[go: up one dir, main page]

Next Issue
Volume 24, October-2
Previous Issue
Volume 24, September-2
 
 
ijms-logo

Journal Browser

Journal Browser

Int. J. Mol. Sci., Volume 24, Issue 19 (October-1 2023) – 628 articles

Cover Story (view full-size image): The incidence of Parkinson’s disease is the fastest growing among neurodegenerative diseases. Compelling evidence implicates inflammation, both in the central nervous system and in the periphery, in the initiation and progression of the disease, although it is unclear as of yet what triggers this response. Gut dysbiosis seems to be a likely candidate for initiating systemic inflammation. The present review discusses the various inflammatory mechanisms, signaling cascades, and mediators involved in the pathophysiology of Parkinson’s disease, as well as strategies that could be developed into disease-modifying treatments after issues related to biomarkers able to identify presymptomatic disease, indications of genetic testing, and timing for the maximum benefit of these therapies are settled. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
19 pages, 4648 KiB  
Article
Unveiling the Anti-Cancer Potential of Onoceranoid Triterpenes from Lansium domesticum Corr. cv. kokosan: An In Silico Study against Estrogen Receptor Alpha
by Ari Hardianto, Sarah Syifa Mardetia, Wanda Destiarani, Yudha Prawira Budiman, Dikdik Kurnia and Tri Mayanti
Int. J. Mol. Sci. 2023, 24(19), 15033; https://doi.org/10.3390/ijms241915033 - 9 Oct 2023
Cited by 1 | Viewed by 1452
Abstract
Breast cancer is a significant global concern, with tamoxifen, the standard treatment, raising long-term safety issues due to side effects. In this study, we evaluated the potential of five onoceranoid triterpenes from Lansium domesticum Corr. cv. kokosan against estrogen receptor alpha (ERα) using [...] Read more.
Breast cancer is a significant global concern, with tamoxifen, the standard treatment, raising long-term safety issues due to side effects. In this study, we evaluated the potential of five onoceranoid triterpenes from Lansium domesticum Corr. cv. kokosan against estrogen receptor alpha (ERα) using in silico techniques. Utilizing molecular docking, Lipinski’s rule of five, in silico ADMET, and molecular dynamics simulations, we assessed the potency of five onoceranoid triterpenes against ERα. Molecular docking indicated competitive binding energies for these triterpenes relative to the active form of tamoxifen (4OHT) and estradiol, an ERα native ligand. Three triterpenes met drug-likeness criteria with favorable ADMET profiles. Notably, 2 demonstrated superior binding affinity in molecular dynamics simulations, outperforming estradiol, closely followed by 3 and 4. Hierarchical clustering on principal components (HCPC) and the spatial distribution of contact surface area (CSA) analyses suggest that these triterpenes, especially 2, may act as antagonist ligands akin to 4OHT. These findings highlight the potential of onoceranoid triterpenes in treating ERα-related breast cancer. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Five onoceranoid triterpenes from <span class="html-italic">Lansium domesticum</span> Corr. cv. <span class="html-italic">kokosan</span>: lansiolic acid (<b>1</b>) [<a href="#B31-ijms-24-15033" class="html-bibr">31</a>], 8,14-secogammacera-7,14-dien-3,21-dione (<b>2</b>) [<a href="#B32-ijms-24-15033" class="html-bibr">32</a>], 8,14-secogammacera-7,14(27)-dien-3,21-dione (<b>3</b>) [<a href="#B32-ijms-24-15033" class="html-bibr">32</a>], kokosanolide B (<b>4</b>) [<a href="#B27-ijms-24-15033" class="html-bibr">27</a>], and 3-hydroxy-8,14-secogamasera-7,14-dien-21-one (<b>5</b>) [<a href="#B33-ijms-24-15033" class="html-bibr">33</a>].</p>
Full article ">Figure 2
<p>Docking poses of ligands in the hydrophobic binding site of ER<span class="html-italic">α</span>. The docking poses were generated through molecular docking. The ligand binding domain of ER<span class="html-italic">α</span> consists of twelve helices (H1–H12), beta sheets <b>1</b> and <b>2</b> (S1, S2), coil separating H11 and H12, and coil2-3, which are in different colors. The hydrophobicity of the molecular surface is represented by brown to white and blue color coding. The brown and color scales denote hydrophobicity and hydrophilicity indices, respectively, whereas the white color shows the balance between both properties.</p>
Full article ">Figure 3
<p>(<b>A</b>) The time evolutions of root-mean-square deviations (RMSDs) of ERα in apo and ligand-binding forms. The different colors represent the different clusters from the principal component analysis clustering (<a href="#app1-ijms-24-15033" class="html-app">Figure S4</a>). The color gradient is from red (cluster 1) to yellow (cluster 2), green (cluster 3), cyan (cluster 4), blue (cluster 5), and magenta (cluster 6). (<b>B</b>) RMSD analysis of each ligand during 500 ns MD simulations, calculated with the cpptraj program in Amber20. Data visualization was carried out using an R package ggplot2 on Jupyter Notebook 6.4.7. The different colors represent the different ligand systems.</p>
Full article ">Figure 4
<p>(<b>A</b>) RMSF plots of ERα in apo and ligand-bound forms. The turquoise line represents the RMSF profile of ERα in the apo form, the red line denotes <b>2</b>-bound ER<span class="html-italic">α</span>, the yellow line is <b>3</b>-bound ERα, the green line is <b>4</b>-bound ERα, the purple line is estradiol-bound ERα, and the magenta line is 4OHT-bound ERα. Colored bars under the RMSF lines correspond to ERα segments. (<b>B</b>) The 3D structures of apo and ligand-bound ERα in b-factor putty representations. The difference in each color represents the range of RMSF values in Angstrom Å. The color gradient is from purple (0 Å) to red (13 Å); the closer the color to the red, the higher the RMSF value.</p>
Full article ">Figure 5
<p>Hierarchical clustering on the factor map of apo and ligand-bound ER<span class="html-italic">α</span>. The clustering is based on the RMSF of apo and ligand-bound ER<span class="html-italic">α</span>. The different colors represent the different clusters from the principal component analysis clustering.</p>
Full article ">Figure 6
<p>Box plots of MMGBSA binding energy for ligand binding to ER<span class="html-italic">α</span>. The Box plots are from data points in <a href="#app1-ijms-24-15033" class="html-app">Figure S6</a>. The Box plots are complemented with a <span class="html-italic">p</span>-value from the Kruskal–Wallis test and pairwise statistical significances from the Games-–Howell test, displayed by asterisk symbols. Asterisk symbols *** and **** represent <span class="html-italic">p</span>-values in magnitudes of 10<sup>−4</sup> and less than 10<sup>−4</sup>, respectively. The box plots depict five summary statistics, including the minimum, first quartile, second quartile, third quartile, and maximum values, of MMGBSA binding energy values. A circled point indicates a potential outlier value. The various colors correspond to Box plots depicting MMGBSA binding energy values for the different ligands.</p>
Full article ">Figure 7
<p>Heatmap of MMGBSA binding energy decomposition for ER<span class="html-italic">α</span> residues interacting with ligands. The difference in each color represents the range of ΔG<sup>0</sup><span class="html-italic"><sub>MMGBSA</sub></span> values in kcal.mol<sup>−1</sup>. The color gradient is from blue (around −7 kcal.mol<sup>−1</sup>) to red (around 0 kcal.mol<sup>−1</sup>); the closer the color to the blue, the better the binding affinity.</p>
Full article ">Figure 8
<p>Contact surface area heatmap for each ligand interacting with ER<span class="html-italic">α</span> residues in the binding sites. The difference in each color represents the range of contact surface area in Å<sup>2</sup>. The color gradient is from blue (0 Å<sup>2</sup>) to red (55 Å<sup>2</sup>); the closer the color to red, the higher the area.</p>
Full article ">Figure 9
<p>Van der Waals interaction of <b>2</b> with residues Thr347 and Leu525 of ER<span class="html-italic">α.</span> The surface contacts are visualized using an atomic charge surface. Compound <b>2</b> is displayed as a green licorice structure. The different colors represent the average atomic charge. The red color represents the negative charge, the blue indicates the positive charge, and the greyish white denotes the uncharged atom.</p>
Full article ">
28 pages, 9231 KiB  
Article
Multi-Omics Profiling of Human Endothelial Cells from the Coronary Artery and Internal Thoracic Artery Reveals Molecular but Not Functional Heterogeneity
by Alexey Frolov, Arseniy Lobov, Marsel Kabilov, Bozhana Zainullina, Alexey Tupikin, Daria Shishkova, Victoria Markova, Anna Sinitskaya, Evgeny Grigoriev, Yulia Markova and Anton Kutikhin
Int. J. Mol. Sci. 2023, 24(19), 15032; https://doi.org/10.3390/ijms241915032 - 9 Oct 2023
Cited by 2 | Viewed by 1852
Abstract
Major adverse cardiovascular events occurring upon coronary artery bypass graft surgery are typically accompanied by endothelial dysfunction. Total arterial revascularisation, which employs both left and right internal thoracic arteries instead of the saphenous vein to create a bypass, is associated with better mid- [...] Read more.
Major adverse cardiovascular events occurring upon coronary artery bypass graft surgery are typically accompanied by endothelial dysfunction. Total arterial revascularisation, which employs both left and right internal thoracic arteries instead of the saphenous vein to create a bypass, is associated with better mid- and long-term outcomes. We suggested that molecular profiles of human coronary artery endothelial cells (HCAECs) and human internal mammary artery endothelial cells (HITAECs) are coherent in terms of transcriptomic and proteomic signatures, which were then investigated by RNA sequencing and ultra-high performance liquid chromatography-mass spectrometry, respectively. Both HCAECs and HITAECs overexpressed molecules responsible for the synthesis of extracellular matrix (ECM) components, basement membrane assembly, cell-ECM adhesion, organisation of intercellular junctions, and secretion of extracellular vesicles. HCAECs were characterised by higher enrichment with molecular signatures of basement membrane construction, collagen biosynthesis and folding, and formation of intercellular junctions, whilst HITAECs were notable for augmented pro-inflammatory signaling, intensive synthesis of proteins and nitrogen compounds, and enhanced ribosome biogenesis. Despite HCAECs and HITAECs showing a certain degree of molecular heterogeneity, no specific markers at the protein level have been identified. Coherence of differentially expressed molecular categories in HCAECs and HITAECs suggests synergistic interactions between these ECs in a bypass surgery scenario. Full article
Show Figures

Figure 1

Figure 1
<p>Volcano plot showing the distribution of proteins in the proteome of human coronary artery endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs). Gray points depict the proteins with log<sub>2</sub> fold change &lt; 1 and FDR-corrected <span class="html-italic">p</span> value &gt; 0.05. Green points depict the proteins with log<sub>2</sub> fold change &gt; 1 and FDR-corrected <span class="html-italic">p</span> value &gt; 0.05. Blue points depict the proteins with log<sub>2</sub> fold change &lt; 1 and FDR-corrected <span class="html-italic">p</span> value &lt; 0.05. Red points depict the proteins with log<sub>2</sub> fold change &gt; 1 and FDR-corrected <span class="html-italic">p</span> value &lt; 0.05 (DEPs).</p>
Full article ">Figure 2
<p>Volcano plot showing the distribution of transcripts in the transcriptome of human coronary artery endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs) under laminar flow. Gray points depict the genes with log<sub>2</sub> fold change &lt; 1 and FDR-corrected <span class="html-italic">p</span> value &gt; 0.05. Green points depict the genes with log<sub>2</sub> fold change &gt; 1 and FDR-corrected <span class="html-italic">p</span> value &gt; 0.05. Blue points depict the genes with log<sub>2</sub> fold change &lt; 1 and FDR-corrected <span class="html-italic">p</span> value &lt; 0.05. Red points depict the genes with log<sub>2</sub> fold change &gt; 1 and FDR-corrected <span class="html-italic">p</span> value &lt; 0.05 (DEGs).</p>
Full article ">Figure 3
<p>Volcano plot showing the distribution of transcripts in the transcriptome of human coronary artery endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs) at static cell culture conditions. Gray points depict the genes with log<sub>2</sub> fold change &lt; 1 and FDR-corrected <span class="html-italic">p</span> value &gt; 0.05. Green points depict the genes with log<sub>2</sub> fold change &gt; 1 and FDR-corrected <span class="html-italic">p</span> value &gt; 0.05. Blue points depict the genes with log<sub>2</sub> fold change &lt; 1 and FDR-corrected <span class="html-italic">p</span> value &lt; 0.05. Red points depict the genes with log<sub>2</sub> fold change &gt; 1 and FDR-corrected <span class="html-italic">p</span> value &lt; 0.05 (DEGs).</p>
Full article ">Figure 4
<p>Fluorescent Western blotting for cell adhesion molecules (VCAM1 and ICAM1), transcription factors of endothelial-to-mesenchymal transition (Snail and Slug, TWIST1, and ZEB1), mechanosensitive transcription factors (KLF2, KLF4, and NRF2), endothelial nitric oxide synthase eNOS, and loading control (GAPDH and CD31) in human coronary artery endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs) cultured at static conditions. (<b>A</b>) VCAM1 (pro-inflammatory cell adhesion molecule, green)/GAPDH (loading control, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>); (<b>B</b>) ICAM1 (pro-inflammatory cell adhesion molecule, green)/CD31 (loading control, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>); (<b>C</b>) Snail and Slug (endothelial-to-mesenchymal transition transcription factor, green)/CD31 (loading control, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>); (<b>D</b>) TWIST1 (endothelial-to-mesenchymal transition transcription factor, green)/ZEB1 (another endothelial-to-mesenchymal transition transcription factor, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>); (<b>E</b>) TWIST1 (endothelial-to-mesenchymal transition transcription factor, green)/KLF2 (atheroprotective mechanosensitive transcription factor, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>); (<b>F</b>) KLF4 (atheroprotective mechanosensitive transcription factor, green)/eNOS (endothelial nitric oxide synthase, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>); (<b>G</b>) NRF2 (atheroprotective mechanosensitive transcription factor, green)/GAPDH (loading control, red), fluorescent Western blot (<b>top</b>), and total protein staining confirming an equal protein loading (<b>bottom</b>). Each band within the groups represent a protein lysate from one experiment (<span class="html-italic">n</span> = 3 experiments in total). Total protein normalisation was conducted by Fast Green FCF staining of the membranes after the fluorescent imaging to ensure the equal protein loading at all blots (in addition to loading controls such as GAPDH or CD31). Fluorescent ladder (L) and molecular weight signatures (kDa) are provided to the left of the HCAECs and HITAECs protein bands. Ratios of 1:100 and 1:200 are dilutions of the antibody against TWIST1, highlighted to show low expression of this protein in the quiescent ECs.</p>
Full article ">Figure 5
<p>Bioinformatic analysis of protein–protein interactions between primary human coronary endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs). (<b>A</b>) Overview of protein–protein interactions between HCAECs (blue colour) and HITAECs (red colour); (<b>B</b>) annotation of the interacting proteins related to the specific molecular terms (yellow colour) in comparison with unannotated overview (the main cluster of interacting proteins is demarcated by red contour); and (<b>C</b>) annotations of clustered interacting proteins in selected molecular terms: integrin-mediated cell adhesion (demarcated by blue contour), basement membrane (demarcated by violet contour), and elastic fiber formation (demarcated by green contour).</p>
Full article ">Figure 6
<p>Bioinformatic analysis of gene–gene interactions between primary human coronary endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs) cultured under laminar flow. (<b>A</b>) Overview of gene–gene interactions between HCAECs (blue colour) and HITAECs (red colour) cultured under laminar flow; (<b>B</b>) annotation of the interacting genes related to the specific molecular terms (yellow colour) in comparison with unannotated overview (the main cluster of interacting genes is demarcated by red contour); and (<b>C</b>) annotations of clustered interacting genes in selected molecular terms: sprouting angiogenesis (demarcated by blue contour) and elastic fiber formation (demarcated by green contour).</p>
Full article ">Figure 7
<p>Bioinformatic analysis of gene–gene interactions between primary human coronary endothelial cells (HCAECs) and human internal thoracic artery endothelial cells (HITAECs) cultured at static conditions. (<b>A</b>) Overview of gene–gene interactions between HCAECs (blue colour) and HITAECs (red colour) cultured under laminar flow; (<b>B</b>) annotation of the interacting genes related to the specific molecular terms (yellow colour) in comparison with unannotated overview; and (<b>C</b>) annotations of clustered interacting genes in selected molecular terms: angiogenesis (demarcated by blue contour) and stimulation of guanylate cyclase by nitric oxide (demarcated by green contour).</p>
Full article ">
25 pages, 12701 KiB  
Article
Comparative Chloroplast Genomics of 21 Species in Zingiberales with Implications for Their Phylogenetic Relationships and Molecular Dating
by Dong-Mei Li, Hai-Lin Liu, Yan-Gu Pan, Bo Yu, Dan Huang and Gen-Fa Zhu
Int. J. Mol. Sci. 2023, 24(19), 15031; https://doi.org/10.3390/ijms241915031 - 9 Oct 2023
Cited by 2 | Viewed by 1370
Abstract
Zingiberales includes eight families and more than 2600 species, with many species having important economic and ecological value. However, the backbone phylogenetic relationships of Zingiberales still remain controversial, as demonstrated in previous studies, and molecular dating based on chloroplast genomes has not been [...] Read more.
Zingiberales includes eight families and more than 2600 species, with many species having important economic and ecological value. However, the backbone phylogenetic relationships of Zingiberales still remain controversial, as demonstrated in previous studies, and molecular dating based on chloroplast genomes has not been comprehensively studied for the whole order. Herein, 22 complete chloroplast genomes from 21 species in Zingiberales were sequenced, assembled, and analyzed. These 22 genomes displayed typical quadripartite structures, which ranged from 161,303 bp to 163,979 bp in length and contained 111–112 different genes. The genome structures, gene contents, simple sequence repeats, long repeats, and codon usage were highly conserved, with slight differences among these genomes. Further comparative analysis of the 111 complete chloroplast genomes of Zingiberales, including 22 newly sequenced ones and the remaining ones from the national center for biotechnology information (NCBI) database, identified three highly divergent regions comprising ccsA, psaC, and psaC-ndhE. Maximum likelihood and Bayesian inference phylogenetic analyses based on chloroplast genome sequences found identical topological structures and identified a strongly supported backbone of phylogenetic relationships. Cannaceae was sister to Marantaceae, forming a clade that was collectively sister to the clade of (Costaceae, Zingiberaceae) with strong support (bootstrap (BS) = 100%, and posterior probability (PP) = 0.99–1.0); Heliconiaceae was sister to the clade of (Lowiaceae, Strelitziaceae), then collectively sister to Musaceae with strong support (BS = 94–100%, and PP = 0.93–1.0); the clade of ((Cannaceae, Marantaceae), (Costaceae, Zingiberaceae)) was sister to the clade of (Musaceae, (Heliconiaceae, (Lowiaceae, Strelitziaceae))) with robust support (BS = 100%, and PP = 1.0). The results of divergence time estimation of Zingiberales indicated that the crown node of Zingiberales occurred approximately 85.0 Mya (95% highest posterior density (HPD) = 81.6–89.3 million years ago (Mya)), with major family-level lineages becoming from 46.8 to 80.5 Mya. These findings proved that chloroplast genomes could contribute to the study of phylogenetic relationships and molecular dating in Zingiberales, as well as provide potential molecular markers for further taxonomic and phylogenetic studies of Zingiberales. Full article
(This article belongs to the Collection Feature Papers in Molecular Genetics and Genomics)
Show Figures

Figure 1

Figure 1
<p>Complete chloroplast genome map of <span class="html-italic">H. bijiangense</span> (GenBank OP805589; the outermost three rings) and CGView comparison of 22 complete chloroplast genomes in Zingiberales (the inter rings with different colors). Genes belonging to different functional groups are shown in different colors in the outermost fIRst ring. Genes shown on the outside of the outermost fIRst ring are transcribed counter-clockwise and on the inside clockwise. Gray arrowheads indicate the direction of the genes. The tRNA genes are indicated by a one-letter code of amino acids with anticodons. The outermost second ring with darker gray corresponds to GC content, whereas the outermost third ring with lighter gray corresponds to AT content of <span class="html-italic">H. bijiangense</span> chloroplast genome. The innermost fIRst black ring indicates the chloroplast genome size of <span class="html-italic">H. bijiangense</span>. The innermost second and third rings indicate GC content and GC skew deviations in the chloroplast genome of <span class="html-italic">H. bijiangense</span>, respectively: GC skew + indicates G &gt; C, and GC skew – indicates G &lt; C. From the innermost fourth color ring to the outwards 25th ring in turn: <span class="html-italic">H. bijiangense</span> OP805589, <span class="html-italic">H. brevicaule</span> OP805581, <span class="html-italic">H. chrysoleucum</span> OP805577, <span class="html-italic">H. coccineum</span> ‘Red’ OP805574, <span class="html-italic">H. flavescens</span> Guangxi OP805591, <span class="html-italic">H. flavescens</span> Yunnan OP805575, <span class="html-italic">H. flavum</span> OP805588, <span class="html-italic">H. kwangsiense</span> OP805586, <span class="html-italic">H. menghaiense</span> OP805587, <span class="html-italic">H. puerense</span> OP805578, <span class="html-italic">H.</span> sp.1 LDM232 OP805583, <span class="html-italic">H.</span> sp.2 LDM222 OP805582, <span class="html-italic">H. tienlinense</span> OP805590, <span class="html-italic">H. villosum</span> var. <span class="html-italic">albifihamentum</span> OP805580, <span class="html-italic">H. villosum</span> Guangxi OP805584, <span class="html-italic">H. villosum</span> var. <span class="html-italic">tenuiflorum</span> OP805576, <span class="html-italic">H. viridibracteatum</span> OP805579, <span class="html-italic">H. yunnanense</span> OP805585, <span class="html-italic">C. makoyana</span> OP805573, <span class="html-italic">C. aurantiflora</span> OP805593, <span class="html-italic">K. parviflora</span> OP805592, <span class="html-italic">S. sanguinea</span> OP805594; chloroplast genome similar and highly divergent locations are represented by continuous and interrupted track lines, respectively. LSC, large single-copy region; SSC, small single-copy region; and IR, inverted repeat.</p>
Full article ">Figure 2
<p>Analysis of long repeats in the 22 newly sequenced complete chloroplast genomes of Zingiberales. (<b>A</b>) Total number of four long repeat types. (<b>B</b>) Length distribution of long repeats in each sequenced chloroplast genome.</p>
Full article ">Figure 3
<p>Types and distribution of SSRs in 22 newly sequenced complete chloroplast genomes of Zingiberales. (<b>A</b>) Number of different SSR types. (<b>B</b>) Number of identified SSR motifs in different repeat class types. SSR, simple sequence repeat.</p>
Full article ">Figure 3 Cont.
<p>Types and distribution of SSRs in 22 newly sequenced complete chloroplast genomes of Zingiberales. (<b>A</b>) Number of different SSR types. (<b>B</b>) Number of identified SSR motifs in different repeat class types. SSR, simple sequence repeat.</p>
Full article ">Figure 4
<p>Heat map for the relative synonymous codon usage values of the 22 newly sequenced complete chloroplast genomes of Zingiberales.</p>
Full article ">Figure 5
<p>Visualization of the alignment of the 22 newly sequenced complete chloroplast genomes of Zingiberales. The chloroplast genome of <span class="html-italic">H. bijiangense</span> is used as the reference. The y-axis represents the percent identity ranging from 50% to 100%. The x-axis depicts sequence coordinates within the chloroplast genome. Purple bars represent exons, sky-blue bars represent untranslated regions (UTRs), red bars represent non-coding sequences (CNS), grey bars represent mRNA and white regions represent sequence differences among 22 analyzed chloroplast genomes.</p>
Full article ">Figure 6
<p>Comparison of the IR/SC borders among 22 newly sequenced complete chloroplast genomes of Zingiberales. LSC, large single-copy region; SSC, small single-copy region; IR, inverted repeat region.</p>
Full article ">Figure 7
<p>Comparison of nucleotide diversity values across the 111 complete chloroplast genomes of Zingiberales. (<b>A</b>) Protein-coding regions. (<b>B</b>) Non-coding regions.</p>
Full article ">Figure 8
<p>Two phylogenetic trees reconstructed from the 115 chloroplast genome sequences of Zingiberales and 4 outgroups using maximum likelihood (ML) and Bayesian inference (BI), respectively. (<b>A</b>) Phylogenetic tree reconstruction using ML method of PhyML v. 3.0. Numbers next to the branches are ML bootstrap support values. (<b>B</b>) Phylogenetic tree reconstruction using the BI method of MrBayes v. 3.2.6. Numbers next to the branches are BI probability support values. The newly sequenced 22 chloroplast genomes in this study are in bold.</p>
Full article ">Figure 8 Cont.
<p>Two phylogenetic trees reconstructed from the 115 chloroplast genome sequences of Zingiberales and 4 outgroups using maximum likelihood (ML) and Bayesian inference (BI), respectively. (<b>A</b>) Phylogenetic tree reconstruction using ML method of PhyML v. 3.0. Numbers next to the branches are ML bootstrap support values. (<b>B</b>) Phylogenetic tree reconstruction using the BI method of MrBayes v. 3.2.6. Numbers next to the branches are BI probability support values. The newly sequenced 22 chloroplast genomes in this study are in bold.</p>
Full article ">Figure 9
<p>Divergence time estimation of Zingiberales based on the 115 chloroplast genome sequences. The fossil and calibration taxa are indicated with red points on the corresponding nodes. The mean divergence time of the nodes is shown at the nodes with blue. The blue bars correspond to 95% HPD of estimated divergence time, with minimum and maximum values, respectively.</p>
Full article ">
13 pages, 2561 KiB  
Article
The Relationship between Clock Genes, Sirtuin 1, and Mitochondrial Activity in Head and Neck Squamous Cell Cancer: Effects of Melatonin Treatment
by César Rodríguez-Santana, Alba López-Rodríguez, Laura Martinez-Ruiz, Javier Florido, Olga Cela, Nazzareno Capitanio, Yolanda Ramírez-Casas, Darío Acuña-Castroviejo and Germaine Escames
Int. J. Mol. Sci. 2023, 24(19), 15030; https://doi.org/10.3390/ijms241915030 - 9 Oct 2023
Cited by 2 | Viewed by 1547
Abstract
The circadian clock is a regulatory system, with a periodicity of approximately 24 h, which generates rhythmic changes in many physiological processes, including mitochondrial activity. Increasing evidence links chronodisruption with aberrant functionality in clock gene expression, resulting in multiple diseases such as cancer. [...] Read more.
The circadian clock is a regulatory system, with a periodicity of approximately 24 h, which generates rhythmic changes in many physiological processes, including mitochondrial activity. Increasing evidence links chronodisruption with aberrant functionality in clock gene expression, resulting in multiple diseases such as cancer. Melatonin, whose production and secretion oscillates according to the light–dark cycle, is the principal regulator of clock gene expression. In addition, the oncostatic effects of melatonin correlate with an increase in mitochondrial activity. However, the direct links between circadian clock gene expression, mitochondrial activity, and the antiproliferative effects of melatonin in cancers, including head and neck squamous cell carcinoma (HNSCC), remain largely unknown. In this study, we analyzed the effects of melatonin on HNSCC cell lines (Cal-27 and SCC9), which were treated with 500 and 1000 µM melatonin. We found that the antiproliferative effect of melatonin is not mediated by the Bmal1 clock gene. Additionally, high doses of melatonin were observed to result in resynchronization of oscillatory circadian rhythm genes (Per2 and Sirt1). Surprisingly, the resynchronizing effect of melatonin on Per2 and Sirt1 did not produce alterations in the oscillation of mitochondrial respiratory activity. These results increase our understanding of the possible antiproliferative mechanisms in melatonin in the treatment of head and neck squamous cell carcinoma and suggest that its antiproliferative effects are independent of clock genes but are directly related to mitochondrial activity. Full article
(This article belongs to the Special Issue Molecular Pathology and Novel Therapies for Head and Neck Cancer)
Show Figures

Figure 1

Figure 1
<p>Relative expression of clock genes in HNSCC cell lines Cal-27 (<b>A</b>,<b>B</b>) and SCC9 (<b>C</b>,<b>D</b>): (<b>A</b>,<b>C</b>) <span class="html-italic">Bmal1</span> and (<b>B</b>,<b>D</b>) <span class="html-italic">Per2</span> after serum shock, control, and aMT (500 and 1000 µM) treatments; <span class="html-italic">n</span> = 3–6 independent experiments. The best cosinor fit is shown as a continuous line at a time of 48 h.</p>
Full article ">Figure 2
<p>Relative expression of <span class="html-italic">Bmal1</span> in Cal-27 cells with 10 nM aMT (<b>A</b>) and 100 µM aMT (<b>B</b>) treatments; <span class="html-italic">n</span> = 3–4 independent experiments. The best cosinor fit is shown as a continuous line at a time of 48 h.</p>
Full article ">Figure 3
<p>Relative expression of <span class="html-italic">Sirt1</span> in HNSCC cell lines Cal-27 (<b>A</b>) and SCC9 (<b>B</b>) after serum shock, control, and aMT (500 and 1000 µM) treatments; <span class="html-italic">n</span> = 3–6 independent experiments. The best cosinor fit is shown as a continuous line at a time of 48 h.</p>
Full article ">Figure 4
<p>Measurement of mitochondrial respiratory activity in intact Cal-27 cells after serum shock, control, and aMT (500 and 1000 µM) treatments. OCR refers to the resting condition (i.e., endogenous substrate-sustained respiration); <span class="html-italic">n</span> = 3–6 independent experiments. The best cosinor fit is shown as a continuous line.</p>
Full article ">Figure 5
<p>Measurement of mitochondrial respiratory activity in intact Cal-27 cells after serum shock, in control and aMT 1000 µM. OCR refers to resting the condition (i.e., endogenous substrate-sustained respiration); <span class="html-italic">n</span> = 3 independent experiments. The best cosinor fit is shown as a continuous line.</p>
Full article ">Figure 6
<p>Effect of melatonin on <span class="html-italic">Bmal1</span>-silenced HNSCC cell lines Cal-27 (<b>left</b>) and SCC-9 (<b>right</b>). (<b>A</b>,<b>B</b>) Histogram showing the <span class="html-italic">Bmal1</span> transcript level attained by q-RT-PCR in HNSCC cells transfected with siRNA-<span class="html-italic">Bmal1</span>. (<b>C</b>,<b>D</b>) The effect of melatonin on cell proliferation assessed by CyQuant assay. Data are presented as the mean ± SEM of <span class="html-italic">n</span> = 3 independent experiments; * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001 vs. mock control; # <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.001 vs. siRNA-<span class="html-italic">Bmal1</span> control.</p>
Full article ">
19 pages, 8161 KiB  
Article
The MADF-BESS Protein CP60 Is Recruited to Insulators via CP190 and Has Redundant Functions in Drosophila
by Larisa Melnikova, Varvara Molodina, Valentin Babosha, Margarita Kostyuchenko, Pavel Georgiev and Anton Golovnin
Int. J. Mol. Sci. 2023, 24(19), 15029; https://doi.org/10.3390/ijms241915029 - 9 Oct 2023
Cited by 2 | Viewed by 1186
Abstract
Drosophila CP190 and CP60 are transcription factors that are associated with centrosomes during mitosis. CP190 is an essential transcription factor and preferentially binds to housekeeping gene promoters and insulators through interactions with architectural proteins, including Su(Hw) and dCTCF. CP60 belongs to a family [...] Read more.
Drosophila CP190 and CP60 are transcription factors that are associated with centrosomes during mitosis. CP190 is an essential transcription factor and preferentially binds to housekeeping gene promoters and insulators through interactions with architectural proteins, including Su(Hw) and dCTCF. CP60 belongs to a family of transcription factors that contain the N-terminal MADF domain and the C-terminal BESS domain, which is characterized by the ability to homodimerize. In this study, we show that the conserved CP60 region adjacent to MADF is responsible for interacting with CP190. In contrast to the well-characterized MADF-BESS transcriptional activator Adf-1, CP60 is recruited to most chromatin sites through its interaction with CP190, and the MADF domain is likely involved in protein–protein interactions but not in DNA binding. The deletion of the Map60 gene showed that CP60 is not an essential protein, despite the strong and ubiquitous expression of CP60 at all stages of Drosophila development. Although CP60 is a stable component of the Su(Hw) insulator complex, the inactivation of CP60 does not affect the enhancer-blocking activity of the Su(Hw)-dependent gypsy insulator. Overall, our results indicate that CP60 has an important but redundant function in transcriptional regulation as a partner of the CP190 protein. Full article
(This article belongs to the Special Issue Molecular Genetics of Drosophila Development)
Show Figures

Figure 1

Figure 1
<p>CP60 is a component of the Su(Hw) insulator complex: (<b>A</b>) Schematic presentation of the Su(Hw) insulator complex. The previously described interactions [<a href="#B58-ijms-24-15029" class="html-bibr">58</a>] between the components of the complex are shown. The CP190 domains are shown as yellow ovals, with four zinc fingers shown as yellow boxes; the Mod(mdg4)-67.2 domains are shown as green ovals; Su(Hw) domains are shown as lilac boxes. Domain abbreviations: CID—CP190 interacting domain; Ac—C-terminal acidic domain; Zn—zinc-finger domain; LZ—leucine zipper; BTB—BTB/POZ domain; Q—glutamine-rich region; DD—dimerization domain; FLYWCH—FLYWCH-type zinc finger; SID—Su(Hw) interacting domain; D—asparagine-rich domain; M—the microtubule- and centrosome-associated domain; E—glutamine-rich C-terminal domain. Bold capital letters indicate the Su(Hw) binding site. (<b>B</b>) Nuclear extract from <span class="html-italic">Drosophila</span> S2 cells was immunoprecipitated with antibodies against CP60, and the immunoprecipitates (IP) were analyzed via Western blotting for the presence of Su(Hw), Mod(mdg4)-67.2, and CP190 proteins. Input is the input fraction (1% of the lysate used for immunoprecipitation); output is the supernatant after immunoprecipitation; IP is the immunoprecipitate; PI is immunoprecipitation with nonspecific IgG. (<b>C</b>) Nuclear extract from <span class="html-italic">Drosophila</span> S2 cells was immunoprecipitated with antibodies against Su(Hw), Mod(mdg4)-67.2, or CP190, and the immunoprecipitates (IP) were analyzed via Western blotting for the presence of CP60 protein. The uncropped images for (<b>B</b>,<b>C</b>) are shown in <a href="#app1-ijms-24-15029" class="html-app">Supplementary Materials Figure S2</a>. (<b>D</b>) Identification of direct interactions between CP60 and other components of the Su(Hw) insulator complex using the yeast two-hybrid assay. The results are summarized in the table with + and—signs referring to a strong interaction or no interaction, respectively. Interactions with pGBT9 or pGAD424 vectors were used as the negative control. Interactions between CP190 and Mod(mdg4)-67.2 with Su(Hw) were used as a positive control.</p>
Full article ">Figure 2
<p>CP190 is responsible for CP60 recruitment to the Su(Hw) chromatin sites: (<b>A</b>) Western blot analysis of RNAi efficiency; wt—S2 cells without treatment; proteins indicated below the line (RNAi) were knocked out by RNAi; antibodies for staining are listed on the left of the panel. Anti-tubulin antibodies (αTub) were used as a loading control. The uncropped images are shown in <a href="#app1-ijms-24-15029" class="html-app">Supplementary Materials Figure S3</a>. (<b>B</b>–<b>E</b>) ChIP-qPCR analysis of binding (<b>B</b>) Su(Hw), (<b>C</b>) CP190, (<b>D</b>) CP60, and (<b>E</b>) Mod(mdg4)-67.2 proteins to the selected Su(Hw) sites in wild-type (wt) S2 cells and after RNAi inactivation of each protein. The <span class="html-italic">ras64B</span> coding region (Ras) was used as a control that does not contain Su(Hw) binding sites. IgG—immunoprecipitation with nonspecific IgG. The percentage recovery of immunoprecipitated DNA (Y axis) was calculated relative to the amount of input DNA. Error bars indicate SDs of quadruplicate PCR measurements from two independent biological samples of chromatin. Asterisks indicate significance levels: * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 (Student’s <span class="html-italic">t</span>-test). Dots on the bar plots indicate the values of individual experiments.</p>
Full article ">Figure 3
<p>Analysis of the in vitro association between Su(Hw) or CP60 and Su(Hw) motifs: (<b>A</b>) Schematic representation of the tested Su(Hw) binding sites. The red squares show the localization of the Su(Hw) motifs. (<b>B</b>) The Adf-1 binding site from <span class="html-italic">bxd</span> PRE and purified Adf-1 protein were used in EMSA as the positive control. DNA fragments without protein are marked with a minus (−) sign. The triangle represents a threefold increase in Adf-1 concentration from 0.05 µg to 0.15 µg. (<b>C</b>) EMSA of the binding of Su(Hw) (0.05 µg) and CP60 (0.05 µg) proteins to DNA fragments 50 A, 62 D, 66 E, and 87 E. The Adf-1 binding site was used as the negative control for the Su(Hw) or CP60 proteins. (<b>D</b>) EMSA of the binding of CP60 and Su(Hw) proteins to the <span class="html-italic">gypsy</span> insulator (<span class="html-italic">gypsy</span>) and two copies of the minimal 125 bp 1A2 region (1A2 125 × 2), and 250 bp (1A2 250) and 450 bp (1A2 450) regions of 1A2 insulator site. The amount of Su(Hw) protein was increased threefold from 0.05 µg to 0.15 µg. Other designations are as in (<b>B</b>). (<b>E</b>) Cooperation of Su(Hw), CP190, and CP60 in the binding to the <span class="html-italic">gypsy</span> insulator. The combination of proteins (0.05 µg) is indicated above in the panel. The absence of the corresponding protein in the EMSA probe is marked as a minus (−) sign. (<b>F</b>) EMSA supershift by CP60 antibodies and different combinations of Su(Hw) insulator proteins (0.05 µg). Ab—CP60 antibodies were added to the probe. Other designations as in (<b>E</b>).</p>
Full article ">Figure 4
<p>Mapping domains of CP60 that are responsible for homodimerization and interaction with CP190 in the yeast two-hybrid assay. The scale at the top of the figure is in amino acids. A schematic representation of the CP60 protein is shown with deletion margins on the left. The MADF domain is shown as a green rectangle. Conserved regions 1–3 are represented by red rectangles. The fourth C-terminal conserved region is shown as a blue rectangle. The interactions are summarized in columns on the right-hand side with + and − referring to the presence and absence of interactions, respectively.</p>
Full article ">Figure 5
<p>Multiple alignment analysis of the CP60 C-terminal conserved region (440–441 aa) sequence with different known BESS domains. Alignment regions are shown as colored boxes or outlines that enclose one or more residue symbols. The residue coloring corresponds to the Clustal X color code (<a href="#app1-ijms-24-15029" class="html-app">Supplementary Table S1</a>).</p>
Full article ">Figure 6
<p>Functional analysis of the <span class="html-italic">Map60</span> gene: (<b>A</b>) CRISPR/Cas9 deletion of the <span class="html-italic">Map60</span> gene. The CP60 coding regions are shown as orange boxes. White rectangles represent the 5′ and 3′UTR. CRISPR targets are shown as vertical bars. Primer sequences used in the CRISPR/Cas9 genome editing are given in <a href="#app1-ijms-24-15029" class="html-app">Supplementary Table S2</a>. The <span class="html-italic">Red</span> reporter (magenta box), controlled by the <span class="html-italic">3xP3</span> promoter (black line), was used for the selection of the <span class="html-italic">Map60</span> deletion. The <span class="html-italic">attP</span> and <span class="html-italic">lox</span> sites were used for genome manipulation and are shown as a green box and vertical yellow arrows, respectively. SV40 terminator is shown as a blue box. The two black arrows represent primers 1 and 2, which were used to test the obtained mutations using PCR. (<b>B</b>) PCR analysis of genomic DNA from <span class="html-italic">Map60<sup>Δ1</sup></span> lines after <span class="html-italic">CRE/loxP</span> excision of marker <span class="html-italic">dsRed</span> cassette (<span class="html-italic">ΔdsRed</span>). The molecular weight in bp is shown on the right. (<b>C</b>) Western blot analysis (8% SDS PAGE) of protein extracts prepared as described previously [<a href="#B72-ijms-24-15029" class="html-bibr">72</a>] from adult three-day-old males of the wild-type (wt) line and three lines homozygous for <span class="html-italic">Map60<sup>Δ1</sup></span>. The membrane was sequentially stained with tested polyclonal rat antibodies against CP60 (αCP60) and antibodies against tubulin (αTub) as loading control. The molecular weight in kDa is shown on the left. The uncropped images are shown in <a href="#app1-ijms-24-15029" class="html-app">Supplementary Materials Figure S7</a>. (<b>D</b>) Polytene chromosomes from the salivary glands of third-instar <span class="html-italic">y<sup>2</sup>w<sup>1118</sup>ct<sup>6</sup></span> (wt) and <span class="html-italic">y<sup>2</sup>w<sup>1118</sup>ct<sup>6</sup></span>; <span class="html-italic">Map60<sup>Δ1</sup>-1/Map60<sup>Δ1</sup></span>-1 (Map60<sup>Δ1</sup>-1) larvae costained with rat anti-CP60 antibodies (green), rabbit anti-CP190 antibodies (red), and DAPI (blue) Scale bars, 10 μm. (<b>E</b>) ChIP-qPCR analysis of Su(Hw), CP190, and CP60 binding to Su(Hw) sites (description in <a href="#ijms-24-15029-f002" class="html-fig">Figure 2</a>) in wt and Map60<sup>Δ1</sup>-1 (Δ1) lines. The <span class="html-italic">ras64B</span> coding region (Ras) that did not contain Su(Hw) binding sites was used as a control. IgG—immunoprecipitation with nonspecific IgG. The percentage recovery of immunoprecipitated DNA (Y axis) was calculated relative to the amount of input DNA. Error bars indicate SDs of quadruplicate PCR measurements from two independent biological samples of chromatin. Significance levels: <span class="html-italic">p</span> &lt; 0.05 (Student’s <span class="html-italic">t</span>-test). Dots on the bar plots indicate the values of individual experiments.</p>
Full article ">Figure 7
<p>Testing the genome-wide distribution of CP60 and its colocalization with the components of the Su(Hw) complex. Polytene chromosomes from the salivary glands of third-instar (<b>A</b>) <span class="html-italic">y<sup>2</sup>w<sup>1118</sup>ct<sup>6</sup></span> (wt) and (<b>B</b>) <span class="html-italic">y<sup>2</sup>w<sup>1118</sup>ct<sup>6</sup></span>; <span class="html-italic">Cp190<sup>2</sup>/Cp190<sup>3</sup></span> (<span class="html-italic">Cp190<sup>2</sup>/Cp190<sup>3</sup></span>) larvae costained with rat anti-CP60 antibodies (green), rabbit anti-CP190 or anti-Su(Hw) antibodies (red), and DAPI (blue). Scale bars, 10 μm.</p>
Full article ">Figure 8
<p>CP190 is responsible for CP60 recruitment to dCTCF/CP190 and BEAF/CP190 sites. ChIP-qPCR analysis of binding of dCTCF, CP190, and CP60 to the (<b>A</b>) dCTCF (Fab3, Fab4, Fab8, and MCP)- and Pita (Fab7)-dependent sites of the Bitorax complex. (<b>B</b>) BEAF-dependent sites from promoter regions of the <span class="html-italic">aurora</span>, <span class="html-italic">cg3281</span>, and <span class="html-italic">janA</span> genes. (<b>C</b>) Promoter and regulatory regions from the Bitorax complex are not bound by CP190 but are enriched in isoforms of the Mod(mdg4) protein. The <span class="html-italic">ras64B</span> coding region (Ras) was used as a control devoid of dCTCF/CP190- and BEAF/CP190 binding sites. IgG—immunoprecipitation with nonspecific IgG. The percentage recovery of immunoprecipitated DNA (Y axis) was calculated relative to the amount of input DNA. Error bars indicate SDs of quadruplicate PCR measurements from two independent biological samples of chromatin. Significance levels: <span class="html-italic">p</span> &lt; 0.01 (Student’s <span class="html-italic">t</span>-test). Dots on the bar plots indicate the values of individual experiments.</p>
Full article ">
15 pages, 2173 KiB  
Review
The Complex Network of ADP-Ribosylation and DNA Repair: Emerging Insights and Implications for Cancer Therapy
by Ziyuan Li, Aiqin Luo and Bingteng Xie
Int. J. Mol. Sci. 2023, 24(19), 15028; https://doi.org/10.3390/ijms241915028 - 9 Oct 2023
Cited by 1 | Viewed by 1782
Abstract
ADP-ribosylation is a post-translational modification of proteins that plays a key role in various cellular processes, including DNA repair. Recently, significant progress has been made in understanding the mechanism and function of ADP-ribosylation in DNA repair. ADP-ribosylation can regulate the recruitment and activity [...] Read more.
ADP-ribosylation is a post-translational modification of proteins that plays a key role in various cellular processes, including DNA repair. Recently, significant progress has been made in understanding the mechanism and function of ADP-ribosylation in DNA repair. ADP-ribosylation can regulate the recruitment and activity of DNA repair proteins by facilitating protein–protein interactions and regulating protein conformations. Moreover, ADP-ribosylation can influence additional post-translational modifications (PTMs) of proteins involved in DNA repair, such as ubiquitination, methylation, acetylation, phosphorylation, and SUMOylation. The interaction between ADP-ribosylation and these additional PTMs can fine-tune the activity of DNA repair proteins and ensure the proper execution of the DNA repair process. In addition, PARP inhibitors have been developed as a promising cancer therapeutic strategy by exploiting the dependence of certain cancer types on the PARP-mediated DNA repair pathway. In this paper, we review the progress of ADP-ribosylation in DNA repair, discuss the crosstalk of ADP-ribosylation with additional PTMs in DNA repair, and summarize the progress of PARP inhibitors in cancer therapy. Full article
(This article belongs to the Special Issue Molecular Mechanism of DNA Replication and Repair, 2nd Edition )
Show Figures

Figure 1

Figure 1
<p>Overview of ADP-ribosylation and its writers and erasers: (<b>A</b>) schematic representation of the structure of poly(ADP)-ribosylation; (<b>B</b>) known ADP-ribosylation writers [<a href="#B16-ijms-24-15028" class="html-bibr">16</a>] and (<b>C</b>) erasers and their catalytic properties (PARG [<a href="#B24-ijms-24-15028" class="html-bibr">24</a>], MacroD1 [<a href="#B25-ijms-24-15028" class="html-bibr">25</a>], MacroD2 [<a href="#B28-ijms-24-15028" class="html-bibr">28</a>], TARG1 [<a href="#B26-ijms-24-15028" class="html-bibr">26</a>], ARH1 [<a href="#B29-ijms-24-15028" class="html-bibr">29</a>], ARH3 [<a href="#B27-ijms-24-15028" class="html-bibr">27</a>], NUDT9 [<a href="#B31-ijms-24-15028" class="html-bibr">31</a>], NUDT16 [<a href="#B32-ijms-24-15028" class="html-bibr">32</a>], ENPP1 [<a href="#B33-ijms-24-15028" class="html-bibr">33</a>], PARP9/14 [<a href="#B34-ijms-24-15028" class="html-bibr">34</a>]).</p>
Full article ">Figure 2
<p>Crosstalk of ADP-ribosylation with other protein post-translational modifications. Summary of interaction relationships for ADP-ribosylation and the remaining five modifications, i.e., ubiquitination, methylation, acetylation, phosphorylation, and SUMOylation. Examples are given to illustrate the relationship between them (a [<a href="#B52-ijms-24-15028" class="html-bibr">52</a>], b [<a href="#B53-ijms-24-15028" class="html-bibr">53</a>], c [<a href="#B54-ijms-24-15028" class="html-bibr">54</a>], d [<a href="#B55-ijms-24-15028" class="html-bibr">55</a>,<a href="#B56-ijms-24-15028" class="html-bibr">56</a>], e [<a href="#B57-ijms-24-15028" class="html-bibr">57</a>], f [<a href="#B62-ijms-24-15028" class="html-bibr">62</a>], g [<a href="#B61-ijms-24-15028" class="html-bibr">61</a>], h [<a href="#B58-ijms-24-15028" class="html-bibr">58</a>,<a href="#B59-ijms-24-15028" class="html-bibr">59</a>], i [<a href="#B60-ijms-24-15028" class="html-bibr">60</a>]).</p>
Full article ">Figure 3
<p>Mechanisms of drug resistance to PARPi. There are numerous reasons for PARPi resistance, including efflux of PARPi, reduced PARP1 trapping, restoration of HR repair, re-establishment of replication bifurcation stability, etc.</p>
Full article ">
36 pages, 9847 KiB  
Article
Imidazo-Pyrazole-Loaded Palmitic Acid and Polystyrene-Based Nanoparticles: Synthesis, Characterization and Antiproliferative Activity on Chemo-Resistant Human Neuroblastoma Cells
by Giulia Elda Valenti, Barbara Marengo, Marco Milanese, Guendalina Zuccari, Chiara Brullo, Cinzia Domenicotti and Silvana Alfei
Int. J. Mol. Sci. 2023, 24(19), 15027; https://doi.org/10.3390/ijms241915027 - 9 Oct 2023
Cited by 1 | Viewed by 1395
Abstract
Neuroblastoma (NB) is a childhood cancer, commonly treated with drugs, such as etoposide (ETO), whose efficacy is limited by the onset of resistance. Here, aiming at identifying new treatments for chemo-resistant NB, the effects of two synthesized imidazo-pyrazoles (IMPs) (4G and 4I [...] Read more.
Neuroblastoma (NB) is a childhood cancer, commonly treated with drugs, such as etoposide (ETO), whose efficacy is limited by the onset of resistance. Here, aiming at identifying new treatments for chemo-resistant NB, the effects of two synthesized imidazo-pyrazoles (IMPs) (4G and 4I) were investigated on ETO-sensitive (HTLA-230) and ETO-resistant (HTLA-ER) NB cells, detecting 4I as the more promising compound, that demonstrated IC50 values lower than those of ETO on HTLA ER. Therefore, to further improve the activity of 4I, we developed 4I-loaded palmitic acid (PA) and polystyrene-based (P5) cationic nanoparticles (P5PA-4I NPs) with high drug loading (21%) and encapsulation efficiency (97%), by a single oil-in-water emulsification technique. Biocompatible PA was adopted as an emulsion stabilizer, while synthesized P5 acted as an encapsulating agent, solubilizer and hydrophilic–lipophilic balance (HLB) improver. Optic microscopy and cytofluorimetric analyses were performed to investigate the micromorphology, size and complexity distributions of P5PA-4I NPs, which were also structurally characterized by chemometric-assisted Fourier transform infrared spectroscopy (FTIR). Potentiometric titrations allowed us to estimate the milliequivalents of PA and basic nitrogen atoms present in NPs. P5PA-4I NPs afforded dispersions in water with excellent buffer capacity, essential to escape lysosomal degradation and promote long residence time inside cells. They were chemically stable in an aqueous medium for at least 40 days, while in dynamic light scattering (DLS) analyses, P5PA-4I showed a mean hydrodynamic diameter of 541 nm, small polydispersity (0.194), and low positive zeta potentials (+8.39 mV), assuring low haemolytic toxicity. Biological experiments on NB cells, demonstrated that P5PA-4I NPs induced ROS-dependent cytotoxic effects significantly higher than those of pristine 4I, showing a major efficacy compared to ETO in reducing cell viability in HTLA-ER cells. Collectively, this 4I-based nano-formulation could represent a new promising macromolecular platform to develop a new delivery system able to increase the cytotoxicity of the anticancer drugs. Full article
Show Figures

Figure 1

Figure 1
<p>IMP derivatives selected for screening.</p>
Full article ">Figure 2
<p>Chemical structure of PA (<b>a</b>) and P5 (<b>b</b>).</p>
Full article ">Figure 3
<p>Cell viability was evaluated in HTLA-230 (<b>a</b>) and HTLA-ER (<b>b</b>) cells exposed to increasing concentrations of <b>4G</b> (0–100 µM) for 24, 48 and 72 h. Bar graphs summarize quantitative data of the means ± SEM of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 vs. Control (Ctr) cells.</p>
Full article ">Figure 4
<p>Cell viability was evaluated in HTLA-230 (<b>a</b>) and HTLA-ER (<b>b</b>) cells exposed to increasing concentrations of <b>4I</b> (0–100 µM) for 24, 48 and 72 h. Bar graphs summarize quantitative data of the means ± SEM of three independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 vs. Control (Ctr) cells.</p>
Full article ">Figure 5
<p>ROS generation was analysed in HTLA-230 (<b>a</b>) and HTLA-ER (<b>b</b>) cells exposed to increasing concentrations (0–100 µM) of <b>4I</b> for 24, 48 and 72 h. Bar graphs summarize quantitative data of the means ± SEM of three independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001 vs. Control (Ctr) cells.</p>
Full article ">Figure 6
<p>ROS DCFH positive cells (%) vs. cells viability in HTLA-230 population (<b>a</b>) and in HTLA-ER population (<b>b</b>).</p>
Full article ">Figure 7
<p>Optical images captured on lyophilized P5 with a 4 × objective (<b>a</b>) and 10 × objective (<b>b</b>), on P5PA NPs with a 10 × objective (<b>c</b>,<b>d</b>) and on P5PA-4I NPs with a 10 × (<b>e</b>,<b>f</b>) and a 40 × objective (<b>g</b>).</p>
Full article ">Figure 8
<p>Optical images captured on sonicated aqueous dispersions of P5 with a 4 × objective (<b>a</b>), 20 × objective (<b>b</b>), and 40 × objective (<b>c</b>), of P5PA NPs with a 4 × objective (<b>d</b>), and a 40 × objective (<b>e</b>,<b>f</b>) and of P5PA-4I NPs with a 4 × objective (<b>g</b>), 10 × objective (<b>h</b>), 20 × objective (<b>i</b>) and a 40 × objective (<b>j</b>).</p>
Full article ">Figure 9
<p>Size distribution by intensity of P5PA (<b>a</b>) and P5PA-4I (<b>c</b>) and zeta potential of P5PA (<b>b</b>) and P5PA-4I (<b>d</b>).</p>
Full article ">Figure 10
<p>Suspension of P5 was prepared in type I deionised water 0.22 µM filtered at the concentration of 4.1 mg/mL (<span class="html-italic">W</span>/<span class="html-italic">V</span>). Representative flow cytometry graphs for ungated dot plot Forward Scatter high (FSC-HLog) vs. Side Scatter high (SSC-HLog), R1 gated particles in red (<b>a</b>); ungated dot plot Forward Scatter Area (FSC-ALog) vs. Forward Scatter High (FSC-HLog), R1 gated particles in red (<b>b</b>); R2 gated dot plot Side Scatter High (SSC-HLog) vs. Forward Scatter High (FSC-HLog) (<b>c</b>); R2 gated contour plot Side Scatter High (SSC-HLog) vs. Forward Scatter High (FSC-HLog) (<b>d</b>). Representative flow cytometry histogram plots showing the Forward Scatter High (FSC-HLog) and the Side Scatter High (SSC-HLog) distributions. Overall, 33,602 events per group were acquired (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 11
<p>Suspension of P5PA was prepared in type I deionised water 0.22 µM filtered at the concentration of 5.3 mg/mL (<span class="html-italic">W</span>/<span class="html-italic">V</span>). Representative flow cytometry graphs for ungated dot plot Forward Scatter high (FSC-HLog) vs. Side Scatter high (SSC-HLog), R1 gated particles in red (<b>a</b>); ungated dot plot Forward Scatter Area (FSC-ALog) vs. Forward Scatter High (FSC-HLog), R1 gated particles in red (<b>b</b>); R2 gated dot plot Side Scatter High (SSC-HLog) vs. Forward Scatter High (FSC-HLog) (<b>c</b>); R2 gated contour plot Side Scatter High (SSC-HLog) vs. Forward Scatter High (FSC-HLog) (<b>d</b>). Representative flow cytometry histogram plots showing the Forward Scatter High (FSC-HLog) and the Side Scatter High (SSC-HLog) distributions. Overall, 33,602 events per group were acquired (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 12
<p>Suspension of P5PA-4I was prepared in type I deionised water 0.22 µM filtered at the concentration of 5.1 mg/mL (<span class="html-italic">W</span>/<span class="html-italic">V</span>). Representative flow cytometry graphs for ungated dot plot Forward Scatter high (FSC-HLog) vs. Side Scatter high (SSC-HLog), R1 gated particles in red (<b>a</b>); ungated dot plot Forward Scatter Area (FSC-ALog) vs. Forward Scatter High (FSC-HLog), R1 gated particles in red (<b>b</b>); R2 gated dot plot Side Scatter High (SSC-HLog) vs. Forward Scatter High (FSC-HLog) (<b>c</b>); R2 gated contour plot Side Scatter High (SSC-HLog) vs. Forward Scatter High (FSC-HLog) (<b>d</b>). Representative flow cytometry histogram plots showing the Forward Scatter High (FSC-HLog) and the Side Scatter High (SSC-HLog) distributions. Overall, 33,602 events per group were acquired (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 13
<p>Representative flow cytometry histogram plots showing the Forward Scatter High (FSC-H Log) (<b>a</b>) and the Side Scatter High (SSC-H Log) (<b>b</b>) distributions of P5, P5PA and P5PA-4I NPs compared to the those of standard beads. Overall, 33,602 events per group were acquired. The colours of the histogram plots are the same of those used in <a href="#ijms-24-15027-f010" class="html-fig">Figure 10</a>, <a href="#ijms-24-15027-f011" class="html-fig">Figure 11</a> and <a href="#ijms-24-15027-f012" class="html-fig">Figure 12</a> and in <a href="#app1-ijms-24-15027" class="html-app">Figures S2 and S3</a>.</p>
Full article ">Figure 14
<p>ATR-FTIR spectra of PA, P5 and P5PA NPs.</p>
Full article ">Figure 15
<p>ATR-FTIR spectra of P5PA NPs (empty NPs), <b>4I</b> and P5PA-4I NPs (<b>4I</b>-loaded NPs).</p>
Full article ">Figure 16
<p>Score plot of PC1 (explaining the 64.9% of variance) vs. PC2 (explaining the 21.8% of variance).</p>
Full article ">Figure 17
<p>Potentiometric titration of P5PA and P5PA-4I NPs (<b>a</b>); first derivatives (dpH/dV) of the titration curves (<b>b</b>). Error bars knowingly omitted to not complicate the image. FD = first derivative.</p>
Full article ">Figure 18
<p>Buffer capacity curves (<b>a</b>) and average buffer capacity (<b>b</b>) of P5, P5PA, P5PA-4I NPs and <span class="html-italic">b</span>-PEI 25 kD.</p>
Full article ">Figure 19
<p>Score plot of PC1 (explaining the 97.6% of variance) vs. PC2 (explaining the 1.3% of variance). Numbers associated with names indicate the number of days after first titration. Overlapped samples indicated very similar titration data and chemical stability.</p>
Full article ">Figure 20
<p>Projection of the external data set of new samples P5PA_20 and P5PA4I_20 (red labels) in the score plot of PC1 (explaining the 97.6% of variance) vs. PC2 (explaining the 1.3% of variance) of the training set of the first ten samples (empty circles).</p>
Full article ">Figure 21
<p>Cumulative release curves of <b>4I</b> from P5PA-4I NPs and from an aqueous suspension of <b>4I,</b> having the same initial <b>4I</b> concentration (4.2 mg/mL).</p>
Full article ">Figure 22
<p>Dose- and time-dependent cytotoxicity activity of <b>4I</b> (light violet bars), and P5PA-4I NPs (red bars) at 24 h (bars without pattern), 48 h (bars with small squares pattern), and 72 h (bars with greater squares pattern) towards HTLA (<b>a</b>) and HTLA-ER (<b>b</b>) NB cells. Significance refers to control (*) or <b>4I</b> (°). Specifically, <span class="html-italic">p</span> &gt; 0.05 ns; ° (vs. <b>4I</b>); <span class="html-italic">p</span> &lt; 0.01 ** (vs. CTR), °° (vs. <b>4I</b>); <span class="html-italic">p</span> &lt; 0.001 *** (vs. CTR), °°° (vs. <b>4I</b>).</p>
Full article ">Figure 22 Cont.
<p>Dose- and time-dependent cytotoxicity activity of <b>4I</b> (light violet bars), and P5PA-4I NPs (red bars) at 24 h (bars without pattern), 48 h (bars with small squares pattern), and 72 h (bars with greater squares pattern) towards HTLA (<b>a</b>) and HTLA-ER (<b>b</b>) NB cells. Significance refers to control (*) or <b>4I</b> (°). Specifically, <span class="html-italic">p</span> &gt; 0.05 ns; ° (vs. <b>4I</b>); <span class="html-italic">p</span> &lt; 0.01 ** (vs. CTR), °° (vs. <b>4I</b>); <span class="html-italic">p</span> &lt; 0.001 *** (vs. CTR), °°° (vs. <b>4I</b>).</p>
Full article ">Figure 23
<p>Dose- and time-dependent cytotoxic activity of <b>4I</b> (blue line), P5PA-4I NPs (pink line) and ETO (green line) at 24 h in HTLA-ER NB cells. Significance refers to control (*). Specifically, <span class="html-italic">p</span> &gt; 0.05 ns; <span class="html-italic">p</span> &lt; 0.01 **; <span class="html-italic">p</span> &lt; 0.001 ***.</p>
Full article ">Figure 24
<p>H<sub>2</sub>O<sub>2</sub> production was analyzed in HTLA-230 (<b>a</b>) and HTLA-ER (<b>b</b>) cells exposed to increasing concentrations (0–50 µM) of both <b>4I</b> (light violet bars), and P5PA-4I NPs (red bars) at 24 h (bars without pattern), 48 h (bars with small squares pattern), and 72 h (bars with greater squares pattern). Bar graphs summarize quantitative data of the means ± SEM of three independent experiments. Significance refers to control (*) or <b>4I</b> (°). Specifically, <span class="html-italic">p</span> &gt; 0.05 ns; <span class="html-italic">p</span> &lt; 0.05 * (vs. CTR), ° (vs. <b>4I</b>); <span class="html-italic">p</span> &lt; 0.01 ** (vs. CTR); <span class="html-italic">p</span> &lt; 0.001 *** (vs. CTR), °°° (vs. 4I).</p>
Full article ">Figure 25
<p>H<sub>2</sub>O<sub>2</sub> production analysed in HTLA-ER cells exposed to increasing concentrations (0–50 µM) of <b>4I</b> (blue line), P5PA-4I NPs (pink line) and ETO (green line) at 24 h. Data has been reported as means ± SEM of three independent experiments. Significance refers to control (*). Specifically, <span class="html-italic">p</span> &gt; 0.05 ns; <span class="html-italic">p</span> &lt; 0.01 **; <span class="html-italic">p</span> &lt; 0.001 ***.</p>
Full article ">Figure 26
<p>Dispersion graphs of HTLA-230 (<b>a</b>) and of HTLA-ER cells (<b>b</b>).</p>
Full article ">Scheme 1
<p>Preparation of P5PA-4I NPs by a single oil-in-water (O/W) emulsion method. DCM = dichloromethane; r.t. = room temperature.</p>
Full article ">
27 pages, 4838 KiB  
Article
Development of Novel Class of Phenylpyrazolo[3,4-d]pyrimidine-Based Analogs with Potent Anticancer Activity and Multitarget Enzyme Inhibition Supported by Docking Studies
by Ahmed K. B. Aljohani, Waheed Ali Zaki El Zaloa, Mohamed Alswah, Mohamed A. Seleem, Mohamed M. Elsebaei, Ashraf H. Bayoumi, Ahmed M. El-Morsy, Mohammed Almaghrabi, Aeshah A. Awaji, Ali Hammad, Marwa Alsulaimany and Hany E. A. Ahmed
Int. J. Mol. Sci. 2023, 24(19), 15026; https://doi.org/10.3390/ijms241915026 - 9 Oct 2023
Cited by 2 | Viewed by 1443
Abstract
Phenylpyrazolo[3,4-d]pyrimidine is considered a milestone scaffold known to possess various biological activities such as antiparasitic, antifungal, antimicrobial, and antiproliferative activities. In addition, the urgent need for selective and potent novel anticancer agents represents a major route in the drug discovery process. [...] Read more.
Phenylpyrazolo[3,4-d]pyrimidine is considered a milestone scaffold known to possess various biological activities such as antiparasitic, antifungal, antimicrobial, and antiproliferative activities. In addition, the urgent need for selective and potent novel anticancer agents represents a major route in the drug discovery process. Herein, new aryl analogs were synthesized and evaluated for their anticancer effects on a panel of cancer cell lines: MCF-7, HCT116, and HePG-2. Some of these compounds showed potent cytotoxicity, with variable degrees of potency and cell line selectivity in antiproliferative assays with low resistance. As the analogs carry the pyrazolopyrimidine scaffold, which looks structurally very similar to tyrosine and receptor kinase inhibitors, the potent compounds were evaluated for their inhibitory effects on three essential cancer targets: EGFRWT, EGFRT790M, VGFR2, and Top-II. The data obtained revealed that most of these compounds were potent, with variable degrees of target selectivity and dual EGFR/VGFR2 inhibitors at the IC50 value range, i.e., 0.3–24 µM. Among these, compound 5i was the most potent non-selective dual EGFR/VGFR2 inhibitor, with inhibitory concentrations of 0.3 and 7.60 µM, respectively. When 5i was tested in an MCF-7 model, it effectively inhibited tumor growth, strongly induced cancer cell apoptosis, inhibited cell migration, and suppressed cell cycle progression leading to DNA fragmentation. Molecular docking studies were performed to explore the binding mode and mechanism of such compounds on protein targets and mapped with reference ligands. The results of our studies indicate that the newly discovered phenylpyrazolo[3,4-d]pyrimidine-based multitarget inhibitors have significant potential for anticancer treatment. Full article
(This article belongs to the Special Issue Natural Products and Synthetic Compounds for Drug Development)
Show Figures

Figure 1

Figure 1
<p>Examples of different generations of EGFR-TK inhibitors.</p>
Full article ">Figure 2
<p>Examples of pyrazolo[3,4-<span class="html-italic">d</span>]pyrimidine-based EGFR-TK inhibitors.</p>
Full article ">Figure 3
<p>Work design. (<b>a</b>) Structure of ATP binding pocket, (<b>b</b>) Mapping of reference EGFR-TK inhibitors to pharmacophore points.</p>
Full article ">Figure 4
<p>Relative cytotoxic activity of the selected molecules upon incubation with the selected cell lines for 48 h (<b>a</b>), 72 h (<b>b</b>), and 24 h (<b>c</b>).</p>
Full article ">Figure 4 Cont.
<p>Relative cytotoxic activity of the selected molecules upon incubation with the selected cell lines for 48 h (<b>a</b>), 72 h (<b>b</b>), and 24 h (<b>c</b>).</p>
Full article ">Figure 4 Cont.
<p>Relative cytotoxic activity of the selected molecules upon incubation with the selected cell lines for 48 h (<b>a</b>), 72 h (<b>b</b>), and 24 h (<b>c</b>).</p>
Full article ">Figure 5
<p>Dot plot of Annexin V/PI double staining of MCF-7 control and treatment with analogue <b>5i</b>. Statistical analysis of the apoptosis percentage of MCF-7 cells after incubation with compounds <b>5i</b> in conc. (3.81 μM) for 24 h. The data are reported as the mean ± SD of three independent experiments in triplicate.</p>
Full article ">Figure 6
<p>DNA content distribution histograms of control and treated cells. Statistical analysis of cell cycle phases percentage of MCF-7 cells after incubation with compound <b>5i</b> (3.81 μM) for 24 h. The data are reported as the mean ± SD of three independent experiments in triplicate.</p>
Full article ">Figure 7
<p>Docking modes of active compound <b>5i.</b> Shown are the predicted complexes of (<b>a</b>) an erlotinib drug with EGFR and (<b>b</b>) a pyridyl-pyrimidine benzimidazole derivative with VGFR.</p>
Full article ">Figure 7 Cont.
<p>Docking modes of active compound <b>5i.</b> Shown are the predicted complexes of (<b>a</b>) an erlotinib drug with EGFR and (<b>b</b>) a pyridyl-pyrimidine benzimidazole derivative with VGFR.</p>
Full article ">Figure 8
<p>Molecular mapping of compound <b>5i</b> (<b>middle</b>) to reference bound ligands within protein targets. Color coding of similar fragments: red, green, and cyan types.</p>
Full article ">Scheme 1
<p>Synthetic protocol of 4-(2-arylidenehydrazinyl)-6methyl-1-phenyl-1<span class="html-italic">H</span>-pyrazolo[3,4-<span class="html-italic">d</span>]pyrimidine (<b>5a</b>–<b>l</b>).</p>
Full article ">Scheme 2
<p>Synthetic protocol of pyrazolopyrimidine derivatives (<b>6</b>–<b>9</b>).</p>
Full article ">
20 pages, 8177 KiB  
Article
The Co-Expression Pattern of Calcium-Binding Proteins with γ-Aminobutyric Acid and Glutamate Transporters in the Amygdala of the Guinea Pig: Evidence for Glutamatergic Subpopulations
by Daniel Kalinowski, Krystyna Bogus-Nowakowska, Anna Kozłowska and Maciej Równiak
Int. J. Mol. Sci. 2023, 24(19), 15025; https://doi.org/10.3390/ijms241915025 - 9 Oct 2023
Viewed by 1438
Abstract
The amygdala has large populations of neurons utilizing specific calcium-binding proteins such as parvalbumin (PV), calbindin (CB), or calretinin (CR). They are considered specialized subsets of γ-aminobutyric acid (GABA) interneurons; however, many of these cells are devoid of GABA or glutamate decarboxylase. The [...] Read more.
The amygdala has large populations of neurons utilizing specific calcium-binding proteins such as parvalbumin (PV), calbindin (CB), or calretinin (CR). They are considered specialized subsets of γ-aminobutyric acid (GABA) interneurons; however, many of these cells are devoid of GABA or glutamate decarboxylase. The neurotransmitters used by GABA-immunonegative cells are still unknown, but it is suggested that a part may use glutamate. Thus, this study investigates in the amygdala of the guinea pig relationships between PV, CB, or CR-containing cells and GABA transporter (VGAT) or glutamate transporter type 2 (VGLUT2), markers of GABAergic and glutamatergic neurons, respectively. The results show that although most neurons using PV, CB, and CR co-expressed VGAT, each of these populations also had a fraction of VGLUT2 co-expressing cells. For almost all neurons using PV (~90%) co-expressed VGAT, while ~1.5% of them had VGLUT2. The proportion of neurons using CB and VGAT was smaller than that for PV (~80%), while the percentage of cells with VGLUT2 was larger (~4.5%). Finally, only half of the neurons using CR (~53%) co-expressed VGAT, while ~3.5% of them had VGLUT2. In conclusion, the populations of neurons co-expressing PV, CB, and CR are in the amygdala, primarily GABAergic. However, at least a fraction of neurons in each of them co-express VGLUT2, suggesting that these cells may use glutamate. Moreover, the number of PV-, CB-, and CR-containing neurons that may use glutamate is probably larger as they can utilize VGLUT1 or VGLUT3, which are also present in the amygdala. Full article
(This article belongs to the Section Molecular Neurobiology)
Show Figures

Figure 1

Figure 1
<p>(Modified) All figures were published in Równiak et al. [<a href="#B27-ijms-24-15025" class="html-bibr">27</a>]. Brightfield photomicrographs highlighting the patterns of parvalbumin (PV, <b>A</b>–<b>A″</b>), calbindin (CB, <b>B</b>–<b>B″</b>), and calretinin (CR, <b>C</b>–<b>C″</b>) immunoreactivity in the amygdala of the male guinea pig. The scale bar applies to all microphotographs, and it corresponds to the length of 2000 μm (<b>A</b>–<b>C</b>) and 200 μm (<b>A′</b>–<b>C′</b>, <b>A″</b>–<b>C″</b>). LAdl–lateral nucleus, lateral part, LAm–lateral nucleus, medial part, LAvl–lateral nucleus, ventral part, BLa–basomedial nucleus, anterior part, BMa–basomedial nucleus, anterior part, MEd–medial nucleus, dorsal part, CEm–central nucleus, medial part, CEl–central nucleus, lateral part, COP–cortical nucleus, PAC–piriform-amygdalar area. Noted that nuclei are indicated by the dotted line.</p>
Full article ">Figure 2
<p>The volume density of vesicular GABA transporter (VGAT, <b>A</b>) and vesicular glutamate transporter type 2 (VGLUT2, <b>B</b>) in the “cortex-like” and “striatum-pallidum-like” amygdala of the guinea pig (n = 5). Note significant differences between “cortex-like” and “stratum-pallidum-like” nuclei (<b>A1</b>–<b>A4</b>). Data are expressed as box-and-whiskers plots, with the “box” depicting the median and the 25th and 75th quartiles and the “whiskers” showing the 5th and 95th percentile. * (<span class="html-italic">p</span> ≤ 0.05), ** (<span class="html-italic">p</span> ≤ 0.01), and *** (<span class="html-italic">p</span> ≤ 0.001) indicate statistically significant differences between studied nuclei analyzed by Student’s t-test, whereas ns means none statistically significant. Elements–immunoreactive somata, fibers, and neuropil. LA—lateral nucleus, BL—basolateral nucleus, BM—basomedial nucleus, CO—cortical nucleus, CE—central nucleus, ME—medial nucleus.</p>
Full article ">Figure 3
<p>Representative color photomicrographs illustrating immunoreactivity of vesicular GABA transporter (VGAT, <b>A</b>–<b>C</b>) in the amygdala of guinea pig. The scale bar applies to all microphotographs, and it corresponds to the length of 200 μm. LA—lateral nucleus, BL—basolateral nucleus, BM—basomedial nucleus as a “cortex-like” amygdala, CE—central nucleus, ME—medial nucleus as a “striatum-pallidum-like” amygdala. Noted that nuclei are indicated by the dotted line.</p>
Full article ">Figure 4
<p>Representative color photomicrographs illustrating the co-expression pattern of PV (<b>A</b>–<b>A‴</b>), CB (<b>B</b>–<b>B‴</b>), and CR (<b>C</b>–<b>C‴</b>) neurons with VGAT in the basolateral (<b>A</b>–<b>A‴</b>), basomedial (<b>B</b>–<b>B‴</b>) and cortical (<b>C</b>–<b>C‴</b>) nuclei as a “cortex-like” amygdala of the guinea pig. Arrows indicate double-labeled cells. Boxed regions in (<b>A″</b>–<b>C″</b>) are shown at a greater magnification in (<b>A‴</b>–<b>C‴</b>). The scale bar applies to all microphotographs, and it corresponds to the length of 50 μm.</p>
Full article ">Figure 5
<p>Representative color photomicrographs illustrating the co-expression pattern of CB (<b>A</b>–<b>A‴</b>) and CR (<b>B</b>–<b>B‴</b>) neurons with VGAT in the medial nucleus (<b>A</b>–<b>B‴</b>) as a “striatum-pallidum-like” amygdala of the guinea pig. Arrows indicate double-labeled cells. Boxed regions in (<b>A″</b>,<b>B″</b>) are shown at a greater magnification in (<b>A‴</b>,<b>B‴</b>). The scale bar applies to all microphotographs, and it corresponds to the length of 50 μm.</p>
Full article ">Figure 6
<p>Representative color photomicrographs illustrating the co-expression pattern of PV (<b>A</b>–<b>A‴</b>), CB (<b>B</b>–<b>B‴</b>), and CR (<b>C</b>–<b>C‴</b>) neurons with VGLUT2 in the basolateral (<b>A</b>–<b>A‴</b>,<b>C</b>–<b>C‴</b>) and cortical (<b>B</b>–<b>B‴</b>) nuclei as a “cortex-like” amygdala of the guinea pig. Arrows indicate double-labeled cells. Boxed regions in (<b>A″</b>–<b>C″</b>) are shown at a greater magnification in (<b>A‴</b>–<b>C‴</b>). The scale bar applies to all microphotographs, and it corresponds to the length of 50 μm.</p>
Full article ">Figure 7
<p>Representative color photomicrographs illustrating the co-expression pattern of CB (<b>A</b>–<b>A‴</b>) and CR (<b>B</b>–<b>B‴</b>) neurons with VGLUT2 in the central (<b>A</b>–<b>A‴</b>) and medial (<b>B</b>–<b>B‴</b>) nuclei as a “striatum-pallidum-like” amygdala of the guinea pig. Arrows indicate double-labeled cells. Boxed regions in (<b>A″</b>,<b>B″</b>) are shown at a greater magnification in (<b>A‴</b>,<b>B‴</b>). The scale bar applies to all microphotographs, and it corresponds to the length of 50 μm.</p>
Full article ">
21 pages, 3592 KiB  
Article
Multiple Sclerosis-Associated Gut Microbiome in the Israeli Diverse Populations: Associations with Ethnicity, Gender, Disability Status, Vitamin D Levels, and Mediterranean Diet
by Zehavit Nitzan, Elsebeth Staun-Ram, Anat Volkowich and Ariel Miller
Int. J. Mol. Sci. 2023, 24(19), 15024; https://doi.org/10.3390/ijms241915024 - 9 Oct 2023
Cited by 2 | Viewed by 1611
Abstract
Microbiome dysbiosis is increasingly being recognized as implicated in immune-mediated disorders including multiple sclerosis (MS). The microbiome is modulated by genetic and environmental factors including lifestyle, diet, and drug intake. This study aimed to characterize the MS-associated gut microbiome in the Israeli populations [...] Read more.
Microbiome dysbiosis is increasingly being recognized as implicated in immune-mediated disorders including multiple sclerosis (MS). The microbiome is modulated by genetic and environmental factors including lifestyle, diet, and drug intake. This study aimed to characterize the MS-associated gut microbiome in the Israeli populations and to identify associations with demographic, dietary, and clinical features. The microbiota from 57 treatment-naive patients with MS (PwMS) and 43 age- and gender-matched healthy controls (HCs) was sequenced and abundance compared. Associations between differential microbes with demographic or clinical characteristics, as well as diet and nutrient intake, were assessed. While there was no difference in α- or β-diversity of the microbiome, we identified 40 microbes from different taxonomic levels that differ in abundance between PwMS and HCs, including Barnesiella, Collinsella, Egerthella, Mitsuokella, Olsenella Romboutsia, and Succinivibrio, all enhanced in PwMS, while several members of Lacnospira were reduced. Additional MS-differential microbes specific to ethnicity were identified. Several MS-specific microbial patterns were associated with gender, vitamin D level, Mediterranean diet, nutrient intake, or disability status. Thus, PwMS have altered microbiota composition, with distinctive patterns related to geographic locations and population. Microbiome dysbiosis seem to be implicated in disease progression, gender-related differences, and vitamin D-mediated immunological effects recognized in MS. Dietary interventions may be beneficial in restoring a “healthy microbiota” as part of applying comprehensive personalized therapeutic strategies for PwMS. Full article
(This article belongs to the Special Issue Molecular Mechanism in Multiple Sclerosis and Related Disorders)
Show Figures

Figure 1

Figure 1
<p><b>General microbiome composition.</b> (<b>A</b>): Relative composition at the class level. (<b>B</b>): Alpha-diversity between MS and HC calculated by Shannon index at the OTU level. (<b>C</b>): Beta-diversity between MS and HC calculated by Bray–Curtis dissimilarity at the OTU level. MS—multiple sclerosis.</p>
Full article ">Figure 2
<p><b>Differential abundant taxa between PwMS and HCs.</b> (<b>A</b>)—Representative graphs of taxa with significant difference in relative abundance, as determined by at least two out of three tools (DESeq2, MetagenomeSeq, and EdgeR) at different taxonomic levels (FDR &lt; 0.1). Each black dot represents the relative abundance of a participant sample. (<b>B</b>)—Linear discriminant analysis (LDA) score of significant different taxa as determined by LDA effect size (LefSE) between PwMS and HCs at different taxonomic levels (<span class="html-italic">p</span> &lt; 0.05). (<b>C</b>)—Venn diagram presenting the overlap of differentially abundant taxa identified in the analysis of all samples, Jewish participants or Arab participants. HC—healthy control, MS—multiple sclerosis.</p>
Full article ">Figure 3
<p><b>Associations between MS-differentially abundant microbiota, EDSS, vitamin D, and gender.</b> (<b>A</b>)—Representative graphs of MS-differentially abundant taxa, which significantly correlates with EDSS (Spearman correlation, adjusted for BMI). (<b>B</b>)—Representative graphs of MS-differentially abundant taxa, which significantly correlates with serum vitamin D level shown as continuous or divided into 3 groups: &lt;50 nMol/L—vitamin D insufficiency, 50–75 nMol/L—vitamin D mild deficiency, &gt;75 nMol/L—vitamin D sufficient (Spearman correlation, adjusted for BMI). (<b>C</b>)—Representative graphs of MS-differentially abundant taxa enriched in PwMS (<b>I</b>) or reduced in PwMS (<b>II</b>), which differ significantly between females and males. Figure presents FDR calculated from EdgeR, * FDR from DeSeq2. Each black dot represents the relative abundance of a participant sample.</p>
Full article ">Figure 4
<p><b>Correlations between MS-differentially abundant microbiota and dietary data.</b> (<b>A</b>)—The distribution of adherence to a Mediterranean diet (MDS) in PwMS and HCs (Kruskal–Wallis test). (<b>B</b>)—Alpha-diversity Shannon index and beta-diversity Bray–Curtis dissimilarity according to high, low, or intermediate MDS (OTU level). (<b>C</b>)—Representative graphs of differentially abundant taxa, enriched in PwMS (<b>I</b>) or reduced in PwMS (<b>II</b>) which significantly correlates with MDS (Spearman correlation). MDS shown as continuous or divided into 3 groups (MDS low (1–6 points), intermediate (7–11 points), high (2–17 points). MDS—Mediterranean diet score.</p>
Full article ">Figure 5
<p><b>Dietary recommendations.</b> Potential beneficial dietary recommendations for modulation and “repair” of altered microbiota abundance in PwMS, based upon correlations between nutrients and MS-differentially abundant taxa.</p>
Full article ">Figure 6
<p><b>Predicted functional pathways.</b> (<b>A</b>). Main KEGG metabolism pathways in the MS and healthy control cohorts (<b>B</b>). KOs discriminating between PwMS and HCs according to LEfSe at FDR &lt; 0.1. KEGG—Kyoto Encyclopedia of Genes and Genomes, LEfSe: linear discriminant analysis effect size.</p>
Full article ">
26 pages, 2408 KiB  
Review
Aggregation, Transmission, and Toxicity of the Microtubule-Associated Protein Tau: A Complex Comprehension
by Jiaxin Hu, Wenchi Sha, Shuangshuang Yuan, Jiarui Wu and Yunpeng Huang
Int. J. Mol. Sci. 2023, 24(19), 15023; https://doi.org/10.3390/ijms241915023 - 9 Oct 2023
Cited by 5 | Viewed by 2466
Abstract
The microtubule-associated protein tau is an intrinsically disordered protein containing a few short and transient secondary structures. Tau physiologically associates with microtubules (MTs) for its stabilization and detaches from MTs to regulate its dynamics. Under pathological conditions, tau is abnormally modified, detaches from [...] Read more.
The microtubule-associated protein tau is an intrinsically disordered protein containing a few short and transient secondary structures. Tau physiologically associates with microtubules (MTs) for its stabilization and detaches from MTs to regulate its dynamics. Under pathological conditions, tau is abnormally modified, detaches from MTs, and forms protein aggregates in neuronal and glial cells. Tau protein aggregates can be found in a number of devastating neurodegenerative diseases known as “tauopathies”, such as Alzheimer’s disease (AD), frontotemporal dementia (FTD), corticobasal degeneration (CBD), etc. However, it is still unclear how the tau protein is compacted into ordered protein aggregates, and the toxicity of the aggregates is still debated. Fortunately, there has been considerable progress in the study of tau in recent years, particularly in the understanding of the intercellular transmission of pathological tau species, the structure of tau aggregates, and the conformational change events in the tau polymerization process. In this review, we summarize the concepts of tau protein aggregation and discuss the views on tau protein transmission and toxicity. Full article
(This article belongs to the Section Molecular Biology)
Show Figures

Figure 1

Figure 1
<p>Tau isoforms. Six isoforms of microtubule-associated protein tau (MAPT) are generated through alternative splicing. In the human brain, splicing of exons 2 and 3 results in three isoforms with 2, 1, or no N-terminus insertion of 29 amino acids (referred to as 2N, 1N, and 0N). Each isoform contains a microtubule binding domain encoded by exon 10 that consists of either three repeats (3R) or four repeats (4R), resulting in the genesis of six isoforms (2N4R, 1N4R, 0N4R, 2N3R, 1N3R, and 0N3R).</p>
Full article ">Figure 2
<p>A stepwise model elucidating the process of tau aggregation. The microtubule-associated protein tau is physiologically associated with the tubulin heterodimer to maintain the stability of microtubules. Following post-translational modifications (PTMs) such as phosphorylation, tau undergoes dissociation from the tubulin heterodimer. The dissociation of tau monomers and the naturally unfolded “paper clip” tau species undergo a series of mis-sorting processes, resulting in the dendritic and somatic mislocalization of tau. This further facilitates PTMs and liquid–liquid phase separation (LLPS). The presence of PTMs and LLPS further facilitates the conformational changes in Tau, thereby exposing its aggregation-prone motifs. The intermolecular interface of tau facilitates the nucleation and formation of tau dimers and trimers, which serve as seeds to promote gradual polymerization through the extension of aggregates by the addition of tau monomers, ultimately resulting in the formation of oligomers and filaments that accompany conformational changes.</p>
Full article ">Figure 3
<p>Post-translational modifications (PTMs) occur on the tau protein. The amino acid sequence of the longest tau isoform, 2N4R, is presented here. Distinctive colored dots are utilized to represent various PTMs, while the amino acids associated with tau pathology are highlighted in yellow.</p>
Full article ">Figure 4
<p>Tau transmission. The transmission of tau is initiated with the release of pathological tau species, including monomers, dimers, oligomers, and NFTs. This process is facilitated by (1) the direct translocation across the plasma membrane, (2) membranous organelle-based unconventional secretion (MOBS), (3) shedding of microvesicles derived from the plasma membrane, and (4) nanotube-mediated transfer. The extracellular tau and tau-containing vesicles can be internalized by (5) translocation across the membrane, (6) micropinocytosis, and (7) endocytosis. Subsequently, they can be (8) released into the recipient cells, (9) seeding the endogenous tau to undergo conformational change and polymerization.</p>
Full article ">Figure 5
<p>The toxic effects of tau aggregates. The deleterious effects of Tau aggregates encompass a variety of biological processes, including destabilization of microtubules and cytoskeletons, impairment of axonal transport and synaptic plasticity, disruption of mitochondrial function and proteostasis, hyperactivation of microglia, and induction of neuronal cell death, among others.</p>
Full article ">
15 pages, 7155 KiB  
Article
Robust AMBER Force Field Parameters for Glutathionylated Cysteines
by Zineb Elftmaoui and Emmanuelle Bignon
Int. J. Mol. Sci. 2023, 24(19), 15022; https://doi.org/10.3390/ijms241915022 - 9 Oct 2023
Viewed by 1033
Abstract
S-glutathionylation is an oxidative post-translational modification, which is involved in the regulation of many cell signaling pathways. Increasing amounts of studies show that it is crucial in cell homeostasis and deregulated in several pathologies. However, the effect of S-glutathionylation on proteins’ structure and [...] Read more.
S-glutathionylation is an oxidative post-translational modification, which is involved in the regulation of many cell signaling pathways. Increasing amounts of studies show that it is crucial in cell homeostasis and deregulated in several pathologies. However, the effect of S-glutathionylation on proteins’ structure and activity is poorly understood, and a drastic lack of structural information at the atomic scale remains. Studies based on the use of molecular dynamics simulations, which can provide important information about modification-induced modulation of proteins’ structure and function, are also sparse, and there is no benchmarked force field parameters for this modified cysteine. In this contribution, we provide robust AMBER parameters for S-glutathionylation, which we tested extensively against experimental data through a total of 33 μs molecular dynamics simulations. We show that our parameter set efficiently describes the global and local structural properties of S-glutathionylated proteins. These data provide the community with an important tool to foster new investigations into the effect of S-glutathionylation on protein dynamics and function, in a common effort to unravel the structural mechanisms underlying its critical role in cellular processes. Full article
(This article belongs to the Section Physical Chemistry and Chemical Physics)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Structure of a glutathionylated cysteine. The glycine, cysteine, and glutamate residues within the glutathionylation moiety (SSG) are labeled in grey. The target cysteine backbone atoms are displayed as if embedded in a protein (without capping). (<b>B</b>) A few examples of proteins regulated by S-glutathionylation, with respect to their function [<a href="#B7-ijms-24-15022" class="html-bibr">7</a>,<a href="#B18-ijms-24-15022" class="html-bibr">18</a>,<a href="#B19-ijms-24-15022" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Boxplots of the RMSD values in Å of each system, considering all atoms (<b>top</b>) or the residues’ backbone atoms (<b>bottom</b>). RMSD values for dimers are shown by monomer (Mono 1 or Mono 2). The black lines inside each boxplot display the median values. The average and standard deviation values are written above each box.</p>
Full article ">Figure 3
<p>Representative structures of the two main conformations of hLyso as calculated from the cluster analysis of MD ensembles. The structures of the crystal reference and from the AF and crys simulations are displayed in gray, cyan, and blue, respectively. The percentage of occurrence of each cluster (primary on the left and secondary on the right) is also given.</p>
Full article ">Figure 4
<p>Boxplots of the S-glutathionylated cysteine (CSG) RMSD values in Å for each system. RMSD values for dimers are shown by monomer (Mono 1 or Mono 2). The black lines inside each boxplot display the median values. Average and standard deviation values are written above each box.</p>
Full article ">Figure 5
<p>Times series of CSG RMSD values in Å of the three replicates for AF simulations of AtGrxC5, hGSTO2, PtGSTL1, PtGSTL3, PtGrx12, PtGSTF1, ScGrx2, and TvGSTO2C, which show better trends. Plots for the dimeric systems (hGSTO2, PtGSTF1, and TvGSTO2C) are paired in the columns, with CSG1 and CSG2 referring to Monomers 1 and 2, respectively.</p>
Full article ">Figure 6
<p>(<b>A</b>) Representative structures of the major conformations of the CSG in AtGrxC5, TvGSTO2C Monomer 1, PtGSTL3, hLyso, ScTrx1, and PtTrxL2.1 as calculated from the cluster analysis of MD ensembles. Structures of the crystal reference and from the AF and crys simulations are displayed in gray, cyan, and blue, respectively. The percentage of occurrence of the clusters is also given. (<b>B</b>) Times series of the CSG RMSD values in Å of the three replicates for the AF (left) and crys (right) simulations of hLyso, ScTrx1, and PtTrxL2.1.</p>
Full article ">Figure 7
<p>(<b>A</b>) CSG structure, highlighting the side chain atoms used to monitor hydrogen bonds. (<b>B</b>–<b>D</b>) Distribution of distances corresponding to the native interactions identified in the reference structures of hLyso, PtTrxL2.1, and ScTrx1, respectively. Both distributions for AF (cyan) and crys (blue) simulations are shown. Cation–<math display="inline"><semantics> <mi>π</mi> </semantics></math> interactions involving Y63 in hLyso are monitored as distances between CSG.N1 and the center of mass of the aromatic ring heavy atoms of Y63 (Y63.Ph).</p>
Full article ">
60 pages, 22427 KiB  
Review
Graphitic Carbon Nitride/Zinc Oxide-Based Z-Scheme and S-Scheme Heterojunction Photocatalysts for the Photodegradation of Organic Pollutants
by Gopal Panthi and Mira Park
Int. J. Mol. Sci. 2023, 24(19), 15021; https://doi.org/10.3390/ijms241915021 - 9 Oct 2023
Cited by 5 | Viewed by 2416
Abstract
Graphitic carbon nitride (g-C3N4), a metal-free polymer semiconductor, has been recognized as an attractive photocatalytic material for environmental remediation because of its low band gap, high thermal and photostability, chemical inertness, non-toxicity, low cost, biocompatibility, and optical and electrical [...] Read more.
Graphitic carbon nitride (g-C3N4), a metal-free polymer semiconductor, has been recognized as an attractive photocatalytic material for environmental remediation because of its low band gap, high thermal and photostability, chemical inertness, non-toxicity, low cost, biocompatibility, and optical and electrical efficiency. However, g-C3N4 has been reported to suffer from many difficulties in photocatalytic applications, such as a low specific surface area, inadequate visible-light utilization, and a high charge recombination rate. To overcome these difficulties, the formation of g-C3N4 heterojunctions by coupling with metal oxides has triggered tremendous interest in recent years. In this regard, zinc oxide (ZnO) is being largely explored as a self-driven semiconductor photocatalyst to form heterojunctions with g-C3N4, as ZnO possesses unique and fascinating properties, including high quantum efficiency, high electron mobility, cost-effectiveness, environmental friendliness, and a simple synthetic procedure. The synergistic effect of its properties, such as adsorption and photogenerated charge separation, was found to enhance the photocatalytic activity of heterojunctions. Hence, this review aims to compile the strategies for fabricating g-C3N4/ZnO-based Z-scheme and S-scheme heterojunction photocatalytic systems with enhanced performance and overall stability for the photodegradation of organic pollutants. Furthermore, with reference to the reported system, the photocatalytic mechanism of g-C3N4/ZnO-based heterojunction photocatalysts and their charge-transfer pathways on the interface surface are highlighted. Full article
(This article belongs to the Special Issue Catalysts: Design, Synthesis, and Molecular Applications)
Show Figures

Figure 1

Figure 1
<p>Structures of (<b>a</b>) triazine and (<b>b</b>) heptazine.</p>
Full article ">Figure 2
<p>Schematic representation of synthesis process of g-C<sub>3</sub>N<sub>4</sub> by thermal polymerization of different nitrogen-rich precursors.</p>
Full article ">Figure 3
<p>(<b>a</b>) High-temperature self-polymerization of urea and thiourea in air to form graphitic carbon nitride and (<b>b</b>) graphitic carbon nitride synthesis pathway using cyanamide as a precursor. Reprinted from Ref. [<a href="#B35-ijms-24-15021" class="html-bibr">35</a>] with permission from Elsevier.</p>
Full article ">Figure 4
<p>Thermal graphitic carbon nitride self-polymerization synthesis route for (<b>a</b>) guanidine thiocyanate and (<b>b</b>) guanidine hydrochloride. Reprinted from Ref. [<a href="#B35-ijms-24-15021" class="html-bibr">35</a>] with permission from Elsevier.</p>
Full article ">Figure 5
<p>(<b>a</b>) Different morphological structures of g-C<sub>3</sub>N<sub>4</sub> and (<b>b</b>) different band structures of some representative photocatalysts.</p>
Full article ">Figure 6
<p>Schematic illustrations of different types of heterojunction photocatalysts along with charge-transfer processes; (<b>a</b>–<b>d</b>) Z-scheme heterojunctions and (<b>e</b>) S-scheme heterojunction.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic illustration of the plausible photocatalytic mechanism of ZnO/Fe<sub>2</sub>O<sub>3</sub>/g-C<sub>3</sub>N<sub>4</sub> heterojunction under visible-light irradiation, (<b>b</b>) SEM image of ACZ, (<b>c</b>) photocatalytic reaction mechanism diagram of Ag<sub>3</sub>PO<sub>4</sub>/g-C<sub>3</sub>N<sub>4</sub>/ZnO composite, (<b>d</b>) TEM image of typical ZnO/C/g-C<sub>3</sub>N<sub>4</sub> nanofiber at 120,000 magnification, and (<b>e</b>) schematic diagram of the two possible proposed mechanisms of charge transfer, as well as the intermediate reactions for degradation of the pollutant over the ZnO/C/g-C<sub>3</sub>N<sub>4</sub> Z-scheme system with carbon at the interface between ZnO and g-C<sub>3</sub>N<sub>4</sub>. Reprinted from Refs. [<a href="#B127-ijms-24-15021" class="html-bibr">127</a>,<a href="#B130-ijms-24-15021" class="html-bibr">130</a>,<a href="#B136-ijms-24-15021" class="html-bibr">136</a>] with permission from Elsevier.</p>
Full article ">Figure 8
<p>Schematic illustration of the synthetic processes of Pbi-ZnO-g-C<sub>3</sub>N<sub>4</sub> (<b>a</b>), SEM image of Pbi-ZnO-g-C<sub>3</sub>N<sub>4</sub> (<b>b</b>), the photocatalytic degradation efficiency of atrazine by Pbi-ZnO and Pbi-ZnO-g-C<sub>3</sub>N<sub>4</sub> under simulated sunlight (<b>c</b>), atrazine degradation over Pbi-ZnO-g-C<sub>3</sub>N<sub>4</sub> during five recycling tests (<b>d</b>), and proposed mechanisms for the formation of reactive oxidative species during biochar- and Pbi-ZnO-g-C<sub>3</sub>N<sub>4</sub>-based photocatalytic processes (<b>e</b>). Reprinted from Ref. [<a href="#B138-ijms-24-15021" class="html-bibr">138</a>] with permission from Elsevier.</p>
Full article ">Figure 9
<p>(<b>a</b>) Schematic illustration of fabrication process of ALD-based g-C<sub>3</sub>N<sub>4</sub>@ZnO photocatalyst, (<b>b</b>) HRTEM image of a randomly selected position from g-C<sub>3</sub>N<sub>4</sub>@ZnO sample, (<b>c</b>) photocatalytic degradation of cephalexin by g-C<sub>3</sub>N<sub>4</sub>, g-C<sub>3</sub>N<sub>4</sub>@ZnO, and ZnO under simulated sunlight irradiation, and (<b>d</b>) schematic illustration of the mechanism of enhanced photocatalytic activity. Reprinted from Ref. [<a href="#B143-ijms-24-15021" class="html-bibr">143</a>] with permission from Elsevier.</p>
Full article ">Figure 10
<p>HRTEM images of (<b>a</b>) Thio-CNZ, (<b>b</b>) Urea-CNZ, and (<b>c</b>) DCDA-CNZ; (<b>d</b>) adsorption in the dark and photocatalytic degradation in the visible-light irradiation of MB over CNZ composite photocatalysts. Photocatalytic degradation curves of (<b>e</b>) CFZ and (<b>f</b>) RB5 by ZnO, gCN, and their heterostructures. Diagram illustrating the possible charge separation on 50-Zn/gCN through the Z-scheme mechanism (<b>g</b>). Reprinted from Refs. [<a href="#B148-ijms-24-15021" class="html-bibr">148</a>,<a href="#B157-ijms-24-15021" class="html-bibr">157</a>] with permission from Elsevier.</p>
Full article ">Figure 11
<p>(<b>a</b>) TEM image of CN/AB/ZO 30 and (<b>b</b>) schematic illustration of the proposed mechanism for the photocatalytic degradation of MB under irradiation by visible light with g-C<sub>3</sub>N<sub>4</sub>/AgBr/ZnO. (<b>c</b>) HR-TEM image of the as-fabricated CdS@ZnO/g-C<sub>3</sub>N<sub>4</sub> nanocomposite and (<b>d</b>) the general mechanism of the photocatalytic performance of the ternary composite CdS@ZnO/g-C<sub>3</sub>N<sub>4</sub>. (<b>e</b>) Schematic illustration of plausible Z-scheme mechanism exposing energy-band diagram with charge separation/transfer pathway in a g-C<sub>3</sub>N<sub>4</sub>/ZnO-Ag<sub>2</sub>O ternary PCs for photodegradation of organic pollutants under visible-light exposure. (<b>f</b>) Charge transfer based on the Z-scheme mechanism in MoS<sub>2</sub>/g-C<sub>3</sub>N<sub>4</sub>/ZnO. Reprinted from Refs. [<a href="#B163-ijms-24-15021" class="html-bibr">163</a>,<a href="#B166-ijms-24-15021" class="html-bibr">166</a>,<a href="#B167-ijms-24-15021" class="html-bibr">167</a>,<a href="#B169-ijms-24-15021" class="html-bibr">169</a>] with permission from Elsevier.</p>
Full article ">Figure 12
<p>(<b>a</b>) Generation of doping level between CB and VB by metal/non-metal doping in a semiconductor photocatalyst, (<b>b</b>) HRTEM images of ZnO/g-C<sub>3</sub>N<sub>4</sub> and (<b>c</b>) Cu-ZnO/g-C<sub>3</sub>N<sub>4</sub>. (<b>d</b>) Photocatalytic mechanism for degradation of atrazine by the synthesized Cu-ZnO/g-C<sub>3</sub>N<sub>4</sub>. (<b>e</b>) X-ray diffraction patterns of ZnO, Ni/ZnO, NiZG-70, and ZG-70 composites and (<b>f</b>) the proposed mechanism for the photocatalytic dye degradation activity of NiZG-70 nanocomposites. Reprinted from Refs. [<a href="#B174-ijms-24-15021" class="html-bibr">174</a>,<a href="#B177-ijms-24-15021" class="html-bibr">177</a>] with permission from Elsevier.</p>
Full article ">Figure 13
<p>(<b>a</b>) A low-scale HR-TEM image of pc-GCN/15-SiO<sub>2</sub>/5-ZnO, (<b>b</b>) schematic diagram of the proposed photocatalytic mechanism of MB degradation by the pcGCN/15-SiO<sub>2</sub>/5-ZnO nanocomposite under WLLI. (<b>c</b>) TEM image of CNZ/BCN (1:1) and (<b>d</b>) schematic illustration of the proposed mechanism for CIP degradation by CNZ/BCN (1:1): Z-scheme-mediated pathway. Reprinted from Refs. [<a href="#B185-ijms-24-15021" class="html-bibr">185</a>,<a href="#B190-ijms-24-15021" class="html-bibr">190</a>] with permission from Elsevier.</p>
Full article ">Figure 14
<p>TEM image of the sample CZg (CuO/ZnO/g-C<sub>3</sub>N<sub>4</sub>) heterostructure (<b>a</b>) and schematic diagram of the charge migration pathway in CuO/ZnO/g-C<sub>3</sub>N<sub>4</sub> heterostructure based on Z-scheme heterojunction approach (<b>b</b>). FESEM image of CZ@T-GCN (<b>c</b>) and schematic illustration for plausible charge relocation and photocatalytic mechanism of AMOX degradation over CZ@T-GCN (<b>d</b>). HRTEM image of CZN1 heterostructure (<b>e</b>), and schematic illustration of charge-transfer mechanism occurring through direct dual Z-scheme pathway in magnetic CN/ZnO/NiFe<sub>2</sub>O<sub>4</sub> system under visible-light irradiation (<b>f</b>). Reprinted from Refs. [<a href="#B191-ijms-24-15021" class="html-bibr">191</a>,<a href="#B194-ijms-24-15021" class="html-bibr">194</a>,<a href="#B195-ijms-24-15021" class="html-bibr">195</a>] with permission from Elsevier.</p>
Full article ">Figure 15
<p>(<b>a</b>) SEM image of 15% NZCN and (<b>b</b>) the possible mechanism for NOR degradation in the 15% NZCN/vis system. (<b>c</b>) Removal rate of BPA by FZCCN-4 photocatalysts over five cycles. Condition: C<sub>0</sub> (BPA) = 10 mg L<sup>−1</sup>; C<sub>0</sub> (catalyst) = 0.8 g L<sup>−1</sup>. (<b>d</b>) Proposed photocatalytic mechanism in the FZCCN-4 photocatalytic system under visible-light irradiation. (<b>e</b>) Work function of ZnO (101) and g-C<sub>3</sub>N<sub>4</sub> (001), (<b>f</b>) the internal electric field and band bending at the interface of ZnO@g-C<sub>3</sub>N<sub>4</sub> after contact, and (<b>g</b>) transfer process of photogenerated carriers in S-scheme heterojunctions. Reprinted from Refs. [<a href="#B198-ijms-24-15021" class="html-bibr">198</a>,<a href="#B201-ijms-24-15021" class="html-bibr">201</a>,<a href="#B205-ijms-24-15021" class="html-bibr">205</a>] with permission from Elsevier.</p>
Full article ">
20 pages, 4935 KiB  
Article
3β-Hydroxy-12-oleanen-27-oic Acid Exerts an Antiproliferative Effect on Human Colon Carcinoma HCT116 Cells via Targeting FDFT1
by Jue Tu, Xiang Meng, Juanjuan Wang, Ziyi Han, Zuoting Yu and Hongxiang Sun
Int. J. Mol. Sci. 2023, 24(19), 15020; https://doi.org/10.3390/ijms241915020 - 9 Oct 2023
Cited by 2 | Viewed by 1578
Abstract
3β-hydroxy-12-oleanen-27-oic acid (ATA), a cytotoxic oleanane triterpenoid with C14-COOH isolated from the rhizome of Astilbe chinensis, has been previously proven to possess antitumor activity and may be a promising antitumor agent. However, its molecular mechanisms of antitumor action were still [...] Read more.
3β-hydroxy-12-oleanen-27-oic acid (ATA), a cytotoxic oleanane triterpenoid with C14-COOH isolated from the rhizome of Astilbe chinensis, has been previously proven to possess antitumor activity and may be a promising antitumor agent. However, its molecular mechanisms of antitumor action were still unclear. This study explored the underlying mechanisms of cytotoxicity and potential target of ATA against human colorectal cancer HCT116 cells via integrative analysis of transcriptomics and network pharmacology in combination with in vitro and in vivo experimental validations. ATA significantly inhibited the proliferation of HCT116 cells in a concentration- and time-dependent manner and induced the cell cycle arrest at the G0/G1 phase, apoptosis, autophagy, and ferroptosis. Transcriptomic analysis manifested that ATA regulated mRNA expression of the genes related to cell proliferation, cell cycle, and cell death in HCT116 cells. The integrated analysis of transcriptomics, network pharmacology, and molecular docking revealed that ATA exerted cytotoxic activity via interactions with FDFT1, PPARA, and PPARG. Furthermore, FDFT1 was verified to be an upstream key target mediating the antiproliferative effect of ATA against HCT116 cells. Of note, ATA remarkably suppressed the growth of HCT116 xenografts in nude mice and displayed an apparent attenuation of FDFT1 in tumor tissues accompanied by the alteration of the biomarkers of autophagy, cell cycle, apoptosis, and ferroptosis. These results demonstrate that ATA exerted in vitro and in vivo antiproliferative effects against HCT116 cells through inducing cell apoptosis, autophagy, and ferroptosis via targeting FDFT1. Full article
(This article belongs to the Section Molecular Immunology)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of 3<span class="html-italic">β</span>-hydroxy-12-oleanen-27-oic acid (ATA), oleanolic acid (OA), and ursolic acid (UA).</p>
Full article ">Figure 2
<p>ATA exhibited cytotoxic effects towards colorectal cancer cells. (<b>A</b>) The IC<sub>50</sub> values of ATA towards seven colorectal cancer and human intestinal epithelial FHs74Int cells via the MTT assay. (<b>B</b>) The cell viabilities of HCT116 cells treated with ATA at the indicated concentrations for 24 h, 48 h, and 72 h via the MTT assay. (<b>C</b>) The morphological changes of ATA-treated HCT116 cells under a fluorescence microscope after incubation with the specific fluorescent dyes AO for apoptosis (up), MDC for autophagy (middle), and DCFH-DA for ROS (below). Scale bars: 20 μm. (<b>D</b>,<b>E</b>) Cell cycle distribution (<b>D</b>) and the percentage (<b>E</b>) in various phases of the HCT116 cells treated with ATA (10 μM and 20 μM) for 24 h via FCM. (<b>F</b>,<b>G</b>) FCM dot plot (<b>F</b>) and the apoptotic percentage (<b>G</b>) of HCT116 cells after treatment with ATA (20 μM and 30 μM) for 24 h. The figures tagged represent the cell numbers for immediate analysis. Data are expressed as means ± SD for three independent experiments. * <span class="html-italic">p &lt;</span> 0.05, ** <span class="html-italic">p &lt;</span> 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. 0 μg/mL.</p>
Full article ">Figure 3
<p>Workflow of the transcriptomic analysis of HCT116 cells treated with ATA.</p>
Full article ">Figure 4
<p>Gene expression profiles in HCT116 cells induced via ATA. (<b>A</b>) Principal component analysis (PCA) of the HCT116 cells treated without (Ctrl) or with ATA (20 μM). (<b>B</b>–<b>F</b>) Volcano plot showing mRNA expression profiles in the HCT116 cells treated without (Ctrl) or with ATA (20 μM) for 3 h (<b>B</b>), 6 h (<b>C</b>), 12 h (<b>D</b>), 18 h (<b>E</b>), and 24 h (<b>F</b>). Red points represent up-regulation, while green points indicate down-regulation; gray points represent normal expression. (<b>G</b>) RT-qPCR validation of 3 up- and 3 down-regulated DEGs.</p>
Full article ">Figure 5
<p>Function and hub genes of the DEGs in HCT116 cells induced via ATA. (<b>A</b>,<b>B</b>) GO functions (<b>A</b>) and KEGG pathways (<b>B</b>) of DEGs using Metascape. (<b>C</b>,<b>D</b>) Four densely connected modules (<b>C</b>) and functional annotation (<b>D</b>) of the DEGs using the MCODE plug-in of Cytoscape. MCODE 1 contains 29 nodes and 252 edges. MCODE 2 contains 43 nodes and 155 edges. MCODE 3 includes 35 nodes and 119 edges. MCODE 4 includes 46 nodes and 156 edges. Red nodes represent up-regulated genes; blue nodes represent down-regulated genes. The genes were laid out in circles according to the node size depending on the betweenness value, where edges indicate straight associations. (<b>E</b>) Upset plot of 11 hub genes using 8 Cytoscape algorithms.</p>
Full article ">Figure 6
<p>Identification of the antitumor targets of ATA. (<b>A</b>) Overlap of the predicted targets of ATA (blue) and disease targets from the transcriptomic data (pink). (<b>B</b>) Function analysis of fifteen common targets via GeneMANIA. (C) Binding pattern diagram of ATA and the target proteins FDFT1, PPARA, and PPARG. The yellow dashed line represents the hydrogen bond interaction. (<b>D</b>–<b>F</b>) After pre-incubation with or without YM-53601 (2.5 μM and 5 µM, (<b>D</b>)), GW6471 (2.5 μM and 5 µM, (<b>E</b>)), or GW9662 (5 μM and 10 µM, (<b>F</b>)) for 30 min, HCT116 cells were treated with ATA (0 μM, 10 μM, and 15 µM) for 48 h. The cell viabilities were detected via the MTT assay. The data are expressed as means ± SD (<span class="html-italic">n</span> = 3). ** <span class="html-italic">p</span> &lt; 0.01 and *** <span class="html-italic">p</span> &lt; 0.001 vs. control (Ctrl). (<b>G</b>) Venn diagram of predicted targets for ATA, OA, and UA. Five genes in red were identified as ATA-specific targets. (<b>H</b>) The cell viabilities of HCT116 cells treated with ATA, UA, and OA for 48 h were detected via the MTT assay and the IC<sub>50</sub> values were calculated using Graphpad Prism 9.0 software.</p>
Full article ">Figure 7
<p>FDFT1 mediated the cytotoxicity of ATA towards HCT116 cells. (<b>A</b>) The gene expression levels of FDFT1 in the HCT116 cells treated with ATA for different times via the RT-qPCR assay. (<b>B</b>,<b>C</b>) The protein expression levels of FDFT1 in the HCT116 cells treated with ATA (7.5, 15 and 30 μM) for 24 h via Western blotting. The figure (<b>B</b>) shown is representative of three independent experiments. The data (<b>C</b>) are expressed as means ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.01 vs. Ctrl (0 μM). (<b>D</b>–<b>F</b>) After pre-incubation with or without the FDFT1 inhibitor YM-53601 (2 and 4 µM) for 30 min, HCT116 cells were treated with ATA (0 µM, 5 µM, and 7.5 µM, D), OA (100 μM, E), or UA (40 μM, F) for 24 h. The cell viabilities were detected via the MTT assay. The data are expressed as means ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 vs. the control (Ctrl). (<b>G</b>–<b>H</b>) After pre-incubation with or without YM-53601 (4 µM, 30 min), the HCT116 cells were treated with ATA (20 μM) for 24 h. The autophagic flux was observed through the mRFP-GFP-LC3 assay. The figure (<b>G</b>) shown is representative of three independent experiments. The cell apoptosis was determined using FCM. The figure (<b>H</b>) is representative of three independent experiments. The apoptotic percentages (<b>I</b>) are expressed as means ± SD (<span class="html-italic">n</span> = 3). *** <span class="html-italic">p &lt;</span> 0.001 vs. Ctrl. (<b>J</b>–<b>L</b>) The mRNA (<b>J</b>) and protein (<b>K</b>,<b>L</b>) expression levels of FDFT1 in HCT116 cells were detected using RT-qPCR and Western blotting after transfection with FDFT1 siRNA for 24 h and 48 h, respectively. siNC—negative control siRNA. The figure (<b>K</b>) shown is representative of three independent experiments. The data (<b>J</b>,<b>L</b>) are expressed as means ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 vs. siNC. (<b>M</b>) The HCT116 cells were transfected with FDFT1 siRNAs for 48 h, followed by exposure to ATA at indicated concentrations for another 48 h. The cell viabilities were detected through the MTT assay. The data are expressed as means ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.001 vs. siNC.</p>
Full article ">Figure 8
<p>ATA inhibited the tumor growth of HCT116 xenografts in nude mice. (<b>A</b>) Tumor anatomy of HCT116 subcutaneous tumor-bearing mice. Scale bar = 10 mm. (<b>B</b>) Growth curve of tumor volume. (<b>C</b>) Average final tumor weights of each group. (<b>D</b>) Body weights of tumor-bearing mice. (<b>E</b>,<b>F</b>) Expression levels of FDFT1, LC3, p62, cyclinD1, pro-caspase3, and GPX4 in tumor tissues from Ctrl and ATA (60 mg/kg) groups via Western blotting. The data were expressed as means ± SD (<span class="html-italic">n</span> = 4). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl.</p>
Full article ">Figure 9
<p>Hypothetical pathway of FDFT1 in mediating the cytotoxic effect of ATA against HCT116 cells.</p>
Full article ">
19 pages, 673 KiB  
Review
Novel Approaches to Treatment of Acute Myeloid Leukemia Relapse Post Allogeneic Stem Cell Transplantation
by Carmine Liberatore and Mauro Di Ianni
Int. J. Mol. Sci. 2023, 24(19), 15019; https://doi.org/10.3390/ijms241915019 - 9 Oct 2023
Cited by 3 | Viewed by 2690
Abstract
The management of patients with acute myeloid leukemia (AML) relapsed post allogeneic hematopoietic stem cell transplantation (HSCT) remains a clinical challenge. Intensive treatment approaches are limited by severe toxicities in the early post-transplantation period. Therefore, hypomethylating agents (HMAs) have become the standard therapeutic [...] Read more.
The management of patients with acute myeloid leukemia (AML) relapsed post allogeneic hematopoietic stem cell transplantation (HSCT) remains a clinical challenge. Intensive treatment approaches are limited by severe toxicities in the early post-transplantation period. Therefore, hypomethylating agents (HMAs) have become the standard therapeutic approach due to favorable tolerability. Moreover, HMAs serve as a backbone for additional anti-leukemic agents. Despite discordant results, the addition of donor lymphocytes infusions (DLI) generally granted improved outcomes with manageable GvHD incidence. The recent introduction of novel targeted drugs in AML gives the opportunity to add a third element to salvage regimens. Those patients harboring targetable mutations might benefit from IDH1/2 inhibitors Ivosidenib and Enasidenib as well as FLT3 inhibitors Sorafenib and Gilteritinib in combination with HMA and DLI. Conversely, patients lacking targetable mutations actually benefit from the addition of Venetoclax. A second HSCT remains a valid option, especially for fit patients and for those who achieve a complete disease response with salvage regimens. Overall, across studies, higher response rates and longer survival were observed in cases of pre-emptive intervention for molecular relapse. Future perspectives currently rely on the development of adoptive immunotherapeutic strategies mainly represented by CAR-T cells. Full article
(This article belongs to the Special Issue Pathophysiology to Novel Therapeutic Approaches for Leukemia)
Show Figures

Figure 1

Figure 1
<p>Available therapies for the treatment of acute myeloid leukemia relapsing post allogeneic stem cell transplantation.</p>
Full article ">Figure 2
<p>Proposed treatment algorithm for AML patients relapsed after allogeneic hematopoietic stem cell transplant. AML: acute myeloid leukemia; HSCT: allogeneic hematopoietic stem cell transplantation; IC: intensive chemotherapy; DLI: donor lymphocytes infusion.</p>
Full article ">
17 pages, 3809 KiB  
Communication
The Effects of Combined Exposure to Bisphenols and Perfluoroalkyls on Human Perinatal Stem Cells and the Potential Implications for Health Outcomes
by Andrea Di Credico, Giulia Gaggi, Ines Bucci, Barbara Ghinassi and Angela Di Baldassarre
Int. J. Mol. Sci. 2023, 24(19), 15018; https://doi.org/10.3390/ijms241915018 - 9 Oct 2023
Cited by 3 | Viewed by 1317
Abstract
The present study investigates the impact of two endocrine disruptors, namely Bisphenols (BPs) and Perfluoroalkyls (PFs), on human stem cells. These chemicals leach from plastic, and when ingested through contaminated food and water, they interfere with endogenous hormone signaling, causing various diseases. While [...] Read more.
The present study investigates the impact of two endocrine disruptors, namely Bisphenols (BPs) and Perfluoroalkyls (PFs), on human stem cells. These chemicals leach from plastic, and when ingested through contaminated food and water, they interfere with endogenous hormone signaling, causing various diseases. While the ability of BPs and PFs to cross the placental barrier and accumulate in fetal serum has been documented, the exact consequences for human development require further elucidation. The present research work explored the effects of combined exposure to BPs (BPA or BPS) and PFs (PFOS and PFOA) on human placenta (fetal membrane mesenchymal stromal cells, hFM-MSCs) and amniotic fluid (hAFSCs)-derived stem cells. The effects of the xenobiotics were assessed by analyzing cell proliferation, mitochondrial functionality, and the expression of genes involved in pluripotency and epigenetic regulation, which are crucial for early human development. Our findings demonstrate that antenatal exposure to BPs and/or PFs may alter the biological characteristics of perinatal stem cells and fetal epigenome, with potential implications for health outcomes at birth and in adulthood. Further research is necessary to comprehend the full extent of these effects and their long-term consequences. Full article
(This article belongs to the Special Issue Endocrine Disruptors Exposure and Human Health)
Show Figures

Figure 1

Figure 1
<p>Analysis of PB and PF effects on hFM-MSC proliferation by impedance and Ki67 measurements. On day, 0 hFM-MSCs (displayed in the box, original magnification 20×, and scale bar 50 μm) were treated with 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, and the impedance values were monitored in real-time up to 48 h. Cells not treated (brown line), treated only with the vehicle (dark blue line), or with lysing agents (purple line) represented the experimental controls. (<b>a</b>) Left panel, absolute impedance values (expressed in ohms, Ω). Graph is representative of three different experiments. Right panel, normalized impedance values (Day 0 = 1) of control (brown line) or samples treated with 0.1 μM of BPA, BPS, PFOS, or PFOA alone or in combination, as indicated. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). (<b>b</b>) Immunocytochemical detection of Ki67 (green fluorescence) in hFM-MSC control cells and after 24 h exposure to vehicle or to 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, as reported. The nuclei were counterstained with DAPI (blue). Original magnification: 40×, scale bar 50 μm. Images are representative of 3 independent experiments. Histogram indicates the % of the positive nuclei in the different experimental conditions. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, BPS + PFOA (ALL).</p>
Full article ">Figure 1 Cont.
<p>Analysis of PB and PF effects on hFM-MSC proliferation by impedance and Ki67 measurements. On day, 0 hFM-MSCs (displayed in the box, original magnification 20×, and scale bar 50 μm) were treated with 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, and the impedance values were monitored in real-time up to 48 h. Cells not treated (brown line), treated only with the vehicle (dark blue line), or with lysing agents (purple line) represented the experimental controls. (<b>a</b>) Left panel, absolute impedance values (expressed in ohms, Ω). Graph is representative of three different experiments. Right panel, normalized impedance values (Day 0 = 1) of control (brown line) or samples treated with 0.1 μM of BPA, BPS, PFOS, or PFOA alone or in combination, as indicated. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). (<b>b</b>) Immunocytochemical detection of Ki67 (green fluorescence) in hFM-MSC control cells and after 24 h exposure to vehicle or to 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, as reported. The nuclei were counterstained with DAPI (blue). Original magnification: 40×, scale bar 50 μm. Images are representative of 3 independent experiments. Histogram indicates the % of the positive nuclei in the different experimental conditions. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, BPS + PFOA (ALL).</p>
Full article ">Figure 2
<p>Effects of exposure to BPs and/or PFs on hFM-MSC mitochondrial health. Immunocytochemical detection of MMP (red fluorescence) in control cells and after 24 h exposure to vehicle only or to 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, as indicated. The nuclei were counterstained with DAPI (blue). Original magnification: 20×, scale bar 100 μm. Images are representative of 3 independent experiments. Histogram indicates the fluorescent intensity (absolute values) in the different experimental conditions. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, BPS + PFOA (ALL).</p>
Full article ">Figure 3
<p>Pluripotency marker and epigenetic regulatory expression in hFM-MSCs upon BPs and/or PFs treatment. The gene expression of the stemness identifiers (NANOG, OCT4, SOX2, KLF4, ESG1, and OVOL1) and epigenetic regulators (DNMT1, DNTM3A, TET1, TET2, and TET3) were detected by qPCR in hFM-MSCs treated with BPs and/or PFs for 24 h, as indicated. The fold changes were determined from the −ΔΔCt values calculated using 18S as a reference gene and normalized to untreated hFM-MSCs as control condition (CTRL). The graphs show the mean ± SD of 3 independent experiments, * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL. Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, BPS + PFOA (ALL).</p>
Full article ">Figure 4
<p>Analysis of PB and PF effects on hAFSC proliferation by impedance and Ki67 measurements. On day 0, hAFSCs (displayed in the box, original magnification 20×, and scale bar 50 μm) were treated with 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, and the impedance values were monitored in real time up to 48 h. Cells not treated (brown line), treated only with the vehicle (dark blue line), or with lysing agents (purple line) represented the experimental controls. (<b>a</b>) Left panel, absolute impedance values (expressed in ohms, Ω). Graph is representative of three different experiments. Right panel, normalized impedance values (Day 0 = 1) of control (brown line) or samples treated with 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, as indicated. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). (<b>b</b>) Immunocytochemical detection of Ki67 (green fluorescence) in hAFSC control cells and after 24 h exposure to vehicle or to 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, as reported. The nuclei were counterstained with DAPI (blue). Original magnification: 40×, scale bar 50 μm. Images are representative of 3 independent experiments. Histogram indicates the % of the positive nuclei in the different experimental conditions. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, BPS + PFOA (ALL).</p>
Full article ">Figure 5
<p>Effects of exposure to BPs and/or PFs on hAFSC mitochondrial health. Immunocytochemical detection of MMP (red fluorescence) in control cells and after 24 h exposure to vehicle only or to 0.1 μM of BPA, BPS, PFOS, or PFOA, alone or in combination, as indicated. The nuclei were counterstained with DAPI (blue). Original magnification: 20×, scale bar 100 μm. Images are representative of 3 independent experiments. Histogram indicates the fluorescent intensity (absolute values) in the different experimental conditions. Data are expressed as mean ± SD (<span class="html-italic">n</span> = 3, * <span class="html-italic">p</span> &lt; 0.05 vs. control). Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, BPS + PFOA (ALL).</p>
Full article ">Figure 6
<p>Pluripotency markers and epigenetic regulatory enzymes expression in hAFSCs upon BPs and/or PFs treatment. The gene expression of the stemness identifiers (NANOG, OCT4, SOX2, KLF4, ESG1, and OVOL1), and epigenetic regulators (DNMT1, DNTM3A, TET1, TET2, and TET3) were detected by qPCR in hAFSCs treated with BPs and/or PFs for 24 h, as indicated. The fold changes were determined from the −ΔΔCt values calculated using 18S as a reference gene and normalized to untreated hAFSCs as control condition (CTRL). The graphs show the mean ± SD of 3 independent experiments, * <span class="html-italic">p</span> &lt; 0.05 vs. CTRL. Bisphenol A (BPA), Bisphenol S (BPS), perfluorooctane sulfonate (PFOS), perfluorooctanoic acid (PFOA), BPA + PFOS, BPS + PFOA, BPS + PFOS, and BPS + PFOA (ALL).</p>
Full article ">
11 pages, 3887 KiB  
Article
Enhancing the Heterologous Expression of a Thermophilic Endoglucanase and Its Cost-Effective Production in Pichia pastoris Using Multiple Strategies
by Wuling Dai, Haofan Dong, Zhaokun Zhang, Xin Wu, Tongtong Bao, Le Gao and Xiaoyi Chen
Int. J. Mol. Sci. 2023, 24(19), 15017; https://doi.org/10.3390/ijms241915017 - 9 Oct 2023
Cited by 1 | Viewed by 1525
Abstract
Although Pichia pastoris was successfully used for heterologous gene expression for more than twenty years, many factors influencing protein expression remain unclear. Here, we optimized the expression of a thermophilic endoglucanase from Thermothielavioides terrestris (TtCel45A) for cost-effective production in Pichia pastoris. To [...] Read more.
Although Pichia pastoris was successfully used for heterologous gene expression for more than twenty years, many factors influencing protein expression remain unclear. Here, we optimized the expression of a thermophilic endoglucanase from Thermothielavioides terrestris (TtCel45A) for cost-effective production in Pichia pastoris. To achieve this, we established a multifactorial regulation strategy that involved selecting a genome-editing system, utilizing neutral loci, incorporating multiple copies of the heterologous expression cassette, and optimizing high-density fermentation for the co-production of single-cell protein (SCP). Notably, even though all neutral sites were used, there was still a slight difference in the enzymatic activity of heterologously expressed TtCel45A. Interestingly, the optimal gene copy number for the chromosomal expression of TtCel45A was found to be three, indicating limitations in translational capacity, post-translational processing, and secretion, ultimately impacting protein yields in P. pastoris. We suggest that multiple parameters might influence a kinetic competition between protein elongation and mRNA degradation. During high-density fermentation, the highest protein concentration and endoglucanase activity of TtCel45A with three copies reached 15.8 g/L and 9640 IU/mL, respectively. At the same time, the remaining SCP of P. pastoris exhibited a crude protein and amino acid content of up to 59.32% and 46.98%, respectively. These findings suggested that SCP from P. pastoris holds great promise as a sustainable and cost-effective alternative for meeting the global protein demand, while also enabling the production of thermophilic TtCel45A in a single industrial process. Full article
(This article belongs to the Special Issue Microbial Enzymes for Biotechnological Applications)
Show Figures

Figure 1

Figure 1
<p>Overall structure of TtCel45A from <span class="html-italic">Thermothielavioides terrestris.</span> (<b>a</b>) The three-dimensional structure of Cel45A from <span class="html-italic">Thermothielavioides terrestris.</span> (<b>b</b>) Six disulfide bonds in the homology model of TtCel45a.</p>
Full article ">Figure 2
<p>Percent of cells successfully transformed with the GFP expression cassette using (<b>a</b>) CRISPR-Cas9 and (<b>b</b>) homologous recombination, as detected via flow cytometry.</p>
Full article ">Figure 3
<p>Comparison of TtCel45A expression in transformants with different copy numbers of the genomically integrated expression cassette. (<b>a</b>) Comparison of endoglucanase activity of transformants with different copy numbers. “*” for <span class="html-italic">p</span> &lt; 0.05, “**” for <span class="html-italic">p</span> &lt; 0.001, and “ns” for no significance. (<b>b</b>) SDS-PAGE of TtCel45A expression in transformants with different copy numbers. Red marker for 72KDa.</p>
Full article ">Figure 4
<p>Analysis of fed-batch fermentation and nutritional content of <span class="html-italic">P. pastoris</span> biomass after fermentation. (<b>a</b>) Fed-batch fermentation of <span class="html-italic">P. pastoris</span> in a 5 L bioreactor. (<b>b</b>) Analysis of endoglucanase activity during fermentation. (<b>c</b>) Amino acid profile of single-cell protein from <span class="html-italic">P. pastoris</span> co-produced in fermentation. (<b>d</b>) High-cell-density fermentation of TtCel45A in a 5 L bioreactor.</p>
Full article ">
19 pages, 3403 KiB  
Review
Autophagy as a Target for Non-Immune Intrinsic Functions of Programmed Cell Death-Ligand 1 in Cancer
by Blanca Estela García-Pérez, Christian Pérez-Torres, Shantal Lizbeth Baltierra-Uribe, Juan Castillo-Cruz and Nayeli Shantal Castrejón-Jiménez
Int. J. Mol. Sci. 2023, 24(19), 15016; https://doi.org/10.3390/ijms241915016 - 9 Oct 2023
Cited by 5 | Viewed by 1818
Abstract
Autophagy is a catabolic process that is essential to the maintenance of homeostasis through the cellular recycling of damaged organelles or misfolded proteins, which sustains energy balance. Additionally, autophagy plays a dual role in modulating the development and progression of cancer and inducing [...] Read more.
Autophagy is a catabolic process that is essential to the maintenance of homeostasis through the cellular recycling of damaged organelles or misfolded proteins, which sustains energy balance. Additionally, autophagy plays a dual role in modulating the development and progression of cancer and inducing a survival strategy in tumoral cells. Programmed cell death-ligand 1 (PD-L1) modulates the immune response and is responsible for maintaining self-tolerance. Because tumor cells exploit the PD-L1–PD-1 interaction to subvert the immune response, immunotherapy has been developed based on the use of PD-L1-blocking antibodies. Recent evidence has suggested a bidirectional regulation between autophagy and PD-L1 molecule expression in tumor cells. Moreover, the research into the intrinsic properties of PD-L1 has highlighted new functions that are advantageous to tumor cells. The relationship between autophagy and PD-L1 is complex and still not fully understood; its effects can be context-dependent and might differ between tumoral cells. This review refines our understanding of the non-immune intrinsic functions of PD-L1 and its potential influence on autophagy, how these could allow the survival of tumor cells, and what this means for the efficacy of anti-PD-L1 therapeutic strategies. Full article
(This article belongs to the Special Issue Recent Research on Autophagy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of PD-L1. PD-L1 is a transmembrane glycoprotein comprising 290 amino acids, grouped into two domains (IgV and IgC), followed by a short flexible stem. It has a hydrophobic transmembrane (TM) domain, as well as a small intracellular (IC) domain of 30 amino acids. The 3D image shown is from the Protein Data Bank and is available at <a href="https://www.rcsb.org/structure/3bis" target="_blank">https://www.rcsb.org/structure/3bis</a> (accessed on 21 September 2023).</p>
Full article ">Figure 2
<p>The main mechanisms that modulate PD-L1 expression in cancer cells: In the tumor microenvironment, several signals can induce PD-L1 overexpression through the activation of downstream effectors. (1) Hypoxia-inducible factor 1α (HIF-1α), which can also be induced by other stimuli, such as growth factors, has been associated with the expression of PD-L1. (2) Some chemotherapeutic agents, such as paclitaxel or cisplatin, through the phosphorylation of ERK ½, induce the expression of PD-L1; however, the complete mechanisms behind this event remain unknown. (3) The activation of mTOR through various signals, such as growth factors, triggers signaling cascades that induce the expression of PD-L1, which is decreased when the mTOR inhibitor rapamycin is used. (4) Cytokines, such as IFN-γ, induce PD-L1 expression through the JAK/STAT pathway. (5) The stimulation of some TLRs induces the activation of canonical pathways to regulate PD-L1 overexpression. Green arrows mean activation vias, red block lines mean inhibition pathways, dotted arrows mean nuclear translocation, black arrow means decreasing.</p>
Full article ">Figure 3
<p>A schematic representation of the bidirectional autophagy/PD-L1 relationship: The <b>right</b> side shows the main findings related to the effects of the activation (<b>A</b>) or inhibition (<b>B</b>) of autophagy on PD-L1 expression. The <b>left</b> side represents the effects of the intrinsic non-immune functions of PD-L1 and the known downstream molecules that culminate in the disruption (<b>C</b>) or activation (<b>D</b>) of autophagy. Green arrows mean activation vias, red block lines mean inhibition pathways, red dotted line means hypothetical pathway, black arrows aiming up mean increasing, black arrows aiming down mean decreasing.</p>
Full article ">
11 pages, 2454 KiB  
Article
Proenkephalin Levels and Its Determinants in Patients with End-Stage Kidney Disease Treated with Hemodialysis and Peritoneal Dialysis
by Wiktoria Grycuk, Zuzanna Jakubowska and Jolanta Małyszko
Int. J. Mol. Sci. 2023, 24(19), 15015; https://doi.org/10.3390/ijms241915015 - 9 Oct 2023
Cited by 1 | Viewed by 1027
Abstract
Recently, proenkephalin A (PENK A) has been shown to reflect glomerular dysfunction and to predict new-onset acute kidney injury and heart failure. While previous studies have investigated PENK A as a biomarker in individuals with preserved renal function, PENK A concentration in patients [...] Read more.
Recently, proenkephalin A (PENK A) has been shown to reflect glomerular dysfunction and to predict new-onset acute kidney injury and heart failure. While previous studies have investigated PENK A as a biomarker in individuals with preserved renal function, PENK A concentration in patients with end-stage kidney disease (ESKD) was not investigated. Plasma PENK A concentration was assessed in 88 patients with ESKD treated with hemodialysis (HD) or peritoneal dialysis (PD), and its associations with kidney function and heart failure indicators were investigated. In HD patients, the difference in PENK A levels before and after hemodialysis, was measured and further assessed for an association with the type of HD membrane used. PENK A levels did not differ significantly between HD and PD patients. In HD patients, the median PENK A concentration was significantly higher before than after hemodialysis (1.368 vs. 2.061, p = 0.003). No correlation was found between PENK A level and urea (p = 0.192), eGFR (p = 0.922), dialysis vintage (p = 0.637), and residual urine output (p = 0.784). Heart failure (p = 0.961), EF (p = 0.361), and NT-proBNP (p = 0.949) were not associated with increased PENK A concentration. PENK A does not reflect renal function and cardiac status in patients with ESKD. Further research is required to establish the clinical utility of the new biomarker in patients with impaired kidney function. Full article
(This article belongs to the Special Issue Renal Dysfunction, Uremic Compounds, and Other Factors 2.0)
Show Figures

Figure 1

Figure 1
<p>Graphs showing the relationship between PENK A levels and kidney function indicators. No correlation was found between proenkephalin A and (<b>a</b>) eGFR (<span class="html-italic">p</span> = 0.922) and (<b>b</b>) urea (<span class="html-italic">p</span> = 0.192); (<b>c</b>) there were no significant differences in PENK A levels depending on the residual urine output (<span class="html-italic">p</span> = 0.784) and (<b>d</b>) dialysis vintage (<span class="html-italic">p</span> = 0.067).</p>
Full article ">Figure 1 Cont.
<p>Graphs showing the relationship between PENK A levels and kidney function indicators. No correlation was found between proenkephalin A and (<b>a</b>) eGFR (<span class="html-italic">p</span> = 0.922) and (<b>b</b>) urea (<span class="html-italic">p</span> = 0.192); (<b>c</b>) there were no significant differences in PENK A levels depending on the residual urine output (<span class="html-italic">p</span> = 0.784) and (<b>d</b>) dialysis vintage (<span class="html-italic">p</span> = 0.067).</p>
Full article ">Figure 2
<p>(<b>a</b>) No correlation was found between PENK A level and NT–proBNP. The comparison between groups showed no significant differences in PENK A levels depending on (<b>b</b>) the range of EF (<span class="html-italic">p</span> = 0.361) or (<b>c</b>) presence of HF (<span class="html-italic">p</span> = 0.961).</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) No correlation was found between PENK A level and NT–proBNP. The comparison between groups showed no significant differences in PENK A levels depending on (<b>b</b>) the range of EF (<span class="html-italic">p</span> = 0.361) or (<b>c</b>) presence of HF (<span class="html-italic">p</span> = 0.961).</p>
Full article ">Figure 3
<p>Figure presenting PENK A levels in patients treated with high-flux (boxplot on the left) and low-flux (boxplot on the right) hemodialysis membranes. PENK A concentration was significantly higher in patients using low-flux membrane vs. patients using high-flux membranes, both before hemodialysis (<span class="html-italic">p</span> = 0.003) and afterwards (<span class="html-italic">p</span> = 0.003): (<b>a</b>) PENK A levels in samples drawn directly before hemodialysis; (<b>b</b>) PENK A levels after hemodialysis.</p>
Full article ">Figure 4
<p>Figure presenting the distribution of residual urine output (mL/24 h) in patients treated with high-flux (green) and low-flux (blue) hemodialysis. No significant differences in residual diuresis were observed between the high-flux and low-flux groups (<span class="html-italic">p</span> = 0.793).</p>
Full article ">
13 pages, 2944 KiB  
Article
Integrated Multi-Omics Analyses Reveal That Autophagy-Mediated Cellular Metabolism Is Required for the Initiation of Pollen Germination
by Xuemei Zhou, Qiuyu Zhang, Yuliang Zhao, Shanshan Ding and Guang-Hui Yu
Int. J. Mol. Sci. 2023, 24(19), 15014; https://doi.org/10.3390/ijms241915014 - 9 Oct 2023
Viewed by 1042
Abstract
Autophagy is an evolutionarily conserved mechanism for degrading and recycling various cellular components, functioning in both normal development and stress conditions. This process is tightly regulated by a set of autophagy-related (ATG) proteins, including ATG2 in the ATG9 cycling system and ATG5 in [...] Read more.
Autophagy is an evolutionarily conserved mechanism for degrading and recycling various cellular components, functioning in both normal development and stress conditions. This process is tightly regulated by a set of autophagy-related (ATG) proteins, including ATG2 in the ATG9 cycling system and ATG5 in the ATG12 conjugation system. Our recent research demonstrated that autophagy-mediated compartmental cytoplasmic deletion is essential for pollen germination. However, the precise mechanisms through which autophagy regulates pollen germination, ensuring its fertility, remain largely unknown. Here, we applied multi-omics analyses, including transcriptomic and metabolomic approaches, to investigate the downstream pathways of autophagy in the process of pollen germination. Although ATG2 and ATG5 play similar roles in regulating pollen germination, high-throughput transcriptomic analysis reveals that silencing ATG5 has a greater impact on the transcriptome than silencing ATG2. Cross-comparisons of transcriptome and proteome analysis reveal that gene expression at the mRNA level and protein level is differentially affected by autophagy. Furthermore, high-throughput metabolomics analysis demonstrates that pathways related to amino acid metabolism and aminoacyl-tRNA biosynthesis were affected by both ATG2 and ATG5 silencing. Collectively, our multi-omics analyses reveal the central role of autophagy in cellular metabolism, which is critical for initiating pollen germination and ensuring pollen fertility. Full article
(This article belongs to the Special Issue Recent Research on Autophagy)
Show Figures

Figure 1

Figure 1
<p>Global transcriptome analysis of WT, <span class="html-italic">ATG2</span>-silenced, and <span class="html-italic">ATG5</span>-silenced pollen (<b>A</b>). Overlap analysis of detected genes in three independent biological replicates. (<b>B</b>). Principal component analysis of the transcriptomes of WT, <span class="html-italic">ATG2</span>-silenced, and <span class="html-italic">ATG5</span>-silenced pollen. (<b>C</b>). Clustering analysis the transcriptome of WT, <span class="html-italic">ATG2</span>-silenced, and <span class="html-italic">ATG5</span>-silenced pollen according to the Pearson’s correlation efficient.</p>
Full article ">Figure 2
<p>Identification of differentially expressed genes between WT and <span class="html-italic">ATG</span>-silenced pollen. (<b>A</b>). Volcano plots displaying differentially expressed genes between <span class="html-italic">ATG2</span>-silenced and WT pollen or <span class="html-italic">ATG5</span>-silenced and WT pollen. Red dots indicate upregulated genes, and blue dots indicate downregulated genes. (<b>B</b>). FPKM values of <span class="html-italic">ATG2</span> and <span class="html-italic">ATG5</span> in WT and their respective RNAi lines. (<b>C</b>). Number of differentially expressed genes between <span class="html-italic">ATG2</span>-silenced and WT pollen or <span class="html-italic">ATG5</span>-silenced and WT pollen. Fold changes are indicated with the distinct colors. Student’s t test, ** <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Identification of downstream pathways of autophagy in the process of pollen germination (<b>A</b>,<b>B</b>). Overlap analysis of upregulated genes (<b>A</b>) and downregulated genes (<b>B</b>) between <span class="html-italic">ATG2</span>/WT and <span class="html-italic">ATG5</span>/WT. (<b>C</b>,<b>D</b>). Heat maps showing the expression levels of the commonly downregulated (<b>C</b>) and commonly upregulated genes (<b>D</b>) in <span class="html-italic">ATG2</span>-silenced and <span class="html-italic">ATG5</span>-silenced pollen. (<b>E</b>,<b>F</b>). GO enrichment analysis of the commonly downregulated (<b>E</b>) and upregulated genes (<b>F</b>) in <span class="html-italic">ATG2</span>-silenced and <span class="html-italic">ATG5</span>-silenced pollen.</p>
Full article ">Figure 4
<p>Comparisons between differentially regulated transcripts and proteins in <span class="html-italic">ATG2</span>-silenced and <span class="html-italic">ATG5</span>-silenced pollen (<b>A</b>,<b>B</b>). Overlap analysis of differentially regulated transcripts and proteins in <span class="html-italic">ATG2</span>-silenced (<b>A</b>) and <span class="html-italic">ATG5</span>-silenced (<b>B</b>) pollen. (<b>C</b>,<b>D</b>). Heat maps showing differentially expressed genes that consistently change at mRNA and protein levels in <span class="html-italic">ATG2</span>-silenced (<b>C</b>) and <span class="html-italic">ATG5</span>-silenced (<b>D</b>) pollen. (<b>E</b>,<b>F</b>). Heat maps showing genes that are differentially regulated at both mRNA and protein levels in <span class="html-italic">ATG2</span>-silenced (<b>E</b>) and <span class="html-italic">ATG5</span>-silenced (<b>F</b>) pollen.</p>
Full article ">Figure 5
<p>Silencing <span class="html-italic">ATGs</span> results in the decreased metabolite levels in pollen (<b>A</b>). OPLS-DA of metabolite levels. The ellipse represents the Hotelling T2 with 95% confidence. (<b>B</b>,<b>C</b>). Volcano plots showing metabolites with different abundances between WT and <span class="html-italic">ATG</span>-silenced pollen. Blue dots indicate downregulated metabolites in <span class="html-italic">ATG</span>-silenced pollen, while pink dots indicate upregulated metabolites in <span class="html-italic">ATG</span>-silenced pollen. (<span class="html-italic">t</span>-test, <span class="html-italic">p</span> &lt; 0.05). (<b>D</b>). Venn diagrams showing the overlaps of metabolites with significantly different contents in <span class="html-italic">ATG</span>-silenced pollen.</p>
Full article ">Figure 6
<p>Significant impact of <span class="html-italic">ATG</span> downregulation on downstream metabolic pathways (<b>A</b>). Heat map depicting the relative metabolite contents. (<b>B</b>,<b>C</b>). Metabolic pathways regulated by <span class="html-italic">ATG2</span> (<b>B</b>) or <span class="html-italic">ATG5</span> (<b>C</b>).</p>
Full article ">
26 pages, 5591 KiB  
Article
Combined Omics Approaches Reveal Distinct Mechanisms of Resistance and/or Susceptibility in Sugar Beet Double Haploid Genotypes at Early Stages of Beet Curly Top Virus Infection
by Paul J. Galewski, Rajtilak Majumdar, Matthew D. Lebar, Carl A. Strausbaugh and Imad A. Eujayl
Int. J. Mol. Sci. 2023, 24(19), 15013; https://doi.org/10.3390/ijms241915013 - 9 Oct 2023
Viewed by 1620
Abstract
Sugar beet is susceptible to Beet curly top virus (BCTV), which significantly reduces yield and sugar production in the semi-arid growing regions worldwide. Sources of genetic resistance to BCTV is limited and control depends upon insecticide seed treatments with neonicotinoids. Through double haploid [...] Read more.
Sugar beet is susceptible to Beet curly top virus (BCTV), which significantly reduces yield and sugar production in the semi-arid growing regions worldwide. Sources of genetic resistance to BCTV is limited and control depends upon insecticide seed treatments with neonicotinoids. Through double haploid production and genetic selection, BCTV resistant breeding lines have been developed. Using BCTV resistant (R) [KDH13; Line 13 and KDH4-9; Line 4] and susceptible (S) [KDH19-17; Line 19] lines, beet leafhopper mediated natural infection, mRNA/sRNA sequencing, and metabolite analyses, potential mechanisms of resistance against the virus and vector were identified. At early infection stages (2- and 6-days post inoculation), examples of differentially expressed genes highly up-regulated in the ‘R’ lines (vs. ‘S’) included EL10Ac5g10437 (inhibitor of trypsin and hageman factor), EL10Ac6g14635 (jasmonate-induced protein), EL10Ac3g06016 (ribosome related), EL10Ac2g02812 (probable prolyl 4-hydroxylase 10), etc. Pathway enrichment analysis showed differentially expressed genes were predominantly involved with peroxisome, amino acids metabolism, fatty acid degradation, amino/nucleotide sugar metabolism, etc. Metabolite analysis revealed significantly higher amounts of specific isoflavonoid O-glycosides, flavonoid 8-C glycosides, triterpenoid, and iridoid-O-glycosides in the leaves of the ‘R’ lines (vs. ‘S’). These data suggest that a combination of transcriptional regulation and production of putative antiviral metabolites might contribute to BCTV resistance. In addition, genome divergence among BCTV strains differentially affects the production of small non-coding RNAs (sncRNAs) and small peptides which may potentially affect pathogenicity and disease symptom development. Full article
(This article belongs to the Special Issue Advances and New Perspectives in Plant-Microbe Interactions 2.0)
Show Figures

Figure 1

Figure 1
<p>Heatmaps showing a subset of differentially expressed (DE) transcripts (<span class="html-italic">p</span> &lt; 0.01) at (<b>A</b>) 2 days post inoculation (dpi), and (<b>B</b>) 6 dpi in the leaves of Beet curly top virus susceptible (Line 19; S) and resistant (Line 13 and Line 4; R) sugar beet lines. Data are Mean ± SE of 4–5 biological replicates and <span class="html-italic">p</span> &lt; 0.01 between ‘S’ and ‘R’ lines; I = infected].</p>
Full article ">Figure 2
<p>Gene ontology (GO) of differentially expressed sugar beet genes in the leaves of Beet curly top virus infected sugar beet plants. (<b>A</b>) 2 dpi and (<b>B</b>) 6 dpi. Data are mean of 3 biological replicates.</p>
Full article ">Figure 3
<p>Pathway enrichment of differentially expressed sugar beet genes in the leaves of Beet curly top virus infected plants. Kyoto Encyclopedia of Genes and Genomes (KEGG) enrichment of sugar beet genes in (<b>A</b>) 2 dpi and (<b>B</b>) 6 dpi. Data are mean of 4–5 biological replicates.</p>
Full article ">Figure 4
<p>Weighted gene co-expression network analysis (WGCNA) of sugar beet genes in the leaves infected with or without Beet curly top virus (BCTV) show distinct clustering pattern in the BCTV susceptible and resistant sugar beet lines. (<b>A</b>) Gene cluster dendrogram, (<b>B</b>) Heatmap showing network of differentially expressed genes, and (<b>C</b>) module-sample relationship. Data are Mean of 4–5 biological replicates.</p>
Full article ">Figure 5
<p>Heatmap of metabolites in the leaves of BCTV susceptible (Line 19; S) and resistant (Line 13 and Line 4; R) sugar beet lines at 2 d and 6 d in control (C; uninfected) and infected (I) samples. Data are Mean ± SE of 4–6 biological replicates.</p>
Full article ">Figure 6
<p>Specific metabolites highly abundant in the leaves of Beet curly top virus (BCTV) susceptible [Line 19; S (represented in blue color)] and resistant [Line 13 (represented in green color) and Line 4 (represented in pink color); R] sugar beet lines at 2 d and 6 d in control (C; uninfected) and infected (I) samples. (<b>A</b>) vitexin 2″-glucoside, (<b>B</b>) 9-methoxy-7-[4-[3 4 5-trihydroxy-6-[[3 4 5-trihydroxy-6-(hydroxymethyl)oxan-2-yl]oxymethyl]oxan-2-yl]oxyphenyl]-[1 3]dioxolo [4 5-g]chromen-8-one, (<b>C</b>) 9-methoxy-7-[4-[(2S 3R 4R 5S 6R)-3 4 5-trihydroxy-6-(hydroxymethyl)oxan-2-yl]oxyphenyl]-[1 3]dioxolo [4 5-g]chromen-8-one, (<b>D</b>) Agnuside, (<b>E</b>) 2-Phenylethyl 2-O-[(2S 3R 4R)-3 4-dihydroxy-4-(hydroxymethyl)tetrahydro-2-furanyl]-beta-D-glucopyranoside, and (<b>F</b>) ursolic acid. Data are Mean ± SE of 4–6 biological replicates; * <span class="html-italic">p</span> &lt; 0.05 between ‘S’ and ‘R’ lines at a specific time point within uninfected control (C) or infected (I) samples.</p>
Full article ">Figure 7
<p>Beet curly top virus (BCTV) strain specific distribution and relative abundance of virus-derived small non-coding RNAs during sugar beet infection. The solid gray blocks and colored blocks denote main open reading frames (ORFs) and predicted small ORFs, respectively. The outer track, which is gray, denotes viral derived small non-coding RNAs, while the colored portion represents small non-coding RNA core sequences showing sequence complementarity to putative sugar beet target genes [V3: movement protein; V2: SS-DS DNA regulator; V1: capsid protein; C3: replication enhancer; C2: pathogenesis enhancement protein; C1: rolling circle replication initiator protein; C4: cell cycle regulator].</p>
Full article ">Figure 8
<p>Divergence with respect to functional elements in the genomes of four Beet curly top virus (BCTV) strains. Higher values of FST at specific locations in the genomes are associated with greater divergence as evident by the peaks [small ORFs: small Open Reading Frames; small non-coding RNA: sncRNA; V3: movement protein; V2: SS-DS DNA regulator; V1: capsid protein; C3: replication enhancer; C2: pathogenesis enhancement protein; C1: rolling circle replication initiator protein; C4: cell cycle regulator].</p>
Full article ">
16 pages, 1500 KiB  
Article
Synthesis and Antiviral and Antitumor Activities of Novel 18β-Glycyrrhetinic Acid Derivatives
by Bo-Wen Pan, Liang-Liang Zheng, Yang Shi, Zhang-Chao Dong, Ting-Ting Feng, Jian Yang, Ying Wei and Ying Zhou
Int. J. Mol. Sci. 2023, 24(19), 15012; https://doi.org/10.3390/ijms241915012 - 9 Oct 2023
Cited by 1 | Viewed by 1293
Abstract
A series of novel derivatives of 18β-glycyrrhetinic acid (GA) were synthesized by introducing aromatic or heterocyclic structures to extend the side chain, thereby enhancing their interaction with amino acid residues in the active pocket of the target protein. These compounds were [...] Read more.
A series of novel derivatives of 18β-glycyrrhetinic acid (GA) were synthesized by introducing aromatic or heterocyclic structures to extend the side chain, thereby enhancing their interaction with amino acid residues in the active pocket of the target protein. These compounds were structurally characterized using 1H NMR, 13C NMR, and HRMS. The compounds were subsequently evaluated for their inhibitory effects on HIV-1 protease and cell viability in the human cancer cell lines K562 and HeLa and the mouse cancer cell line CT26. Towards HIV-1 protease, compounds 28 and 32, which featured the introduction of heterocyclic moieties at the C3 position of GA, exhibited the highest inhibition, with inhibition rates of 76% and 70.5%, respectively, at 1 mg/mL concentration. Further molecular docking suggests that a 3-substituted polar moiety would be likely to enhance the inhibitory activity against HIV-1 protease. As for the anti-proliferative activities of the GA derivatives, incorporation of a thiazole heterocycle at the C3- position in compound 29 significantly enhanced the effect against K562 cells with an IC50 value of 8.86 ± 0.93 µM. The introduction of electron-withdrawing substituents on the C3-substituted phenyl ring augmented the anti-proliferative activity against Hela and CT26 cells. Compound 13 exhibited the highest inhibitory activity against Hela cells with an IC50 value of 9.89 ± 0.86 µM, whereas compound 7 exerted the strongest inhibition against CT26 cells with an IC50 value of 4.54 ± 0.37 µM. These findings suggest that further modification of GA is a promising path for developing potent novel anti-HIV and anticancer therapeutics. Full article
(This article belongs to the Special Issue Antiviral Drug Targets: Structure, Function, and Drug Design 2.0)
Show Figures

Figure 1

Figure 1
<p>Structure of compound <b>12</b> determined by X-ray crystallography.</p>
Full article ">Figure 2
<p>Putative binding of compound <b>28</b> (<b>a</b>) and compound <b>32</b> (<b>b</b>) to HIV-1 protease. Carbon atoms were colored blue (<b>a</b>) or green (<b>b</b>), while oxygen, hydrogen, nitrogen, and sulfur atoms were colored red, grey, blue, and yellow, respectively. A green dotted line was used to show the π-cation interaction, and hydrogen bonds were shown as yellow dotted lines.</p>
Full article ">Scheme 1
<p>Synthesis of compounds 1–35. Reagents and conditions: a. Dess-Martin Periodinane, DCM, rt, 3 h. b. K<sub>2</sub>CO<sub>3</sub>, KI, DMF, C<sub>2</sub>H<sub>5</sub>Br, rt, 4 h. c. NH<sub>2</sub>OH·HCl, NaHCO<sub>3</sub>, C<sub>2</sub>H<sub>5</sub>OH, 90 °C, 14 h, 80%. d. Halogenated hydrocarbons, NaH, Anhydrous THF, rt.</p>
Full article ">
20 pages, 3094 KiB  
Article
Blood-Based Transcriptomic Biomarkers Are Predictive of Neurodegeneration Rather Than Alzheimer’s Disease
by Artur Shvetcov, Shannon Thomson, Jessica Spathos, Ann-Na Cho, Heather M. Wilkins, Shea J. Andrews, Fabien Delerue, Timothy A. Couttas, Jasmeen Kaur Issar, Finula Isik, Simranpreet Kaur, Eleanor Drummond, Carol Dobson-Stone, Shantel L. Duffy, Natasha M. Rogers, Daniel Catchpoole, Wendy A. Gold, Russell H. Swerdlow, David A. Brown and Caitlin A. Finney
Int. J. Mol. Sci. 2023, 24(19), 15011; https://doi.org/10.3390/ijms241915011 - 9 Oct 2023
Cited by 1 | Viewed by 1829
Abstract
Alzheimer’s disease (AD) is a growing global health crisis affecting millions and incurring substantial economic costs. However, clinical diagnosis remains challenging, with misdiagnoses and underdiagnoses being prevalent. There is an increased focus on putative, blood-based biomarkers that may be useful for the diagnosis [...] Read more.
Alzheimer’s disease (AD) is a growing global health crisis affecting millions and incurring substantial economic costs. However, clinical diagnosis remains challenging, with misdiagnoses and underdiagnoses being prevalent. There is an increased focus on putative, blood-based biomarkers that may be useful for the diagnosis as well as early detection of AD. In the present study, we used an unbiased combination of machine learning and functional network analyses to identify blood gene biomarker candidates in AD. Using supervised machine learning, we also determined whether these candidates were indeed unique to AD or whether they were indicative of other neurodegenerative diseases, such as Parkinson’s disease (PD) and amyotrophic lateral sclerosis (ALS). Our analyses showed that genes involved in spliceosome assembly, RNA binding, transcription, protein synthesis, mitoribosomes, and NADH dehydrogenase were the best-performing genes for identifying AD patients relative to cognitively healthy controls. This transcriptomic signature, however, was not unique to AD, and subsequent machine learning showed that this signature could also predict PD and ALS relative to controls without neurodegenerative disease. Combined, our results suggest that mRNA from whole blood can indeed be used to screen for patients with neurodegeneration but may be less effective in diagnosing the specific neurodegenerative disease. Full article
(This article belongs to the Special Issue Molecular Aspects of the Neurodegenerative Brain Diseases)
Show Figures

Figure 1

Figure 1
<p>Principal component analysis (PCA) of AD reference dataset, GSE97760.</p>
Full article ">Figure 2
<p>K-means clustering of top 1000 dysregulated genes in GSE97760 identified using principal component analysis. (<b>A</b>) Principal component (PC) 1. Red n = 255, yellow n = 271, green n = 250, and blue n = 224 genes. (<b>B</b>) PC2. Red n = 239, yellow n = 216, green n = 261, and blue n = 284 genes. Gene names are listed in <a href="#app1-ijms-24-15011" class="html-app">Supplementary Tables S1 and S2</a>.</p>
Full article ">Figure 3
<p>Receiver operating characteristic (ROC) curve of the random forest model’s performance for PC1 gene expression cluster in (<b>A</b>) dataset A, GSE63061, and (<b>B</b>) dataset B, GSE63060. AUC: area under the curve.</p>
Full article ">Figure 4
<p>Receiver operating characteristic (ROC) curve of the random forest model’s performance for PC2 metabolic process cluster in (<b>A</b>) dataset A, GSE63061, and (<b>B</b>) dataset B, GSE63060. AUC: area under the curve.</p>
Full article ">Figure 5
<p>Feature selection of the top-performing central gene nodes using recursive feature elimination (RFE) for (<b>A</b>) PC1, gene expression cluster, and (<b>B</b>) PC2, metabolic process cluster. The blue dot indicates peak performance of the model where features were identified from. (<b>C</b>,<b>D</b>) The eight top-performing genes for the (<b>C</b>) PC1 gene expression cluster and (<b>D</b>) PC2 metabolic process cluster.</p>
Full article ">Figure 6
<p>Flow chart of the methodological pipeline. Abbreviations: PC: principal component; ND: neurodegenerative diseases.</p>
Full article ">
41 pages, 1260 KiB  
Article
A Catalog of Coding Sequence Variations in Salivary Proteins’ Genes Occurring during Recent Human Evolution
by Lorena Di Pietro, Mozhgan Boroumand, Wanda Lattanzi, Barbara Manconi, Martina Salvati, Tiziana Cabras, Alessandra Olianas, Laura Flore, Simone Serrao, Carla M. Calò, Paolo Francalacci, Ornella Parolini and Massimo Castagnola
Int. J. Mol. Sci. 2023, 24(19), 15010; https://doi.org/10.3390/ijms241915010 - 9 Oct 2023
Cited by 1 | Viewed by 1269
Abstract
Saliva houses over 2000 proteins and peptides with poorly clarified functions, including proline-rich proteins, statherin, P-B peptides, histatins, cystatins, and amylases. Their genes are poorly conserved across related species, reflecting an evolutionary adaptation. We searched the nucleotide substitutions fixed in these salivary proteins’ [...] Read more.
Saliva houses over 2000 proteins and peptides with poorly clarified functions, including proline-rich proteins, statherin, P-B peptides, histatins, cystatins, and amylases. Their genes are poorly conserved across related species, reflecting an evolutionary adaptation. We searched the nucleotide substitutions fixed in these salivary proteins’ gene loci in modern humans compared with ancient hominins. We mapped 3472 sequence variants/nucleotide substitutions in coding, noncoding, and 5′-3′ untranslated regions. Despite most of the detected variations being within noncoding regions, the frequency of coding variations was far higher than the general rate found throughout the genome. Among the various missense substitutions, specific substitutions detected in PRB1 and PRB2 genes were responsible for the introduction/abrogation of consensus sequences recognized by convertase enzymes that cleave the protein precursors. Overall, these changes that occurred during the recent human evolution might have generated novel functional features and/or different expression ratios among the various components of the salivary proteome. This may have influenced the homeostasis of the oral cavity environment, possibly conditioning the eating habits of modern humans. However, fixed nucleotide changes in modern humans represented only 7.3% of all the substitutions reported in this study, and no signs of evolutionary pressure or adaptative introgression from archaic hominins were found on the tested genes. Full article
(This article belongs to the Special Issue Recent Advances in Salivary Gland and Their Function 2.0)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of basic proline-rich genes and encoded proteins: PRB1 (<b>A</b>), PRB2 (<b>B</b>), PRB3 (<b>C</b>), PRB4 (<b>D</b>). For each protein, the genetic allelic variants (S, small; M, medium; L, large; and VL, very large) are shown on the left-sided column; the resulting alternative proteoforms are shown on the right-sided column as blocks, with the corresponding symbol on top. Vertical dashed lines indicate the pro-protein convertase cleavage sites with corresponding Arg (R) residues’ positions. The P enclosed in a circle denotes phosphorylation sites; aminoacidic substitutions are shown for selected isoforms. See text for additional details.</p>
Full article ">Figure 2
<p>Schematic representation of acidic proline-rich proteins (<b>A</b>) and cystatins (<b>B</b>). For each protein, the genetic allelic variants (S, small; M, medium; L, large; and VL, very large) are shown on the left-sided column; the resulting alternative proteoforms are shown on the right-sided column as blocks with corresponding symbols on top. All cystatin alternative proteoforms feature two disulfide bridges (indicated by brackets between Cys), oxidation (ox), and phosphorylation (P) sites. Vertical dashed lines indicate the pro-protein convertase cleavage sites with corresponding Arg (R) residues’ positions. The P enclosed in a circle denotes phosphorylation sites; ox: oxidation sites; p-E: N-terminal pyroglutamic acid; aminoacidic substitutions are shown for selected isoforms. See text for additional details.</p>
Full article ">Figure 3
<p>Nucleotide substitutions in salivary protein genes. The pie chart shows the type and number of 3472 nucleotide substitutions across the 17 tested salivary genes. In particular, the 428 substitutions found in coding regions included 307 nonsynonymous changes across all the 17 genes tested. See text for additional details.</p>
Full article ">Figure 4
<p>Predicted archaic hominins’ PRB-1 (panel (<b>a</b>)) and PRB-2 (panels (<b>b</b>–<b>d</b>)) protein variants.</p>
Full article ">
23 pages, 7622 KiB  
Article
Keras/TensorFlow in Drug Design for Immunity Disorders
by Paulina Dragan, Kavita Joshi, Alessandro Atzei and Dorota Latek
Int. J. Mol. Sci. 2023, 24(19), 15009; https://doi.org/10.3390/ijms241915009 - 9 Oct 2023
Cited by 1 | Viewed by 1770
Abstract
Homeostasis of the host immune system is regulated by white blood cells with a variety of cell surface receptors for cytokines. Chemotactic cytokines (chemokines) activate their receptors to evoke the chemotaxis of immune cells in homeostatic migrations or inflammatory conditions towards inflamed tissue [...] Read more.
Homeostasis of the host immune system is regulated by white blood cells with a variety of cell surface receptors for cytokines. Chemotactic cytokines (chemokines) activate their receptors to evoke the chemotaxis of immune cells in homeostatic migrations or inflammatory conditions towards inflamed tissue or pathogens. Dysregulation of the immune system leading to disorders such as allergies, autoimmune diseases, or cancer requires efficient, fast-acting drugs to minimize the long-term effects of chronic inflammation. Here, we performed structure-based virtual screening (SBVS) assisted by the Keras/TensorFlow neural network (NN) to find novel compound scaffolds acting on three chemokine receptors: CCR2, CCR3, and one CXC receptor, CXCR3. Keras/TensorFlow NN was used here not as a typically used binary classifier but as an efficient multi-class classifier that can discard not only inactive compounds but also low- or medium-activity compounds. Several compounds proposed by SBVS and NN were tested in 100 ns all-atom molecular dynamics simulations to confirm their binding affinity. To improve the basic binding affinity of the compounds, new chemical modifications were proposed. The modified compounds were compared with known antagonists of these three chemokine receptors. Known CXCR3 compounds were among the top predicted compounds; thus, the benefits of using Keras/TensorFlow in drug discovery have been shown in addition to structure-based approaches. Furthermore, we showed that Keras/TensorFlow NN can accurately predict the receptor subtype selectivity of compounds, for which SBVS often fails. We cross-tested chemokine receptor datasets retrieved from ChEMBL and curated datasets for cannabinoid receptors. The NN model trained on the cannabinoid receptor datasets retrieved from ChEMBL was the most accurate in the receptor subtype selectivity prediction. Among NN models trained on the chemokine receptor datasets, the CXCR3 model showed the highest accuracy in differentiating the receptor subtype for a given compound dataset. Full article
(This article belongs to the Special Issue G Protein-Coupled Receptors 2023)
Show Figures

Figure 1

Figure 1
<p>The validation of the inactive-state CXCR3 model through an analysis of micro- and macroswitches. The inactive-state model of CXCR3 (blue-to-red) was superposed on active-state chemokine receptor structures (gray): 7O7F (CCR5) [<a href="#B88-ijms-24-15009" class="html-bibr">88</a>] for TM6 and the tryptophan toggle switch W6.48 with F6.44 from the PIF motif [<a href="#B89-ijms-24-15009" class="html-bibr">89</a>], and 6WWZ (CCR6) for comparison of the tyrosine toggle switch Y7.53 and the ionic lock including R3.50 from the DRY motif. The residues have been labeled using the Ballesteros–Weinstein numbering system [<a href="#B90-ijms-24-15009" class="html-bibr">90</a>].</p>
Full article ">Figure 2
<p>Results of the MD simulations for CCR2 for four different compounds proposed by virtual screening. (<b>Top</b>) The interactions between the receptor and the ligand obtained after 100 ns of the simulation. The receptor was shown in the red-to-blue color scheme; yellow dashed lines—hydrogen bonds; blue dashed lines—pi-pi stacking. The residues have been labeled using Ballesteros–Weinstein numbering system [<a href="#B90-ijms-24-15009" class="html-bibr">90</a>]. (<b>Bottom</b>) The RMSD plots obtained for each of the ligands over the 100 ns simulation, as well as the average RMSD with its fluctuation range.</p>
Full article ">Figure 3
<p>Results of the MD simulations for CCR3 for five proposed compounds. (<b>Top</b>) The interactions between the receptor and the ligand obtained after 100 ns of the simulation. The receptor was shown in the red-to-blue color scheme; yellow dashed lines—hydrogen bonds; blue dashed lines—pi-pi stacking; green dashed lines—pi-cation; purple dashed lines—salt bridges. The residues have been labeled using Ballesteros–Weinstein numbering system [<a href="#B90-ijms-24-15009" class="html-bibr">90</a>]. (<b>Bottom</b>) The RMSD plots obtained for each of the ligands over the 100 ns simulation, as well as the average RMSD with its fluctuation range.</p>
Full article ">Figure 4
<p>Results of the MD simulations for CXCR3 for five proposed compounds. (<b>Top</b>) The interactions between the receptor and the ligand obtained for 100 ns of the simulation. The receptor was shown in the red-to-blue color scheme; yellow dashed lines—hydrogen bonds; blue dashed lines—pi-pi stacking. The residues have been labeled using Ballesteros–Weinstein numbering system [<a href="#B90-ijms-24-15009" class="html-bibr">90</a>]. (<b>Bottom</b>) The RMSD computed for each of the ligands over the 100 ns simulation, as well as the average RMSD with its fluctuation range.</p>
Full article ">
11 pages, 2781 KiB  
Article
Fast and High-Efficiency Synthesis of Capsanthin in Pepper by Transient Expression of Geminivirus
by Zhimin Lin, Muhammad Moaaz Ali, Xiaoyan Yi, Lijuan Zhang and Shaojuan Wang
Int. J. Mol. Sci. 2023, 24(19), 15008; https://doi.org/10.3390/ijms241915008 - 9 Oct 2023
Cited by 1 | Viewed by 1219
Abstract
The color of the chili fruit is an important factor that determines the quality of the chili, as red chilies are more popular among consumers. The accumulation of capsanthin is the main cause of reddening of the chili fruit. Capsanthin is an important [...] Read more.
The color of the chili fruit is an important factor that determines the quality of the chili, as red chilies are more popular among consumers. The accumulation of capsanthin is the main cause of reddening of the chili fruit. Capsanthin is an important metabolite in carotenoid metabolism, and its production level is closely linked to the expression of the genes for capsanthin/capsorubin synthase (CCS) and carotenoid hydroxylase (CrtZ). We reported for the first time that the synthesis of capsanthin in chili was enhanced by using a geminivirus (Bean Yellow Dwarf Virus). By expressing heterologous β-carotenoid hydroxylase (CrtZ) and β-carotenoid ketolase (CrtW) using codon optimization, the transcription level of the CCS gene and endogenous CrtZ was directly increased. This leads to the accumulation of a huge amount of capsanthin in a very short period of time. Our results provide a platform for the rapid enhancement of endogenous CCS activity and capsanthin production using geminivirus in plants. Full article
(This article belongs to the Special Issue Advances in Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Visualization of GFP expression in <span class="html-italic">N. benthamiana</span> plants under UV light. (<b>a</b>). Normal tobacco plants under UV light; (<b>b</b>). GFP protein expression on tobacco leaves 3 days after injection under UV light. I, inoculated leaves; U, uninoculated leaves.</p>
Full article ">Figure 2
<p>The <span class="html-italic">CrtW</span> and <span class="html-italic">CrtZ</span> genes are expressed in a geminivirus expression system under the CAMV-35S promoter and the strong rbcs promoter. (<b>A</b>) The p1300BR vector was used for <span class="html-italic">CrtW</span> gene expression; (<b>B</b>) the p1300BZ vector was used for <span class="html-italic">CrtZ</span> gene expression.</p>
Full article ">Figure 3
<p>The <span class="html-italic">CrtZ</span> and <span class="html-italic">CrtW</span> genes promote the ripening of the pepper while increasing the expression of the <span class="html-italic">CCS</span> gene. (<b>A</b>) The phenotypic observation of p1300BZ and P1300BR, with Mock as a control; (<b>B</b>) RT-qPCR expression analysis of the <span class="html-italic">CCS</span> genes in p1300BR, p1300BZ, and Mock; (<b>C</b>) RT-qPCR analysis of the expression of the genes <span class="html-italic">CrtW</span>, <span class="html-italic">CrtZ0</span>, <span class="html-italic">LCYB</span>, <span class="html-italic">PSY</span> in p1300BR and Mock; (<b>D</b>) RT-qPCR analysis of the expression of the genes <span class="html-italic">CrtZ</span>, <span class="html-italic">CrtZ0</span>, <span class="html-italic">LCYB</span>, <span class="html-italic">PSY</span> in p1300BZ and Mock. All gene expression analyses were performed using ubiquitin as an internal reference.</p>
Full article ">Figure 4
<p>Expression of <span class="html-italic">CrtW</span> gene significantly enhances capsanthin production in pepper in a geminivirus system. (<b>A</b>) Color comparison of three different pepper powders in cell lysate, including Mock, p1300BZ, and p1300BR; (<b>B</b>) carotenoid production of zeaxanthin, violaxanthin, capsorubin, capsanthin, and antheraxanthin in Mock and p1300BR; (<b>C</b>) values were calculated from the peak area of the HPLC chromatogram.</p>
Full article ">Figure 5
<p>Synthesis pattern of capsanthin under a geminivirus expression system. CrtW—β-carotenoid ketolase; CrtY—lycopene β-cyclase; CrtZ—β-carotenoid hydroxylase; CCS—capsanthin/capsorubin synthase; ZEP—zeaxanthin epoxidase.</p>
Full article ">
22 pages, 3832 KiB  
Article
PNA6, a Lactosyl Analogue of Angiotensin-(1-7), Reverses Pain Induced in Murine Models of Inflammation, Chemotherapy-Induced Peripheral Neuropathy, and Metastatic Bone Disease
by Maha I. Sulaiman, Wafaa Alabsi, Lajos Szabo, Meredith Hay, Robin Polt, Tally M. Largent-Milnes and Todd W. Vanderah
Int. J. Mol. Sci. 2023, 24(19), 15007; https://doi.org/10.3390/ijms241915007 - 9 Oct 2023
Viewed by 1332
Abstract
Pain is the most significant impairment and debilitating challenge for patients with bone metastasis. Therefore, the primary objective of current therapy is to mitigate and prevent the persistence of pain. Thus, cancer-induced bone pain is described as a multifaceted form of discomfort encompassing [...] Read more.
Pain is the most significant impairment and debilitating challenge for patients with bone metastasis. Therefore, the primary objective of current therapy is to mitigate and prevent the persistence of pain. Thus, cancer-induced bone pain is described as a multifaceted form of discomfort encompassing both inflammatory and neuropathic elements. We have developed a novel non-addictive pain therapeutic, PNA6, that is a derivative of the peptide Angiotensin-(1-7) and binds the Mas receptor to decrease inflammation-related cancer pain. In the present study, we provide evidence that PNA6 attenuates inflammatory, chemotherapy-induced peripheral neuropathy (CIPN) and cancer pain confined to the long bones, exhibiting longer-lasting efficacious therapeutic effects. PNA6, Asp-Arg-Val-Tyr-Ile-His-Ser-(O-β-Lact)-amide, was successfully synthesized using solid phase peptide synthesis (SPPS). PNA6 significantly reversed inflammatory pain induced by 2% carrageenan in mice. A second murine model of platinum drug-induced painful peripheral neuropathy was established using oxaliplatin. Mice in the oxaliplatin-vehicle treatment groups demonstrated significant mechanical allodynia compared to the oxaliplatin-PNA6 treatment group mice. In a third study modeling a complex pain state, E0771 breast adenocarcinoma cells were implanted into the femur of female C57BL/6J wild-type mice to induce cancer-induced bone pain (CIBP). Both acute and chronic dosing of PNA6 significantly reduced the spontaneous pain behaviors associated with CIBP. These data suggest that PNA6 is a viable lead candidate for treating chronic inflammatory and complex neuropathic pain. Full article
(This article belongs to the Special Issue New Advance on Molecular Targets for the Treatment of Pain)
Show Figures

Figure 1

Figure 1
<p>The chemical structure of PNA6.</p>
Full article ">Figure 2
<p>Analytical HPLC chromatogram of pure PNA6.</p>
Full article ">Figure 3
<p>Mass spectroscopy spectrum of pure PNA6 shows single charged ion ([M + H]<sup>+</sup> 1212.35) and double charged ion ([M + 2H]<sup>2+</sup> 606.81).</p>
Full article ">Figure 4
<p>PK analysis of PNA6 in serum (<b>A</b>) and brain parenchyma (<b>B</b>). Data are presented as mean ± standard error (SEM).</p>
Full article ">Figure 5
<p>Effects of PNA6 on E0771 cancer cell viability in vitro. (<b>A</b>) E0771 adenocarcinoma cells do not have a significant change in viability when treated with varying concentrations of PNA6 for 24 h; the cell viability was tested using the XTT assay. One-Way ANOVA with Bonferroni post hoc correction (n = 5). (<b>B</b>) Timeline of 1, 2, 3, and 4 h post PNAS application at 1µM utilizing the XTT assay. There was no significant effect on the cell viability of E0771 cells in the presence of PNA6 from 1 to 4 h.</p>
Full article ">Figure 6
<p>Acute model of inflammatory pain. Significant mechanical allodynia was detected after 2% λ-carrageenan injection. (<b>A</b>) Timeline of experiment. (<b>B</b>) Mice injected intra-hind paw with saline control did not show a significant decrease in mechanical thresholds in the ipsilateral hind paw. Neither PNA6 nor saline given intraperitoneally 3 h posts i-paw injection resulted in a significant change over the 3 h testing period. (<b>C</b>) 2% λ-carrageenan injection significantly decreased hind-paw thresholds 3 h post-injection. PNA6 (1 mg/kg i.p.) significantly reversed the mechanical hypersensitivity induced by 2% λ-carrageenan in the ipsilateral hind paw at 60, 90, and 120 min (n = 10–12 mice in each group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01). Saline (1 mL/kg, i.p.) did not significantly affect 2% λ-carrageenan-induced mechanical hypersensitivity. Results were analyzed by repeated measures (RM) two-way ANOVA followed by the Bonferroni post hoc test.</p>
Full article ">Figure 7
<p>Chronic effect of PNA6 on CIPN. Timeline of experiment (<b>A</b>). PNA6 does not impair animal locomotion at all testing times (<b>B</b>). Animals experienced a significant decrease in mechanical thresholds due to the treatment of oxaliplatin (4 mg/kg i.p on days 1, 2, 8, and 9). Animals with oxaliplatin demonstrated a significant increase in the paw withdrawal thresholds on days 7, 10, 11, 13, and 14 after PNA6 (1 mg/kg i.p.) treatment for 14 consecutive days (<b>C</b>). The area under the curve was calculated, demonstrating significant inhibition of CIPN by PNA6 (<b>D</b>). Values represent mean ± SEM, n = 12 per group, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0 0.0001. Results were analyzed by repeated measures (RM) two-way ANOVA followed by Bonferroni’s multiple comparison test.</p>
Full article ">Figure 8
<p>Acute administration of PNA6 in established cancer-induced bone pain attenuates spontaneous pain behavior in CIBP. Animals display no pain behavior before surgery (BL = baseline). Instead, animals demonstrate increased flinching, guarding, and decreased limb use post-cancer surgery inoculation (CIBP). (<b>A</b>) Time-response curve (TRC) of flinching behavior after PNA6 1 mg/kg at times 30, 60, 90, and 120 min on day seven after surgery. (<b>B</b>) Guarding behavior time-response curve after 1 mg/kg at times 30, 60, 90, and 120 min on day seven after surgery. (<b>C</b>) Time-response curve of limb use at 30, 60, 90, and 120 min on day seven after surgery. Saline had no significant effect on all three behaviors. The time of peak effect for PNA6 was 60 min after administration. n = 10–12 mice in each group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01. Results were analyzed using RM, two-way ANOVA followed by Bonferroni post hoc test.</p>
Full article ">Figure 9
<p>Sustained administration of PNA6 attenuates cancer pain in the murine model of cancer-induced bone pain. Spontaneous pain behavior (<b>A</b>) flinching, (<b>B</b>) guarding, and (<b>C</b>) limb use measured before and at regular intervals following cancer inoculation in female C57BL/6J mice. At D7 post-surgery, animals inoculated with E0771 breast adenocarcinoma cells had significantly elevated spontaneous pain behavior in mice in comparison to the sham mice group. PNA6 at D7, 10, and 13 significantly attenuated flinching and guarding in mice with cancer compared to saline-treated mice (<b>A</b>,<b>B</b>). PNA6 at D7 and 10 significantly increased limb use in mice with cancer compared to saline-treated mice (<b>C</b>). MasR1 antagonist, A779, reversed the antinociception of PNA6 in the murine model of cancer-induced bone pain. To investigate receptor dependence, animals were dosed post-surgery from days 7 to 13 with A779 1 mg/ kg 30 min before administration of PNA6. Spontaneous pain behaviors flinching (<b>D</b>), guarding (<b>E</b>), and limb use (<b>F</b>) were recorded as previously described. Daily administration of A779 reversed the effect of PNA6 in the CIBP model. In all behavioral tests, n = 10–12 mice in each group, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001, ns meant no significant difference. Results were analyzed using RM two-way ANOVA followed by Bonferroni post hoc test. Limb Uue scale: 0 = complete lack of use, 1 = partial non-use, 2 = limp with guard, 3 = limping, 4 = normal.</p>
Full article ">Figure 10
<p>Sustained administration of PNA6 did not alter bone integrity. (<b>A</b>) Representative radiograph of the right femur of mice D14 post-surgery. Radiographs were obtained before surgery and at D14 post-surgery to monitor cancer-induced lesions. (<b>B</b>) Quantification of bone scores (n = 10–12/treatment groups). Bone scores were determined on a 5-point scale by blinded observers. The scoring system was as follows: 0 = normal bone, 1 = 1–3 lesions with no fracture, 2 = 4–6 lesions with no fracture, 3 = uni-cortical, full thickness fracture, and 4 = bi-cortical, full thickness fracture. Data are expressed as mean ± S.E.M. ** <span class="html-italic">p</span> &lt; 0.01, ns meant no significant difference (one-way ANOVA, Tukey’s HSD post hoc).</p>
Full article ">
34 pages, 3227 KiB  
Review
Insulin Receptor Isoforms and Insulin Growth Factor-like Receptors: Implications in Cell Signaling, Carcinogenesis, and Chemoresistance
by Mariam Ahmed Galal, Samhar Samer Alouch, Buthainah Saad Alsultan, Huda Dahman, Nouf Abdullah Alyabis, Sarah Ammar Alammar and Ahmad Aljada
Int. J. Mol. Sci. 2023, 24(19), 15006; https://doi.org/10.3390/ijms241915006 - 9 Oct 2023
Cited by 6 | Viewed by 2476
Abstract
This comprehensive review thoroughly explores the intricate involvement of insulin receptor (IR) isoforms and insulin-like growth factor receptors (IGFRs) in the context of the insulin and insulin-like growth factor (IGF) signaling (IIS) pathway. This elaborate system encompasses ligands, receptors, and binding proteins, giving [...] Read more.
This comprehensive review thoroughly explores the intricate involvement of insulin receptor (IR) isoforms and insulin-like growth factor receptors (IGFRs) in the context of the insulin and insulin-like growth factor (IGF) signaling (IIS) pathway. This elaborate system encompasses ligands, receptors, and binding proteins, giving rise to a wide array of functions, including aspects such as carcinogenesis and chemoresistance. Detailed genetic analysis of IR and IGFR structures highlights their distinct isoforms, which arise from alternative splicing and exhibit diverse affinities for ligands. Notably, the overexpression of the IR-A isoform is linked to cancer stemness, tumor development, and resistance to targeted therapies. Similarly, elevated IGFR expression accelerates tumor progression and fosters chemoresistance. The review underscores the intricate interplay between IRs and IGFRs, contributing to resistance against anti-IGFR drugs. Consequently, the dual targeting of both receptors could present a more effective strategy for surmounting chemoresistance. To conclude, this review brings to light the pivotal roles played by IRs and IGFRs in cellular signaling, carcinogenesis, and therapy resistance. By precisely modulating these receptors and their complex signaling pathways, the potential emerges for developing enhanced anti-cancer interventions, ultimately leading to improved patient outcomes. Full article
(This article belongs to the Special Issue The Role of the IGF Axis in Disease, 3nd Edition)
Show Figures

Figure 1

Figure 1
<p>The formation of IR isoforms: (1) Gene expression: <span class="html-italic">INSR</span> gene encoding IR is transcribed from DNA in the nucleus. (2) Pre-mRNA splicing involving introns’ removal and exclusion/inclusion of E11 to form mature mRNA molecules that encode either IR-A or IR-B isoforms; this process is regulated by specific splicing factors that bind to <span class="html-italic">cis</span>-acting elements in the pre-mRNA. (3) mRNA export from the nucleus to the cytoplasm through nuclear pores to serve as templates for protein synthesis. (4) mRNA Translation for IR synthesis. (5) Post-translational modifications (PTMs) in the endoplasmic reticulum and Golgi apparatus, including glycosylation, disulfide bond formation, and proteolytic cleavage. (6) Protein trafficking and secretion: mature IR isoforms are transported from the ER to the plasma membrane, where they are inserted and anchored by transmembrane domains. This process occurs through vesicular transport and fusion with the plasma membrane. Once inserted into the plasma membrane, IR isoforms are available for ligand binding and downstream signaling.</p>
Full article ">Figure 2
<p>Structure of IR receptors. (<b>A</b>) Illustration showing different IR domains encoded by the 22 exons. IR has two main subunits: α and β. The α subunit contains 5 main domains, L1 (AA 28–174) CR (AA 182–339), and L2 (AA 340–497), and 2 Fibronectin subunits: FnIII-1 (residue 624–726) and FnIII-2 (757–842). The two α-subunits are linked by a disulfide bond between the two Cys 524 in the first FnIII domain. One to three of the triplet Cys at 682, 683, and 685 in the insert within the second FnIII domain are also involved in α-α disulfide bridges. There is a single disulfide bridge between α and β subunits between Cys 647 in the insert domain and Cys 872. The β subunit details are explained in the main text. Teal arrow shows 6 O-glycosylations, whereas the pink arrows imply the N-glycosylation. (<b>B</b>) IR-B amino acid sequence colored to identify each domain sequence presented in (<b>A</b>). (<b>C</b>) 3D structure of IR-B showing the Λ-shaped structure when no ligand is bound to it. (<b>D</b>) Structural differences between IR-A and IR-B. JM: Juxtamembrane, TK: Tyrosine Kinase. The 3D models were created using Swiss Model.</p>
Full article ">Figure 3
<p>Hypothetical mechanisms of IR receptor activation upon ligand binding as proposed in the literature. (<b>A</b>) Ward et al. [<a href="#B40-ijms-24-15006" class="html-bibr">40</a>]; (<b>B</b>) Lee et al. [<a href="#B99-ijms-24-15006" class="html-bibr">99</a>]; (<b>C</b>) Kavran et al. [<a href="#B100-ijms-24-15006" class="html-bibr">100</a>]; (<b>D</b>) Maruyama et al. [<a href="#B97-ijms-24-15006" class="html-bibr">97</a>].</p>
Full article ">Figure 4
<p>An illustration of the downstream pathways upon activation of IR and IGF-IR in physiological conditions. Specific differences between the receptors’ signaling pathways are explained in the text. Red circles are phosphorylation; green lines are excitatory signals; red lines are inhibitory signals.</p>
Full article ">
19 pages, 6727 KiB  
Article
Genome-Wide Evolutionary Characterization and Expression Analysis of Major Latex Protein (MLP) Family Genes in Tomato
by Zhengliang Sun, Liangzhe Meng, Yuhe Yao, Yanhong Zhang, Baohui Cheng and Yan Liang
Int. J. Mol. Sci. 2023, 24(19), 15005; https://doi.org/10.3390/ijms241915005 - 9 Oct 2023
Cited by 2 | Viewed by 1397
Abstract
Major latex proteins (MLPs) play a key role in plant response to abiotic and biotic stresses. However, little is known about this gene family in tomatoes (Solanum lycopersicum). In this paper, we perform a genome-wide evolutionary characterization and gene expression analysis [...] Read more.
Major latex proteins (MLPs) play a key role in plant response to abiotic and biotic stresses. However, little is known about this gene family in tomatoes (Solanum lycopersicum). In this paper, we perform a genome-wide evolutionary characterization and gene expression analysis of the MLP family in tomatoes. We found a total of 34 SlMLP members in the tomato genome, which are heterogeneously distributed on eight chromosomes. The phylogenetic analysis of the SlMLP family unveiled their evolutionary relationships and possible functions. Furthermore, the tissue-specific expression analysis revealed that the tomato MLP members possess distinct biological functions. Crucially, multiple cis-regulatory elements associated with stress, hormone, light, and growth responses were identified in the promoter regions of these SlMLP genes, suggesting that SlMLPs are potentially involved in plant growth, development, and various stress responses. Subcellular localization demonstrated that SlMLP1, SlMLP3, and SlMLP17 are localized in the cytoplasm. In conclusion, these findings lay a foundation for further dissecting the functions of tomato SlMLP genes and exploring the evolutionary relationships of MLP homologs in different plants. Full article
Show Figures

Figure 1

Figure 1
<p>Phylogenetic analysis of MLP direct homologs in different plants. The MLP family is divided into three groups, represented by different colors. At, Arabidopsis; Sl, tomato; Md, apple; Cs, cucumber.</p>
Full article ">Figure 2
<p>Phylogenetic analysis, gene structure, and conserved motifs of tomato MLP genes. (<b>a</b>) Construction of NJ tree consisting of 34 SlMLP protein sequences. (<b>b</b>) Distribution of conserved motifs in MLP proteins. (<b>c</b>) Exon/intron structure of the SlMLP gene. Different colored boxes represent different themes. The length of the motifs can be estimated using the scale at the bottom.</p>
Full article ">Figure 3
<p>Distribution of the 34 SlMLP genes on the 8 chromosomes. Vertical bars represent chromosomes, and chromosome numbers are at the top of each chromosome. The scale on the left represents the chromosome length (Mb).</p>
Full article ">Figure 4
<p>Syntenic analysis of tomato and three other plant MLP genes. All homozygous blocks between the two genomes are indicated by gray lines in the background, and homozygous MLP gene pairs are marked by red lines. The numbers indicate the order of the chromosomes.</p>
Full article ">Figure 5
<p>Cis-element analysis of the SlMLP promoters. (<b>a</b>–<b>d</b>) SlMLP cis-elements of promoters are classified into different groups. The elements are indicated by differently colored boxes. (<b>e</b>) The proportion of cis-regulatory elements in the promoter region of the SlMLP gene. (<b>f</b>) Number of cis-elements in four groups for each SlMLP promoter.</p>
Full article ">Figure 6
<p>Expression patterns of SlMLPs in various tomato tissues. The qRT-PCR data were log2-normalized to construct the heat map using TBtools (1.0) software. The roots, stems, shoot apices, and leaves from plants at the 4 true-leaf stage, fully opened flowers, fruits at 7 DPA (days post-anthesis), and seeds from fruits at the mature red stage were used for this analysis.</p>
Full article ">Figure 7
<p>qRT-PCR analyses of SlMLP genes under different abiotic stresses. The relative expression levels of SMLPs under heat (42 °C), cold (4 °C), salt (NaCl), and drought (PEG-6000) are marked in orange, blue, brown, and green colors, respectively. Each value is the mean ± SD of three biological replicates and vertical bars indicate standard deviation. Asterisks represent significant differences in gene expression between abiotic stress treatments (different time points) and the control (0 h) by Student’s <span class="html-italic">t</span>-tests (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 8
<p>Subcellular localization of SlMLP proteins. GFP signals showing the subcellular localization of the selected SlMLP proteins. Bars = 20 μm.</p>
Full article ">Figure 9
<p>Predicted interaction network of SlMLP proteins. The different data sources of predicted interactions are indicated by lines with different colors.</p>
Full article ">
19 pages, 5873 KiB  
Article
An Investigation of the JAZ Family and the CwMYC2-like Protein to Reveal Their Regulation Roles in the MeJA-Induced Biosynthesis of β-Elemene in Curcuma wenyujin
by Yuyang Liu, Shiyi Wu, Kaer Lan, Qian Wang, Tingyu Ye, Huanan Jin, Tianyuan Hu, Tian Xie, Qiuhui Wei and Xiaopu Yin
Int. J. Mol. Sci. 2023, 24(19), 15004; https://doi.org/10.3390/ijms241915004 - 9 Oct 2023
Cited by 3 | Viewed by 1224
Abstract
β-Elemene (C15H24), a sesquiterpenoid compound isolated from the volatile oil of Curcuma wenyujin, has been proven to be effective for multiple cancers and is widely used in clinical treatment. Unfortunately, the β-elemene content in C. wenyujin is very [...] Read more.
β-Elemene (C15H24), a sesquiterpenoid compound isolated from the volatile oil of Curcuma wenyujin, has been proven to be effective for multiple cancers and is widely used in clinical treatment. Unfortunately, the β-elemene content in C. wenyujin is very low, which cannot meet market demands. Our previous research showed that methyl jasmonate (MeJA) induced the accumulation of β-elemene in C. wenyujin. However, the regulatory mechanism is unclear. In this study, 20 jasmonate ZIM-domain (JAZ) proteins in C. wenyujin were identified, which are the core regulatory factors of the JA signaling pathway. Then, the conservative domains, motifs composition, and evolutionary relationships of CwJAZs were analyzed comprehensively and systematically. The interaction analysis indicated that CwJAZs can form homodimers or heterodimers. Fifteen out of twenty CwJAZs were significantly induced via MeJA treatment. As the master switch of the JA signaling pathway, the CwMYC2-like protein has also been identified and demonstrated to interact with CwJAZ2/3/4/5/7/15/17/20. Further research found that the overexpression of the CwMYC2-like gene increased the accumulation of β-elemene in C. wenyujin leaves. Simultaneously, the expressions of HMGR, HMGS, DXS, DXR, MCT, HDS, HDR, and FPPS related to β-elemene biosynthesis were also up-regulated by the CwMYC2-like protein. These results indicate that CwJAZs and the CwMYC2-like protein respond to the JA signal to regulate the biosynthesis of β-elemene in C. wenyujin. Full article
(This article belongs to the Special Issue Advances in Molecular Plant Sciences)
Show Figures

Figure 1

Figure 1
<p>Sequence logos and multiple sequence alignment of TIFY domains (<b>A</b>) and Jas domains (<b>B</b>) of CwJAZ proteins in <span class="html-italic">C. wenyujin</span>. The conserved TIFY/Jas/Degron/NLS motifs were outlined with a red straight line in sequence logos and a black straight line in alignment of the amino acid sequences, respectively. Each stack height indicates the conservation of the sequence at the corresponding position. Each letter height within each stack indicates the relative frequency of the corresponding amino acid. The amino acid residues with a black, purple, and gray background represent 100%, at least 80%, and at least 60% identity, respectively.</p>
Full article ">Figure 2
<p>Phylogenetic relationships and motif compositions of JAZ proteins in <span class="html-italic">C. wenyujin</span>. (<b>A</b>) The phylogenetic tree of CwJAZ proteins was constructed using the neighbor-joining method with a bootstrap of 1000 replicates via MEGA 11. (<b>B</b>) Schematic diagrams of motif compositions. Different motifs for CwJAZ proteins were indicated by different colored boxes and numbered 1–10.</p>
Full article ">Figure 3
<p>Phylogenetic tree of JAZ proteins from <span class="html-italic">C. wenyujin</span>, <span class="html-italic">A. thaliana</span>, <span class="html-italic">O. sativa</span>, and <span class="html-italic">Z. mays</span>. The phylogenetic tree was created using MEGA 11 software coupled with the neighbor-joining method with 1000 bootstrap replicates. Overall, 20 CwJAZs (red circle), 12 AtJAZs (blue triangle), 15 OsJAZs (pink rhombus), and 29 ZmJAZs (green square) were classified into four groups (groups <b>I</b>–<b>IV</b>).</p>
Full article ">Figure 4
<p>Expression levels of <span class="html-italic">CwJAZs</span> under MeJA treatment. Three biological replicates and three technical replicates were performed. Vertical bars refer to ±SE (<span class="html-italic">n</span> = 3). Asterisks indicate significant differences (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 5
<p>Homo- and heteromeric interactions of 14 CwJAZ proteins via yeast two-hybrid assay. Recombinant plasmids were transformed into yeast strain AH109. Then, co-transformants were screened with SD/-Leu/-Trp medium (<b>A</b>) and SD/-Leu/-Trp/-His/-Ade medium with X-α-gal (<b>C</b>). (<b>B</b>) The co-transformants with pGADT7-LargeT and pGBKT7-P53 were used as positive controls, while those with pGADT7-LargeT and pGBKT7-LaminC were used as negative controls. Three independent experiments were performed.</p>
Full article ">Figure 6
<p>Bioinformatics analysis of the CwMYC2-like protein. (<b>A</b>) A phylogenetic tree containing CwMYC2-like and 14 MYC2 proteins in other species. (<b>B</b>) Schematic diagram of the CwMYC2-like protein domain. (<b>C</b>) Amino acid sequence alignment between the CwMYC2-like protein and other species of MYC2 proteins. The outlined boxes indicate conserved domains. The amino acid residues with a black, purple, and gray background represent 100%, at least 80%, and at least 60% identity, respectively. The asterisk is the marker of an integer.</p>
Full article ">Figure 7
<p>Analysis of expression pattern, transcriptional activation, and interactions of CwMYC2-like protein. (<b>A</b>) Expression level of <span class="html-italic">CwMYC2-like</span> gene in different tissues of <span class="html-italic">C. wenyujin</span>. Leave-3M and Flower-3M: leaves and flower obtained from 3-month-old <span class="html-italic">C. wenyujin</span> plant. Leave-6M and Rhizome-6M: leaves and rhizomes obtained from 6-month-old <span class="html-italic">C. wenyujin</span> plant. (<b>B</b>) Expression levels of <span class="html-italic">CwMYC2-like</span> gene in the leaf under MeJA treatment. Three biological replicates and three technical replicates were performed. Vertical bars refer to ±SE (<span class="html-italic">n</span> = 3). Asterisks indicate significant differences (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01). (<b>C</b>) Analysis of the transactivation activity of CwMYC2-like protein in yeast. Recombinant plasmids were transformed into yeast strain AH109, and then the transformant strains were screened with SD/-Trp, SD/-His/-Trp/+X-a-gal and SD/-Trp/-His/-Ade/+X-a-gal media. (<b>D</b>) The interactions between CwJAZ proteins and CwMYC2-like protein. Three independent experiments were performed.</p>
Full article ">Figure 8
<p>The biological function of CwMYC2-like protein in <span class="html-italic">C. wenyujin</span>. (<b>A</b>) Expression level of <span class="html-italic">CwMYC2</span>-<span class="html-italic">like</span> gene in control (EV) and overexpression (OE) lines. (<b>B</b>) Gas chromatography–mass spectrometry (GC-MS) analysis of β-elemene content extracted from EV and OE leaves. (<b>C</b>) The content of β-elemene in control and OE lines. Three biological replicates were performed. Vertical bars refer to ±SE (<span class="html-italic">n</span> = 3). Asterisks indicate significant differences (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 9
<p>Expression levels of structure genes related to β-elemene biosynthesis. Three biological replicates and three technical replicates were performed. Vertical bars refer to ±SE (<span class="html-italic">n</span> = 3). Asterisks indicate significant differences (* <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01). EV and OE lines represent <span class="html-italic">A. tumefaciens</span> GV3101 carrying the pBI121-GUS empty vector and GV3101 carrying the pBI121-CwMYC2-like-GUS recombinant plasmid, respectively. HMGS, 3-hydroxy-3-methylglutaryl-CoA; HMGR, 3-hydroxy-3-methylglutaryl-CoA reductase; DXS, 1-deoxy-<span class="html-small-caps">d</span>-xylulose 5-phosphate synthase; DXR, 1-deoxy-<span class="html-small-caps">d</span>-xylulose 5-phosphate reductoisomerase; MCT, 2-<span class="html-italic">C</span>-methyl-<span class="html-small-caps">d</span>-erythritol 4-phosphate cytidylyltransferase; HDS, 1-hydroxy-2-methyl-2-butenyl 4-diphosphate synthase; HDR, 4-hydroxy-3-methylbut-2-enyldiphosphatereductase; FPPS, farnesyl diphosphate.</p>
Full article ">
Previous Issue
Back to TopTop