<b>E</b>) seborrheic lichenoid keratosis, (<b>F</b>) solar lentigo.</p>
Full article ">Figure 2
<p>Distribution of dermatologists’ expertise in dermoscopy.</p>
Full article ">Figure 3
<p>Distribution of the specific subareas of the face.</p>
Full article ">Figure 4
<p>ROC curves for the model on the validation and testing samples and the pattern recognition diagnoses of the dermatologists.</p>
Full article ">Figure 5
<p>Loss value and mean recall and accuracy for each epoch in the training and validation sample.</p>
Full article ">Figure 6
<p>Confusion matrix of the CNN model on the testing sample.</p>
Full article ">Figure 7
<p>Management distributions according to dermatologists compared to the scores of the model for LM/LMM (<b>top</b>) and for other aPFLs (<b>bottom</b>) [<a href="#B23-bioengineering-11-01036" class="html-bibr">23</a>].</p>
Full article ">
18 pages, 1334 KiB  
Article
Studying the Stability of Anthocyanin Pigments Isolated from Juices of Colored-Fleshed Potatoes
by Agnieszka Tkaczyńska, Esther Sendra, Nuria Jiménez-Redondo and Elżbieta Rytel
Int. J. Mol. Sci. 2024, 25(20), 11116; https://doi.org/10.3390/ijms252011116 - 16 Oct 2024
Viewed by 286
Abstract
The aim of this study was to obtain extracts of anthocyanin pigments from red and purple-fleshed potato juices characterized by stable color. For this purpose, potato juices were pasteurized at different temperatures or fruit and vegetable concentrates were added to them. Color stability [...] Read more.
The aim of this study was to obtain extracts of anthocyanin pigments from red and purple-fleshed potato juices characterized by stable color. For this purpose, potato juices were pasteurized at different temperatures or fruit and vegetable concentrates were added to them. Color stability tests of the obtained pigments were carried out in model pH and temperature conditions and after adding to natural yogurt. Both the pasteurization process and the addition of fruit and vegetable concentrates to the potato juices positively affected their color and its stability in time. However, the pasteurization of the potato juices had a negative effect on the content of biologically active compounds, in contrast to the juices stabilized with the addition of fruit and vegetable concentrates. Anthocyanin pigments from red-fleshed potato juices were more stable than those isolated from the purple-fleshed potato juices. The results of model tests of the anthocyanin pigment concentrates from the colored-flesh potatoes and natural yoghurts with their addition confirmed the high stability of the tested concentrates. Full article
(This article belongs to the Section Bioactives and Nutraceuticals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sensory evaluation of yogurts without (natural yogurt) and with added pigments obtained from pasteurized and unpasteurized juices. 0—lowest score; 10—highest score.</p>
Full article ">Figure 2
<p>Sensory evaluation of yogurts without (natural yogurt) and with added pigments obtained from juices with added fruit and vegetable concentrates. 0—lowest score; 10—highest score.</p>
Full article ">
21 pages, 7014 KiB  
Article
Molecular Mechanism of Mok I Gene Overexpression in Enhancing Monacolin K Production in Monascus pilosus
by Zhiwei Huang, Lishi Xiao, Wenlan Mo, Yaru Zhang, Yiyang Cai, Simei Huang, Zhiting Chen and Chuannan Long
J. Fungi 2024, 10(10), 721; https://doi.org/10.3390/jof10100721 - 16 Oct 2024
Viewed by 204
Abstract
Monascus species are capable of producing various active metabolites, including monacolin K (MK) and pigments. Studies have shown that the overexpression of the mok I gene from the MK synthesis gene cluster in Monascus species can significantly increase MK production; however, the molecular [...] Read more.
Monascus species are capable of producing various active metabolites, including monacolin K (MK) and pigments. Studies have shown that the overexpression of the mok I gene from the MK synthesis gene cluster in Monascus species can significantly increase MK production; however, the molecular mechanism has not yet been fully elucidated. Therefore, this study focused on the mok I gene of Monascus pilosus to construct overexpression strains of the mok I gene, resulting in high-yield MK production. Sixteen positive transformants were obtained, seven of which produced 9.63% to 41.39% more MK than the original strain, with no citrinin detected in any of the transformants. The qRT-PCR results revealed that the expression levels of mok I in the transformed strains TI-13, TI-24, and TI-25 increased by more than 50% compared to the original strain at various fermentation times, with the highest increase being 10.9-fold. Furthermore, multi-omics techniques were used to analyze the molecular mechanisms underlying enhanced MK production in transformed strains. The results indicated that mok I overexpression may enhance MK synthesis in M. pilosus by regulating the expression of key genes (such as MAO, HPD, ACX, and PLC) and the synthesis levels of key metabolites (such as delta-tocopherol and alpha-linolenic acid) in pathways linked to the biosynthesis of cofactors, the biosynthesis of unsaturated fatty acids, tyrosine metabolism, ubiquinone and other terpenoid-quinone biosynthesis, alpha-linolenic acid metabolism, and glycerophospholipid metabolism. These findings provide a theoretical basis for further study of the metabolic regulation of MK in Monascus species and for effectively enhancing their MK production. Full article
(This article belongs to the Special Issue Monascus spp. and Their Relative Products)
Show Figures

Figure 1

Figure 1
<p>HPLC chromatograms of MK from the transformed strains and the original strain of <span class="html-italic">M. pilosus</span>. (<b>A</b>) Acidic form of MK standard. (<b>B</b>) Lactone form of MK standard. (<b>C</b>) Samples of transformed strains of <span class="html-italic">M. pilosus</span>. (<b>D</b>) Sample of the original strain of <span class="html-italic">M. pilosus</span>.</p>
Full article ">Figure 2
<p>MK yield of the transformed strains and the original strain of <span class="html-italic">M. pilosus</span>. Different letters in different strains indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Identification of MK in the samples of transformed strain by HPLC/MS. (<b>A</b>) Mass spectrogram of lactone MK standard under positive ion mode (M + H)<sup>+</sup> and (M + Na)<sup>+</sup>. (<b>B</b>) Mass spectrogram of lactone MK standard under negative ion mode (M + COOH)<sup>−</sup>. (<b>C</b>) Mass spectrogram of acid MK in the sample of the transformed strain under positive ion mode (M + Na)<sup>+</sup>; (<b>D</b>) Mass spectrogram of lactone MK in the sample of the transformed strain under negative ion mode (M − H)<sup>−</sup>.</p>
Full article ">Figure 4
<p>HPLC chromatograms of citrinin from the transformed strains of <span class="html-italic">M. pilosus</span> and commercially available RYR. (<b>A</b>) Citrinin standard. (<b>B</b>) Transformed strain of <span class="html-italic">M. pilosus</span> samples. (<b>C</b>) Commercially available RYR samples.</p>
Full article ">Figure 4 Cont.
<p>HPLC chromatograms of citrinin from the transformed strains of <span class="html-italic">M. pilosus</span> and commercially available RYR. (<b>A</b>) Citrinin standard. (<b>B</b>) Transformed strain of <span class="html-italic">M. pilosus</span> samples. (<b>C</b>) Commercially available RYR samples.</p>
Full article ">Figure 5
<p>Expression levels of <span class="html-italic">mok I</span> gene in the transformed strains and the original strain.</p>
Full article ">Figure 6
<p>Analysis of differentially expressed genes in spore samples from different groups. (<b>A</b>) PCA graph of spore samples. (<b>B</b>) Venn diagram of differentially expressed genes in spore samples of each group. (<b>C</b>) Volcano plots of gene expression differences among spore samples from different groups. Red represents upregulated genes and blue represents downregulated genes.</p>
Full article ">Figure 7
<p>GO, EggNOG, and KEGG functional annotation analysis of differentially expressed genes in spore samples from different groups. (<b>A</b>–<b>C</b>) EggNOG functional annotation. (<b>D</b>–<b>F</b>) GO functional annotation. (<b>G</b>–<b>I</b>) KEGG functional annotation. (<b>A</b>,<b>D</b>,<b>G</b>) TI2504 vs. CK04. (<b>B</b>,<b>E</b>,<b>H</b>) TI2508 vs. CK08. (<b>C</b>,<b>F</b>,<b>I</b>) TI2512 vs. CK12.</p>
Full article ">Figure 8
<p>Sankey-bubble diagram for GO and KEGG enrichment analysis of differential genes in spore samples from different groups. (<b>A</b>–<b>C</b>) GO enrichment. (<b>D</b>–<b>E</b>) KEGG enrichment. (<b>A</b>,<b>D</b>) TI2504 vs. CK04. (<b>B</b>) TI2508 vs. CK08. (<b>C</b>,<b>E</b>) TI2512 vs. CK12. The top 20 enrichment pathways were showed in the bubble chart, and the top 5 differential genes were showed in the Sankey plot sorted based on the <span class="html-italic">p</span>-value.</p>
Full article ">Figure 9
<p>PCA and OPLS-DA analysis of metabolites in RYR samples. (<b>A</b>) PCA score chart. (<b>B</b>–<b>D</b>) OPLS-DA score charts of TI2504 vs. CK04, TI2508 vs. CK08, and TI2512 vs. CK12.</p>
Full article ">Figure 10
<p>Volcano plot and VIP value analysis of differential metabolites in RYR samples from different groups. (<b>A</b>–<b>C</b>) Volcano plots. (<b>D</b>–<b>F</b>) VIP value analysis charts. (<b>A</b>,<b>D</b>) TI2504 vs. CK04. (<b>B</b>,<b>E</b>) TI2508 vs. CK08. (<b>C</b>,<b>F</b>) TI2512 vs. CK12. The top 30 metabolites were showed in the VIP chart sorted based on the VIP value. Red represents upregulated metabolites and blue represents downregulated metabolites.</p>
Full article ">Figure 11
<p>KEGG topology analysis of differential metabolites in RYR samples from different groups. (<b>A</b>) TI2504 vs. CK04. (<b>B</b>) TI2508 vs. CK08. (<b>C</b>) TI2512 vs. CK12. The top 5–6 enrichment pathways were labeled in the bubble charts based on the impact value and <span class="html-italic">p</span>-value.</p>
Full article ">Figure 12
<p>Enrichment map of KEGG pathway between differentially expressed genes in spore samples and differential metabolites in RYR samples from different groups. (<b>A</b>) TI2504 vs. CK04. (<b>B</b>) TI2508 vs. CK08. (<b>C</b>) TI2512 vs. CK12.</p>
Full article ">
14 pages, 1550 KiB  
Article
Non-Invasive Detection of Nitrogen Deficiency in Cannabis sativa Using Hand-Held Raman Spectroscopy
by Graham Antoszewski, James F. Guenther, John K. Roberts, Mickal Adler, Michael Dalle Molle, Nicholas S. Kaczmar, William B. Miller, Neil S. Mattson and Heather Grab
Agronomy 2024, 14(10), 2390; https://doi.org/10.3390/agronomy14102390 - 16 Oct 2024
Viewed by 267
Abstract
Proper crop management requires rapid detection methods for abiotic and biotic stresses to ensure plant health and yield. Hemp (Cannabis sativa L.) is an emerging economically and environmentally sustainable crop capable of yielding high biomass. Nitrogen deficiency significantly reduces hemp plant growth, [...] Read more.
Proper crop management requires rapid detection methods for abiotic and biotic stresses to ensure plant health and yield. Hemp (Cannabis sativa L.) is an emerging economically and environmentally sustainable crop capable of yielding high biomass. Nitrogen deficiency significantly reduces hemp plant growth, affecting photosynthetic capacity and ultimately decreasing yield. When symptoms of nitrogen deficiency are visible to humans, there is often already lost yield. A real-time, non-destructive detection method, such as Raman spectroscopy, is therefore critical to identify nitrogen deficiency in living hemp plant tissue for fast, precise crop remediation. A two-part experiment was conducted to investigate portable Raman spectroscopy as a viable hemp nitrogen deficiency detection method and to compare the technique’s predictive ability against a handheld SPAD (chlorophyll index) meter. Raman spectra and SPAD readings were used to train separate nitrogen deficiency discrimination models. Raman scans displayed characteristic spectral markers indicative of nitrogen deficiency corresponding to vibrational modes of carotenoids, essential pigments for photosynthesis. The Raman-based model consistently predicted nitrogen deficiency in hemp prior to the onset of visible stress symptoms across both experiments, while SPAD only differentiated nitrogen deficiency in the second experiment when the stress was more pronounced. Our findings add to the repertoire of plant stresses that hand-held Raman spectroscopy can detect by demonstrating the ability to provide assessments of nitrogen deficiency. This method can be implemented at the point of cultivation, allowing for timely interventions and efficient resource use. Full article
(This article belongs to the Section Precision and Digital Agriculture)
Show Figures

Figure 1

Figure 1
<p>Agilent Resolve spectrometer scanning the upper node, leaflet 2, of ‘TJ’s CBD’ hemp.</p>
Full article ">Figure 2
<p>The mean and standard deviation (shaded region) of Raman spectra of hemp leaf samples after 7 days under N-deficient (n = 36) versus complete nutrition (n = 36) in the two-cultivar trial [<a href="#B39-agronomy-14-02390" class="html-bibr">39</a>].</p>
Full article ">Figure 3
<p>Variable importance in projection scores for the three-component PLS-DA model for (<b>a</b>) early and (<b>b</b>) later-stage nitrogen deficiency detection. A VIP score &gt; 1 implies that the wavelength contributes significant information towards the model’s predictions.</p>
Full article ">Figure 4
<p>Photos of hemp leaf samples from the first experiment after (<b>a</b>) 7 days of nitrogen-deficient nutrient solution and (<b>b</b>) 7 days of complete solution. (<b>c</b>) Differences in mean SPAD readings between the two trial periods. Error bars represent measured standard deviation.</p>
Full article ">
19 pages, 818 KiB  
Article
Exploring Eye, Hair, and Skin Pigmentation in a Spanish Population: Insights from Hirisplex-S Predictions
by Belén Navarro-López, Miriam Baeta, Victoria Suárez-Ulloa, Rubén Martos-Fernández, Olatz Moreno-López, Begoña Martínez-Jarreta, Susana Jiménez, Iñigo Olalde and Marian M. de Pancorbo
Genes 2024, 15(10), 1330; https://doi.org/10.3390/genes15101330 (registering DOI) - 16 Oct 2024
Viewed by 218
Abstract
Background/Objectives: Understanding and predicting human pigmentation traits is crucial for individual identification. Genome-wide association studies have revealed numerous pigmentation-associated SNPs, indicating genetic overlap among pigmentation traits and offering the potential to develop predictive models without the need for analyzing large numbers of SNPs. [...] Read more.
Background/Objectives: Understanding and predicting human pigmentation traits is crucial for individual identification. Genome-wide association studies have revealed numerous pigmentation-associated SNPs, indicating genetic overlap among pigmentation traits and offering the potential to develop predictive models without the need for analyzing large numbers of SNPs. Methods: In this study, we assessed the performance of the HIrisPlex-S system, which predicts eye, hair, and skin color, on 412 individuals from the Spanish population. Model performance was calculated using metrics including accuracy, area under the curve, sensitivity, specificity, and positive and negative predictive value. Results: Our results showed high prediction accuracies (70% to 97%) for blue and brown eyes, brown hair, and intermediate skin. However, challenges arose with the remaining categories. The model had difficulty distinguishing between intermediate eye colors and similar shades of hair and exhibited a significant percentage of individuals with incorrectly predicted dark and pale skin, emphasizing the importance of careful interpretation of final predictions. Future studies considering quantitative pigmentation may achieve more accurate predictions by not relying on categories. Furthermore, our findings suggested that not all previously established SNPs showed a significant association with pigmentation in our population. For instance, the number of markers used for eye color prediction could be reduced to four while still maintaining reasonable predictive accuracy within our population. Conclusions: Overall, our results suggest that it may be possible to reduce the number of SNPs used in some cases without compromising accuracy. However, further validation in larger and more diverse populations is essential to draw firm conclusions and make broader generalizations. Full article
(This article belongs to the Section Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Distribution of eye, hair, and skin pigmentation categories in our sample collection.</p>
Full article ">Figure 2
<p>Real examples of eye (<b>A</b>), hair (<b>B</b>), and skin (<b>C</b>) pigmentation represent our main categories (with a representation greater than 8%). (<b>A</b>) From left to right: blue, intermediate, and brown eyes. (<b>B</b>) From left to right: blond, brown, and black hair. (<b>C</b>) From left to right: pale and intermediate skin.</p>
Full article ">Figure 3
<p>ROC curves for eye (<b>A</b>), hair (<b>B</b>), and skin (<b>C</b>) pigmentation prediction models.</p>
Full article ">
Back to TopTop