[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,509)

Search Parameters:
Keywords = glutamine

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
28 pages, 1604 KiB  
Review
The Metabolomic Footprint of Liver Fibrosis
by Diren Beyoğlu, Yury V. Popov and Jeffrey R. Idle
Cells 2024, 13(16), 1333; https://doi.org/10.3390/cells13161333 (registering DOI) - 11 Aug 2024
Abstract
Both experimental and clinical liver fibrosis leave a metabolic footprint that can be uncovered and defined using metabolomic approaches. Metabolomics combines pattern recognition algorithms with analytical chemistry, in particular, 1H and 13C nuclear magnetic resonance spectroscopy (NMR), gas chromatography–mass spectrometry (GC–MS) [...] Read more.
Both experimental and clinical liver fibrosis leave a metabolic footprint that can be uncovered and defined using metabolomic approaches. Metabolomics combines pattern recognition algorithms with analytical chemistry, in particular, 1H and 13C nuclear magnetic resonance spectroscopy (NMR), gas chromatography–mass spectrometry (GC–MS) and various liquid chromatography–mass spectrometry (LC–MS) platforms. The analysis of liver fibrosis by each of these methodologies is reviewed separately. Surprisingly, there was little general agreement between studies within each of these three groups and also between groups. The metabolomic footprint determined by NMR (two or more hits between studies) comprised elevated lactate, acetate, choline, 3-hydroxybutyrate, glucose, histidine, methionine, glutamine, phenylalanine, tyrosine and citrate. For GC–MS, succinate, fumarate, malate, ascorbate, glutamate, glycine, serine and, in agreement with NMR, glutamine, phenylalanine, tyrosine and citrate were delineated. For LC–MS, only β-muricholic acid, tryptophan, acylcarnitine, p-cresol, valine and, in agreement with NMR, phosphocholine were identified. The metabolomic footprint of liver fibrosis was upregulated as regards glutamine, phenylalanine, tyrosine, citrate and phosphocholine. Several investigators employed traditional Chinese medicine (TCM) treatments to reverse experimental liver fibrosis, and a commentary is given on the chemical constituents that may possess fibrolytic activity. It is proposed that molecular docking procedures using these TCM constituents may lead to novel therapies for liver fibrosis affecting at least one-in-twenty persons globally, for which there is currently no pharmaceutical cure. This in-depth review summarizes the relevant literature on metabolomics and its implications in addressing the clinical problem of liver fibrosis, cirrhosis and its sequelae. Full article
(This article belongs to the Section Cellular Metabolism)
Show Figures

Figure 1

Figure 1
<p>The de novo synthesis of ascorbic acid showing potential enzymes involved whose mRNA expression was determined using quantitative RT-PCR. Intermediates are G-6-P, glucose 6-phosphate; F-6-P, fructose 6-phosphate; G-1-P, glucose 1-phosphate; UDPG, uridine diphosphate glucose; UDPGA, uridine diphosphate glucuronic acid. The enzymes are GCK, glucokinase (HK4; EC 2.7.1.1); ADPGK, ADP-dependent glucokinase (EC 2.7.1.147); HK1, hexokinase 1 (EC 2.7.1.1); HK2, hexokinase 2 (EC 2.7.1.1); GPI, glucose 6-phosphate isomerase (EC 5.3.1.9); PGM1, phosphoglucomutase 1 (EC 5.4.2.2); UGP2, UDP-glucose pyrophosphorylase 2 (EC 2.7.7.9); UGDH, UDP glucose 6-dehydrogenase (EC 1.1.1.22); UGT1A1, UDP glucuronosyltransferase family 1 member A1 (EC 2.4.1.17); AKR1A4, aldo-keto reductase family 1, member A1 (aldehyde reductase; EC 1.1.1.2); AKR1B3, aldo-keto reductase family 1, member B3 (aldose reductase; EC 1.1.1.21); RGN, regucalcin (gluconolactonase; EC 3.1.1.17); GULO, gulonolactone oxidase (EC 1.1.3.8). Adapted from [<a href="#B164-cells-13-01333" class="html-bibr">164</a>] with permission.</p>
Full article ">Figure 2
<p>The treatments or the chemical constituents contained therein for experimental fibrosis in rodents. Picroside I [<a href="#B166-cells-13-01333" class="html-bibr">166</a>] and (–)-phylligenin [<a href="#B167-cells-13-01333" class="html-bibr">167</a>] were administered as such. Berberine is one of the isoquinoline alkaloids present in <span class="html-italic">Corydalis saxicola</span> Bunting [<a href="#B168-cells-13-01333" class="html-bibr">168</a>]. The ecdysone insect molting hormones are likely significant constituents of the ethanol extract of <span class="html-italic">Periplanata americana</span> (American cockroach) that comprises Ganlong capsules [<a href="#B169-cells-13-01333" class="html-bibr">169</a>]. Gypenoside XVII is a saponin extract derived from <span class="html-italic">Gynostemma pentaphyllum</span> [<a href="#B135-cells-13-01333" class="html-bibr">135</a>,<a href="#B141-cells-13-01333" class="html-bibr">141</a>]. Amarogentin is a secoiridoid glycoside from gentian root [<a href="#B138-cells-13-01333" class="html-bibr">138</a>]. Herbarulide and dankasterone A are from the popular edible fungus <span class="html-italic">Flammulina velutipes</span> [<a href="#B136-cells-13-01333" class="html-bibr">136</a>]. Forsythin and forsythiaside A are from Forsythiae fructus [<a href="#B137-cells-13-01333" class="html-bibr">137</a>].</p>
Full article ">Figure 3
<p>Venn diagram showing the upregulated metabolites in mouse, rat and human serum, urine, liver and feces discovered by NMR, GC–MS and LC–MS. Metabolites were included in each section if they had been reported in two or more studies. Note that no metabolites were discovered universally by all three analytical platforms or in common by GC–MS and LC–MS.</p>
Full article ">
16 pages, 2841 KiB  
Review
Vitamin C-Dependent Uptake of Non-Heme Iron by Enterocytes, Its Impact on Erythropoiesis and Redox Capacity of Human Erythrocytes
by Xia Pan, Martin Köberle and Mehrdad Ghashghaeinia
Antioxidants 2024, 13(8), 968; https://doi.org/10.3390/antiox13080968 - 9 Aug 2024
Viewed by 370
Abstract
In the small intestine, nutrients from ingested food are absorbed and broken down by enterocytes, which constitute over 95% of the intestinal epithelium. Enterocytes demonstrate diet- and segment-dependent metabolic flexibility, enabling them to take up large amounts of glutamine and glucose to meet [...] Read more.
In the small intestine, nutrients from ingested food are absorbed and broken down by enterocytes, which constitute over 95% of the intestinal epithelium. Enterocytes demonstrate diet- and segment-dependent metabolic flexibility, enabling them to take up large amounts of glutamine and glucose to meet their energy needs and transfer these nutrients into the bloodstream. During glycolysis, ATP, lactate, and H+ ions are produced within the enterocytes. Based on extensive but incomplete glutamine oxidation large amounts of alanine or lactate are produced. Lactate, in turn, promotes hypoxia-inducible factor-1α (Hif-1α) activation and Hif-1α-dependent transcription of various proton channels and exchangers, which extrude cytoplasmic H+-ions into the intestinal lumen. In parallel, the vitamin C-dependent and duodenal cytochrome b-mediated conversion of ferric iron into ferrous iron progresses. Finally, the generated electrochemical gradient is utilized by the divalent metal transporter 1 for H+-coupled uptake of non-heme Fe2+-ions. Iron efflux from enterocytes, subsequent binding to the plasma protein transferrin, and systemic distribution supply a wide range of cells with iron, including erythroid precursors essential for erythropoiesis. In this review, we discuss the impact of vitamin C on the redox capacity of human erythrocytes and connect enterocyte function with iron metabolism, highlighting its effects on erythropoiesis. Full article
(This article belongs to the Special Issue Blood Cells and Redox Homeostasis in Health and Disease)
Show Figures

Figure 1

Figure 1
<p><b>Transmembrane H<sub>2</sub>S diffusion and Band-3 mediated Cl</b><sup>−</sup><b>/HS</b><sup>−</sup> <b>exchange in hRBCs.</b> Methemoglobin (Hb-Fe<sup>3+</sup>)-mediated H<sub>2</sub>S degradation ensures the maintenance of physiological plasma and tissue concentration of free H<sub>2</sub>S. The Cl<sup>−</sup>/HS<sup>−</sup>/H<sub>2</sub>S cycle is also efficiently involved in net acid (H<sup>+</sup>-ions) efflux; for more details, see the following review [<a href="#B15-antioxidants-13-00968" class="html-bibr">15</a>]. For interactions of hRBCs with endogenous cells and pathogens, see the main text.</p>
Full article ">Figure 2
<p><b>Uptake and systemic circulation of non-heme iron require several carriers located in the cell membrane of human enterocytes, plasma protein transferrin, and transferrin receptor.</b> Erythropoiesis requires liver- and kidney-dependent production of erythropoietin.</p>
Full article ">Figure 3
<p>(<b>A</b>) <b>Inverse correlation between DMT1 abundance and pH along the small intestinal. (B) GLUT-1-dependent influx of glucose into human enterocytes.</b> DMT1: divalent metal transporter 1; GLUT-1: glucose transporter-1.</p>
Full article ">Figure 4
<p><b>GLUT-1-mediated influx of oxidized form of vitamin C (DHA) into the mature hRBCs</b>. The interplay between DHA, PPP, GSH, AA, and the subsequent reduction in vitamin E prevents lipid peroxidation. As a result, cell membrane integrity is maintained and in vivo hemolysis of erythrocytes is minimized. Recently, a link between iron metabolism, lipid peroxidation, and hemolysis was found in stored human and mice erythrocytes [<a href="#B113-antioxidants-13-00968" class="html-bibr">113</a>,<a href="#B114-antioxidants-13-00968" class="html-bibr">114</a>]. Vitamin E is located inside the cell membrane.</p>
Full article ">Figure 5
<p><b>Iron transport into erythroid precursors</b>. This process comprises the endocytosis of Transferrin-bound iron, DMT1B-II-mediated Fe<sup>2+</sup> export from acidified endosomes into the cytoplasm. MT: mitochondria.</p>
Full article ">
18 pages, 4636 KiB  
Article
Exploring the Role of Bergamot Polyphenols in Alleviating Morphine-Induced Hyperalgesia and Tolerance through Modulation of Mitochondrial SIRT3
by Sara Ilari, Saverio Nucera, Lucia Carmela Passacatini, Federica Scarano, Roberta Macrì, Rosamaria Caminiti, Stefano Ruga, Maria Serra, Luigino Antonio Giancotti, Filomena Lauro, Concetta Dagostino, Valeria Mazza, Giovanna Ritorto, Francesca Oppedisano, Jessica Maiuolo, Ernesto Palma, Valentina Malafoglia, Carlo Tomino, Vincenzo Mollace and Carolina Muscoli
Nutrients 2024, 16(16), 2620; https://doi.org/10.3390/nu16162620 - 9 Aug 2024
Viewed by 298
Abstract
Morphine is an important pain reliever employed in pain management, its extended utilize is hindered by the onset of analgesic tolerance and oxidative stress. Long-term morphine administration causes elevated production of reactive oxygen species (ROS), disrupting mitochondrial function and inducing oxidation. Sirtuin 3 [...] Read more.
Morphine is an important pain reliever employed in pain management, its extended utilize is hindered by the onset of analgesic tolerance and oxidative stress. Long-term morphine administration causes elevated production of reactive oxygen species (ROS), disrupting mitochondrial function and inducing oxidation. Sirtuin 3 (SIRT3), a mitochondrial protein, is essential in modulating ROS levels by regulating mitochondrial antioxidant enzymes as manganese superoxide dismutase (MnSOD). Our investigation focused on the impact of SIRT3 on hyperalgesia and morphine tolerance in mice, as evaluating the antioxidant effect of the polyphenolic fraction of bergamot (BPF). Mice were administered morphine twice daily for four consecutive days (20 mg/kg). On the fifth day, mice received an acute dose of morphine (3 mg/kg), either alone or in conjunction with BPF or Mn (III)tetrakis (4-benzoic acid) porphyrin (MnTBAP). We evaluated levels of malondialdehyde (MDA), nitration, and the activity of SIRT3, MnSOD, glutamine synthetase (GS), and glutamate 1 transporter (GLT1) in the spinal cord. Our findings demonstrate that administering repeated doses of morphine led to the development of antinociceptive tolerance in mice, accompanied by increased superoxide production, nitration, and inactivation of mitochondrial SIRT3, MnSOD, GS, and GLT1. The combined administration of morphine with either BPF or MnTBAP prevented these effects. Full article
(This article belongs to the Special Issue Effects of Natural Bioactives on Pain and Neuroinflammation)
Show Figures

Figure 1

Figure 1
<p>Acute morphine injection. Administration of morphine (3 mg/kg), in mice, generated a considerable near-maximal antinociceptive response persisting for 60 min. Values are reported as mean ± SEM, based on 15 mice; * <span class="html-italic">p</span> &lt; 0.0001 vs. morphine 0 mg/kg.</p>
Full article ">Figure 2
<p>A significant loss to the antinociceptive effect of the acute injection of morphine was observed in animals that received repeated administration of morphine over 4 days. Concurrent administration of morphine with BPF (5–50 mg/kg) (<b>A</b>) or MnTBAP (5–30 mg/kg) (<b>B</b>) over a period of 4 days inhibited the development of tolerance in a dose-dependent manner. * <span class="html-italic">p</span> &lt; 0.001 compared to vehicle (Veh); † <span class="html-italic">p</span> &lt; 0.01; †† <span class="html-italic">p</span> &lt; 0.001 compared to vehicle + morphine.</p>
Full article ">Figure 3
<p>(<b>A</b>) Repeated administration of morphine for 4 days in mice caused an increased production of superoxide in the spinal cord compared to the control group (vehicle), as demonstrated by the oxidation of HE. Co-administration of morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg) was able to reduce the increase in ethidium and therefore in superoxide. Original magnification, ×10. Scale Bar 100 µm. Micrographs illustrate results from at least three distinct animals. (<b>B</b>) Persistent morphine treatment induced protein nitration in the spinal cord. Co-administration of morphine with BPF (25 mg/kg) and MnTBAP (10 mg/kg) inhibited nitrotyrosine formation. Original magnification, ×10. Scale Bar 100 µm. Micrographs illustrate results from at least three distinct animals in experiments conducted on separate days.</p>
Full article ">Figure 4
<p>Increased MDA levels in spinal cord represents the presence of oxidative stress during morphine tolerance in mice. Mice that received morphine for 4 days showed an amount of MDA level. Co-administration of morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg) resulted in a substantial decrease in MDA. The results are shown as the mean ± SEM for 6 mice. * <span class="html-italic">p</span> &lt; 0.0001 versus vehicle (Veh); † <span class="html-italic">p</span> &lt; 0.01 versus vehicle + morphine.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>) Nitration of GS and GLT1 proteins in spinal cord tissues as assessed by immunoprecipitation. Administering morphine for 4 days in combination with BPF (25 mg/kg) or MnTBAP (10 mg/kg) prevented the nitration of GS and GLT1. In these conditions, actin expression appeared statistically similar across the lanes. The reported data include densitometric analyses for all animals per group. GS, nitrated GS, GLT1 and nitrated GLT1 were first normalized to actin and then these values were used to obtain GS nitrated/GS and GLT1 nitrated/GLT1 ratio. The data are presented as the mean ± SEM for 6 mice; * <span class="html-italic">p</span> &lt; 0.001 versus Veh; † <span class="html-italic">p</span> &lt; 0.001 versus morphine.</p>
Full article ">Figure 6
<p>Chronic morphine administration in mice induced nitration on mitochondrial proteins as shown by WES methodology. Co-administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg) attenuated mitochondrial proteins nitration. No statistically significant difference for the TOM20 value was identified in the lanes under these conditions. (<b>A</b>,<b>B</b>) Lanes and (<b>D</b>) electropherogram are representative of the results from six animals; (blue lanes: morphine groups; pink lanes: BPF group; green lanes: MnTBAP group; and light grey: vehicle group). (<b>C</b>) The reported data include densitometric analyses for all animals per group. The results are presented as the mean ± SEM for six mice. * <span class="html-italic">p</span> &lt; 0.05 versus Veh; † <span class="html-italic">p</span> &lt; 0.05 versus vehicle + morphine.</p>
Full article ">Figure 7
<p>(<b>A</b>) Nitration of MnSOD protein in spinal cord tissues as assessed by immunoprecipitation. Combined treatment with morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg), for four consecutive days, prevented MnSOD nitration. In these conditions, prohibitin expression appeared statistically similar across the lanes. Densitometric analyses for all animals in each group are reported. MnSOD and nitrated MnSOD were first normalized with prohibitin and then these values were used to obtain MnSOD nitrated/MnSOD ratio. (<b>B</b>) Nitration on MnSOD is linked to inactivation of its biological function, which is restored following the administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg). The results are presented as the mean ± SEM for six mice; * <span class="html-italic">p</span> &lt; 0.001 compared to Veh; † <span class="html-italic">p</span> &lt; 0.001 compared to morphine.</p>
Full article ">Figure 8
<p>(<b>A</b>) Nitration of SIRT3 protein in spinal cord tissues is identified by immunoprecipitation. Combined treatment with morphine and BPF (25 mg/kg) or MnTBAP (10 mg/kg), for 4 consecutive days, blocked SIRT3 nitration. Prohibitin levels appeared statistically similar across the lanes. Densitometric analyses for all animals in each group are reported. MnSOD and nitrated MnSOD were initially normalized using prohibitin, and these values were then utilized to calculate the MnSOD nitrated/MnSOD ratio. (<b>B</b>) SIRT3 activation, expressed in arbitrary fluorescence units (AFU), is restored following the administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg). The results are presented as the mean ± SEM for six mice; * <span class="html-italic">p</span> &lt; 0.001 compared to Veh; † <span class="html-italic">p</span> &lt; 0.001 compared to morphine.</p>
Full article ">Figure 9
<p>(<b>A</b>) SIRT3 inhibition induces acetylation on mitochondrial proteins during morphine tolerance in mice as shown by WES methodology. Co-administration of BPF (25 mg/kg) or MnTBAP (10 mg/kg) attenuated mitochondrial proteins acetylation. No statistically significant difference in TOM20 value was observed between the lanes under these conditions. (<b>A</b>,<b>B</b>) Lanes and (<b>D</b>) electropherogram are representative of results from six animals (green lanes: morphine groups; blue lanes: BPF group; dark grey lanes: MnTBAP group; and light grey lanes: vehicle group). (<b>C</b>) Densitometric analyses for all animals in each group are reported. Values are presented as the mean ± SEM for six mice. * <span class="html-italic">p</span> &lt; 0.05 vs Veh; † <span class="html-italic">p</span> &lt; 0.05 vs. vehicle + morphine.</p>
Full article ">
10 pages, 831 KiB  
Article
Hyperglycemia Stimulates the Irreversible Catabolism of Branched-Chain Amino Acids and Generation of Ketone Bodies by Cultured Human Astrocytes
by Eduard Gondáš, Eva Baranovičová, Jakub Šofranko and Radovan Murín
Biomedicines 2024, 12(8), 1803; https://doi.org/10.3390/biomedicines12081803 - 8 Aug 2024
Viewed by 227
Abstract
Astrocytes are considered to possess a noticeable role in brain metabolism and, as a partners in neuron–glia cooperation, to contribute to the synthesis, bioconversion, and regulation of the flux of substrates for neuronal metabolism. With the aim of investigating to what extent human [...] Read more.
Astrocytes are considered to possess a noticeable role in brain metabolism and, as a partners in neuron–glia cooperation, to contribute to the synthesis, bioconversion, and regulation of the flux of substrates for neuronal metabolism. With the aim of investigating to what extent human astrocytes are metabolizing amino acids and by which compounds are they enriching their surroundings, we employed a metabolomics analysis of their culture media by 1H-NMR. In addition, we compared the composition of media with either 5 mM or 25 mM glucose. The quantitative analysis of culture media by 1H-NMR revealed that astrocytes readily dispose from their milieu glutamine, branched-chain amino acids, and pyruvate with significantly high rates, while they enrich the culture media with lactate, branched-chain keto acids, citrate, acetate, ketone bodies, and alanine. Hyperglycemia suppressed the capacity of astrocytes to release branched-chain 2-oxo acids, while stimulating the generation of ketone bodies. Our results highlight the active involvement of astrocytes in the metabolism of several amino acids and the regulation of key metabolic intermediates. The observed metabolic activities of astrocytes provide valuable insights into their roles in supporting neuronal function, brain metabolism, and intercellular metabolic interactions within the brain. Understanding the complex metabolic interactions between astrocytes and neurons is essential for elucidating brain homeostasis and the pathophysiology of neurological disorders. The observed metabolic activities of astrocytes provide hints about their putative metabolic roles in brain metabolism. Full article
(This article belongs to the Section Neurobiology and Clinical Neuroscience)
Show Figures

Figure 1

Figure 1
<p>Representative <sup>1</sup>H-NMR spectra of branched-chain amino acids (BCAAs), leucine, isoleucine, and valine (<b>A</b>); branched-chain keto acids (BCKAs), 2-oxoisocaproate (<b>B</b>), 2-oxoisovalerate, and 3-methyl-2-oxopentanonate (<b>C</b>), as well as of acetone (<b>D</b>). Spectra were recorded for culture media supplemented with 10% FBS before incubation (black line) or after 24 h incubation of astrocytes in media either with 5 mM glucose (green line) or 25 mM (blue line) initial glucose level. The ranges of chemical shifts (δ) for BCAA, BCKA, and acetone are depicted.</p>
Full article ">Figure 2
<p>The quantification of the specific uptake or release of metabolites from culture medium either with 5 mM glucose (black column) or with 25 mM (white column) supplemented with 10% FBS on cultured human astrocytes by <sup>1</sup>H-NMR analysis, except 3-hydroxybutyrate (3-OHB), which has been estimated enzymatically. Both groups were incubated for 24 h. The specific release (negative value) and uptake of the metabolites from the culture medium were quantified by standardizing them to the total cellular protein mass and incubation time (nmol*h<sup>−1</sup>*mg<sup>−1</sup>). The estimation of the specific uptake of leucin (Leu), isoleucine (Ile), and valine (Val) (<b>A</b>) from the culture medium; the specific release of BCKA, namely, α-ketoisocaproate (KIC), α-keto-methylvalerate (KMV), and α-ketoisovalerate (KIV) (<b>B</b>); the specific release of 3-OHB, and acetone (<b>C</b>) into the culture medium; and the calculation of KIC/leu, KMV/Ile, and BCKA/BCAA ratios (<b>D</b>) are depicted. All values represent the mean ± SEM from six independent experiments, and * represents the value of <span class="html-italic">p</span> ≤ 0.05, and ** <span class="html-italic">p</span> ≤ 0.01.</p>
Full article ">
19 pages, 956 KiB  
Article
Feed Restriction in Angus Steers Impacts Ruminal Bacteria, Its Metabolites, and Causes Epithelial Inflammation
by Qianming Jiang, Matheus Castilho Galvão, Abdulrahman S. Alharthi, Ibrahim A. Alhidary, Mateus P. Gionbelli, Joshua C. McCann and Juan J. Loor
Ruminants 2024, 4(3), 387-405; https://doi.org/10.3390/ruminants4030028 - 3 Aug 2024
Viewed by 267
Abstract
We identified alterations in the ruminal microbiome, metabolome, and epithelial inflammatory response due to moderate feed restriction (FR). Ruminal digesta and epithelial biopsies from seven ruminally cannulated Angus steers were initially collected during ad libitum access to feed (PRE). After a 10 day [...] Read more.
We identified alterations in the ruminal microbiome, metabolome, and epithelial inflammatory response due to moderate feed restriction (FR). Ruminal digesta and epithelial biopsies from seven ruminally cannulated Angus steers were initially collected during ad libitum access to feed (PRE). After a 10 day recovery, steers underwent a 3-day FR period (FRP) at 25% intake of PRE followed by a 15 day recovery (POST) phase with ad libitum access to feed. At the end of FRP and POST, ruminal digesta and epithelial biopsies were collected again for microbial DNA and tissue RNA extraction. RT-qPCR was applied for relative microbial abundance and RNA extraction. Metabolite profiling of digesta was performed via GC-MS. The abundance of Succinivibrio dextrinosolvens, Streptococcus bovis, and Bifidobacteria spp. (N124) was higher (p < 0.05) during FRP than PRE and POST, while Lactobacillus spp. (C25), Escherichia coli (EC42405), Fibrobacter succinogenes, and Megaspheara elsdenii abundances were lower in FRP than PRE (p < 0.05). The TNF and TLR2 mRNA abundance was greater in FRP than PRE (p < 0.05). Among 15 detected amino acids, glutamine, isoleucine, lysine, phenylalanine, threonine, and valine were lower (p < 0.05) in FRP than PRE. Metabolite pathway analysis revealed alterations in amino acid, fatty acid, vitamin, and energy metabolism during FRP (p < 0.05). The mRNA of the proinflammatory genes TNF and TLR2 in the epithelium peaked (p < 0.05) at FRP and remained higher at POST. Results indicated that a short FR influenced ruminal bacteria, reduced concentrations of most metabolites, and triggered an inflammatory response. Full article
Show Figures

Figure 1

Figure 1
<p>Relative universal 1 and 2 16S rRNA copies per µL of ruminal digesta from beef steers (<span class="html-italic">n</span> = 7) before (PRE) and during a 3-day feed-restriction period (FRP) at 25% intake of PRE and after a 15-day recovery (POST). <sup>a,b</sup> Means differ (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">Figure 2
<p>mRNA abundance (relative units) of the proinflammatory genes tumor necrosis factor (<span class="html-italic">TNF</span>) and toll-like receptor 2 (<span class="html-italic">TLR2</span>) in ruminal epithelium from beef steers during a pre-feed-restriction period (PRE), a 3-day 25% feed-restriction period (FRP), and a post-feed-restriction period (POST) (<span class="html-italic">n</span> = 7). <sup>a,b</sup> Means differ (<span class="html-italic">p</span> ≤ 0.05).</p>
Full article ">
22 pages, 6491 KiB  
Article
Formononetin Defeats Multidrug-Resistant Cancers by Induction of Oxidative Stress and Suppression of P-Glycoprotein
by Ying-Tzu Chang, I-Ting Wu, Ming-Jyh Sheu, Yu-Hsuan Lan and Chin-Chuan Hung
Int. J. Mol. Sci. 2024, 25(15), 8471; https://doi.org/10.3390/ijms25158471 - 2 Aug 2024
Viewed by 352
Abstract
Multidrug resistance (MDR) remains the most difficult problem facing conventional chemotherapy for cancers. Astragalus membranaceus is a historically traditional Chinese medicine. One of its bioactive components, formononetin, exhibits antitumor effects on various cancers. However, the effects of formononetin on MDR cancers have not [...] Read more.
Multidrug resistance (MDR) remains the most difficult problem facing conventional chemotherapy for cancers. Astragalus membranaceus is a historically traditional Chinese medicine. One of its bioactive components, formononetin, exhibits antitumor effects on various cancers. However, the effects of formononetin on MDR cancers have not been evaluated. Therefore, we investigated the defense’s effects of formononetin on MDR. We used rhodamine 123 and doxorubicin efflux assays to analyze the inhibition kinetics of P-glycoprotein (P-gp) mediated-efflux. Cell viability was detected by sulforhodamine B assay, and the synergistic effects of formononetin combined with chemotherapeutic agents were further calculated using CompuSyn software. Molecular docking was performed with iGEMDOCK. We discovered that formononetin considerably induced oxidative stress and the disruption of mitochondrial membrane potential in MDR cancer cells. Furthermore, formononetin inhibits the P-gp efflux function by ATPase stimulation and the uncompetitive inhibition of P-gp-mediated effluxes of rhodamine 123 and doxorubicin. The molecular docking model indicates that formononetin may bind to P-gp by strong hydrogen bonds at Arginine (Arg) 489 and Glutamine (Gln) 912. Formononetin exhibits significant synergistic effects with vincristine and doxorubicin toward MDR cancer cells, and it synergistically suppressed tumor growth in vivo with paclitaxel. These results suggest that formononetin should be seen as a potential candidate for the adjuvant therapy of MDR cancers. Full article
(This article belongs to the Section Molecular Oncology)
Show Figures

Figure 1

Figure 1
<p>Inhibitory effect of formononetin on P-gp efflux of <span class="html-italic">ABCB1</span>/Flp-In<sup>TM</sup>-293 cells. (<b>A</b>) Chemical structure of formononetin. (<b>B</b>) P-gp overexpressing <span class="html-italic">ABCB1</span>/Flp-In<sup>TM</sup>-293 cells were incubated with serial-dose formononetin or verapamil, a positive control, for 30 min. The intracellular calcein fluorescence was determined to indicate the inhibition of P-gp. Each data are expressed as the mean ± standard error of at least two experiments, each performed in triplicate. Verapamil has used as positive control of P-gp inhibitor. The abbreviations of verapamil and formononetin are VER and FMN, respectively. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control group.</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>B</b>) The effects of FMN on P-gp ATPase activity was detected and data were analyzed as RLUs. After P-gp membranes were incubated with or without FMN, the unmetabolized ATP transformed into a luminescence. FMN could significantly stimulate basal and verapamil-stimulated P-gp ATPase activity. Verapamil (200 µM) was used as positive control. (<b>A</b>,<b>B</b>) Data were presented as the difference between Na3VO<sub>4</sub>-treated samples. Each data are expressed as the mean ± standard error of at least two experiments, each performed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control group. (<b>C</b>) P-gp substrates were identified by MDR1 shift assay, UIC2 fluorescence was increased during the binding of substrate on P-gp. Vinblastine (22.5 µM), a P-gp standard substrate, was used as positive control. The abbreviations of verapamil and formononetin are VER and FMN, respectively.</p>
Full article ">Figure 3
<p>The kinetic interactions of formononetin on rhodamine 123 or doxorubicin. Michaelis–Menten kinetics of P-gp efflux were determined by the extracellular fluorescence of rhodamine 123 (<b>A</b>) or doxorubicin (<b>C</b>). Lineweaver–Burk plot analysis of formononetin inhibitory mechanism on rhodamine 123 and doxorubicin efflux is shown in (<b>B</b>,<b>D</b>), respectively. Each datum is expressed as the mean ± standard error of at least two experiments, each performed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to nontreatment control.</p>
Full article ">Figure 3 Cont.
<p>The kinetic interactions of formononetin on rhodamine 123 or doxorubicin. Michaelis–Menten kinetics of P-gp efflux were determined by the extracellular fluorescence of rhodamine 123 (<b>A</b>) or doxorubicin (<b>C</b>). Lineweaver–Burk plot analysis of formononetin inhibitory mechanism on rhodamine 123 and doxorubicin efflux is shown in (<b>B</b>,<b>D</b>), respectively. Each datum is expressed as the mean ± standard error of at least two experiments, each performed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to nontreatment control.</p>
Full article ">Figure 4
<p>Effects of formononetin combined with chemotherapeutic agents on cell viabilities of drug-sensitive HeLaS3 cells and MDR KBvin cells. (<b>A</b>,<b>B</b>) The combination effects of formononetin and vincristine or doxorubicin on cell viability by SRB assay in cervical cancer HeLaS3 cells and MDR KBvin cells. Cells were pretreated with drugs alone or compound drug for 72 h. (<b>C</b>) Normalized isobologram for non-constant ratio combination of formononetin and chemotherapeutic drugs. The co-treatment of vincristine or doxorubicin with formononetin for 72 h at different combined concentrations. The actual drug dose was normalized with its corresponding IC<sub>50</sub> and used to determine the synergetic effect of the co-treatments in KBvin cells. The line on the isobologram denotes the half effect from each drug. Antagonism, additive, or synergism effects were indicated above, on, or below the line, respectively. ⊡ formononetin 25 µg/mL + vincristine 1000 nM/doxorubicin 10,000 nM ⊙ formononetin 25 µg/mL + vincristine 100 nM/doxorubicin 1000 nM; ▽ formononetin 20 µg/mL + vincristine 1000 nM; △ formononetin 20 µg/mL + vincristine 100 nM; ⟐ formononetin 20 µg/mL + vincristine 1000 nM; ╳ formononetin 20 µg/mL + vincristine 1000 nM. Data presented as mean ± SE of at least two experiments, each in duplicate. * indicates <span class="html-italic">p</span> value &lt; 0.05 compared with doxorubicin only or vincristine-only group. The abbreviations of vincristine, doxorubicin, and formononetin are VIN, DOX, and FMN, respectively.</p>
Full article ">Figure 4 Cont.
<p>Effects of formononetin combined with chemotherapeutic agents on cell viabilities of drug-sensitive HeLaS3 cells and MDR KBvin cells. (<b>A</b>,<b>B</b>) The combination effects of formononetin and vincristine or doxorubicin on cell viability by SRB assay in cervical cancer HeLaS3 cells and MDR KBvin cells. Cells were pretreated with drugs alone or compound drug for 72 h. (<b>C</b>) Normalized isobologram for non-constant ratio combination of formononetin and chemotherapeutic drugs. The co-treatment of vincristine or doxorubicin with formononetin for 72 h at different combined concentrations. The actual drug dose was normalized with its corresponding IC<sub>50</sub> and used to determine the synergetic effect of the co-treatments in KBvin cells. The line on the isobologram denotes the half effect from each drug. Antagonism, additive, or synergism effects were indicated above, on, or below the line, respectively. ⊡ formononetin 25 µg/mL + vincristine 1000 nM/doxorubicin 10,000 nM ⊙ formononetin 25 µg/mL + vincristine 100 nM/doxorubicin 1000 nM; ▽ formononetin 20 µg/mL + vincristine 1000 nM; △ formononetin 20 µg/mL + vincristine 100 nM; ⟐ formononetin 20 µg/mL + vincristine 1000 nM; ╳ formononetin 20 µg/mL + vincristine 1000 nM. Data presented as mean ± SE of at least two experiments, each in duplicate. * indicates <span class="html-italic">p</span> value &lt; 0.05 compared with doxorubicin only or vincristine-only group. The abbreviations of vincristine, doxorubicin, and formononetin are VIN, DOX, and FMN, respectively.</p>
Full article ">Figure 5
<p>Mechanism of formononetin MDR reversal ability on cancer cells. (<b>A</b>) The effects of formononetin on vincristine-induced cytotoxicity was assessed by apoptosis assay. Apoptotic cells were stained with 5 µL of Annexin V–FITC and propidium iodide (PI) and analyzed by flow cytometry. (<b>B</b>) <span class="html-italic">ABCB1</span> mRNA expression was determined by real-time RT PCR. Cells were pretreated with formononetin 10 µg/mL or 25 µg/mL in HeLaS3 and KBvin for 72 h. Each datum is expressed as the mean ± standard error of at least two experiments, each performed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control group. The abbreviations of vincristine and formononetin are VIN and FMN, respectively.</p>
Full article ">Figure 5 Cont.
<p>Mechanism of formononetin MDR reversal ability on cancer cells. (<b>A</b>) The effects of formononetin on vincristine-induced cytotoxicity was assessed by apoptosis assay. Apoptotic cells were stained with 5 µL of Annexin V–FITC and propidium iodide (PI) and analyzed by flow cytometry. (<b>B</b>) <span class="html-italic">ABCB1</span> mRNA expression was determined by real-time RT PCR. Cells were pretreated with formononetin 10 µg/mL or 25 µg/mL in HeLaS3 and KBvin for 72 h. Each datum is expressed as the mean ± standard error of at least two experiments, each performed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control group. The abbreviations of vincristine and formononetin are VIN and FMN, respectively.</p>
Full article ">Figure 6
<p>The docking results showed the superimposition of docked poses of compounds in the P-gp-binding pocket of the 3D structure of formononetin, rhodamine 123, and doxorubicin on P-gp: (<b>A</b>) formononetin, (<b>B</b>) rhodamine 123, (<b>C</b>) doxorubicin.</p>
Full article ">Figure 6 Cont.
<p>The docking results showed the superimposition of docked poses of compounds in the P-gp-binding pocket of the 3D structure of formononetin, rhodamine 123, and doxorubicin on P-gp: (<b>A</b>) formononetin, (<b>B</b>) rhodamine 123, (<b>C</b>) doxorubicin.</p>
Full article ">Figure 7
<p>Intracellular ROS production and mitochondria membrane potential changes in the MDR KBvin cells. The intracellular ROS production was detected in the HeLaS3 (<b>A</b>) and KBvin cells (<b>B</b>). Formononetin with or without chemotherapeutic drugs were treated for 1 h. Menadione was used as a positive control. The mitochondria membrane potential changes were measured in the HeLaS3 (<b>C</b>) and KBvin cells (<b>D</b>). The cells were treated with formononetin or chemotherapeutic drugs only or formononetin combined with chemotherapeutic drug for 6 h. Menadione was used as a positive control. Each datum is expressed as the mean ± standard error of at least two experiments, each per-formed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control. N-acetylcysteine was applied as the antioxidant, and the cell viabilities were further evaluated by SRB assay in HeLaS3 (<b>E</b>) and KBvin cells (<b>F</b>). * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to with doxorubicin only or vincristine only group. # indicates <span class="html-italic">p</span> &lt; 0.05 compared with formononetin combined doxorubicin or vincristine.</p>
Full article ">Figure 7 Cont.
<p>Intracellular ROS production and mitochondria membrane potential changes in the MDR KBvin cells. The intracellular ROS production was detected in the HeLaS3 (<b>A</b>) and KBvin cells (<b>B</b>). Formononetin with or without chemotherapeutic drugs were treated for 1 h. Menadione was used as a positive control. The mitochondria membrane potential changes were measured in the HeLaS3 (<b>C</b>) and KBvin cells (<b>D</b>). The cells were treated with formononetin or chemotherapeutic drugs only or formononetin combined with chemotherapeutic drug for 6 h. Menadione was used as a positive control. Each datum is expressed as the mean ± standard error of at least two experiments, each per-formed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control. N-acetylcysteine was applied as the antioxidant, and the cell viabilities were further evaluated by SRB assay in HeLaS3 (<b>E</b>) and KBvin cells (<b>F</b>). * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to with doxorubicin only or vincristine only group. # indicates <span class="html-italic">p</span> &lt; 0.05 compared with formononetin combined doxorubicin or vincristine.</p>
Full article ">Figure 7 Cont.
<p>Intracellular ROS production and mitochondria membrane potential changes in the MDR KBvin cells. The intracellular ROS production was detected in the HeLaS3 (<b>A</b>) and KBvin cells (<b>B</b>). Formononetin with or without chemotherapeutic drugs were treated for 1 h. Menadione was used as a positive control. The mitochondria membrane potential changes were measured in the HeLaS3 (<b>C</b>) and KBvin cells (<b>D</b>). The cells were treated with formononetin or chemotherapeutic drugs only or formononetin combined with chemotherapeutic drug for 6 h. Menadione was used as a positive control. Each datum is expressed as the mean ± standard error of at least two experiments, each per-formed in triplicate. * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to control. N-acetylcysteine was applied as the antioxidant, and the cell viabilities were further evaluated by SRB assay in HeLaS3 (<b>E</b>) and KBvin cells (<b>F</b>). * denotes <span class="html-italic">p</span> &lt; 0.05 as compared to with doxorubicin only or vincristine only group. # indicates <span class="html-italic">p</span> &lt; 0.05 compared with formononetin combined doxorubicin or vincristine.</p>
Full article ">Figure 8
<p>Formononetin combined with paclitaxel suppressed MDR KBvin cell growth in a xenotransplantation model. (<b>A</b>) Formononetin did not express significant toxicity at 48 hpi. (<b>B</b>) In an MDR KBvin xenograft model, we demonstrated that formononetin synergistically suppressed tumor size in combination with paclitaxel. The intensity of red fluorescence is proportional to the tumor size. Scale bar represents 1 mm. * <span class="html-italic">p</span> &lt; 0.05 compared with the control; @ <span class="html-italic">p</span> &lt; 0.05 compared to paclitaxel only group. hpf: hours post-fertilization; hpi: hours post-treatment or post-injection. The abbreviations of paclitaxel and formononetin are PXL and FMN, respectively.</p>
Full article ">Figure 8 Cont.
<p>Formononetin combined with paclitaxel suppressed MDR KBvin cell growth in a xenotransplantation model. (<b>A</b>) Formononetin did not express significant toxicity at 48 hpi. (<b>B</b>) In an MDR KBvin xenograft model, we demonstrated that formononetin synergistically suppressed tumor size in combination with paclitaxel. The intensity of red fluorescence is proportional to the tumor size. Scale bar represents 1 mm. * <span class="html-italic">p</span> &lt; 0.05 compared with the control; @ <span class="html-italic">p</span> &lt; 0.05 compared to paclitaxel only group. hpf: hours post-fertilization; hpi: hours post-treatment or post-injection. The abbreviations of paclitaxel and formononetin are PXL and FMN, respectively.</p>
Full article ">
15 pages, 3602 KiB  
Article
Flux Calculation for Primary Metabolism Reveals Changes in Allocation of Nitrogen to Different Amino Acid Families When Photorespiratory Activity Changes
by Nils Friedrichs, Danial Shokouhi and Arnd G. Heyer
Int. J. Mol. Sci. 2024, 25(15), 8394; https://doi.org/10.3390/ijms25158394 - 1 Aug 2024
Viewed by 270
Abstract
Photorespiration, caused by oxygenation of the enzyme Rubisco, is considered a wasteful process, because it reduces photosynthetic carbon gain, but it also supplies amino acids and is involved in amelioration of stress. Here, we show that a sudden increase in photorespiratory activity not [...] Read more.
Photorespiration, caused by oxygenation of the enzyme Rubisco, is considered a wasteful process, because it reduces photosynthetic carbon gain, but it also supplies amino acids and is involved in amelioration of stress. Here, we show that a sudden increase in photorespiratory activity not only reduced carbon acquisition and production of sugars and starch, but also affected diurnal dynamics of amino acids not obviously involved in the process. Flux calculations based on diurnal metabolite profiles suggest that export of proline from leaves increases, while aspartate family members accumulate. An immense increase is observed for turnover in the cyclic reaction of glutamine synthetase/glutamine-oxoglutarate aminotransferase (GS/GOGAT), probably because of increased production of ammonium in photorespiration. The hpr1-1 mutant, defective in peroxisomal hydroxypyruvate reductase, shows substantial alterations in flux, leading to a shift from the oxoglutarate to the aspartate family of amino acids. This is coupled to a massive export of asparagine, which may serve in exchange for serine between shoot and root. Full article
(This article belongs to the Special Issue Plant Respiration in the Light and Photorespiration)
Show Figures

Figure 1

Figure 1
<p>Diurnal profiles of metabolite concentrations in µmol g<sup>−1</sup> fresh weight over time in hours from light on for metabolites that showed significant genotype × daytime interaction in a two-way ANOVA (<span class="html-italic">p</span> &lt; 0.05, n = 5) at least under one CO<sub>2</sub> treatment. Graphs show diurnal dynamics for Col-0 (black dashed lines) and the <span class="html-italic">hpr1-1</span> mutant (red lines for eCO<sub>2</sub>; blue lines for aCO<sub>2</sub>) at elevated CO<sub>2</sub> (<b>A</b>–<b>G</b>) and after a sudden shift to ambient CO<sub>2</sub> level (<b>H</b>–<b>N</b>). Lines connect means of measurements at time points 0, 2, 4, 6, 8, 16 and 24 h starting from light on. Light off was at 8 h. Error bars show standard error of the mean (n = 5). Results of the ANOVA for all metabolites are given in <a href="#app1-ijms-25-08394" class="html-app">Supplemental S1 and S2</a>.</p>
Full article ">Figure 2
<p>Model used for simulating primary metabolism. Black arrows show routes of carbon; dashed arrows show routes of nitrogen. Gly: glycine, Ser: serine; HP: sugar phosphates, ex4: export of Ser; AA1: pool of cysteine, pyruvate family and aromatic amino acids; MF: pool of malate and fumarate; exp1: sugar export, exp3: export of AA1, Cit: citrate, Asf: aspartate family; KgF: oxoglutarate family; Glu: glutamate; αKG: oxoglutarate, exp5: export of Asf; exp2: export of KgF, Gln: glutamine, NH4<sup>+</sup>: ammonium. Reactions 1: net photosynthesis; 2: starch build and degradation; 3: synthesis of sucrose, including hydrolysis to hexoses (sucrose cycling); 4: export; 5: synthesis of AA1; 6: export; 7: photorespiratory Gly production; 8. Gly decarboxylation and Ser synthesis; 9: hydroxypyruvate reduction; 10: export or import of Ser; 11: synthesis of malate and fumarate; 12: respiration; 13: synthesis of citrate; 14: synthesis of Asf; 15: export; 16, 17: GS/GOGAT cycle; 18: synthesis of KgF; 19: export.</p>
Full article ">Figure 3
<p>Measured and simulated metabolite levels in µmol g<sup>−1</sup> fresh weight for Col-0 (black) and <span class="html-italic">hpr1-1</span> (red) plants grown constantly at elevated CO<sub>2</sub> level (1000 ppm). Closed circles at time points 0, 2, 4, 6, 8, 16 and 24 are means of five-fold replication. Open circles were created by a smoothing spline (see <a href="#sec2-ijms-25-08394" class="html-sec">Section 2</a>). Lines are results of simulations with the best fitting parameter set. For abbreviations of metabolites, see <a href="#ijms-25-08394-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 4
<p>Measured and simulated metabolite levels in µmol g<sup>−1</sup> fresh weight for Col-0 (black) and <span class="html-italic">hpr1-1</span> (blue) plants shifted from elevated to ambient CO<sub>2</sub> level (~450 ppm). Closed circles at time points 0, 2, 4, 6, 8, 16 and 24 are means of five-fold replication. Open circles were created by a smoothing spline (see <a href="#sec2-ijms-25-08394" class="html-sec">Section 2</a>). Lines are results of simulations with the best fitting parameter set. For abbreviations of metabolites, see <a href="#ijms-25-08394-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>Calculated flux trajectories in µmol g<sup>−1</sup> h<sup>−1</sup> for Col-0 (dashed black lines) and <span class="html-italic">hpr1-1</span> (red lines) constantly grown at elevated CO<sub>2</sub> level (1000 ppm). All fluxes are termed based on reaction products, except f_hpr, which is the HPR reaction, and f_shmt, which is a proxy for reactions leading from Ser to Gly. All exports out of the model scope have the prefix ex_. Numbers in parentheses refer to the numbering of reactions in <a href="#ijms-25-08394-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 6
<p>Calculated flux trajectories in µmol g-1 h-1 for Col-0 (dashed black lines) and hpr1-1 (blue lines) shifted from elevated to ambient CO<sub>2</sub> level (~450 ppm). All fluxes are termed based on reaction products, except f_hpr, which is the HPR reaction, and f_shmt, which is a proxy for reactions leading from Ser to Gly. All exports out of the model scope have the prefix ex_. Numbers in parentheses refer to the numbering of reactions in <a href="#ijms-25-08394-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 7
<p>Schemes of carbon allocation during day (<b>A</b>–<b>D</b>) and night (<b>E</b>–<b>H</b>) for Col-0 (right) and hpr1-1 (left). (<b>A</b>,<b>B</b>,<b>E</b>,<b>F</b>): eCO<sub>2</sub>; (<b>C</b>,<b>D</b>,<b>G</b>,<b>H</b>): aCO<sub>2</sub>. The width of arrows indicates flux into a metabolite pool integrated over day and night, respectively.</p>
Full article ">
17 pages, 3189 KiB  
Article
Functional Insights into the Sphingolipids C1P, S1P, and SPC in Human Fibroblast-like Synoviocytes by Proteomic Analysis
by Thomas Timm, Christiane Hild, Gerhard Liebisch, Markus Rickert, Guenter Lochnit and Juergen Steinmeyer
Int. J. Mol. Sci. 2024, 25(15), 8363; https://doi.org/10.3390/ijms25158363 - 31 Jul 2024
Viewed by 316
Abstract
The (patho)physiological function of the sphingolipids ceramide-1-phosphate (C1P), sphingosine-1-phosphate (S1P), and sphingosylphosphorylcholine (SPC) in articular joints during osteoarthritis (OA) is largely unknown. Therefore, we investigated the influence of these lipids on protein expression by fibroblast-like synoviocytes (FLSs) from OA knees. Cultured human FLSs [...] Read more.
The (patho)physiological function of the sphingolipids ceramide-1-phosphate (C1P), sphingosine-1-phosphate (S1P), and sphingosylphosphorylcholine (SPC) in articular joints during osteoarthritis (OA) is largely unknown. Therefore, we investigated the influence of these lipids on protein expression by fibroblast-like synoviocytes (FLSs) from OA knees. Cultured human FLSs (n = 7) were treated with 1 of 3 lipid species—C1P, S1P, or SPC—IL-1β, or with vehicle. The expression of individual proteins was determined by tandem mass tag peptide labeling followed by high-resolution electrospray ionization (ESI) mass spectrometry after liquid chromatographic separation (LC-MS/MS/MS). The mRNA levels of selected proteins were analyzed using RT-PCR. The 3sphingolipids were quantified in the SF of 18 OA patients using LC-MS/MS. A total of 4930 proteins were determined using multiplex MS, of which 136, 9, 1, and 0 were regulated both reproducibly and significantly by IL-1β, C1P, S1P, and SPC, respectively. In the presence of IL-1ß, all 3 sphingolipids exerted ancillary effects. Only low SF levels of C1P and SPC were found. In conclusion, the 3 lipid species regulated proteins that have not been described in OA. Our results indicate that charged multivesicular body protein 1b, metal cation symporter ZIP14, glutamine-fructose-6-P transaminase, metallothionein-1F and -2A, ferritin, and prosaposin are particularly interesting proteins due to their potential to affect inflammatory, anabolic, catabolic, and apoptotic mechanisms. Full article
(This article belongs to the Special Issue Proteomics and Its Applications in Disease 3.0)
Show Figures

Figure 1

Figure 1
<p>Venn diagram of proteins that are reproducibly regulated by (<b>A</b>) C1P (red), C1P in the presence of IL-1ß (yellow), and IL-1ß alone (green); (<b>B</b>) S1P (blue), S1P in the presence of IL-1ß (orange), and IL-1ß alone (green); (<b>C</b>) SPC in the presence of IL-1ß (violet) and IL-1ß alone (green); and (<b>D</b>) IL-1ß in the presence of C1P (yellow), S1P (orange), or SPC (violet). The number of proteins that can be reproducibly regulated is shown. In <a href="#app1-ijms-25-08363" class="html-app">Tables S1 and S2</a>, the AR of each protein is listed together with the results of the statistical analysis.</p>
Full article ">Figure 2
<p>C1P and S1P reproducibly and significantly regulate the level of proteins in FLSs. Protein levels of (<b>A</b>) stromelysin, charged multivesicular body protein 1b, metal cation symporter ZIP14; (<b>B</b>) cPLA2, cytochrome c oxidase subunit 1, SOD; and (<b>C</b>) long-chain-fatty-acid–CoA ligase 4, ICAM-1, and glutamine fructose-6-Ptransaminase were quantified by MS in duplicate in the 7 biological replicates. The dot plots represent the data obtained from the resulting 14 replicates and illustrate the x-fold abundance of the proteins in the treated FLS cells in comparison to that of only vehicle-treated controls (which are normalized to 1 and shown as a dotted line). The mean value ± SD is represented by lines within each figure (<span class="html-italic">n</span> = 7). * 0.05 ≥ <span class="html-italic">p</span> &gt; 0.01; ** 0.01 ≥ <span class="html-italic">p</span> &gt; 0.001; *** <span class="html-italic">p</span> ≥ 0.001.</p>
Full article ">Figure 3
<p>The biological processes of FLS being altered by C1P. The 9 proteins were reproducibly upregulated by more than 1.2-fold by C1P in FLS during 48 h of treatment, and <a href="#app1-ijms-25-08363" class="html-app">Table S1</a> provides further data on these proteins. The Go Slim categories for proteins were generated by Proteome Discoverer 2.5 software using the Gene Ontology (GO) database.</p>
Full article ">Figure 4
<p>The molecular functions of FLS being altered by C1P. The 9 proteins were reproducibly upregulated by more than 1.2-fold by C1P in FLS during 48 h of treatment, and <a href="#app1-ijms-25-08363" class="html-app">Table S1</a> provides further data on these proteins. The Go Slim categories for proteins were generated by Proteome Discoverer 2.5 software using the Gene Ontology (GO) database.</p>
Full article ">Figure 5
<p>The cellular localization of 9 proteins being regulated by C1P. The 9 proteins were reproducibly upregulated by more than 1.2-fold by C1P in FLS during 48 h of treatment, and <a href="#app1-ijms-25-08363" class="html-app">Table S1</a> provides further data on these proteins. The Go Slim categories for proteins were generated by Proteome Discoverer 2.5 software using the Gene Ontology (GO) database.</p>
Full article ">Figure 6
<p>Levels of C1P 16:0 ad SPC 18:1;O2 in knee SF of patients with early-stage or late-stage knee OA. Sphingolipids were quantified by LC-MS/MS in extracts of SF obtained from 8 patients with eOA (black circle) and 10 patients with lOA (blue open circle). Data presented indicate mean ± SD of lipid concentration in SF (N = 8 or 10).</p>
Full article ">
18 pages, 17205 KiB  
Article
Circadian Rhythm and Nitrogen Metabolism Participate in the Response of Boron Deficiency in the Root of Brassica napus
by Ling Liu, Xianjie Duan, Haoran Xu, Peiyu Zhao, Lei Shi, Fangsen Xu and Sheliang Wang
Int. J. Mol. Sci. 2024, 25(15), 8319; https://doi.org/10.3390/ijms25158319 - 30 Jul 2024
Viewed by 298
Abstract
Boron (B) deficiency has been shown to inhibit root cell growth and division. However, the precise mechanism underlying B deficiency-mediated root tip growth inhibition remains unclear. In this study, we investigated the role of BnaA3.NIP5;1, a gene encoding a boric acid channel, [...] Read more.
Boron (B) deficiency has been shown to inhibit root cell growth and division. However, the precise mechanism underlying B deficiency-mediated root tip growth inhibition remains unclear. In this study, we investigated the role of BnaA3.NIP5;1, a gene encoding a boric acid channel, in Brassica napus (B. napus). BnaA3.NIP5;1 is expressed in the lateral root cap and contributes to B acquisition in the root tip. Downregulation of BnaA3.NIP5;1 enhances B sensitivity in B. napus, resulting in reduced shoot biomass and impaired root tip development. Transcriptome analysis was conducted on root tips from wild-type B. napus (QY10) and BnaA3.NIP5;1 RNAi lines to assess the significance of B dynamics in meristematic cells during seedling growth. Differentially expressed genes (DEGs) were significantly enriched in plant circadian rhythm and nitrogen (N) metabolism pathways. Notably, the circadian-rhythm-related gene HY5 exhibited a similar B regulation pattern in Arabidopsis to that observed in B. napus. Furthermore, Arabidopsis mutants with disrupted circadian rhythm (hy5/cor27/toc1) displayed heightened sensitivity to low B compared to the wild type (Col-0). Consistent with expectations, B deficiency significantly disrupted N metabolism in B. napus roots, affecting nitrogen concentration, nitrate reductase enzyme activity, and glutamine synthesis. Interestingly, this disruption was exacerbated in BnaA3NIP5;1 RNAi lines. Overall, our findings highlight the critical role of B dynamics in root tip cells, impacting circadian rhythm and N metabolism, ultimately leading to retarded growth. This study provides novel insights into B regulation in root tip development and overall root growth in B. napus. Full article
(This article belongs to the Special Issue The Gene, Genomics, and Molecular Breeding in Cruciferae Plants 2.0)
Show Figures

Figure 1

Figure 1
<p>Downregulated expression of <span class="html-italic">BnaA3NIP5;1</span> enhances boron sensitivity. (<b>A</b>) The phenotypic characteristics and statistical analysis were conducted on primary root length and lateral root number in QY10 and NSQ lines cultured on plates for 10 days under 0.1 μM B. The study involved a minimum of three replicates. (<b>B</b>) Additionally, longitudinal sections (<b>C</b>–<b>E</b>) and diameter statistics (<b>F</b>) of the root tips from QY10 and NSQ lines in scenario (<b>A</b>) were examined (<span class="html-italic">n</span> ≥ 3). (<b>G</b>) Furthermore, the phenotypes of QY10 and NSQ lines were observed after 14 days of culture in nutrient solution under 0.25 μM B. (<b>H</b>–<b>J</b>) Microstructural analysis of the root tips from QY10 and NSQ lines (in scenario (<b>G</b>)) revealed the non-root-hair zone (NRHZ) length. (<b>K</b>) The NRHZ length was statistically analyzed (<span class="html-italic">n</span> ≥ 3), and the reported values represent means ± standard deviation. Letters denote significant differences between different treatments based on Duncan’s test (<span class="html-italic">p</span> &lt; 0.05). The abbreviation ‘NSQ’ refers to the <span class="html-italic">BnaA3NIP5;1</span> RNAi lines.</p>
Full article ">Figure 2
<p>Differentially expressed genes (DEGs) among different groups. (<b>A</b>–<b>C</b>) We conducted principal component analysis (PCA), sample correlation analysis, and generated a gene coexpression Venn diagram using data from nine samples. These samples included NB_QY10 (normal boron QY10, 100 μM B), LB_QY110 (low-boron QY10, 0.25 μM B), and LB-NSQ (low-boron NSQ, 0.25 μM B), each with three biological replicates. (<b>D</b>) We quantified the differentially expressed genes (DEGs) between various groups. (<b>E</b>) Additionally, we created a Venn diagram to compare DEGs among LB_QY10 versus NB_QY10, LB_NSQ versus NB_QY10, and LB_NSQ versus LB_QY10. (<b>F</b>) Finally, we determined the count of upregulated and downregulated DEGs in the union of the DEGs identified in scenario (<b>E</b>).</p>
Full article ">Figure 3
<p>qRT-PCR verification transcriptome data results of selected 10 DEGs. The gene expression profiles were assessed using RNA-seq (<b>left panel</b>) and qRT-PCR (<b>right panel</b>) in NB_QY10, LB_QY10, and LB_NSQ. The study involved three pools, each containing 30–50 root tips, and the reported values represent means ± SD.</p>
Full article ">Figure 4
<p>KEGG functional enrichment analysis of DEGs in different groups. (<b>A</b>) KEGG enrichment analysis of DEGs in LB_QY10 vs. NB_QY10. (<b>B</b>) KEGG enrichment analysis of DEGs in LB_NSQ vs. NB_QY10. (<b>C</b>) KEGG enrichment analysis of DEGs in LB_NSQ vs. LB_QY10. (<b>D</b>) KEGG enrichment analysis of DEGs in LB_NSQ vs. LB_QY10 vs NB_QY10. (<b>E</b>) KEGG enrichment analysis of up-regulated DEGs in LB_NSQ vs. LB_QY10 vs NB_QY10. (<b>F</b>) KEGG enrichment analysis of down-regulated DEGs in LB_NSQ vs. LB_QY10 vs NB_QY10. The dot size and color correspond to the number of genes and the corrected <span class="html-italic">p</span>-value, respectively. The gene ratio represents the proportion of differentially expressed genes (DEGs) annotated within a specific pathway term relative to the total number of genes annotated in that term. A higher gene ratio signifies greater pathway intensity. The padj value, ranging from 0 to 1, reflects the corrected <span class="html-italic">p</span>-value, with lower values indicating stronger significance. Only the top 20 enriched pathways are displayed, and the red box highlights the pathways of particular interest.</p>
Full article ">Figure 5
<p>Expression patterns of circadian rhythm pathway genes in <span class="html-italic">B. napus</span> and <span class="html-italic">Arabidopsis</span>. (<b>A</b>) The gene identity (ID), gene annotation (name), and differential expression multiples of circadian rhythm pathway genes were investigated under two conditions: LB_QY10 versus NB_QY10 and LB_NSQ versus NB_QY10 using RNA-seq. (<b>B</b>) The expression pattern of <span class="html-italic">AtHY5</span> in <span class="html-italic">Arabidopsis</span> Col-0 roots was examined after 12 days of growth on agar plates under normal (100 μM B) and low-boron (0.1 μM B) conditions. The study involved three pools, each containing 20 roots, (<span class="html-italic">n</span> = 3). and the reported values represent means ± standard deviation. Letters denote significant differences between different treatments based on Duncan’s test (<span class="html-italic">p</span> &lt; 0.05). Additionally, GUS staining was performed on 7-day-old <span class="html-italic">ProAtHY5: GUS</span> seedlings cultured on agar plates with normal boron ((<b>C</b>,<b>E</b>,<b>G</b>,<b>I</b>,<b>K</b>) (100 μM B)) and boron-deficient ((<b>D</b>,<b>F</b>,<b>H</b>,<b>J</b>,<b>L</b>) (0.1 μM B)) conditions.</p>
Full article ">Figure 6
<p><span class="html-italic">hy5/cor27/toc1</span> mutants are more sensitive to boron deficiency than Col-0. (<b>A</b>) The phenotypes of wild-type (Col-0) and <span class="html-italic">hy5/cor27/toc1</span> mutant plants were observed after 12 days of growth on agar plates under both normal (100 μM B) and boron-deficient (0.1 μM B) conditions. (<b>B</b>) Microstructure images of the root tips from Col-0 and <span class="html-italic">hy5/cor27/toc1</span> mutants were captured under boron deficiency (0.1 μM B) conditions in scenario (<b>A</b>). The root tips were stained with propidium iodide (PI), and the upper row displays the view after PI staining, while the lower row shows the view under bright field. Arrows indicate the location of root hairs. (<b>C</b>) Statistical analysis was performed on primary root length, meristematic zone length, elongation zone length, and the number of lateral roots in Col-0 and <span class="html-italic">hy5/cor27/toc1</span> mutants from scenario (<b>A</b>) (<span class="html-italic">n</span> ≥ 10). The reported values represent means ± standard deviation. Letters denote significant differences between different treatments based on Duncan’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Nitrogen metabolism participates in the response of <span class="html-italic">B. napus</span> roots to B deficiency. (<b>A</b>) The gene identity (ID), gene annotation (name), and differential expression multiples of nitrogen metabolism pathway genes were analyzed in RNA-seq data for LB_QY10 versus NB_QY10 and LB_NSQ versus NB_QY10. (<b>B</b>) A schematic diagram illustrates transcriptome changes in the main pathways of nitrate absorption and assimilation. Genes in green font indicate decreased expression during the low-boron response, while genes in red represent increased expression. (<b>C</b>–<b>F</b>) Nitrogen concentration, nitrate reductase (NR), glutamine synthase (GS), and glutamate dehydrogenase (GDH) were determined in roots of Y10 and NSQ lines cultured in nutrient solution for 14 days under 100 and 0.25 μM B. The values represent means ± SD (standard deviation), and letters indicate significant differences between different treatments based on Duncan’s test (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>A working model illustrating the involvement of circadian rhythms and nitrogen metabolism in B deficiency-mediated root tip growth inhibition. Boron (B) deficiency activates the activity of circadian-rhythm-related transcription factors (<b>on the left</b>) and reduces the transcription level and enzyme activity of nitrogen metabolism-related genes (<b>on the right</b>, where green font represents decrease and red font represents increase). The gray box indicates cells, and the orange circle represents B.</p>
Full article ">
15 pages, 7083 KiB  
Article
Metal Ion Binding to Human Glutaminyl Cyclase: A Structural Perspective
by Giusy Tassone, Cecilia Pozzi and Stefano Mangani
Int. J. Mol. Sci. 2024, 25(15), 8279; https://doi.org/10.3390/ijms25158279 - 29 Jul 2024
Viewed by 260
Abstract
Glutaminyl-peptide cyclotransferases (QCs) convert the N-terminal glutamine or glutamate residues of protein and peptide substrates into pyroglutamate (pE) by releasing ammonia or a water molecule. The N-terminal pE modification protects peptides/proteins against proteolytic degradation by amino- or exopeptidases, increasing their stability. Mammalian QC [...] Read more.
Glutaminyl-peptide cyclotransferases (QCs) convert the N-terminal glutamine or glutamate residues of protein and peptide substrates into pyroglutamate (pE) by releasing ammonia or a water molecule. The N-terminal pE modification protects peptides/proteins against proteolytic degradation by amino- or exopeptidases, increasing their stability. Mammalian QC is abundant in the brain and a large amount of evidence indicates that pE peptides are involved in the onset of neural human pathologies such as Alzheimer’s and Huntington’s disease and synucleinopathies. Hence, human QC (hQC) has become an intensively studied target for drug development against these diseases. Soon after its characterization, hQC was identified as a Zn-dependent enzyme, but a partial restoration of the enzyme activity in the presence of the Co(II) ion was also reported, suggesting a possible role of this metal ion in catalysis. The present work aims to investigate the structure of demetallated hQC and of the reconstituted enzyme with Zn(II) and Co(II) and their behavior in the presence of known inhibitors. Furthermore, our structural determinations provide a possible explanation for the presence of the mononuclear metal binding site of hQC, despite the presence of the same conserved metal binding motifs present in distantly related dinuclear aminopeptidase enzymes. Full article
Show Figures

Figure 1

Figure 1
<p>The chemical structure of PBD-150 and SEN-177.</p>
Full article ">Figure 2
<p>(<b>A</b>) Active site view of apo-hQC-2X (in cartoon, light cyan carbons; residues in sticks). A water molecule, WatC, occupies the metal cofactor pocket, forming H-bonds with Asp159, Glu202, His330, and a further solvent molecule, Wat2. (<b>B</b>) Active site residues and water molecules surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ). Oxygen and nitrogen atoms are colored red and blue, respectively. Water molecules are represented as red spheres and H-bonds as blue dashed lines.</p>
Full article ">Figure 3
<p>Active site view of the superimposition between apo-hQC-2X (in cartoon, light cyan carbons; residues in sticks; water molecules in cyan) and hQC-2X in complex with Zn(II) ions (gold carton, carbons, and water molecules; Zn(II) ion in gray; PDB id 4UY9 [<a href="#B36-ijms-25-08279" class="html-bibr">36</a>]). The comparison shows the high conservation of the catalytic cavity configuration upon metal removal.</p>
Full article ">Figure 4
<p>(<b>A</b>) Active site view of the structure of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (gray sphere). Coordination bonds are colored gray. (<b>B</b>) Zn(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the zinc K-edge (dark gray wire, contoured at 5 σ). (<b>C</b>) Active site view of the structure of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (magenta sphere). Coordination and H-bonds are colored magenta and blue, respectively. (<b>D</b>) Co(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ). Oxygen atoms are colored red and nitrogen blue. Water molecules are represented as red spheres.</p>
Full article ">Figure 5
<p>(<b>A</b>) Binding of PBD-150 (in sticks, orange carbon atoms) to the active site of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (magenta sphere). Coordination and H-bonds are colored magenta and blue, respectively. (<b>B</b>) PBD-150, Co(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ). (<b>C</b>–<b>E</b>) Active site view of PBD-150 and Co(II) ion surrounded by the omit map (dark green wire, contoured at 2.5 σ) and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ), respectively, in chain A (panel (<b>C</b>)), B (panel (<b>D</b>)), and C (panel (<b>E</b>)). Oxygen atoms are colored red, nitrogen blue, and sulfur yellow. Water molecules are represented as red spheres.</p>
Full article ">Figure 6
<p>(<b>A</b>) Binding of SEN-177 (in sticks, green carbon atoms) to the active site of hQC-2X (in cartoon, light cyan carbons; residues in ticks) in complex with Co(II) ions (magenta sphere). Coordination and H-bonds are colored magenta and blue, respectively. (<b>B</b>) SEN-177, Co(II) ion and its coordinating residues are surrounded by the 2<span class="html-italic">F<sub>o</sub></span> − <span class="html-italic">F<sub>c</sub></span> map (blue wire, contoured at 1.5 σ), and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ). (<b>C</b>–<b>E</b>) Active site view of SEN-177 and Co(II) ion surrounded by the omit map (dark green wire, contoured at 2.5 σ) and the anomalous map computed at the cobalt K-edge (purple wire, contoured at 5 σ), respectively, in chain A (panel (<b>C</b>)), B (panel (<b>D</b>)), and C (panel (<b>E</b>)). Oxygen atoms are colored red, nitrogen blue, sulfur yellow, and halogen gray. Water molecules are represented as red spheres.</p>
Full article ">Figure 7
<p>Active site view of the structural comparison between the complexes (<b>A</b>) hQC-2X–Co(II) – PBD-150 (light cyan cartoon and carbons; Co(II) ion as magenta sphere and PBD-150 in sticks, cyan carbons) and hQC-2X–Zn(II)–PBD-150 (gold cartoon and carbons; Zn(II) ion as gray sphere and PBD-150 in sticks, gold carbons; PDB id 4YWY [<a href="#B36-ijms-25-08279" class="html-bibr">36</a>]); (<b>B</b>) hQC-2X–Co(II)–SEN-177 (light cyan cartoon and carbons; Co(II) ion as magenta sphere and SEN-177 in sticks, cyan carbons) and hQC-2X–Zn(II) –SEN-177 (gold cartoon and carbons; Zn(II) ion as gray sphere and SEN-177 in sticks, gold carbons; PDB id 6GBX [<a href="#B32-ijms-25-08279" class="html-bibr">32</a>]). The comparisons highlight the similar binding modes of the inhibitors in the complexes with Zn(II) and Co(II) metal ions.</p>
Full article ">Figure 8
<p>(<b>A</b>) Superimposition of Zn1-loaded hQC-2X (pale cyan sticks and blue labels) with ApAP (green sticks and green labels). Although distant in phylogeny and in sequence, the remarkable structural coincidence of the metal binding sites can readily be appreciated. (<b>B</b>) Surface representation (coded by electrostatic potential) of the active site cavity of ApAP with the two Zn(II) ions (green) protruding in the cavity. (<b>C</b>) Surface representation (coded by electrostatic potential) of the active site cavity of hQC-2X with the Zn(II) ion (white) protruding in the cavity. The described H-bonds between are shown as black dashed lines, while the Ser160-Zn2 distance is shown as red dashed line. The difference in the cavity volume due to the Ser160-Asp248 H-bond can be appreciated.</p>
Full article ">Scheme 1
<p>Schematic reactions catalyzed by hQC. Data regarding the optimum pH for the catalyzed reactions are from Schilling et al. [<a href="#B5-ijms-25-08279" class="html-bibr">5</a>].</p>
Full article ">
20 pages, 4420 KiB  
Article
Plasma Metabolomics Study on the Impact of Different CRF Levels on MetS Risk Factors
by Xiaoxiao Fei, Qiqi Huang and Jiashi Lin
Metabolites 2024, 14(8), 415; https://doi.org/10.3390/metabo14080415 - 27 Jul 2024
Viewed by 417
Abstract
To investigate the metabolomic mechanisms by which changes in cardiorespiratory fitness (CRF) levels affect metabolic syndrome (MetS) risk factors and to provide a theoretical basis for the improvement of body metabolism via CRF in people with MetS risk factors, a comparative blood metabolomics [...] Read more.
To investigate the metabolomic mechanisms by which changes in cardiorespiratory fitness (CRF) levels affect metabolic syndrome (MetS) risk factors and to provide a theoretical basis for the improvement of body metabolism via CRF in people with MetS risk factors, a comparative blood metabolomics study of individuals with varying levels of CRF and varying degrees of risk factors for MetS was conducted. Methods: Ninety subjects between the ages of 40 and 45 were enrolled, and they were categorized into low-MetS (LM ≤ two items) and high MetS (HM > three items) groups, as well as low- and high-CRF (LC, HC) and LCLM, LCLM, LCHM, and HCHM groups. Plasma was taken from the early morning abdominal venous blood. LC-MS was conducted using untargeted metabolomics technology, and the data were statistically and graphically evaluated using SPSS26.0 and R language. Results: (1) There were eight common differential metabolites in the HC vs. LC group as follows: methionine (↓), γ-aminobutyric acid (↑), 2-oxoglutatic acid (↑), arginine (↑), serine (↑), cis-aconitic acid (↑), glutamine (↓), and valine (↓); the HM vs. LM group are contrast. (2) In the HCHM vs. LCLM group, trends were observed in 2-oxoglutatic acid (↑), arginine (↑), serine (↑), cis-aconitic acid (↑), glutamine (↓), and valine (↓). (3) CRF and MetS risk factors jointly affect biological metabolic pathways such as arginine biosynthesis, TCA cycle, cysteine and methionine metabolism, glycine, serine, and threonine metabolism, arginine and proline metabolism, and alanine, aspartate, and glutamate metabolism. Conclusion: The eight common differential metabolites can serve as potential biomarkers for distinguishing individuals with different CRF levels and varying degrees of MetS risk factors. Increasing CRF levels may potentially mitigate MetS risk factors, as higher CRF levels are associated with reduced MetS risk. Full article
(This article belongs to the Special Issue Interactions between Exercise Physiology and Metabolism)
Show Figures

Figure 1

Figure 1
<p>Flowchart of different metabolic syndrome MetS and cardiorespiratory fitness CRF groupings. Participants were categorized into the low-risk group and the high-risk group (HM) if they had ≥3 MetS-related risk factors; CRF was categorized into low CRF level (LC), medium CRF level (MC), and high CRF level (HC) using the 3-quartile method; participants were classified into four groups based on the CRF level and the number of MetS-related risk factors: low CRF and low MS risk factor group (LCLM), low CRF and high MS risk factor group (LCHM), high CRF and low MS risk factor group (HCLM), and high CRF and high MS risk factor group (HCHM).</p>
Full article ">Figure 2
<p>(<b>a</b>) Total signal intensity in the amine/phenol-based channel sub-sample analysis; (<b>b</b>) Background peak mass distribution in the amine/phenol-based channel sub-sample analysis; (<b>c</b>) Retention time assay results; (<b>d</b>) Distribution of identified metabolites in different tiers of the database.</p>
Full article ">Figure 3
<p>PLS-DA Score Plot: (<b>a</b>) LC vs. MC vs. HC group; (<b>b</b>) LM vs. HM group (<b>c</b>) LCLM vs. HCLM; (<b>d</b>) LCHM vs. HCHM; (<b>e</b>) LCLM vs. LCHM; (<b>f</b>) HCLM vs. HCHM; (<b>g</b>) LCLM vs. HCHM; (<b>h</b>) LCHM vs. HCLM.</p>
Full article ">Figure 4
<p>The OPLS-DA score plots (<b>A</b>–<b>J</b>) depict the scores for each group, while the response ranking test plots (<b>a</b>–<b>j</b>) display the results of the response ranking test. In the ranking test, the horizontal axis represents the correlation between the <span class="html-italic">Y</span> values of the random grouping and the <span class="html-italic">Y</span> values of the original grouping, while the vertical axis represents the R<sup>2</sup> and Q<sup>2</sup> scores.</p>
Full article ">Figure 5
<p>Differential metabolite volcano plot (<b>a</b>) HC vs. LC, (<b>b</b>) HM vs. LM, (<b>c</b>) HCLM vs. LCLM, (<b>d</b>) HCHM vs. LCHM, (<b>e</b>) LCHM vs. LCLM, (<b>f</b>) HCHM vs. LCLM. The horizontal axis represents the fold change in metabolite expression across different subgroups [log2(FoldChange)], while the vertical axis indicates the significance level of differences [−log10 (<span class="html-italic">p</span>-value)]. Each point on the plot represents a metabolite, with red indicating a significant increase, blue indicating a significant decrease, and black indicating no significant difference. Plasma metabolites were visualized on a volcano plot based on their fold change (FC) values, <span class="html-italic">p</span>-values, and <span class="html-italic">q</span>-values. VIP value indicates the contribution of each variable to the PLS-DA model. Differential metabolites were identified based on VIP values &gt; 1, fold change (FC) &gt; 1.2 or &lt;0.83, <span class="html-italic">p</span> &lt; 0.05, and <span class="html-italic">q</span> &lt; 0.25.</p>
Full article ">Figure 6
<p>Venn diagram analysis of potential biomarkers in different groups.</p>
Full article ">Figure 7
<p>Histogram of trends in common differential metabolites by group. “*” indicating differential metabolite FC &gt; 1.2 or &lt;0.83, <span class="html-italic">p</span> &lt; 0.05, and <span class="html-italic">q</span> &lt; 0.25; “**” indicating differential metabolite FC &gt; 1.2 or &lt;0.83, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">q</span> &lt; 0.10; black column: HCHM group vs. LCLM group; grey column: HC group vs. LC group; white column: HM group vs. LM group. (<b>a</b>) 2-oxoglutaric acid; (<b>b</b>) L-arginine; (<b>c</b>) L-serine; (<b>d</b>) cis-aconitic acid; (<b>e</b>) L-glutamine; (<b>f</b>) L-valine.</p>
Full article ">Figure 8
<p>Bubble diagram of metabolic pathways of differential metabolites in each group. (<b>a</b>–<b>f</b>) represent the metabolic pathway bubble diagrams for each group, (<b>a</b>) HC vs. LC group; (<b>b</b>) HM vs. LM group; (<b>c</b>) HCLM vs. LCLM group; (<b>d</b>) HCHM vs. LCHM group; (<b>e</b>) LCHM group vs. LCLM group (<b>f</b>) HCHM vs. LCLM group (1): arginine biosynthesis; (2): tricarboxylic acid cycle(TCA); (3): cysteine and methionine metabolism; (4): glycine, serine, and threonine metabolism; (5): arginine and proline metabolism; (6): alanine, aspartic acid, and glutamate metabolism; (7): aminotransferase-tRNA biosynthesis; (8): Butyric acid metabolism; (9): glyoxylate and dicarboxylic acid metabolism; (10): purine metabolism; (11): glutathione metabolism; (12): glycerophospholipid metabolism; (13): D-glutamine and D-glutamate metabolism; (14): pyruvate metabolism; (15): glycolysis/glycohydrogenation; (16): taurine and hypotaurine metabolism. The graph’s <span class="html-italic">X</span>-axis represents the pathway influence factor, and the <span class="html-italic">Y</span>-axis shows the enrichment analysis’s <span class="html-italic">p</span>-value. The bigger the circle, the more influential it is; the darker the circle’s color, the lower the <span class="html-italic">p</span>-value and the more significant the enrichment.</p>
Full article ">
12 pages, 1241 KiB  
Review
Specific Biomarkers in Spinocerebellar Ataxia Type 3: A Systematic Review of Their Potential Uses in Disease Staging and Treatment Assessment
by Alexandra E. Soto-Piña, Caroline C. Pulido-Alvarado, Jaroslaw Dulski, Zbigniew K. Wszolek and Jonathan J. Magaña
Int. J. Mol. Sci. 2024, 25(15), 8074; https://doi.org/10.3390/ijms25158074 - 24 Jul 2024
Viewed by 419
Abstract
Spinocerebellar ataxia type 3 (SCA3) is the most common type of disease related to poly-glutamine (polyQ) repeats. Its hallmark pathology is related to the abnormal accumulation of ataxin 3 with a longer polyQ tract (polyQ-ATXN3). However, there are other mechanisms related to SCA3 [...] Read more.
Spinocerebellar ataxia type 3 (SCA3) is the most common type of disease related to poly-glutamine (polyQ) repeats. Its hallmark pathology is related to the abnormal accumulation of ataxin 3 with a longer polyQ tract (polyQ-ATXN3). However, there are other mechanisms related to SCA3 progression that require identifying trait and state biomarkers for a more accurate diagnosis and prognosis. Moreover, the identification of potential pharmacodynamic targets and assessment of therapeutic efficacy necessitates valid biomarker profiles. The aim of this review was to identify potential trait and state biomarkers and their potential value in clinical trials. Our results show that, in SCA3, there are different fluid biomarkers involved in neurodegeneration, oxidative stress, metabolism, miRNA and novel genes. However, neurofilament light chain NfL and polyQ-ATXN3 stand out as the most prevalent in body fluids and SCA3 stages. A heterogeneity analysis of NfL revealed that it may be a valuable state biomarker, particularly when measured in plasma. Nonetheless, since it could be a more beneficial approach to tracking SCA3 progression and clinical trial efficacy, it is more convenient to perform a biomarker profile evaluation than to rely on only one. Full article
Show Figures

Figure 1

Figure 1
<p>Flow diagram of EPPI v6 study selection following PRISM and STROBE guidelines.</p>
Full article ">Figure 2
<p>SCA 3 and biofluid markers. NfL = neurofilament light chain; CSF = cerebrospinal fluid; Aβ<sub>42</sub> = β-amyloid protein at amino acid 42; GFAP = glial fibrillary acidic protein; UCHL1 = ubiquitin carboxy-terminal hydrolase L1; pNfH = phosphorylated neurofilament heavy; IGF-1 = insulin grown factor 1; IGFBPs = IGF-binding proteins; NSE = neuron-specific enolase; CHIP = carboxyl terminus of Hsp-70 protein; SOD = superoxide dismutase; and GSH-Px = glutathione peroxidase.</p>
Full article ">Figure 3
<p>SCA3 stages are indicated by gray dashed lines. The Y axis represents the levels of biomarkers, and the X axis depicts disease progression (<b>a</b>). CSF biomarkers in SCA3 stages. Disease progression is shown along the X axis, and biomarker levels along the Y axis. Stages are indicated by gray dashed lines (<b>b</b>).</p>
Full article ">
17 pages, 4790 KiB  
Article
Integration of Illumina and PacBio HiFi Sequencing Reveals a Three-Linear-Molecule Mitogenome with RNA-Editing Sites and Phylogeny in Arrow Bamboo (Fargesia qinlingensis)
by Hao Wu, Xue Li, Ke Qu, Lele Yang, Tao Su, Lijun Yong, Mei Han and Fuliang Cao
Forests 2024, 15(7), 1267; https://doi.org/10.3390/f15071267 - 20 Jul 2024
Viewed by 580
Abstract
Arrow bamboo (Fargesia qinlingensis) is endemic to the Qinling Mountains and has remarkable adaptive resilience to changing climates. However, its complete mitogenome remains unknown. Using the Illumina and PacBio HiFi sequencing platforms, we found that the mitogenome assembly of the F. [...] Read more.
Arrow bamboo (Fargesia qinlingensis) is endemic to the Qinling Mountains and has remarkable adaptive resilience to changing climates. However, its complete mitogenome remains unknown. Using the Illumina and PacBio HiFi sequencing platforms, we found that the mitogenome assembly of the F. qinlingensis has a multi-branched skeleton comprising three linear molecules (M1, M2, and M3), with a length of 442,368 bp and a GC content of 44.05%. Thirty-five unique PCGs were identified in the complete mitogenome, including twenty-four core structural genes, eleven noncore structural genes, three rRNAs, and sixteen tRNAs. The GCU for alanine and CAA for glutamine represented the most significant frequency (RSCU = 1.55) in the codon usage preference. A total of 51, 28, and 14 SSRs were determined on M1, M2, and M3, respectively. The mitogenome contained 149 pairs of dispersed repeats with lengths greater than 30 bp, the most abundant of which were 82 forward and 67 palindromic repeats. A long repeat sequence (14,342 bp) was characterized in mediating mitogenome recombination. DNA transfer analyses suggested that 44 MTPTs (30,943 bp, 6.99%) originated from the plastome. Among the 482 potential C-U/T RNA-editing sites predicted in 35 PCGs, ccmFn (38 times) and ccmC (36 times) shoed the highest frequency. Collinearity and phylogenetic trees revealed the close relationship between F. qinlingensis and Bambusa oldhamii. The primary features of the mitogenome of F. qinlingensis will help decipher the functional mitochondrial traits related to growth performance and climate resilience. Moreover, our findings provide insights into the evolution, environmental adaptation, and sustainable use of subalpine bamboo resources in the Qinling Mountains. Full article
(This article belongs to the Special Issue Genomic Analysis of Growth and Stress Adaptation in Forest Trees)
Show Figures

Figure 1

Figure 1
<p>Schematic structure of the mitogenome of <span class="html-italic">F. qinlingensis</span> determined using the Bandage software. (<b>A</b>) Draft of the mitogenome assembly with six branch contigs. (<b>B</b>) Multi-branched mitogenome with overlapping nodes colored purple, blue, and yellow. Repetitive sequences are colored red. Green regions indicate the DNA sequences transferred from the plastome. (<b>C</b>) Simplified mitogenome of three overlapping linear molecules (M1, M2, and M3), supported by the PacBio HiFi long-read sequencing data.</p>
Full article ">Figure 2
<p>Schematic map of three linearized mitogenomes and gene annotations in <span class="html-italic">F. qinlingensis</span>. Different annotated gene families are shown above/below the lines in three linear molecules: M1, M2, and M3.</p>
Full article ">Figure 3
<p>Codon usage preference in 35 PCGs based on the RSCU. <span class="html-italic">X</span>-axis indicates the codon families for amino acids. <span class="html-italic">Y</span>-axis represents the RSCU value. RSCU, relative synonymous codon usage.</p>
Full article ">Figure 4
<p>Repetitive sequences identified in the three molecules, M1, M2, and M3, of <span class="html-italic">F. qinlingensis</span>. Cyclic maps illustrate the distributions of three types of linked repetitive sequences, including dispersed repeats (palindromic repeats in orange and forward repeats in blue), tandem repeats (black lines in the second ring), and SSRs (black lines in the outer ring).</p>
Full article ">Figure 5
<p>Plastome and plastid DNA migration in the mitogenome. (<b>A</b>) Assembled circular plastome, showing the sequence length and annotated genes. (<b>B</b>) MTPTs identified in the <span class="html-italic">F. qinlingensis</span> mitogenome. Green ribbons represent homologous MTPTs transferred between two organelles. MTPT, mitochondrial plastid DNA.</p>
Full article ">Figure 6
<p>Prediction of RNA-editing sites in 35 PCGs and the experimental evaluation. (<b>A</b>) PCR amplification was conducted using DNA and cDNA as templates. (<b>B</b>) Experimental confirmation of the RNA-editing sites in selected PCGs (<span class="html-italic">atp9-191</span>, <span class="html-italic">cox2-353</span>, <span class="html-italic">-608</span>, <span class="html-italic">nad1-216</span>, and <span class="html-italic">nad7-224</span>), showing the exemplified C to T changes (dotted line framed) via Sanger sequencing.</p>
Full article ">Figure 7
<p>Collinearity and evolutionary relationships between <span class="html-italic">F. qinlingensis</span> and other plants. (<b>A</b>) Synteny analyses of seven mitogenomes within Pooideae and Bambusoideae. Pink-curved ribbons represent DNA inversion regions, and gray-curved ribbons indicate sites with high homology. Collinear blocks with sequence lengths &lt; 0.5 kb were not retained. (<b>B</b>) Phylogenetic relationships between <span class="html-italic">F. qinlingensis</span> and 28 genetically related species shown using full-length mitogenome sequences. This unrooted tree was constructed using the neighbor-joining method.</p>
Full article ">
9 pages, 700 KiB  
Article
Perioperative Glutamine Supplementation May Restore Atrophy of Psoas Muscles in Gastric Adenocarcinoma Patients Undergoing Gastrectomy
by Jin-Ming Wu, Hsing-Hua Tsai, Shang-Ming Tseng, Kao-Lang Liu and Ming-Tsan Lin
Nutrients 2024, 16(14), 2301; https://doi.org/10.3390/nu16142301 - 17 Jul 2024
Viewed by 724
Abstract
Background: Sarcopenia, characterized by degenerative skeletal muscle loss, is increasingly linked to poor surgical outcomes. Glutamine, an immune-modulating formula, may stimulate muscle protein synthesis and inhibit degradation. We used the psoas major muscle area (PMMA) at the third lumbar vertebra, normalized for height [...] Read more.
Background: Sarcopenia, characterized by degenerative skeletal muscle loss, is increasingly linked to poor surgical outcomes. Glutamine, an immune-modulating formula, may stimulate muscle protein synthesis and inhibit degradation. We used the psoas major muscle area (PMMA) at the third lumbar vertebra, normalized for height (PMMA index), as a skeletal muscle indicator. This study investigates whether perioperative glutamine supplementation mitigates psoas muscle atrophy. Methods: We enrolled gastric adenocarcinoma (GA) patients undergoing gastrectomy. Computed tomography assessed the psoas muscle short axis. Muscle atrophy was estimated by changes between preoperative and three-month post-gastrectomy scans. Perioperative glutamine supplementation (PGS) comprised five-day parenteral plus one-month oral use. Propensity score matching minimized potential bias. A linear regression model predicted the association. Results: Of 516 patients analyzed (2016–2019), 100 (19.4%) received PGS. After propensity score matching, each group contained 97 cases. The PGS group showed a significantly higher median PMMA index change than the non-PGS group (0.3 vs. −0.3 cm2/m2, p = 0.004). Multivariate analysis revealed that PGS was significantly associated with increased PMMA index (coefficient = 0.60; 95% CI: 0.19–1.01; p = 0.005). Conclusions: PGS may help restore psoas muscle atrophy in GA patients undergoing gastrectomy. The underlying mechanisms likely relate to glutamine’s role in protein metabolism and immune function. Further studies are needed to elucidate these mechanisms fully. Full article
Show Figures

Figure 1

Figure 1
<p>Flow diagram for the inclusion and exclusion criteria in this study.</p>
Full article ">Figure 2
<p>Distribution of the estimated propensity score for receiving perioperative parenteral glutamine supplementation between two groups before and after matching.</p>
Full article ">
26 pages, 8101 KiB  
Article
Enhanced Removal of Rhodamine b Dye from Aqueous Media via Adsorption on Facilely Synthesized Zinc Ferrite Nanoparticles
by Asma S. Al-Wasidi and Salwa AlReshaidan
Inorganics 2024, 12(7), 191; https://doi.org/10.3390/inorganics12070191 - 12 Jul 2024
Viewed by 548
Abstract
This paper studies the synthesis, characterization, and application of ZnFe2O4 nanoparticles for the removal of rhodamine b dye from aqueous media. Utilizing the combustion procedure, ZnFe2O4 nanoparticles were synthesized using two different fuels: glutamine (SG) and L-arginine [...] Read more.
This paper studies the synthesis, characterization, and application of ZnFe2O4 nanoparticles for the removal of rhodamine b dye from aqueous media. Utilizing the combustion procedure, ZnFe2O4 nanoparticles were synthesized using two different fuels: glutamine (SG) and L-arginine (SA). In addition, the synthesized ZnFe2O4 nanoparticles were characterized through various techniques, including Fourier transform infrared (FTIR), X-ray diffraction (XRD), field emission scanning electron microscope (FE-SEM), energy-dispersive X-ray (EDX), high resolution transmission electron microscope (HR-TEM), and Brunauer-Emmett-Teller (BET) surface area analysis. XRD analysis verified the creation of a ZnFe2O4 cubic spinel structure without any contaminants, revealing average crystallite sizes of 43.72 and 29.38 nm for the SG and SA samples, respectively. The FTIR spectra exhibited peaks indicative of metal-oxygen bond stretching, verifying the presence of a spinel formation. Elemental analysis via EDX confirmed the stoichiometric composition typical of zinc ferrite. In addition, FE-SEM imaging displayed that the SG and SA samples are composed of particles with irregular and spherical shapes, measuring average diameters of 135.11 and 59.89 nm, respectively. Furthermore, the BET surface area of the SG and SA samples is 60 and 85 m2/g, respectively. The maximum adsorption capacity of the SA sample (409.84 mg/g) towards rhodamine b dye was higher than that of the SG sample (279.33 mg/g), which was ascribed to its larger surface area and porosity. Kinetic and equilibrium studies revealed that the adsorption process of rhodamine b dye onto the SG and SA samples followed the Langmuir isotherm and pseudo-second-order model. Thermodynamic analysis indicated that the adsorption process was spontaneous, exothermic, and physical. The study concludes that ZnFe2O4 nanoparticles synthesized using L-arginine (SA) exhibit enhanced rhodamine b dye removal efficiency due to their smaller size, increased surface area, and higher porosity compared to those synthesized with glutamine (SG). The optimum conditions for the adsorption process of rhodamine b dye were found to be at pH 10, a contact time of 70 min, and a temperature of 298 K. These findings underscore the potential of L-arginine-synthesized ZnFe2O4 nanoparticles for effective and sustainable environmental cleanup applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD analysis of the SG (<b>A</b>) and SA (<b>B</b>) products.</p>
Full article ">Figure 2
<p>FTIR spectra of the SG (<b>A</b>) and SA (<b>B</b>) products.</p>
Full article ">Figure 3
<p>EDX patterns of the SG (<b>A</b>) and SA (<b>B</b>) products.</p>
Full article ">Figure 4
<p>FE-SEM analysis of the SG (<b>A</b>) and SA (<b>B</b>) products.</p>
Full article ">Figure 5
<p>HR-TEM analysis of the SG (<b>A</b>) and SA (<b>B</b>) products.</p>
Full article ">Figure 6
<p>N<sub>2</sub> adsorption-desorption analysis of the SG (<b>A</b>) and SA (<b>B</b>) products.</p>
Full article ">Figure 7
<p>Influence of pH on the adsorption of the rhodamine b dye by the SG and SA products (<b>A</b>). The point of zero charge of the SG and SA samples (<b>B</b>).</p>
Full article ">Figure 8
<p>Influence of contact time on the adsorption of the rhodamine b dye by the SG and SA samples.</p>
Full article ">Figure 9
<p>Linear kinetic modeling of rhodamine b dye adsorption onto the SG and SA samples by the pseudo-first-order (<b>A</b>) and pseudo-second-order (<b>B</b>) kinetic plots.</p>
Full article ">Figure 10
<p>Nonlinear kinetic modeling of rhodamine b dye adsorption onto the SG and SA samples by the pseudo-first-order (<b>A</b>) and pseudo-second-order (<b>B</b>) kinetic plots.</p>
Full article ">Figure 11
<p>Influence of temperature on the adsorption of the rhodamine b dye by the SG and SA samples.</p>
Full article ">Figure 12
<p>Thermodynamic analysis of rhodamine b dye adsorption on the SG and SA samples via Van’t Hoff plots.</p>
Full article ">Figure 13
<p>Impact of initial dye concentration on the adsorption of the rhodamine b dye by the SG and SA samples.</p>
Full article ">Figure 14
<p>Equilibrium linear modeling of rhodamine b dye adsorption onto the SG and SA samples by the Langmuir (<b>A</b>) and Freundlich (<b>B</b>) equilibrium plots.</p>
Full article ">Figure 15
<p>Equilibrium nonlinear modeling of rhodamine b dye adsorption onto the SG and SA samples by the Langmuir (<b>A</b>) and Freundlich (<b>B</b>) equilibrium plots.</p>
Full article ">Figure 16
<p>Effect of desorption of rhodamine b dye ions from the SG and SA adsorbents. Experimental conditions: contact time = 70 min, initial dye concentration = 250 mg/L, volume of dye = 100 mL, pH = 10, adsorbent dose = 0.05 g, temperature = 298 K, concentration of HCl eluting agent range = 3–9 M, volume of HCl = 50 mL, desorption time = 30 min.</p>
Full article ">Figure 17
<p>Effect of reusability of the SG and SA adsorbents on the sorption of rhodamine b dye. Experimental conditions: initial dye concentration = 250 mg/L, volume of dye solution = 100 mL, amount of adsorbent = 0.05 g, pH = 10, contact adsorption time = 70 min, adsorption temperature = 298 K, concentration of HCl eluting agent = 9 M, volume of HCl = 50 mL, desorption time = 30 min.</p>
Full article ">Scheme 1
<p>Production of zinc ferrite (ZnFe<sub>2</sub>O<sub>4</sub>) nanoparticles by the combustion method using glutamine (<b>A</b>) and L-arginine (<b>B</b>) fuels.</p>
Full article ">Scheme 2
<p>Elimination mechanism of rhodamine b dye by ZnFe<sub>2</sub>O<sub>4</sub> nanoparticles.</p>
Full article ">Scheme 3
<p>Practical procedure for the fabrication of SG (<b>A</b>) and SA (<b>B</b>) samples by the combustion method.</p>
Full article ">Scheme 4
<p>Experimental influences for the study of rhodamine b dye adsorption by the SG and SA products.</p>
Full article ">
Back to TopTop