[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

Search Results (6,162)

Search Parameters:
Keywords = cell wall

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 1343 KiB  
Article
Effect of Oscillating Magnetic Fields (OMFs) and Pulsed Electric Fields (PEFs) on Supercooling Preservation of Atlantic Salmon (Salmo salar L.) Fillets
by Dongyoung Lee, Jinwen Tang, Seung Hyun Lee and Soojin Jun
Foods 2024, 13(16), 2525; https://doi.org/10.3390/foods13162525 - 13 Aug 2024
Abstract
Salmon, rich in protein and omega-3 fatty acids, has a short shelf life of 1 to 3 days when stored at 2 to 8 °C. Freezing, used for long-term preservation, often results in ice crystal formation. Ice crystals can cause structural damage, leading [...] Read more.
Salmon, rich in protein and omega-3 fatty acids, has a short shelf life of 1 to 3 days when stored at 2 to 8 °C. Freezing, used for long-term preservation, often results in ice crystal formation. Ice crystals can cause structural damage, leading to cell wall rupture, which can affect the texture and cause nutrient loss. Ultimately, this process reduces the overall quality of the salmon. Supercooling, which cools food below its freezing temperature without forming ice crystals, offers an alternative. This study investigated the effects of oscillating magnetic fields (OMFs) and pulsed electric fields (PEFs) on ice crystal formation during salmon supercooling. The results showed that using OMFs and PEFs in supercooling reduced the storage temperature of salmon, maintaining a similar thiobarbituric acid reactive substances (TBARS) value to that of frozen and refrigerated samples. There was no significant difference in meat color between the fresh and frozen samples, and drip loss weight was comparable between the fresh and supercooled samples. The microbiological counts were the lowest in the supercooled samples compared to the frozen and refrigerated ones. These findings suggest that supercooling storage with OMFs and PEFs can mitigate quality degradation in salmon typically associated with freezing. Full article
(This article belongs to the Special Issue Application of Thermal/Non-thermal Technologies in the Food Field)
Show Figures

Figure 1

Figure 1
<p>The supercooling chamber has electromagnet coils for the OMF and titanium electrodes for the PEF. The sample was located between a pair of electrodes.</p>
Full article ">Figure 2
<p>Temperature plots of refrigerated, frozen, and supercooled salmon fillets.</p>
Full article ">Figure 3
<p>Mean values of drip loss in samples after 10 days of storage under refrigerated, frozen, and supercooled conditions. Different letters indicate significant differences within the same storage time (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Mean values of Warner–Bratzler shear force of samples after 10 days of storage under refrigerated, frozen, and supercooled conditions. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Mean values of degree of lipid oxidation of samples after 10 days of storage under refrigerated, frozen, and supercooled conditions. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Mean value of APC during 10 days of storage under refrigerated, frozen, and supercooled conditions. Different letters indicate significant differences according to Tukey’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
22 pages, 7839 KiB  
Article
Allele-Specific Hormone Dynamics in Highly Transgressive F2 Biomass Segregants in Sugarcane (Saccharum spp.)
by Noor-ul Ain, Habiba and Ray Ming
Plants 2024, 13(16), 2247; https://doi.org/10.3390/plants13162247 - 13 Aug 2024
Abstract
Sugarcane holds global promise as a biofuel feedstock, necessitating a deep understanding of factors that influence biomass yield. This study unravels the intricate dynamics of plant hormones that govern growth and development in sugarcane. Transcriptome analysis of F2 introgression hybrids, derived from the [...] Read more.
Sugarcane holds global promise as a biofuel feedstock, necessitating a deep understanding of factors that influence biomass yield. This study unravels the intricate dynamics of plant hormones that govern growth and development in sugarcane. Transcriptome analysis of F2 introgression hybrids, derived from the cross of Saccharum officinarum “LA Purple” and wild Saccharum robustum “MOL5829”, was conducted, utilizing the recently sequenced allele-specific genome of “LA Purple” as a reference. A total of 8059 differentially expressed genes were categorized into gene models (21.5%), alleles (68%), paralogs (10%), and tandemly duplicated genes (0.14%). KEGG analysis highlighted enrichment in auxin (IAA), jasmonic acid (JA), and abscisic acid (ABA) pathways, revealing regulatory roles of hormone repressor gene families (Aux/IAA, PP2C, and JAZ). Signaling pathways indicated that downregulation of AUX/IAA and PP2C and upregulation of JAZ repressor genes in high biomass segregants act as key players in influencing downstream growth regulatory genes. Endogenous hormone levels revealed higher concentrations of IAA and ABA in high biomass, which contrasted with lower levels of JA. Weighted co-expression network analysis demonstrated strong connectivity between hormone-related key genes and cell wall structural genes in high biomass genotypes. Expression analysis confirmed the upregulation of genes involved in the synthesis of structural carbohydrates and the downregulation of inflorescence and senescence-related genes in high biomass, which suggested an extended vegetative growth phase. The study underscores the importance of cumulative gene expression, including gene models, dominant alleles, paralogs, and tandemly duplicated genes and activators and repressors of disparate hormone (IAA, JA, and ABA) signaling pathways are the points of hormone crosstalk in contrasting biomass F2 segregants and could be applied for engineering high biomass acquiring varieties. Full article
(This article belongs to the Section Plant Genetics, Genomics and Biotechnology)
Show Figures

Figure 1

Figure 1
<p>Partial least squares-discriminant analysis (PLS-DA), Hierarchical clustering, and volcano plot of DEGs: (<b>A</b>) PLS-DA score plot of FPKM data of DEGs generated by using preprocessed original data shows the clustering of extreme biomass segregants. Component 1 (26%) clearly distinguishes the two biomass groups. (<b>B</b>) (<b>a</b>) Hierarchical clustering heatmap of the DEGs based on FPKM expression values, (<b>b</b>) Six clusters indicating the up and down regulated genes identified in heatmap. (<b>C</b>) Volcano plot overall shows the range of log2FC of DEGs, i.e., 24 to −26 from right to left.</p>
Full article ">Figure 2
<p>KEGG and GO analysis of DEGs: (<b>A</b>) Bar plot for the KEGG categories and enriched DEGs in KEGG analysis. The abscissa depicts enriched pathways while gene number is plotted on an ordinate axis. (<b>B</b>) GO shows the enrichment of potential DEGs in different categories, i.e., biological processes, cellular components, and molecular function.</p>
Full article ">Figure 3
<p>Heatmap of log2FC values in HB segregants. Red and black values in legend show down and upregulated genes. (<b>A</b>) shows ubiquitin-mediated signaling of auxin and jasmonic acid signaling pathways. (<b>B</b>) depicts the expression patterns of hormone-responsive growth-related genes in cell wall and terminal developmental phases, i.e., inflorescence and senescence. <span class="html-italic">S-40</span>: Senescence regulator, <span class="html-italic">ATHB</span>: ARABIDOPSIS THALIANA HOMEOBOX 7, <span class="html-italic">KNAT1</span>: BREVIPEDICELLUS, <span class="html-italic">PHOX2/ARIX</span>: Transcription factor PHOX2/ARIX, <span class="html-italic">ELF3</span>: EARLY FLOWERING 3 (<span class="html-italic">ELF3</span>), <span class="html-italic">FPF1</span>: FLOWERING PROMOTING FACTOR 1, <span class="html-italic">AGAMOUS-LIKE 12</span>: AGAMOUS-LIKE 12, <span class="html-italic">XXT</span>: galactosyl transferase GMA12/MNN10 family, <span class="html-italic">EXPANSIN</span>: EXPANSIN, <span class="html-italic">CESA</span>: cellulose synthase subfamily, <span class="html-italic">GAE</span>: GDP-mannose 4,6 dehydratase, <span class="html-italic">CSLA02</span>: belongs to the glycosyltransferase 2 family, <span class="html-italic">XTH</span>: Xyloglucan endohydrolysis (<span class="html-italic">XEH</span>) and or endotransglycosylation (<span class="html-italic">XET</span>), and <span class="html-italic">UGT</span>: UDP-glycosyltransferase family.</p>
Full article ">Figure 4
<p>Weighted gene co-expression network analysis (WGCNA) of DEGs between HB and LB segregants. (<b>A</b>) eigengene modules, module–trait relationship, and module–module relationship. Module–trait heatmap shows Pearson’s correlation of all the modules with samples, whereas in the module–module relationship, progressive saturation in blue and red color points to high co-expression interconnectedness. Additionally, it shows dendrogram of module clustering, with green and red horizontal lines representing threshold (0.25, 0.3). (<b>B</b>) Cytoscape network shows co-expression network of blue module, which depicts highly upregulated genes in HB segregants (<b>C</b>) Cystoscape network of overrepresented DEGs in module “greenyellow”. Size and color of nodes are proportional to weights, whereas edge colors correspond to module names.</p>
Full article ">Figure 5
<p>Heatmap generated using log2FC. Expression dynamics of TFs and PKs involved in high biomass samples. Green and blue scales represent up and downregulated TFs in HB samples, whereas grey color indicates blanks.</p>
Full article ">Figure 6
<p>(<b>A</b>) Concentrations of the endogenous hormones, i.e., ABA, JA, and IAA, in the leaves of high and low biomass F2 segregants. Bar charts present the means with error bars showing standard errors, different letters are based on one-way ANOVA and LSD tests at α = 0.05. (<b>B</b>) Linear regression model of hormone content and qPCR values of the genes identified in RNA-Seq. analysis in the respective signaling pathway.</p>
Full article ">Figure 7
<p>Confirmation of FPKM by qPCR expression. Green bars represent the relative expression in qPCR, and orange lines represent FPKM values in transcriptome for corresponding genes. Values on the <span class="html-italic">y</span>-axis indicate relative expression levels of qPCR (<b>left</b>) and RNA-Seq (<b>right</b>). Error bars show standard error of the means at (<span class="html-italic">p</span> &lt; 0.05), and “r” is indicative of correlation between qPCR and FPKM expression values.</p>
Full article ">
17 pages, 2154 KiB  
Review
Plasmodesmata Function and Callose Deposition in Plant Disease Defense
by Jingsheng Chen, Xiaofeng Xu, Wei Liu, Ziyang Feng, Quan Chen, You Zhou, Miao Sun, Liping Gan, Tiange Zhou and Yuanhu Xuan
Plants 2024, 13(16), 2242; https://doi.org/10.3390/plants13162242 - 13 Aug 2024
Viewed by 4
Abstract
Callose, found in the cell walls of higher plants such as β-1,3-glucan with β-1,6 branches, is pivotal for both plant development and responses to biotic and abiotic stressors. Plasmodesmata (PD), membranous channels linking the cytoplasm, plasma membrane, and endoplasmic reticulum of adjacent cells, [...] Read more.
Callose, found in the cell walls of higher plants such as β-1,3-glucan with β-1,6 branches, is pivotal for both plant development and responses to biotic and abiotic stressors. Plasmodesmata (PD), membranous channels linking the cytoplasm, plasma membrane, and endoplasmic reticulum of adjacent cells, facilitate molecular transport, crucial for developmental and physiological processes. The regulation of both the structural and transport functions of PD is intricate. The accumulation of callose in the PD neck is particularly significant for the regulation of PD permeability. This callose deposition, occurring at a specific site of pathogenic incursion, decelerates the invasion and proliferation of pathogens by reducing the PD pore size. Scholarly investigations over the past two decades have illuminated pathogen-induced callose deposition and the ensuing PD regulation. This gradual understanding reveals the complex regulatory interactions governing defense-related callose accumulation and protein-mediated PD regulation, underscoring its role in plant defense. This review systematically outlines callose accumulation mechanisms and enzymatic regulation in plant defense and discusses PD’s varied participation against viral, fungal, and bacterial infestations. It scrutinizes callose-induced structural changes in PD, highlighting their implications for plant immunity. This review emphasizes dynamic callose calibration in PD constrictions and elucidates the implications and potential challenges of this intricate defense mechanism, integral to the plant’s immune system. Full article
(This article belongs to the Special Issue Plant Pathology and Epidemiology for Grain, Pulses, and Cereal Crops)
Show Figures

Figure 1

Figure 1
<p>Significance of callose in the physiology of plants and fungi, including its role in pollen maturation, pollen-selective fertilization, gametophyte development, PD permeability regulation, cell plate formation, sieve pore size, and biotic and abiotic stresses process, and it is a component of fungal cell walls.</p>
Full article ">Figure 2
<p>Enzymatic regulation of callose accumulation in plants. GSLs/CalS, enzymes presumed to facilitate the catalysis of 1,3-β glucan polymer synthesis from UDP glucose; BGs, responsible for the degradation of callose; PDCB, enhances the stability of callose through binding interactions.</p>
Full article ">Figure 3
<p>Dynamic responses of PD-related proteins to pathogenic intrusions: viral, fungal, and bacterial. MPs are known to facilitate viral particle transport via PD; when a PD presents a barrier to viral invasion in the host, the virus generates MPs specifically targeted at the PD, creating an aperture in the PD that allows the viral particle to pass through. Effectors secreted by bacteria and fungi during invasion of the host can travel through stomata to other parts of the host’s body. The distribution of effectors is strongly hindered when callose builds up in the PD neck region to decrease the PD pore size. Augmented callose deposition leads to PD constriction, thereby impeding the transit of bacteria, fungi, viruses, and their corresponding effectors.</p>
Full article ">Figure 4
<p>Signaling pathways regulating callose deposition and PD permeability upon fungal and bacterial infection in plants.</p>
Full article ">
16 pages, 673 KiB  
Review
Production of Bioactive Peptides from Microalgae and Their Biological Properties Related to Cardiovascular Disease
by Ranitha Fernando, Xiaohong Sun and H. P. Vasantha Rupasinghe
Macromol 2024, 4(3), 582-597; https://doi.org/10.3390/macromol4030035 (registering DOI) - 12 Aug 2024
Viewed by 140
Abstract
Microalgae are a substantial group of unicellular prokaryotic and eukaryotic marine organisms. Due to their high protein content of 50–70%, microalgae have the potential to become a sustainable alternative protein source, as well as aiding in the development of bioactive peptide-based nutraceuticals. A [...] Read more.
Microalgae are a substantial group of unicellular prokaryotic and eukaryotic marine organisms. Due to their high protein content of 50–70%, microalgae have the potential to become a sustainable alternative protein source, as well as aiding in the development of bioactive peptide-based nutraceuticals. A series of major steps are involved in the production of peptides from microalgae, which include the disruption of the microalgal cell wall, the hydrolysis of proteins, and the extraction or isolation of peptides derived from hydrolysis. Physical methods of cell wall disruptions are favored due to the ability to obtain high-quality protein fractions for peptide production. Bioactive peptides are protein fragments of two to twenty amino acid residues that have a beneficial impact on the physiological functions or conditions of human health. Strong scientific evidence exists for the in vitro antioxidant, antihypertensive, and anti-atherosclerotic properties of microalgal peptides. This review is aimed at summarizing the methods of producing microalgal peptides, and their role and mechanisms in improving cardiovascular health. The review reveals that the validation of the physiological benefits of the microalgal peptides in relation to cardiovascular disease, using human clinical trials, is required. Full article
9 pages, 724 KiB  
Brief Report
Cell Wall Profiling of the Resurrection Plants Craterostigma plantagineum and Lindernia brevidens and Their Desiccation-Sensitive Relative, Lindernia subracemosa
by John P. Moore, Brock Kuhlman, Jeanett Hansen, Leonardo Gomez, Bodil JØrgensen and Dorothea Bartels
Plants 2024, 13(16), 2235; https://doi.org/10.3390/plants13162235 - 12 Aug 2024
Viewed by 259
Abstract
Vegetative desiccation tolerance has evolved within the genera Craterostigma and Lindernia. A centre of endemism and diversification for these plants appears to occur in ancient tropical montane rainforests of east Africa in Kenya and Tanzania. Lindernia subracemosa, a desiccation-sensitive relative of Craterostigma [...] Read more.
Vegetative desiccation tolerance has evolved within the genera Craterostigma and Lindernia. A centre of endemism and diversification for these plants appears to occur in ancient tropical montane rainforests of east Africa in Kenya and Tanzania. Lindernia subracemosa, a desiccation-sensitive relative of Craterostigma plantagineum, occurs in these rainforests and experiences adequate rainfall and thus does not require desiccation tolerance. However, sharing this inselberg habitat, another species, Lindernia brevidens, does retain vegetative desiccation tolerance and is also related to the resurrection plant C. plantagineum found in South Africa. Leaf material was collected from all three species at different stages of hydration: fully hydrated (ca. 90% relative water content), half-dry (ca. 45% relative water content) and fully desiccated (ca. 5% relative water content). Cell wall monosaccharide datasets were collected from all three species. Comprehensive microarray polymer profiling (CoMPP) was performed using ca. 27 plant cell-wall-specific antibodies and carbohydrate-binding module probes. Some differences in pectin, xyloglucan and extension epitopes were observed between the selected species. Overall, cell wall compositions were similar, suggesting that wall modifications in response to vegetative desiccation involve subtle cell wall remodelling that is not reflected by the compositional analysis and that the plants and their walls are constitutively protected against desiccation. Full article
(This article belongs to the Special Issue New Perspectives on the Plant Cell Wall)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Monosaccharide compositional analysis of the total AIR isolated from leaf material of <span class="html-italic">Craterostigma plantagineum</span> (<b>a</b>), and <span class="html-italic">Lindernia brevidens</span> (<b>b</b>). White bars represent hydrated leaves, mid-grey shaded bars represent partially hydrated leaves and shaded bars represent desiccated leaves. Monosaccharide codes are for arabinose (Ara), rhamnose (Rha), fucose (Fuc), xylose (Xyl), mannose (Man), galactose (Gal), galacturonic acid (GalUA), glucose (Glc) and glucuronic acid (GlcUA). Error bars represent the standard error (SE) of the mean of four biological samples with two technical replicates per biological sample. Statistically significant differences, based on one-way ANOVA variance testing, are indicated on the bar graphs as an asterisk.</p>
Full article ">Figure 2
<p>Comprehensive microarray polymer profiling (CoMPP) analysis of plant-leaf cell wall fractions from CDTA-extractable material (<b>a</b>) and NaOH-extractable material (<b>b</b>) isolated from <span class="html-italic">Craterostigma plantagineum</span>, <span class="html-italic">Lindernia brevidens</span> and <span class="html-italic">Lindernia subracemosa</span> leaves that were hydrated (H), partially hydrated (PD) or desiccated (D). The heatmaps indicate the relative abundance of plant cell wall glycan-associated epitopes present in the AIR, and the colour intensity is correlated to the mean spot signals. The values in the heatmap are the mean spot signals from three experiments. The highest signal in the entire data set was set to 100, and all other data were adjusted accordingly.</p>
Full article ">
20 pages, 12372 KiB  
Article
Influence of Anatomical Spatial Architecture of Pinus devoniana on Pressure Gradients Inferred from Coupling Three-Dimensional CT Imaging and Numerical Flow Simulations
by Juan Gabriel Rivera-Ramos, José Cruz de León, Dante Arteaga, Raúl Espinoza-Herrera, Erica Arreola García, Manuel Arroyo-Albiter and Luis Olmos
Forests 2024, 15(8), 1403; https://doi.org/10.3390/f15081403 - 10 Aug 2024
Viewed by 367
Abstract
Conifer forests in Michoacán are facing climate change. Pinus devoniana Lindley, with natural distribution in the state, has shown certain adaptability, and knowing the influence of anatomy in the flow system is essential to delimit how it contributes to safety margins and water [...] Read more.
Conifer forests in Michoacán are facing climate change. Pinus devoniana Lindley, with natural distribution in the state, has shown certain adaptability, and knowing the influence of anatomy in the flow system is essential to delimit how it contributes to safety margins and water efficiency. For this, the pressure gradients in the cell lumens and their ramifications were analyzed by numerical simulations of flow throughout the real microstructure. Xylem were evaluated in radial, tangential and longitudinal directions. With the skeletonization of lumens and their constrictions, a branching system of interconnection between tracheids, ray cells, intercellular chambers, extensions, and blind pits were identified. In the simulation, the branched system bypasses the longitudinal fluid passage through the pores in membranes of pairs of pits to redirect it through the direct path branching, contributing to safety margins and water efficiency. Thus, resilience at low pressures because of the lower pressure drop in the extensions. The interface between the branching system and the cell lumens are sites of higher pressure gradient, more conducive to water-vapor formation or air leakage in the face of the lowest pressure system. The flow lines move along easy paths, regardless of the simulated flow direction. Deposits in the cell extensions were shown to be attached to the S3 layer of the cell wall, leaving the center of the duct free to flow. It is concluded that the spatial architecture of the xylem anatomy of Pinus dvoniana is a factor in the resilience at low pressures due to high water stress of the species. Full article
Show Figures

Figure 1

Figure 1
<p>Sample preparation: (<b>a</b>) <span class="html-italic">Pinus devoniana</span> tree, (<b>b</b>) sample extraction, and (<b>c</b>) specimens used for the study.</p>
Full article ">Figure 2
<p>CT image processing: (<b>a</b>) initial image, (<b>b</b>) 3D-filtered initial image, (<b>c</b>) binary image, (<b>d</b>) tangential slice, (<b>e</b>) radial slice view of uncompleted fibers due to the angle between the fiber inclination and the crop section, (<b>f</b>) cross section, (<b>g</b>) the yellow rectangle is the ROI extracted from a 3D image acquired with 4 µm voxel, (<b>h</b>) color distribution.</p>
Full article ">Figure 3
<p>Reconstruction of continuity of cell cavities, spaces and cell extensions in microstructure of <span class="html-italic">Pinus devoniana</span> wood CT images: (<b>a</b>) continuity between lumen, pit, extensions spaces and extensions chambers, (<b>b</b>) types of extensions, (<b>c</b>) continuity in crossing fields, (<b>d</b>) chambers between extensions spaces, (<b>e</b>) checking of branches in transverse microstructure, (<b>f</b>) checking of branches in tracheid overlap zone microstructure; Parenchyma cavity (PC), extension (E), tracheid cavity (TC), blind pits (BP), chambers in BS (Ch), branched system (BS).</p>
Full article ">Figure 4
<p>CT images of the microstructure of <span class="html-italic">Pinus devoniana</span> wood: (<b>a</b>) cell cavities in the microstructure, (<b>b</b>) <span class="html-italic">Pinus devoniana</span> cell wall layers identified with CT, (<b>c</b>) connectivity of extensions and wall roughness, (<b>d</b>) tracheid corner extension; middle lamella (ML), extension (E), wood layer (S2) and (S3), parenchyma cavity (PC), tracheid cavity (TC), pit (P).</p>
Full article ">Figure 5
<p>Pressure gradient plots simulated of <span class="html-italic">Pinus devoniana</span> wood microstructure issues from CT images: (<b>a</b>) flow pressure drop on entering the tracheid, (<b>b</b>) pressure gradient in longitudinal flow direction, (<b>c</b>) different pressure gradient on the extensions, (<b>d</b>) similar pressure gradient between epithelial cells, tracheids, pits, and parenchyma; extension (E), tracheid (T), pits (P).</p>
Full article ">Figure 6
<p>Pressure plots in BS of <span class="html-italic">Pinus devoniana</span> wood microstructure from CT images: (<b>a</b>) longitudinal flow pressure drop in tracheids and extensions, (<b>b</b>) pressure drop in BS in a high pressure system; extension (E), tracheid (T), pits (P).</p>
Full article ">Figure 7
<p>Pressure plots in radial flow of <span class="html-italic">Pinus devoniana</span> wood microstructure from CT images: (<b>a</b>) flow pressure drop on entering the tracheid, (<b>b</b>) pressure gradient in radial flow direction; tracheid (T), pits (P).</p>
Full article ">Figure 8
<p>Pressure plots in tangential flow of <span class="html-italic">Pinus devoniana</span> wood microstructure from CT images: (<b>a</b>) pits in crossing fields, (<b>b</b>) pressure gradient in extensions in tangential flow direction; extension (E).</p>
Full article ">Figure 9
<p>Connection of blind pits with extensions on <span class="html-italic">Pinus devoniana</span> wood from CT images: (<b>a</b>) continuity of the tracheid lumen through the extensions, (<b>b</b>) connection of the extensions with the ray; extension (E), tracheid (T).</p>
Full article ">Figure 10
<p>Pressure plots and flow lines inside of the microstructure of <span class="html-italic">Pinus devoniana</span> wood from CT images: (<b>a</b>) pressure within chambers linking the extension connected in turn to blind pit chambers between two tracheids, (<b>b</b>) longitudinal flow lines avoiding the pit pairs and continuing through the extensions; extension (E), tracheid (T), branched system (BS).</p>
Full article ">Figure 11
<p>Microstructure of <span class="html-italic">Pinus devoniana</span> wood in 4 µm voxel from CT images; (<b>a</b>) early wood and late wood, (<b>b</b>) obstacles to flow, (<b>c</b>) relationship between pressure drop and diameter of cell lumens. Late Xylem (XL), Early Xylem (XE).</p>
Full article ">Figure 12
<p>Radial flow in microstructure of <span class="html-italic">Pinus devoniana</span> wood from CT images.</p>
Full article ">Figure 13
<p>Tangential flow in microstructure of <span class="html-italic">Pinus devoniana</span> wood from CT images: (<b>a</b>) flow lines in ray-extension continuation, (<b>b</b>) periodic flow in tracheid rays and extensions; resiniferous channel (Rch), extension (E).</p>
Full article ">
19 pages, 5194 KiB  
Article
Effect of Polyethylene Glycol with Different Molecular Weights on the Properties of Mytilaria laosensis Timber
by Linhan He, Xiaoling Liu, Yin Kuang, Liuliu Wang, Songwu Chen, Yufen Luo and Yao Chen
Forests 2024, 15(8), 1401; https://doi.org/10.3390/f15081401 - 10 Aug 2024
Viewed by 208
Abstract
Mytilaria laosensis, a common fast-growing tree species in southern China, boasts excellent growth speed and attractive color and texture. However, due to its short growth cycle and high proportion of juvenile wood, it typically exhibits poor dimensional stability and low strength, which [...] Read more.
Mytilaria laosensis, a common fast-growing tree species in southern China, boasts excellent growth speed and attractive color and texture. However, due to its short growth cycle and high proportion of juvenile wood, it typically exhibits poor dimensional stability and low strength, which significantly limits its practical applications. This study uses vacuum impregnation to modify M. laosensis wood with polyethylene glycol (PEG), focusing on the effects and mechanisms of PEG with different molecular weights on wood properties. The results indicate that PEG enters the wood cell walls through capillary action and diffusion, forming hydrogen bonds with the free hydroxyl groups on cellulose and hemicellulose, which keeps the cell walls swollen and enhances dimensional stability. Post modification, the dimensional stability of M. laosensis wood improved, with an anti-swelling efficiency ranging from 61.43% to 71.22%, showing an initial increase followed by a decrease with increasing PEG molecular weight. The optimal PEG molecular weight for anti-swelling efficiency was 1500 Da, achieving 71.22%. The flexural modulus of elasticity and flexural strength of the treated wood also first decreased and then increased with increasing PEG molecular weight. Among them, the PEG1000-treated material showed the best performance, with the flexural modulus of elasticity increased by about 29% and the flexural strength increased by about 5% compared to untreated wood. Additionally, PEG, having a higher pyrolysis temperature than wood, raised the initial pyrolysis temperature and maximum pyrolysis rate temperature of M. laosensis wood, thus improving its thermal stability. These findings provide scientific evidence and technical support for the efficient utilization and industrialization of M. laosensis wood, promoting its widespread application and industrial development. Full article
(This article belongs to the Section Wood Science and Forest Products)
Show Figures

Figure 1

Figure 1
<p>Flow chart of PEG-modified <span class="html-italic">M. laosensis</span> timber.</p>
Full article ">Figure 2
<p>WPG of <span class="html-italic">Mytilaria laosensis</span> modified by PEG of different molecular weights. Significant differences between treatments are indicated by different letters.</p>
Full article ">Figure 3
<p>(<b>a</b>) Dry shrinkage rate and (<b>b</b>) wet swelling rate of <span class="html-italic">Mytilaria laosensis</span> in different treatment conditions.</p>
Full article ">Figure 4
<p>Relationship between PEG molecular weight and ASE. Significant differences between treatments are indicated by different letters.</p>
Full article ">Figure 5
<p>(<b>a</b>) The schematic diagram of VCC and ASE; (<b>b</b>) relationships with VCC and principal causes; (<b>c</b>) VCC values of PEG-treated woods.</p>
Full article ">Figure 6
<p>The change in water absorption of PEG-modified wood with different PEG molecular weight.</p>
Full article ">Figure 7
<p>Modulus of bending elasticity of PEG-treated woods with different molecular weight. Significant differences between treatments are indicated by different letters.</p>
Full article ">Figure 8
<p>Bending strength of PEG-treated woods with different molecular weight. Significant differences between treatments are indicated by different letters.</p>
Full article ">Figure 9
<p>(<b>a</b>) TG and (<b>b</b>) DTG curves of untreated and PEG-treated wood.</p>
Full article ">Figure 10
<p>FTIR spectra of PEG-treated samples with different molecular weights.</p>
Full article ">Figure 11
<p>XRD spectra of PEG-treated samples with different molecular weights.</p>
Full article ">Figure 12
<p>SEM images of untreated wood (<b>a</b>,<b>b</b>), PEG400-treated wood (<b>c</b>,<b>d</b>), PEG600-treated wood (<b>e</b>,<b>f</b>), PEG800-treated wood (<b>g</b>,<b>h</b>), PEG1000-treated wood (<b>i</b>,<b>j</b>), PEG1500-treated wood (<b>k</b>,<b>l</b>), and PEG2000-treated wood (<b>m</b>,<b>n</b>) in cross-sections (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>,<b>m</b>) and longitudinal sections (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>,<b>n</b>).</p>
Full article ">Figure 13
<p>Mechanism of the PEG-modified wood.</p>
Full article ">
14 pages, 3333 KiB  
Article
Discovery of Antibacterial Compounds with Potential Multi-Pharmacology against Staphylococcus Mur ligase Family Members by In Silico Structure-Based Drug Screening
by Mio Teshima, Kohei Monobe, Saya Okubo and Shunsuke Aoki
Molecules 2024, 29(16), 3792; https://doi.org/10.3390/molecules29163792 - 10 Aug 2024
Viewed by 313
Abstract
Staphylococcus aureus (S. aureus) is a major bacterial infection in humans, leading to severe disease and causing death. The stagnation of antibiotic development in recent decades has made it difficult to combat drug-resistant infections. In this study, we performed an in [...] Read more.
Staphylococcus aureus (S. aureus) is a major bacterial infection in humans, leading to severe disease and causing death. The stagnation of antibiotic development in recent decades has made it difficult to combat drug-resistant infections. In this study, we performed an in silico structure-based drug screening (SBDS) targeting the S. aureus MurE (saMurE) enzyme involved in cell wall synthesis of S. aureus. saMurE is an enzyme that is essential for the survival of S. aureus but not present in humans. SBDS identified nine saMurE inhibitor candidates, Compounds 19, from a structural library of 154,118 compounds. Among them, Compound 2 showed strong antibacterial activity against Staphylococcus epidermidis (S. epidermidis) used as a model bacterium. Amino acid sequence homology between saMurE and S. epidermidis MurE is 87.4%, suggesting that Compound 2 has a similar inhibitory effect on S. aureus. Compound 2 showed an IC50 value of 301 nM for S. epidermidis in the dose-dependent growth inhibition assay. Molecular dynamics simulation showed that Compound 2 binds stably to both S. aureus MurD and S. aureus MurF, suggesting that it is a potential multi-pharmacological pharmacological inhibitor. The structural and bioactivity information of Compound 2, as well as its potential multiple-target activity, could contribute to developing new antimicrobial agents based on MurE inhibition. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>saMurE inhibitor screening strategy.</p>
Full article ">Figure 2
<p>Growth inhibitory effect of Compounds <b>1</b>–<b>9</b> on bacteria (<span class="html-italic">S. epidermidis</span>). A total of 0.3% DMSO and 100 μM ampicillin (AMP) were used as samples for comparison. Compounds <b>1</b>–<b>9</b> (100 μM). The vertical axis is the mean +/− SEM of the results of four independent experiments. Dunnett’s test: ****; <span class="html-italic">p</span> &lt; 0.0001; **; <span class="html-italic">p</span> &lt; 0.0021; *; <span class="html-italic">p</span> &lt; 0.0332; n.s. = not significant.</p>
Full article ">Figure 3
<p>Determination of 50% growth inhibition concentration against bacteria (<span class="html-italic">S. epidermidis</span>). The vertical axis shows the relative bacterial growth rate. The horizontal axis shows the molar concentration of Compound <b>2</b>.</p>
Full article ">Figure 4
<p>MDS results for the saMurE–Compound <b>2</b> complex: (<b>A</b>) Transition of ligand RMSD value (nm). Ligand RMSD values were calculated by comparison with the post-equilibration pose. (<b>B</b>) Radius (nm) of gyration during MDS. (<b>C</b>) Number of intermolecular hydrogen bonds.</p>
Full article ">Figure 5
<p>ProLIF analysis for the saMurE–Compound <b>2</b> interaction: (<b>A</b>) Interacting residues throughout the MDS timeframe. In the screening process, the piperazine group of Compound <b>2</b> is protonated. (<b>B</b>) A major (≥60% probability of presence) interaction residue group was observed throughout the entire period.</p>
Full article ">Figure 6
<p>Toxicity verification of Compound <b>2</b> on mammalian-derived cells: (<b>A</b>) COS-7 cells; (<b>B</b>) HepG2 cells. Negative control was 0.3% DMSO and positive control was 50 μM triclosan (TCS). Concentration of Compound <b>2</b> was 100 μM. Dunnett’s test: ***; <span class="html-italic">p</span> &lt; 0.0002; n.s. = not significant.</p>
Full article ">Figure 7
<p>MDS results for saMur ligase–Compound <b>2</b> complexes: (<b>A</b>) ligand RMSD values (nm); (<b>B</b>) radius (nm) of gyration during MDS; (<b>C</b>) number of intermolecular hydrogen bonds. saMurC (black), saMurD (red), saMurF (green).</p>
Full article ">
12 pages, 1929 KiB  
Article
Targeting N-Acetylglucosaminidase in Staphylococcus aureus with Iminosugar Inhibitors
by Janja Sluga, Tihomir Tomašič, Marko Anderluh, Martina Hrast Rambaher, Gregor Bajc, Alen Sevšek, Nathaniel I. Martin, Roland J. Pieters, Marjana Novič and Katja Venko
Antibiotics 2024, 13(8), 751; https://doi.org/10.3390/antibiotics13080751 - 10 Aug 2024
Viewed by 409
Abstract
Bacteria are capable of remarkable adaptations to their environment, including undesirable bacterial resistance to antibacterial agents. One of the most serious cases is an infection caused by multidrug-resistant Staphylococcus aureus, which has unfortunately also spread outside hospitals. Therefore, the development of new [...] Read more.
Bacteria are capable of remarkable adaptations to their environment, including undesirable bacterial resistance to antibacterial agents. One of the most serious cases is an infection caused by multidrug-resistant Staphylococcus aureus, which has unfortunately also spread outside hospitals. Therefore, the development of new effective antibacterial agents is extremely important to solve the increasing problem of bacterial resistance. The bacteriolytic enzyme autolysin E (AtlE) is a promising new drug target as it plays a key role in the degradation of peptidoglycan in the bacterial cell wall. Consequently, disruption of function can have an immense impact on bacterial growth and survival. An in silico and in vitro evaluation of iminosugar derivatives as potent inhibitors of S. aureus (AtlE) was performed. Three promising hit compounds (1, 3 and 8) were identified as AtlE binders in the micromolar range as measured by surface plasmon resonance. The most potent compound among the SPR response curve hits was 1, with a KD of 19 μM. The KD value for compound 8 was 88 μM, while compound 3 had a KD value of 410 μM. Full article
(This article belongs to the Special Issue Recent Advances in Antimicrobial Drug Discovery, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) 3D binding model of the NAG-NAM central unit on the AtlE surface (PDB ID: 4PI7); (<b>b</b>) 2D modeled interactions of the NAG-NAM central unit with AtlE (red residue represents the hydrogen bond acceptor, green residue represents the hydrogen bond donor).</p>
Full article ">Figure 2
<p>(<b>a</b>) 3D binding model of compounds <b>1</b>, <b>3</b> and <b>8</b> on the AtlE surface (PDB ID: 4PI7); (<b>b</b>) 2D modeled interactions of compounds <b>1</b>, <b>3</b> and <b>8</b> with AtlE (red residues represent the hydrogen bond acceptors, green residues represent the hydrogen bond donor, and yellow residues represent the hydrophobic interactions).</p>
Full article ">Figure 3
<p>Representative SPR sensorgram (response curves) for compound 1-deoxynojirimycin at two different concentrations.</p>
Full article ">Figure 4
<p>(<b>a</b>) Representative SPR sensorgrams (response curves) and (<b>b</b>) representative saturation curves with evaluated <span class="html-italic">K</span><sub>D</sub> for compounds <b>1</b>, <b>3</b> and <b>8</b> at different concentrations. SPR analysis of compound <b>1</b>, <b>3</b> and <b>8</b> interactions with the immobilized AtlE. Compounds were injected across immobilized AtlE in serial dilutions for 60 s at a rate of 30 mL/min, and the dissociation was followed for 50 s. Sensorgrams are shown along with the apparent equilibrium dissociation constant (<span class="html-italic">K<sub>D</sub></span>) determined from the response curves as a function of the compound concentration injected across AtlE. <span class="html-italic">K<sub>D</sub></span> values are the mean ± standard deviation of three titrations. The data were fitted to the steady-state affinity binding model.</p>
Full article ">
9 pages, 4547 KiB  
Case Report
Renal Angiomyolipoma with Tumor Thrombus in the Inferior Vena Cava and Right Atrium Accompanied by Renal Cell Carcinoma: A Case Report
by Fan Shu, Yichang Hao, Ye Yan, Min Lu, Lulin Ma, Shaohui Deng, Liyuan Ge and Shudong Zhang
Medicina 2024, 60(8), 1293; https://doi.org/10.3390/medicina60081293 - 10 Aug 2024
Viewed by 211
Abstract
Background: Renal angiomyolipoma (AML) without local invasion is generally considered benign. However, it may extend to the renal sinus, even the renal vein, or the inferior vena cava (IVC). In patients with non-tuberous sclerosis complex, coexistence of renal cell carcinoma (RCC) and renal [...] Read more.
Background: Renal angiomyolipoma (AML) without local invasion is generally considered benign. However, it may extend to the renal sinus, even the renal vein, or the inferior vena cava (IVC). In patients with non-tuberous sclerosis complex, coexistence of renal cell carcinoma (RCC) and renal AML is uncommon. Case presentation: A 72-year-old woman was incidentally found to have a solitary right renal mass with an IVC thrombus extending into the right atrium during a routine health checkup. Robot-assisted laparoscopic radical nephrectomy and thrombectomy were successfully performed through adequate preoperative examination and preparation. Two tumor lesions were found and pathologically confirmed as renal AML and RCC, and the tumor thrombus was derived from the renal AML. During the one-year follow-up period, no signs of recurrence or metastatic disease were observed. Conclusions: Renal AML with a tumor thrombus in the IVC and right atrium accompanied by RCC may occur, although rarely. In clinical practice, if preoperative manifestations differ from those of common diseases, rare diseases must be considered to avoid missed diagnoses. In addition, adequate examination and multidisciplinary discussions before making a diagnosis are necessary. For a level 4 tumor thrombus with no infringement of the venous wall, adoption of robot-assisted minimally invasive surgery, without extracorporeal circulation technology, is feasible. Full article
(This article belongs to the Section Oncology)
Show Figures

Figure 1

Figure 1
<p>Abdominal contrast-enhanced computed tomography. (<b>A</b>) Rounded and high-density region (red arrow, renal cell carcinoma) measuring approximately 2.5 cm × 2.2 cm × 2.1 cm near the capsule area of the kidney (coronal plane). (<b>B</b>) Hilar angiomyolipoma lesion (red arrow) with a density close to that of adipose tissue, which was not identified preoperatively (coronal plane). (<b>C</b>) Low-density, linearly elongated tumor thrombus (red arrow) in the inferior vena cava (coronal plane). (<b>D</b>) Renal cell carcinoma lesion (red arrow) near the capsule of the kidney (horizontal plane). (<b>E</b>) Angiomyolipoma lesion (red arrow) in the renal hilum (horizontal plane).</p>
Full article ">Figure 2
<p>Color ultrasound Doppler of the inferior vena cava. (<b>A</b>) Hypoechoic nodule with a cross-sectional area of 2.41 cm × 2.13 cm in the right kidney (The colored dashed lines represented the two diametral lines of a cross section in different directions). (<b>B</b>) Linearly slender tumor thrombus along the direction of the extension of the inferior vena cava (The colored dashed line marked a distance traveled by the tumor thrombus). (<b>C</b>) The maximum diameter (colored dashed line) of the tumor thrombus in the inferior vena cava is approximately 0.8 cm. (<b>D</b>) Tumor thrombus protruding into the right atrium, with a large, round, and blunt tail (yellow arrow); RA: right atrium, RV: right ventricle.</p>
Full article ">Figure 3
<p>Kidney and tumor thrombus specimens. (<b>A</b>) Intact kidney resected along with the tumor thrombus. The tumor thrombus was segmented, and the middle segment was slender (red arrow). (<b>B</b>) The kidney dissected along the opposite side of the hilum. A round lesion (RCC) was observed on cross-section; the transverse section area was grayish brown and the focal area closely adhered to the perirenal fat. (<b>C</b>,<b>D</b>) The kidney dissected along the hilar side. A fat-rich lesion (AML) was found (blue arrow); it had a cord-like growth, and its cross-section area appeared golden brown.</p>
Full article ">
27 pages, 6716 KiB  
Article
Comparative Metabolome and Transcriptome Analysis of Rapeseed (Brassica napus L.) Cotyledons in Response to Cold Stress
by Xinhong Liu, Tonghua Wang, Ying Ruan, Xiang Xie, Chengfang Tan, Yiming Guo, Bao Li, Liang Qu, Lichao Deng, Mei Li and Chunlin Liu
Plants 2024, 13(16), 2212; https://doi.org/10.3390/plants13162212 - 9 Aug 2024
Viewed by 231
Abstract
Cold stress affects the seed germination and early growth of winter rapeseed, leading to yield losses. We employed transmission electron microscopy, physiological analyses, metabolome profiling, and transcriptome sequencing to understand the effect of cold stress (0 °C, LW) on the cotyledons of cold-tolerant [...] Read more.
Cold stress affects the seed germination and early growth of winter rapeseed, leading to yield losses. We employed transmission electron microscopy, physiological analyses, metabolome profiling, and transcriptome sequencing to understand the effect of cold stress (0 °C, LW) on the cotyledons of cold-tolerant (GX74) and -sensitive (XY15) rapeseeds. The mesophyll cells in cold-treated XY15 were severely damaged compared to slightly damaged cells in GX74. The fructose, glucose, malondialdehyde, and proline contents increased after cold stress in both genotypes; however, GX74 had significantly higher content than XY15. The pyruvic acid content increased after cold stress in GX74, but decreased in XY15. Metabolome analysis detected 590 compounds, of which 32 and 74 were differentially accumulated in GX74 (CK vs. cold stress) and XY15 (CK vs. cold stressed). Arachidonic acid and magnoflorine were the most up-accumulated metabolites in GX74 subjected to cold stress compared to CK. There were 461 and 1481 differentially expressed genes (DEGs) specific to XY15 and GX74 rapeseeds, respectively. Generally, the commonly expressed genes had higher expressions in GX74 compared to XY15 in CK and cold stress conditions. The expression changes in DEGs related to photosynthesis-antenna proteins, chlorophyll biosynthesis, and sugar biosynthesis-related pathways were consistent with the fructose and glucose levels in cotyledons. Compared to XY15, GX74 showed upregulation of a higher number of genes/transcripts related to arachidonic acid, pyruvic acid, arginine and proline biosynthesis, cell wall changes, reactive oxygen species scavenging, cold-responsive pathways, and phytohormone-related pathways. Taken together, our results provide a detailed overview of the cold stress responses in rapeseed cotyledons. Full article
(This article belongs to the Special Issue Genetics and Genomics of Crop Breeding and Improvement)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Visual comparison of injury to the cotyledons of cold-sensitive (XY15) and -tolerant (GX74) <span class="html-italic">B. rapa</span> in response to cold stress (LW) compared to control (CK). (<b>B</b>) Transmission electron microscopic observations of mesophyll cells of XY15 and GX74 before and after cold stress treatment. Mi = mitochondria, CP = chloroplast, red arrow = cell wall, orange arrow = starch grains, green arrow = osmiophilic granules, blue arrow = crenellate, purple arrow = vacuole, and yellow arrow = lamellar. The red circles show that the outer membrane of the mesophyll cells is dissolved/ruptured.</p>
Full article ">Figure 2
<p>Physiological performance of cold-tolerant (GX74) and -sensitive (XY15) genotypes of <span class="html-italic">B. rapa</span> cotyledons before (LW) and after (18 h) of cold stress (LW). (<b>A</b>). Fructose content. (<b>B</b>). Glucose content. (<b>C</b>). Malondialdehyde content. (<b>D</b>). Hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) content. (<b>E</b>). Relative electrical conductivity. (<b>F</b>). Pyruvic acid content. (<b>G</b>). Proline content. Bars of a genotype with * and ns indicate significant and non-significant difference, respectively, between CK and LW. Error bars represent standard error of means of three replicates.</p>
Full article ">Figure 3
<p>Metabolome analysis of <span class="html-italic">B. rapa</span> (XY15 and GX74) cotyledons challenged with cold stress. (<b>A</b>) Hierarchical heatmap clusters, (<b>B</b>) principal component analysis, and (<b>C</b>) Pearson’s correlation coefficient analyses of the detected metabolites in negative (upper panel) and positive modes (lower panel). (<b>D</b>–<b>F</b>) Differential metabolome analysis of <span class="html-italic">B. rapa</span> (XY15 and GX74) cotyledons in response to cold stress. (<b>A</b>) Venn diagram showing differential metabolites in GX74 (CKvsLW) and XY15 (CKvsLW). Log2 fold change of the highly up- and down-accumulated metabolites in (<b>B</b>) GX74 (CKvsLW) and (<b>C</b>) XY15 (CKvsLW).</p>
Full article ">Figure 4
<p>Overview of gene expression in <span class="html-italic">B. rapa</span> cotyledons in response to cold stress. (<b>A</b>) Distribution of average TPM and (<b>B</b>) principal component analysis in GX74 and XY15 before (CK) and after cold stress (LW). The numbers (1–3) with treatments indicate replicates. (<b>C</b>) Summary of differentially expressed genes and (<b>D</b>) Venn diagram of differentially expressed genes in GX74 and XY15 before and after cold stress. The up and down arrows indicate the number of up- or downregulated genes in cold treated samples compared to CK. KEGG pathway (top 20) enrichment barplots of DEGs in (<b>E</b>) GX74 (CKvsLW) and (<b>F</b>) XY15 (CKvsLW). <span class="html-italic">x</span>- and <span class="html-italic">y</span>-axis in (<b>E</b>,<b>F</b>) represent no. of DEGs and KEGG pathways. The color bars represent <span class="html-italic">p</span>-value (adjusted). The lower the <span class="html-italic">p</span>-value, the more significant the enrichment results.</p>
Full article ">Figure 5
<p>Effects of cold stress on gene expression related to physiological and biochemical changes, ion homeostasis, and ROS scavenging. The heatmaps are based on the log2 fold change values. The left and right columns of the heatmap represent XY15 (CKvsLW) and GX74 (CKvsLW), respectively. The colors of the boxes in the pathway correspond to the borders of the heatmaps.</p>
Full article ">Figure 6
<p>Expression changes in sugar biosynthesis-related KEGG pathways. The genes given as red semi-circles were differentially expressed in the <span class="html-italic">B. rapa</span> (XY15 and GX74) cotyledons before and after cold stress treatment. The heatmap on the right panel represents log2 fold change values of the DEGs enriched in the KEGG pathways; gene id is followed by a number (given as |x|) corresponding to the number given in the red semi-circle. The number is followed by gene annotation according to KEGG database.</p>
Full article ">Figure 7
<p>Expression changes in auxin, abscisic acid, and cold-responsive genes in <span class="html-italic">B. rapa</span> (XY15 and GX74) cotyledons in response to cold stress. The heatmaps represent log2 fold change values.</p>
Full article ">Figure 8
<p>Quantitative real-time PCR analysis of <span class="html-italic">B. napus</span> genes in XY15 and GX74 before and after cold stress. The bars are means of three replicates. The error bars represent standard deviation.</p>
Full article ">Figure 9
<p>Contrasting cold stress responses of GX74 and XY15 cotyledons as revealed by full-length transcriptome analyses. The red, green, and black dots show the up-, down-, and not regulated genes within each pathway. The reduction in red and green color intensity of the dots indicates a higher and lower number of DEGs involved in each pathway, respectively. The dark yellow circles indicate the metabolites were differentially accumulated related to those pathways.</p>
Full article ">
20 pages, 5565 KiB  
Article
Biocatalytic Screening of the Oxidative Potential of Fungi Cultivated on Plant-Based Resources
by Alina Kinner, Stephan Lütz and Katrin Rosenthal
AppliedChem 2024, 4(3), 282-301; https://doi.org/10.3390/appliedchem4030018 - 8 Aug 2024
Viewed by 240
Abstract
The environmental impacts of the postindustrial era, which rely on fossil fuels, have compelled a reconsideration of the future of energy and chemical industries. Fungi are a valuable resource for improving a circular economy through the enhanced valorization of biomass and plant waste. [...] Read more.
The environmental impacts of the postindustrial era, which rely on fossil fuels, have compelled a reconsideration of the future of energy and chemical industries. Fungi are a valuable resource for improving a circular economy through the enhanced valorization of biomass and plant waste. They harbor a great diversity of oxidative enzymes, especially in their secretome. Enzymatic breakdown of the plant cell wall complex and lignocellulosic biomass yields sugars for fermentation and biofuel production, as well as aromatic compounds from lignin that can serve as raw materials for the chemical industry. To harness the biocatalytic potential, it is essential to identify and explore wild-type fungi and their secretomes. This study successfully combined genome mining and activity screening to uncover the oxidative potential of a collection of underexploited ascomycetes and basidiomycetes. The heme peroxidase and laccase activities of four promising candidates, Bipolaris victoriae, Colletotrichum sublineola, Neofusicoccum parvum and Moesziomyces antarcticus, were investigated to gain a deeper insight into their enzyme secretion. Furthermore, a plant-based medium screening with the phytopathogen C. sublineola revealed that soybean meal is a beneficial component to trigger the production and secretion of enzymes that catalyze H2O2-dependent oxidations. These results demonstrate that understanding fungal secretomes and their enzymatic potential opens exciting avenues for sustainable biotechnological applications across various industries. Full article
Show Figures

Figure 1

Figure 1
<p>Taxonomic classification of the mined fungi, belonging to the phyla <span class="html-italic">Ascomycota</span> and <span class="html-italic">Basidiomycota</span>, based on NCBI Common Tree and visualized using the online tool ‘iTOL: Interactive Tree Of Life’. PROSITE entries (DyP: PS51404, heme peroxidase: PS00435 + PS00436, Laccase: PS00079 + PS00080, UPO: PS51405) were matched with protein sequences stored on UniProtKB and only sequences containing a signal peptide were considered. Strains highlighted in gray were selected for activity screening during shake flask cultivation. DyP, dye de-colorizing peroxidase; UPO, unspecific peroxygenase.</p>
Full article ">Figure 2
<p>Enzyme assays for the detection and quantification of different oxidoreductase activities by oxidation of ABTS (<b>A</b>), NBD (<b>B</b>), and veratryl alcohol (<b>C</b>). ABTS, 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid); NBD, 5-nitro-1,3-benzodioxole; and UPO, unspecific peroxygenase.</p>
Full article ">Figure 3
<p>Product concentration of veratraldehyde and 4-nitrocatechol. Values given are means and average deviations for duplicates. Peroxidase and peroxygenase activity were determined by the oxidation of 1 mM veratryl alcohol to veratraldehyde (black gradation) for 30 min and 1 mM NBD to 4-nitrocatechol (red gradation) for 20 min with 1 mM H<sub>2</sub>O<sub>2</sub> using culture supernatant during 250 mL shake flask cultivation of the fungi in SG medium at 24 °C for four weeks. The assays were each performed during two different cultivations, in which the fungi were cultivated in duplicates. Due to solid growth of the fungi, some samples could not be taken during the cultivation (marked with n.d.). Product concentration was quantified by LC-MS using an external standard (15–500 µM). All UniProtKB hits for the selected PROSITE entries of this genome mining are shown in brackets. n.d., not determined.</p>
Full article ">Figure 4
<p>Shake flask cultivation of <span class="html-italic">B. victoriae</span>, <span class="html-italic">C. sublineola</span>, <span class="html-italic">M. antarcticus</span>, and <span class="html-italic">N. parvum</span> in SG medium. Values given are means and average deviations for duplicates. (<b>A</b>) pH of the culture supernatant (dashed line) and conversion of 0.5 mM veratryl alcohol to veratraldehyde (circles; with H<sub>2</sub>O<sub>2</sub> in black, without H<sub>2</sub>O<sub>2</sub> in grey) for 30 min at 21 °C and oxidation of 0.3 mM ABTS (rectangles; with H<sub>2</sub>O<sub>2</sub> in dark green, without H<sub>2</sub>O<sub>2</sub> in light green) for 15 min at 25 °C. (<b>B</b>) The volumetric activity of the culture supernatant was determined by oxidation of ABTS for 15 min at 25 °C. (<b>C</b>) SDS-PAGE analysis of the 20-fold concentrated fungal culture supernatants. (<b>D</b>) Fungal cultures in SG medium during 500 mL shake flask cultivation at 24 °C and 100 rpm on day 6, 9, 12, 15, 19, and 22.</p>
Full article ">Figure 5
<p>Shake flask cultivation of <span class="html-italic">C. sublineola</span> in four complex plant-based media (SG, S, C, and BS medium). Values given are means and average deviations for duplicates. (<b>A</b>) pH of the culture supernatant (dashed line) and conversion of 0.5 mM veratryl alcohol to veratraldehyde (circles; with H<sub>2</sub>O<sub>2</sub> in black, without H<sub>2</sub>O<sub>2</sub> in grey) for 30 min at 21 °C and oxidation of 0.3 mM ABTS (rectangles; with H<sub>2</sub>O<sub>2</sub> in dark green, without H<sub>2</sub>O<sub>2</sub> in light green) for 15 min at 25 °C. Glucose levels were monitored using glucose test strips. The day of depletion in SG medium is indicated by a thin dashed line. (<b>B</b>) The volumetric activity of the culture supernatant was determined by oxidation of ABTS for 15 min at 25 °C. (<b>C</b>) SDS-PAGE analysis of the 20-fold concentrated fungal culture supernatants during the 26 days of cultivation. (<b>D</b>) Fungal cultures during 1000 mL (SG medium)- or 250 mL (S, C, and BS medium)-shake flask cultivation at 24 °C and 100 rpm after inoculation and on day 3, 5, 7, 9, 11, 12, 13, 15, 18, 20, and 26.</p>
Full article ">Figure A1
<p>Cell wet weight of fungi cultivated in different plant-based media. (<b>A</b>) Cells were harvested after cultivation in SG medium at 24 °C and 100 rpm in 100 mL shake flasks for 28 days. (<b>B</b>) Four selected fungi were cultivated again in SG medium at 24 °C and 100 rpm in 500 mL shake flasks for 22 days. (<b>C</b>) Cultivation of <span class="html-italic">C. sublineola</span> in four complex plant-based media in 1000 mL (SG) or 250 mL shake flasks (S, BS, and C) at 24 °C and 100 rpm for 26 days. Values given are means and average deviations for two replicates.</p>
Full article ">
15 pages, 6966 KiB  
Article
Xylogenesis Responses to a Mediterranean Climate in Holm Oak (Quercus ilex L.)
by Iqra Liyaqat, Angela Balzano, Francesco Niccoli, Jerzy Piotr Kabala, Maks Merela and Giovanna Battipaglia
Forests 2024, 15(8), 1386; https://doi.org/10.3390/f15081386 - 8 Aug 2024
Viewed by 493
Abstract
Quercus ilex L., an evergreen oak species typical of the western and central Mediterranean basin, is facing decline and dieback episodes due to the increase in the severity and frequency of heat waves and drought events. Studying xylogenesis (the wood formation process) is [...] Read more.
Quercus ilex L., an evergreen oak species typical of the western and central Mediterranean basin, is facing decline and dieback episodes due to the increase in the severity and frequency of heat waves and drought events. Studying xylogenesis (the wood formation process) is crucial for understanding how trees respond with their secondary growth to environmental conditions and stress events. This study aimed to characterize the wood formation dynamics of Quercus ilex and their relationship with the meteorological conditions in an area experiencing prolonged drought periods. Cambial activity and xylem cell production were monitored during the 2019 and 2020 growing seasons in a Q. ilex forest located at the Vesuvius National Park (southern Italy). The results highlighted the significant roles of temperature and solar radiation in stimulating xylogenesis. Indeed, the correlation tests revealed that temperature and solar radiation positively influenced growth and cell development, while precipitation had an inhibitory effect on secondary wall formation. The earlier cell maturation in 2020 compared to 2019 underscored the impact of global warming trends. Overall, the trees studied demonstrated good health, growth and adaptability to local environmental fluctuations. This research provides novel insights into the intra-annual growth dynamics of this key Mediterranean species and its adaptation strategies to climatic variability, which will be crucial for forest management in the context of climate change. Full article
(This article belongs to the Section Forest Ecology and Management)
Show Figures

Figure 1

Figure 1
<p>Study site located within Vesuvius National Park, Naples. Red triangle indicates the <span class="html-italic">Quercus ilex</span> stand.</p>
Full article ">Figure 2
<p>Weather conditions of the study site during the monitoring period of 2019 and 2020. In red, the maximum temperature; in black, the average temperature; in grey, the minimum temperature. The blue bars represent precipitation.</p>
Full article ">Figure 3
<p>Number of cambial cells and width of different developmental xylem zones in <span class="html-italic">Quercus ilex</span> trees in 2019 (<b>A</b>–<b>D</b>) and 2020 (<b>E</b>–<b>H</b>): cambial cells (CCs), enlarging post-cambial cells (PCs), cells developing secondary walls (SW) cells and mature (MT) cells with a lignified secondary wall. Mean values are shown for the days of the year (DOY) when the sampling was performed.</p>
Full article ">Figure 4
<p>Final tree ring width (TRW) containing fully mature cells formed for the years 2019 and 2020. Scale bar = 200 µm.</p>
Full article ">Figure 5
<p>Non-collapsed phloem (NCP) width in 2019 (<b>A</b>) and 2020 (<b>B</b>). Mean values are shown on the days of the year (DOY) when the sampling was performed.</p>
Full article ">Figure 6
<p>Correlations between xylogenesis and meteorological conditions. CC = cambial cells (number of cells), PC = post-cambial cells (width, μm), SW = secondary wall-forming cells (width, μm), MT = mature cells (width, μm). Meteorological variables: mean radiation (MJ/day), maximum temperature (°C), mean temperature (°C), minimum temperature (°C), and total precipitation (mm/day). Positive correlations are displayed in red, negative correlations in blue, non-significant ones in grey.</p>
Full article ">
23 pages, 3779 KiB  
Article
Vancomycin-Conjugated Polyethyleneimine-Stabilized Gold Nanoparticles Attenuate Germination and Show Potent Antifungal Activity against Aspergillus spp.
by Aishwarya Nikhil, Atul Kumar Tiwari, Ragini Tilak, Saroj Kumar, Prahlad Singh Bharti, Prem C. Pandey, Roger J. Narayan and Munesh Kumar Gupta
Appl. Sci. 2024, 14(16), 6926; https://doi.org/10.3390/app14166926 - 7 Aug 2024
Viewed by 342
Abstract
Antifungal drug resistance in filamentous fungi, particularly Aspergillus species, is increasing worldwide. Therefore, new antifungal drugs or combinations of drugs are urgently required to overcome this public health situation. In the present study, we examined the antifungal activity of vancomycin-functionalized AuNPs. These functionalized [...] Read more.
Antifungal drug resistance in filamentous fungi, particularly Aspergillus species, is increasing worldwide. Therefore, new antifungal drugs or combinations of drugs are urgently required to overcome this public health situation. In the present study, we examined the antifungal activity of vancomycin-functionalized AuNPs. These functionalized AuNPs were characterized, and their antifungal activity and associated killing mechanism were investigated using conventional methodologies against the conidia of A. fumigatus and A. flavus. The differential antifungal activity of vancomycin-functionalized Au-NPs against the conidia of Aspergillus species is dependent on structural differences in the conidial cell wall. The results demonstrated potent fungicidal activity against A. fumigatus, with a MIC value of 4.68 µg/mL, 93% germination inhibition, and 38.4% killing rate within 8 h of exposure. However, the activity against A. flavus was fungistatic; a MIC value of 18.7 µg/mL and 35% conidial germination inhibition, followed by 28.4% killing rate, were noted under similar conditions. Furthermore, endogenous reactive oxygen species (ROS) accumulation was 37.4 and 23.1% in conidial populations of A. fumigatus and A. flavus, respectively. Raman spectroscopy analysis confirmed the possible (but not confirmed) binding of functionalized AuNPs with the chitin and galactomannan components of the cell wall. A potential strategy that involves the exploration of antibacterial drugs using AuNPs as efficient drug carriers may also be appropriate for countering emerging drug resistance in filamentous fungi. Full article
(This article belongs to the Section Nanotechnology and Applied Nanosciences)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Physical characterization of PEI-AuNP@Van nanoparticles. (<b>a</b>) UV-Vis spectrum, (<b>b</b>) TEM micrograph, (<b>c</b>) high resolution image of nanoparticles, (<b>d</b>) mean nanoparticle size, (<b>e</b>) XRD diffractogram, and (<b>f</b>) dynamic light scattering data. (<b>B</b>) Zeta potential distribution histograms for non-functionalized gold nanoparticles (PEI-AuNPs) and vancomycin-functionalized gold nanoparticles (PEI-AuNP@Van).</p>
Full article ">Figure 1 Cont.
<p>(<b>A</b>) Physical characterization of PEI-AuNP@Van nanoparticles. (<b>a</b>) UV-Vis spectrum, (<b>b</b>) TEM micrograph, (<b>c</b>) high resolution image of nanoparticles, (<b>d</b>) mean nanoparticle size, (<b>e</b>) XRD diffractogram, and (<b>f</b>) dynamic light scattering data. (<b>B</b>) Zeta potential distribution histograms for non-functionalized gold nanoparticles (PEI-AuNPs) and vancomycin-functionalized gold nanoparticles (PEI-AuNP@Van).</p>
Full article ">Figure 2
<p>Fluorescence spectroscopic confirmation of vancomycin functionalization of PEI-stabilized gold nanoparticles. Adopted and modified under CC BY (2023) [<a href="#B33-applsci-14-06926" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>Antifungal evaluation plates (<b>a</b>) against <span class="html-italic">A. fumigatus</span> and (<b>b</b>) <span class="html-italic">A. flavus</span> (NC = PEI-AuNPs control; NP = PEI-AuNP@Van; PC = positive control). (<b>c</b>) MIC values of non-functionalized AuNPs (PEI-AuNPs), vancomycin-functionalized AuNPs (PEI-AuNP@Van), positive control (voriconazole), vancomycin, and negative control (DW) against <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>.</p>
Full article ">Figure 4
<p>Conidial germination inhibition assay.</p>
Full article ">Figure 5
<p>Confocal micrograph of PEI-AuNP@Van treated conidia of <span class="html-italic">Aspergillus</span>. (<b>a</b>) Bright field image of <span class="html-italic">A. fumigatus</span> treated with PEI-AuNP@Van and (<b>b</b>) stained with PI; (<b>c</b>) blue panel showing surface adsorbed functionalized nanoparticles, and (<b>d</b>) merged panel; (<b>e</b>) bright field image of <span class="html-italic">A. flavus</span> treated with PEI-AuNP@Van and (<b>f</b>) stained with PI; (<b>g</b>) blue panel showing surface adsorbed functionalized nanoparticles, and (<b>h</b>) merged panel.</p>
Full article ">Figure 6
<p>Histograms of dead conidia after 8 h of exposure to PEI-AuNP@Van. (<b>a</b>,<b>c</b>) represent the untreated controls of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>, respectively; (<b>b</b>,<b>d</b>) represent PEI-AuNP@Van-treated conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>.</p>
Full article ">Figure 7
<p>Endogenous ROS accumulation in PEI-AuNP@Van-treated conidia. (<b>a</b>,<b>c</b>) show the untreated control conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>, respectively; (<b>b</b>,<b>d</b>) represent the PEI-AuNP@Van-treated conidia.</p>
Full article ">Figure 8
<p>Raman spectrums of voriconazole and PEI-AuNP@Van-treated conidia of <span class="html-italic">A. flavus</span> and <span class="html-italic">A. fumigatus</span> along with an untreated control. (<b>ai</b>) Untreated conidia of <span class="html-italic">A. flavus</span>, (<b>aii</b>) treated with voriconazole, (<b>aiii</b>) treated with PEI-AuNP@Van, (<b>bi</b>) untreated conidia of <span class="html-italic">A. fumigatus</span>, (<b>bii</b>) treated with voriconazole, and (<b>biii</b>) treated with PEI-AuNP@Van.</p>
Full article ">Figure 9
<p>TEM micrograph of PEI-AuNP@Van-treated conidia of <span class="html-italic">A. fumigatus</span> and <span class="html-italic">A. flavus</span>: (<b>a</b>) represents the untreated conidia of <span class="html-italic">A. fumigatus</span>, and (<b>b</b>) treated with PEI-AuNP@Van; (<b>c</b>) represents the untreated conidia of <span class="html-italic">A. flavus</span>, and (<b>d</b>) treated with PEI-AuNP@Van.</p>
Full article ">
26 pages, 2855 KiB  
Article
Transcriptome Profiling and Weighted Gene Correlation Network Analysis Reveal Hub Genes and Pathways Involved in the Response to Polyethylene-Glycol-Induced Drought Stress of Two Citrus Rootstocks
by Emanuele Scialò, Angelo Sicilia, Alberto Continella, Alessandra Gentile and Angela Roberta Lo Piero
Biology 2024, 13(8), 595; https://doi.org/10.3390/biology13080595 - 7 Aug 2024
Viewed by 401
Abstract
Agriculture faces the dual challenge of increasing food production and safeguarding the environment. Climate change exacerbates this challenge, reducing crop yield and biomass due to drought stress, especially in semi-arid regions where Citrus plants are cultivated. Understanding the molecular mechanisms underlying drought tolerance [...] Read more.
Agriculture faces the dual challenge of increasing food production and safeguarding the environment. Climate change exacerbates this challenge, reducing crop yield and biomass due to drought stress, especially in semi-arid regions where Citrus plants are cultivated. Understanding the molecular mechanisms underlying drought tolerance in Citrus is crucial for developing adaptive strategies. Plants of two citrus rootstocks, Carrizo Citrange and Bitters (C22), were grown in aerated half-strength Hoagland’s nutrient solution. Post-acclimation, the plants were exposed to a solution containing 0% (control) or 15% PEG-8000 for 10 days. Leaf malonyl dialdehyde (MDA) and hydrogen peroxide (H2O2) content were measured to assess the reached oxidative stress level. Total RNA was extracted, sequenced, and de novo-assembled. Weighted Gene Correlation Network Analysis (WGCNA) was conducted to examine the relationship between gene expression patterns and the levels of MDA and H2O2 used as oxidative stress indicators. Plant visual inspection and MDA and H2O2 contents clearly indicate that Bitters is more tolerant than Carrizo towards PEG-induced drought stress. RNA-Seq analysis revealed a significantly higher number of differentially expressed genes (DEGs) in Carrizo (6092) than in Bitters (320), with most being associated with drought sensing, ROS scavenging, osmolyte biosynthesis, and cell wall metabolism. Moreover, the WGCNA identified transcription factors significantly correlated with MDA and H2O2 levels, thus providing insights into drought-coping strategies and offering candidate genes for enhancing citrus drought tolerance. Full article
(This article belongs to the Section Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Malondialdehyde (MDA) (<b>A</b>) and hydrogen peroxide (H<sub>2</sub>O<sub>2</sub>) (<b>B</b>) content in treated (PEG) and control (CK) plants of two citrus rootstock genotypes after 10 days of PEG treatment. Each point represents the mean value of three replicates. Different letters indicate significantly different values (ANOVA, <span class="html-italic">p</span> &lt; 0.05); CAR, Carrizo Citrange; C22, Bitters.</p>
Full article ">Figure 2
<p>Volcano plot showing the differentially expressed genes (DEGs) in the CAR_PEG vs. CAR_CK (<b>A</b>) and the C22_PEG vs. C22_CK (<b>B</b>) comparisons. Red dots represent the upregulated genes with statistical significance, the blue dots represent the downregulated genes with statistical significance, and the grey dots (ns) are DEGs with −log10padj &lt; 1.3, adopting a log<sub>2</sub> Fold Change threshold of 1 (2.0 fold change). The X-axis is the gene expression change, and the Y-axis is the <span class="html-italic">p</span>-value adjusted after normalisation.</p>
Full article ">Figure 3
<p>Gene Ontology (GO) enrichment analysis for the DEGs in the CAR_PEG vs. CAR_CK (<b>A</b>) and the C22_PEG vs. C22_CK (<b>B</b>) comparisons. The X-axis indicates the -log10(FDR), and the Y-axis indicates the GO terms within each category. Black dots indicate significantly enriched terms (FDR &lt; 0.05), while grey dots indicate non-significantly enriched terms (FDR ≥ 0.05). Symbols indicate the GO category (circles indicate the Molecular function category, triangles indicate the Biological process category, and squares indicate the Cellular component category). The dot size indicates the Gene Ratio.</p>
Full article ">Figure 4
<p>Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment analysis for the DEGs in the CAR_PEG vs. CAR_CK (<b>A</b>) and the C22_PEG vs. C22_CK (<b>B</b>) comparisons. The X-axis indicates the −log10(padj), and the Y-axis indicates the KEGG pathways. Black dots indicate significantly enriched terms (padj &lt; 0.05), while grey dots indicate non-significantly enriched terms (padj ≥ 0.05). The dot size indicates the Gene Ratio.</p>
Full article ">Figure 5
<p>Distribution of the ten most abundant families of transcription factors responsive to drought stress in the CAR_PEG vs. CAR_CK (<b>A</b>) and the C22_PEG vs. C22_CK (<b>B</b>) comparisons.</p>
Full article ">Figure 6
<p>Heatmap of the correlation between modules and MDA and H<sub>2</sub>O<sub>2</sub> levels.</p>
Full article ">Figure 7
<p>Diagrams illustrating the abundance and distribution of eigengenes associated with the grey60 module across various traits and comparisons.</p>
Full article ">Figure 8
<p>Diagrams illustrating the abundance and distribution of eigengenes associated with the turquoise and darkturquoise modules across various traits and comparisons.</p>
Full article ">
Back to TopTop